10.8
CiteScore
 
5.2
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
10.8
CiteScore
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
10 (
1_suppl
); S261-S273
doi:
10.1016/j.arabjc.2012.07.032

Comparative study of N-[(4-methoxyphenyl) (morpholin-4-yl)methyl]acetamide (MMPA) and N-[morpholin-4-yl(phenyl)methyl]acetamide (MPA) as corrosion inhibitors for mild steel in sulfuric acid solution

Department of Chemistry, Jamal Mohamed College (Autonomous), Tiruchirappalli 20, Tamil Nadu, India

⁎Corresponding author. anwarchemsathiq@yahoo.co.in (M. Anwar Sathiq)

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University.

Abstract

Two Mannich bases namely, N-[(4-methoxyphenyl)(morpholin-4-yl)methyl]acetamide (MMPA) and N-[morpholin-4-yl(phenyl)methyl]acetamide (MPA) were synthesized and their influence on the inhibition of corrosion of mild steel in 1.0 M H2SO4 was investigated by weight loss, potentiodynamic polarization, electrochemical impedance spectroscopy (EIS), scanning electron microscopy (SEM) and FT-IR spectroscopy. The weight loss measurements showed that these inhibitors have excellent inhibiting effect at a concentration of 0.01 M. The inhibitor efficiency was found to depend on both concentration and molecular structure of the inhibitor. Both the compounds have been found to be relatively good inhibitors. Potentiodynamic polarization curves revealed that the studied inhibitors represent a mixed-type, predominantly cathodic control. An equivalent circuit is suggested based on an analysis of EIS data. The negative value of standard free energy of adsorption in the presence of inhibitor suggests spontaneous adsorption of inhibitors on the mild steel surface. The activation energy of corrosion and other thermodynamic parameters were calculated to elaborate the mechanism of corrosion inhibition. The Temkin isotherm was found to provide an accurate description of the adsorption behavior of the inhibitors. Surface analysis using scanning electron microscope (SEM) shows a significant morphological improvement on the mild steel surface with the addition of inhibitors. FT-IR spectra revealed the interaction between inhibitor molecules and mild steel surface.

Keywords

Mannich base
Mild steel
Adsorption isotherm
Electrochemical impedance spectroscopy
Thermodynamic parameters
Activation energy
1

1 Introduction

Corrosion is a naturally occurring phenomenon which deteriorates a metallic material or its properties because of a reaction with its environment. Corrosion can cause dangerous and expensive damage to everything from pipelines, bridges and public buildings to vehicles, water and waste water systems, and even home appliances. Among numerous anticorrosion measures, corrosion inhibitor is widely used and acts as one of the most economical and effective ways (Zhang et al., 2011; El-Taib Heakal et al., 2011; Sherif et al., 2010; Amin et al., 2010). Inhibitors often work by adsorbing themselves onto the metallic surface, protecting it by forming a film. The strength of the adsorption bond is the dominant factor for organic inhibitors. Their effectiveness depends on the chemical composition, their chemical structure and their affinity for the metal surface. Of great interest is to find a relationship between the inhibitive properties and the molecular structure of organic compounds (Jamal Abdul Nasser, 2005; Jamal Abdul Nasser and Anwar Sathiq, 2007; Lazarova et al., 2008; Li et al., 2008; Elewady, 2008; Ramesh Saliyan and Vasudeva Adhikari, 2009; Popova et al., 2011).

A perusal of the literature reveals corrosion inhibitors derived from Mannich base constitute an important and potential class. However, very little work has been done on Mannich base derivatives as corrosion inhibitors (Saukaitis and Gardner, 1970; Singh and Quraishi, 2010;).

In continuation of the earlier work on the development of Mannich bases as acid corrosion inhibitors (Jamal Abdul Nasser and Anwar Sathiq, 2011), the authors have studied the corrosion inhibiting behavior of Mannich bases, namely N-[(4-methoxyphenyl)(morpholin-4-yl)methyl]acetamide (MMPA) and N-[morpholin-4-yl(phenyl)methyl]acetamide (MPA) on the corrosion of mild steel in sulfuric acid solutions.

2

2 Materials and methods

2.1

2.1 Materials

Mild steel strips with the composition carbon = 0.07%; sulfur = 0%; phosphorus = 0.008%; silicon = 0%; manganese = 0.34% and iron = remainder and size of 4 × 1 × 0.025 cm were used for weight loss and effect of temperature studies. Mild steel cylindrical rods of the same composition embedded in polytetrafluoroethylene (PTFE) with an exposed area of 1 cm2 were used for potentiodynamic polarization and impedance measurements. The electrode was polished using a sequence of emery papers of different grades and then degreased with acetone. N-[(4-methoxyphenyl)(morpholin-4-yl)methyl]acetamide (MMPA) and N-[morpholin-4-yl(phenyl)methyl]acetamide were synthesized (Raman et al., 2004), and purified by recrystallization from ethanol to analytical purity grade. Its purity was confirmed by elemental analysis and characterized by FT-IR and NMR spectroscopy. The name and molecular structure of studied Mannich bases are given in Fig. 1.

Name and molecular structure of studied Mannich bases.
Figure 1
Name and molecular structure of studied Mannich bases.

The acid solution (1.0 M H2SO4) was prepared by the dilution of an analytical grade H2SO4 with double distilled water. All tests were conducted at different temperatures in magnetically stirred solutions.

2.2

2.2 Weight loss measurements

Weight loss measurements were done according to the method described previously (Jamal Abdul Nasser, 2005; Jamal Abdul Nasser and Anwar Sathiq, 2011). Weight loss measurements were performed at 303 ± 1 K for 2 h by immersing the mild steel coupons into acid solution (100 mL) without and with various amounts of inhibitor. After the elapsed time, the specimens were taken out, washed, dried, and weighed accurately. All the tests were performed in triplicate and the average values were reported. All the concentrations of an inhibitor for weight loss and electrochemical study were taken in M.

The surface coverage (θ) and inhibition efficiency (IE%) were determined by using the following equation:

(1)
θ = W o - W 1 W o
(2)
I . E . ( % ) = W o - W 1 W o X 100
Where, W1 and Wo are the weight loss values in the presence and absence of inhibitors, respectively.

2.3

2.3 Effect of temperature

The loss in weights was calculated at different temperatures from 303 to 333 K. Each experiment was duplicated to get good reproducibility. Weight loss measurements were performed in 1.0 M H2SO4 with and without the addition of the inhibitor at their best inhibiting concentrations. Percentage inhibitions of the inhibitor at various temperatures were calculated.

2.4

2.4 Potentiodynamic polarization measurements

Both cathodic and anodic polarization curves were recorded (mVs−1) using the corrosion measurement system BAS (Model: 100 A) computerized electrochemical analyzer and PL-10 digital Plotter. A platinum foil and Hg| Hg2SO4| 1.0 M H2SO4 electrode was used as auxiliary and reference electrodes respectively. The Tafel polarization curves were obtained by changing the electrode potential automatically from ± 0.200 V at open circuit potential with a scan rate 1.0 mV s−1.

2.5

2.5 Electrochemical impedance measurements

Electrochemical measurements were carried out in the conventional three-electrode, cylindrical Pyrex glass cell with a capacity of 1000 mL. A saturated calomel electrode (SCE) and a platinum disc electrode were used, respectively, as reference and auxiliary electrodes. The temperature was controlled at 303 K by a thermostat. The working electrode was abraded with silicon carbide paper (grade P1200), degreased with AR grade ethanol and acetone, and rinsed with double distilled water before use. All potentials are reported versus SCE. EIS measurements were carried out in a frequency range of 100,000–0.01 Hz with an amplitude of 10 mV using AC signals at open circuit potential.

2.6

2.6 Standard free energy of adsorption

The standard adsorption free energy changes in 0.01 M concentration (best inhibition) of MMPA and MPA at different temperatures (313 K to 333 K) were calculated (Quraishi et al., 2010) using the equation ΔG0 = −RT ln (K 55.5) kJ/mole, Where ΔG0 = adsorption free energy, R = Gas Constant T = Temperature and K = Adsorptive equilibrium constant. The value of 55.5 is the concentration of water in solution expressed in mol L−1. The value of K was calculated from K = θ/C (1−θ). Where θ = Surface coverage (Inhibition efficiency/100) and C = inhibitor concentration.

2.7

2.7 SEM analysis

The specimens used for surface morphological examination were immersed in acid containing various concentrations of inhibitor and blank for 2 h. Then they were removed, rinsed quickly with rectified spirit, and dried. The analysis was performed on HITACHI-model S-3000 H SEM.

2.8

2.8 FT-IR spectroscopy

The percentage transmission is recorded against wave number. The mild steel specimens were immersed in various test solutions for a period of 2 h. After 2 h, the specimens were taken out and dried. The surface film was scratched carefully and its FT-IR spectra were recorded using Perkin–Elmer make model spectrum RXI instrument.

3

3 Results and discussion

3.1

3.1 Weight loss measurements

The effect of concentration of MMPA and MPA on the corrosion of mild steel in 1.0 M H2SO4 is given in Table 1. The corrosion rate decreased considerably with an increase in concentration of both the inhibitors and reached the minimum value in the range of 0.1 M concentrations. It is obvious that inhibition efficiency values for the two tested Mannich bases increases with the increase in inhibitor concentration, and this increase in the inhibition efficiency, at a given inhibitor concentration, enhances in the following order: MPA < MMPA. The greater inhibitive power of MMPA is due to the presence of electron donating p-OCH3 group that increases the electron density of the phenyl ring. The inhibition action of Mannich bases can be explained by considering the following mechanism: Fe (Inh)ads, reaction intermediates (Bockris and Drazic, 1962): Fe + Inh ↔ Fe (Inh)ads ↔ Fen+ + ne + Inh. At first, when there is not enough Fe (Inh)ads to cover the metal surface, because the inhibitor concentration is low or because the adsorption rate is slow, metal dissolution takes place in sites on the mild steel surface free of Fe (Inh)ads. With high inhibitor concentration, a compact and coherent inhibitor over the film is formed on the mild steel which reduces chemical attacks on the metal.

Table 1 Corrosion rate, inhibition efficiency and surface coverage of mild steel immersed in 1.0 M H2SO4 for various concentrations of inhibitors obtained by weight loss method at 303 ± 1 K.
Name of the inhibitor Concentration of inhibitor (M) Corrosion rate (mpy) I.E. (%) Surface coverage (θ)
MMPA Blank 2.5496
0.0000001 1.2748 50.00 0.5000
0.000001 0.9915 61.11 0.6111
0.000005 0.8410 67.01 0.6701
0.00001 0.7082 72.22 0.7222
0.00005 0.6551 74.31 0.7431
0.0001 0.5666 78.13 0.7813
0.0005 0.3807 85.10 0.8510
0.001 0.2833 88.89 0.8889
0.005 0.2213 91.32 0.9132
0.01 0.1505 94.10 0.9410
MPA Blank 2.5496
0.0000001 1.8231 28.47 0.2847
0.000001 1.4873 41.67 0.4167
0.000005 1.1951 53.13 0.5313
0.00001 0.8587 66.32 0.6632
0.00005 0.7082 72.22 0.7222
0.0001 0.6108 76.04 0.7604
0.0005 0.4515 82.29 0.8229
0.001 0.4072 84.03 0.8403
0.005 0.2921 88.54 0.8854
0.01 0.2213 91.32 0.9132

3.2

3.2 Effect of temperature

The effect of temperature on inhibition reaction is highly complex, because many changes may occur on the metal surface such as rapid etching, rupture, desorption of inhibitor and the decomposition and/or rearrangement of inhibitor (Ramesh Saliyan and Vasudeva Adhikari, 2008). The effect of temperature on the corrosion inhibition with and without inhibitor is shown in Table 2. It can be seen that the weight loss increases with temperature in the absence and presence of inhibitor. Adsorption and desorption of inhibitor molecules continuously occur at the metal surface and an equilibrium exists between two processes at a particular temperature. With the increase of temperature, the equilibrium between the adsorption and desorption processes is shifted leading to a higher desorption rate than adsorption until equilibrium is again established at a different value of equilibrium constant. It explains the lower inhibition efficiency at higher temperatures (Chaudhary et al., 2007: Sanyal and Kumar, 2010).

Table 2 Values of corrosion rate, inhibition efficiency and surface coverage for different temperatures in the presence of 0.01 M concentrations of inhibitors in 1.0 M H2SO4.
Name of the inhibitor Temperature (K) Corrosion rate (mpy) I.E. (%) Surface coverage (θ)
MMPA 303 0.1505 94.10 0.9410
308 0.1771 93.17 0.9317
313 0.2125 92.11 0.9211
318 0.3364 87.80 0.8780
323 0.4692 83.99 0.8399
328 0.6108 80.62 0.8062
333 0.9295 77.50 0.7750
MPA 303 0.2213 91.32 0.9132
308 0.2390 90.78 0.9078
313 0.2833 89.47 0.8947
318 0.3984 85.67 0.8567
323 0.6285 78.55 0.7855
328 0.9738 69.10 0.6910
333 1.4076 65.93 0.6510

In order to calculate activation parameters for the corrosion process, Arrhenius Eq. (3) and transition state Eq. (4) were used (Quraishi et al., 2010).

(3)
Log ( CR ) = - Ea / 2.303 RT + log λ
(4)
CR = RT / Nh exp ( - Δ H a o / RT ) exp ( Δ S a o / R )
Where Ea is the apparent activation energy, λ is the Arrhenius pre-exponential factor, CR is the corrosion rate, Δ H a o the enthalpy of activation, Δ S a o the entropy of activation, N the Avogadro's number, R the universal gas constant and T the absolute temperature.

The apparent activation energy (Ea) at the optimum concentration of MMPA and MPA determined by linear regression between log (CR) and 1000/T in 1 M H2SO4 medium has been shown in Fig. 2. The results are shown in Table 3.

The Arrhenius plots of log Rate vs 1000/T for the effect of temperature on the performance of MMPA and MPA at the optimum inhibition concentration of 0.01 M.
Figure 2
The Arrhenius plots of log Rate vs 1000/T for the effect of temperature on the performance of MMPA and MPA at the optimum inhibition concentration of 0.01 M.
Table 3 The values of activation parameters Ea, Δ H a o and Δ S a o for mild steel in 1.0 M H2SO4 in the absence and presence of inhibitors.
Name of the inhibitor Ea (kJ mol−1) Δ H a o (kJ mol−1) Δ S a o (J mol−1 K−1)
Blank 45.19 5.14 −220.58
MMPA 46.17 49.24 −99.40
MPA 52.78 51.78 −88.65

Radovici classifies the inhibitors into three groups according to temperature effects:

  1. Inhibitors whose inhibition efficiency (I.E.) decreases with temperature increase. The value of the apparent activation energy Ea found is greater than that in the uninhibited solution.

  2. Inhibitors whose I.E. does not change with temperature variation. The apparent activation energy does not change with the presence or absence of inhibitors.

  3. Inhibitors in whose presence the I.E. increases with the temperature increase while the value of Ea for the corrosion process is smaller than that obtained in the uninhibited solution. This is an indication, according to the literature, for a specific type of adsorption of the inhibitors (Radovici, 1965).

From the present set of Mannich bases in 1 M H2SO4, the inhibition efficiency decreases with temperature and the Ea values are greater as compared to blank acid value (Table 3). The increase in the apparent activation energy may be interpreted as physical adsorption that occurs in the first stage (Quraishi et al., 2010; Parameswari et al., 2010).

The relationship between log (CR/T) and 1000/T is shown in Fig. 3. Straight lines are obtained with a slope ( - Δ H a o / 2.303 R ) and an intercept of log ( R / Nh + Δ S a o / 2.303 R ) from which the values of Δ H a o and Δ S a o were calculated and is presented in Table 3.

The plots of log (CR/T) vs 1000/T for the effect of temperature on the performance of MMPA and MPA at the optimum inhibition concentration of 0.01 M.
Figure 3
The plots of log (CR/T) vs 1000/T for the effect of temperature on the performance of MMPA and MPA at the optimum inhibition concentration of 0.01 M.

Inspection of these data revealed that activation parameters for dissolution reaction of mild steel in 1.0 M H2SO4 in the presence of Mannich bases are higher (49.24 kJ mol−1 for MMPA and 51.78 for MPA) than those in the absence of inhibitors (5.14 kJ mol−1). The positive signs of Δ H a o reflect the endothermic nature of the mild steel dissolution process suggesting that the dissolution of mild steel is slow (Guan et al., 2004) in the presence of inhibitors.

The thermodynamic values obtained are the algebraic sum of the adsorption of organic molecules and desorption of water molecules (Branzol et al., 2000). Hence, the gain in entropy is attributed to the increase in solvent entropy and to more positive water desorption enthalpy (Ateya et al., 1984). The less negative value (when compared to free acid) of entropy of activation for Mannich bases also suggests that an increase in disordering takes place in going from reactants to the metal/solution interface, which is the driving force for the adsorption of inhibitors onto the mild steel surface.

3.3

3.3 Potentiodynamic polarization studies

The values of corrosion potential (Ecorr), corrosion current density (Icorr), and anodic (ba) and cathodic (bc) Tafel slopes can be evaluated from anodic and cathodic regions of the Tafel plots. The linear Tafel segments of anodic and cathodic curves were extrapolated to corrosion potential to obtain corrosion current densities (Icorr).

The inhibition efficiency was evaluated from the measured Icorr values using the relationship: Inhibition efficiency = i corr o - i corr o / i corr o × 100 Where, i corr o and i corr i are values of corrosion current density in the absence and in the presence of inhibitor, respectively.

Figs. 4 and 5 represent the potentiodynamic polarization curves of mild steel in 1.0 M H2SO4 in the absence and presence of various concentrations of the MMPA and MPA. It can be seen from Figs. 4 and 5 that, in the presence of inhibitors, the curves are shifted to lower current regions, showing the inhibition tendency of the Mannich bases. There was no definite trend observed in the Ecorr values in the presence of both the Mannich bases. In the present study, shift in Ecorr values is in the range of 0.200–0.250 V suggesting that they all act as mixed type inhibitors (Singh and Quraishi, 2010).

Potentiodynamic polarization curve for mild steel in 1 M H2SO4 in the absence and presence of various concentrations of MMPA.
Figure 4
Potentiodynamic polarization curve for mild steel in 1 M H2SO4 in the absence and presence of various concentrations of MMPA.
Potentiodynamic polarization curve for mild steel in 1 M H2SO4 in the absence and presence of various concentrations of MPA.
Figure 5
Potentiodynamic polarization curve for mild steel in 1 M H2SO4 in the absence and presence of various concentrations of MPA.

The values of various electrochemical parameters derived by Tafel polarization of the inhibitors are given in Table 4. Investigation Table 4 revealed that the values of ba change slightly in the presence of both the inhibitors, whereas a more pronounced change occurs in the values of bc, indicating that both anodic and cathodic reactions are affected but the effect on the cathodic reactions is more prominent. Thus, both the MMPA and MPA acted as mixed type corrosion inhibitors, but predominantly cathodic inhibitor (Jamal Abdul Nasser and Anwar Sathiq, 2011).

Table 4 Potentiodynamic polarization parameters and inhibition efficiency values for mild steel in 1.0 M H2SO4 in the absence and presence of different concentrations of MMPA and MPA.
Name of the inhibitor C (M) Ecorr (V) Tafel slope Icorr (A/cm2) I.E. (%)
ba (V/dec) bc (V/dec)
MMPA Blank −0.4559 0.1251 0.1311 2.74 × 10−3
0.000001 −0.4652 0.1291 0.1417 1.38 × 10−3 49.64
0.0001 −0.4622 0.1359 0.1613 6.50 × 10−4 76.28
0.01 −0.4725 0.1417 0.1819 1.81 × 10−4 93.39
MPA Blank −0.4559 0.1251 0.1311 2.74 × 10−3
0.000001 −0.4482 0.1282 0.1462 1.61 × 10−3 41.61
0.0001 −0.4721 0.1326 0.1612 6.20 × 10−4 75.84
0.01 −0.4742 0.1372 0.1732 2.60 × 10−4 90.51

3.4

3.4 Electrochemical impedance spectroscopy

The corrosion of mild steel in 1.0 M H2SO4 solution in the absence and presence of MMPA and MPA were investigated by EIS at the open circuit potential condition. Nyquist plots for mild steel obtained at the interface in the absence and presence of inhibitors at optimum concentrations are given in Figs. 5 and 6.

Impedance diagram for mild steel in 1 M H2SO4 in the absence and presence of various concentration of MMPA.
Figure 6
Impedance diagram for mild steel in 1 M H2SO4 in the absence and presence of various concentration of MMPA.

The Nyquist diagram obtained with 1.0 M H2SO4 shows only one capacitive loop and the diameter of the semicircle increases on increasing the inhibitor concentration suggesting that the formed inhibitive film was strengthened by the addition of inhibitors. All the obtained plots show only one semicircle and they were fitted using one time constant equivalent model (Randle's model) with double layer capacitance (Cdl), corrosion current density (Icorr) and charge transfer resistance (Rt). All the main parameters deduced from the impedance technique are given in Table 5.

Table 5 Impedance parameters and inhibition efficiency values for mild steel in 1.0 M H2SO4 in the absence and presence of different concentrations of inhibitors.
Name of the inhibitor C (M) Rt (ohm cm2) Icorr (mA/cm2) Cdl (μF/cm2) I.E. (%)
MMPA Blank 14.61 5.54 81.66
0.000001 68.24 2.20 66.32 60.29
0.0001 100.12 1.23 27.47 77.80
0.01 102.14 0.34 18.18 93.86
MPA Blank 14.61 5.54 81.66
0.000001 21.65 2.53 67.72 40.43
0.0001 32.12 1.21 29.84 74.91
0.01 45.40 0.43 19.86 91.15

Deviation from perfect circular shape, often known as frequency dispersion, was attributed to surface roughness and inhomogeneities of the solid surface (Lenderink et al., 1993). The change of icorr and Rt values can be related to the gradual replacement of water molecules and /or hydroxyl ions by Mannich bases molecules on the mild steel surface and consequently to a decrease in the number of active sites necessary for the corrosion reaction.

Moreover, the values of double layer capacitance (Cdl) decreased with increasing inhibitor concentrations. The decrease in Cdl is probably due to the decrease in local dielectric constant and/or an increase in the thickness of the protective layer at electrode surface, enhancing, therefore, the corrosion resistance of the studied mild steel.

The inhibition efficiency was evaluated from the measured Icorr values using the relationship. Inhibition efficiency = i corr o - i corr i / i corr o × 100 Where, i corr o and i corr i are values of corrosion current density in the absence and in the presence of inhibitor, respectively.

The results indicate a good agreement between the values of corrosion inhibition efficiency from weight loss, potentiodynamic polarization and impedance measurements. It is concluded that the corrosion rate depends on the chemical nature of the electrolyte rather than the applied technique (Ostovari et al., 2009). The differences are sometime as high as >2–3% but the order of magnitude is the same for all the methods.

3.5

3.5 Adsorption isotherm

The establishment of isotherms that describe the adsorptive behavior of a corrosion inhibitor is an important part of its study, as they can provide important clues to the nature of metal–inhibitor interaction. Several adsorption isotherms were assessed. The simplest, is the Langmuir isotherm, based on the assumption that all adsorption sites are equivalent and that particle binding occurs independently from nearby sites being occupied or not. Interactions between adsorbed species complicate the problem by making the energy of adsorption a function of surface coverage. One of the isotherms that include this possibility is the Temkin isotherm. The Temkin adsorption isotherm was found to provide the best description of the adsorption behavior of the Mannich bases: this has the following surface coverage, θ, and the bulk concentration, C, relationship: f (θ,x) exp(−2a θ) = KC. Where f (θ,x) is the configurational factor, which depends on the physical model and the assumptions underlying the derivation of the isotherm, θ is the degree of surface coverage, x is the size ratio, “a” is the molecular interaction parameter, C is the inhibitor concentration in the electrolyte and K is the equilibrium constant for adsorption. (See Fig. 7).

Impedance diagram for mild steel in 1 M H2SO4 in the absence and presence of various concentration of MPA.
Figure 7
Impedance diagram for mild steel in 1 M H2SO4 in the absence and presence of various concentration of MPA.

The equation expressing Temkin's adsorption isotherm is: Kads C = ef θ Where Kads is the adsorption equilibrium constant, θ is the degree of surface coverage, C is the molar concentration of inhibitor (mol L−1) and “f” is the molecular interaction constant. The plots of log C against θ displayed a straight line for the inhibitor tested. The linear plot with high correlation coefficient (0.9872 and 0.9500 for MMPA and MPA, respectively) clearly revealed that the surface adsorption process of inhibitors on the mild steel surface obeys the Temkin adsorption (Fig. 8).

Temkin adsorption isotherm plot for the adsorption of MMPA and MPA on mild steel in 1.0 M H2SO4 solution.
Figure 8
Temkin adsorption isotherm plot for the adsorption of MMPA and MPA on mild steel in 1.0 M H2SO4 solution.

3.6

3.6 Standard free energy of adsorption

The negative values of Δ G ads 0 ensure the spontaneity of the adsorption process and stability of the adsorbed layer on the mild steel surface (Keles et al., 2008: Singh and Quraishi, 2010). It is usually accepted that the value of Δ G ads 0 around −20 kJ mol−1 or lower indicates the electrostatic interaction between charged metal surface and charged organic molecules in the bulk of the solution while those around −40 kJ mol−1 or higher involve charge sharing or charge transfer between the metal surface and organic molecules (Moretti et al., 1996). In the present study, the Δ G ads o values obtained for the Mannich bases on mild steel in 1.0 M H2SO4 solution are ranging between −28 and −27 kJ mol−1 for MMPA and −27 to −25 kJ mol−1 for MPA, which are higher than −20 kJ mol−1 and lower than −40 kJ mol−1 (Table 6); this indicates that the adsorption is neither typical physisorption nor typical chemisorption but it is a complex mixed type that is the adsorption of inhibitor molecule on the mild steel surface in the present study involves both physisorption and chemisorption and chemisorption is the predominant mode of adsorption.

Table 6 Standard free energy of adsorption and adsorption equilibrium constant of 0.01 M concentrations of inhibitors in 1.0 M H2SO4 on mild steel at different temperatures.
Name of the inhibitor Temperature (K) Kads (M−1) Δ G ads o (kJ mol−1)
MMPA 303 1594.92 −28.70
308 1364.13 −28.77
313 1167.43 −28.83
318 725.62 −28.03
323 524.61 −27.60
328 415.99 −27.40
333 344.44 −27.29
MPA 303 1052.07 −27.65
308 984.60 −27.93
313 849.67 −28.00
318 597.84 −27.52
323 366.20 −26.64
328 223.62 −25.71
333 193.51 −25.69

3.7

3.7 SEM analysis

SEM photographs of the metal sample in the absence and presence of MMPA and MPA are shown in Figs. 9, 9a and 9b, respectively. The inhibited metal surface is smoother than the uninhibited surface indicating a protective layer of adsorbed inhibitor preventing acid attack.

SEM image of mild steel in 1 M H2SO4 solution.
Figure 9
SEM image of mild steel in 1 M H2SO4 solution.
SEM image of surface of mild steel after immersion for 2 h in 1 M H2SO4 in the presence of 0.01 M MMPA.
Figure 9a
SEM image of surface of mild steel after immersion for 2 h in 1 M H2SO4 in the presence of 0.01 M MMPA.
SEM image of surface of mild steel after immersion for 2 h in 1 M H2SO4 in the presence of 0.01 M MPA.
Figure 9b
SEM image of surface of mild steel after immersion for 2 h in 1 M H2SO4 in the presence of 0.01 M MPA.

3.8

3.8 FT-IR spectroscopy

FT-IR spectra were recorded to understand the interaction of inhibitor molecules with the metal surface (Tamil Selvi et al., 2003). Figs. 10 and 11 show the IR spectra of pure MMPA and MPA and Figs. 10a and 11a represent the spectra of the scraped samples obtained from the metal surfaces after corrosion experiments in 1.0 M H2SO4, respectively.

FT-IR spectrum of pure sample of MMPA.
Figure 10
FT-IR spectrum of pure sample of MMPA.
FT-IR spectrum of pure MPA.
Figure 11
FT-IR spectrum of pure MPA.
FT-IR spectrum of scraped sample of MMPA obtained from the metal surface after corrosion experiment in 1 M H2SO4.
Figure 10a
FT-IR spectrum of scraped sample of MMPA obtained from the metal surface after corrosion experiment in 1 M H2SO4.
FT-IR spectrum of scraped sample of MPA obtained from the metal surface after corrosion experiment in 1 M H2SO4.
Figure 11a
FT-IR spectrum of scraped sample of MPA obtained from the metal surface after corrosion experiment in 1 M H2SO4.

In pure MMPA, the IR bands observed at 3399.18, 1610.55 and 1109.57 cm−1 have been assigned to –NH, amide –C⚌O and –C–N–C of morpholine group, respectively (Table 7) (Geary, 1971; Raman et al., 2004).

Table 7 FT-IR spectral data for pure and scraped sample of MMPA and MPA.
Name of the inhibitor Function groups Original frequency cm−1 Shifted frequency cm−1
MMPA –NH 3399.18 3394.18
–CO– 1610.55 1620.50
–C–N-C– 1109.57 1103.66
MPA –NH 3429.84 3405.52
–CO– 1713.12 1706.89
–C–N–C– 1112.18 1113.30

In the scraped sample obtained from the mild steel surface after corrosion experiments in 1.0 M H2SO4, the IR bands that appeared at 3394.18, 1620.50 and 1103.66 cm−1 have been assigned to –NH, amide –C⚌O and –C–N–C of morpholine group respectively (Table 7). All the bands displayed substantial shifts with fairly low intensity indicating coordination through the oxygen of amide moiety and nitrogen of morpholine ring (Kerridge, 1988; Raman et al., 2004).

In pure MPA, the IR bands observed at 3429.84, 1713.12 and 1112.18 cm−1 have been assigned to –NH, amide –C⚌O and –C–N–C of morpholine group, respectively (Table 7) (Geary, 1971; Raman et al., 2004).

Similarly, in the scraped sample obtained from the mild steel surface after corrosion experiments in 1.0 M H2SO4, the IR bands that appeared at 3405.52, 1706.89 and 1113.30 cm−1 have been assigned to –NH, amide –C⚌O and –C–N–C of the morpholine group respectively (Table 7). All the bands displayed substantial shifts with fairly low intensity indicating coordination through the oxygen of amide moiety and nitrogen of morpholine ring (Kerridge, 1988; Raman et al., 2004).

3.9

3.9 Mechanism of corrosion inhibition

From the results obtained from different electrochemical and weight loss measurements, it was concluded that both the Mannich bases MMPA and MPA inhibit the corrosion of mild steel in 1.0 M H2SO4 by adsorption at mild steel/solution interface.

It is a general assumption that the adsorption of organic inhibitors at the metal surface interface is the first step in the mechanism of the inhibitor action. Organic molecules may be adsorbed on the metal surface in four types, namely

  1. Electrostatic interaction between the charged molecules and the charged metal,

  2. Interaction of unshared electron pairs in the molecule with the metal,

  3. Interaction of p-electrons with the metal and/or

  4. A combination of types (a–c) (Schweinsberg et al., 1988; Shorky et al., 1998; Singh and Quraishi, 2010).

In general, both the Mannich bases, MMPA and MPA contain one amide linkage, one aromatic ring and a residue of morpholine ring.

The inhibition of active dissolution of the metal is due to the adsorption of the inhibitor molecules on the metal surface forming a protective film. The inhibitor molecules can be adsorbed onto the metal surface through electron transfer from the adsorbed species to the vacant electron orbital of low energy in the metal to form a co-ordinate type link.

The inhibition efficiency depends on many factors (Noor, 2005; Bentiss et al., 2009) including the number of adsorption centers, mode of interactions with metal surface, molecular size and structure.

It is well known that iron has co-ordinate affinity toward nitrogen, sulfur and oxygen-bearing ligands (Donald, 1990; Snyder et al., 1989; Elayyachy et al., 2006; Lece et al., 2008). Hence, adsorption on iron can be attributed to co-ordination through amide linkage, hetero atom (N and O) and π-electrons of aromatic ring.

If one considers the structures of investigated compounds (Fig. 1) several potential sources of inhibitor–metal interaction can be identified. In both the Mannich bases, there are the unshared electron pairs on N and O, capable of forming a co-ordination σ-bond with iron (Shriver et al., 1994). Further, the double bonds in the molecule allow back donation of metal d-electrons to the π∗-orbital and this type of interaction cannot occur with amines.

The better inhibition performance of MMPA than MPA is due to the presence of p-OCH3 group in its structure, because of electro donating mesomeric effect, p-OCH3 group increases the electron density of aromatic ring and makes the π-electrons more available to interact with mild steel surface. Thus, MMPA is more effectively adsorbed. In the above facts, the investigated Mannich bases follow type (d) mechanism. Thus, the inhibition efficiency of the studied compounds follows the order MPA < MMPA. The results of weight loss, EIS and potentiodynamic polarization follow the same trend.

4

4 Conclusions

  1. Two Mannich bases namely, MMPA and MPA act as good inhibitors for the corrosion of mild steel in 1.0 M H2SO4 solution.

  2. Inhibition efficiency increases with increase in the concentration of the studied inhibitor but decreases with rise in temperature.

  3. Both MMPA and MPA act as mixed-type, but predominantly cathodic inhibitors in 1.0 M H2SO4.

  4. EIS spectra exhibit one capacitive loop, and the presence of each inhibitor enhances Rt values while reduces Cdl values.

  5. The inhibition efficiency of the studied compounds follows the order MPA < MMPA. The results of weight loss, EIS and potentiodynamic polarization follow the same trend.

  6. Activation energy (Ea) value is higher for the inhibited acids than uninhibited acids showing the temperature dependence of inhibition efficiency. The increase in the apparent activation energy may be interpreted as physical adsorption that occurs in the first stage.

  7. The negative values of Δ G ads 0 indicate that the spontaneous and strong adsorption of Mannich bases on the surface of mild steel in 1.0 M H2SO4.-

  8. In the present investigation the calculated values of Δ G ads 0 are greater than −20 kJ mol−1 but less than −40 kJ mol−1, indicating that the adsorption of mechanism of both the inhibitors in 1.0 M H2SO4 solutions at the studied temperatures may be a combination of both physisorption and chemisorption (comprehensive adsorption).

  9. In the present investigation, Δ H ads 0 values are larger than the common physical adsorption heat, but smaller than the common chemical adsorption heat, once again emphasizing that both physical and chemical adsorption is involved.

  10. The protective film formation of the Mannich bases in 1.0 M H2SO4 via the chemical bond formation in between the mild steel and the inhibitors is confirmed by FT-IR spectroscopic technique.

  11. Surface analysis using scanning electron microscope (SEM) shows a significant morphological improvement on the mild steel surface with addition of inhibitors.

Acknowledgments

The authors thank the Director of ICP centre, CECRI, Karaikudi, for the kind permission for providing the facilities of electrochemical studies and Prof. Dr. V. Muthupandi, MME, NIT, Trichy, for providing the SEM facilities and Principal, Jamal Mohamed College (Autonomous), Trichy 20, for providing necessary facilities and encouragement.

References

  1. , , , , . Corros. Sci.. 2010;52:1684.
  2. , , , . Corros. Sci.. 1984;24:509.
  3. , , , , , , , , . Corros. Sci.. 2009;51:1628.
  4. , , . Electrochim. Acta.. 1962;7:293.
  5. , , , . Mater. Chem. Phys.. 2000;65:288.
  6. , , , . J. Sci. Ind. Res.. 2007;66:835.
  7. , . Chem. Rev.. 1990;90:585.
  8. , , , . Corros. Sci.. 2006;48:2470.
  9. , . Int. J. Electrochem. Sci.. 2008;3:1149.
  10. , , , . Mater. Chem. Phys.. 2011;125:26.
  11. , . Coord. Chem. Rev.. 1971;7:81.
  12. , , , . Mater. Chem. Phys.. 2004;86:59.
  13. , , . Bull. Electrochem.. 2007;23:237.
  14. , , . Arabian. J. Chem 2011
    [CrossRef]
  15. , . Bull. Electrochem.. 2005;21:305.
  16. , , , , . Colloids. Surf. A. Physicochem. Eng. Aspects.. 2008;320:138.
  17. , . Chem. Soc. Rev.. 1988;7:181.
  18. , , , , , . J. Appl. Electrochem.. 2008;38:1391.
  19. , , , . Corros. Sci.. 2008;50:1460.
  20. , , , . Electrochim. Acta.. 1993;38:1989.
  21. , , , , , . J. Appl. Electrochem.. 2008;38:289.
  22. , , , , . Electrochim. Acta.. 1996;41:1971.
  23. , . Corros. Sci.. 2005;47:33.
  24. , , , , , . Corros. Sci.. 2009;51:1935.
  25. , , , , . Port. Electrochim. Acta.. 2010;28:189.
  26. , , , , . Corros. Sci.. 2011;53:679.
  27. , , , . Corros. Sci.. 2010;52:1472.
  28. , , , . Indian J. Chem.. 2004;43A:2357.
  29. , , . Bull. Mater. Sci.. 2008;31:699.
  30. , , . Indian J. Chem. Technol.. 2009;16:162.
  31. Radovici, O., 1965. Proceedings of the 2nd European Symposium on Corrosion inhibitors, Ferrara, 178.
  32. , , . J. Indian Chem. Soc.. 2010;87:189.
  33. Saukaitis, A. J., Gardner, G. S., 1970. US Patent, Derivatives of rosin amines, 2, 758, 970.
  34. , , , , . Corros. Sci.. 1988;28:33.
  35. , , , . Electrochim. Acta.. 2010;55:3657.
  36. , , , , , , . Corros. Sci.. 1998;40:2173.
  37. , , , . Inorganic Chemistry (2nd ed.). Oxford: Oxford University Press; . p 239
  38. , , . J. Appl. Electrochem.. 2010;40:1293.
  39. , , , , . J. Am. Chem. Soc.. 1989;111:5214.
  40. , , , . J. Appl. Electrochem.. 2003;33:1175.
  41. , , , , , , . Corros. Sci.. 2011;53:147.
Show Sections