Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Corrosion Science 129 (2017) 91–101

Contents lists available at ScienceDirect

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

1,2,3-Triazole derivatives as corrosion inhibitors for mild steel in acidic T


medium: Experimental and computational chemistry studies

Qi Maa, Sijun Qia, Xiaohong Hea, Yongming Tanga, , Gang Lub
a
School of Chemistry and Molecular Engineering, Nanjing Tech University, Nanjing 211816, PR China
b
College of Materials Science and Engineering, Nanjing Tech University, Nanjing 210009, PR China

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

Keywords: (1-Benzyl-1H-1,2,3-triazole-4-yl)methanol (BTM) and (1-(pyridin-4-ylmethyl)-1H-1,2,3- triazole-4-yl)methanol


A. Mild steel (PTM) were prepared and investigated as corrosion inhibitors for mild steel in 1.0 M HCl. It is found that PTM
B. EIS functions as more effective inhibitor than BTM in the concentration range of 0.2 −1.0 mM. Computational
B. AFM chemistry studies show that the triazole derivatives can adsorb on the mild steel surface by sharing the lone pair
B. Modelling studies
electrons of N atoms with iron atoms or by accepting electrons from iron surfaces. Due to its strong interaction
C. Acid inhibition
with the mild steel surface in aqueous system, the pyridine segment should be responsible for the higher in-
hibition efficiency of PTM.

1. Introduction inhibition efficiency at a low concentration [35–37]. However, there is


still a lack of interaction mechanism between 1,2,3-triazole derivatives
Corrosion inhibition efficacy of organic inhibitors strongly depends and metal surface.
on their adsorption properties [1,2]. It has been reported that adsorp- In the present study, we synthesize two new 1,2,3-triazole deriva-
tion is mainly related to the presence of π-electrons of aromatic rings tives, (1-benzyl-1H-1,2,3-triazole-4-yl)methanol (BTM) and (1-(pyridin-
and heteroatoms in the molecular structures [3–6]. Based on that, the 4-ylmethyl)-1H-1,2,3- triazole-4-yl)methanol (PTM) as shown in Fig. 1,
organic compounds containing nitrogen, sulfur and oxygen atoms have and investigate their corrosion inhibition for mild steel in 1 M HCl so-
attracted more and more attention in the studies on corrosion inhibition lution. Based on the theoretical study, furthermore, we attempt to
of mild steel in acidic media [7–11]. A typical example is triazole de- clarify the interaction between the triazole derivatives and iron surface
rivatives, especially 1,2,4-triazole derivative. Actually, unmodified and to understand the difference in inhibition efficiency between BTM
1,2,4-triazole is not a good corrosion inhibitor for mild steel in acidic and PTM.
solution [12]. A common pathway to enhance the inhibition efficiency
is to modify the structures of the triazole compounds with various 2. Experimental and computational details
substituents. Lagrenée et al. synthesized a serial of 3,5-disubstituted
1,2,4-triazole compounds and found the triazole derivatives exhibit 2.1. Synthesis of BTM and PTM
excellent corrosion inhibition efficacy [13–20]. It has been observed
that, moreover, 3 and/or 4-substituted triazole derivatives by thiol- or BTM and PTM were prepared through a modified procedure re-
amino-group can function as highly efficient corrosion inhibitors be- ported previously [38,39], as shown in Fig. 1.
cause the modifications of those groups make stronger adsorptive layers
of the triazole compounds on mild steel [21–30]. In addition, Schiff’s 2.1.1. Benzyl azide
base derivatives of triazole have also been investigated as corrosion To a solution of benzyl chloride (2.291 mL, 0.02 mol) and water
inhibitors of mild steel in acidic medium [31–34]. (4 mL) in DMSO (20 mL), sodium azide (2.600 g, 0.04 mol) was added.
Regardless of the wide investigation on 1,2,4-triazole compounds as The resulting mixture was stirred at room temperature for 24 h, and
corrosion inhibitors of mild steel in acidic solutions, much less attention then the reaction mixture was extracted with petroleum ether
has been paid to 1,2,3-triazole derivatives. Recently, several studies (30 mL × 4). The combined organic layer was washed with water
have shown that 1,2,3-triazole derivatives with various substituents can (30 mL × 2) and saturated brine (30 mL), respectively, and then was
strongly adsorb on mild steel surface and achieve adequate corrosion dried over anhydrous Na2SO4. After removing the desiccant by


Corresponding author.
E-mail address: tangym@njtech.edu.cn (Y. Tang).

http://dx.doi.org/10.1016/j.corsci.2017.09.025
Received 26 February 2017; Received in revised form 26 September 2017; Accepted 28 September 2017
Available online 04 October 2017
0010-938X/ © 2017 Elsevier Ltd. All rights reserved.
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Fig. 1. Synthesis routes of BTM and PTM.

filtration, the mixture was evaporated in vacuo to give colorless oily coupons with the size of 5.0 cm × 2.5 cm × 0.2 cm. Weight loss tests
liquid. The product was used without further purification in the next were carried out in non-deaerated solutions. The coupons were fully
reaction. Yield 2.370 g (87%). 1H NMR (400 MHz, CDCl3): δ 7.40-7.30 immersed in 1.0 M HCl solutions containing various concentrations of
(m, 5H, ArH), 4.32 (s, 2H, CH2) the inhibitors at 25 ± 0.1 °C for 8 h. After being withdrawn from the
test solutions, the mild steel specimens were washed with deionized
2.1.2. (1-benzyl-1H-1,2,3-triazole-4-yl)methanol (BTM) water, scrubbed with absorbent cotton to remove the corrosion pro-
Propargyl alcohol (0.640 mL, 11 mmol) was added to a mixture of ducts on the surfaces and then washed with anhydrous ethanol, fol-
benzyl azide (1.330 g, 10 mmol), CuSO4 (0.500 g, 2 mmol) and sodium lowed by drying at room temperature. Finally, the specimens were
ascorbate (0.790 g, 4 mmol) in tBuOH:H2O (10 mL:10 mL). The re- weighed for the calculation of corrosion rate.
sulting yellow mixture was stirred at room temperature for 16 h under After weight loss tests, the surfaces of the specimens were im-
an atmosphere of nitrogen. Then the reaction mixture was diluted and mediately examined by both atomic force microscope (AFM, Nanosurf
extracted with ethyl acetate (30 mL × 3). The combined organic layers FlexAFM, Switzerland) and scanning electron microscope (SEM, Hitachi
were washed with water (30 mL × 2) and saturated brine (20 mL), S-4800, Japan). For AFM measurements, the surface morphology on a
respectively. The organic phase was then dried over anhydrous MgSO4, 50 μm × 50 μm scale was scanned in tapping mode using silicon can-
filtered and evaporated under reduced pressure to give acicular solid. tilevers with a resonance frequency of 120–250 kHz and a force con-
Yield: 1.342 g (71%). 1H NMR (400 MHz, CDCl3): δ 7.45 (s, 1H, NeCH), stant of 20–100 N/m. All images were processed using a C3000 soft-
7.38-7.34 (m, 3H, CHAr), 7.27-7.24 (m, 2H, CHAr), 5.49 (s, 2H, Ar-CH2), ware of Nanosurf. For SEM measurements, an accelerating voltage of
4.73 (s, 2H, CH2), 3.58 (bs, 1H, OH). 5000 V was applied.

2.1.3. 4-(azidomethyl)pyridine
2.2.2. Electrochemical tests
To a solution of 4-(bromomethyl)pyridine hydrobromide (1.000 g,
Electrochemical impedance spectroscopy (EIS) measurements were
3.95 mmol) and potassium carbonate (0.550 g, 3.95 mmol) in DMF
carried out using the classical three-electrode system where the counter
(10 mL). Then sodium azide (0.390 g, 5.93 mmol) was added. The re-
electrode was a platinum electrode and a saturated calomel electrode
action mixture was stirred at room temperature for 3 h. Afterwards the
(SCE) was used as reference. To reduce Ohmic polarization, a Luggin
resulting reaction mixture was diluted and extracted with CH2Cl2
capillary was connected to the reference. An epoxy-encapsulated mild
(40 mL × 3). The combined organic layers were washed with water
steel electrode with the exposed area of 0.785 cm2 was employed as
(30 mL x 2), dried over anhydrous Na2SO4, filtered and evaporated
working electrode. Prior to use, the mild steel electrode was abraded
under reduced pressure to give oily liquid. Yield: 0.470 g (88%). 1H
with SiC sandpapers (up to 2000 grit), rinsed with deionized water and
NMR (400 MHz, CDCl3): δ 8.63-8.62 (m, 2H, CH-NeCH), 7.27-7.24 (m,
ethanol successively and finally dried under nitrogen. Before EIS mea-
2H, CH-CeCH), 4.40 (s, 2H, N3eCH2).
surements, the mild steel electrode was immersed in the test solutions
under unstirred condition for 60 min until a steady potential. EIS
2.1.4. (1-(pyridin-4-ylmethyl)-1H-1,2,3-triazole-4-yl)methanol (PTM)
measurements were performed at open-circuit potential (OCP) over the
Propargyl alcohol (0.256 mL, 4.4 mmol) was added to a mixture of
range of 0.01 Hz to 100 kHz with the amplitude of 5 mV for the sine
4-(azidomethyl)pyridine (0.536 g, 4 mmol), CuSO4 (0.128 g, 0.8 mmol)
wave signal. All EIS data were analyzed by Zview software.
and sodium ascorbate (0.316 g, 1.6 mmol) in tBuOH:H2O
(10 mL:10 mL). The resulting yellow mixture was stirred at 50 °C for
16 h under nitrogen, and then was diluted and extracted with ethyl 2.3. Quantum chemical method
acetate (30 mL × 3). The organic layer was dried over anhydrous
MgSO4, filtered and evaporated under reduced pressure to give white Density functional theory (DFT) calculations were performed to
solid. Yield: 0.228 g (30%). 1H NMR (400 MHz, CDCl3): δ 8.56 (d, 2H, optimize the geometries of inhibitor molecules using Dmol3 module in
CH-NeCH), 7.57 (s, 1H, NeCH-C), 7.10 (d, 2H, CH-CeCH), 4.80 (s, 2H, Materials Studio software 6.0. In the DFT calculations, the generalized
N3eCH2), 3.97 (bs, 1H, OH). gradient approximation (GGA) of the Perdew-Burke-Ernzerhof (PBE)
formula was used for the electronic exchange-correlation potential
2.2. Corrosion inhibition tests [40], and all electron calculations were performed with double-numeric
basis set (DNP 4.4) [41]. Geometry optimizations were performed
2.2.1. Weight loss measurements and surface morphology without any symmetry constraint and vibrational analysis was carried
Mild steel specimens containing 0.17 wt.% C, 0.20 wt.% Si, 0.37 wt. out to ensure the optimized structures being the minimum point on
% Mn, 0.03 wt.% S, 0.01 wt.% P, and balance iron were cut into potential energy surface.

92
Q. Ma et al. Corrosion Science 129 (2017) 91–101

2.4. Molecular dynamic simulations corrosive medium. It is clear, furthermore, that the inhibition efficiency
of PTM is higher than that of BTM, in particular at the low con-
Discover module of Materials Studio software 6.0 was used to model centrations, suggesting that the stronger adsorption of PTM on mild
the interaction between inhibitors and iron surface. Prior to building steel surface than BTM.
the simulation box, a bcc unit cell of iron was optimized to a minimum To further verify the efficacy of the compounds for corrosion in-
point of energy, and then a p(6 × 6) supercell of Fe (1 0 0) containing hibition of mild steel, AFM measurements were performed to qualita-
six atomic layers were built on the basis of the optimized unit cell. A tively characterize the topographical change in the absence and pre-
slab containing water and inhibitor molecules was added to the Fe (1 0 sence of inhibitors. Fig. 2 shows the 2-D and 3-D images of mild steel
0) surface by the “build layers” tool. The constructed box with periodic surfaces abraded before immersion as well as after 8 h immersion
boundary conditions has a size of 17.2 Å × 17.2 Å × 30.3 Å. MD si- without and with 1 mM inhibitors. The abraded mild steel surface is
mulation was performed using COMPASS force field at 298 K, NVT relatively smooth with distinct abrading scratches (Fig. 2a and b). After
ensemble, with a time step of 0.1 fs and simulation time of 100 ps as all the immersion in 1.0 M HCl in the absence of inhibitor, however, the
Fe atoms were fixed at the bulk positions and both inhibitors and water mild surface is highly damaged with deep holes and pits (Fig. 2c and d).
molecules were allowed to fully relax. After simulations, the interaction In comparison, the surface topographies of the mild steel surfaces were
energy (Eint) between inhibitor molecule and Fe (1 0 0) plane was significantly improved by the additions of BTM and PTM and even the
calculated as follows [42]: abrading scratches can be still made out after 8 h immersion
(Fig. 2e–h), which is attributed to the formation of compact adsorptive
Eint = Etot − (Esurf/H2O + Emol ) (1)
layers of BTM and PTM and consequently isolation of mild steel surface
where Etot is total energy of simulation system, Esurf/H2O is energy of iron from the corrosive solution.
surface together with water molecules, and Emol is energy of free in- Fig. 2. AFM images of mild steel specimens before and after im-
hibitive molecule. mersion in 1.0 M HCl. (a) and (b): before immersion, (c) and (d): after
immersion in uninhibited solution, (e) and (f): after immersion with
3. Results and discussion 0.8 mM BTM, (g) and (h): after immersion with 0.8 mM PTM.
In addition to AFM, the corrosion samples were examined by SEM as
3.1. Weight loss and surface analysis well. The abrading scratches can be distinctly observed from the surface
micrograph of mild steel before immersion (Fig. 3a). The surface of
Table 1 shows the corrosion rate (v), inhibition efficiency (ƞ) and mild steel in the uninhibited solution is drastically damaged with a
surface coverage (θ) in the absence and presence of 1,2,3-triazole ragged morphology (Fig. 3b). From Fig. 3c and d, the aggressive effect
compounds for mild steel in 1.0 HCl solution at 25 °C. The corrosion of hydrochloric acid is strongly mitigated by the presence of inhibitors.
rate (v) is calculated as follows: After immersion, the surfaces in the inhibited solutions remain rela-
tively even, and the abrading scratches can still be observed. Further-
w0 − w
v= more, it is evident that the mild steel surface is better protected by PTM
St (2)
than by BTM, indicating that PTM is more effective in inhibiting cor-
where w0 and w are the weight values without and with inhibitor, re- rosion of mild steel than BTM in the present case.
spectively, S is the surfaces area of mild steel coupons and t is the im-
mersion time. The inhibition efficiency and surface coverage are cal- 3.2. Electrochemical impedance spectroscopy
culated according to the following equations:
v 0 − v1 The corrosion inhibition properties of BTM and PTM for mild steel
η% = × 100 in 1.0 M HCl are examined by EIS at the open circuit potential after the
v0 (3)
immersion of 60 min. Prior to EIS measurements, OCP curves of mild
η steel are recorded in the absence and presence of the inhibitors, as
θ=
100 (4) shown in Fig. 4. The OCP values slights increase with time in the be-
where v0 and v1 are the corrosion rate values in the absence and pre- ginning of the immersion. After about 14 min, a steady potential,
sence of inhibitors, respectively. −0.528 V for the uninhibited solution and −0.506 V for the inhibited
From Table 1, with the increase in the concentration of inhibitors, solutions, respectively, is reached for the three systems. The addition of
the corrosion rate of mild steel decreases significantly and the corrosion the inhibitors shifts the open circuit potential toward positive around
inhibition efficiency increases gradually. The inhibition efficiency va- 20 mV relative to the OCP in the blank solution.
lues for both compounds are more than 95% at the concentration of Fig. 5 shows the Nyquist plots of mild steel obtained in 1.0 M HCl
0.8 mM. The good corrosion inhibition performance is attributed to the solutions with and without the inhibitors, and the corresponding Bode
formation of compact adsorptive layers protecting mild steel from the plots are presented in Fig. 6. As shown in Fig. 5, the impedance spectra
in the absence of and presence of inhibitors are presented only in the
Table 1 first quadrant, indicating that there is only one time constant corre-
Weight-loss measurements for mild steel in 1.0 M HCl. sponding to the charge transfer process. That is to say, the presence of
BTM and PTM in the corrosive solution does not change the corrosion
Compounds Conc. (mM) v (g m−2 h−1) η (%) θ mechanism of mild steel in 1.0 M HCl. Moreover, the capacitive loops
0 74.31 – –
are slightly depressed as semi-circular shape with center under the real
axis. The deformation of the capacitance semicircle, referred to as fre-
BMT 0.2 31.06 58.2 0.582 quency dispersion, is attributed to the surface heterogeneity [43,44].
0.4 17.61 76.3 0.763 For solid electrodes in strongly aggressive media, the surface hetero-
0.6 8.62 88.4 0.884
0.8 2.67 96.4 0.964
geneity may result from surface roughness, impurities, dislocations,
1.0 1.63 97.8 0.978 grain boundaries, fractality, distribution of the activity centers, in-
PMT 0.2 8.10 89.1 0.891 hibitor’s adsorption, and formation of porous layers [45]. The same
0.4 3.94 94.7 0.947 conclusion can be drawn from the Bode plots in Fig. 6 where only a
0.6 1.79 97.6 0.976
phase peak is observed at the middle frequency range in the absence
0.8 1.34 98.2 0.982
1.0 0.74 99.0 0.990 and presence of inhibitors and the deviations of the phase peak values
from 90° are attributed to the dispersing effect. It is apparent that the

93
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Fig. 2. AFM images of mild steel specimens before and after immersion in 1.0 M HCl. (a) and (b): before immersion, (c) and (d): after immersion in uninhibited solution, (e) and (f): after
immersion with 0.8 mM BTM, (g) and (h): after immersion with 0.8 mM PTM.

size of the capacitive loop in Fig. 5 as well as the low frequency im- ZCPE = Y 0−1 (jω)−n (5)
pedance modulus in Fig. 6 increases as the concentration of the in- −1 n −2
where Y0 is a proportionality coefficient, Ω s cm , ω is the angular
hibitors increases and consequently the protection efficacy increases,
frequency and n has the meaning of a phase shift. For n = 0, CPE re-
indicating that the formation of more integrated protective adsorption
presents a resistance, for n = 1 a capacitance and for n = −1 an in-
layer with the increase in the concentration of inhibitors [46]. The
ductance. The “double layer capacitance” values (Cdl) are calculated as
impedance data are fitted using the circuit in Fig. 7. Rs is the solution
follows:
resistance, Rp represents the polarization resistance and CPE is constant
phase element to replace a double layer capacitance (Cdl) to obtain a Cdl = Y0 (ωm"e;)n − 1 (6)
more accurate fit of experimental data set. The impedance of a constant
phase element is given in Equation (5): where ωm" is the angular frequency corresponding to the maximum
imaginary part of the impedance spectrum. The impedance parameters

Fig. 3. SEM images of mild steel specimens before


and after immersion in 1.0 M HCl. (a): before im-
mersion, (b): after immersion in uninhibited solu-
tion, (c): after immersion with 0.8 mM BTM, (d):
after immersion with 0.8 mM PTM.

94
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Fig. 4. Open circuit potential curves for mild steel in 1.0 M HCl in the
absence and presence of inhibitors.

the corrosion of metals. The comparative investigation reveals that PTM


exhibits higher corrosion inhibition efficiency than BTM at the same
concentration, in good accordance with the results from the weight-loss
measurements. It should be noted, however, that there is a slight dif-
ference in inhibition efficiency value between EIS and weight loss tests,
which may be related to the difference in experimental condition such
as immersion time [47]. In addition, it can be seen from Table 2 that the
electric double layer capacitance decreases with the concentrations of
inhibitors, which is attributed to the replacement of the water mole-
cules adsorbed on the electrode surface by the inhibitor molecules,
leading to the decrease of the dielectric constant or the increase of the
thickness of the double electricity layer [48,49].

3.3. Adsorption isotherm

Basic information on the interaction between the inhibitors and the


metal surface can be provided by adsorption isotherm [15]. The surface
coverage values listed in Table 1 are graphically tested to allow fitting
of a suitable adsorption isotherm. The plots of Cinh/θ against Cinh for
BTM and PTM present straight lines with almost unit slope (Fig. 8),
indicating that the adsorptions of the compounds on the mild steel
surface obey the Langmuir adsorption isotherm:
Cinh 1
= + Cinh
θ K ads (8)
where Kads is the equilibrium constant of the adsorption process. The
0
standard adsorption free energy ( ΔGads ) can be estimated from the
following equation:

Fig. 5. Nyquist plots for mild steel electrode in 1.0 M HCl without and with inhibitors. 1 ΔG 0
Solid lines show fitted lines.
K ads = exp ⎜⎛− ads ⎞⎟
55.5 ⎝ RT ⎠ (9)
The 0
ΔGads values are −31.3 and −36.0 kJ/mol for BTM and PTM, re-
obtained from the fitting of experimental data are listed in Table 2. The
spectively. From the structures of the two compounds, there are
inhibition efficiency (ƞ) is calculated as follows:
abundant lone-pair electrons on the N atoms of triazole segments which
Rp − R p0 may be donated to vacant d-orbital of Fe atoms to form coordinate
η% = × 100 bonds. Also, the aromatic segments can enhance the adsorption of those
Rp (7)
molecules by donating electrons to the iron surface or accepting elec-
where R p0 and Rp are the values of polarization resistance in the unin- trons from the iron surface. In addition, the adsorption free energy
hibited and inhibited solutions, respectively. value of PTM is more negative than that of BTM, suggesting the
As seen from Table 2, Rp value reflecting the difficulty level of stronger interaction between PTM and the iron surface. An interesting
corrosion reaction increases significantly with the concentration of in- fact is that the sole difference in structure between BTM and PTM
hibitors, indicating that the addition of inhibitors effectively inhibits molecules is the structures of aromatic segments. As shown in Fig. 1, the

95
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Fig. 6. Bode plots for mild steel electrode in 1.0 M HCl without and with inhibitors. Solid lines show fitted lines.

and why the higher inhibition efficiency is achieved by PTM instead of


BTM. To answer the questions, quantum chemical calculations and
molecular dynamic simulations are used to study the interaction be-
tween the triazole derivatives and iron surfaces.
The optimized structures of BTM and PTM by DFT calculations
taking account of solvent effect are displayed in Fig. 9 and both highest
occupied molecular orbital (HOMO) and lowest unoccupied molecular
orbital (LUMO) are also analyzed. It can be observed from Fig. 9 that
the two segments of those molecules, triazole ring and benzene ring for
BTM and pyridine ring for PTM, are noncoplanar. Furthermore, HOMOs
Fig. 7. Equivalent circuit used to fit the impedance data in Fig. 5.
are much more concentrated on aromatic rings rather than triazole
rings, especially for PTM the distribution of HOMO on triazole ring is
aromatic segment of PTM is a pyridine ring instead of the benzene ring very insignificant compared to that on pyridine ring. This suggests that
of BTM. It is expected, therefore, that the pyridine segment should be the aromatic ring segments play the more important role in donating
responsible for the stronger adsorption of PTM than BTM since the N electrons to the metal atoms when the compounds adsorb on iron. For
atoms of pyridine segment have a strong ability to donate electrons to BTM the LUMO density on the triazole ring is significantly higher than
mild steel surface. More evidence on the interaction between the on the benzene ring while for PTM LUMOs are dominantly concentrated
compounds and the iron surface can be found in the following com- on the pyridine ring rather than the triazole ring. That is to say, the
putational studies. triazole ring segment of BTM has stronger electron-accepting ability
than the benzene ring segment. Nevertheless, the pyridine ring segment
3.4. Quantum chemical calculations of PTM acts as electron-accepting part as well as electron-donating part
in adsorption of PTM on metal. Frontier molecular orbital energies of
Corrosion tests show that both BTM and PTM can effectively protect the two molecules (EHOMO and ELUMO) are listed in Table 3. According
mild steel from corrosion in acidic media and, moreover, PTM exhibits to the frontier molecular orbital theory, a high EHOMO value for a mo-
better inhibition capability than BTM at the same concentration. The lecule implies the tendency of the molecule to donate electrons to the
next questions are that how the compounds function on iron surface appropriate vacant d-orbital of the metal atoms. On the other hand, a

96
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Table 2
EIS parameters for mild steel in 1.0 M HCl.

Compounds Conc. (mM) Rs (Ω cm2) Rp (Ω cm2) Y0 × 104 (Ω−1 sn cm−2) n Cdl (μF cm−2) η (%)

0 0.79 4.58 6.84 0.79 262 –

BMT 0.2 0.89 9.55 5.54 0.78 203 52.0


0.4 0.76 16.97 4.79 0.72 161 73.0
0.6 0.53 19.44 4.11 0.75 154 76.4
0.8 0.57 33.52 2.97 0.70 114 86.3
1.0 0.47 42.72 2.61 0.69 97 89.3
PMT 0.2 0.98 29.05 2.98 0.84 178 84.2
0.4 0.99 37.83 2.60 0.80 135 87.9
0.6 0.97 42.38 1.99 0.80 104 89.2
0.8 0.96 45.37 1.90 0.78 95 89.9
1.0 0.97 46.67 1.84 0.78 91 90.2

low ELUMO value for a molecule indicates the relative tendency of the accepted due to the extremely small number [54]. The parameter ΔN is
molecule to accept electrons from metallic orbital during back-donation a measure of the electron transfer between molecule and metal.
[50,51]. From Table 3, only slight difference (less than 0.05 eV) in Namely, the electrons are transferred from molecule to metal if
EHOMO value between BTM and PTM is observed, indicating a similar ΔN > 0and vice versa if ΔN < 0 [55,56]. From Table 3, it is apparent
ability to donate electrons for the two derivatives. However, the ELUMO that in the case of PTM the electrons are transferred from iron to PTM
value of PTM is 0.459 eV lower than that of BTM, indicating that PTM molecule due to the negative ΔN value, but an opposite effect is ob-
has a stronger tendency to accept electrons from metallic orbital than tained for the adsorption of BTM on iron. Combined with the results
BTM as they adsorb on iron. Depending on the stronger electron-ac- from frontier molecular orbital analysis, accordingly, it is concluded
cepting ability, therefore, PTM can more powerfully adsorb on the that the stronger ability of PTM to accept electrons should be mainly
metallic surface to achieve better corrosion inhibition effectiveness responsible for the higher inhibition efficiency compared to BTM.
compared to BTM.
Although the above result may partly explain why PTM exhibits
higher inhibition efficiency than BTM, it is still necessary to clarify 3.5. Molecular dynamic simulations
whether electrons are transferred from inhibitor molecule to metal or
otherwise as the inhibitors adsorb on the metallic surface. A parameter, Quantum chemical calculations show the noncoplanar conforma-
fraction of electrons transferred (ΔN) as defined in Equation (10), is tions between triazole segment and aromatic segment for both BTM and
introduced to explain the electron transfer between inhibitors and iron PTM. As they adsorb on metallic surface, however, we are wondering if
[42,52]. the molecules can maintain the noncoplanar conformations on iron. MD
simulations can provide a good insight into the interaction between
ϕ − χinh inhibitor and iron surface. MD simulations are first performed in a
ΔN =
2(γFe − γinh ) (10) vacuum slab without any solvation effect, and the corresponding stable
configurations for BTM and PTM systems are presented in Fig. 10. In the
where ϕ is work function of iron surface having a value of 3.91 eV for stable configurations, both BTM and PTM molecules are nearly parallel
Fe (1 0 0) plane [53], χinh is absolute electronegativity of inhibitor, to the Fe (1 0 0) plane and exhibit an approximately planar con-
χinh = − (EHOMO + ELUMO)/2, and γ is global hardness of inhibitor or formation. The corresponding interaction energy values are listed in
iron: for inhibitor γinh = − (EHOMO − ELUMO)/2, and for iron γFe ≈ 0is Table 4. It can be seen that the Eint value of BTM/Fe system is slightly

Fig. 8. Adsorption isotherms for BTM and PTM on mild steel.

97
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Fig. 9. Optimized structures and molecular orbital densities of BTM (a


and b) and PTM (c and d). Left: HOMO, right: LUMO.

Table 3 Table 4
Quantum chemical parameters of BTM and PTM. Interaction energies between inhibitors and Fe (1 0 0) plane.

Molecules EHOMO (eV) ELUMO (eV) ΔN μ (debye) Systems Eint (kJ/mol)

BTM −6.127 −1.335 0.037 6.22 Fe (1 0 0)/BTM −407.9


PTM −6.096 −1.794 −0.008 7.54 Fe (1 0 0)/PTM −402.2
Fe (1 0 0)/BTM + H2O −454.6
Fe (1 0 0)/PTM + H2O −568.0
more negative than that of PTM/Fe system, which indicates that BTM
molecules can more strongly adsorb on iron surface to achieve better
inhibition effectiveness than PTM molecules. Apparently this result is in an aqueous system. The triazole segment of BMT molecule still remains
disagreement with the experimental results. That is to say, in the pre- parallel to the Fe (1 0 0) plane but the benzene segment exhibits a tilted
sent case MD simulations in the vacuum slab is not a reliable measure to orientation against the metallic plane. It is inferred that the hydro-
study the interaction mechanism between inhibitors and metallic sur- phobicity of benzene segment should be responsible for the adsorption
face. In fact, the adsorption of inhibitor occurs at a liquid/solid inter- orientation. In aqueous solution, water molecules or hydrated ions can
face which is totally overlooked by the MD simulations in the vacuum preferentially adsorb on iron surface with hydrophilicity to form a
slab. Therefore, solvation effect is taken into consideration in the fol- stable hydration layer. As shown in Fig. 11, the distribution density of
lowing MD simulations. water molecules on iron surface is significantly higher than that in bulk
Fig. 11 displays the equilibrium configurations for BTM and PTM in solution. As a consequence, the hydration layer hinders the interaction

Fig. 10. Equilibrium configurations of BTM (a and b) and PTM (c and


d) in vacuum slabs. Top: sideview, bottom: topview.

98
Q. Ma et al. Corrosion Science 129 (2017) 91–101

Fig. 11. Equilibrium configurations of BTM (a and b) and PTM (c and


d) in aqueous systems. Top: sideview, bottom: topview.

between benzene segment and iron atoms even though the quantum
chemical calculations show that the segment has a strong tend to do-
nate electrons to vacant d-orbital of iron atoms. However, a parallel
mode is observed for the adsorption of PTM in the aqueous system, and
the coplanar conformation between the triazole segment and the pyr-
idine segment is attained. This should be attributed to the strong in-
teraction between the pyridine segment and iron atoms. From the
quantum chemical calculations, the pyridine segment acts as electron
donator as well as electron acceptor to form an adsorptive bond with
iron atoms. In addition, due to its strong hydrophilicity, pyridine seg-
ment can more easily approach the iron surface than the benzene seg-
ment of BTM. Quantum chemical calculations show, actually, that the
dipole moment of PTM is significantly higher than that of BTM
(Table 3). Compared to BTM, evidently, the adsorption mode of PTM
can block more active corrosion sites on iron surface at the same con-
centration, partly explaining the better inhibition effectiveness of PTM
than of BTM. The Eint values between inhibitors and Fe (1 0 0) plane in
the aqueous systems can be found in Table 4. The interaction energy
between PTM and Fe (1 0 0) plane is more negative than that between
BTM and Fe (1 0 0) plane, indicating that PTM can more tightly adsorb
on iron surface and further that PTM can achieve a better corrosion
inhibition effectiveness than BTM. This modelling conclusion is in good
agreement with that of corrosion inhibition tests.
In order to further clarify the difference between adsorptions of
BTM and PTM on iron, radial distribution function (RDF) between N Fig. 12. RDF curves of N-Fe in the simulation systems. Bottom for N on triazole segment
of BTM, middle for N on triazole segment of PTM and top for N on pyridine segment of
atoms and Fe atoms is analyzed after MD simulations (Fig. 12). The
PTM.
highest peak of RDF curve represents the most probable distance be-
tween N atoms and Fe atoms. The distances between Fe atoms and N
atoms on the triazole segments for BTM and PTM are 3.15 Å and 3.25 Å, indicating a very strong interaction between pyridine segment and iron
respectively, confirming the close contact between triazole segments of surface. This is in good agreement with the result of quantum chemical
those molecules and iron surface. For PTM, additionally, the distance calculations where it is mentioned that the pyridine segment play a key
between N atom on pyridine segment and Fe atoms is 3.05 Å, slightly role in donating electrons to metal atoms as well as accepting electrons
less than the distance between triazole segment and iron atoms, from metal atoms.

99
Q. Ma et al. Corrosion Science 129 (2017) 91–101

4. Conclusions (2013) 123–133.


[18] M.E. Belghiti, Y. Karzazi, A. Dafali, B. Hammouti, F. Bentiss, I.B. Obot, I. Bahadur,
E.E. Ebenso, Experimental quantum chemical and Monte Carlo simulation studies of
Both BTM and PTM at a concentration more than 0.8 mM can ef- 3,5-disubstituted-4-amion-1,2,4-triazoles as corrosion inhibitors on mild steel in
fectively prevent corrosion of mild steel from the acidic medium. In acidic medium, J. Mol. Liquids 218 (2016) 281–293.
[19] F. Bentiss, M. Traisnel, L. Gengembre, M. Lagrenée, A new triazole derivatives as
comparison with BTM, however, PTM exhibits higher inhibition effi- inhibitor of the acid corrosion of mild steel: electrochemical studies, weight loss
ciency in the concentration range of this study and more negative determination, SEM and XPS, Appl. Surf. Sci. 152 (1999) 237–249.
standard adsorption free energy. Although the two compounds can both [20] B.E. Mehdi, B. Mernari, M. Traisnel, F. Bentiss, M. Lagrenée, Synthesis and com-
parative study of the inhibitive effect of some new triazole derivatives towards
adsorb on metal surface through the electron donation or acceptance corrosion of mild steel in hydrochloric acid solution, Mater. Chem. Phys. 77 (2002)
between the triazole segments and iron atoms, the interaction of the 489–496.
pyridine segment of PTM with the iron surface is stronger than that of [21] M.A. Quraishi, R. Sardar, Aromatic triazoles as corrosion inhibitors for mild steel in
acidic environments, Corrosion 58 (2002) 748–755.
the benzene segment of BTM, resulting in the parallel mode of PTM
[22] A.Y. Musa, A.A.H. Kadhum, A.B. Mohamad, M.S. Takriff, Experimental and theo-
molecules to the mild steel surface and further better blocking effect of retical study on the inhibition performance of triazole compounds for mild steel,
PTM for the mild steel surface than of BTM. Corrosi. Sci. 52 (2010) 3331–3340.
[23] B.D. Mert, M.E. Mert, G. Kardaş, B. Yazıcı, Experimental and theoretical in-
vestigation of 3-amino-1,2,4-triazole-5-thiol as a corrosion inhibitor for carbon steel
Acknowlegements in HCl medium, Corrosi. Sci. 53 (2010) 4265–4272.
[24] A. Lesar, I. Milošev, Density functional study of the corrosion inhibition properties
This work was supported by Natural Science Foundation of Jiangsu of 1,2,4-triazole and its amino derivatives, Chem. Phys. Lett. 483 (2009) 198–203.
[25] H.H. Hassan, E. Abdelghani, M.A. Amin, Inhibition of mild steel corrosion in hy-
Province (BK20161540). drochloric acid solution by triazole derivatives Part I. Polarization and EIS studies,
Electrochim. Acta 52 (2007) 6359–6366.
Appendix A. Supplementary data [26] H.H. Hassan, Inhibition of mild steel corrosion in hydrochloric acid solution by
triazole derivatives Part II. Time and temperature effects and thermodynamic
treatments, Electrochim. Acta 53 (2007) 1722–1730.
Supplementary data associated with this article can be found, in the [27] M.K. Awad, M.R. Mustafa, M.M.A. Elnga, Computational simulation of the mole-
online version, at http://dx.doi.org/10.1016/j.corsci.2017.09.025. cular structure of some triazoles as inhibitors for the corrosion of metal surface, J.
Mol. Struct.: Theochem. 959 (2010) 66–74.
[28] M.A. Quraishi, H.K. Sharma, 4-Amino-3-butyl-1,2,4-triaole: a new corrosion in-
References hibitor for mild steel in sulphuric acid, Mater. Chem. Phys. 78 (2002) 18–21.
[29] J. Cruz, E. Garcia-Ochoa, M. Castro, Experimental and theoretical study of the 3-
amino-1,2,4-triazole and 2-aminothiazole corrosion inhibitors in carbon steel,
[1] R. Yıldız, An electrochemical and theoretical evaluation of 4,6-diamino-2- pyr-
J.Electrochem. Soc. 150 (2003) B26–B35.
imidinethiol as a corrosion inhibitor for mild steel in HCl solutions, Corros. Sci. 90
[30] S. John, A. Joseph, Electroanalytical surface morphological and theoretical studies
(2015) 544–553.
on the corrosion inhibition behavior of different 1,2,4-triazole precursors on mild
[2] A. Zarrouk, B. Hammouti, T. Lakhlifi, M. Traisnel, H. Vezin, F. Bentiss, New 1H-
steel in 1 M hydrochloric acid, Mater. Chem. Phys. 133 (2012) 1083–1091.
pyrrole-2,5-dione derivatives as efficient organic inhibitors of carbon steel corro-
[31] H.-L. Wang, R.-B. Liu, J. Xin, Inhibiting effects of some mercapto-triazole deriva-
sion in hydrochloric acid medium: electrochemical, XPS and DFT studies, Corros.
tives on the corrosion of mild steel in 1.0 M HCl medium, Corros. Sci. 46 (2004)
Sci. 90 (2015) 572–584.
2455–2466.
[3] A. Dutta, S. Kr Saha, P. Banerjee, D. Sukul, Correlating electronic structure with
[32] K.R. Ansari, M.A. Quraishi, A. Singh, Schiff’s base of pyridyl substituted triazoles as
corrosion inhibition potentiality of some bis-benzimidazole derivatives for mild
new and effective corrosion inhibitors for mild steel in hydrochloric acid solution,
steel in hydrochloric acid: combined experimental and theoretical studies, Corros.
Corros. Sci. 79 (2014) 5–15.
Sci. 98 (2015) 541–550.
[33] W. Li, Q. He, C. Pei, B. Hou, Experimental and theoretical investigation of the ad-
[4] L. Li, X. Zhang, S. Gong, H. Zhao, Y. Bai, Q. Li, L. Ji, The discussion of descriptors for
sorption behavior of new triazole derivatives as inhibitors for mild steel corrosion in
the QSAR model and molecular dynamics simulation of benzimidazole derivatives
acid media, Electrochim. Acta 52 (2007) 6386–6394.
as corrosion inhibitors, Corros. Sci. 99 (2015) 76–88.
[34] D. Gopi, K.M. Govindaraju, L. Kavitha, Investigation of triazole derived Schiff’s
[5] G. Sığırcık, T. Tüken, M. Erbil, Assessment of the inhibition efficiency of 3,4-dia-
bases as corrosion inhibitors for mild steel in hydrochloric acid medium, J. Appl.
minobenzonitrile against the corrosion of steel, Corros. Sci. 102 (2016) 437–445.
Electrochem. 40 (2010) 1349–1356.
[6] N. Kıcır, G. Tansuğ, M. Erbil, T. Tüken, Investigation of ammonium (2,4-di-
[35] Q. Deng, N.-N. Ding, X.-L. Wei, L. Cai, X.-P. He, Y.-T. Long, G.-R. Chen, K. Chen,
methylphenyl)-dithiocarbamate as a new, effective corrosion inhibitor for mild
Identification of diverse 1,2,3-triazole-connected benzyl glycoside-serine/threonine
steel, Corros. Sci. 105 (2016) 88–99.
conjugates as potent corrosion inhibitors for mild steel in HCl, Corros. Sci. 64
[7] H.M.A. El-Lateef, Experimental and computational investigation on the corrosion
(2012) 64–73.
inhibition characteristics of mild steel by some novel synthesized imines in hy-
[36] A. Espinoza-Vázquez, G.E. Negrón-Silva, R. González-Olvera, D. Angeles-Beltrán,
drochloric acid solutions, Corros. Sci. 92 (2015) 104–117.
H. Herrera-Hernández, M. Romero-Romo, M. Palomar-Pardavé, Mild steel corrosion
[8] D. Daoud, T. Douadi, H. Hamani, S. Chafaa, M. Al-Noaimi, Corrosion inhibition of
inhibition in HCl by di-alkyl and di-1,2,3-triazole derivatives of uracil and thymine,
mild steel by two new S-heterocyclic compounds in 1 M HCl: experimental and
Mater Chem. Phys. 145 (2014) 407–417.
computational study, Corros. Sci. 94 (2015) 21–37.
[37] T. Zhang, S. Cao, H. Quan, Z. Huang, S. Xu, Synthesis and corrosion inhibition
[9] P. Mourya, P. Singh, A.K. Tewari, R.B. Rastogi, M.M. Singh, Relationship between
performance of alkyl triazole derivatives, Res. Chem. Intermed. 41 (2015)
structure and inhibition behavior of quinolinium salts for mild steel corrosion:
2709–2724.
experimental and theoretical approach, Corros. Sci. 95 (2015) 71–87.
[38] J.M. Aizpurua, I. Azcune, R.M. Fratila, E. Balentova, M. Sagartzazu-Aizpurua,
[10] M. Farsak, H. Keleş, M. Keleş, A new corrosion inhibitor for protection of low
J.I. Miranda, Click synthesis of nonsymmetrical bis(1,2,3-triazoles), Org. Lett. 12
carbon steel in HCl solution, Corros. Sci. 98 (2015) 223–232.
(2010) 1584–1587.
[11] E. Gutiérrez, J.A. Rodríguez, J. Cruz-Borbolla, J.G. Alvarado-Rodríguez,
[39] M. Piccinno, G. Aragay, F.Y. Mihan, P. Ballester, A.D. Cort, Unexpected emission
P. Thangarasu, Development of a predictive model for corrosion inhibition of
properties of a 1,8-naphthalimide unit covalently appended to a Zn–salophen, Eur
carbon steel by imidazole and benzimidazole derivatives, Corros. Sci. 108 (2016)
J. Inorg. Chem. 16 (2015) 2664–2670.
23–35.
[40] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made
[12] K.F. Khaled, Molecular simulation, quantum chemical calculations and electro-
simple, Phys. Rev. B 77 (1996) 3865–3868.
chemical studies for inhibition of mild steel by triazoles, Electrochim. Acta 53
[41] B. Delley, Ground-state enthalpies: evaluation of electronic structure approaches
(2008) 4384–3492.
with emphasis on the density functional method, J. Phys. Chem. A 110 (2006)
[13] B. Mernari, H.E. Attari, M. Traisnel, F. Bentiss, M. Lagrenée, 3,5-Bis (n-pyridyl)-4-
13632–13639.
amino-1,2,4-triazoles on the corrosion for mild steel in 1 M HCl medium, Corrosi.
[42] D. Zhang, Y. Tang, S. Qi, D. Dong, H. Cang, G. Lu, The inhibition performance of
Sci. 40 (1998) 391–399.
long-chain alkyl-substituted benzimidazole derivatives for corrosion of mild steel in
[14] F. Bentiss, M. Lagrenée, M. Traisnel, J.C. Hornez, The corrosion inhibition of mild
HCl, Corros. Sci. 102 (2016) 517–522.
steel in acidic media by a new triazole derivative, Corrosi. Sci. 41 (1999) 789–803.
[43] K. Jüttner, Electrochemical impedance spectroscopy (EIS) of corrosion process on
[15] M. Lagrenée, B. Mernari, M. Bouanis, M. Traisnel, F. Bentiss, Study of the me-
inhomogeneous surfaces, Electrochim. Acta 35 (1990) 1501–1508.
chanism and inhibiting efficiency of 3,5-bis(4-methylthiophenyl)-4H-1,2,4-triazole
[44] T. Pajkossy, Impedance of roughness capacitive electrodes, J. Electroanal. Chem.
on mild steel corrosion in acidic media, Corrosi. Sci. 44 (2002) 573–588.
364 (1994) 111–125.
[16] F. Bentiss, C. Jama, B. Mernari, H.E. Attari, L.E. Kadi, M. Lebrini, M. Traisnel,
[45] A. Popova, S. Raicheva, E. Sokolova, M. Christov, Frequency dispersion of the in-
M. Lagrenée, Corrosion control of mild steel using 3,5-bis (4-methyloxyphenyl)-4-
terfacial impedance at mild steel corrosion in acid media in the presence of ben-
amino-1,2,4-triazoles in normal hydrochloric acid medium, Corrosi. Sci. 51 (2009)
zimidazole derivatives, Langmuir 12 (1996) 2083–2089.
1628–1635.
[46] R. Solmaz, Investigation of the inhibition effect of 5-((E)-4-phenylbuta-1,3- dieny-
[17] M. Tourabi, K. Nohair, M. Traisnel, C. Jama, F. Bentiss, Electrochemical and XPS
lideneamino)-1,3,4-thiadiazole-2-thiol Schiff base on mild steel corrosion in hy-
studies of the corrosion inhibition of carbon steel in hydrochloric acid pickling
drochloric acid, Corros. Sci. 52 (2010) 3321–3330.
solutions by 3, 5-bis (2-thienylmethyl)-4-amino-1, 2, 4-triazole, Corrosi. Sci. 75

100
Q. Ma et al. Corrosion Science 129 (2017) 91–101

[47] Y. Tang, F. Zhang, S. Hu, Z. Cao, Z. Wu, W. Jing, Novel benzimidazole derivatives as Corros. Sci. 99 (2015) 1–30.
corrosion inhibitors of mild steel in the acidic media. Part I: Gravimetric, electro- [52] Z. Cao, Y. Tang, H. Cang, J. Xu, G. Lu, W. Jing, Novel benzimidazole derivatives as
chemical, SEM and XPS studies, Corros. Sci. 74 (2013) 271–282. corrosion inhibitors of mild steel in the acidic media. Part II: theoretical studies,
[48] Y. Tang, W. Yang, X. Yin, Y. Liu, R. Wan, J. Wang, Phenyl-substituted amino Corros. Sci. 83 (2014) 292–298.
thiadiazoles as corrosion inhibitors for copper in 0.5 M H2SO4, Mater. Chem. Phys. [53] A. Kokalj, On the HSAB based estimate of charge transfer between adsorbates and
116 (2009) 479–483. metal surfaces, Chem. Phys. 393 (2012) 1–12.
[49] X. Wang, H. Yang, F. Wang, An investigation of benzimidazole derivative as cor- [54] W. Yang, R.G. Parr, Hardness softness, and the fukui function in the electronic
rosion inhibitor for mild steel in different concentration HCl solutions, Corros. Sci. theory of metals and catalysis, Proc. Natl. Acad. Sci. U. S. A. 82 (1985) 6723–6726.
53 (2011) 113–121. [55] A. Kokalj, Is the analysis of molecular electronic structure of corrosion inhibitors
[50] N. Yilmaz, A. Fitoz, Ü. Ergun, K.C. Emregül A combined electrochemical and the- sufficient to predict the trend of their inhibition performance, Electrochim. Acta 56
oretical study into the effect of 2-((thiazole-2-ylimino)methyl)phenol as a corrosion (2010) 745–755.
inhibitor for mild steel in a highly acidic environment, Corros. Sci. 111 (2016) [56] N. Kovačević, A. Kokalj, Analysis of molecular electronic structure of imidazole-
110–120. and benzimidazole-based inhibitors: a simple recipe for qualitative estimation of
[51] I.B. Obot, D.D. Macdonald, Z.M. Gasem, Density functional theory (DFT) as a chemical hardness, Corros. Sci. 53 (2011) 909–921.
powerful tool for designing new organic corrosion inhibitors. Part 1: An overview,

101

You might also like