You are on page 1of 12

Journal of Electroanalytical Chemistry 791 (2017) 83–94

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry

journal homepage: www.elsevier.com/locate/jelechem

Inhibition of the corrosion of X70 and Q235 steel in CO2-saturated brine


by imidazoline-based inhibitor
Huan-huan Zhang a, Kewei Gao a,b, Luchun Yan a, Xiaolu Pang a,⁎
a
Department of Materials Physics and Chemistry, University of Science and Technology Beijing, Beijing 100083, China
b
Key Laboratory for Environmental Fracture (MOE), University of Science and Technology Beijing, Beijing 100083, China

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

Article history: Potentiodynamic polarization, electrochemical impedance spectroscopy, and weight loss methods were used to
Received 23 September 2016 study the inhibition of the corrosion of X70 and Q235 steels in CO2-saturated 3 wt% NaCl solution with various
Received in revised form 20 February 2017 concentrations of imidazoline-based inhibitor. The inhibition efficiencies for X70 and Q235 reached 97.8% and
Accepted 26 February 2017
98.6%, respectively, with sufficient concentration of inhibitor. The results indicated that, the inhibition efficiencies
Available online 1 March 2017
for X70 and Q235 were dramatically different with low concentration of inhibitor. Compared with X70, Q235 ex-
Keywords:
hibited higher inhibition efficiency and stronger adsorption ability because of the larger amount and more homo-
Inhibition efficiency geneous distribution of carbides. The adsorption of the inhibitor on the metal surface consisted of a mix of
Langmuir adsorption isotherm physisorption and chemisorption and was observed to follow the Langmuir adsorption isotherm.
Chemical adsorption © 2017 Elsevier B.V. All rights reserved.
Potential of zero charge

1. Introduction without loss of efficiency (99.51%). Okafor et al. [11] observed that the
extent of inhibition on the corrosion of mild steel was dependent on
Corrosion has a critical economic impact in the petroleum industry. the concentration of 2-undecyl-1,1-dicarboxylethyl quaternary
The most predominant corrosion is the one caused by CO2, which result- imidazoline (CQI) and the temperature. These researchers proposed
ed in losses of millions of dollars per year [1–4]. From the viewpoint of that CQI inhibited the corrosion reaction via chemical adsorption on
materials cost, the use of carbon steel for oil and gas pipelines is the the metal/solution interface in 3% NaCl solution and via a combination
most convenient option. X70 and Q235 steels are widely used as pipe- of chemical and columbic attraction in 3% Na2SO4 solutions.
line materials in the petroleum industry because of their cost-effective- It is generally accepted that imidazoline derivatives inhibit corrosion
ness. However, carbon steel is prone to corrosion in a CO2 environment. via adsorption at the metal–solution interface, subsequently forming a
When the environment is too aggressive for carbon steel, the use of in- protective layer to retard the corrosion process of the steels [12–14].
hibitors is the most convenient option to reduce corrosion problems [5– The adsorption ability of inhibitors on a metal surface depends on the
7]. Imidazoline and its derivatives have been widely used to protect the inhibitor molecular structure, inhibitor concentration, electrolyte chem-
pipelines from CO2 corrosion in petroleum industries because of their istry, and nature and surface charge of the metal [15–17]. The potential
high inhibition effectiveness and environmental friendliness [8–11]. of zero charge (Epzc) describes the condition when the electric charge
Ortega-Sotelo et al. [8] observed that carboxyamido imidazoline in- density on a surface is zero, which is a concept relating to the phenom-
hibitor reached its highest corrosion protection for X70 steel in a CO2 enon of adsorption. The surface charge of a metal can be determined by
environment at a concentration of 8.1 × 10−5 mol/l, lower or higher in- the position of the open-circuit potential with respect to the respective
hibitor concentrations than 8.1 × 10−5 mol/l increased the corrosion Epzc [18].
rate. Aiad et al. [9] noted that, for X52 steel in a CO2 environment, the Steels may have different microstructures depending on the chemi-
localized corrosion attack severity increased with increasing CO2 pres- cal composition and on the fabrication process. These various micro-
sure from 10 to 60 bar and decreased with increasing inhibitor concen- structural components (ferrite, pearlite) can affect not only the
tration from 10 to 100 ppm. Farelas et al. [10] observed that 1-(2- corrosion resistance of the material but also the performance of inhibi-
aminoethyl)-2(heptadec-8-enyl)-bis-imidazoline formed a more com- tors. Okafor et al. [17] noted that an imidazoline-based inhibitor had dif-
pact inhibitor layer on mild steel in CO2-saturated 3 wt% NaCl solution, ferent inhibition efficiencies for N80 and P110 carbon steel; the authors
compared with 1-(2-hydroxyethyl)-2(heptadec-8-enyl)-imidazoline; pointed out that the results were mainly ascribed to the difference in
the former inhibitor could be used at a lower concentration (10 ppm) grain size and structural modification, however, they did not discuss
the issue in detail. Oblonsky et al. [19] observed that
⁎ Corresponding author. octadecyldimethylbenzylammonium chloride (ODBAC) adsorbed
E-mail address: pangxl@mater.ustb.edu.cn (X. Pang). strongly onto the ferrite–pearlite microstructure and weakly onto the

http://dx.doi.org/10.1016/j.jelechem.2017.02.046
1572-6657/© 2017 Elsevier B.V. All rights reserved.
84 H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94

martensitic microstructure; the results were attributed to the more sta-


ble passive film on the martensitic steel preventing optimal adsorption
of the inhibitor. Lopez et al. [20] noted that the presence of 100 ppm
benzimidazole improved the corrosion resistance of the annealed sam-
ples (ferrite + homogeneous distribution of globular carbides), where-
as the opposite effect was observed for the Q&T samples
(ferrite + laminar carbides). Several authors have studied the effect of
steel microstructure on the inhibition behavior; however, there is not
a general agreement on this issue.
In the present study, the inhibition performance of an imidazoline-
based inhibitor on widely used pipeline steels was investigated in
CO2-saturated 3 wt% NaCl solution. X70 and Q235 steel, mostly used
as pipeline materials. Despite of their great importance in field applica-
tions, the relationship between the microstructure of the steels and the
inhibition performance of the inhibitor has barely been considered. The
main driving force behind this study was to explore the differences be-
tween the inhibition efficiencies of an imidazoline-based inhibitor for
different pipeline steels. Potentiodynamic polarization, electrochemical
impedance spectroscopy (EIS), and weight loss methods were used to
study the inhibition of the corrosion of the two steels in CO2-saturated
3 wt% NaCl solution for various concentrations of the imidazoline-
based inhibitor. This work investigated the adsorption thermodynamics
of the inhibitor as well as the activation energies. In addition, the Epzc
was employed to study the surface charge of the metal. This work also
aimed to reveal the corrosion inhibition mechanism of the
imidazoline-based inhibitor.

Fig. 1. Microstructural images of X70 steel (a) and Q235 steel (b).

2. Materials and methods


2.2. Electrochemical studies
2.1. Materials and medium
The samples were embedded in epoxy resin with an exposed area of
The chemical compositions of the X70 and Q235 steels, listed in 1 cm2. Before the measurements, the samples were abraded with
Table 1, were analyzed by the inductive coupled plasma (ICP). Optical 800 grit SiC paper, washed with distilled water, and rinsed with ethanol,
microscope (Leica DM 2500 M) and scanning electron microscopy respectively. Electrochemical experiments were performed using an
(SEM; Zeiss Evo18, Germany) were respectively employed to obtain mi- electrochemistry workstation (CS310, Wuhan Corrtest Instrument Co.,
crographs of the microstructures. Before the microstructural analysis, Ltd., China) in a typical three-electrode setup under atmospheric pres-
the X70 and Q235 samples were abraded using up to 2000 grit SiC sure. Three-electrode jacketed test cells with a working volumes of
paper and polished with a 2.5 μm diamond abrasive. Then, the samples 250 ml were used as the containers. An Ag/AgCl electrode was used as
were immersed in an etchant (3 ml of 68% analytical nitric acid and the reference electrode, and platinum foil was used as the counter elec-
98 ml of anhydrous, ethyl alcohol) for 10 s, washed with distilled trode. The samples of X70 and Q235 samples were used as the working
water and rinsed with alcohol, and dried in a stream of hot air. The mi- electrodes.
crostructures of the steels are depicted in Fig. 1. The Q235 steel sample The measurements were performed at 313 K and 333 K without the
has a pronounced homogeneous distribution of carbides, whereas only inhibitor and with various concentrations of the inhibitor. The inhibitor
a few carbides are observed in the X70 specimen. The main reason for was added at the beginning of the experiments. During the electro-
this discrepancy is that the carbon content of the Q235 steel is much chemical measurements, electrolyte solution was bubbled using a 1-
higher. bar CO2 flow with a high purity (99.99%) to ensure saturation through-
The samples were machined to dimensions of out the experiment.
10 mm × 10 mm × 3 mm. The corrosion medium was 3 wt% NaCl solu- In all the experiments, the working electrode samples were allowed
tion containing various concentrations of an imidazoline-based inhibi- to reach their stable open-circuit potentials (Eocp), which were reached
tor. The NaCl solution was prepared using analytical grade reagents after for 1 h. Electrochemical impedance spectroscopy (EIS) measure-
and distilled water. The imidazoline-based product used as the inhibitor ments were performed in a frequency range of 100,000 Hz to 0.1 Hz
was previously synthesized in our laboratory [21], and its molecular using a perturbation of 10 mV amplitude at the Eocp. The potentiody-
structure is shown in Fig. 2. Because of their biodegradability, namic polarization was performed in the potential range of −250 mV
imidazoline-based inhibitors are ‘green’ inhibitors and safety non-toxic- to +300 mV vs. Eocp at a scan rate of 0.5 mV/s. The polarization curves
ity [22]. The solution was de-aerated with high purity N2 before the were also obtained after 1 h immersion at the open-circuit potential.
tests. In this study, the inhibitor was used in a concentration range of Each test was repeated three times to ensure the reproducibility of the
4.8780 × 10−6 mol/l to 4.8780 × 10−5 mol/l. results.

Table 1
Chemical compositions of X70 and Q235 steel (wt%).

Element C Si Mn Mo Cr Nb Ti V Ni Cu Fe

X70 0.063 0.18 1.5 0.17 0.019 0.025 0.016 0.026 0.12 0.11 Balance
Q235 0.18 0.25 0.5 Balance
H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94 85

Table 2
Experimental conditions for weight loss measurements.

Test Temperature Pressure Samples Inhibitor Immersion


(K) (1 MPa) (×10−5 mol/l) time (h)

1 333 1 6 samples of X70 0 96


2 333 1 6 samples of Q235 0 96
3 333 1 6 samples of X70 4.878 96
4 333 1 6 samples of Q235 4.878 96
5 333 1 3 samples of 0.4878 96
X70 + 3 samples of
Q235

equations [28]:

Fig. 2. Molecular structure of the synthesized inhibitor. 8:76  104 Δm


corrosion rate ¼ ð4Þ
Sρt

The inhibition efficiency (IE) of the inhibitor was determined using corrosion rate0 −corrosion rateinh
IE ¼  100% ð5Þ
the following equations [23]: corrosion rate0

where S is the electrode area (cm2), t is the immersion time (96 h), Δm
i0corr −icorr
IE ¼  100% ð1Þ is the weight-loss (g), and ρ is the steel density (g/cm3); corrosion rate0
i0corr and corrosion rateinh are the corrosion rates (mm/y) of the samples in
the absence and presence of the inhibitor, respectively.
After the corrosion tests, surface morphologies analysis of the sam-
Rct −R0ct
IE ¼  100% ð2Þ ples in the solution without and with inhibitor was performed using
Rct
SEM.

where i0corr and icorr are the corrosion current densities without and with 2.4. Potential of zero charge measurements
the inhibitor, respectively; R0ct and Rct are the values of the charge trans-
fer resistance without and with the inhibitor, respectively. Potential of zero charge is a concept of fundamental importance in
The values of surface coverage (θ) of the inhibitor was calculated surface science that relates to the phenomenon of adsorption and de-
using the following equation [24]: scribes the condition when the electric charge density on a surface is
zero [18]. EIS measurements were used to evaluate the potential of
icorr zero charge (Epzc) of the steels in CO2-saturated 3 wt% NaCl solution
θ ¼ 1− ð3Þ without and with the addition of 4.8780 × 10− 6 mol/l inhibitor at
i0corr
333 K. The electrochemical impedance spectra were recorded after 1 h
of immersion at different potentials with an AC amplitude of 10 mV,
The adsorption isotherm can provide useful information about the by scanning the frequency from 100,000 Hz to 0.1 Hz. EIS measure-
inhibition mechanism [25–26]. The inhibitor concentration (C) and sur- ments were performed at potentials from −200 mV to + 200 mV vs.
face coverage θ (obtained from polarization curves) were employed to Eocp, and the measurements were recorded at intervals of 20 mV. The
determine the best-fitting adsorption isotherm and explain the adsorp- differential capacities (CPEdl) were determined by fitting the measured
tion process of the inhibitor on the sample surfaces. data. In general, the minimum value on the CPEdl vs. applied potential
Eapp curve is considered as the Epzc value of the electrode.
2.3. Weight loss measurements
3. Results and discussions
Weight loss measurements were conducted in a high-temperature
and high-pressure magnetic drive autoclave. The sample surfaces 3.1. Polarization curves
were abraded with a series of SiC paper (up to 800 grit), washed with
distilled water, and rinsed with alcohol. The samples were then weight- Fig. 3 presents the open circuit potential plots of the X70 and Q235
ed on an electronic balance with a precision of 0.1 mg and stored until steels in CO2-saturated 3 wt% NaCl solution with various concentrations
use. Before the experiments, the samples were fixed to a of the inhibitor at studied temperatures.
polytetrafluoroethylene holder with an exposed area of Fig. 4 presents the polarization curves of the X70 and Q235 steels in
10 mm × 10 mm. CO2-saturated 3 wt% NaCl solution containing various concentrations of
The test medium was introduced into the autoclave at room temper- the inhibitor at various temperatures.
ature and ambient pressure. The test samples were completely im- The relevant electrochemical parameters are listed in Table 3, in-
mersed in 1000 ml electrolyte solution. Prior to the experiments, the cluding the Eocp, corrosion current density (icorr), corrosion potential
test solution and autoclave were purged with a 1-bar CO2 (99.99%) (Ecorr), anodic Tafel slope (Ba), and cathodic Tafel slope (Bc). The IE
flow for 120 min to remove oxygen within the system. Experimental and θ values are also included in Table 3.
conditions are listed in Table 2. The data listed in Table 3 indicated that, with increasing the concen-
After the immersion experiments, the six samples were removed tration of inhibitor, icorr decreased and IE increased for both steels at the
and cleaned with distilled water and alcohol. Three of the samples studied temperatures. The curves in Fig. 4 indicated that both the anodic
were pickled using a chemical cleanup method to remove the corrosion and cathodic current densities decreased distinctly with the addition of
product scale (GB/T 16545-1996, China, idt ISO 8407, 1991) [27] and the inhibitor for X70 and Q235 steels. The values of Bc and Ba in Table 3
then weighted on a balance with a precision of 0.1 mg. The general cor- change with inhibitor concentration, indicating that the inhibitors con-
rosion rate and inhibition efficiency were calculated using the following trolled both the anodic and cathodic corrosion reactions. In addition,
86 H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94

Fig. 3. Open circuit potential-time plots for X70 and Q235 steel in CO2-saturated 3 wt% NaCl solution with various concentrations of inhibitor at studied temperatures: (a) X70-313 K; (b)
X70-333 K; (c) Q235-313 K; (d) Q235-333 K.

Fig. 4. Polarization curves for X70 and Q235 steel in CO2-saturated 3 wt% NaCl solution containing various concentrations of inhibitor at various temperatures: (a) X70-313 K; (b) X70-
333 K; (c) Q235-313 K; (d) Q235-333 K.
H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94 87

Table 3
Electrochemical parameters, inhibition efficiency, and surface coverage for X70 and Q235 steels in CO2-saturated 3 wt% NaCl containing different concentrations of the inhibitor.

T, K C, 10−5 mol/l Eocp, mV Ecorr, mV (mV) icorr, μA/cm2 (μA/cm2) Ba, mV/dec −Bc, mV/dec IE, % θ

X70
313 0 −672 ±9 −667 ±4 144.7 ± 5.8 152 ± 8 572 ± 11 –
0.4878 −651 ± 11 −629 ±4 61.1 ± 4.7 104 ± 6 361 ± 9 57.8 0.588
1.2195 −624 ±9 −631 ±4 11.4 ± 1.5 84 ± 8 182 ± 6 92.2 0.922
2.4390 −610 ±5 −613 ±4 7.7 ± 0.7 73 ± 7 177 ± 8 94.7 0.957
4.8780 −605 ±8 −606 ±4 3.7 ± 0.9 77 ± 5 151 ± 8 97.4 0.974
333 0 −667 ±9 −663 ±5 225.6 ± 6.9 102 ± 5 451 ± 9 –
0.4878 −634 ±5 −639 ±7 88.1 ± 5.4 96 ± 5 339 ± 6 60.9 0.619
1.2195 −629 ±7 −605 ±5 16.9 ± 3.2 74 ± 6 180 ± 7 92.5 0.935
2.4390 −612 ±9 −608 ±4 8.4 ± 1.3 70 ± 4 148 ± 5 96.3 0.963
4.8780 −593 ± 12 −610 ±5 4.9 ± 0.7 66 ± 4 153 ± 4 97.8 0.988

Q235
313 0 −681 ±6 −685 ±4 128.2 ± 5.3 147 ± 8 673 ± 12 – –
0.4878 −637 ±9 −627 ±5 10.5 ± 0.4 97 ± 8 206 ± 6 91.9 0.929
1.2195 −613 ±8 −628 ±5 7.7 ± 0.8 79 ± 5 216 ± 4 94.0 0.940
2.4390 −609 ±8 −630 ±4 3.3 ± 0.3 85 ± 4 223 ± 7 97.5 0.985
4.8780 −602 ± 11 −631 ±4 2.2 ± 0.4 72 ± 4 209 ± 8 98.3 0.983
333 0 −681 ±4 −682 ±5 219.8 ± 4.9 136 ± 9 393 ± 9 – –
0.4878 −630 ±9 −625 ±5 12.7 ± 1.2 58 ± 5 173 ± 5 94.2 0.942
1.2195 −641 ±6 −644 ±5 8.4 ± 0.9 54 ± 4 168 ± 4 96.2 0.962
2.4390 −629 ±8 −646 ±5 4.6 ± 0.4 61 ± 4 170 ± 4 97.7 0.98
4.8780 −620 ±5 −648 ±4 3.0 ± 0.4 59 ± 5 160 ± 6 98.6 0.99

Eocp shifted in the positive direction for both steels with the addition of inhibitor concentration at the studied temperatures. The results may be
the inhibitor, as shown in Fig. 3. In the present study, the shift in Eocp attributed to the increase in the surface coverage by the inhibitor, which
values was b85 mV, suggesting that the synthesized inhibitor acted as led to an increase in the inhibition efficiency (Table 4). The decrease in
a mixed-type inhibitor [29]. The shift in Eocp can be explain by the active CPEdl suggests that the organic inhibitor molecules may adsorb at the
sites blocking effect that occurs when an inhibitor is added [30]. For CO2 steel surface by replacing the water molecules, decreasing the extent
corrosion, the anodic reaction is the oxidation of iron, and the cathodic of metal dissolution [36]. In addition, the Rct values increased with in-
reaction is the reduction of hydrogen [31]. Therefore, the inhibitor could creasing temperature, resulting in an increase in the inhibition efficien-
reduce the anodic dissolution of steels as well as retard the cathodic hy- cies. These results agree with those obtained from the polarization
drogen evolution reaction. The results indicate that the inhibitor exhib- curves. The results for Q235 steel in Table 4 show the same trend as
ited an excellent inhibiting effect for the corrosion of X70 and Q235 the experimental results discussed above.
steels at the studied temperatures. Noted that the inhibition efficiencies Table 4 indicates that, with the same inhibitor concentration at
for X70 steel were always lower than those for Q235 steel under the 313 K, the Rct values complied with the order of Q235 N X70, and the in-
same conditions. hibition efficiencies followed the order of Q235 N X70. The maximum IE
values were 95.8% and 97.0% for X70 and Q235, respectively. The inhib-
3.2. Electrochemical impedance spectroscopy (EIS) results itor behaved more efficiently inhibited the corrosion of Q235 steel.
When the experiments were performed at 333 K, the results showed
Nyquist and Bode plots for the X70 and Q235 steel in CO2-saturated the same trend as discussed above. These results again confirmed that
3 wt% NaCl solution without and with various concentrations of the in- the inhibitor in this study exhibited good inhibitive performance for
hibitor after 1 h of immersion time at various temperatures are present- X70 and Q235 in CO2-saturated 3 wt% NaCl solution. The EIS results
ed in Fig. 5. were consistent with the polarization measurements.
It is apparent that there is only one semicircle for the X70 and Q235
samples exposed to CO2-saturated 3 wt% NaCl solution without and 3.3. Adsorption isotherm
with various concentrations of inhibitor in Fig. 5(a), (d), (g), (j) and
(n). In general, a single capacitive loop implies one time constant, as ob- The type of adsorption isotherm is determined by inhibitor concen-
served in Fig. 5(c), (f), (i), (m), and (q), which is most likely associated tration (C) and θ. In this study, the θ values of the inhibitor were evalu-
with the charge transfer process [28,32–35]. The equivalent circuit ated based on polarization results and are listed in Table 3.
model used to fit the experimental results is shown in Fig. 6, where Rs To describe the adsorption of the imidazoline-based inhibitor on X70
represents the electrolyte resistance, Rct represents the charge transfer and Q235 steels surfaces, Temkin, Frumkin, Freundlich, and Langmuir
resistance, and CPEdl represents the constant phase element of the dou- isotherms were tested. However, the best agreement was obtained for
ble layer. In this study, CPE was used to replace a pure double layer ca- the Langmuir isotherm. The Langmuir adsorption isotherm is expressed
pacitance to obtain more accurate fitting results [34]. by the following modified equation [37]:
The double layer capacitance was estimated using the following
equation [1,26]: C f
¼ þ fC ð7Þ
θ K
n−1
CPEdl ¼ Y 0 ð2π f max Þ ð6Þ
where K is the equilibrium constant of the inhibitor molecular adsorp-
where fmax is the frequency of the imaginary component maximum tion process, and f is the slope of the straight line.
in the Nyquist plot and Y0 represents magnitude of the CPE, −1 ≤ n ≤ 1. The plots of C / θ vs. C produced straight lines at 313 K and 333 K, as
In all cases, the fitted data had an average error lower than 10%. The observed in Fig. 7. Table 5 presents the correlation coefficients (R2,
characteristic parameters are listed in Table 4. which is employed to determine the best-fitting isotherm) and slopes.
Using the results of X70 steel as an example, as observed in Table 4, The derivation of the slopes from unity can be attributed to the interac-
the Rct values increased and the CPEdl values decreased with increasing tions between the adsorbed species on the steels surfaces and the
88 H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94

change of the adsorption heat with increasing surface coverage [37]. The K reflects strong adsorption ability of the inhibitor molecules onto the
K values were determined from the intercepts of the lines in Fig. 7. metal surface [25,38]. In addition, the Q235 steel had larger K values
The data listed in Table 5 indicated that the K values increased with at the studied temperatures, indicating that the adsorption ability of
increasing temperature. The K value represents the binding force of in- the inhibitor onto Q235 was stronger than that onto X70. The K value
hibitor molecules adsorbing onto the metal surface, and a large value of is related to the standard free energy of adsorption (ΔG0ads) by the

Fig. 5. Nyquist and Bode plots of X70 and Q235 steel in CO2-saturated 3 wt% NaCl solution without and with different concentrations of inhibitor at various temperatures: (a)–(c), X70,
313 K; (d)–(f), X70, 333 K; (g)–(i), Q235, 313 K; (j)–(m), Q235, 333 K; (n)–(q), without inhibitor.
H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94 89

Fig. 5 (continued).

following equation [39]: The calculated values of ΔG0ads are listed in Table 5. The negative
  values obtained indicate that the inhibitor spontaneously adsorbed
1 −ΔG0ads onto the surfaces of the X70 and Q235 steels at the studied tempera-
K¼ exp ð8Þ tures. The absolute values of ΔG0ads for the X70 and Q235 steels were
55:5 RT
larger than 40 kJ/mol, indicating that the adsorption mechanism of the
where 55.5 is the concentration of water in solution expressed in mole, inhibitor onto X70 and Q235 was mainly chemical adsorption at 313 K
R is the molar gas constant, and T is the corresponding temperature in K. and 333 K [40–41].
90 H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94

The heat of adsorption (Qads) was evaluated from θ and temperature


[25], as follows:

     
θ2 θ1 T1T 2
Q ads ¼ 2:303R log − log ð10Þ
1−θ2 1−θ1 T 2 −T 1

where θ1 and θ2 are the values of θ at temperatures T1 (313 K) and T2


Fig. 6. Equivalent circuit model used in the fitting of impedance data for the studied
(333 K).
inhibitor.
The calculated Qads values of the X70 and Q235 steels were
7.34 kJ/mol and 11.60 kJ/mol, respectively. The obtained positive Qads
3.4. Activation energy and adsorption heat values indicate that the adsorption of imidazoline-based inhibitor on
the steels was endothermic, and the results are in agreement with the
The polarization curves of the X70 and Q235 steels in CO2-saturated phenomenon of the proposed chemisorption mechanism [43].
3 wt% NaCl solution with and without 4.8780 × 10−5 mol/l inhibitor at
313 K and 333 K are presented in Fig. 8(a) and (b), respectively. 3.5. Weight loss measurements
As observed in Table 3, for X70 and Q235 steels, the corrosion cur-
rent densities increased with increasing temperature in the absence To further estimate the different inhibition effect of the imidazoline-
and presence of the inhibitor. The results can be attributed to the accel- based inhibitor on X70 and Q235 steel, weight loss measurements were
eration of all the chemical, electrochemical, and transport with increas- employed for both steels in 3 wt% NaCl solution with 1 MPa partial pres-
ing temperature [42]. In addition, the values of IE and θ increased with sure of CO2 at 333 K for 96 h.
increasing temperature for both steels. The samples surfaces of the X70 and Q235 steel were completely
To clarify the effect of temperature on the inhibition characteristics covered by corrosion products in the absence of the inhibitor, as ob-
of the inhibitor, the apparent activation energies (Ea) of the corrosion served in Fig. 9(a) and (b), respectively. This finding results from the
process without and with 4.8780 × 10−5 mol/l inhibitor were deter- dissolution of metal and the formation of corrosion products. When
mined from a modified form of the Arrhenius equation as follows [1]: the X70 and Q235 samples were separately immersed in
4.8780 × 10− 5 mol/l inhibitor (Test 1 and Test 2), both steels were
  well protected, as observed in Fig. 9(c) and (d), respectively. The results
icorr;2 Ea 1 1 are attributed to the adsorption of inhibitor onto the samples surfaces in
log ¼ − ð9Þ
icorr;1 2:303R T 1 T 2 a short time, consequently forming a protective layer between the raw
materials and corrosive medium [44]. In this case, the inhibition effi-
ciencies for X70 and Q235 were 97.31% and 97.44%, respectively, as ob-
where icorr,1 and icorr,2 are the corrosion current densities at tempera- served in Fig. 10.
tures T1 (313 K) and T2 (333 K), respectively, and R is the molar gas When the X70 and Q235 samples were simultaneously immersed in
constant. the solution with inhibitor at low concentration (4.8780 × 10−6 mol/l,
The calculated Ea values of the X70 and Q235 steels without the in- Test 5), the surface morphology of X70 was severely damaged, as ob-
hibitor were 19.25 kJ/mol and 23.36 kJ/mol, respectively. However, served in Fig. 11(a). In contrast, the surface of Q235 steel was well
the presence of 4.8780 × 10−5 mol/l inhibitor resulted in a decrease of protected, as observed Fig. 11(b). The inhibition efficiencies for X70
the Ea values to 12.17 kJ/mol and 13.44 kJ/mol for X70 and Q235 steels, and Q235 were 8.35% and 95.10%, respectively, as observed in Fig. 10, in-
respectively. This decrease can be interpreted as a chemical adsorption dicating that most of the inhibitor molecules adsorbed onto the Q235
of the inhibitor on the surfaces of the X70 and Q235 steel [15,24,38]. steel surfaces.

Table 4
EIS results and inhibition efficiency for X70 and Q235 steels in CO2-saturated 3 wt% NaCl containing different concentrations of the inhibitor.

T, K C, 10−5 mol/l YCPE, Ω−1cm2sn10−6 (mV) fmax, Hz (μA/cm2) n Rct, Ω/cm2 CPEdl, μF/cm2 IE, %

X70
313 0 1186 ± 48 4.250 0.85 ± 0.03 112 ± 4 725 –
0.4878 856 ± 37 3.360 0.84 ± 0.03 280 ± 4 526 58.1
1.2195 187 ± 12 1.670 0.87 ± 0.04 913 ± 5 138 87.7
2.4390 130 ± 8 1.670 0.86 ± 0.02 1355 ± 7 94 91.7
4.8780 102 ± 9 1.040 0.88 ± 0.02 2689 ± 8 82 95.8
333 0 1574 ± 57 6.820 0.84 ± 0.04 73 ± 2 863 –
0.4878 1208 ± 46 4.250 0.86 ± 0.03 174 ± 3 763 60.0
1.2195 270 ± 13 2.100 0.85 ± 0.05 897 ± 5 184 91.9
2.4390 179 ± 9 2.100 0.88 ± 0.04 1240 ± 6 132 94.1
4.8780 133 ± 10 1.670 0.87 ± 0.01 1799 ± 5 98 95.9

Q235
313 0 611 ± 25 5.360 0.86 ± 0.03 132 ± 3 374 –
0.4878 76 ± 5 2.100 0.85 ± 0.02 1474 ± 5 52 91.0
1.2195 70 ± 6 1.320 0.88 ± 0.04 2601 ± 7 55 94.9
2.4390 74 ± 6 0.827 0.84 ± 0.02 3310 ± 5 57 96.0
4.8780 64 ± 5 0.654 0.83 ± 0.02 4397 ± 8 51 97.0
333 0 939 ± 30 5.360 0.88 ± 0.04 77 ± 4 616 –
0.4878 82 ± 5 2.100 0.85 ± 0.04 1134 ± 5 56 93.2
1.2195 85 ± 5 1.320 0.85 ± 0.02 1696 ± 6 62 95.5
2.4390 75 ± 4 1.040 0.87 ± 0.03 2914 ± 8 59 97.4
4.8780 71 ± 6 0.827 0.86 ± 0.03 3158 ± 5 57 97.6
H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94 91

Fig. 7. Langmuir adsorption plots for X70 and Q235 steel in CO2-saturated 3 wt% NaCl solution containing different concentrations of inhibitor from polarization results: (a) X70-313 K, (b)
X70-333 K, (c) Q235-313 K, (d) Q235-333 K.

It can be concluded that the imidazoline-based inhibitor had an ex- thus higher for various concentrations of inhibitor at the studied
cellent inhibition effect on the X70 and Q235 steels for the concentra- temperatures.
tion of 4.8780 × 10−5 mol/l inhibitor. In addition, the immersion test
result confirms that the inhibitor molecules exhibited stronger adsorp- 3.6. Potential of zero charge and inhibition mechanism
tion ability onto Q235 steel than that onto X70 steel; this result is con-
sistent with the previously discussed equilibrium constant (K) values. The adsorption of an inhibitor on a metal surface depends on not
The polarization and weight loss measurements confirmed that the in- only the physicochemical properties of the inhibitor but also the
hibitor behaved better on Q235 steel than on X70 steel at the studied employed electrolyte and the surface charge of the metal. Epzc can be
temperatures. The different adsorption ability of the inhibitor on X70 used to investigate and explain the adsorption mechanism of an inhib-
and Q235 can be attributed to the significant distinct microstructures. itor [35].
It has been proposed that, compared with the alloyed Fe matrix, or- EIS was used to determine Epzc of X70 and Q235 steels in CO2-satu-
ganic inhibitor molecules will be more preferentially and strongly rated 3 wt% NaCl solution without and with 4.8780 × 10−6 mol/l inhib-
adsorbed onto the exposed carbides (Fe3C) through Lewis–acid interac- itor at 333 K after 1 h of exposure time. The surface charge of the metal
tions [45]. Therefore, the amount and distribution of carbides should not can be determined by comparing Eocp with Epzc, which can be deter-
be overlooked. As observed in Fig. 1, the X70 steel mainly consists of fer- mined using the following equation:
rite, and the Q235 steel has a uniform microstructure consisting of fer-
rite and a homogeneous distribution of fine carbides. Therefore, based EΓ ¼ Eocp −Epzc ð11Þ
on the electrochemical measurements and immersion test results, the
inhibitor had stronger power to adsorb onto the Q235 surface than
onto the X70 surface, and the inhibition efficiencies for Q235 were where EΓ is the “rational” corrosion potential [36].
Plots of CPEdl vs. the applied potential (Eapp) are presented in Fig. 12.
In general, a minimum value on the CPEdl vs. Eapp curve is considered as
the Epzc value of the electrode. As shown in Table 6, positive values of EΓ
Table 5 in the blank and inhibited solutions indicated that the X70 and Q235
Adsorption parameters obtained from the Langmuir adsorption isotherm at different
surfaces were positively charged in the solution without and with the
temperatures.
addition of the inhibitor [36]. Therefore, the following mechanism for
T(K) Steel R2 f K(l/mol) ΔG0ads(kJ/mol) the corrosion of X70 and Q235 in CO2-saturated 3 wt% NaCl solution
313 X70 0.99588 0.96599 375,119 −43.85 with the inhibitor is proposed.
333 0.99676 0.96815 429,647 −47.03 It is well known that, in the solution, a protonation of the synthe-
313 Q235 0.99994 1.00696 1,934,583 −48.12 sized imidazoline-based inhibitor will occur at the S atom as follows
333 0.99999 1.00751 3,093,758 −52.50
[46]:
92 H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94

Fig. 10. Corrosion rates and inhibition efficiencies of X70 and Q235 steel in 3 wt% NaCl
solution with various concentrations of inhibitor at 333 K.

ð12Þ

It has been reported that the cathodic process of carbon steel CO2
corrosion is the hydrogen evolution reaction. In this study, in the
Fig. 8. Polarization curves for X70 (a) and Q235 (b) steel in CO2-saturated 3 wt% NaCl inhibited solution, the protonated imidazoline-based inhibitor can be
solution in the absence and presence of 4.8780 × 10−5 mol/l inhibitor at 313 K and 333 K. adsorbed at the cathodic sites in competition with hydrogen ions, there-
by reducing H2 gas evolution [47].

Fig. 9. Surface morphologies of the steels separately immersed in 3% NaCl solution without and with the inhibitor at 333 K: (a) X70-0 mol/l; (b) Q235-0 mol/l; (c) X70-
4.8780 × 10−5 mol/l; (d) Q235-4.8780 × 10−5 mol/l.
H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94 93

Table 6
Excess charge on X70 and Q235 surfaces in CO2-saturated 3 wt% NaCl solution without and
with inhibitor at 333 K.

Steel Solution Eocp, vs. Ag/AgCl Epzc, vs. Ag/AgCl Excess


(mV) (mV) charge

X70 Without inhibitor −663 −763 +


With 4.878 × 10−6 mol/l −637 −737 +
inhibitor
Q235 Without inhibitor −681 −781 +
With 4.878 × 10−6 mol/l −625 −685 +
inhibitor

[35–36,48–49]. In addition to this type of physical adsorption, as


discussed above, the synthesized inhibitor molecules can adsorb on
the X70 surface directly by sharing a lone electron pair in N, S atoms
and/or π electrons in the benzimidazole ring with the unoccupied d or-
bital of iron atoms to form a coordinate covalent bond (chemisorption)
[46,50]. Furthermore, because various orientations of the “d” orbital of
iron exist, the double bonds in the benzimidazole ring allow back dona-
tion of iron d-electrons to the π*-orbital of the inhibitor to form a feed-
back bond [47–48]. The inhibition mechanism mentioned above also
applies to Q235 steel.
Therefore, it can be concluded that the imidazoline-based inhibitor
Fig. 11. Surface morphologies of X70 (a) and Q235 (b) steel simultaneously immersed in adsorbed on X70 and Q235 surfaces via the forms of physisorption
3 wt% NaCl solution with 4.8780 × 10−6 mol/l inhibitor at 333 K. and chemisorption. Among the above adsorption processes, physical
adsorption was the necessary first step, which then allowed the chem-
For the anodic reaction, in the presence of 4.8780 × 10−6 mol/l in- ical adsorptions to proceed [47]. The large negative values of ΔG0ads
hibitor, positive excess charge was carried on X70 surfaces. In this also suggest chemisorption was the dominant mechanism.
case, it was very difficult for the protonated inhibitor to approach the
metal surface. In contrast, the protonated inhibitor could be adsorbed 4. Conclusions
onto the positively charged surface via anions (Cl−, HCO2− 2−
3 , CO3 ) in
the solution. In this step, the anions acted as the connecting bridges be- This study estimated the ability of an imidazoline-based inhibitor to
tween the positively charged metal surface and protonated inhibitor minimize CO2 corrosion damage of X70 and Q235 in 3 wt% NaCl solution

Fig. 12. Plots of CPEdl versus Eapp in CO2-saturated 3 wt% NaCl solution at 333 K.
94 H. Zhang et al. / Journal of Electroanalytical Chemistry 791 (2017) 83–94

at 313 K and 333 K. The electrochemical results revealed that the protec- [13] C.B. Verma, A. Singh, G. Pallikonda, M. Chakravarty, M.A. Quraishi, I. Bahadur, E.E.
Ebenso, J. Mol. Liq. 209 (2015) 306–319.
tion efficiency increased with increasing inhibitor concentration and [14] X. Liu, P.C. Okafor, Y.G. Zheng, Corros. Sci. 51 (2009) 744–751.
slightly increased with increasing temperature. The inhibitor acted as [15] M.P. Desimone, G. Gordillo, S.N. Simison, Corros. Sci. 53 (2011) 4033–4043.
a mixed-type inhibitor and suppressed both the anodic and cathodic [16] D.A. Lopez, S.N. Simison, S.R. de Sanchez, Corros. Sci. 47 (2005) 735–755.
[17] P.C. Okafor, C.B. Liu, Y.J. Zhu, Y.G. Zheng, Ind. Eng. Chem. Res. 50 (2011) 7273–7281.
corrosion reactions. The results also indicated that the amount and dis- [18] A.O. Yuce, B.D. Mert, G. Kardas, B. Yazici, Corros. Sci. 83 (2014) 310–316.
tribution of carbides (Fe3C) on the sample had a remarkable effect on [19] L.J. Oblonsky, G.R. Chesnut, T.U. Devine, Corrosion 51 (1995) 891–900.
the performance of the inhibitor. The inhibitor showed better corro- [20] D.A. Lopez, S.N. Simison, S.R. de Sanchez, Electrochim. Acta 48 (2003) 845–854.
[21] H.H. Zhang, X.L. Pang, M. Zhou, C. Liu, L. Wei, K.W. Gao, Appl. Surf. Sci. 356 (2015)
sion-inhibiting performance for Q235 steel than for X70 steel. Epzc mea- 63–72.
surements revealed that the anions formed connecting bridges between [22] J. Cruz, L.M.R. Martinez-Aguilera, R. Salcedo, M. Castro, Int. J. Quantum Chem. 85
the steel surfaces and protonated inhibitor molecules. The adsorption of (2001) 546–556.
[23] M.A. Amin, K.F. Khaled, Q. Mohsen, H.A. Arida, Corros. Sci. 52 (2010) 1684–1695.
the inhibitor on the metal surface was a mix of physisorption and chem-
[24] F. Bentiss, M. Lobrini, M. Lagrenee, Corros. Sci. 47 (2005) 2915–2931.
isorption. The adsorption of the inhibitor was observed to follow the [25] I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, Corros. Sci. 51 (2009) 1868–1875.
Langmuir adsorption isotherm. The present work showed that the [26] V.V. Torres, V.A. Rayol, M. Magalhaes, G.M. Viana, L.C.S. Aguiar, S.P. Machado, H.
imidazoline-based inhibitor was an efficient corrosion inhibitor and Orofino, E. Delia, Corros. Sci. 79 (2014) 108–118.
[27] J.W. Tang, Y.W. Shao, J.B. Guo, G.Z. Meng, F.H. Wang, Corros. Sci. 52 (2010)
that the steel characteristics could affect the inhibitor efficiency. Further 2050–2058.
research should be conducted to provide more information on the effect [28] S.A. Umoren, I.B. Obot, A. Madhankumar, Z.M. Gasem, Carbohydr. Polym. 124 (2015)
of the microstructure on the inhibition efficiency. It would be particular- 280–291.
[29] Z. Tao, W. He, S. Wang, S. Zhang, G. Zhou, Corros. Sci. 60 (2012) 205–213.
ly useful to investigate the effect of corrosion time on the inhibition ef- [30] C. Cao, Corros. Sci. 38 (1996) 2073–2082.
ficiencies of imidazoline-based inhibitors for different steels. [31] M. Nordsveen, S. Nesic, R. Nyborg, A. Stangelend, Corrosion 59 (2003) 443–456.
[32] M. Behpour, S.M. Ghoreishi, Corros. Sci. 52 (2010) 4046–4057.
[33] T. Hong, Y.H. Sun, W.P. Jepson, Corros. Sci. 44 (2002) 101–112.
Acknowledgements [34] W.W. Zhang, R. Ma, H.H. Liu, Y. Liu, S. Liu, L. Liu, J. Mol. Liq. 222 (2016) 671–679.
[35] K.G. Zhang, B. Xu, W.Z. Yang, X.S. Yin, Y. Liu, Y.Z. Chen, Corros. Sci. 90 (2015)
This work was supported by the Beijing Natural Science Foundation 284–295.
[36] K. Mallaiya, R. Subramaniam, S.S. Srikandan, S. Gowri, A. Selvaraj, Electrochim. Acta
(2131004) and National Natural Science Foundation of China 56 (2011) 3857–3863.
(51271024). [37] M.A. Hegazy, J. Mol. Liq. 208 (2015) 227–236.
[38] S.A.M. Refaey, F. Taha, A.M. Abd EI-Malak, Appl. Surf. Sci. 236 (2004) 175–185.
[39] M.A. Migahed, M.A. Hegazy, A.M. Al-Sabagh, Corros. Sci. 61 (2012) 10–18.
References
[40] E. Khamis, F. Bellucci, R.M. Latanision, E.S.H. EI-Ashry, Corrosion 47 (1991) 677–686.
[41] S. Bilgic, M. Sahin, Mater. Chem. Phys. 70 (2001) 290–295.
[1] P.C. Okafor, X. Liu, Y.G. Zheng, Corros. Sci. 51 (2009) 761–768.
[42] S. Nesic, Corros. Sci. 49 (2007) 4308–4338.
[2] B. Wang, M. Du, J. Zhang, C.J. Gao, Corros. Sci. 53 (2011) 353–361.
[43] S.A. Umoren, I.B. Obot, E.E. Ebenso, P.C. Okafor, E.E. Oguzie, Anti-Corros. Methods
[3] M.E.O. Maartinez, J.M. Flores, J. Genesca, J. Loss Prev. Process Ind. 35 (2015) 19–28.
Mater. 53 (2006) 277–282.
[4] A. Singh, Y.H. Lin, E.E. Ebenso, W.Y. Liu, J. Pan, B. Huang, J. Ind. Eng. Chem. 24 (2015)
[44] F.J. Liu, M. Du, J. Zhang, M. Qiu, Corros. Sci. 51 (2009) 102–109.
219–228.
[45] D.A. Lopez, W.H. Schreiner, S.R. de Sanchez, S.N. Simison, Appl. Surf. Sci. 236 (2004)
[5] I. Jevremović, M. Singer, M. Achour, D. Blumer, T. Baugh, V. Misković-Stanković, S.
77–97.
Nešić, Corrosion 69 (2013) 186–192.
[46] G.A. Zhang, C.F. Chen, M.X. Lu, C.W. Cai, Y.S. Wu, Mater. Chem. Phys. 105 (2007)
[6] S.H. Yoo, Y.W. Kim, K. Chung, S.Y. Baik, J.S. Kim, Corros. Sci. 59 (2012) 42–54.
331–340.
[7] J.M. Zhao, G.H. Chen, Electrochim. Acta 69 (2012) 247–255.
[47] X.M. Wang, H.Y. Yang, F.H. Wang, Corros. Sci. 53 (2011) 113–121.
[8] D.M. Ortega-Sotelo, J.G. Gonzalez-Rodriguez, M.A. Neri-Flores, M. Casales, L.
[48] B. Xu, W.Z. Yang, Y. Liu, X.S. Yin, W.N. Gong, Y.Z. Chen, Corros. Sci. 78 (2014)
Martinez, A. Martinez-Villafane, J. Solid State Electrochem. 15 (2011) 1997–2004.
260–268.
[9] A.H. Mustafa, B.A. Wahjoedi, M.C. Ismail, J. Mater. Eng. Perform. 22 (2013)
[49] S.R. Kumar, I. Danaee, M.R. Avei, M. Vijayan, J. Mol. Liq. 212 (2015) 168–186.
1748–1755.
[50] J. Zhang, M. Du, H.H. Yu, N. Wang, Acta Phys. -Chim. Sin. 25 (2009) 525–531.
[10] F. Farelas, A. Ramire, Int. J. Electrochem. Sci. 5 (2010) 797–814.
[11] P.C. Okafor, C.B. Liu, X. Liu, Y.G. Zheng, J. Appl. Electrochem. 39 (2009) 2535–2543.
[12] A. Biswas, S. Pal, G. Udayabhanu, Appl. Surf. Sci. 353 (2015) 173–183.

You might also like