1-s2.0-S0010938X14005095-main_2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Corrosion Science 90 (2015) 572–584

Contents lists available at ScienceDirect

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

New 1H-pyrrole-2,5-dione derivatives as efficient organic inhibitors


of carbon steel corrosion in hydrochloric acid medium: Electrochemical,
XPS and DFT studies
A. Zarrouk a, B. Hammouti a, T. Lakhlifi b, M. Traisnel c, H. Vezin d, F. Bentiss c,e,⇑
a
Laboratoire de Chimie Analytique et Environnement, LCAE-URAC18, Faculté des Sciences, Université Mohammed Ier, B.P. 4808, M-60000 Oujda, Morocco
b
Département de Chimie, Faculté des Sciences, Université Moulay Ismail, Meknès, B.P. 4010, Morocco
c
UMET-PSI, CNRS UMR 8207, ENSCL, Université Lille I, CS 90108, F-59652 Villeneuve d’Ascq Cedex, France
d
Laboratoire de Spectrochimie Infrarouge et Raman (LASIR), UMR-CNRS 8516, Bâtiment C5, F-59655 Villeneuve d’Ascq Cedex, France
e
Laboratoire de Catalyse et de Corrosion des Matériaux (LCCM), Faculté des Sciences, Université Chouaib Doukkali, B.P. 20, M-24000 El Jadida, Morocco

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

Article history: New 1H-pyrrole-2,5-dione derivatives, namely 1-phenyl-1H-pyrrole-2,5-dione (PPD) and 1-(4-methyl-
Received 7 July 2014 phenyl)-1H-pyrrole-2,5-dione (MPPD) were synthesised and their inhibitive action against the corrosion
Accepted 31 October 2014 of carbon steel in 1 M HCl solution were investigated at 308 K by weight loss, potentiodynamic polariza-
Available online 8 November 2014
tion curves, and electrochemical impedance spectroscopy (EIS) methods. The results showed that the
investigated 1H-pyrrole-2,5-dione derivatives are good corrosion inhibitors for carbon steel in 1 M HCl
Keywords: medium, their inhibition efficiency increased with inhibitor concentration, and MPPD is slightly more
A. Carbon steel
effective than PPD. Potentiostatic polarization study showed that PPD and MPPD are mixed-type inhib-
A. Acid solutions
B. XPS
itors in 1 M HCl. Impedance experimental data revealed a frequency distribution of the capacitance,
C. Acid inhibition simulated as constant phase element. The results obtained from electrochemical and weight loss studies
were in reasonable agreement. The adsorption of MPPD and PPD on steel surface obeyed Langmuir’s
adsorption isotherm. Thermodynamic data and XPS analysis clearly indicated that the adsorption
mechanism of 1H-pyrrole-2,5-dione derivatives on carbon steel surface in 1 M HCl solution is mainly con-
trolled by a chemisorption process. Quantum chemical calculations using the Density Functional Theory
(DFT) were performed on 1H-pyrrole-2,5-dione derivatives to determine the relationship between
molecular structures and their inhibition efficiencies.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction these molecules adsorb on the metal surface by displacing water


molecules on the surface and forming a protective film [12–14].
Acid solutions are generally used for the removal of rust and The adsorption of these molecules depends mainly on certain
scale in industrial processes. Inhibitors are generally used in these physicochemical properties of the inhibitor molecule such as func-
processes to control the metal dissolutions. Hydrochloric acid is tional groups, steric factors, aromaticity, electron density at the
widely used in the pickling of steel and different alloys. Most pop- donor atoms and p orbital character of donating electrons and
ular inhibitors are organic compound containing N, S, and O atoms. the electronic structure of the molecules [14,15]. Regarding the
Organic compound containing functional electronegative groups adsorption of inhibitor on the metal surface, two types of interac-
and p electrons in triple or conjugated double bonds are usually tions are responsible. One is physical adsorption, which involves
good inhibitors. The study of corrosion process and their inhibition electrostatic forces between ionic charges or dipoles of the
by organic compounds is a very active field of research [1–8]. The adsorbed species and the electric charge at metal/solution inter-
efficiency of these molecules depends mainly on their abilities to face. The other is chemical adsorption, which involves charge shar-
adhere onto the metal surfaces [9–11]. The researches show that ing or charge transfer from inhibitor molecules to the metal surface
to form coordinate types of bond [16,17]. Theoretical calculations
have been used recently to explain the mechanism of corrosion
⇑ Corresponding author at: UMET-PSI, CNRS UMR 8207, ENSCL, Université Lille I,
inhibition, which proved to be a very powerful tool in this direction
CS 90108, F-59652 Villeneuve d’Ascq Cedex, France. Tel.: +33 320 434 482; fax: +33
320 436 584. [18–22]. The geometry of inhibitor molecule in its ground state,
E-mail address: fbentiss@enscl.fr (F. Bentiss). the nature of its molecular orbitals, HOMO (Highest Occupied

http://dx.doi.org/10.1016/j.corsci.2014.10.052
0010-938X/Ó 2014 Elsevier Ltd. All rights reserved.
A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584 573

Molecular Orbital) and LUMO (Lowest Unoccupied Molecular Orbi- 600 and 1200); rinsed with distilled water, degreased in acetone
tal) are directly involved in the corrosion inhibition activity. The in an ultrasonic bath immersion for 5 min, washed again with
selection of appropriate inhibitors mainly depends on the type of bidistilled water and then dried at room temperature before use.
acid, its concentration, temperature, the presence of dissolved The acid solutions (1 M HCl) were prepared by dilution of an ana-
inorganic and/or organic substances even in minor amounts and, lytical reagent grade 37% HCl with double-distilled water. The con-
of course, on the type of metallic material supposed to be protected centration range of 1H-pyrrole-2,5-dione derivatives employed
[17]. was 0.001–1 mM.
The aim of this paper is to explore, for the first time, the use of
two 1H-pyrrole-2,5-dione derivatives as corrosion inhibitors for 2.2. Corrosion tests
carbon steel in 1 M HCl medium. The choice of 1H-pyrrole-2,5-
dione derivatives was based on the consideration that these 2.2.1. Weight loss measurements
organic compounds contain many p-electrons and hetero atoms The gravimetric measurements were carried out at definite time
like one nitrogen and two oxygen atoms which induce greater interval of 6 h at room temperature using an analytical balance
adsorption on the steel surface compared with other organic inhib- (precision ± 0.1 mg). The carbon steel specimens used have a rect-
itors. The corrosion inhibitive activity of these organic compounds angular form (length = 1.6 cm, width = 1.6 cm, thickness = 0.2 cm).
was examined successively via weight loss, potentiodynamic Gravimetric experiments were carried out in a double glass cell
polarization curves, electrochemical impedance spectroscopy equipped with a thermostated cooling condenser containing
(EIS), isotherm calculations, and X-ray photoelectron spectroscopy 100 mL of non-de-aerated test solution. After immersion period,
(XPS) techniques. Molecular modelling has been also conducted in the steel specimens were withdrawn, carefully rinsed with bidis-
order to determine the correlation between the corrosion inhibi- tilled water, ultrasonic cleaning in acetone, dried at room temper-
tion properties of the investigated 1H-pyrrole-2,5-dione deriva- ature and then weighted. Triplicate experiments were performed
tives and their molecular structures. in each case and the mean value of the weight loss was calculated.

2.2.2. Electrochemical measurements


2. Experimental details Electrochemical experiments were conducted using an
electrochemical measurement system Tacussel Radiometer PGZ
2.1. Materials 100 controlled by a PC supported by Voltamaster.4 Software. A
conventional three-electrode cylindrical Pyrex glass cell was used.
The tested inhibitors, namely 1-phenyl-1H-pyrrole-2,5-dione The temperature is thermostatically controlled. The working elec-
(PPD) and 1-(4-methylphenyl)-1H-pyrrole-2,5-dione (MPPD) were trode was carbon steel with the surface area of 1 cm2. A saturated
synthesised according to a previously described experimental pro- calomel electrode (SCE) was used as a reference. All potentials
cedure [23]. The structures of the PPD and MPPD were confirmed were given with reference to this electrode. The counter electrode
by FTIR, 1H, 13C NMR, and mass spectroscopy. The molecular struc- was a platinum plate of surface area of 1 cm2. Before starting the
tures of these 1H-pyrrole-2,5-dione derivatives are shown in Fig. 1. experiments, the working electrode was immersed in test solution
PPD: Yield 86.8%; m.p. 127–128.5 °C; IR (KBr, cm1): 1685 (tCO, for 30 min at open circuit potential (OCP) to reach a steady state,
imide); 1H NMR (dimethyl-d6 sulfoxide; 300 MHz): D (ppm) 6.32 the time necessary to reach a quasi-stationary value for the
(d, 1Holefin, J = 12.06 Hz); 6.50 (d, 1Holefin, J = 12.06 Hz); 7.10 (dd, open-circuit potential. After measuring the Eocp, the electrochemi-
1Harom.para, J = 7.38 Hz); 7.34 (dd, 2Harom. meta, J = 7.64 Hz); 7.64 cal measurements were performed. All electrochemical tests have
(2d, 2Harom. ortho, J = 7.31 Hz); 13C NMR (dimethyl-d6 sulfoxide; been performed in aerated solutions at 308 K. The response of
75 MHz): D (ppm) 127 (Carom. para); 129 (Carom. ortho); 130 (Carom. the electrochemical system to ac excitation with a frequency rang-
meta); 132 (Carom.AN); 138 (C@Colefin); 163 (C@O). ing from 105 Hz to 101 Hz and peak to peak amplitude of 10 mV
MPPD: Yield 63.8%, m.p. 158–160 °C; IR (KBr, cm1): 1690 was measured with data density of 10 points per decade. All elec-
(tCO), imide); 1H NMR (dimethyl-d6 sulfoxide; 300 MHz): D trochemical impedance spectroscopy diagrams were recorded at
(ppm) 2.20 (s, 3H, CH3); 6.32 (d, 1Holefin, J = 12.23 Hz); 6.50 (d, the open circuit potential, i.e., at the corrosion potential Ecorr. The
1H olefin, J = 12.23 Hz); 7.14 (d. 2Harom. ortho, J = 8.40 Hz); 7.54 (d. impedance data were analyzed and fitted with the simulation
2Harom. meta, J = 8.40 Hz); 13C NMR (dimethyl-d6 sulfoxide; ZView 2.80, equivalent circuit software. After ac impedance test,
75 MHz): D (ppm) 20.9 (CH3); 120.1 (Carom. para); 130.0 (Carom. the potentiodynamic Tafel measurements were scanned from
ortho); 131.3 (Carom. meta); 132.1 (Carom.AN); 136.3 (C@Colefin); cathodic to the anodic direction, E = Ecorr ± 200 mV, with a scan rate
163.5 (C@O). of 1 mV s1. The potentiodynamic data were analyzed using the
The steel used in this study is a carbon steel (Euronorm: C35E polarization VoltaMaster 4 software.
carbon steel and US specification: SAE 1035) with a chemical com-
position (in wt%) of 0.370% C, 0.230% Si, 0.680% Mn, 0.016% S, 2.3. X-ray photoelectron spectroscopy (XPS)
0.077% Cr, 0.011% Ti, 0.059% Ni, 0.009% Co, 0.160% Cu and the
remainder iron (Fe). The carbon steel samples were pre-treated X-ray photoelectron spectroscopy (XPS) spectra were recorded
prior to the experiments by grinding with emery paper SiC (120, 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 of
around 1 mm. The analyser was operated in constant pass
O O energy of 40 eV using an analysis area of approximately
700 lm  300 lm. Charge compensation was applied to compen-
sate for the charging effects that occurred during the analysis.
N CH3 N The C 1s (285.0 eV) binding energy (BE) was used as internal refer-
ence. The spectrometer BE scale was initially calibrated against the
O Ag 3d5/2 (368.2 eV) level. Pressure was in the 1010 torr range dur-
O
MPPD PPD ing the experiments. Quantification and simulation of the experi-
mental photopeaks were carried out using CasaXPS software.
Fig. 1. The structures of the studied 1H-pyrrole-2,5-dione derivatives. Quantification took into account a non-linear Shirley background
574 A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584

subtraction [24]. XPS analyses were practiced on pure 1H-pyrrole- Analyse of the obtained results in Table 1 reveals that the cor-
2,5-dione derivatives and on protected steel surface with theses rosion rate (CR), decreases with increase in inhibitor concentration
organic inhibitors. The carbon steel sample (1 cm2) was pre- in both cases (PPD and MPPD). At the same time, the inhibition effi-
treated by the same procedure as for the gravimetric test. After ciency, gWL (%), was enhanced by the inhibitor concentration
6 h of immersion at 308 K, the carbon steel sheet was rinsed with reaching a maximum of 97.5% at 1 mM in the case of MPPD.
acetone and ultra pure water. Indeed, when the concentration of inhibitors was less than 1 mM,
gWL increased sharply with an increase in concentration, but a fur-
ther raise inhibitor concentration (0.1–1 mM) caused no apprecia-
2.4. Quantum chemical calculations
ble change in inhibitive performance of these inhibitors (Table 1).
This behavior can be due to the fact that the adsorption coverage
All theoretical ab initio calculations were performed by means
of inhibitor on steel surface increases with the inhibitor concentra-
of the Gaussian 03 program on 18 processors Linux Transtec Clus-
tion. The good inhibitive performance of 1H-pyrrole-2,5-dione
ter with parallel G03 Linda version. The molecular structures of
derivatives may be explained on the basis of adsorption of the
PPD and MPPD (neutral and diprotonated forms) were fully and
these molecules (physisorption and/or chemisorption). The MPPD
geometrically optimized using the functional hybrid B3LYP Density
and PPD molecules can easily protonated to form cation-ionic
Functional Theory formalism (DFT) and the 6-31+G (2d,2p) orbital
forms in HCl solution and therefore adsorb at the metal surface
basis set for all atoms. Self Consistent Reaction Field (SCRF) calcu-
through the already adsorbed chloride ions. Also, the adsorption
lations in consideration of the solvent (water) effect were done
of the neutral 1H-pyrrole-2,5-dione molecules could occur through
using the Polarized Continuum Method (PCM) with UAHF set of
to the formation of links between the d-orbital of iron atoms,
radii for solvent. The molecular structures were fully and
involving the displacement of water molecules from the metal sur-
geometrically optimized with tight SCF in solvent. Four main
face, and the lone sp2 electron pairs present on the N, and O atoms
quantum related parameters were considered: the energy of the
and p-orbitals in aromatic ring, blocking the active sites in the steel
highest occupied molecular orbital (EHOMO), the energy of the
surface and therefore decreasing the corrosion rate.
lowest unoccupied molecular orbital (ELUMO), energy band gap
At a concentration 1 mM of each studied 1H-pyrrole-2,5-dione
DE = ELUMO  EHOMO and the dipole moment (l).
derivative, the inhibition efficiency, attains 97.5%, and 94.4% for
MPPD and PPD, respectively. Thus, we deduce that MPPD is the
3. Results and discussion best inhibitor of these two tested compounds. The variation in
inhibitive efficiency mainly depends on the type and the nature
3.1. Weight loss study of the substituents present in the inhibitor molecule. The ability
of the molecule to adsorb on the steel surface was dependent on
Corrosion inhibition performance of organic inhibitors can be the group in para position in phenyl substituent. Indeed, it appears
evaluated by using weight loss method. This method of monitoring that replacement of H atom in para position in phenyl substituent
corrosion rate is useful because of its simple application and reli- of PPD molecule by a CH3 group in MPPD arises a sleight enhance-
ability. The corrosion parameters obtained by conducting weight ment in the inhibition efficiency. This enhanced efficiency may be
loss measurements for carbon steel in the absence and presence due to the introduction of the electron releasing on the pyrrole-
of different concentration of 1H-pyrrole-2,5-dione derivatives 2,5-dione moiety, giving therefore, a favorable electron density
(MPPD and PPD) in 1 M HCl after 6 h of immersion at 308 K are for preferential adsorption interactions. Hence it facilitates greater
given in Table 1. The experiments were done by triplicate and adsorption of MPPD on carbon steel surface than PPD, leading to
the average value of the weight loss was used to calculate the higher inhibition efficiency of MPPD than PPD. The same effect
corrosion rate (CR), in mg cm2 h1 and the inhibition efficiency, has been observed for disubstituted 1,3,4-oxadiazole [26], 1,3,4-
gWL, in percent (%). These were calculated from Eqs. (1) and (2) thiadiazole [27] and disubstituted-4H-1,2,4-triazole derivatives
[25], respectively: [28]. This difference in effectiveness could be explained by compu-
tational calculations using DFT method (see the last part of this
Wb  Wa work).
CR ¼ ð1Þ
At 
wi
gWL ð%Þ ¼ 1   100 ð2Þ 3.2. Tafel polarization study
w0

where Wb and Wa are the specimen weight before and after immer- Tafel polarization curves for carbon steel in 1 M HCl for differ-
sion in the tested solution, w0 and wi are the values of corrosion ent concentrations of PPD and MPPD are shown in Fig. 2. The
weight losses of carbon steel in uninhibited and inhibited solutions, related electrochemical parameters such as corrosion potential
respectively, A the total area of the one steel specimen (cm2) and t is (Ecorr), corrosion current density (Icorr), and Tafel cathodic constant
the exposure time (h). (bc) were listed in Table 2. The Icorr was determined by Tafel

Table 1
Corrosion parameters obtained from weight loss measurements for carbon steel in 1 M HCl containing various concentration of 1H-pyrrole-2,5-dione derivatives at 308 K.

Inhibitor Concentration (mM) CR (mg cm2 h1) gWL (%) h


Blank – 1.070 ± 0.005 – –
MPPD 0.001 0.206 ± 0.003 80.7 ± 0.3 0.807
0.01 0.124 ± 0.003 88.4 ± 0.3 0.884
0.1 0.068 ± 0.002 93.6 ± 0.1 0.936
1 0.027 ± 0.001 97.5 ± 0.1 0.975
PPD 0.001 0.213 ± 0.005 80.1 ± 0.4 0.801
0.01 0.136 ± 0.005 87.3 ± 0.4 0.873
0.1 0.083 ± 0.003 92.2 ± 0.3 0.922
1 0.060 ± 0.002 94.4 ± 0.2 0.944
A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584 575

0.1 of the constant cathodic Tafel slopes, bc, for the two studied inhib-
MPPD
itors suggest that the cathodic process, the hydrogen evolution
reaction, is activation controlled and the addition of 1H-pyrrole-
0.01
2,5-dione derivatives does not modify the mechanism of this pro-
cess [30,31]. In acidic medium the reduction of H+ ions at the car-
1E-3 bon steel surface takes place mainly through a charge transfer
mechanism. The inhibitor molecules are first adsorbed onto the
I (A cm )
-2

carbon steel surface and, therefore, impedes by merely blocking


1E-4 the reaction sites of the carbon steel surface. However, for poten-
Blank tial higher than 250 mVSCE, the presence of PPD and MPPD did
10-3 M not change the current-vs.-potential characteristics in anodic
1E-5
10-4 M domain (Fig. 2). This potential can be defined as the desorption
10-5 M potential. The phenomenon may be explained by the equality of
10-6 M the rate of adsorption of inhibitor and that of the metal oxidation
1E-6
leading to a desorption of the inhibitor molecules from the elec-
-0.7 -0.6 -0.5 -0.4 -0.3 trode surface [32]. From Table 2, it is also clear that there is a shift
E (V/SCE) toward cathodic region in the values of corrosion potential (Ecorr).
In the literature [33,34], it has been reported that if the displace-
0.1 ment in Ecorr(inh) is bigger than 85 mVSCE from Ecorr, the inhibitor
PPD can be seen as a cathodic or anodic type; and if displacement in
0.01 Ecorr(inh) is less than 85 mVSCE, the inhibitor can be seen as mixed
type. In our study the maximum displacement in Ecorr value was
72.2 mVSCE toward cathodic region, which indicates that these
1E-3
investigated 1H-pyrrole-2,5-dione derivatives (PPD and MPPD)
I (A cm )

are mixed type inhibitors. It can be observed from Table 2 that


-2

1E-4 the jcorr values decrease considerably in the presence of PPD and
Blank
MPPD and decrease with increasing the inhibitor concentration.
1E-5 10 -3 M Correspondingly, gTafel (%) values increase with increasing the
10 -4 M inhibitor concentration reaching a maximum value at 1 mM in
10 -5 M
the both cases of inhibitors. It is also evident that MPPD presents
1E-6 10 -6 M
the slightly better performance than PPD, which can be correlated
-0.7 -0.6 -0.5 -0.4 -0.3 to the slight difference between the two inhibitor molecules.
E (V/SCE) Indeed, the introduction of CH3 group in the case of MPPD arises
an enhancement in the inhibition efficiency. Tafel polarization also
Fig. 2. Polarization curves for carbon steel in 1 M HCl containing different confirms the inhibiting nature of 1H-pyrrole-2,5-dione derivatives
concentrations of 1H-pyrrole-2,5-dione derivatives. and the inhibition efficiencies values, calculated from the potentio-
dynamic polarization method results show the same trend as those
obtained from weight loss measurements (Tables 1 and 2).
extrapolation of only the cathodic polarization curve alone, which
usually produces a longer and better defined Tafel region [29]. The
jcorr values were used to calculate the inhibition efficiency, gTafel (%) 3.3. Ac impedance study
(listed in Table 2), using the following equation [8]:
The corrosion inhibition property of PPD and MPPD on carbon
Icorr  IcorrðiÞ steel was also examined by electrochemical impedance spectros-
gTafel ð%Þ ¼  100 ð3Þ copy (EIS). Fig. 3 shows the Nyquist diagrams of carbon steel
Icorr
obtained at open-circuit potential in 1 M HCl solution in the
where Icorr and Icorr(i) are the corrosion current densities for steel absence and presence of different concentrations of 1H-pyrrole-
electrode in the uninhibited and inhibited solutions, respectively. 2,5-dione derivatives (MPPD and PPD). As shown in Fig. 3, when
As can been seen from Fig. 2, it is clear that both anodic and the 1H-pyrrole-2,5-dione derivatives are added to the corrosive
cathodic reactions of corrosion process were inhibited when the solution, the impedance diagrams are larger than in the blank solu-
1H-pyrrole-2,5-dione derivatives were added to the acid solution. tion. The size of the impedance diagram increases as the concen-
The parallel cathodic Tafel lines in Fig. 2 and he small variation tration rises and consequently the protection efficiency increases,

Table 2
Polarization parameters and the corresponding inhibition efficiencies for the corrosion of carbon steel in 1 M HCl containing different concentrations of 1H-pyrrole-2,5-dione
derivatives at 308 K.

Inhibitor Concentration (mM) Ecorr (mV/SCE) bc (mV dec1) Icorr (lA cm2) gTafel (%)
Blank – 475.9 176.0 1077.8 –
MPPD 0.001 548.1 166.4 212.4 80.3
0.01 513.4 168.7 129.8 88.0
0.1 497.1 172.1 74.6 93.1
1 472.4 184.8 32.2 97.0
PPD 0.001 545.5 165.4 221.2 79.5
0.01 520.5 168.9 143.0 86.7
0.1 492.3 170.5 88.6 91.8
1 473.0 188.9 61.5 94.3
576 A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584

750
MPDD Blank
10-3 M
600 10-4 M
10-5 M
10-6 M
Fig. 4. Electrical equivalent circuit used for modeling the interface carbon steel/1 M
450
HCl solution without and with 1H-pyrrole-2,5-dione derivatives.
-Zi (Ω cm2)

10 Hz
15.8 Hz
300 5 Hz
[39,40]. The charge transfer resistance is corresponding to the cor-
2.5 Hz
40 Hz rosion reaction at metal substrate/solution interface, whose value
12.5 Hz
is a measure of electron transfer across the surface and is propor-
150 125Hz
tional to corrosion rate. Excellent fit with this model was obtained
5 Hz 1 Hz 0.1 Hz
1 Hz for all experimental data. As an example, the Nyquist and Bode
plots of both experimental and simulated data of carbon steel in
0
1 M HCl solution containing 0.01 mM of MPPD are shown in
0 150 300 450 600 750 900
Fig. 5. It is clear that the measured impedance plots are in accor-
Z r (Ω cm2 ) dance with those calculated by the used equivalent circuit model.
Table 3 contains all the impedance parameters obtained from the
500 simulation of experimental impedance data, including Rct, A and
PPD Blank n. In the Table 3 are shown also the calculated ‘‘double layer capac-
-3
10 M itance’’ values (Cdl), using the following equation [41,42]:
400 -4
10 M  1=n
-5
10 M C dl ¼ AR1n
ct ð5Þ
-6
10 M
300 and the relaxation time constant (s) of charge-transfer process
using the Eq. (7) [42]:
-Zi (Ω cm2)

200 25 Hz 10 Hz

63.3 Hz -200
4 Hz
2 Hz
a Experimentental curve
200 Hz 12.5 Hz FitResult
100
40 Hz 5 Hz 0.1 Hz

0
Ω cm2)

25 Hz
40 Hz
0 100 200 300 400 500 600 12.5 Hz
-100
Zi (Ω

Z r (Ω cm2 )
100 Hz
5 Hz
Fig. 3. Nyquist diagrams for carbon steel in 1 M HCl containing different concen- 1.5 Hz
316 Hz
trations of 1H-pyrrole-2,5-dione derivatives at 308 K.
0.1 Hz

due to the adsorption of inhibitor molecules on the metal surface


[35]. One single capacitive loop is observed on the Nyquist repre- 0
0 100 200 300
sentations which means one phenomenon only occurred. This is
Zr (Ω cm2)
in accord with observation of Bode plots representation (not
given). The capacitive loop is related to the charge transfer process
of the metal corrosion and double layer behavior. The depressed 10 3 -75
b Experiment
semi-circle in Nyquist representation is generally attributed to FitResult
the frequency dispersion as well as inhomogeneities, roughness
of metal surface and mass transport process [36]. Thus, in these sit-
10 2
uations pure double layer capacitors are better described by a
Phase angle (Degree)

transfer function with constant phase elements (CPE) to get a more -50
|Z|(Ω cm2)

accurate fit of experimental data set [37,38]. Its impedance is given


by Eq. (4):
10 1
1 n
Z CPE ¼ A ðixÞ ð4Þ

where A is proportionality coefficient (in X–1 sn cm–2), x is the -25

angular frequency and i is the imaginary number, n is an exponent 10 0


related to the phase shift and can be used as a measure of surface
irregularity. For ideal electrodes, the CPE is equal to an ideal capac-
itor when n = 1.
The impedance spectra were fitted by a simple Randles circuit 10-1 0
10 -1 100 10 1 102 10 3 10 4 105
(Fig. 4), which consists of Rs solution resistance, Rct charge transfer
Frequency (Hz)
resistance and CPE constant phase elements for the double layer. A
constant phase element (CPE) is used instead of a pure capacitor to Fig. 5. EIS Nyquist and Bode diagrams for carbon steel/1 M HCl + 0.01 mM of MPDD
compensate for non-ideal capacitive response of the interface interface: (. . .) experimental; (—) fitted data using structural model in Fig. 5.
A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584 577

Table 3
Impedance data of carbon steel in absence and presence of different concentrations of 1H-pyrrole-2,5-dione derivatives.

Inhibitor Concentration (mM) Rs (X cm2) Rct (X cm2) n A  104 (sn X1 cm2) Cdl (lF cm2) s (s) gZ (%)
Blank – 1.67 30.9 0.89 1.658 85.1 0.00262 –
MPPD 0.001 0.80 178.5 0.83 0.567 22.1 0.00397 82.8
0.01 0.95 279.8 0.81 0.541 20.5 0.00574 89.0
0.1 0.74 395.5 0.82 0.457 19.0 0.00750 92.2
1 0.71 882.4 0.84 0.336 17.9 0.01583 96.5
PPD 0.001 1.02 180.1 0.82 0.585 21.5 0.00387 82.7
0.01 1.00 263.8 0.81 0.552 20.5 0.00540 88.3
0.1 1.08 359.8 0.82 0.474 19.4 0.00698 91.4
1 0.92 467.9 0.83 0.409 18.2 0.00851 93.4

s ¼ C dl Rct ð6Þ whole investigated concentration range and the protection effi-
ciency gZ (%) increases with increasing the concentration in the
Analyse of the impedance results in Table 4 shows that the charge- both cases; the maximum inhibition efficiency was achieved at
transfer resistance value, Rct, increases with the concentration of 1 mM. The comparative study reveals that MPPD is more effective
1H-pyrrole-2,5-dione derivatives (MPPD and PPD) and reaches a than PPD and the difference in inhibitive efficiency mainly depends
maximum value of 882.4 X cm2 at 1 mM in the case of MDDP. on the type and the nature of the substituent present in the inhib-
The increase in Rct values demonstrates the improved protection itor molecule. The g (%) values obtained from the ac impedance
effects of these 1H-pyrrole-2,5-dione derivatives, and a slow cor- technique are comparable and run parallel with those obtained
roding system, due to the gradual replacement of water molecules from the weight loss measurements and the potentiodynamic
by 1H-pyrrole-2,5-dione molecules on the surface and consequently polarization method (Tables 1–3).
to a decrease in the number of active sites necessary for the corro-
sion reaction [43,44]. The value of the proportional factor A of CPE 3.4. Adsorption isotherm and surface analysis
varies in a regular manner with inhibitor concentration. However,
the values of the phase shift (n) did not vary significantly, confirm- If it is assumed, that adsorption of the organic molecules or ions
ing therefore that the charge transfer controlled dissolution mech- on the metal surface is the cause for the inhibitor action in acidic
anism of carbon steel in 1 M HCl without and with inhibitors. medium (1 M HCl), then the surface coverage (h) can be estimated
After addition of 1H-pyrrole-2,5-dione derivatives in the corrosive using from the inhibitor efficiency as h = g (%)/100, if one assumes
solution, n values (ranges from 0.81 to 0.84) decrease, when com- that the values of g (%) do no differ substantially from surface cov-
pared to that obtained in pure 1 M HCl (0.89). This shows an erage [46]. An attempt was made to find a suitable adsorption iso-
increase of the surface inhomogeneity as a result of the inhibitor’s therm, which can describe the concentration dependence of the
adsorption [45]. It is worth mentioning that the value of the relax- inhibitor efficiency. Adsorption isotherms were determined, using
ation time constant (s) slowly increases with 1H-pyrrole-2,5-dione the data of the gravimetric data, which are collected in a time
derivatives concentration as well and the time of adsorption pro- interval of 6 h, considered sufficient for adsorption equilibrium to
cess becomes therefore much higher which means a slow adsorp- be achieved [28,46]. Several adsorption isotherms (Langmuir, Tem-
tion process [46,47]. MPPD has the highest value of s, implying kin, Frumkin, . . .) were assessed and the Langmuir adsorption iso-
that MPPD was the slowest in the adsorption processes. However, therm was found to fit well with the experimental data obtained
the addition of 1H-pyrrole-2,5-dione derivatives (MPPD and PPD) for 1H-pyrrole-2,5-dione derivatives (MPPD and PPD). The Lang-
to the corrosive solution decreases the double layer capacitance muir isotherm is given by the following equation [50]:
(Cdl) (Table 3). For example, the double layer capacitance (Cdl) value
C inh 1
decrease from 85.1 lF cm–2 in the blank solution to 17.9 lF cm–2 in ¼ þ C inh ð7Þ
h K ads
the case of 1 mM of MPPD signifying that the charge and discharge
rates to the metal–solution interface is greatly decreased. This where h is the fractional surface coverage; Cinh is the inhibitor con-
shows that there is agreement between the amount of charge that centration; and Kads is the equilibrium constant of the adsorption
can be stored (i.e. capacitance) and the discharge velocity in the process. Fig. 6 showed the dependence of Cinh/h against Cinh. To
interface (s) [48]. The double layer between the charged metal sur-
face and the solution is considered as an electrical capacitor. The
adsorption of 1H-pyrrole-2,5-dione molecules on the carbon steel 1.4
surface decreases its electrical capacity because they displace the
1.2
water molecules and other ions originally adsorbed on the surface.
The decrease in this capacity with increase in MPPD and PPD con- 1
C inh /θ (mM)

centrations may be attributed to the formation of a protective layer


on the electrode surface [49]. 0.8
The inhibition efficiency, gZ (%), is calculated from Rct as
0.6
previously described [7]. The inspection of the obtained results
indicates that MPPD and PPD exhibit inhibitor properties in the 0.4

0.2 MPPD
Table 4
Thermodynamic parameters for the adsorption of 1H-pyrrole-2,5-dione derivatives PPD
on the carbon steel in 1 M HCl at 308 K. 0
0 0.2 0.4 0.6 0.8 1
Inhibitor R2 Slope Kads (M1) DGoads (kJ mol1) C inh (mM)
MPPD 0.999 1.02 5  105 43.9
PPD 0.999 1.05 1  106 45.2 Fig. 6. Langmuir isotherm adsorption model of 1H-pyrrole-2,5-dione derivatives on
the carbon steel surface in 1 M HCl.
578 A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584

Fe 2p3/2 a Cl 2p b

Intensity (a.u.)

Intensity (a.u.)
718 716 714 712 710 708
202 200 198 196
Binding energy (eV)
Binding energy (eV)

O 1s c C 1s d
Intensity (a.u.)
Intensity (a.u.)

536 534 532 530 528 294 292 290 288 286 284 282
Binding energy (eV) Binding energy (eV)

N 1s e
Intensity (a.u.)

404 403 402 401 400 399 398 397 396


Binding energy (eV)

Fig. 7. High-resolution X-ray photoelectron deconvoluted profiles of a – Fe 2p3/2, b – Cl 2p, c – O 1s, d – C 1s, and e – N 1s for MPPD treated carbon steel.
A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584 579

Fe 2p3/2 a Cl 2p b

Intensity (a.u.)
Intensity (a.u.)

720 718 716 714 712 710 708 706 704 203 202 201 200 199 198 197 196
Binding energy (eV) Binding energy (eV)

O 1s c C 1s d
Intensity (a.u.)

Intensity (a.u.)

536 535 534 533 532 531 530 529 528 292 290 288 286 284 282

Binding energy (eV) Binding energy (eV)

N 1s e
Intensity (a.u.)

403 402 401 400 399 398 397 396


Binding energy (eV)

Fig. 8. High-resolution X-ray photoelectron deconvoluted profiles of a – Fe 2p3/2, b – Cl 2p, c – O 1s, d – C 1s, and e – N 1s for PPD treated carbon steel.
580 A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584

calculate the adsorption parameters, the straight line was drawn the carbon steel surface after 6 h of immersion in the corrosive
using the least squares method. The experimental (points) and cal- solution containing 1 mM of MPPD and PPD, respectively. For com-
culated isotherms (lines) were plotted in Fig. 6. A very good fit was parison purpose, the N 1s XPS spectra were also performed from
observed with the regression coefficient up to 0.99 in the both cases the pure 1H-pyrrole-2,5-dione derivatives. The obtained XPS spec-
and the obtained line had slopes very close to unity (Table 4), which tra (Fe 2p, Cl 2p, O 1s, N 1s and C 1s core levels), given in Figs. 7–9,
suggests that the experimental data are well described by Langmuir show complex forms, which were assigned to the corresponding
isotherm and exhibit single-layer adsorption characteristic. This species through a deconvolution fitting procedure. The binding
kind of isotherm involves the assumption of no interaction between energy (BE, eV) and the corresponding quantification (%) of each
the adsorbed species on the electrode surface [44]. component are summarized in Table 5.
From the intercepts of the straight lines Cinh/h  axis, the Kads The Fe 2p spectra for carbon steel surface covered with 1H-pyr-
values were calculated and given in Table 4. Kads is related to the role-2,5-dione derivatives depict a double peak profile located at a
standard Gibbs free energy of adsorption, DGoads, according to [51]: binding energy (BE) around 711 eV (Fe 2p3/2) and 725 eV (Fe 2p1/2)
  together with an associated ghost structure on the high energy side
1 DGoads
K ads ¼ exp ð8Þ showing the subsequent oxidation of the steel surface. The decon-
55:55 RT
volution of the high resolution Fe 2p3/2 XPS spectrum consists in
where R is the universal gas constant and T is the absolute temper- three main peaks (Figs. 7a and 8a and Table 5). These peaks may
ature. The value 55.55 in the above equation is the concentration of be assigned as being due to iron in environments associated with
water in solution in mol/l. The calculated DGoads values, using Eq. (8), iron oxide and hydroxide. Indeed, the first peak located at
were also given in Table 4. Generally speaking, the adsorption type is 710.5 eV for MPPD, and 710.3 for PPD, was assigned to ferric oxy-
regarded as physisorption if the absolute value of DGoads was of the des such as Fe2O3 (i.e., Fe3+ oxide) and/or Fe3O4 (i.e., Fe2+/Fe3+
order of 20 kJ mol1 or lower. The inhibition behavior is attributed mixed oxide) [60], while that located at 711.5 eV for MPPD, and
to the electrostatic interaction between the organic molecules and 711.3 eV for PPD, may be associated to ferric hydroxide species
iron atom. When the absolute value of DGoads is of the order of such as FeOOH [61]. The last peak at 713.3 eV for MPPD, and
40 kJ mol1 or higher, the adsorption could be seen as chemisorp- 713.0 for PPD, was partly ascribed to the satellite of Fe(III) [62]
tion. In this process, the covalent bond is formed by the charge shar- and partly related to the presence of a small concentration of FeCl3
ing or transferring from the inhibitor molecules to the metal surface on the steel surface due to the testing environment [63,64]. Indeed,
[52–55]. It is difficult to distinguish between chemisorption and the signal of Cl 2p was observed on the steel surface in presence of
physisorption only based on these criteria, especially when charged 1H-pyrrole-2,5-dione derivatives (Figs. 7b and 8b and Table 5), but
species are adsorbed. The possibility of Coulomb interactions the low intensity of Cl 2p peak reflects the presence of small chlo-
between adsorbed cations and specifically adsorbed anions can ride content on the steel surface. The Cl 2p core-level is best
increase the Gibbs energy even if no chemical bond appears [53]. resolved with at least two spin–orbit-split doublets (Cl 2p1/2 and
According the obtained values of DGoads (Table 4), it can be suggested Cl 2p3/2) [63], with binding energy for Cl 2p3/2 peak lying at about
that the adsorption of the 1H-pyrrole-2,5-dione derivatives (MPPD 198.8 eV for MPPD, and 198.7 eV for PPD. This former can be attrib-
and PPD) is mainly due to chemisorption. This behavior is in accor- uted to ClAFe bond in FeCl3 as mentioned previously [63]. For PPD,
dance with the findings of other researchers in the case of the 4- one additional minor component (707.1 eV) in Fe 2p3/2 XPS spec-
amino-1,2,4-triazoles [8,28,56], the 1,3,4-thiadiazoles [57,58], trum (Fig. 8a) appear to be associated with a metallic iron
1,3,4-oxadiazoles [44], and the macrocyclic polyether compounds [65,66]. However, the metallic iron signal is not observed in the
[59]. case of MPPD, suggesting that the inhibitive layer formed in the
In order to provide insight into the chemical nature of the inhib- case of this inhibitor is thicker.
itor/carbon-steel interface, and to elucidate the adsorption mecha- The O 1s spectra for carbon steel surface after immersion in 1 M
nism of 1H-pyrrole-2,5-dione derivatives (MPPD and PPD), X-ray HCl solution containing 1H-pyrrole-2,5-dione derivatives may be
photoelectron spectroscopy (XPS) analyses were carried out on fitted into two main peaks (Figs. 7c and 8c and Table 5). These

N 1s MPPD N 1s PPD
Intensity (a.u.)
Intensity (a.u.)

403 402 401 400 399 398 404 403 402 401 400 399 398 397
Binding energy (eV) Binding energy (eV)

Fig. 9. High-resolution X-ray photoelectron deconvoluted profile N 1s for pure MPPD and PPD.
A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584 581

two peaks at 530.3 eV and 531.8 eV for MPPD, and 530.2 eV and
531.7 eV for PPD, are attributed to OAFe and HOAFe in the iron
oxides and hydroxides, respectively [67–70], in agreement with
the presence of the iron oxide/hydroxide layer as detected in the
Fe 2p3/2 spectra. The second peak at 531.8 eV for MPPD, and
531.7 eV for PPD, may be also assigned to oxygen double bonded
to carbon (C@O), which is present in 1H-pyrrole-2,5-dione deriva-
tives (MPPD and PPD) [71,72]. Its presence demonstrates that these
studied molecules are adsorbed on the steel surface. This is in
agreement with the C 1s region, where peaks due to CAH, CAC,
C@C, CAN and C@O bonds were obtained. Indeed, the correspond-
ing C 1s spectra of MPPD and PPD can be deconvoluted into three
components with different intensities (Figs. 7d and 8d and Table 5).
The first and most intense, at 285.0 eV for both inhibitors, is mainly MPPD MPPDH2+
attributed to the presence of contaminant hydrocarbons and to the
CAC, C@C and CAH bonds of MPPD and PPD [45,73], which carbon
rich structure (Fig. 1). The second component at 286.4 eV for
MPPD, and 286.5 eV for PPD, may be associated with the presence
of CAN and C@O groups in the 1H-pyrrole-2,5-dione derivatives
[74]. The last and less intense component at higher binding
energy (289.2 eV for MPPD and 288.8 for PPD eV) can be mainly
attributed to NAC@O (imide group) in the pyrrole moiety and
can also be related to the carbon atom of the C+AO [73], and so
it may be the result of the protonation of the carbonyl groups of
1H-pyrrole-2,5-dione derivatives in 1 M HCl medium.
The high-resolution N 1s spectra of the carbon steel sample
treated by 1H-pyrrole-2,5-dione derivatives depict one peak
(Figs. 7e and 8e), providing thus the evidence that the investigated
PPD PPDH2+
1H-pyrrole-2,5-dione derivatives (MPPD and PPD) were chemically Fig. 10. Lowest energy optimized conformations in solvent of neutral and diprot-
adsorbed on the steel surface. Indeed, the N 1s analyse of polished onated forms of PPD and MPPD.
carbon steel, previously described [75], was showed the absence of
the nitrogen in this substrate (spectrum not given) [75]. Focusing appearance of NAFe bond clearly indicates that complex formed
on the N 1s spectra (Figs. 7e and 8e), it is clear that the nitrogen directly on the basis of donor acceptor interactions between N
content in the case of PPD is low compared to that of MPPD on atoms of the 1H-pyrrole-2,5-dione derivatives and the vacant d
the steel surface indicating that PPD is less adsorbed on the steel orbitals of iron, forming therefore an organo-metalic complex.
surface than that MPPD. Fitting these data reveals two compo- The conclusions drawn from the XPS data provide therefore
nents; those can be assigned, respectively, to two chemically dis- direct evidence of the adsorption of the 1H-pyrrole-2,5-dione
tinct nitrogen atoms. The most intense component, located at derivatives (MPPD and PPD) onto carbon steel surface and corrob-
400.3 eV for MPPD and 400.2 eV for PPD, is assigned to NAC pyrro- orate the thermodynamic study. Furthermore, the comparison of
lic, appears at somewhat lower position than the one observed in the Fe 2p3/2 XPS results for carbon steel after immersion in 1 M
the case of pure 1H-pyrrole-2,5-dione derivatives (Fig. 9). Indeed, HCl medium containing these organic inhibitors with that for
a single peak appeared at 400.7 eV for both pure compounds untreated surface steel [75], shows that the addition of these
(MPPD and PPD) witch is attributed to the NAC pyrrolic [76]. The organic inhibitors in the corrosive solution promotes the formation
less intense component, located at 398.8 eV for MPPD and PPD, of a stable and insoluble oxide layer (Fe2O3, FeOOH) that can
may be attributed to NAFe, du to the coordination of the nitrogen reduce ions diffusion, and therefore improves the corrosion resis-
atom in the pyrrole moiety with the iron atom of steel surface. The tance of carbon steel in 1 M HCl medium.

Table 5
Binding energies (eV), relative intensity and their assignment for the major core lines observed for MPPD and PPD treated carbon steel surface.

Substrate C 1s N 1s O 1s Cl 2p Fe 2p3/2
BE (eV) Assignment BE (eV) Assignment BE (eV) Assignment BE (eV) Assignment BE (eV) Assignment
MPPD treated carbon 285.0 CAC/C@C/ 398.8 NAFe 530.3 Fe2O3/ 198.8 Cl 2p3/2 710.5 Fe2O3/Fe3O4
steel (65%) CAH (22%) (53%) Fe3O4 (63%) (16%)
286.4 CAN/C@O 400.3 CAN 531.8 C@O/ 200.3 Cl 2p1/2 711.5 FeOOH
(15%) (78%) pyrrolic (47%) FeOOH (37%) (41%)
289.2 NAC@O/ – – – – – – 713.3 Satellite of Fe(III)/
(20%) C+AO (43%) FeCl3
PPD treated carbon 285.0 CAC/C@C/ 398.8 NAFe 530.2 Fe2O3/ 198.7 Cl 2p3/2 707.1 Fe0
steel (65%) CAH (20%) (41%) Fe3O4 (65%) (2%)
286.5 CAN/C@O 400.2 CAN 531.7 C@O/ 200.3 Cl 2p1/2 710.3 Fe2O3/Fe3O4
(19%) (80%) pyrrolic (59%) FeOOH (35%) (14%)
288.8 NAC@O/ – – – – – – 711.3 FeOOH
(16%) C+AO (42%)
– – – – – – – – 713.0 Satellite of Fe(III)/
(42%) FeCl3
582 A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584

HOMO-MPPD HOMO-MPPDH 2+

LUMO-MPPD LUMO-MPPDH2+

Fig. 11. The frontier molecular orbital density distributions (HOMO and LUMO) along the MPPD (neutral and diprotonated forms).

3.5. Quantum chemical study to understand if any structural differences can be related to the
observed differences of the corrosion inhibition efficiency. In 1 M
To study the relationship between molecular structure and HCl medium, the 1H-pyrrole-2,5-dione derivatives could be easily
inhibitive effect of the investigated 1H-pyrrole-2,5-dione protonated. Consequently, all the theoretical quantum calculations
derivatives (PPD and MPPD), a quantitative structure and were performed with neutral and diprotonated forms of PPD and
activity relationship method (QSAR) was used. Indeed, we have MPPD using more energetically stable conformations in solvent
performed the Density Functional Theory (DFT) method in order (Fig. 10).

HOMO-PPD HOMO-PPDH 2+

LUMO-PPD LUMO-PPDH 2+

Fig. 12. The frontier molecular orbital density distributions (HOMO and LUMO) along the PPD (neutral and diprotonated forms).
A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584 583

Table 6 4. Conclusions
Calculated quantum chemical indices of diprotonated 1H-pyrrole-2,5-dione deriva-
tives in solvent.
Concluding the experimental part, it was clearly demonstrated
Molecule EHOMO (eV) ELUMO (eV) DE gap (eV) l (Debye) that all techniques used, especially electrochemical techniques
PDDH2+
2 4.807 0.647 4.160 0.37 and XPS, are able to characterize and to follow the corrosion inhi-
MPDDH2+2 4.794 0.573 4.221 0.13 bition process promoted by 1H-pyrrole-2,5-dione derivatives. It
was shown that the synthesized 1H-pyrrole-2,5-dione derivatives
(PPD and MPPD) exhibited excellent inhibition properties for the
carbon steel corrosion in 1 M HCl at 308 K, but MPPD has the better
The reactive ability of the inhibitor is considered to be closely performance. The results obtained from weight loss measure-
related to their frontier molecular orbitals, the HOMO and LUMO. ments, polarization curves and ac impedance study were in
A typical electron density distribution of HOMO and LUMO for reasonable agreement. The potentiodynamic polarization curves
MPPD and PPD (neutral and diprotonated forms) are shown in showed that the 1H-pyrrole-2,5-dione derivatives prevents metal
Figs. 11 and 12. From these figures, it could be seen that MPPD dissolution and also hydrogen evolution reactions (mixed-type
and PPD have similar HOMO and LUMO distributions, and the inhibitors). Ac impedance experimental data revealed frequency
diprotonation of these molecules shows a complete inversion of distribution of the capacitance, simulated as constant phase ele-
the HOMO and LUMO distribution compared to the neutral forms ment. The adsorption of 1H-pyrrole-2,5-dione derivatives was suc-
(Figs. 11 and 12). In the case of MPPDH+2 and PPDH+2, the HOMO cessfully described by the Langmuir adsorption isotherm and the
on both molecules is distributed along 1H-pyrrole-2,5-dione moi- corresponding values of DGoads revealed that the adsorption mecha-
ety, however the LUMO density is mainly focused around of phenyl nism of these inhibitors on carbon steel surface in 1 M HCl solution is
ring. This is due to the presence of nitrogen and oxygen atoms mainly due to chemisorption. XPS analysis corroborated the thermo-
together with several p-electrons on the entire molecule. Thus, dynamic results and showed evidence of the chemisorption of PPD
the unoccupied d orbitals of Fe atom can accept electrons from and MPPD on the carbon steel surface in 1 M HCl medium. QSAR
inhibitor molecule to form coordinate bond. Also the inhibitor mol- approach, using DFT method, was applied to explain the relationship
ecule can accept free electrons from the metal by using their anti- between the structure of the 1H-pyrrole-2,5-dione derivatives (PPD
bond orbitals to form stable chelates (feedback bond). and MPPD) and their inhibition effect. The computational chemistry
The calculated quantum chemical indices (EHOMO, ELUMO, DEgap results revealed that the better inhibition efficiency obtained by
and l) for diprotonated forms are reported in Table 6. Thus an MPPD can be explained by its low dipole moment value.
attempt has been made to correlate corrosion inhibition efficiency
(dependent variable) and the set of some quantum chemical indi-
ces (independent variables). EHOMO is often associated with the
References
electron donating ability of a molecule; high values of EHOMO are
likely to indicate the tendency of the molecule to donate electrons [1] G. Schmitt, Br. Corros. J. 19 (1984) 165–176.
to appropriate acceptor molecules with lower energy MO. ELUMO, [2] G. Trabanelli, Corrosion 47 (1991) 410–419.
on the other hand, indicates the ability of the molecule to accept [3] Y.I. Kuznetsov, Organic Inhibitors of Corrosion of Metals, Springer, 1996.
[4] V.S. Sastri, Corrosion Inhibitors – Principles and Applications, Wiley,
electrons [77,78]. The binding ability of the inhibitor to the metal Chichester, England, 1998.
surface increases with increasing HOMO and decreasing LUMO [5] M. Lagrenée, B. Mernari, M. Bouanis, M. Taisnel, F. Bentiss, Corros. Sci. 44
energy values. Thus, the lower the value of ELUMO, the most proba- (2002) 573–588.
[6] R. Coughlin, Corrosion inhibitors, in: J.J. Florio, D.J. Miller (Eds.), Handbook of
ble it is that the molecule would accept electrons. Moreover, the Coatings Additives, second ed., Marcel Dekker, New York, 2004, pp. 127–144.
gap between the HOMO and LUMO energy levels of the molecule [7] H. Zarrok, A. Zarrouk, B. Hammouti, R. Salghi, C. Jama, F. Bentiss, Corros. Sci. 64
is an important parameter that determines the reactivity of the (2012) 243–252.
[8] M. Tourabi, K. Nohair, M. Traisnel, J. Jama, F. Bentiss, Corros. Sci. 75 (2013)
inhibitor molecule toward the adsorption on the metallic surface. 123–133.
As DE decreases (most especially for the cationic species), the reac- [9] P. Mohan, G. Paruthimal Kalaignan, J. Mater. Sci. Technol. 29 (2013) 1096–
tivity of the molecule increases leading to increase in the inhibition 1100.
[10] N.D. Nam, Q.V. Bui, M. Mathesh, M.Y.J. Tan, M. Forsyth, Corros. Sci. 76 (2013)
efficiency of the molecule [79,80]. Dipole moment has been
257–266.
regarded as also an important index to correlate with the inhibi- [11] R. Solmaz, E. Altunbasß Sßahin, A. Doner, G. Kardasß, Corros. Sci. 53 (2011) 3231–
tion efficiency. However, there still exist the controversies with 3240.
respect to the relationship between dipole moment and inhibition [12] M.A. Deyab, Corros. Sci. 49 (2007) 2315–2328.
[13] S.A. Abd El-Maksoud, A.S. Fouda, Mater. Chem. Phys. 93 (2005) 84–90.
efficiency until now. [14] K.F. Khaled, Electrochim. Acta 48 (2003) 2493–2503.
In this study, the calculated orbital energy EHOMO and ELUMO, and [15] A. Popova, M. Christov, S. Raicheva, E. Sokolova, Corros. Sci. 46 (2004) 1333–
the band gap (DE) values for the both protonated molecules 1350.
[16] G. Avci, Colloids Surf. A 317 (2008) 730–736.
(PPDH2+ and MPPDH2+) do not exhibit important differences [17] E.A. Noor, A.H. Al-Moubaraki, Mater. Chem. Phys. 110 (2008) 145–154.
(Table 6). Such results could not explain the differences observed [18] G. Gece, S. Bilgic, Corros. Sci. 52 (2010) 3304–3308.
in the corrosion inhibition efficiency. However, a difference can [19] J. Radilla, G.E. Negron-Silva, M. Palomar-Pardave, M. Romero-Romo, M. Galvan,
Electrochim. Acta 112 (2013) 577–586.
be observed in the case of dipole moment (l) values. Effectively, [20] T. Arslan, F. Kandemirli, E.E. Ebenso, I. Love, H. Alemu, Corros. Sci. 51 (2009)
Table 6 shows a higher dipole moment (l) for PPDH2+ in compar- 35–47.
ison to MPPDH2+. Authors are not unanimous about the influence [21] L. Herrag, H. Hammouti, S. Elkadiri, A. Aouniti, C. Jama, H. Vezin, F. Bentiss,
Corros. Sci. 52 (2010) 3042–3051.
of the dipole moment on the corrosion inhibition. Some authors [22] H. Zarrok, H. Oudda, A. Zarrouk, R. Salghi, B. Hammouti, M. Bouachrine, Der
showed that an increase of the dipole moment leads to decrease Pharm. Chem. 3 (2011) 576–590.
of inhibition and vice versa, [81,82]. In contrast, the increase of [23] V. Rodeschini, N.S. Simpkins, F. Zhangi, Org. Synth. 11 (2009) 1028–1036.
[24] D.A. Shirley, Phys. Rev. B 5 (1972) 4709–4714.
the dipole moment can lead to increase of inhibition and vice versa
[25] K. Boumhara, F. Bentiss, M. Tabyaoui, J. Costa, J.-M. Desjobert, A. Bellaouchou,
[26], which could be related to the dipole–dipole interaction of A. Guenbour, B. Hammouti, S.S. Al-Deyab, Int. J. Electrochem. Sci. 9 (2014)
molecules and metal surface. The lower value of l obtained of 1187–1206.
MPPDH2+ is coherent with the first explanation suggesting that [26] F. Bentiss, B. Mernari, N. Chaibi, M. Traisnel, H. Vezin, M. Lagrenée, Corros. Sci.
44 (2002) 2271–2289.
lower value of the dipole moment in this case will favor accumula- [27] F. Bentiss, M. Lebrini, M. Lagrenée, M. Traisnel, A. Elfarouk, H. Vezin,
tion of the inhibitor in the surface layer. Electrochim. Acta 52 (2007) 6865–6872.
584 A. Zarrouk et al. / Corrosion Science 90 (2015) 572–584

[28] F. Bentiss, M. Bouanis, B. Mernari, M. Traisnel, H. Vezin, M. Lagrenée, Appl. Surf. [56] F. Bentiss, C. Jama, B. Mernari, H. El Attari, L. El Kadi, M. Lebrini, M. Traisnel, M.
Sci. 253 (2007) 3696–3704. Lagrenée, Corros. Sci. 51 (2009) 1628–1635.
[29] E. McCafferty, Corros. Sci. 47 (2005) 3202–3215. [57] F. Bentiss, M. Lebrini, M. Lagrenée, Corros. Sci. 47 (2005) 2915–2931.
[30] M. Lebrini, M. Lagrenée, H. Vezin, L. Gengembre, F. Bentiss, Corros. Sci. 47 [58] F. Bentiss, B. Mernari, M. Traisnel, H. Vezin, M. Lagrenée, Corros. Sci. 53 (2011)
(2005) 485–505. 487–495.
[31] F. Xu, J. Duan, S. Zhang, B. Hou, Mater. Lett. 62 (2008) 4072–4074. [59] M. Lebrini, M. Lagrenée, H. Vezin, M. Traisnel, F. Bentiss, Corros. Sci. 49 (2007)
[32] A.A. Aksut, W.J. Lorenz, F. Mansfeld, Corros. Sci. 22 (1982) 611–619. 2254–2269.
[33] E.S. Ferreira, C. Giancomelli, F.C. Giacomelli, A. Spinelli, Mater. Chem. Phys. 83 [60] M.A. Pech-Canul, P. Bartolo-Perez, Surf. Coat. Technol. 184 (2004) 133–140.
(2004) 129–134. [61] F.Z. Bouanis, F. Bentiss, M. Traisnel, C. Jama, Electrochim. Acta 54 (2009) 2371–
[34] H. Ashassi-Sorkhabi, M.R. Majidi, K. Seyyedi, Appl. Surf. Sci. 225 (2004) 176– 2378.
185. [62] A. Galtayries, R. Warocquier-Clérout, M.-D. Nagel, P. Marcus, Surf. Interface
[35] R. Solmaz, Corros. Sci. 52 (2010) 3321–3330. Anal. 38 (2006) 186–190.
[36] I.D. Raistrick, D.R. Franceschetti, J.R. Macdonald, in: E. Barsoukov, J.R. [63] F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, in: J. Chastain (Ed.),
Macdonald (Eds.), Impedance Spectroscopy, Theory, Experimental and Handbook of X-Ray Photoelectron Spectroscopy, Perkin-Elmer Corp.,
Applications, second ed., John Wiley & Sons, New Jersey, 2005. Minnesota, USA, 1992.
[37] A. Popova, E. Sokolova, S. Raicheva, M. Christov, Corros. Sci. 45 (2003) 33–58. [64] V.S. Sastri, M. Elboujdaini, J.R. Romn, J.R. Perumareddi, Corrosion 52 (1996)
[38] J.R. Macdonald, W.B. Johanson, in: J.R. Macdonald (Ed.), Theory in Impedance 447–452.
Spectroscopy, John Wiley & Sons, New York, 1987. [65] R. Devaux, D. Vouagner, A.M. De Becdelievre, C. Duret-Thual, Corros. Sci. 36
[39] C.H. Hsu, F. Mansfeld, Corrosion 57 (2001) 747–748. (1994) 171–186.
[40] D.A. Lopez, S.N. Simison, S.R. deSanchez, Electrochim. Acta 48 (2003) 845–854. [66] V. Di castro, S. Ciampi, Surf. Sci. 331 (1995) 294–299.
[41] H. Ma, X. Cheng, G. Li, S. Chen, Z. Quan, S. Zhao, L. Niu, Corros. Sci. 42 (2000) [67] N. Ochoa, F. Moran, N. Pébère, B. Tribollet, Corros. Sci. 47 (2005) 593–604.
1669–1683. [68] P. Bommersbach, C. Alemany-Dumont, J.P. Millet, B. Normand, Electrochim.
[42] A. Popova, M. Christov, A. Vasilev, Corros. Sci. 49 (2007) 3290–3302. Acta 51 (2005) 1076–1084.
[43] F.B. Growcock, R.J. Jasinski, J. Electrochem. Soc. 136 (1989) 2310–2314. [69] K. Aramaki, Corros. Sci. 48 (2006) 3298–3308.
[44] M. Outirite, M. Lagrenée, M. Lebrini, M. Traisnel, C. Jama, H. Vezin, F. Bentiss, [70] T. Shimura, K. Aramaki, Corros. Sci. 48 (2006) 3784–3801.
Electrochim. Acta 55 (2010) 1670–1681. [71] R.D. Boyd, J. Verran, K.E. Hall, C. Underhill, S. Hibbert, R. West, Appl. Surf. Sci.
[45] A. Popova, S. Raicheva, E. Sokolova, M. Christov, Langmuir 12 (1996) 2083– 172 (2001) 135.
2089. [72] A.G. Kannan, N.R. Choudhury, N.K. Dutta, Polymer 48 (2007) 7078–7086.
[46] M.S. Morad, Corros. Sci. 50 (2008) 436–448. [73] D. Briggs, M.P. Seah, Practical Surface Analysis by Auger and X-ray
[47] M. Lebrini, F. Bentiss, N. Chihib, C. Jama, J.P. Hornez, M. Lagrenée, Corros. Sci. Photoelectron Spectroscopy, John Wiley & Sons Ltd., Sussex, 1983
50 (2008) 2914–2918. (Section 9.4 and Appendix 2).
[48] J. Morales Roque, T. Pandiyan, J. Cruz, E. García-Ochoa, Corros. Sci. 50 (2008) [74] S.R. Kelemen, M. Afeworki, M.L. Gorbaty, Energy Fuels 16 (2002) 1450–1462.
614–624. [75] F.Z. Bouanis, F. Bentiss, S. Bellayer, M. Traisnel, J.B. Vogt, C. Jama, Mater. Chem.
[49] B. Qian, J. Wang, M. Zheng, B. Hou, Corros. Sci. 75 (2013) 184–192. Phys. 127 (2011) 329–334.
[50] S. Garai, S. Garai, P. Jaisankar, J.K. Singh, A. Elango, Corros. Sci. 60 (2012) 193– [76] P. Morales-Gil, M.S. Walczak, R.A. Cottis, J.M. Romero, R. Lindsay, Corros. Sci.
204. 85 (2014) 109–114.
[51] J. Flis, T. Zakroczymski, J. Electrochem. Soc. 143 (1996) 2458–2464. [77] D.K. Yadav, B. Maiti, M.A. Quraishi, Corros. Sci. 52 (2010) 3586–3598.
[52] M. Ehteshamzadeha, A.H. Jafari, E. Naderia, M.G. Hosseini, Mater. Chem. Phys. [78] S. Martinez, Mater. Chem. Phys. 77 (2002) 97–102.
113 (2009) 986–993. [79] K. Sayın, D. Karakasß, Corros. Sci. 77 (2013) 37–45.
[53] A.K. Singh, M.A. Quraishi, Corros. Sci. 53 (2011) 1288–1297. [80] S. Martinez, I. Stagljar, J. Mol. Struct. 640 (2003) 167–174.
[54] M.J. Bahrami, S.M.A. Hosseini, P. Pilvar, Corros. Sci. 52 (2010) 2793–2803. [81] N. Khalil, Electrochim. Acta 48 (2003) 2635–2640.
[55] M. Behpour, S.M. Ghoreishi, N. Mohammadi, N. Soltani, M. Salavati-Niasari, [82] M. Mahdavian, S. Ashhari, Electrochim. Acta 55 (2010) 1720–1724.
Corros. Sci. 52 (2010) 4046–4057.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy