You are on page 1of 17

Res Chem Intermed (2012) 38:2309–2325

DOI 10.1007/s11164-012-0547-4

Poly(4-vinylpyridine-hexadecyl bromide) as corrosion


inhibitor for mild steel in acid chloride solution

S. Belkaid • K. Tebbji • A. Mansri • A. Chetouani •

Belkheir Hammouti

Received: 18 January 2012 / Accepted: 22 March 2012 / Published online: 29 April 2012
 Springer Science+Business Media B.V. 2012

Abstract The influence of the addition of poly(4-vinylpyridine-hexadecyl bro-


mide) P4VP-Alkyl 50 % newly synthesized on the corrosion of mild steel in molar
hydrochloric acid has been investigated by weight-loss measurements combined
with linear potential scan voltammetry (I–E) and electrochemical impedance
spectroscopy (EIS). The polymer reduces the corrosion rate and the inhibition
efficiency (E %) of P4VP-Alkyl 50 % increases with its concentration and attains
95 % at 300 mg/L. E % obtained from cathodic Tafel plots, EIS, and gravimetric
methods were in good agreement. The inhibitor was adsorbed on the iron surface
according to the Langmuir adsorption isotherm model. Polarization measurements
also show that the compound acts as a cathodic inhibitor.

Keywords Steel  Poly(4-vinylpyridine)  Inhibition  Corrosion 


Hydrochloric acid

Introduction

An important practical application of such phenomena is corrosion inhibition and


numerous investigations have been performed on the inhibition of steel, iron, and
iron-based alloy by using of organic compounds [1–19]. The results show that the

S. Belkaid  A. Mansri
LAEPO, Département de chimie, Faculté des sciences, Université de Tlemcen, B.P.119,
13000 Tlemcen, Algeria

K. Tebbji  A. Chetouani  B. Hammouti (&)


LCAE-URAC18, Faculté des Sciences, Université Mohammed Premier, Oujda, Morocco
e-mail: hammoutib@yahoo.fr

A. Chetouani
Laboratoire de chimie physique, Centre Pédagogique Régional, Oujda, Morocco

123
2310 S. Belkaid et al.

inhibitory efficiency increases with inhibitor concentrations. Therefore, they


minimize the direct interaction between the metal and corrosive agents. In some
cases, the coordination of the inhibitor molecules to the surface is weak, and their
presence in the corrosive solutions required maintaining the desired concentration of
these agents to the metal surface [9–16, 20–27].
Numerous works have been devoted to the corrosion-inhibiting effect of aqueous
soluble polymers on metallic materials in acid chloride solution. It was found in
most of this research that the inhibitive power of these polymers is related
structurally to the various active centers of adsorption as cyclic rings and
heteroatoms as oxygen and nitrogen [1, 5, 8, 28–36]. The effect of derivers P4VP
compounds on the corrosion behavior of iron and steel in acidic solutions has been
well invested in our laboratories [15]. It is shown that the protective properties of
such compounds depend upon their ability to reduce the corrosion rate.
We also work on basic polymers P4VP after quaternization. The polyamine
quaternization kinetic depends on numerous effects like steric and electrostatic
phenomena or on solvating systems. These effects play different roles according to
the reagent nature. This is why a high number of literature results are divergent.
Moreover, these factors are dependent because the variation of one factor leads to
the variation of others.
The aim of this paper is to study the corrosion behavior of mild steel in molar
hydrochloric acid without and with addition of poly(4-vinylpyridine-hexadecyl
bromide) P4VP-Alkyl 50 %, newly synthesized as inhibitor for the mild steel
corrosion in 1 M HCl at 298 K. Weight-loss measurements combined with linear
potential scan voltammetry (I–E) and electrochemical impedance spectroscopy
(EIS) were performed in order to complete and to compare the results obtained.

Experimental

Synthesis of the poly(4-vinylpyridine-hexadecyl bromide) compound

A total of 0.05 mol of P4VP was dissolved in 50 mL of absolute ethanol and then
introduced into a 100-mL flask. A known amount of C16H33Br was added. The
mixture was stirred for 25 days at 70 C. After evaporation of excess solvent, the
resulting product was dissolved in dichloromethane and precipitated in a large
amount of octane. The product was dried at 70 C for 24 h. The obtained copolymer
[4VP-C16Br-50] is characterized by conductivity, 1H NMR and TG analysis, as
described below.

Conductivity measurements

The rate of quaternization of the copolymer [4VP-C16Br-50] was estimated by


conductivity measurements of Br– ions, by using a 10–3 N silver nitrate solution.
The equivalent volume of the solution of AgNO3 which acts calibrated solutions
that permit to calculate the quantities of bromide ions. This calibration technique is
realized by conduit-metre (CDM75 Tacussel).

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2311

1
H NMR spectrum (CH3OD) The Proton nuclear magnetic resonance (1H NMR)
spectra of [4VP-C16Br-X] is realized with a DMX-500 (Bruker Company,
Germany).

1
H chemical shifts (CH3OD)
0.9 ppm (CH3 of the end chain alkyl)
1.1–2.5 ppm (CH and CH2 of the backbone and the chain alkyl)
4.2–4.7 ppm (?N–CH2 methylene directly linked to the pyridinium group)
6.4–7.2 ppm (CH-b of the non-quaternized pyridine cycle)
7.3–7.8 ppm (CH-b of the quaternized pyridine cycle)
8.0–8.5 ppm (CH-a of the non-quaternized pyridine cycle)
7.3–7.8 ppm (CH-a of the quaternized pyridine cycle)

Viscosity determination

The molecular viscosimetric weight is determined in ethanol as solvent. For the


determination of the intrinsic viscosity, we used the equation established for the
poly(4-vinylpyridine) [37] using approximation concerning the charged copolymer
nature poly(4-vinylpyridine-g-hexadecyl bromide) 50 %. [g] = 4.48819 9 10–4
Mw0.590.
The molecular viscosimetric weight value calculated by using this equation is
Mv = 3.78.105 g/mole, which is slightly higher than that estimated by conducti-
metry technique, Mw = 3.12.105 g/mole.

Thermogravimetric (TG) analysis

TG experiments were carried out on a 2,950 TGA V5.4A instrument. Measurements


were performed under air atmosphere and heating rate of 10 C/min in the
temperature range between 30 and 800 C. The TG analysis allows the quantifi-
cation of the amount of C16H33Br quaternized onto P4VP copolymer. The pure
P4VP decomposes in a narrow range of temperature beyond 330 C and up to
414 C.
TG analysis allows the quantification of the amount of C16H33Br quaternized
onto P4VP copolymer. The pure P4VP decomposes in a narrow range of
temperature beyond 330 C and up to 414 C. The first loss of weight of 3–7 %
was observed between 30 and 200 C. This represents the removal of residual water
molecules and traces of solvent from the copolymers.
At 200 C, a sudden and rapid weight loss appeared, corresponding to 80.5–81 %
amount of C16H33Br. Such behavior is typical of quaternized P4VP [38] and
corresponds to rapid dequaternization by breaking of the C–N? link. This first step
was followed at about 360 C by a second decomposition step corresponding to the
slower degradation of the backbone.
The molecular weight was estimated as 312,000 g/mol by the viscosity technique
using ethanol as solvent.

123
2312 S. Belkaid et al.

Weight loss measurements

Gravimetric measurements were carried out in a double-walled glass cell equipped


with a thermostat-cooling condenser. The solution volume was 100 mL. The steel
specimens used have a square form (1.5 9 1.5 cm). Prior to immersion, the mild
steel samples were polished with different emery papers, degreased with acetone,
washed thoroughly with doubly distilled water, and finally dried in air. The
immersion time for the weight loss was 1 h at different temperatures. The
aggressive solutions (1 M HCl) were prepared by diluting analytically grade 37 %
HCl with doubly distilled water. The chemical composition of mild steel is given in
Table 1.

Electrochemical measurements

Electrochemical measurements were carried out in a conventional three-electrode


electrolysis cylindrical Pyrex glass cell. The working electrode (WE) in the form of
disc cut from steel has a geometric area of 1 cm2 and is embedded in
polytetrafluoroethylene (PTFE). A saturated calomel electrode (SCE) and a disc
platinum electrode were used as reference and auxiliary electrodes, respectively.
The temperature was thermostatically controlled at 298 ± 1 K. The WE was
abraded with silicon carbide paper (grade P1200), degreased with AR grade ethanol
and acetone, and rinsed with double-distilled water before use.
Running on an IBM compatible personal computer, the 352 Soft CorrTM III
Software communicates with EG&G Instruments potentiostat–galvanostat model
263A at a scan rate of 0.5 mV/s. Before recording the cathodic polarization curves,
the steel electrode is polarized at –800 mV for 10 min. For anodic curves, the
potential of the electrode is swept from its corrosion potential after 30 min at free
corrosion potential, to more positive values. The test solution is deaerated with pure
nitrogen. Gas bubbling is maintained through the experiments.

EIS

The EIS measurements were carried out with the electrochemical system (Tacussel),
which included a digital potentiostat model Volta lab PGZ 100 computer at Ecorr
after immersion in solution without bubbling, the circular surface of steel exposing
of 1 cm2 to the solution were used as working electrode. After the determination of
steady-state current at a given potential, sine wave voltage (10 mV) peak to peak, at
frequencies between 100 kHz and 10 mHz were superimposed on the rest potential.
Computer programs automatically controlled the measurements performed at rest

Table 1 Chemical composition of mild steel


Elements Fe C Si P Mn S Al

% en masse 99.21 0.21 0.38 0.08 0.05 0.05 0.01

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2313

potentials after 30 min of exposure. The impedance diagrams are given in the
Nyquist representation. Values of Rt and Cdl were obtained from Nyquist plots.

Results and discussion

Weight loss measurements

The corrosion rate in 1 M HCl (Wcorr) and at various concentrations of the tested
poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % compound (Wcorr) was
determined after 1 h of immersion. Values of corrosion rate and inhibition
efficiencies are given in Table 2. In the case of the weight loss method, the relation
determines the inhibition efficiency Ew and is determined by the relation:
 
Wcorr
Ew % ¼ 1  o :100 ð1Þ
Wcorr
where Wcorr and Wcorr are the corrosion rates of iron samples in the absence and
presence of the inhibitor, respectively.
It is clear that the addition of poly(4-vinylpyridine-hexadecyl bromide) P4VP-
Alkyl 50 % reduces the corrosion rate in the HCl 1 M solution. The inhibitory effect
increases then with the increase of P4VP-Alkyl 50 % concentrations. E % reaches a
maximum of 95 % at 300 mg/L for poly(4-vinylpyridine-hexadecyl bromide).
We may conclude that poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl
50 % is an inhibitor of mild steel corrosion in 1 M HCl solution and E % reaches a
maximum of 95 % at 300 mg/L.

Influence of temperature

The effect of temperature on the inhibited acid–metal reaction is very complex,


because many changes occur on the metal surface such as rapid etching, desorption
of inhibitor, and the inhibitor itself may undergo decomposition. The change of the
corrosion rate at selected concentrations of the poly(4-vinylpyridine-hexadecyl
bromide) during 1 h of immersion with the temperature was studied in 1 M HCl,
both in absence and presence of poly(4-vinylpyridine-hexadecyl bromide). For this

Table 2 Weight loss of steel in


Inhibitor Concentration (mg/L) W (mg cm–2 h–1) EW (%)
1 M HCl at different
concentrations of P4VP-Alkyl
1 M HCl - 1.06 -
50 % and the corresponding
inhibition efficiencies P4VP-Alkyl 5 0.651 39
10 0.511 52
25 0.341 68
50 0.173 84
100 0.139 87
200 0.073 93
300 0.057 95

123
2314 S. Belkaid et al.

Table 3 Variation of inhibition efficiencies according to temperature


Temperature (K) EW (%) at h EW (%) at h EW (%) at h EW (%) at h
300 mg/L 200 mg/L 100 mg/L 50 mg/L

308 93 0.93 91 0.91 86 0.86 82 0.82


313 96 0.96 94 0.94 88 0.88 83 0.83
323 95 0.95 93 0.93 89 0.89 82 0.82
333 92 0.92 90 0.90 85 0.85 79 0.79
343 91 0.91 88 0.88 82 0.82 76 0.76

purpose, gravimetric experiments were performed at different temperatures


(308–343 K); the corresponding results are given in Table 3.
The fractional surface coverage h can easily be determined from weight-loss
measurements by the ratio E(%)/100; if one assumes that the values of E(%) do not
differ substantially from h. It is clear from Table 3 that the inhibition efficiency
decreases about 3 % only with the rise in temperature in a domain (308–343 K) in
the presence of the inhibitor at different concentrations of poly(4-vinylpyridine-
hexadecyl bromide) P4VP-Alkyl 50 %. Thus, we can conclude that in the presence
of the P4VP Alkyl 50 % molecule, the corrosion rate of steel and the inhibitive
efficiency of the polymer is seen to be almost constant with the rise of temperature
indicating the nature of adsorption mechanism.
Table 3 shows that the corrosion rates increase in the absence and presence
P4VP-Alkyl 50 %. This increase was more pronounced with the rise of temperature
for blank solution. In the presence of the tested products, it was also noted that
inhibition efficiencies E % is almost independent upon the temperature (Table 3)
and the poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % keeps the
behavior of an inhibitor of iron corrosion in 1 M HCl solution at high temperature.
E % reaches a maximum of 91 % at 300 mg/L for P4VP-Alkyl 50 %.
Figure 1 shows the Arrhenius plots (the logarithm of steel corrosion rate vs.
1000/T) for the corrosion rate for both the blank and the solution of poly
(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % inhibitor. The relation can
determine the apparent activation energies:
Ea 0
 Ea
logðWcorr Þ ¼ þ A And log Wcorr ¼ þA ð2Þ
RT RT
Ea and Ea’ are the apparent activation energies with and without P4VP-Alkyl 50 %,
respectively.
Kinetic parameters, such as enthalpy and entropy of corrosion process, may be
evaluated from the effect of temperature. An alternative formulation of the
Arrhenius equation (Eq. 2) and transition state equation (Eq. 3) [39, 40]:
 0  
RT DSa DHa0
W¼ exp exp  ð3Þ
Nh R RT
where R is the universal gas constant, h is Plank’s constant, N is Avogadro’s
number, DS0a is the entropy of activation and DHa0 is the enthalpy of activation.

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2315

ln(W)
0
1M HCL
poLy 300
Poly200
-2 Poly100
Ply50

35.0 35.2 35.4 35.6 35.8 36.0 36.2


1000/T

Fig. 1 Arrhenius plots of the corrosion rate for both the blank and the solution of P4VP-Alkyl 50 %

This result permits verification of the known thermodynamic reaction between


the Ea and DHa0 as shown in Table 4:
DHa0 ¼ Ea  RT ð4Þ

Table 4 presents the calculated values of Ea, DS0a and DHa0 in inhibited and unin-
hibited acid. It is observed that the activation energy is higher in the presence of
poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % inhibitor than in its
absence.
The value of 64.13 kJ mol–1 obtained for the activation energy Ea of the
corrosion process in 1 M HCl lies in the range of the most frequently cited values,
the majority of which are grouped around 60 kJ mol–1 [27]. It can also be seen from
Table 4 that Ea decreased with increasing concentration of P4VP Alkyl 50 %, but all
values of Ea in the range of the studied concentration were higher than those in the
absence of P4VP Alkyl 50 %. The decrease in apparent activation energy occurs
from a shift of the net corrosion reaction from the uncovered surfaces.

Table 4 Thermodynamic parameters at different concentrations


Concentration Ea DH*a DS*a Ea-DHa0
of polymer (mg/L) (kJ mole–1) (kJ mole–1) (kJ mole–1 K–1) (kJ mol–1)

1 M HCl 64.133 61.432 –46.769 2.701


25 71.309 68.608 –32.930 2.701
50 71.963 69.262 –36.174 2.701
100 72.281 69.58 –37.993 2.701
200 75.909 73.208 –30.765 2.701
300 77.929 75.228 –26.990 2.701

123
2316 S. Belkaid et al.

The increase of Ea indicates the adsorption of P4VP-Alkyl 50 % compound on


the electrode surface [9, 41]. The presence of inhibitor causes a change of almost 10
kJ/mol in the values of apparent activation energy at various chemical components
of P4VP-Alkyl 50 % at different concentrations. Ea values were determined from
the slopes of these plots, we then find Ea = 64.13 and 77.92 kJ mol–l at maximum
concentration (300 mg/L of P4VP-Alkyl 50 %).
Inspection of these data from Table 4 and Fig. 3 reveals that the DHa0 values for
dissolution reaction of mild steel in 1 M HCl in the presence of P4VP Alkyl 50 % at
300 mg/L are higher (68.608–75.228 kJ mol–1) than that of in the absence of
inhibitors (61.432 kJ mol–1). This proves the endothermic nature of the mild steel
dissolution process, also suggesting that the dissolution of mild steel is slow in the
presence of an inhibitor. One can notice in Fig. 2, that Ea and DHa0 values vary in the
same way [9–12, 22–25].
One can notice that Ea and DHa0 values vary in the same way. This result allows
verification of the known thermodynamic reaction between the Ea and DHa0 as
shown in Table 4:
DHa0 ¼ Ea  RT
The negative values of entropies imply that the activated complex in the rate-
determining step represents an association rather than a dissociation step, meaning
that a decrease in disordering takes place on going from reactants to the activated
complex [9, 10, 22, 23, 39, 40].
Moreover, DS0a values are more positive in 1 M HCl solutions containing an
inhibitor than that obtained in the uninhibited solution. This behavior can be
explained as a result of the replacement process of water molecules during
adsorption of P4VP Alkyl 50 % on the steel surface. This observation is in
agreement with the findings of other workers [37, 38, 42–48].

80
78 P4VP-Alkyl
Ea et ΔH (kJ/mole)

76
74
72
70
*

68
66 Ea
64
*
62 ΔH
60
0 50 100 150 200 250 300
C (mg/l)

Fig. 2 Variation of Ea and DHa0 according to inhibitor concentrations

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2317

Adsorption isotherm

Adsorption isotherms are very important for the understanding of the mechanism of
organo-electrochemical reaction. The most frequently used isotherms are Langmuir,
Frumkin, Temin, Parsons, etc. [10–16, 22–25]. The type of the adsorption isotherms
provides information about the interaction among the adsorbed molecules them-
selves and also their interactions with the electrode surface. Several adsorption
isotherms were assessed and the Langmuir adsorption isotherm was found to be the
best description of the adsorption behavior of the studied inhibitor:
C 1
¼ þC ð5Þ
h K
 
1 DG0ads
K¼ exp  ð6Þ
55:5 RT
C is the inhibitor concentration; h the fraction of the surface covered determined by
E/100, K is the equilibrium constant of the adsorption process which is related to the
standard Gibbs energy of adsorption and DG0ads the standard free energy of
adsorption. R is the universal gas constant, T the thermodynamic temperature, and
the value of 55.5 is the concentration of water in the solution in mol/L.
Figure 3 shows the dependence of the ratio C/h as function of C at different
temperatures. Taking the fist case, the obtained plot is linear with a slope 1.02 to
close to unity at 308 K. The regression coefficient is R = 0.99989. The intercept
permit the calculation of the equilibrium constant K, which is 32525613.92.
To calculate the adsorption parameters, least-squares linear optimization
procedure was applied at different temperatures. The experimental (points) and
calculated isotherms (lines) are plotted in Fig. 3. The results are presented in
Table 5. A very good fit is observed with the regression coefficients up to 0.999,
which suggests that the experimental data are well described by Langmuir isotherm.
This result leads to evaluating DG0ads ¼ 54:252 kJ=mol at 308 K. This value of
less than –40 kJ/mol indicates that the inhibitor interacts on the steel surface by
electrostatic effect.

-6
1.2x10
308 K
-7
9.0x10
C/θ

-7
6.0x10

-7
3.0x10

0.0

0.0 4.0x10
-7
8.0x10
-7
1.2x10
-6

C (mol / L)

Fig. 3 Langmuir adsorption isotherm at 308 K

123
2318 S. Belkaid et al.

Table 5 van’t Hoff, Gibbs–


Equation of van’t Hoff Gibbs–Helmholtz
Helmholtz, and thermodynamic
parameters
DHads (kJ/mol) –9.407 –9.407

The calculated DG0ads values, using Eq. 6, were also given in Table 6. The large
negative values of DG0ads ensure the spontaneity of the adsorption process and the
stability of the adsorbed layer on the mild steel surface [49, 50] as well as a strong
interaction between the P4VP Alkyl 50 % molecules and the metal surface [39].
We recall that orders of magnitude for low energy liaisons type (van der Waals)
are between 0.5 and 30 kJ mol–1 and for high energy liaisons (covalent, metallic,
and ionic) are 100 kJ mol–1. Also, the energies above 30 kJ mol–1 are significant
and above 100 kJ mol–1 become very strong. These binding energies are relatively
high (–60 kJ/mol), which is why the chemisorptions can take place only at high
temperatures. It requires passing a significant energy barrier. Generally, values of
DG0ads up to –20 kJ mol–1 are consistent with physisorption, while those around
–40 kJ mol–1 or higher are associated with chemisorptions as a result of the sharing
or transfer of electrons from organic molecules to the metal surface to form a
coordinate type of metal bonds [40]. Here, the calculated DG0ads values are around
–55 kJ mol–1, indicating that the adsorption mechanism of P4VP Alkyl 50 % on
mild steel in 1 M HCl solution at the studied temperatures is both electrostatic-
adsorption (ionic) and chemisorptions (molecular) [51].
These results also indicate that the presence of the inhibitor increases the
inhibition efficiency without changing the adsorption mechanism. The extent of
inhibition is directly related to the increase of the adsorption layer, which is a
sensitive function of the molecular structure. Thermodynamically, DG0ads is related
0
to the standard enthalpy and entropy of the adsorption process, DHads and DS0ads ,
respectively, via Eq. (7) [39, 40, 47]:
DG0ads ¼ DHads
0
 TDS0ads ð7Þ
0

The standard enthalpy of adsorption DHads can be calculated according to the
van’t Hoff equation (Eq. 8):

Table 6 Adsorption thermodynamic parameters of steel inhibitor at different temperatures


Temperature Linear regression Slope DG0ads 0
DHads DS0ads
(K) coefficient (r) (kJ mol–1) (kJ mol–1) (J mol–1 K–1)

308 0.9999 1.04 –54.252 –9.407 –145.601


313 0.9999 1.01 –55.085 –145.936
323 0.9999 1.01 –56.698 –146.412
333 0.9999 1.04 –58.094 –146.207
343 0.9999 1.05 –59.319 –145.516

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2319

0
DHads
ln Kads ¼  þ constant ð8Þ
RT
we also like the calculations of other thermodynamic parameters by relation, Gibbs–
Helmholtz, and thermodynamic equation. The calculated values are depicted in
0
Table 5. Inspection of the data in Table 5 shows that the values of DHads by relation
van’t Hoff, Gibbs–Helmholtz, and thermodynamic equations give the same results
and its negative sign is usually characteristic of a strong interaction and a highly
0
efficient adsorption [9–16]. Since the value DHads is negative, the adsorption of
inhibitor molecules onto the mild steel surface is accompanied by a liberation
of energy (exothermic effect). We also note in Table 6 that the negative values of
DG0ads ensure and confirm the spontaneity of the adsorption process and stability of
the adsorbed layer on the steel surface [10, 12, 51]. The result is confirmed by van’t
Hoff, Gibbs–Helmholtz, and thermodynamic theory and calculations.
The obtained values of DG0ads show the regular dependence of DG0ads on
temperature (Table 6), indicating a good correlation among thermodynamic param-
eters. However, a slight decrease is noted in the absolute value of DG0ads with an
increase in temperature. This behavior is explained by the fact that the adsorption is
somewhat unfavorable with increasing experimental temperature, indicating that the
adsorbed molecule has a free reduced movement and the chemisorptions have the
major contribution in the inhibition mechanism [10, 12, 22, 43–45].
The value of DS0ads is negative, meaning that the poly(4-vinylpyridine-hexadecyl
bromide) P4VP-Alkyl 50 % molecules move freely in the bulk solution (are chaotic)
before adsorption, while as adsorption progresses, the inhibitor molecules adsorbed
onto the mild steel surface become more orderly, resulting in a decrease in entropy
[9–12, 22–24]. This order may more probably be explained by the possibility of
formation of steel-inhibitor complex on the metal surface [19, 43–45] (Scheme 1).
0
DHads and DS0ads for the adsorption of poly(4-vinylpyridine-hexadecyl bromide)
P4VP-Alkyl 50 %, on steel surface can be also deduced from Eq. (6). The
calculated DHads0
in this case is –9.407 kJ mol–1, confirming the exothermic
behavior of the P4VP-Alkyl adsorption on the steel surface. We can draw the
following conclusions: The values of DG0ads and DHads 0
are negative; this means that
the adsorption process takes place easily and the adsorption layer on steel is stable.
There is equilibrium between the phases of the two molecules and the adsorption is

Scheme 1 Structure of poly


(4-vinylpyridine-hexadecyl n
bromide) compound

N + Br -

C16 H 33

123
2320 S. Belkaid et al.

Scheme 2 Representation of the recovery of a surface monolayer of P4VP Alkyl 50 %

localized and does not give rise to the formation of a monolayer. The heat of
adsorption is independent of the rate of recovery of the solid surface (Scheme 2) and
the adsorbed molecules increase quickly before reaching a value characterizing the
complete coverage with a monolayer of molecules.
The mechanism for corrosion inhibition of mild steel in 1 M HCl by P4VP-Alkyl
50 % may be explained on the basis of adsorption behavior. The adsorption of the
P4VP Alkyl 50 % molecule.
The adsorption of the neutral P4VP-Alkyl 50 % molecules could occur due to the
formation of links between the d-orbital of iron atoms, involving the displacement
of water molecules from the metal surface and the lone electron pairs. Positively
charged species can also protect the positively charged metal surface acting with a
negatively charged intermediate, such as acid anions adsorbed on the metal surface.
It is shown that the compounds, having a higher electron density load around the
nitrogen atoms, exhibit protective properties which depends upon their ability to
reduce the corrosion rate. The difference in inhibitive efficiency mainly depends on
the length of the carbonic chain joining the hetero ring in the polymer structure.
Indeed, the number of alkyl groups greatly influences the inhibition properties and
can be related to the flexibility of the molecule, therefore influencing the adsorption
process. We can then conclude that with higher alkyl chain length, the inhibition
becomes less effective. The poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl
50 %, as other compounds studied in our laboratory, containing electronegative
function groups, p electrons conjugated and heteroatom are usually inhibitors. They
reported that the nitrogen atoms is the adsorption centers for their interaction with
the metal surface [1, 9, 10, 22, 23, 42, 49, 50, 52, 53] (Scheme 3) This fact proves
that the feedback bonds are formed between d orbitals of steel and inhibitor
molecules. Forming of feedback bonds increases the chemical adsorption of
inhibitor molecules on the steel surface and so increases the inhibition efficiency.

Polarization measurements

Current–potential characteristics resulting from cathodic and anodic polarization


curves of steel in 1 M HCl in presence of the poly(4-vinylpyridine-hexadecyl
bromide) P4VP-Alkyl 50 % at various concentrations are evaluated. The cathodic
Tafel plots of poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % are
shown in Fig. 4. Table 7 shows electrochemical parameters and inhibition
efficiencies (EI) are determined by:
 
Icorr
EI % ¼ 1  o :100 ð9Þ
Icorr

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2321

Schema 3 Schematic representation of the mode of adsorption of P4VP-Alkyl 50%

100

10
I (mA/cm )
2

1M HCl
1 5 mg/l
10 mg/l
0.1 25 mg/l
50 mg/l
0.01 100 mg/l
300 mg/l
1E-3
-800 -700 -600 -500 -400 -300 -200 -100 0
E (mV/SCE)

Fig. 4 Cathodic and anodic polarization curves of iron in 1 M HCl at different content of poly
(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 %

0
Icorr and Icorr are the corrosion current density values with and without the inhibitor,
respectively, determined by extrapolation of cathodic Tafel lines to the corrosion
potential.
From results obtained in Table 7, we note a decrease of corrosion current
densities when the increase of the concentration of the inhibitor. The inhibiting
action is more pronounced with an inhibitor and their inhibition efficiency attains a
maximum value of 91 % at 300 mg/L. The parallel Tafel curves obtained indicate
that hydrogen evolution reaction is activation-controlled and the addition of poly
(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % does not modify the
mechanism of this process [10, 12].
The anodic curves in the presence of poly(4-vinylpyridine-hexadecyl bromide)
P4VP-Alkyl 50 % show that the molecule’s effect depends on the electrode
potential. It seems that the inhibitor compounds tested have no effect from an
overvoltage higher than Ecorr.

123
2322 S. Belkaid et al.

Table 7 Polarization parameters and the corresponding inhibition efficiency of mild steel corrosion in
1 M HCl containing different concentrations of P4VP-Alkyl at 308 K
Inhibitor Concentration (mg/L) Ecorr (mV/SCE) bc (mV/dec) Icorr (lA/cm2) EI (%)

Blanc 1 M HCl –445 153 305 -


P4VP-Alkyl 5 –444 179 197 35
10 –448 184 149 51
25 –451 188 103 66
50 –455 190 63 79
100 –456 192 49 84
200 –458 188 37 88
300 –461 192 28 91

The addition of poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 %


does not affect the Ecorr value. This indicates that the inhibitor acts as cathodic
inhibitor.
Being weakly basic, the inhibitor, rapidly protonated in acid solutions, exists in
their cationic form. Due to electrostatic attraction, the inhibitors are strongly
adsorbed onto the electron-rich areas blocking the cathodic sites. This is in
agreement with the increase of the cathodic over-potential and shift of the steady
corrosion potential to less noble direction in presence of inhibitors. Effectively, the
heterocyclic-containing nitrogen in the polymer structure and electronegative
functional groups, p electrons and the conjugated double bands have been proved to
perform as very good inhibitors for the corrosion of pure iron in acidic solution [9,
10, 12, 22, 23].

EIS

To complete and to compare the results obtained previously, the corrosion behavior
of mild steel, in chloride acid solution with and without inhibitors, was investigated
by EIS at 298 K. The charge transfer-resistances (Rt) values were calculated from
the difference in impedance at low and high frequencies. The double-layer
capacitance (Cdl) was obtained at the frequency fm at which the imaginary
component of the impedance is maximal (Zi, max) by Eq. (4):
1
Cdl ¼ ð10Þ
2pfm:Rt
The inhibition efficiency obtained from the charge transfer resistance was deter-
mined by:
R0t  Rt
ERt ð%Þ ¼  100 ð11Þ
R0t
where Rt and R0 t are, respectively, the transfer-resistance values without and with
the addition of an inhibitor.

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2323

P4VP-Alkyl 1M HCl
800
5 mg/l
Zim (ohm.cm ) 10 mg/l
2
25 mg/l
600
50 mg/l
100 mg/l
400 300 mg/l

200

0 200 400 600 800 1000 1200 1400


2
Z re (ohm.cm )

Fig. 5 Impedance diagrams of iron in 1 M HCl at Ecorr with and without P4VP-Alkyl 50 % at different
concentrations

Typical Nyquist diagrams obtained in the presence of poly(4-vinylpyridine-


hexadecyl bromide) P4VP-Alkyl 50 % compound at 10–3 M are shown in Fig. 5.
The deduced impedance parameters as transfer resistance Rt (X cm2), frequency fm
(Hz), double-layer capacitance Cdl (F/cm2) and corresponding inhibition efficiency
(ERt %) are shown in Table 8. It is seen from Fig. 5 that the impedance diagrams
show semi-circles. We remark that the increase of Rt and decrease of double-layer
capacitance (Cdl) and the efficiency increases when the concentration of poly
(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % increases.
The obtained impedance diagrams show almost a semi-circular appearance,
indicating a charge transfer process mainly controls the corrosion of mild steel. In
fact, the presence of poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 %
compound enhances the value of Rt in acidic solution. Values of double-layer
capacitance are also brought down to the maximum extent in the presence of P4VP-
Alkyl 50 % and the decrease in the values of Cdl follows the order similar to that
obtained for Icorr in this study. The decrease in Cdl may be due to the adsorption of
this compound on the metal surface, leading to the formation of film from acidic
solution [13, 15, 16].

Table 8 Characteristic parameters evaluated from EIS diagrams with and without P4VP-Alkyl 50 % at
different concentrations
Inhibitor C (mg/L) RS (X cm2) Rt (X cm2) fmax (Hz) Cdl (lF/cm2) ERt (%)

1 M HCl – 2.71 115 15.82 87.51 –


P4VP-Alkyl 5 4.08 174 12.5 73.21 34
25 3.31 285 8.93 62.57 60
50 4.17 483 6.33 52.09 76
100 3.08 652 5 48.85 82
200 3.04 843 4.46 42.31 86
300 3.14 1172 4.46 30.44 90

123
2324 S. Belkaid et al.

Conclusions

• poly(4-vinylpyridine-hexadecyl bromide) P4VP-Alkyl 50 % inhibits the corro-


sion of mild steel in 1 M HCl.
• The weight loss, EIS, and polarization curves are in reasonably good agreement.
• The inhibition efficiency of poly(4-vinylpyridine-hexadecyl bromide) P4VP-
Alkyl 50 % decreases slightly with temperature and the addition of P4VP-Alkyl
leads to an increase of activation corrosion energy.
• The adsorption of P4VP-Alkyl on the steel surface in chloride acid obeys the
Langmuir adsorption isotherm model.

References

1. H. Bhandari, V. Choudhary, S.K. Dhawan, Synth. Met. 161, 753 (2011)


2. I. Elouali, B. Hammouti, A. Aouniti, Y. Ramli, M. Azougagh, E.M. Essassi, M. Bouachrine, J. Mater.
Environ. Sci. 1, 1 (2010)
3. M.B. Nemer, Y.L. Xiong, A.E. Ismail, J.H. Jang, Chem. Geol. 280, 26 (2011)
4. J. Soltis, D. Krouse, N. Laycock, Corros. Sci. 53, 2152 (2011)
5. M.A. Salam, S.S. Al-Juaid, A.H. Qusti, A.A. Hermas, Synth. Met. 161, 153 (2011)
6. L. Speckert, G.T. Burstein, Corros. Sci. 53, 534 (2011)
7. T. Stephenson, A. Kubis, M. Derakhshesh, M. Hazelton, C. Holt, P. Eaton, B. Newman, A. Hoff,
M. Gray, D. Mitlin, Energy Fuels 25, 4540 (2011)
8. S.A. Umoren, Y. Li, F.H. Wang, Corros. Sci. 53, 1778 (2011)
9. A. Chetouani, M. Daoudi, B. Hammouti, T. Ben Hadda, M. Benkaddour, Corros. Sci. 48, 2987 (2006)
10. A. Chetouani, B. Hammouti, T. Benhadda, M. Daoudi, Appl. Surf. Sci. 249, 375 (2005)
11. A. Chetouani, B. Hammouti, Bull. Electrochem. 20, 343 (2004)
12. C. Chetouani, K. Medjahed, K.E. Sid-Lakhdar, B. Hammouti, M. Benkaddour, A. Mansri, Corros.
Sci. 46, 2421 (2004)
13. A. Chetouani, A. Aouniti, B. Hammouti, N. Benchat, T. Benhadda, S. Kertit, Corros. Sci. 45, 1675
(2003)
14. A. Chetouani, B. Hammouti, Bull. Electrochem. 19, 23 (2003)
15. A. Chetouani, K. Medjahed, K.E. Benabadji, B. Hammouti, S. Kertit, A. Mansri, Prog. Org. Coat. 46,
312 (2003)
16. A. Chetouani, B. Hammouti, A. Aouniti, N. Benchat, T. Benhadda, Prog. Org. Coat. 45, 373 (2002)
17. O.K. Abiola, M.O. John, P.O. Asekunowo, P.C. Okafor, O.O. James, Green Chem. Lett. Rev. 4, 273
(2011)
18. D. Bankiewicz, E. Alonso-Herranz, P. Yrjas, T. Lauren, H. Spliethoff, M. Hupa, Energy Fuels 25,
3476 (2011)
19. M. Behpour, S.M. Ghoreishi, N. Mohammadi, M. Salavati-Niasari, Corros. Sci. 53, 3380 (2011)
20. K. Laarej, M. Bouachrine, S. Radi, S. Kertit, B. Hammouti, J. Chem. 7, 419 (2010)
21. K. Barouni, L. Bazzi, R. Salghi, M. Mihit, B. Hammouti, A. Albourine, S. El Issami, Mater. Lett. 62,
3325 (2008)
22. M. Bouklah, A. Attayibat, S. Kertit, A. Ramdani, B. Hammouti, Appl. Surf. Sci. 242, 399 (2005)
23. A. Ouchrif, M. Zegmout, B. Hammouti, A. Dafali, M. Benkaddour, A. Ramdani, S. Elkadiri, Prog.
Org. Coat. 53, 292 (2005)
24. M. Zerfaoui, B. Hammouti, H. Oudda, M. Benkaddour, S. Kertit, Bull. Electrochem. 20, 433 (2004)
25. M. Zerfaoui, H. Oudda, B. Hammouti, S. Kertit, M. Benkaddour, Prog. Org. Coat. 51, 134 (2004)
26. M.A. Amin, K.F. Khaled, Q. Mohsen, H.A. Arida, Corros. Sci. 52, 1684 (2010)
27. N. Bertrand, C. Desgranges, D. Poquillon, M.C. Lafont, D. Monceau, Oxid. Met. 73, 139 (2010)
28. K. Kamaraj, V. Karpakam, S. Sathiyanarayanan, G. Venkatachari, Mater. Chem. Phys. 122, 123
(2010)

123
Poly(4-vinylpyridine-hexadecyl bromide) as corrosion inhibitor 2325

29. G. Bereket, E. Hur, Prog. Org. Coat. 65, 116 (2009)


30. Z.H. Guo, K. Shin, A.B. Karki, D.P. Young, R.B. Kaner, H.T. Hahn, J. Nanopart. Res. 11, 1441
(2009)
31. N.A. Negm, M.F. Zaki, J. Dispers. Sci. Technol. 30, 649 (2009)
32. R.M. Coleman, J. Fuoss, Am. Chem. Soc. 11, 5472 (1955)
33. J. Morcellet, C. Loucheux, Makromol. Chem. 176, 315 (1975)
34. S. Pradny, Sevcik. Makromol. Chem. 188, 2875 (1987)
35. S. Dragan, I. Petrariu, M. Dima, J. Polym. Sci. Polym. Chem. Ed. 38, 2881 (1981)
36. M. Prady, S. Sevcik, Makromol. Chem. 187, 2191 (1986)
37. E. Choukchou-B, I. Benabadji, J. François, A. Mansri, Eur. Polym. J. 39, 297 (2003)
38. A. Mansri, Y. Frère, C. Chevino, Ph. Gramain, Des. Monomers Polym. 3, 55 (2000)
39. F.Z. Bouanis, C. Jama, M. Traisnel, F. Bentiss, Corros. Sci. 52, 3180 (2010)
40. F. Bentiss, C. Jama, B. Mernari, H. El Attari, L. El Kadi, M. Lebrini, M. Traisnel, M. Lagrenee,
Corros. Sci. 51, 1628 (2009)
41. S. Kertit, B. Hammouti, Appl. Surf. Sci. 93, 59 (1996)
42. F. Gesmundo, F. Viani, Mater. Chem. Phys. 20, 513 (1988)
43. A. Doner, R. Solmaz, M. Ozcan, G. Kardas, Corros. Sci. 53, 2902 (2011)
44. A. Doner, G. Kardas, Corros. Sci. 53, 4223 (2011)
45. M. Dudukcu, Materials and Corrosion-Werkstoffe Und Korrosion 62, 264 (2011)
46. B.D. Mert, M.E. Mert, G. Kardas, B. Yazici, Corros. Sci. 53, 4265 (2011)
47. M. Lebrini, M. Traisnel, M. Lagrenee, B. Mernari, F. Bentiss, Corros. Sci. 50, 473 (2008)
48. K. Tebbji, H. Oudda, B. Hammouti, M. Benkaddour, S.S. Al-Deyab, A. Aouniti, S. Radi,
A. Ramdani, Res. Chem. Intermed. 37, 985 (2011)
49. D.P. Schweinsberg, V. Ashworth, Corros. Sci. 28, 539 (1988)
50. A.B. Tadros, B.A. Abdelnabey, J. Electroanal. Chem. 246, 433 (1988)
51. F. Bentiss, M. Lagrenee, M. Traisnel, B. Mernari, H. Elattari, J. Appl. Electrochem. 29, 1073 (1999)
52. V. Hluchan, B.L. Wheeler, N. Hackerman, Werkstoffe Und Korrosion-Mater. Corros 39, 512 (1988)
53. K. Juttner, W.J. Lorenz, Werkstoffe Und Korrosion-Mater. Corr. 39, 561 (1988)

123

You might also like