Вы находитесь на странице: 1из 8

Corrosion Science 126 (2017) 247–254

Contents lists available at ScienceDirect

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

Synergistic inhibition effects of octadecylamine and tetradecyl trimethyl MARK


ammonium bromide on carbon steel corrosion in the H2S and CO2 brine
solution

Chen Zhanga, Jingmao Zhaoa,b,
a
College of Material Science and Engineering, Beijing University of Chemical Technology, Beijing 100029, China
b
Beijing Key Laboratory of Electrochemical Process and Technology for Materials, Beijing 100029, China

A R T I C L E I N F O A B S T R A C T

Keywords: The corrosion inhibition performances of octadecylamine (OCT) and tetradecyl trimethyl ammonium bromide
A. Acid solution (TTAB) for carbon steels in H2S and CO2 brine solution were investigated by weight loss, potentiodynamic
A. Carbon steel polarization, molecular dynamic simulation and XPS analysis. TTAB showed better inhibition effects on H2S and
B. Polarization CO2 induced corrosion than OCT. Synergistic corrosion inhibition effect was found between OCT and TTAB with
B. Weight loss
proper mass ratios and the best inhibition performance was achieved with 10 mg L−1 OCT and 20 mg L −1
B. XPS
TTAB. The synergism was correlative to the fractional free volume of the inhibitor film of OCT and TTAB.
C. Acid corrosion

1. Introduction inhibitors by many scholars in recent years [22–29]. However, it is


rarely used for the study of the synergistic inhibition effects.
The combined corrosion effects of CO2 and H2S on carbon steels is a In the present work, the synergistic inhibition effects of octadecy-
serious issue in the oil and gas production. Various advanced corrosion lamine (OCT) and tetradecyl trimethyl ammonium bromide (TTAB) on
protection technologies have been developed [1,2], among which the carbon steel corrosion in the H2S and CO2 co-existing brine solution
injection of corrosion inhibitors is one of the most economical and ef- were determined by weight-loss method and potentiodynamic polar-
fective methods to minimize the CO2 and H2S induced corrosion of ization measurements. XPS analysis and molecular dynamic (MD) si-
carbon steels. Organic compounds containing heteroatoms and aro- mulation were used to conjecture the synergistic inhibition mechanism.
matic rings, such as imidazoline [3–6], amine [7,8], quinoline [9], and
so on, are effective corrosion inhibitors. They are able to form an iso- 2. Experimental
lation film on the metal surface to retard the corrosion in aggressive
solutions [10–12]. 2.1. Materials
Generally, a single inhibitor can rarely achieve the desired corrosion
inhibitions. Synergism of two or more inhibitors can usually enhance All tests were performed on Q235 steels composed of 0.19 wt.% C,
the corrosion inhibition and reduce costs [13]. Previous studies have 0.59 wt.% Mn, 0.3 wt.% Si, 0.044 wt.% P, 0.05 wt.% S and bal. Fe. The
been focused on the synergistic corrosion inhibition effects of inhibitors steel was cut into coupons with the sizes of 50 mm × 10 mm × 3 mm
on carbon steels in inorganic acid solutions, such as HCl, H2SO4, H3PO4, for weight loss test, 10 mm × 10 mm × 3 mm for polarization mea-
and so on [13–19], and CO2-saturated brine solutions [20,21]. The surements, and 5 mm × 5 mm × 2 mm for XPS analysis. The coupons
synergistic corrosion inhibition of steels in H2S and CO2 coexisting were wet abraded with silicon carbide abrasive papers up to 3000 grit,
brine solution has rarely been reported. degreased in acetone, rinsed with ethanol, and dried by a hot air stream
The synergistic inhibition effects of various inhibitors are usually Analytical grade octadecylamine (OCT) and tetradecyl trimethyl
determined by weight loss method, electrochemical measurements, ammonium bromide (TTAB) were used as the corrosion inhibitors in
surface analysis and other techniques to screen and develop new high present work. Their chemical structures are shown in Fig. 1. An ag-
efficiency composite inhibitors. Molecular simulation techniques, gressive solution of 3.5 wt.% NaCl solution was prepared with analy-
which offset the deficiencies of experimental methods, have also been tical grade sodium chloride and distilled water.
used to evaluate the corrosion inhibition performance of various


Corresponding author at: No. 15 of Beisanhuan East Road, Chao Yang District, Beijing 100029, China.
E-mail address: jingmaozhao@126.com (J. Zhao).

http://dx.doi.org/10.1016/j.corsci.2017.07.006
Received 6 August 2016; Received in revised form 11 July 2017; Accepted 12 July 2017
Available online 14 July 2017
0010-938X/ © 2017 Elsevier Ltd. All rights reserved.
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

2.3. Potentiodynamic polarization measurements

The potentiodynamic polarization measurements were performed


on a Gamry interface 1000 electrochemical system at 60 ± 1 °C. A
conventional three-electrode cell with a platinum foil as the counter
electrode and a saturated calomel electrode as the reference electrode
was used. The test solutions for the polarization measurements were
prepared as described in the weight loss test except that the solution
volume was 200 ml. To achieve a stable status of the working electrode,
the working electrode was immersed in the test solution for 1 h before
the measurements to ensure an open circuit potential (OCP) fluctuation
less than ± 5 mV (Fig. 2). The potentials were then scanned from −200
to +200 mV vs. OCP at a scan rate of 0.16 mV s−1. The electrochemical
Fig. 1. Chemical structures of TTAB (a) and OCT (b).
parameters were fitted in the software Cview [30]. The inhibition ef-
ficiency (ηp) of the corrosion inhibitor was defined as follows:
2.2. Weight loss
icorr − icorr (inh)
ηp = × 100%
Three identical coupons were weighed by an electric balance with icorr (3)
precision of ± 0.1 mg and immersed in 700 ml aggressive solutions where icorr and icorr(inh) are the corrosion current densities of the steel in
containing various concentrations of the inhibitors. The solutions were the aggressive solutions with and without inhibitors added, respec-
purged with pure N2 gas for 1 h to remove the oxygen, then H2S tively.
(20 ml min−1) and CO2 (20 ml min−1) were simultaneously introduced Each measurement was repeated at least three times to obtain ac-
into the solution for 1 h. The H2S/CO2 brine solutions containing the curate data.
coupons were seal with silicone sealants and incubated at 60 ± 1 °C
for 72 h. The coupons were then taken out, immersed in a Clarke’s 2.4. Molecular dynamics (MD) simulation
solution (6 g 1,3,5,7-tetraazatricyclo [3.3.1.1(3,7)]decane + 500 ml
distilled water and hydrochloric acid to make 1000 ml) for 5 min to A simplified amorphous model was constructed to calculate fraction
remove the corrosion products, washed under running water, degreased free volume (FFV) with Materials Studio software in the present work
in acetone, rinsed with ethanol, dried, and re-weighed accurately. The (Fig. 3). The simulation was conducted in four steps:
corrosion rate (VCR) of the carbon steel was calculated by Eq. (1).
1. A cubic model containing 50 inhibitor molecules in three-dimen-
ΔW
VCR = sional periodic boundary conditions was constructed by the
Sm × t (1)
Amorphous Cell module in Materials Studio. The initial density of
the model was set to 0.6 g cm−3 and the model was then minimised
where ΔW is the average weight loss of the three coupons, Sm is the
by the smart minimisation method until the convergence reached to
total surface area of the coupon, and t is the immersion time.
0.042 kJ mol−1.
The inhibition efficiency (ηw) of inhibitors was defined as follows:
2. To evaluate the stable density of the model, a 500 ps MD simulation
VCR (inh) with a time step of 1 fs was carried out at NPT ensemble. The final
ηw = (1 − ) × 100% density of the model was defined as the average density of the last
VCR (uninh) (2)
200 frames that were recorded at every one frame per ps.
where VCR(inh) and VCR(uninh) are the corrosion rates of the steel in the 3. Step 1 was repeated with the density of the model set to the value
solution with and without inhibitors, respectively. obtained in step 2.
4. A 1000 ps MD simulation with a time step of 1 fs was carried out at
NVT ensemble to calculate the FFV of the model by the Atom

Fig. 2. OCP curves of Q235 carbon steel in H2S and CO2 co-existing 3.5 wt.
% NaCl solutions containing different inhibitors at 60 °C.

248
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

391–414 eV with a step size of 0.05 eV. All binding energies were ca-
librated with the binding energy of C1s peak at 284.6 eV. The data were
fitted and analysed in XPSPEAK version 4.1 software.

3. Results and discussion

3.1. Weight loss

Weight loss experiments were conducted at 60 °C to investigate the


synergistic corrosion inhibition performances of OCT and TTAB in the
brine solution containing H2S and CO2. As shown in Fig. 4a for the
inhibition efficiencies (ηw) of OCT and TTAB, the ηw values of both OCT
and TTAB increased with the increase of concentration. Fig. 4b shows
the ηw measured in the aggressive solutions containing various con-
centrations of TTAB and OCT with a total concentration of 30 mg L−1.
As can be seen, the ηw of the composite inhibitor increased with the
portion of TTAB (C%TTAB) and remained steady at ∼95% as C%TTAB
increased to over 67%.
To compare the synergistic degrees of the different combinations of
TTAB and OCT, the synergistic parameter, S, was calculated as follows
Fig. 3. Model for FFV calculation. [32,33]:
1 − ηA − ηB + ηAηB
S=
1 − ηAB (4)
Volumes and Surfaces Tool in Materials Studio. The FFV values of
nine structures chosen every five frames in the last forty frames were where ηA, ηB and ηAB are the inhibition efficiencies of inhibitor A, in-
calculated and the maximum and minimum values were omitted. hibitor B and the mixture of A + B, respectively. Generally, S = 1 in-
The mean value of the rest seven values was defined as the FFV dicates that there is no synergistic effect between inhibitor A and B.
value of the model. Alternatively, S > 1 if the synergistic effects manifest themselves and
S < 1 if there is an antagonistic effect between inhibitor A and B. The
COMPASS force field was used in all simulation steps [31]. The si- synergistic corrosion inhibition effects of OCT and TTAB were found at
mulation temperature and pressure of the model were controlled at C%TTAB = 50–83% and peaked at C%TTAB = 67%. In addition, the ηw
333 K and 1 × 105 Pa by Andersen thermostat and Berendsen barostat, of 10 mg L−1 OCT + 20 mg L−1 TTAB is higher than the sum of those
respectively. The long-range coulomb and Van der Waals interactions of 10 mg L−1 OCT and 20 mg L−1 TTAB, indicating the good corrosion
were handled by the Ewald and Atom Based methods, respectively. The inhibition synergistic effect between OCT and TTAB.
cutoff was set to 1.25 nm.
3.2. Potentiodynamic polarization
2.5. X-ray photoelectron spectroscopy (XPS) analysis
To further confirm the synergistic corrosion inhibition effect of OCT
The coupons were immersed in CO2 and H2S co-existing solutions and TTAB, potentiodynamic polarization experiments were performed
containing 10 mg L−1 OCT and 20 mg L−1 TTAB at 60 ± 1 °C for 5, at 60 °C. Fig. 5 shows the polarization curves of mild steels in the ag-
15, 25 and 40 min, respectively, taken out from the solutions, dried by a gressive solutions containing different inhibitors. As can be seen, the
hot air, and kept in a desiccator before use. XPS spectra were recorded inhibitors can affect both anodic dissolution and cathodic reactions on
on a PHI-5300ESCA spectrometer (Perkin–Elmer, USA) equipped with the steel surface. The corrosion potential (Ecorr) of the steel in the
an Al Kα excitation source. The survey scan was set to 0–1155 eV with a presence of composite inhibitors obviously shifted toward positive di-
step size of 1 eV. The high-resolution spectra of N1s were achieved over rection in the TTAB concentration (cTTAB) range of 20–30 mg L−1 and

Fig. 4. ηw vs concentration plots of TTAB and OCT (a) and ηw vs TTAB content plot and S vs TTAB content plot of the composite inhibitor at a total addition concentration of 30 mg L−1
(b) for the Q235 carbon steel measured in 3.5 wt.% NaCl solution containing H2S and CO2 at 60 °C.

249
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

Fig. 5. Polarization curves of Q235 carbon steel measured in H2S and CO2
co-existing 3.5 wt.% NaCl solutions containing various concentrations of
TTAB and OCT at 60 °C.

slightly shifted at cTTAB below 20 mg L−1. It has been reported that corrosive particles in the film with large cavities is faster than that in
TTAB and OCT are anodic-type and mixed-type inhibitors, respectively the film with small cavities. Therefore, the size of the cavities in the
[34,35]. Therefore, the composite inhibitors evolved from mixed-type inhibitor film is an important factor determining the inhibition per-
to anodic-type with the increase of cTTAB, which eventually retarded the formance of the inhibitor. To determine the free volume inside an in-
anodic dissolution of the mild steel [36,37]. hibitor film, the fractional free volume (FFV) was calculated as follows
Table 1 shows the electrochemical parameters fitted from the po- [40,41]:
larization curves. It is clear that ηp of OCT and TTAB composite in-
hibitor is significantly higher than those of OCT or TTAB alone. The ηp Vfree
FFV = × 100%
of the composite inhibitors increased dramatically with the increasing Vfree + Voccupy (5)
of cTTAB to 20 mg L−1 and gradually increased with the increasing of
cTTAB from 20 mg L−1 to 30 mg L−1. Although the absolute ηp values where Vfree is the total volume of the large cavities where the corrosive
differ from those obtained by the weight loss method, their changing particles can enter. The small cavities where the corrosive particles
trends with the increase of cTTAB obtained by these two methods are cannot diffuse into are excluded from the Vfree. Voccupy is the volume of
consistent. the film excluding Vfree. The FFV of an inhibitor film depends on the
radiuses of the corrosive particles and the properties of the inhibitor. A
low FFV indicates fewer cavities with smaller sizes where the corrosive
3.3. Fractional free volume (FFV) particles can diffuse, and thus the good inhibition performance of the
inhibitor. In the present work, the corrosive particles including H3O+,
Organic inhibitors can usually adsorb on the metal surface sponta- HS−, Cl−, HCO3−, and H2O were selected as the probe species to
neously and form an inhibitor film to retard corrosion. Co-adsorption of evaluate the FFV of the inhibitor films. The Van der Waals radiuses
two or more inhibitors, which is usually conducive to the formation of a (Rvdw) of the corrosive particles were calculated by the radial dis-
more compact inhibitor film on the metal surface, may occur if the tribution function (Table 2) [24]. The calculated Rvdw are consistent to
corrosion inhibitors are adsorption-type and have a synergistic corro- the experimental results reported in the literatures, suggesting that the
sion inhibition effect [38,39]. Although the inhibitor film can prevent radial distribution function method adopted in the present work is
the contact of metal surface with corrosive media to some extent, cor- credible.
rosive particles may still be able to diffuse to the metal surface through The inhibitor films were then established with OCT and TTAB at
large cavities in the inhibitor film [25]. The diffusion rate of the mass ratios of 1:0, 1:0.5, 1:1, 1:2, 1:3, 1:5, 1:8 and 0:1, respectively, and
their FFV were calculated. As shown in Fig. 6, the FFV of the films
Table 1 constructed with 50–90% TTAB in the composite inhibitor are rela-
Electrochemical parameters fitted from the polarization curves of Q235 steel in H2S and tively low for all probe species. The lowest FFV was achieved with 75%
CO2 co-existing 3.5 wt.% NaCl solutions containing various concentrations of TTAB and
TTAB i.e. a 1:3 mass ratio of OCT to TTAB.
OCT at 60 °C.

cOCT cTTAB βa (mV/ βc (mV/dec) Ecorrvs.SCE icorr ηp (%) Table 2


(mg/L) (mg/L) dec) (mV) (μA/ Calculated and experimental Van der Waals radiuses of corrosive particles.
cm2)
Corrosive Particle Calculated (nm) Experimental (nm)
0 0 54.38 −231.33 −736.31 141.8 –
0 30 62.48 −149.1 −600.24 3.35 97.64 H2O 0.1365 0.1450 [42]
5 25 57.66 −186.16 −628.73 1.37 99.03 H3O+ 0.1405 –
10 20 68.51 −100.25 −654.65 1.72 98.79 Cl− 0.1725 0.1810 [43]
15 15 61.41 −155.56 −701.08 21.2 85.05 HCO3− 0.1835 –
20 10 58.68 −176.29 −724.08 48.9 65.51 HS− 0.1885 –
25 5 58.09 −181.91 −731.82 84.2 40.62
30 0 51.78 −294.05 −716.63 48.1 66.08

250
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

Fig. 6. FFV vs TTAB content curves for different probe species.

The S curve of the composite inhibitor peaked at 70% TTAB and C1s, and S2p peaks were observed in the wide-scan spectrum. The C1s
showed an opposite changing tendency to FFV values with the increase and N1s signals were mainly attributed to the OCT and TTAB molecules
of C%TTAB (Fig. 4b vs 6), indicating that there is a negative correlation adsorbed on the steel surface. The signals of O1s and S2p were due to
between FFV and S although their extreme points were at different the oxide, oxycarbide, and sulfide of iron. Fe2p peak can be ascribed to
C%TTAB values. Based on this negative correlation, the optimal mass the substrate and its corrosion products.
ratio range of the inhibitors in a composite inhibitor may be con- Fig. 8 shows the high-resolution spectra of N1s of the inhibitor films
jectured out by theoretical calculations. However, further work is prepared at a 1:3 mass ratio of OCT to TTAB. Two peaks at 402.3 eV
needed to develop a method for accurately determining the optimal and 399.8 eV were observed, which were assigned to the −N+(CH3)3
compositions of different composite inhibitors. in TTAB [44,45] and the −NH2 in OCT, respectively [46]. The ratio of
peak areas of −N+(CH3)3 to −NH2 and other parameters obtained
3.4. X-ray photoelectron spectroscopy (XPS) analysis from the N1s spectra are listed in Table 3. The peak areas of
−N+(CH3)3 and −NH2 of the inhibitor film formed in 5 min are almost
To further investigate the co-adsorption of OCT and TTAB, XPS equal, indicating that TTA+ cations and OCT molecules almost are
spectra of the carbon steel surfaces covered with different inhibitor equally adsorbed on the steel surface in the initial immersion stage. The
films were recorded. The inhibitor films were prepared by immersing peak area of −NH2 gradually became larger than that of −N+(CH3)3
the steel coupons in the H2S and CO2 co-existing brine solution con- with the increasing immersion time, indicating that more OCT mole-
taining 10 mg L−1 OCT and 20 mg L−1 TTAB at 60 °C for 5, 15, 25, 40, cules were adsorbed on the steel surface than TTA+ cations in the late
and 60 min respectively. Fig. 7 shows a typical wide-scan spectrum of immersion stage.
the inhibitor film prepared by a 40 min immersion. The F2p, O1s, N1s,

Fig. 7. Wide-scan spectrum of the Q235 carbon steel immersed in H2S and
CO2 co-existing 3.5 wt.% NaCl solution containing 10 mg L−1 OCT and
20 mg L−1 TTAB for 40 min at 60 °C.

251
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

Fig. 8. High-resolution XPS spectra of N1s of the Q235 carbon steel coupons immersed in H2S and CO2 co-existing 3.5 wt.% NaCl solution containing 10 mg L−1 OCT and 20 mg L−1
TTAB for 5 min (a), 15 min (b), 25 min (c), and 40 min (d) at 60 °C.

3.5. The synergistic inhibition mechanism of OCT and TTAB molecules simultaneously adsorbed on the steel surface to form an inner
inhibitor film in the initial immersion stage (Fig. 9b). With the
In the H2S and CO2 co-existing solution, HS− ions might pre- prolonging of the immersion, both OCT and TTAB tended to adsorb on
ferentially adsorb on the carbon steel surface and negatively charged the inner film to form an outer inhibitor film. However, the TTA+ ions
the surface [47,48]. The lone pair electrons of N atom in OCT tend to in the outer film were easily desorbed due to the high water-solubility
bind the hydrogen ions to form amide cations in the acidic solution of TTAB [54]. Therefore, the relative OCT content on the steel surface
[49,50]. TTAB molecules could be hydrolysed to TTA+ cations in the was much higher than that of TTA+ in the later immersion stage. The
test solution. The quaternary ammonium ions and the protonated pri- ideal structure of the inhibitor film is illustrated in Fig. 9c [55]. The
mary amine could simultaneously bind to the HS− layer adsorbed on molecular dynamics simulation indicated that this inhibitor film had a
the metal surface through electrostatic interaction [51–53], which was low FFV for all corrosive particles and thus could effectively prevent
a physisorption process. Subsequently, the N atoms of both TTAB and corrosive particles from diffusing to the steel surface. In addition, FFV is
OCT donated their lone pair electrons to the empty d-orbitals in Fe negatively correlative to S, which could be used to establish a low-cost
atoms to form coordinate covalent bonds between the inhibitors and the method to explore new high-effective composite inhibitors. However,
metal surface, which was a chemisorption process. The physisorption further work is needed to establish an effective method to screen
and chemisorption of TTA+ cations and protonated OCT molecules on composite inhibitors by the correlation between FFV and S.
the steel surface are demonstrated in Fig. 9a.
As demonstrated by the XPS analysis, TTA+ cations and OCT 4. Conclusion

Table 3 In the present work, the synergistic corrosion inhibition effect of


Parameters of the high-resolution spectra of N1 s of the Q235 steel in the H2S and CO2 co- OCT and TTAB on the carbon steels in the H2S and CO2 co-existing
existing 3.5 wt.% NaCl solution containing 10 mg L−1 OCT and 20 mg L−1 TTAB at 60 °C.
brine solution was reported for the first time. Their combination was
Immersion Time -N+(CH3)3 Peak -NH2 Peak Ratio of Peak Area able to effectively retard the H2S and CO2 induced corrosion of carbon
(min) Area Area (-NH2:-N+(CH3)3) steel. TTAB and OCT are anodic-type and mixed-type inhibitors, re-
spectively. Their combinations evolved from mixed-type to anodic-type
5 2710 2692 0.993
with the increase of TTAB concentration. The molecular dynamics si-
15 2543 4468 1.757
25 2023 5526 2.732 mulation indicated that the FFV of the inhibitor film formed with OCT
40 1322 4464 3.377 and TTAB was negatively correlated to the degree of synergistic cor-
rosion inhibition effect, which could be used to screen new composite
corrosion inhibitors. The XPS measurements showed that TTA+ cations
and OCT molecules simultaneously adsorbed on the steel surface in the

252
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

Fig. 9. Schematic diagram of the interactions between the inhibitors and


metal surface (a), co-adsorption process of TTAB and OCT (b), and the
ideal structure of the inhibitor film (c) formed in H2S and CO2 co-existing
3.5 wt.% NaCl solution containing OCT and TTAB at 60 °C.

initial immersion stage, and more OCT molecules adsorbed on the References
surface than TTA+ cations in the late immersion stage.
[1] W. He, O.Ø. Knudsen, S. Diplas, Corrosion of stainless steel 316L in simulated
formation water environment with CO2-H2S-Cl−, Corros. Sci. 51 (2009)
2811–2819.
Acknowledgment [2] G.A. Zhang, Y. Zeng, X.P. Guo, F. Jiang, D.Y. Shi, Z.Y. Chen, Electrochemical cor-
rosion behavior of carbon steel under dynamic high pressure H2S/CO2 environ-
The authors thank National Natural Science Foundation of China for ment, Corros. Sci. 65 (2012) 37–47.
[3] F. Fei, J. Hu, J. Wei, Q. Yu, Z. Chen, Corrosion performance of steel reinforcement
the financial supports for this work (No. 51471021). in simulated concrete pore solutions in the presence of imidazoline quaternary
ammonium salt corrosion inhibitor, Constr. Build. Mater. 70 (2014) 43–53.
[4] M. Yadav, D. Behera, U. Sharma, Nontoxic corrosion inhibitors for N80 steel in

253
C. Zhang, J. Zhao Corrosion Science 126 (2017) 247–254

hydrochloric acid, Arab. J. Chem. 9 (2016) S1487–S1495. for the corrosion of cold rolled steel in HNO3 solution, Corros. Sci. 118 (2017)
[5] Y.Z. Li, Corros. Sci. (2017), http://dx.doi.org/10.1016/j.corsci.2017.06.021. 202–216.
[6] Y. Zuo, Y. Li, Y. Tan, Y. Wang, J. Zhao, The effects of thioureido imidazoline and [30] S.D. Zhu, A.Q. Fu, J. Miao, Z.F. Yin, G.S. Zhou, J.F. Wei, Corrosion of N80 carbon
NaNO2 on passivation and pitting corrosion of X70 steel in acidic NaCl solution, steel in oil field formation water containing CO2 in the absence and presence of
Corros. Sci. 120 (2017) 99–106. acetic acid, Corros. Sci. 53 (2011) 3156–3165.
[7] B. Wu, H.L. Wang, V.J. Cee, B.A. Lanman, T. Nixey, L. Pettus, A.B. Reed, R.P. Wurz, [31] H. Sun, COMPASS: an ab initio force-field optimized for condensed-phase appli-
N. Guerrero, C. Sastri, J. Winston, J.R. Lipford, M.R. Lee, C. Mohr, K.L. Andrews, cations overview with details on alkane and benzene compounds, J. Phys. Chem. B
A.S. Tasker, Discovery of 5-(1H-indol-5-yl)-1,3,4-thiadiazol-2-amines as potent PIM 102 (1998) 7338–7364.
inhibitors, Bioorg. Med. Chem. Lett. 25 (2015) 775–780. [32] M. Hosseini, S.F.L. Mertens, M.R. Arshadi, Synergism and antagonism in mild steel
[8] Y. Loidreau, E. Deau, P. Marchand, M. Nourrisson, C. Logé, G. Coadou, N. Loaëc, corrosion inhibition by sodium dodecylbenzenesulphonate and hexamethylenete-
L. Meijer, T. Besson, Synthesis and molecular modelling studies of 8-arylpyrido tramine, Corros. Sci. 45 (2003) 1473–1489.
[3′,2′:4,5] thieno [3,2-d] pyrimidin-4-amines as multitarget Ser/Thr kinases in- [33] T. Murakawa, S. Nagaura, N. Hackerman, Coverage of iron surface by organic
hibitors, Eur. J. Med. Chem. 92 (2015) 124–134. compounds and anions in acid solutions, Corros. Sci. 7 (1967) 79–89.
[9] J.M. Zhao, H.B. Duan, R.J. Jiang, Synergistic corrosion inhibition effect of quinoline [34] A. Singh, Y. Lin, I.B. Obot, E.E. Ebenso, Macrocyclic inhibitor for corrosion of N80
quaternary ammonium salt and Gemini surfactant in H2S and CO2 saturated brine steel in 3.5% NaCl solution saturated with CO2, J. Mol. Liq. 219 (2016) 865–874.
solution, Corros. Sci. 91 (2015) 108–119. [35] E.S. Ferreira, C. Giacomelli, F.C. Giacomelli, A. Spinelli, Evaluation of the inhibitor
[10] X. Li, X. Xie, Adsorption and inhibition effect of two aminopyrimidine derivatives effect of L-ascorbic acid on the corrosion of mild steel, Mater. Chem. Phys. 83
on steel surface in H2SO4 solution, J. Taiwan Inst. Chem. Eng. 45 (2014) (2004) 129–134.
3033–3045. [36] M. Mobin, M. Rizvi, Inhibitory effect of xanthan gum and synergistic surfactant
[11] P.M. Dasami, K. Parameswari, S. Chitra, Corrosion inhibition of mild steel in 1 M additives for mild steel corrosion in 1 M HCl, Carbohydr. Polym. 136 (2016)
H2SO4 by thiadiazole Schiff bases, Measurement 69 (2015) 195–201. 384–393.
[12] S.W. Xie, Z. Liu, G.C. Han, W. Li, J. Liu, Z. Chen, Molecular dynamics simulation of [37] S.M.A. El Haleem, S.A. El Wanees, A. Bahgat, Environmental factors affecting the
inhibition mechanism of 3,5-dibromo salicylaldehyde Schiff’s base, Comput. Theor. corrosion behaviour of reinforcing steel. VI. Benzotriazole and its derivatives as
Chem. 1063 (2015) 50–62. corrosion inhibitors of steel, Corros. Sci. 87 (2014) 321–333.
[13] X. Li, S. Deng, H. Fu, G. Mu, Synergism between rare earth cerium(IV) ion and [38] T. Wang, X. Tian, Y. Yang, Y.W. Li, J. Wang, M. Beller, H. Jiao, Co-adsorption and
vanillin on the corrosion of cold rolled steel in 1.0 M HCl solution, Corros. Sci. 50 mutual interaction of nCO + mH2 on the Fe(110) and Fe(111) surfaces, Catal.
(2008) 3599–3609. Today 261 (2016) 82–92.
[14] M.A. Amin, Q. Mohsen, O.A. Hazzazi, Synergistic effect of I− ions on the corrosion [39] J. Liu, R. Zhu, T. Xu, Y. Xu, F. Ge, Y. Xi, J. Zhu, H. He, Co-adsorption of phosphate
inhibition of Al in 1.0 M phosphoric acid solutions by purine, Mater. Chem. Phys. and zinc(II) on the surface of ferrihydrite, Chemosphere 144 (2016) 1148–1155.
114 (2009) 908–914. [40] A. Bondi, Van der Waals volumes and radii, J. Phys. Chem. 68 (1964) 441–451.
[15] X. Li, S. Deng, H. Fu, G. Mu, Synergistic inhibition effect of rare earth cerium(IV) [41] F. Pan, F. Peng, Z. Jiang, Diffusion behavior of benzene/cyclohexane molecules in
ion and anionic surfactant on the corrosion of cold rolled steel in H2SO4 solution, poly(vinyl alcohol)-graphite hybrid membranes by molecular dynamics simulation,
Corros. Sci. 50 (2008) 2635–2645. Chem. Eng. Sci. 62 (2007) 703–710.
[16] X. Li, S. Deng, H. Fu, X. Xie, Synergistic inhibition effects of bamboo leaf extract/ [42] A. Lehmann, G. König, K.H. Rieder, Water absorption on perfect CaF2 (111) studied
major components and iodide ion on the corrosion of steel in H3PO4 solution, with He scattering: experimental evidence for ordering of nanoclusters, Phys. Rev.
Corros. Sci. 78 (2014) 29–42. Lett. 73 (1994) 3125.
[17] A. Khamis, M.M. Saleh, M.I. Awad, Synergistic inhibitor effect of cetylpyridinium [43] P. Linus, General Chemistry, third ed., Freeman and Company Press, San Francisco,
chloride and other halides on the corrosion of mild steel in 0.5 M H2SO4, Corros. 1970.
Sci. 66 (2013) 343–349. [44] Z.J. Yu, E.T. Kang, K.G. Neoh, Electroless plating of copper on polyimide films
[18] I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, The synergistic inhibitive effect and some modified by surface grafting of tertiary and quaternary amines polymers, Polymer
quantum chemical parameters of 2,3-diaminonaphthalene and iodide ions on the 43 (2002) 4137–4146.
hydrochloric acid corrosion of aluminium, Corros. Sci. 51 (2009) 276–282. [45] J. Xu, X. Han, H. Liu, Y. Hu, Synthesis and optical properties of silver nanoparticles
[19] Y. Hao, L.A. Sani, T. Ge, Q. Fang, The synergistic inhibition behaviour of tannic acid stabilized by gemini surfactant, Colloids Surf. A 273 (2006) 179–183.
and iodide ions on mild steel in H2SO4 solutions, Corros. Sci. 123 (2017) 158–169. [46] L.T. Weng, C. Poleunis, P. Bertrand, V. Carlier, M. Sclavons, P. Franquinet,
[20] C. Zhang, J.M. Zhao, Synergistic inhibition effect of imidazoline ammonium salt R. Legras, Sizing removal and functionalization of the carbon fiber surface studied
and sodium dodecyl sulfate in CO2 system, Acta Phys. Chim. Sin. 30 (2014) by combined TOF SIMS and XPS, J. Adhes. Sci. Technol. 9 (1995) 859–871.
677–685. [47] X.L. Cheng, H.Y. Ma, J.P. Zhang, X. Chen, S.H. Chen, H.Q. Yang, Corrosion of iron in
[21] J.M. Zhao, G.H. Chen, The synergistic inhibition effect of oleic-based imidazoline acid solutions with hydrogen sulfide, Corrosion 54 (1998) 369–376.
and sodium benzoate on mild steel corrosion in a CO2-saturated brine solution, [48] H.Y. Ma, X.L. Cheng, S.H. Chen, G.Q. Li, X. Chen, S.B. Lei, H.Q. Yang, Theoretical
Electrochim. Acta 69 (2012) 247–255. interpretation on impedance spectra for anodic iron dissolution in acidic solutions
[22] A.Y. Musa, R.T.T. Jalgham, A.B. Mohamad, Molecular dynamic and quantum che- containing hydrogen sulfide, Corrosion 54 (1998) 634–640.
mical calculations for phthalazine derivatives as corrosion inhibitors of mild steel in [49] I.B. Obot, N.O. Obi-Egbedi, 2,3-Diphenylbenzoquinoxaline: a new corrosion in-
1 M HCl, Corros. Sci. 56 (2012) 176–183. hibitor for mild steel in sulphuric acid, Corros. Sci. 52 (2010) 282–285.
[23] D. Wang, B. Xiang, Y. Liang, S. Song, C. Liu, Corrosion control of copper in 3.5 wt.% [50] A. Popova, M. Christov, S. Raicheva, E. Sokolova, Adsorption and inhibitive prop-
NaCl solution by domperidone: experimental and theoretical study, Corros. Sci. 85 erties of benzimidazole derivatives in acid mild steel corrosion, Corros. Sci. 46
(2014) 77–86. (2004) 1333–1350.
[24] Y. Yan, X. Wang, Y. Zhang, P. Wang, X. Gao, J. Zhang, Molecular dynamics simu- [51] M.V. Azghandi, A. Davoodi, G.A. Farzi, A. Kosari, Water-base acrylic terpolymer as
lation of corrosive species diffusion in imidazoline inhibitor films with different a corrosion inhibitor for SAE1018 in simulated sour petroleum solution in stagnant
alkyl chain length, Corros. Sci. 73 (2013) 123–129. and hydrodynamic conditions, Corros. Sci. 64 (2012) 44–54.
[25] J. Zhang, W. Yu, L. Yu, Y. Yan, G. Qiao, S. Hu, Y. Ti, Molecular dynamics simulation [52] Y. Qi, H. Luo, S. Zheng, C. Chen, D. Wang, Effect of immersion time on the hydrogen
of corrosive particle diffusion in benzimidazole inhibitor films, Corros. Sci. 53 content and tensile properties of A350LF2 steel exposed to hydrogen sulphide en-
(2011) 1331–1336. vironments, Corros. Sci. 69 (2013) 164–174.
[26] Y. Qiang, S. Zhang, L. Guo, X. Zheng, B. Xiang, S. Chen, Experimental and theo- [53] B. Tribollet, J. Kittel, A. Meroufel, F. Ropital, F. Grosjean, E.M.M. Sutter, Corrosion
retical studies of four allyl imidazolium-based ionic liquids as green inhibitors for mechanisms in aqueous solutions containing dissolved H2S. Part 2: model of the
copper corrosion in sulfuric acid, Corros. Sci. 119 (2017) 68–78. cathodic reactions on a 316L stainless steel rotating disc electrode, Electrochim.
[27] L.L. Liao, Corros. Sci. 124 (2017) 167–177, http://dx.doi.org/10.1016/j.corsci. Acta 124 (2014) 46–51.
2017.05.020. [54] J.M. Zhao, F. Gu, T. Zhao, R.J. Jiang, Corrosion inhibition performance of imida-
[28] Z. Salarvand, M. Amirnasr, M. Talebian, K. Raeissi, S. Meghdadi, Enhanced corro- zoline derivatives with different pedant chains under three flow rates in high-
sion resistance of mild steel in 1 M HCl solution by trace amount of 2-phenyl- pressure CO2 environment, Res. Chem. Intermed. 42 (2016) 5753–5764.
benzothiazole derivatives: experimental, quantum chemical calculations and mo- [55] M.J. Ondrechen, S. Gozashti, X.M. Wu, An electronic mechanism for electron
lecular dynamics (MD) simulation studies, Corros. Sci. 114 (2017) 133–145. pairing in antiferromagnetic bridged mixed-valence systems, J. Chem. Phys. 96
[29] X. Li, S. Deng, T. Lin, X. Xie, G. Du, 2-Mercaptopyrimidine as an effective inhibitor (1992) 3255–3261.

254

Вам также может понравиться