You are on page 1of 11

Corrosion Science 75 (2013) 123–133

Contents lists available at SciVerse ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Electrochemical and XPS studies of the corrosion inhibition


of carbon steel in hydrochloric acid pickling solutions
by 3,5-bis(2-thienylmethyl)-4-amino-1,2,4-triazole
M. Tourabi a, K. Nohair a, M. Traisnel b, C. Jama b, F. Bentiss a,b,⇑
a
Laboratoire de Catalyse et de Corrosion des Matériaux (LCCM), Faculté des Sciences, Université Chouaib Doukkali, BP 20, M-24000 El Jadida, Morocco
b
UMET-ISP, CNRS UMR 8207, ENSCL, Université Lille Nord de France, BP 90108, F-59652 Villeneuve d’Ascq Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: Corrosion inhibition of carbon steel in normal hydrochloric acid solution at 30 °C by new 4-amino-1,2,4-
Received 25 March 2013 triazole derivative, namely 3,5-bis(2-thienylmethyl)-4-amino-1,2,4-triazole (2-TMAT) has been studied
Accepted 28 May 2013 by electrochemical impedance spectroscopy (EIS) and polarization techniques. The experimental results
Available online 5 June 2013
have showed that this organic compound revealed a good corrosion inhibition and that the inhibition effi-
ciency is increased with the inhibitor concentration. Two time constants determined by the charge-trans-
Keywords: fer and the adsorption of the inhibitor, respectively, can be readily outlined. Potentiodynamic
A. 4-Amino-1,2,4-triazole
polarization showed that 2-TMAT is a mixed type of inhibitor. The adsorption of 2-TMAT on the carbon
A. Carbon steel
A. HCl
steel surface in 1 M HCl solution obeyed to the Langmuir isotherm with a very high negative value of the


B. EIS standard Gibbs free energy of adsorption DGads (chemisorption). X-ray photoelectron spectroscopy (XPS)
C. Adsorption analyses were carried out to establish the mechanism of corrosion inhibition of carbon steel in 1 M HCl
medium by 3,5-bis(2-thienylmethyl)-4-amino-1,2,4-triazole (2-TMAT).
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction 3,5-bis(2-thienylmethyl)-4-amino-1,2,4-triazole (2-TMAT), is


tested for the first time as corrosion inhibitor on carbon steel in
Acid solutions are generally used for the removal of rust and 1 M HCl medium at 30 °C. The inhibitive activity of 2-TMAT is
scale in industrial processes. Hydrochloric acid is widely used in examined successively via electrochemical impedance spectros-
the pickling of steel and different alloys. Inhibitors are generally copy (EIS), polarization, isotherm calculations, scanning electron
used in these processes to control the metal dissolutions. Most microscopy (SEM) and X-ray photoelectron spectroscopy (XPS)
well-known acid corrosion inhibitors are organic compounds con- techniques.
taining nitrogen, sulfur, and oxygen atoms; nitrogen-containing
organic compounds are known to be efficient corrosion inhibitors 2. Experimental details
in hydrochloric acid solutions, while sulfur-containing compounds
are sometimes preferred for sulfuric acid solutions [1,2]. Among 2.1. Materials
them, N-heterocyclic compounds are considered to be the most
effective corrosion inhibitors for steel in acid media [3–5]. The tested inhibitor, namely 3,5-bis(2-thienylmethyl)-4-amino-
Previously, some 1,2,4-triazole derivatives have been reported 1,2,4-triazole (2-TMAT), was synthesised according to a previously
as potential corrosion inhibitors [6–33]. Our previous investiga- described experimental procedure [34,35]. The molecular structure
tions showed that the inhibitive effectiveness of these compounds of 2-TMAT is shown in Fig. 1. The structure of the 2-TMAT was con-
is affected mostly by the molecular structure, the position of the firmed by 1H and 13C NMR, mass spectroscopy and elemental
substituents, the temperature and the acidic media [11,16– analysis:
18,22,23]. In this work we continue our investigations on 1,2,4- Yield 87%; mp 144 °C; 1H NMR (dimethyl-d6 sulfoxide;
aminotriazole derivatives as inhibitors of steel corrosion in acid 300 MHz): d (ppm) 4.34 (s, 4H); 6.89 (m, 4H); 7.16 (d, 2H); 7.26
media. The investigated 4-amino-1,2,4-triazole derivative, namely (s, 2H); 13C NMR (dimethyl-d6 sulfoxide; 75 MHz): d (ppm)
25.68, 125.27, 126.61, 127.56, 137.74, 153.96; MALDI-TOFMS: m/
⇑ Corresponding author at: UMET-ISP, CNRS UMR 8207, ENSCL, Université Lille z 277 (M+1). Anal. Calcd. for C12H12N4S2: C, 52.17; H, 4.35; N,
Nord de France, BP 90108, F-59652 Villeneuve d’Ascq Cedex, France. Tel.: +33 320 20.29; S, 23.19. Found: C, 52.29; H, 4.22; N, 20.41; S, 23.05.
337 746; fax: +33 320 436 814. The material used in this study is a carbon steel (Euronorm:
E-mail address: fbentiss@enscl.fr (F. Bentiss). C35E carbon steel and US specification: SAE 1035) with a chemical

0010-938X/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2013.05.023
124 M. Tourabi et al. / Corrosion Science 75 (2013) 123–133

Instruments SI 1287 potentiostat monitored by a personal com-


puter via a GPIB Interface and CorrWare 2.80 software were used
NH2 to run the tests and to collect the experimental data. Before each
S Tafel experiment, the carbon steel electrode was pre-polarized at
N 800 mVSCE for 10 min and thereafter the WE was allowed to cor-
rode freely and its open circuit potential (OCP) was recorded as a
function of time up to 1 h, the time necessary to reach a quasi-sta-
N S tionary value for the open-circuit potential. This steady-state OCP
N
corresponds to the corrosion potential (Ecorr) of the WE. For LPR
Fig. 1. Chemical structure of 3,5-bis(2-thienylmethyl)-4-amino-1,2,4-triazole (2- measurements, the WE was polarized only in the range ±20 mV
TMAT). vs. Ecorr at a scan rate of 0.125 mV s1. After studying the open cir-
cuit potential and LPR tests, the potentiodynamic Tafel measure-
composition (in wt%) of 0.370% C, 0.230% Si, 0.680% Mn, 0.016% S, ments were scanned from cathodic to the anodic direction,
0.077% Cr, 0.011% Ti, 0.059% Ni, 0.009% Co, 0.160% Cu and the E = Ecorr ± 200 mV, with a scan rate of 0.5 mV s1. The Tafel and
remainder iron (Fe). The steel samples were pre-treated prior to LRP data were analyzed and fitted using the polarization CorrView
the experiments by grinding with emery paper SiC (120, 600 and 2.80 software (Scribner Associates, Inc.).
1200); rinsed with distilled water, degreased in acetone in an
ultrasonic bath immersion for 5 min, washed again with bidistilled
2.3. Scanning electron microscopy (SEM)
water and then dried at room temperature before use. The acid
solutions (1 M HCl) were prepared by dilution of an analytical re-
The morphologies of the uninhibited and inhibited carbon steel
agent grade 37% HCl with doubly distilled water. The solubility
surface were analyzed by scanning electron microscope (JEOL
of the 2-TMAT is about 104 M in 1 M HCl. The concentration range
5300). The energy of the acceleration beam employed was 6 kV.
of 2-TMAT employed was 5  106 M to 104 M in order to com-
pare their corrosion inhibition properties with those of similar het-
erocyclic compounds. 2.4. X-ray photoelectron spectroscopy (XPS)

X-ray photoelectron spectroscopy (XPS) spectra were recorded


2.2. Corrosion tests
by a XPS KRATOS, AXIS UltraDLD spectrometer with the monochro-
matized Al Ka X-ray source (hm = 1486.6 eV) and an X-ray beam
Electrochemical impedance spectroscopy (EIS), linear polariza-
of around 1 mm. The analyzer was operated in constant pass
tion resistance (LPR) and potentiodynamic polarization are three
energy of 40 eV using an analysis area of approximately
classical electrochemical techniques which were used to deter-
700 lm  300 lm. Charge compensation was applied to compen-
mine the corrosion inhibitor characteristics of 2-TMAT in 1 M HCl.
sate for the charging effects that occurred during the analysis.
Ac impedance tests were performed in a polymethyl methacry-
The C 1s (285.0 eV) binding energy (BE) was used as internal refer-
late (PMMA) cell with a capacity of 1000 ml at 30 ± 1 °C, using a
ence. The spectrometer BE scale was initially calibrated against the
thermostat. A saturated calomel electrode (SCE) was used as the
Ag 3d5/2 (368.2 eV) level. Pressure was in the 1010 torr range
reference; while a platinum leaf was used as a counter electrode.
during the experiments. XPS spectra were deconvoluted using a
All potentials are reported vs. SCE. The working electrode was pre-
non-linear least squares algorithm with a Shirley base line and a
pared from a square sheet of carbon steel such that the area ex-
Gaussian–Lorentzian combination. CasaXPS software was used
posed to solution was 7.55 cm2. All impedance spectra were
for all data processing. XPS analyses were practiced on pure 2-
recorded at the steady state of open circuit potential, i.e., at the
TMAT and on protected steel surface with 2-TMAT. The disk carbon
corrosion potential, after 24 h of the exposure of the working elec-
steel (1 cm2) was pre-treated by the same procedure as for the
trode to the solution (no deaeration and no stirring).
gravimetric test. After 24 h of immersion at 30 °C, the carbon steel
Ac impedance measurements were performed using a potentio-
sheet was rinsed with acetone and ultra pure water.
stat Solartron SI 1287 and a Solartron 1255B frequency response
analyzer and ZPlot 2.80 software was used to run the tests and
to collect the experimental data. The response of the electrochem- 3. Results and discussion
ical system to ac excitation with a frequency ranging from 105 Hz
to 102 Hz and peak to peak amplitude of 10 mV was measured 3.1. Corrosion inhibition evaluation
with data density of 10 points per decade. All EIS diagrams were
recorded at the open circuit potential, i.e., at the corrosion poten- 3.1.1. Ac impedance study
tial Ecorr. One EIS spectrum is recorded usually within 10 min. The electrochemical impedance spectroscopy (EIS) is a good
The impedance data were analyzed and fitted with the simulation technique in investigating corrosion inhibition processes. It pro-
ZView 2.80, equivalent circuit software. vides information on both the resistive and capacitive behavior
Linear polarization resistance (LPR) and potentiodynamic polar- at interface and makes it possible to evaluate the performance of
ization measurements were performed in a conventional three- the tested compounds as possible inhibitors against metals’ corro-
electrode cylindrical Pyrex glass cell. The temperature is thermo- sion. Figs. 2 and 3 show a typical set of impedance diagrams for
statically controlled. Pure carbon steel specimen was used as the carbon steel in 1 M HCl without and with various concentrations
working electrode, a platinum electrode as the counter electrode of 2-TMAT. The Nyquist diagram in the absence of inhibitor
and a saturated calomel electrode (SCE) as the reference electrode. (Fig. 2a) shows only one depressed capacitive loop at the higher
The working electrode (WE) in the form of disk cut from steel has a frequency range (HF) with one capacitive time constant in the
geometric area of 1 cm2 and is embedded in polytetrafluoroethyl- Bode-phase plot (Fig. 2b). In this case, the standard Randles’ circuit
ene (PTFE). A fine Luggin capillary was placed close to the working model of Fig. 4a, which has been previously used [36], fits well our
electrode to minimize IR drop. All test solutions were de-aerated in experimental results. In this equivalent circuit, Rs is the solution
the cell by using pure nitrogen for 10 min prior to the experiment. resistance, Rct present the charge transfer resistance whose value
During each experiment, the test solution was mixed with a mag- is a measure of electron transfer across the surface and CPE is
netic stirrer and the gas bubbling was maintained. Solartron the constant phase element. The constant phase element, CPE, is
M. Tourabi et al. / Corrosion Science 75 (2013) 123–133 125

-40 -300
HCl 1 M a 10-4 M a
510-5 M
10-5 M
510-6 M

-30 -200

Zi (Ωcm )
2
50.1 Hz 25.1 Hz

100 Hz 10 Hz
Z i (Ωcm )
2

7.9 Hz 5 Hz
-20 199.5 Hz
-100 12.6 Hz
2.5 Hz
10 Hz 1 Hz
5 Hz 0.5 Hz 0.1 Hz
25.1 Hz 25.1 Hz
0.01Hz
2.5 Hz 2 Hz
-10 50.1 Hz 0
1 Hz
0 100 200 300 400 500
100 Hz
0.1 Hz Zr (Ωcm2)

0 103 -75
0 10 20 30 40 50 10-4 M
510-5 M
b
2
Z r (Ωcm ) 10-5 M
510-6 M
-55
10
2 -50
HCl 1 M b

Phase angle (Degree)


102

-40 -35

|Z| (Ωcm2)
Phase angle (Degree)

-30 -15
101
|Z| (Ωcm2)

10
1 -20
5

-10
100 25
10-2 10-1 100 101 102 103 104 105

0 Frequency (Hz)

Fig. 3. (a) Nyquist and (b) Bode diagrams for carbon steel in 1 M HCl containing
different concentrations of 2-TMAT after 24 h of immersion period at 30 °C. Key: h
100 10 – 5  106 M, s – 105 M, r – 5  105 M, } – 104 M.
-2 -1 0 1 2 3 4 5
10 10 10 10 10 10 10 10
Frequency (Hz)
a depressed semicircle at the high frequency part of the spectrum
Fig. 2. (a) Nyquist and (b) Bode diagrams for carbon steel in 1 M HCl after 24 h of
immersion period at 30 °C.
was observed and a second time constant appeared at low fre-
quency region (Fig. 3). The increase in size of the semicircle with
introduced in the circuit instead of a pure double layer capacitor in 2-TMAT concentration means that the inhibitor effect increases
order to take into account the electrode surface heterogeneity as well. The first capacitive loop (the larger) appearing at high fre-
resulting from surface roughness, impurities, dislocations, grain quency region can be attributed to the charge transfer, the second
boundaries, adsorption of inhibitors, formation of porous layers one (the smaller) at low frequencies can be related to the adsorp-
[37,38] and therefore to give a more accurate fit [39]. The imped- tion of inhibitor molecules on the metal surface and/or all other
ance of the CPE is expressed as: accumulated kinds at the metal/solution interface (inhibitor mole-
n
cules, corrosion products, etc.) [43]. In this case, the structural
Z CPE ¼ A1 ðixÞ ð1Þ model of Fig. 4a is inadequate (see Fig. 5, representative example)
and a two-time-constant model should be used to describe the
where A is the magnitude of CPE (in X1 sn cm2), x is the sine
electrochemical impedance spectra in the presence of 2-TMAT, as
wave modulation angular frequency (in rad s1), i2 = 1 is the imag-
presented in Fig. 4b, that permits the determination of the adsorp-
inary number, and n is an empirical exponent (0 6 n 6 1) which
tion parameters Ra and CPEa (Aa and na, respectively). In Fig. 4b, Rs
measures the deviation from the ideal capacitive behavior [40,41].
represent the solution resistance, Rct is the charge transfer resis-
The corresponding fitting results are listed in Table 1. In this table
tance and CPEd represent the capacitance of the high frequency
are also shown the calculated double layer capacitance (Cdl) derived
semicircle that can be attributed to the charge transfer process.
from the CPE parameters according to the following equation [42]:
Ra represents the resistance of the adsorbed inhibitor and CPEa is
C dl ¼ ðAd R1n d 1=nd
Þ ð2Þ the capacitance of the inhibitor film due to the inhibitor’s adsorp-
ct
tion on the steel surface. The time constant of the adsorption pro-
and the relaxation time constant (sd) of charge-transfer process cess is sa = Ca Ra, the adsorption capacitance being again replaced
using the following equation [42]: by a CPE, and Ca is calculated using Eq. (2). The experimental data
were very well fitted according to the proposed equivalent circuit
sd ¼ C dl Rct ð3Þ
and as an example; the calculated curves are presented in Fig. 6.
The addition of 2-TMAT to the aggressive solution leads to a This model is valid for all concentrations of 2-TMAT and it is well
change of the impedance diagrams in both shape and size, in which representative of the phenomena which may occur in the
126 M. Tourabi et al. / Corrosion Science 75 (2013) 123–133

Rs CPEd

Rct CPEa

Ra

(a) (b)
Fig. 4. Electrical equivalent circuit used for modeling the interface carbon steel/1 M HCl solution (a) without and (b) with 2-TMAT.

Table 1
Impedance parameters and inhibition efficiency values for carbon steel after 24 h immersion period in 1 M HCl containing different concentrations of 2-TMAT at 30 °C.

Conc. Rs (X cm2) Rct (X cm2) 106 Ad nd Cdl sd (s) Ra (X cm2) 104 Aa na Ca sa (s) Rp E
(M) (X1 sn cm2) (lF cm2) (X1 sn cm2) (mF cm2) (X cm2) (%)
Blank 5.70 ± 0.02 37.05 ± 0.13 844.4 ± 0.09 0.879 ± 0.001 524.1 0.0194 – – – – – 37.05 –
5  106 6.36 ± 0.02 155.5 ± 0.9 45.20 ± 0.09 0.906 ± 0.003 27.0 0.0042 97.4 ± 1.6 19.1 ± 1.1 0.64 ± 0.02 0.75 0.073 252.9 85.3
1  105 6.49 ± 0.02 160.7 ± 1.0 32.61 ± 0.09 0.927 ± 0.003 21.6 0.0035 130.9 ± 1.2 12.9 ± 0.8 0.61 ± 0.02 0.42 0.054 291.6 87.3
5  105 6.40 ± 0.02 208.0 ± 2.0 41.79 ± 0.08 0.916 ± 0.003 26.6 0.0055 160.0 ± 1.6 15.7 ± 1.2 0.55 ± 0.02 0.51 0.081 368.0 89.7
1  104 6.38 ± 0.02 279.4 ± 5.6 41.20 ± 0.08 0.913 ± 0.003 26.9 0.0075 179.8 ± 4.6 15.5 ± 1.1 0.61 ± 0.03 0.67 0.121 459.2 91.9

investigated steel/2-TMAT/HCl system, both in the HF and in the LF with inhibitor concentration). A large Rct is associated with a
parts of the spectra. The main fitted parameters are given in Ta- slower corroding system [38,44]. Furthermore, the decrease in
ble 1. Estimates of the margins of error calculated for the parame- the Cdl with increase in 2-TMAT concentrations may be attributed
ters are also presented in this Table. The data are found to be to the formation of a protective layer on the carbon steel surface
sufficiently well fitted within the limits of experimental error [45]. The increase of the nd value after addition of 2-TMAT in the
and reproducibility of data. Inspection of the fitting results (Ta- corrosive solution can corroborate this assumption. Indeed, the
ble 1), shows clearly that Rct and Cdl values have opposite trend
at the whole concentration range (Rct increases and Cdl decreases
-300

-300
Experimental curve
FitResult
a
Experimental curve a
FitResult
-200
Zi (Ωcm )
2

-200
Zi (Ωcm )

25.1 Hz 10 Hz
2

50.1 Hz 5 Hz
25.1 Hz 10 Hz -100
2.5 Hz
5 Hz 100 Hz
50.1 Hz 1 Hz
-100 2.5 Hz 0.1 Hz
100 Hz
1 Hz 0.01 Hz
0.1 Hz
0
0.01 Hz 0 100 200 300 400 500
0 Zr (Ωcm2)
0 100 200 300 400 500

Zr (Ωcm2) 3
10 -75
Experimental curve
3
FitResult b
10 -75
Experimental curve
FitResult
b -55

-55
2
10
Phase angle (Degree)

2
10 -35
Phase angle (Degree)

|Z| (Ωcm2)

-35
|Z| (Ωcm2)

-15
-15 10
1
1
10

5
5

0
0
25 10 25
10 -2 -1 0 1 2 3 4 5
10 -2
10 -1
10 0
10 1
10 2
10 3
10 4
10 5 10 10 10 10 10 10 10 10
Frequency(Hz) Frequency (Hz)
5
Fig. 5. (a) EIS Nyquist and (b) Bode diagrams for carbon steel/1 M HCl + 5  10 M Fig. 6. (a) EIS Nyquist and (b) Bode diagrams for carbon steel/1 M HCl + 5  105 M
of 2-TMAT interface: (  ) experimental; (–) fitted data using EEC in Fig. 4a. of 2-TMAT interface: (  ) experimental; (–) fitted data using EEC in Fig. 4b.
M. Tourabi et al. / Corrosion Science 75 (2013) 123–133 127

Table 2
Some 1,2,4-triazole derivatives investigated as corrosion inhibitors by other authors (the values of inhibition efficiency were obtained by EIS method at 0.1 mM in 1 M HCl).

Inhibitor Inhibition efficiency (%) Reference


3,5-Bis(2-thienylmethyl)-4-amino-1,2,4-triazole (2-TMAT) 91.9 This work
3,5-Bis(4-methoxyphenyl)-4-amino-1,2,4-triazole (4-MAT) 94.8 [24]
4-Amino-5-phenyl-4H-1,2,4-trizole-3-thiol (APTT) 90 [47]
3,5-Di(m-tolyl)-4-amino-1,2,4-triazole (m-DTAT) 86.9 [17]
3,5-Di(m-tolyl)-4H-1,2,4-triazole (m-DTHT) 63.8 [17]
3,5-Diphenyl-4-amino-1,2,4-triazole (DAT) 92.8 [10]
3,5-Dibenzyl-4-amino-1,2,4-triazoles (BAT) 91.1 [10]
3,5-Bis(2-thienyl)-4-amino-1,2,4-triazole (2-TAT) 97.5 [8]
3,5-Bis(4-methylthiophenyl)-4H-1,2,4-triazole (4-MTHT) 99.1 [15]
3,5-Diphenyl-4H-1,2,4-triazole (DHT) 84.1 [22]
3,5-Bis(4-pyridyl)-4H-1,2,4-triazole (4-PHT) 93.9 [22]
4-Amino-5-(2-hydroxy)phenyl-4H-1,2,4,-triazole-3-thiol (AHPTT) 70.5 [48]
4-Amino-5-styryl-4H-1,2,4,-triazole-3-thiol (ASTT) 76.7 [48]
4-Salicylideneamino-3-methyl-1,2,4-triazole-5-thione(SAMTT) 76 [49]
4-(2,4-Dihydroxybenzylideneamino)-3-methyl-1H-1,2,4-triazole-5(4H)-thione (DBAMTT) 79 [49]

value of nd grows as well (0.906–0.927), when compared to that 0


10
obtained in pure 1 M HCl (0.879). This can be attributed to a certain
decrease in the initial surface in homogeneity resulting from the 10-1
adsorption of 2-TMAT molecules on the most active adsorption
-2 a
centers at the carbon steel surface [38]. Also, the addition of 2- 10
TMAT to the corrosive solution increases slightly the time constant
-3
(sd) value with opposite trend of Cdl (Table 1). For example, when A cm2) 10 b c
the 2-TMAT concentration is increased to 104 M in HCl medium, d
10-4
-
I (A/

the interface (sd) parameter increases from 0.0194 to 0.075 s while


the Cdl value decreases from 524 to 26.9 lF cm2, signifying that -5 e a: Blank
10
the charge and discharge rates to the metal–solution interface is b: 5×10-6 M
c: 1×10-5 M
greatly decreased. This shows that there is agreement between 10-6 d: 5×10-5 M
the amount of charge that can be stored (i.e., capacitance) and e: 1×10-4 M
the discharge velocity in the interface (sd) [46]. In the case of 2- 10
-7

-0.70 -0.65 -0.60 -0.55 -0.50 -0.45 -0.40 -0.35 -0.30 -0.25 -0.20
TMAT addition, the resistance Ra has small values compared to
E (V/SCE)
Rct values (the total resistance Rp is dominated by Rct) and increases
significantly with 2-TMAT concentration. However, the Ca values Fig. 7. Polarization curves for carbon steel in 1 M HCl containing different
have opposite trend as that of Ra values at the whole concentration concentrations of 2-TMAT. Key: (a) Blank, (b) 5  106 M, (c) 105 M, (d)
range, while practically no change is observed in the value of na. 5  105 M, and (e) 104 M.
The values of na are lower than that of nd, probably connected with
energy dissipation within the adsorbed layer [42]. The time con-
their maximum value at 0.1 mM (Rp = 459.2 X cm2, gZ (%) = 91.9).
stants sa shows a marked tendency to increase with concentration
The comparison of the inhibitive performance of 2-TMAT for the
and its values are much greater than sd. This behavior can be re-
steel corrosion in 1 M HCl with other 1,2,4-triazole derivatives, pre-
lated to the adsorption process.
viously described in the same conditions, is given in Table 2. A sur-
In the absence of inhibitor molecules, the ac resistance polariza-
vey of these data shows that 2-TMAT can be considered as excellent
tion, Rp, consists of only charge transfer resistance (Rct), while in
organic inhibitor for the corrosion of carbon steel in 1 M HCl.
the presence of inhibitor, the sum of Rct and Ra are equivalent to
Rp. The related inhibition efficiency, gZ (%), is calculated from
Rp = (Rct + Ra) using the following equation: 3.1.2. Polarization studies
Potentiodynamic polarization measurements were carried out
RPðiÞ  Rp in order to gain knowledge concerning the kinetics of the cathodic
gZ ð%Þ ¼  100 ð4Þ
RPðiÞ and anodic reactions. Fig. 7 presents the results of the effect of
2-TMAT concentration on the cathodic and anodic polarization
where Rp and Rp(i) are the ac polarization resistance of carbon steel curves of carbon steel in 1 M HCl, respectively. It could be observed
electrode in the uninhibited and inhibited solutions, respectively. It that the anodic and cathodic reactions are affected by the addition
is obvious that the increase in inhibitor concentration enhances Rp, of 2-TMAT. Indeed, the addition of this 4-amino-1,2,4-triazole
and consequently improves the inhibition efficiency till reaching derivative to HCl solution reduces the anodic dissolution of steel

Table 3
Polarization parameters and the corresponding inhibition efficiency for the corrosion of carbon steel in 1 M HCl containing different concentrations of 2-TMAT at 30 °C.

Inhibitor conc. (M) Tafel data LPR data


Ecorr vs. SCE (mV) jcorr (lA cm2) ba (mV dec1) bc (mV dec1) gTafel (%) Rp (X cm2) gLRP (%)
Blank 451 717 110 108 – 30.9 –
5  106 469 140 65 117 80.5 150.4 79.5
1  105 475 71 68 111 90.1 208.8 85.2
5  105 466 50 57 105 93.0 243.9 87.3
1  104 476 40 68 113 94.4 398.6 92.2
128 M. Tourabi et al. / Corrosion Science 75 (2013) 123–133

and also retards the cathodic hydrogen evolution reaction. The that this studied inhibitor was first adsorbed onto the metal sur-
inhibiting effect of 2-TMAT may be related to its adsorption and face and impeded by merely blocking the reaction sites of the me-
formation of a barrier film on the electrode surface [50]. tal surface without affecting the anodic reaction mechanism [52].
Electrochemical corrosion kinetic parameters, i.e., corrosion The values of Rp increase and jorr values decrease considerably with
potential (Ecorr), corrosion current density (jcorr), cathodic and ano- the increase of the concentration of 2-TMAT, with a slight negative
dic Tafel slopes (bc, ba), obtained from the extrapolation of the shift in corrosion potential (maximum displacement  25 mV)
polarization curves are given in Table 3. The data were fitted using compared to that of uninhibited solution. From these data one
non-linear least square algorithm based on the Tafel-LEV (Leven- can conclude that 2-TMAT is of mixed type with predominantly
berg–Marquardt) method used along with CorrView 2.80 software. cathodic action.
Table 3 also included percentage inhibition efficiency, gTafel (%), The inhibition efficiency values calculated by LPR and potentio-
which was calculated from the following equation [50]: dynamic polarization curves are comparable (Table 3). The coinci-
dence between these two methods can suggest that no significant
jcorr  jcorrðiÞ surface changes occur during the polarization measurements at
gTafel ð%Þ ¼  100 ð5Þ
jcorr ±200 mV vs. Ecorr. The polarization data also confirm the ac imped-
ance results, i.e., 2-TMAT is a good inhibitor and its inhibition effi-
where jcorr and jcorr(i) are the corrosion current densities for steel ciency depends on the inhibitor concentration. The values obtained
electrode in the uninhibited and inhibited solutions, respectively. by means of the polarization experiments are somewhat lower
The LPR experiments were carried out in order to exclude the influ- than in the ac impedance tests but the trends are the same. These
ence of the surface changes which may occur during polarization at differences are probably due to a greater extent to the different
higher overpotentials in the case of potentiodynamic polarization exposure time in the corrosive solution.
method. The polarization resistance (Rp) obtained by LPR method
is used to calculate the corrosion current (jcorr) based on the
3.2. Adsorption isotherm and surface analysis
Stern–Geary kinetics equation [51]. In this case, the inhibition effi-
ciency, gLRP (%), is calculated from Rp using the following equation:
Scanning electron microscopy (SEM) micrographs (Fig. 8a–c) of
R  Rp carbon steel strips were recorded in order to see the changes oc-
gLRP ð%Þ ¼ PðiÞ  100 ð6Þ curred during corrosion process before and after immersion in
RPðiÞ
1 M HCl with and without corrosion inhibitor (2-TMAT). Parallel
where Rp and Rp(i) are the polarization resistance of carbon steel features on the polished steel surface before exposure to the corro-
electrode in the uninhibited and inhibited solutions, respectively. sive solution were observed, which are associated with abrading
The parallel cathodic Tafel curves in Fig. 7 suggest that the scratches (Fig. 8a). Fig. 8b and c show the steel surface after 24 h
hydrogen evolution is activation-controlled and the reduction of immersion in 1 M HCl without and with 2-TMAT. The resulting
mechanism is not affected by the presence of the inhibitor. The val- of the high resolution SEM micrograph (Fig. 8b) shows that the
ues of bc show a slight change with increasing inhibitor concentra- steel surface was strongly damaged in the absence of the 2-TMAT
tion, indicating the influence of the aminotriazole derivative on the with the increased number and depth of the pits. However, in pres-
kinetics of hydrogen evolution. This may probably be due to a dif- ence of 2-TMAT (Fig. 8c), the surface was remarkably improved and
fusion or barrier effect [51]. The values of the slopes of the anodic less pits and cracks were observed in comparison to the steel
Tafel lines, ba, also change with the addition of 2-TMAT suggesting surface in absence of inhibitor. This improvement in surface

Fig. 8. SEM micrographs of the carbon steel surface: (a) Metallic surface after being polished, (b) metallic surface after 24 h immersion in 1 M HCl and (c) metallic surface
after 24 h immersion in 1 M HCl with 1  104 M of 2-TMAT.
M. Tourabi et al. / Corrosion Science 75 (2013) 123–133 129

morphology is probably due to the formation of protective film of considered sufficient for adsorption equilibrium to be achieved
2-TMAT on steel surface which is responsible for the corrosion [55]. Several adsorption isotherms (Langmuir, Temkin, Frumkin,
inhibition. Indeed, 2-TMAT has a strong tendency to adhere to . . .) were assessed and the Langmuir adsorption isotherm was
the steel surface and can be regarded as good inhibitor for steel found to fit well with the experimental data obtained for 2-TMAT.
corrosion in normal hydrochloric medium. The high inhibitive per- The Langmuir isotherm is given by the following equation [56]:
formance suggests a strong bonding of the 2-TMAT derivative on
the metal surface due to presence of lone pairs from heteroatom C inh 1
¼ þ C inh ð7Þ
(nitrogen and sulfur) and p-orbitals, blocking the active sites and h K ads
therefore decreasing the corrosion rate.
In hydrochloric medium, the 4-amino-1,2,4-triazole molecules where h is the fractional surface coverage; Cinh is the inhibitor con-
(2-TMAT) give cationic species with respect to the pKa of the pro- centration; and Kads is the equilibrium constant of the adsorption
tonated function as shown in the following scheme: process. The plot of Cinh/h vs. Cinh, in Fig. 9, yields straight line with

S
NH2 S
+
NH 3
N
+ H+ N

N N
S N N
S
I II

It is assumed that Cl ions are first adsorbed on the metal sur- the regression coefficient (R2) very higher than 0.99, which suggests
face and the net positive charge on the steel surface enhances the that the experimental data are well described by Langmuir iso-
specific adsorption of chloride ions [15,24]. Due to the electrostatic therm and exhibit single-layer adsorption characteristic. This kind
attraction, the cationic forms of inhibitor (2-TMATH+) are adsorbed of isotherm involves the assumption of no interaction between
through electrostatic interactions (physical adsorption) between the adsorbed species on the electrode surface [36,57]. This behavior
positively charged nitrogen atoms (form II) and negatively charged can be explained by the fact that the 2-TMAT molecule possesses
carbon steel surface [53]. The adsorption of the 2-TMAT molecules non-planar structure that can make important steric effects, and
can also occur due to the formation of links between the d-orbital therefore a strong repulsive interaction between adsorbed 2-TMAT
of iron atoms, involving the displacement of water molecules from molecules in the adsorbed layer can occur (formation of single-
the metal surface, by the lone sp2 electron pairs present on the layer).
nitrogen and sulfur atoms and p-orbitals (chemisorption) [53,54]. From the intercepts of the straight lines Cinh/h – axis (Fig. 9), the
If it is assumed, that adsorption of the organic molecules or ions Kads can be calculated; Kads = 1.35  106 M1. Kads is related to the

on the metal surface is the cause for the inhibitor action in both standard Gibbs free energy of adsorption, DGads , according to the
acid media, then the surface coverage (h) can be estimated using following equation [58]:
from the inhibitor efficiency as h = g(%)/100 [53]. An attempt was
 
made to find a suitable adsorption isotherm, which can describe 1 DGoads
the concentration dependence of the inhibitor efficiency. Adsorp- K ads ¼ exp ð8Þ
55:55 RT
tion isotherms were determined, using the data of the ac imped-
ance data, which are collected in a time interval of 24 h,
where R is the universal gas constant and T is the absolute temper-
ature. The value 55.55 in the above equation is the concentration of

0.12 water in solution in mol/l. The calculated DGads value is

45.68 kJ mol1. The high values of Kads and DGads refer to high
adsorption and good inhibiting effect [59,60]. Generally, the energy
0.1 
values of DGads around 20 kJ mol1 or less negative are associated
with an electrostatic interaction between charged inhibitor mole-
0.08
cules and charged metal surface, physisorption; those of
C inh /θ (mM)

40 kJ mol1 or more negative involve charge sharing or transfer


0.06 from the inhibitor molecules to the metal surface to form a coordi-
nate type bond i.e., chemisorptions [61–64]. It is difficult to distin-
0.04 guish between chemisorption and physisorption only based on
these criteria, especially when charged species are adsorbed. The
possibility of Coulomb interactions between adsorbed cations and
0.02
specifically adsorbed anions can increase the Gibbs energy even if
no chemical bond appears [62]. According the obtained value of
0 

0 0.02 0.04 0.06 0.08 0.1 0.12


DGads (45.68 kJ mol1), it can be suggested that the adsorption of
the 2-TMAT is mainly due to chemisorption. This behavior is in
Cinh (mM)
accordance with the findings of other researchers in the case of
Fig. 9. Langmuir adsorption plot for carbon steel in 1 M HCl containing 2-TMAT at the thiadiazoles [53,65], 1,3,4-oxadiazoles [36], 1,2,4-triazoles
30 °C. [22,24] and the macrocyclic polyether compounds [66].
130 M. Tourabi et al. / Corrosion Science 75 (2013) 123–133

C 1s a N 1s b

Intensity (a.u.)

Intensity (a.u.)
289 288 287 286 285 284 283 282 404 403 402 401 400 399 398 397
Binding Energy (eV) Binding Energy (eV)

S 2p c
Intensity (a.u.)

S 2p 3/2

S 2p 1/2

168 167 166 165 164 163 162


Binding Energy (eV)

Fig. 10. The XPS deconvoluted profiles of (a) C 1s, (b) N 1s and (c) S 2p for pure 2-TMAT.

C 1s a
N 1s
b
Intensity a.u.)
Intensity (a.u.)

290 288 286 284 282 403 402 401 400 399 398
Binding Energy (eV) Binding Energy (eV)

Fe 2p 3/2
S 2p c O 1s d e
Intensity (a.u.)
Intensity (a.u.)
Intensity (a.u.)

170 168 166 164 162 160 535 533 531 529 527 717 715 713 711 709 707

Binding Energy (eV) Binding Energy (eV) Binding Energy (eV)

Fig. 11. The XPS deconvoluted profiles of (a) C 1s, (b) N 1s, (c) S 2p, (d) O 1s and (e) Fe 2p3/2 for 2-TMAT treated carbon steel.
M. Tourabi et al. / Corrosion Science 75 (2013) 123–133 131

Table 4
Binding energies (eV), relative intensity and their assignment for the major core lines observed pure 2-TMAT and 2-TMAT treated carbon steel substrate.

Substrate C 1s N 1s S 2p O 1s Fe 2p3/2
BE (eV) Assignment BE (eV) Assignment BE (eV) Assignment BE (eV) Assignment BE (eV) Assignment
Pure 2-TMAT 285.0 CAC/C@C/ 399.5 NH2 group/ 164.2 (60 %) Thiophen (S – – – –
(60%) CAH (40%) RACN 2p3/2)
286.2 CAN/C@N/ 400.6 @NA 165.4 (40 %) Thiophen (S – – – –
(40%) CAS (34%) structure 2p1/2)
– – 401.6 @N+H – – – – – –
(26%)
2-TMAT treated 285.0 CAC/C@C/ 399.6 @NA/RACN 161.8 (20 %) Disulfide 529.6 Fe2O3/ 707.8 Fe0
carbon steel (50%) CAH (39%) (FeS2) (30 %) Fe3O4 (1.2 %)
286.4 CAN/CAS 400.9 @NAFe 163.7 (27 %)/ Thiophen 531.1 FeOOH 711.5 Fe2O3/Fe3O4/
(26%) (49%) 164.9 (20 %) (>S) (54 %) (51.8 %) FeSO4/FeOOH
+
288.4 C@N 401.8 NHþ +
3 /@N H
168.5 (33 %) >S+ 532.6 Adsorbed 713.0 (31 FeCl3
(24%) (12%) (16 %) H2O %)
– – – – – – – – 714.6 (16 Satellite of Fe(III)
%)

In order to confirm this assumption (chemisorption of 2-TMAT)


and elucidate the nature of the organic thin film formed on the H H
steel surface, X-ray photoelectron spectroscopy (XPS) analyses N
are given and discussed below. For comparison purpose, the XPS S
spectra were obtained from the pure 2-TMAT, and the steel surface N
which was treated by 104 M of 2-TMAT after 24 h of immersion in
1 M HCl. The XPS spectra, shown in Figs. 10 and 11, were obtained
N S
both from pure 2-TMAT (C 1s, N 1s and S 2p) and the 2-TMAT-trea- N
δ(+)
ted carbon steel surface (C 1s, N 1s, S 2p, O 1s and Fe 2p). All XPS
spectra show complex forms, which were assigned to the corre- H H δ(−)
sponding species through a deconvolution fitting procedure. The N
binding energies (BE, eV) and the corresponding quantification S
(%) of each peak component are summarized in Table 4. N
The deconvoluted C 1s spectrum for pure 2-TMAT shows two
main peaks (Fig. 10a, Table 4). The first peak located at approx.
N S
285.0 eV has the largest contribution and is attributed to the N
CAC, C@C and CAH aromatic bonds [24]. The second peak at
Fig. 12. Intermolecular hydrogen bonding of 2-TMAT.
286.2 eV is attributed to the CAN and C@N bonds in the triazole
ring [24] and the CAS bond in the thienyl rings [67]. After acidic
immersion, the deconvoluted C 1s spectrum for 2-TMAT-treated nitrogen atom of another aminotriazole molecule as illustrated in
carbon steel shows three main peaks (Fig. 11a, Table 4). Indeed, Fig. 12. The high-resolution N 1s spectrum of the carbon steel sam-
the first component at a BE  285.0 eV is attributed to the CAC, ple treated by 2-TMAT shows peaks indicative of different chemical
C@C and CAH aromatic bonds which the largest contribution; forms of N presence on the steel surface (Fig. 11b, Table 4). The
the second component at 286.4 eV may be assigned to the carbon appearance of nitrogen species on the protected steel surface pro-
atoms bonded to: (i) nitrogen in CAN and C@N bonds in the tria- vides thus the evidence that the investigated 4-amino-1,2,4-tria-
zole ring and (ii) sulfur in CAS bond in the thienyl ring; the last zole (2-TMAT) was chemically adsorbed on the steel surface.
component at higher binding energy (located at approx. Moreover, the N 1s analyze of untreated carbon steel shows the ab-
288.4 eV) may be ascribed to the carbon atom of the C@N+ in the sence of the nitrogen in this substrate [72]. Similarly, three compo-
triazole ring [67], resulting probably from the protonation of the nents located at BE  399.6, 400.9 and 401.8 eV were used to fit the
@NA structure in the triazole ring and/or the coordination of nitro- spectrum of treated steel and the relative proportion of nitrogen
gen in triazole ring with the steel surface. species are quite different from those observed with pure 2-TMAT.
For the pure 2-TMAT molecule, the N 1s spectrum can be The first component is attributed to CAN and the unprotonated N
decomposed into three components centered at 399.5 eV, atom (@NA structure) in the triazole ring [69] while the second is
400.6 eV and 401.6 eV (Fig. 10b, Table 4). The first component mainly attributed to the coordinate nitrogen in triazole ring with
has the largest contribution and is mainly are attributable to the the steel surface. Indeed, the NAFe bond complex can cause the
bonds of CAN and the unprotonated N atoms (@NA structure) in peak shift to higher binding energy compared to uncoordi-
the triazole ring [68,69]. The second peak may be assigned to neu- nated @NA structure [73]. The same behavior was observed in
tral amino nitrogen (˜NANH2 in aminotriazole) [70]. The last peak our previously papers concerning the corrosion inhibition of mild
at higher binding energy (401.6 eV) is attributed to positively steel using 3,5-bis(n-pyridyl)-1,2,4-oxadiazoles [36], 3,5-bis(4-
charged nitrogen, and could be related to protonated nitrogen methoxyphenyl)-4-amino-1,2,4-triazole [24] and 3,6-bis(3-pyri-
atoms (@N+HA) in the triazole ring [12]. Indeed, the molecular dyl)pyridazine [50]. The last component at 401.8 eV is attributable
conformation of 2-TMAT can yield intermolecular hydrogen bond- to the oxidized nitrogen as a result of the protonation of the nitro-
ing which leads to a positive polarization of the nitrogen atom and gen in the amino group ðANHþ 3 Þ, which leads to a positive polariza-
therefore an increased binding energy [71]. Thus, the presence of tion of the nitrogen atom, and therefore a core-level chemical shift
the positively charged nitrogens (@N+HA structure) can be ex- to higher binding energy is produced [74].
plained by hydrogen bonding between one hydrogen atom bonded The S 2p core-level of the pure 2-TMAT is best resolved with at
to a nitrogen atom of the amino group and lone electron pair of a least two spin-orbit-split doublets (S 2p3/2 and S 2p1/2), with
132 M. Tourabi et al. / Corrosion Science 75 (2013) 123–133

binding energy (BE) for S 2p3/2 peak lying at about 164.2 eV analysis corroborate the thermodynamic results and give evidence
(Fig. 10c, Table 4). This former can be assigned to the neutral thio- of the chimisorption of 2-TMAT on the carbon steel surface in 1 M
phen as mentioned previously [75,76]. After immersion in 1 M HCl HCl medium.
containing 2-TMAT, the S 2p spectrum shows the presence of sulfur
on the steel surface but in low content (Fig. 11c). This spectrum
Acknowledgement
could be deconvoluted into four main components centered at
161.9 eV, 163.7 eV, 164.9 eV and 168.5 eV. The first one can be
The authors gratefully acknowledge the EGIDE for its financial
attributed to the sulfide species [77,78]; the second and third are
support under Volubilis MA/10/234.
attributed to the neutral thiophen and the last peak observed at
168.5 eV may be assigned to the SAFe bound complex and/or the
sulfur atoms in a more positive environment [76]. References
The O 1s spectrum for carbon steel surface after immersion in
[1] V.S. Sastri, Corrosion Inhibitors – Principles and Applications, Wiley,
1 M HCl solution containing 2-TMAT could be fitted into three
Chichester, England, 1998.
main peaks (Fig. 11d, Table 4). The first peak, at approx. [2] G. Schmitt, Br. Corros. J. 19 (1984) 165–176.
529.6 eV, is ascribed to O2 and could be related to oxygen atoms [3] G. Trabanelli, Corrosion 47 (1991) 410.
bonded to Fe3+ in the Fe2O3 and/or Fe3O4 oxides [79]. The second [4] Y.I. Kuznetsov, Organic Inhibitors of Corrosion of Metals, Springer, 1996.
[5] R. Coughlin, Corrosion inhibitors, in: J.J. Florio, D.J. Miller (Eds.), Handbook of
peak, observed at a BE  531.1 eV, is attributed to OH, can be Coatings Additives, second ed., Marcel Dekker, New York, 2004, pp. 127–144.
associated to the presence of hydrous iron oxides, such as FeOOH [6] B. Mernari, H. El Attari, M. Traisnel, F. Bentiss, M. Lagrenée, Corros. Sci. 40
[79]. Finally, the third peak at 532.9 eV may be assigned to oxygen (1998) 391–399.
[7] F. Bentiss, M. Traisnel, L. Gengembre, M. Lagrenée, Appl. Surf. Sci. 152 (1999)
of adsorbed water [80]. 237–249.
The Fe 2p spectrum for carbon steel surface covered with 2- [8] F. Bentiss, M. Lagrenée, M. Traisnel, J.C. Hornez, Corros. Sci. 41 (1999) 789–803.
TMAT exhibits two doublets, 711 eV (Fe 2p3/2) and 724.1 eV (Fe [9] F. Bentiss, M. Lagrenée, M. Traisnel, B. Mernari, H. Elattari, J. Appl. Electrochem.
29 (1999) 1073–1078.
2p1/2), with an associated ghost structure on the high energy side [10] F. Bentiss, M. Traisnel, H. Vezin, M. Lagrenée, Ind. Eng. Chem. Res. 39 (2000)
showing the subsequent oxidation of the steel surface. The decon- 3732–3736.
volution of the high resolution Fe 2p3/2 XPS spectrum consists in [11] F. Bentiss, M. Traisnel, M. Lagrenée, Br. Corros. J. 35 (2000) 315–320.
[12] F. Bentiss, M. Traisnel, L. Gengembre, M. Lagrenée, Appl. Surf. Sci. 161 (2000)
four peaks (Fig. 11e, Table 4). The small peak at lower bending en- 194–202.
ergy (707.8 eV) is attributable to metallic iron as previously re- [13] M.A. Quraishi, D. Jamal, Mater. Chem. Phys. 68 (2001) 283–287.
ported [81,82]. The second peak at a BE  711.5 eV assigned to [14] F. Bentiss, M. Bouanis, B. Mernari, M. Taisnel, M. Lagrenée, J. Appl. Electrochem.
32 (2002) 671–672.
Fe3+ as mentioned in [83], was attributed to ferric compounds such
[15] M. Lagrenée, B. Mernari, M. Bouanis, M. Taisnel, F. Bentiss, Corros. Sci. 44
as Fe2O3 (i.e., Fe3+ oxide), FeOOH (i.e., oxyhydroxyde) [84,85], while (2002) 573–588.
that located at around 713.0 eV is attributed to the presence of a [16] F. Bentiss, M. Lagrenée, B. Elmehdi, B. Mernari, M. Traisnel, Corrosion 58 (2002)
small concentration of FeCl3 on the metal surface [74,86]. The last 399–407.
[17] B. El Mehdi, B. Mernari, M. Traisnel, F. Bentiss, M. Lagrenée, Mater. Chem. Phys.
peak at BE  714.6 may be ascribed to the satellite of Fe(III) [87]. 77 (2003) 489–496.
The comparison of the Fe2p3/2 XPS results for carbon steel after [18] F. Bentiss, M. Traisnel, H. Vezin, M. Lagrenée, Corros. Sci. 45 (2003) 371–380.
immersion in 1 M HCl medium containing 2-TMAT with that for [19] H.-L. Wang, H.-B. Fan, J.-S. Zheng, Mater. Chem. Phys. 77 (2003) 655–661.
[20] M.A. Quraishi, H.K. Sharma, Mater. Chem. Phys. 78 (2003) 18–21.
untreated surface steel, previously described [72], shows that the [21] L. Wang, Corros. Sci. 48 (2006) 608–616.
aggressive solution induces a significant decrease in Fe0 amount [22] F. Bentiss, M. Bouanis, B. Mernari, M. Traisnel, H. Vezin, M. Lagrenée, Appl. Surf.
in favor of oxidized species (Fe2+ and Fe3+) indicating therefore that Sci. 253 (2007) 3696–3704.
[23] M. Lebrini, M. Traisnel, M. Lagrenée, B. Mernari, F. Bentiss, Corros. Sci. 50
the oxide layer thickness is increasing. The formation of a stable (2008) 473–479.
and insoluble oxide layer (FeOOH), which can reduce ions diffu- [24] F. Bentiss, C. Jama, B. Mernari, H. El Attari, L. El Kadi, M. Lebrini, M. Traisnel, M.
sion, may be benefice in improving the corrosion resistance of car- Lagrenée, Corros. Sci. 51 (2009) 1628–1635.
[25] S. Zhang, Z. Tao, S. Liao, F. Wu, Corros. Sci. 52 (2010) 3126–3132.
bon steel in HCl solutions. [26] A.Y. Musa, A.A.H. Kadhum, A.B. Mohamad, M.S. Takriff, A.R. Daud, S.K.
On the basis of XPS analyses, the obtained results give evidence Kamarudin, Corros. Sci. 52 (2010) 526–533.
of chemical interactions between the 2-TMAT inhibitor and carbon [27] A.Y. Musa, A.A.H. Kadhum, A.B. Mohamad, M.S. Takriff, Corros. Sci. 52 (2010)
3331–3340.
steel surface. Indeed, the presence of nitrogen species on the steel
[28] A.Y. Musa, A.B. Mohamad, A.A.H. Kadhum, M.S. Takriff, Int. J. Electrochem. Sci.
surface, such as @NA structure, thiophen and NAC, confirms that 6 (2011) 2758–2766.
the investigated aminotriazole was chemisorbed on the carbon [29] H. Zarrok, A. Zarrouk, B. Hammouti, R. Salghi, C. Jama, F. Bentiss, Corros. Sci. 64
steel surface and corroborates the thermodynamic study. Thus 2- (2012) 243–252.
[30] A. Zarrouk, B. Hammouti, S.S. Al-Deyab, R. Salghi, H. Zarrok, C. Jama, F. Bentiss,
TMAT can be regarded as good inhibitor for carbon steel corrosion Int. J. Electrochem. Sci. 7 (2012) 5997–6011.
in normal hydrochloric acid medium. [31] W. Qafsaoui, H. Takenouti, Corros. Sci. 52 (2010) 3667–3676.
[32] Z. Khiati, A.A. Othman, M. Sanchez-Moreno, M.-C. Bernard, S. Joiret, E.M.M.
Sutter, V. Vivier, Corros. Sci. 53 (2011) 3092–3099.
[33] B.D. Mert, M.E. Mert, G. Kardas, B. Yazici, Corros. Sci. 53 (2011) 4265–4272.
4. Conclusions [34] F. Bentiss, M. Lagrenée, M. Traisnel, B. Mernari, H. Elattari, J. Heterocycl. Chem.
36 (1999) 149–152.
[35] F. Bentiss, M. Lagrenée, D. Barbry, Tetrahedron Let. 41 (2000) 1539–1541.
2,5-Bis(2-thienylmethyl)-4-amino-1,2,4-triazole (2-TMAT) [36] M. Outirite, M. Lagrenée, M. Lebrini, M. Traisnel, C. Jama, H. Vezin, F. Bentiss,
shows excellent inhibition properties for the corrosion of carbon Electrochim. Acta 55 (2010) 1670–1681.
steel in 1 M HCl at 30 °C, and the inhibition efficiency, g (%), in- [37] A. Popova, E. Sokolova, S. Raicheva, M. Christov, Corros. Sci. 45 (2003) 33–58.
[38] F.B. Growcock, J.H. Jasinski, J. Electrochem. Soc. 136 (1989) 2310–2314.
creases with increase in the inhibitor concentration. The EIS spec- [39] J.R. Macdonald, W.B. Johanson, in: J.R. Macdonald (Ed.), Theory in Impedance
tra are well described by the proposed structural models. Based on Spectroscopy, John Wiley & Sons, New York, 1987.
the Tafel polarization results, 2-TMAT can be classified as mixed [40] D.A. Lopez, S.N. Simison, S.R. de Sanchez, Electrochim. Acta 48 (2003) 845–854.
[41] Z.B. Stoynov, B.M. Grafov, B. Savova-Stoynova, V.V. Elkin, Electrochemical
inhibitor. The ac impedance and polarization methods are in good
Impedance, Nauka, Moscow, 1991.
agreement. The corrosion process is inhibited by the adsorption of [42] A. Popova, M. Christov, A. Vasilev, Corros. Sci. 49 (2007) 3290–3302.
2-TMAT on the steel surface and the adsorption of the inhibitor fits [43] R. Solmaz, G. Kardasß, M. Çulha, B. Yazıcı, M. Erbil, Electrochim. Acta 53 (2010)

a Langmuir isotherm model. The calculated value of DGads can sug- 5941–5952.
[44] M. Lebrini, F. Bentiss, N. Chihib, C. Jama, J.P. Hornez, M. Lagrenée, Corros. Sci.
gest that the adsorption mechanism of 2-TMAT on carbon steel 50 (2008) 2914–2918.
surface in 1 M HCl solution is mainly due to chemisorption. XPS [45] M. Kedam, O.R. Mattos, H. Takenouti, J. Electrochem. Soc. 128 (1981) 266–274.
M. Tourabi et al. / Corrosion Science 75 (2013) 123–133 133

[46] J. Morales Roque, T. Pandiyan, J. Cruz, E. García-Ochoa, Corros. Sci. 50 (2008) [68] E.T. Kang, K.G. Neoh, K.L. Tan, Surf. Interface Anal. 19 (1992) 33–37.
614–624. [69] M. Lebrini, M. Lagrenée, M. Traisnel, L. Gengembre, H. Vezin, F. Bentiss, Appl.
[47] A.Y. Musa, A.A.H. Kadhum, M.S. Takriff, A.R. Daud, S.K. Kamarudin, N. Surf. Sci. 253 (2007) 9267–9276.
Muhamad, Corros. Eng. Sci. Technol. 45 (2010) 163–168. [70] O. Olivares-Xometl, N.V. Likhanova, M.A. Domínguez-Aguilar, J.M. Hallen, L.S.
[48] M.A. Quraishi, K.R. Sudheer, Eno E. Ebenso, Int. J. Electrochem. Sci. 7 (2012) Zamudio, E. Arce, Appl. Surf. Sci. 252 (2006) 2139–2152.
7476–7492. [71] A. Welle, J.D. Liao, K. Kaiser, M. Grunze, U. Mäder, N. Blank, Appl. Surf. Sci. 119
[49] M.S. Kumar, S.L. Ashok Kumar, A. Sreekanth, Ind. Eng. Chem. Res. 51 (2012) (1997) 185–198.
5408–5418. [72] F.Z. Bouanis, F. Bentiss, S. Bellayer, M. Traisnel, J.B. Vogt, C. Jama, Mater. Chem.
[50] F. Bentiss, M. Outirite, M. Traisnel, H. Vezin, M. Lagrenée, B. Hammouti, S.S. Al- Phys. 127 (2011) 329–334.
Deyab, C. Jama, Int. J. Electrochem. Sci. 7 (2012) 1699–1723. [73] O. Olivares, N.V. Likhanova, B. Gómez, J. Navarrete, M.E. Llanos-Serrano, E.
[51] I. Ahamad, R. Prasad, M.A. Quraishi, Corros. Sci. 52 (2010) 1472–1481. Arce, J.M. Hallen, Appl. Surf. Sci. 252 (2006) 2894–2909.
[52] S.S. Abdel-Rehim, M.A.M. Ibrahim, K.F. Khaled, Mater. Chem. Phys. 70 (2001) [74] F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, in: J. Chastain (Ed.),
268–273. Handbook of X-ray Photoelectron Spectroscopy, Perkin–Elmer Corp.,
[53] F. Bentiss, B. Mernari, M. Traisnel, H. Vezin, M. Lagrenée, Corros. Sci. 53 (2011) Minnesota, USA, 1992.
487–495. [75] M. Lebrini, M. Lagrenée, H. Vezin, L. Gengembre, F. Bentiss, Corros. Sci. 47
[54] M.M. El-Naggar, Corros. Sci. 49 (2007) 2226–2236. (2005) 485–505.
[55] A. Popova, M. Christov, A. Zwetanova, Corros. Sci. 49 (2007) 2131–2143. [76] E.T. Kang, K.G. Neoh, K.L. Tan, Phys. Rev. B 44 (1991) 10461–10469.
[56] M. Bouklah, B. Hammouti, M. Lagrenée, F. Bentiss, Corros. Sci. 48 (2006) 2831– [77] N.R. Urban, K. Ernst, S. Bernasconi, Geochim. Cosmochim. Acta 63 (1999) 837–
2842. 853.
[57] M. Elayyachy, A. El Idrissi, B. Hammouti, Corros. Sci. 48 (2006) 2470–2479. [78] J. Riga, J.J. Verbist, P. Josseaux, A.K. Mesmaeker, Surf. Interface Anal. 7 (1985)
[58] F. Bentiss, M. Lebrini, H. Vezin, F. Chai, M. Traisnel, M. Lagrenée, Corros. Sci. 51 163–168.
(2009) 2165–2173. [79] W. Temesghen, P.M.A. Sherwood, Anal. Bioanal. Chem. 373 (2002) 601–608.
[59] C.M. Goulart, A. Esteves-Souza, C.A. Martinez-Huitle, C.J.F. Rodrigues, M.A.M. [80] K. Babić-Samardžija, C. Lupu, N. Hackerman, A.R. Barron, A. Luttge, Langmuir
Maciel, A. Echevarria, Corros. Sci. 67 (2013) 281–291. 21 (2005) 12187–12196.
[60] R. Solmaza, G. Kardas, M. Culha, B. Yazıcı, M. Erbil, Electrochim. Acta 53 (2008) [81] R. Devaux, D. Vouagner, A.M. De Becdelievre, C. Duret-Thual, Corros. Sci. 36
5941–5952. (1994) 171–186.
[61] M. Ehteshamzadeha, A.H. Jafari, E. Naderia, M.G. Hosseini, Mater. Chem. Phys. [82] V. Di castro, S. Ciampi, Surf. Sci. 331 (1995) 294–299.
113 (2009) 986–993. [83] N. Nakayamaand, A. Obuchi, Corros. Sci. 45 (2003) 2075–2092.
[62] A.K. Singh, M.A. Quraishi, Corros. Sci. 53 (2011) 1288–1297. [84] M.A. Pech-Canul, P. Bartolo-Perez, Surf. Coat. Technol. 184 (2004) 133–140.
[63] M.J. Bahrami, S.M.A. Hosseini, P. Pilvar, Corros. Sci. 52 (2010) 2793–2803. [85] F.Z. Bouanis, F. Bentiss, M. Traisnel, C. Jama, Electrochim. Acta 54 (2009) 2371–
[64] M. Behpour, S.M. Ghoreishi, N. Mohammadi, N. Soltani, M. Salavati-Niasari, 2378.
Corros. Sci. 52 (2010) 4046–4057. [86] V.S. Sastri, M. Elboujdaini, J.R. Romn, J.R. Perumareddi, Corrosion 52 (1996)
[65] F. Bentiss, M. Lebrini, M. Lagrenée, Corros. Sci. 47 (2005) 2915–2931. 447–452.
[66] M. Lebrini, M. Lagrenée, H. Vezin, M. Traisnel, F. Bentiss, Corros. Sci. 49 (2007) [87] A. Galtayries, R. Warocquier-Clérout, M.-D. Nagel, P. Marcus, Surf. Interface
2254–2269. Anal. 38 (2006) 186–190.
[67] J.F. Watts, J. Wolstenholme, An Introduction to Surface Analysis by XPS and
AES, John Wiley and Sons Inc., UK, 2003.

You might also like