Luo2017 XPS Trés Importent

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 42

Accepted Manuscript

Title: Corrosion inhibition of mild steel in simulated seawater


solution by a green eco-friendly mixture of glucomannan (GL)
and bisquaternary ammonium salt (BQAS)

Authors: Xiaohu Luo, Xinyu Pan, Song Yuan, Shuo Du,


Caixia Zhang, Yali Liu

PII: S0010-938X(16)30925-8
DOI: http://dx.doi.org/doi:10.1016/j.corsci.2017.06.013
Reference: CS 7117

To appear in:

Received date: 11-10-2016


Revised date: 11-6-2017
Accepted date: 12-6-2017

Please cite this article as: Xiaohu Luo, Xinyu Pan, Song Yuan, Shuo Du, Caixia Zhang,
Yali Liu, Corrosion inhibition of mild steel in simulated seawater solution by a green
eco-friendly mixture of glucomannan (GL) and bisquaternary ammonium salt (BQAS),
Corrosion Sciencehttp://dx.doi.org/10.1016/j.corsci.2017.06.013

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Corrosion inhibition of mild steel in simulated seawater
solution by a green eco-friendly mixture of glucomannan
(GL) and bisquaternary ammonium salt (BQAS)

Xiaohu Luo a,b, Xinyu Pan a, Song Yuan a, Shuo Du a,Caixia Zhang a, and Yali Liua,*

a
State Key Laboratory for Chemo/Biosensing and Chemometrics, College of
Chemistry and Chemical Engineering, Hunan University, Changsha, 410082, P. R.
China
b
School of Chemistry and Chemical engineering, Qiannan Normal University for
Nationalities, Duyun, 558000, P. R. China

Graphical abstract:

Schematic illustration of the adsorption formation of GL molecules and BQAS

2017/6/12 -1-
molecules on the mild steel surface in the simulated seawater.

Highlights:
 The mixture of the GL and BQAS can act as a green eco-friendly inhibitor.
 Corrosion inhibition performance of the mixture was determined.
 Corrosion inhibition performance of the mixture was much better than that of
monocomponent.
 Adsorption kinetic of the mixture was discussed.

Abstract: The corrosion inhibition effect of a green eco-friendly mixture comprised


of glucomannan (GL) and bisquaternary ammonium salt (BQAS) on the corrosion
protection of mild steel in simulated seawater is investigated by electrochemical
measurements and surface characterization. It is found that the mixture of GL and
BQAS is a mixed-type inhibitor and can provide effective corrosion inhibition for
mild steel. The corrosion inhibition is achieved by the chemical as well as physical
adsorption of the mixture on the surface of the steel. It is also found that the
adsorption fairly obeys Langmuir isotherm.

Keywords: corrosion; mild steel; seawater; green inhibitor

1. Introduction
With fast development of manufacture and heavy industry such as chemical
production, iron, print circuit board and power generation, fresh water has been
widely used to remove the redundant heat from heat transfer in cooling systems.
Moreover, explosive population has also intensified the consumption of fresh water,
which has caused the shortage of fresh water source in recent years [1]. The utilization
of seawater can be seen as effective method to ease the pressure on the water for
industrial application in the cooling systems. However, the natural seawater contains

2017/6/12 -2-
some corrosive mediums such as chloride ion, sulfate ions, carbonate ions, hydrogen
carbonate ions, organic acids and microorganism, thus, may causing a range of
problems, such as corrosion [2-6]. As is well known, corrosion not only results in a
significant decrease of heat exchange efficiency in the cooling system, but also leads
to the fast deterioration of equipment materials which causes serious economic loss
and safety issues [7-10].
Corrosion inhibitors such as organic inhibitors have been considered as the most
common method to control or inhibit the corrosion of mild steel in the seawater
environments [11,12]. Commonly, corrosion protection by the inhibitors is mostly
attributed to the modification of the mild steel surface by the adsorption of inhibitor
molecules and the subsequent formation of a protective layer [13-15].
Green, eco-friendly natural polymeric substances extracted from seeds or leaves as
the corrosion inhibitors have attracted considerable attention and interest due to its
low cost, eco-friendly nature and practical use. Additionally, to be compliant with the
environmental laws, many researchers has diverted to use green eco-friendly natural
products, such as fruit peel aqueous extracts [15], lignin terpolymer [16], and cassava
strarches [17], as the corrosion inhibitors to reduce the corrosion of various steels and
non-ferrous metals. Results reveal that these natural products exhibit superior
corrosion protection performance to mild steel, copper and aluminum. On the other
hand, synthetic polymers have also been used as corrosion inhibitors for various
metals [18,19]. For example, the polyacrylic acid, as a synthetic polymer corrosion
inhibitor, can bind active metal surface and act as adsorption center due to the oxygen
heteroatom in the presence of the carboxyl group [20]. Moreover, other compounds
have been found to exhibit good corrosion protection to metals, including Schiff bases
[21], substituted thioureas [22], and saccharides [23].
Glucomannan (GL) (Fig. 1a) as a green natural product with the jelly property is
contained in the cell walls of amorphophallus konjac, which is very abundant in China
and contains poly(D-galacturonic acid). GL, as a green food additive, is extensively
used in the frozen food industry due to its green and eco-friendly property. The unique
properties of poly(D-galacturonic acid) includes its polyfunctionality, biodegradable
2017/6/12 -3-
nature, nontoxicity, flexible structural network. Poly(D-galacturonic acid) derivatives
have been assessed as a green inhibitor for X60 pipeline steel in acidic media [24].
Symmetric bisquaternary ammonium salt (BQAS) as a synthetic polymer are
composed of two hydrophilic groups and two hydrophobic groups per molecule and
separated by a spacer (Fig. 1b), which is also widely used as a surfactant in industry
due to its low cost and nontoxicity. Recently, the application of dissymmetric
quaternary ammonium salt has been investigated by some researchers, indicating that
quaternary ammonium salt has good corrosion inhibition performance for metals in
the various corrosive media [25,26]. However, the application of symmetric BQAS as
corrosion inhibitor is not reported.
Synergism is an effective method to decrease the dosage of inhibitors and to
diversify the application of the inhibitors in corrosive media. Previous studies report
that the addition of multiple inhibitors to the seawater media could significantly
increase the inhibition efficiency and decrease the dosage of inhibitors compared with
monocomponent [27-29]. The objective of this study is to investigate the inhibitive
performance of a green eco-friendly mixture of GL and BQAS for the mild steel in
simulated seawater solution by using the open circuit potential (OCP), weight loss,
potentiodynamic polarization and electrochemical impedance spectroscopy (EIS)
measurement with the corrosion kinetic parameters being determined. Surface
morphology of the tested samples is characterized by scanning electron microscopy
(SEM) and atomic force microscopy (AFM). Additionally, the adsorption mechanism
of GL and BQAS at the mild steel surface is also investiagted. The effects of the
temperature, concentration and immersion time on the inhibitive performance are
further studied.

2. Experimental
2.1 Materials
The GL and BQAS were of reagent grade and were purchased from Feixiang Group
(Jiangsu, China), they were used without further purification. The chemical structures
of the two compounds are shown in Fig. 1.

2017/6/12 -4-
The composition (wt.%) of mild steel used is summarized in Table 1. As for weight
loss measurement, the mild steel was cut into 3 cm × 2 cm × 0.1 cm size, whereas the
mild steel of size 3 cm × 2 cm × 0.1 cm with 1.5 cm2 exposed surface areas was used
as working electrode for open circuit potential, potentiodynamic polarization and
electrochemical impedance spectroscopy (EIS) measurements. Prior to the
measurements, these samples should be mechanically polished with 150, 400, 800,
1200 and 1600 grit grinding paper, and then cleaned with ethanol, washed with
distilled water and dried in vaccum at 30℃ [10,15].
2.2 Electrolytic solution
All experiments were carried out in simulated seawater solution in the absence and
presence of different concentrations of GL and/or BQAS as corrosion inhibitors. The
composition (wt.%) of the simulated seawater is shown in Table 2. All simulated
solutions were prepared from analytical grade reagents and distilled water. All
analytical grade reagents purchased from Feixiang Group (Jiangsu, China) were used
without further purification. The temperature of the simulated solution was controlled
by using an aqueous thermostated bath.
2.3 Weight loss measurements
Weight loss measurements were carried out according to the ASTM standard G
31-72 [30]. The mild steel specimens were immersed in simulated solution with
various concentrations of GL and/or BQAS. The specimens were weighted before and
after immersion in the simulated solution for 4 h. After immersion, the specimens
were rinsed with distilled water and ethanol, and then were dried in vaccum at 30℃
[15,22]. The weight loss was measured using the analytical balance, with the mean
weight loss and corresponding standard deviation being recorded.
2.4 Potentiodynamic polarization measurements
All electrochemical measurements were carried out in a conventional
three-electrode cell, where a platinum foil was the counter electrode, a saturated
calomel electrode (SCE) acted as the reference electrode and the mild steel with 1.5
cm2 exposed surface area was used as the working electrode. Before each
electrochemical measurement, the sample should be immersed in simulated solution
2017/6/12 -5-
at open circuit potential (OCP) for 1 h in order to attain a stable state.
In the case of potentdiodynamic polarization study, polarization curves were
obtained by scanning the potential range between -0.25 V and +0.25 V with respect to
the OCP at a scanning rate of 0.5 mV s-1. The corrosion potential (Ecorr) and corrosion
current density (icorr) were acquired by extrapolated the linear Tafel segments of the
cathodic curves and anodic Tafel lines to the point of intersection [31,32]. The values
of anodic Tafel slope (βa) and cathodic Tafel slope (βc) were determined from the
theoretically calculated straight line and experimental line, respectively.
2.5 Electrochemical impedance spectroscopy
Electrochemical impedance spectroscopy (EIS) measurement was carried out by
using CHI600 instrument (made in China) with an AC amplitude of 5 mV for the
frequency range from 100 kHz to 0.01 Hz at the OCP. All the impedance data were
analyzed and pertinent EIS parameters were calculated by Zsimpwin software. To
ensure the reliability and repeatability for the data, five measurements were carried
out for each condition.
2.6 X-ray photoelectron spectroscopy
X-ray photoelectron spectroscopy (XPS) was carried out using a PHI 5700 ESCA
spectrometer with a monochromatic Al Kα X-ray source (hυ=1486.7 eV). Binding
energies were calibrated based on the adventitious C 1s peak at 285.0 eV. The
pressure was at the 10-10 Pa. The conditions used for all of the survey scans were as
follows: pass energy of 150 eV, energy range of 1100 eV, step size of 0.5 eV, X-ray
spot size of 700 × 400 µm and sweep time of 240 s. For the high-resolution spectra an
energy range up to 40 eV was used, depending on the peak being examined, with a
pass energy of 10 eV and a step size of 0.05 eV. The XPS data were achieved from the
surface directly without Ar ion sputtering.
2.7 Surface characterization
2.7.1 Scanning Electron Microscopy
The surface of the mild steel after exposure to simulated solution with and without
the corrosion inhibitors for different immersion times was characterized by scanning
electron microscopy (SEM) using Nova nanosem 430 with 20 kV acceleration
2017/6/12 -6-
voltage.
2.7.2 Atomic Force Microscopy
The surface topographic features of mild steel immersed in the simulated solution
in the absence and presence of GL and/or BQAS was further characterized by atomic
force microscopy (AFM). The measurements were carried out by contact mode AFM
(Dimension Edge, Bruker). Micro-fabricated Si3N4 cantilever as the support with
20-50 nm radius of curvature was used. After various immersion time in simulated
solution with and without GL and BQAS, the mild steel specimens were washed with
distilled water and ethanol, dried in vacuum drying oven at 40℃ before the AFM.

3. Result and discussion


3.1 Open circuit potential measurement
The variations in OCP with time of mild steel in simulated seawater in the absence
and presence of inhibitors at 25 ℃, pH = 8.0 are presented in Fig. 2. The OCP value
of mild steel without the inhibitors decreases with time initially, then stabilizes at
around -0.554 V/SCE after 40 min of immersion. The initial decrease may be
attributed to the pitting corrosion occurred at the surface of the mild steel. The OCP
values for the mild steel in the presence of BQAS and GL are shifted to noble
direction, and are at around -0.514 V/SCE and -0.502 V/SCE, respectively, indicating
that the mild steel is protected by the BQAS or GL. Moreover, the OCP value in
simulated seawater containing the mixture of BQAS and GL stabilizes at about -0.471
V/SCE, a more noble potential value compared with that in the presence of BQAS or
GL alone, indicating the synergistic effect of BQAS and GL.
3.2 Weight loss measurement
Corrosion protection performance of the inhibitor was determined by weight loss
measurement. In the all weight loss measurements, five specimens were measured and
the average values are reported. The relationship between the weight loss
measurements and different inhibitor concentrations in simulated seawater at 25 ℃,
pH = 8.0 for 4 h of immersion time is shown in Fig. 3. Corrosion rate (CR) as
expressed in mg cm-2 h-1 is determined according to the following equation [33,34]:

2017/6/12 -7-
Δm
CR 
s  Δt
where Δm, s and Δt are average weight loss (mg), exposed coupon area (cm-2) and the
immersion time (h) in the simulated seawater, respectively. Based on the results of the
corrosion rate, the inhibition efficiency (ηw%) is calculated through the following
equation:

C0  Cin
ηw %  100%
C0

where C0 and Cin are the corrosion rate for mild steel in simulated seawater at 25 ℃,
pH = 8.0 for 4 h of immersion time without and with inhibitors, respectively. From the
Fig. 3a, the inhibition efficiency of mild steel increases gradually with the
concentration initially, then keeps stable. The BQAS and GL exhibit the maximum
inhibition efficiency with approximately 77.5% and 85.4% at the concentration of
around 80 mg L-1, respectively, whereas the maximum inhibition efficiency for the
mixture reaches about 98.8% at the concentration of 80 mg L-1. As is shown in Fig. 3b,
the corrosion rate for mild steel decreases with the increase of the concentration, and
the minimum corrosion rate values for the BQAS, GL and mixture at the
concentration of 80 mg L-1 are around 0.694, 0.402 and 0.085 mg cm-2 h-1,
respectively. These results strongly demonstrate that the corrosion inhibition
performance for the mixture is much better than that of monocomponent, which can
be interpreted by the synergistic effect of the GL and BQAS. This successful
corrosion inhibition action for the mixture is obtained by the formation of protective
layer at the mild steel surface.
3.3 Potentiodynamic polarization
The potentiodynamic polarization curves of the mild steel in simulated seawater
with and without inhibitors are shown in Fig. 4. The potentiodynamic polarization
was carried out to determine the effect of inhibitors on the anodic dissolution of the
mild steel and cathodic reduction. Polarization parameters including the values of
anodic Tafel slope (βa), cathodic Tafel slope (βc), corrosion potential (Ecorr) and
corrosion current density (icorr) are listed in Table 3. From Fig. 4 and Table 3, it is

2017/6/12 -8-
found that compared to the blank solution, the value of icorr decreases significantly and
the value of Ecorr shifts notably towards positive direction by addition of GL and
BQAS to the simulated seawater, which implies that both the cathodic and anodic
reactions are suppressed by the inhibitors. In other words, GL and BQAS could
reduce anodic dissolution of mild steel and suppress the cathodic reaction. Moreover,
the value of icorr for the mixture of GL and BQAS is around 0.164 μA cm-2, much
lower than that for GL (i.e. 2.21 μA cm-2) or BQAS alone (i.e. 3.57 μA cm-2). The Ecorr
value for the mixture is approximately -0.472 V/SCE, more noble than that for GL (i.e.
-0.503 V/SCE) or BQAS alone (i.e. -0.513 V/SCE). The inhibition efficiency (ηi%) of
the mixture is up to 99.1%, much higher than that of GL (i.e. 87.4%) and BQAS (i.e.
79.7%). These results suggest that the inhibitive performance of the mixture is better
than that of monocomponent. The inhibition efficiency (ηi%) is calculated according
to the following equation:

i corr 0   i corr inh


ηi %  100%
i corr 0 

where icorr(0) and icorr(inh) are the corrosion current densities without and with the
addition of inhibition. It is suggested that if the displacement in Ecorr is > 85 mV with
respect to the Ecorr of blank solution, suggesting the chemical inhibitor can be seen as
an anodic or cathodic type inhibitor, and if the displacement in Ecorr is < 85 mV, the
inhibitor can be seen as a mixed type [35,36]. In the present study the largest
displacement in Ecorr obtained at the concentration of 40 mg L-1 GL and 40 mg L-1
BQAS is about 81.2 mV, which is less than 85 mV, indicating that the mixed
inhibitors act as a mixed type inhibitor. Additionally, the prominent changes in the
anodic and cathodic Tafel slopes in the simulated seawater with the mixture of
inhibitors may be explained by that the GL and BQAS molecules adsorbed on the
mild steel surface by inter- and intra- molecular forces, forming a protective layer,
which affects both the cathodic and anodic reaction mechanism and can act as a
physical barrier to prevent diffusion of the corrosive species to the mild steel surface.
3.4 Eletrochemical impedance spectroscopy
To investigate the corrosion inhibition process in terms of capacitive as well as
2017/6/12 -9-
resistive behavior at the mild steel/solution interface, elctrochemical impedance
spectroscopy (EIS) studies were carried out. Fig. 5 illustrates Nyquist plots from AC
impedance measurement of the mild steel in the simulated seawater in the absence
and presence of inhibitors at 25 ℃. The diameter for Nyquist plots significantly
increases with the addition of inhibitors to the simulated seawater, indicating the
strengthening of protective layers. The impedance diagrams exhibit a depressed
semicircle in the center under the real axis, which may be due to the roughness of
metal surface as well as inhomogeneties of the mild steel raised from the corrosion
process and the frequency dispersion [37]. Also, the semicircular shape obtained in
the blank solution and in the presence of inhibitors is different, indicating that the
corrosion mechanism is changed. The Nyquist plot of the mild steel in the blank
seawater is consisted of two capacitive loops. The high frequency capacitive loop is
attributed to the formation of an oxide layer or to oxide layer itself [37,38], and the
low frequency capacitive loop is attributed to the bulk relaxation of species in the
oxide layer, or to surface roughness [38,39]. The corresponding structural model of
the interface is illustrated in Fig. 6a. R1 presents the polarisation resistance, L refers to
the inductance, and R1 + R2 presents the charge-trasfer resistance (Rct).
A single capacitive loop is observed for the mild steel in simulated seawater with
inhibitors, indicating only one process occuring. The corresponding equivalent circuit
is displayed in Fig. 6b, which is consisted of the solution resistance (Rs) as well as
charge transfer resistance (Rct) and the constant phase element of double layer (CPE).
In general, the double layer capacitance (Cdl) value is influenced by the roughness of
metal surface, which could be simulated using the constant phase element (CPE) [40].
CPE is deduced according to the following equation [41]:

1
ZCPE 
Y0 αω 
n

where Y0 represents a proportionality factor and the parameter ‘n’ shows phase shit,
which indicates different physical phenomenon, including surface roughness, inhibitor
adsorption and/or desorption and porous film formation, and the values of ‘n’ lie
between 0 and 1 [42,43]. According to the above equivalent circuit, the
2017/6/12 - 10 -
electrochemical parameters including charge transfer resistance (Rct), the double layer
capacitance (Cdl), n, goodness of fit (λ2) are calculated using the Zsimpwin software,
and listed in Table 4. the Cdl was simulated by using the constant phase element, and
the values of Cdl were obtained by determining a frequency at which the imaginary
component of the impedance is a maximum fmax from the following equation:

Cdl  Y0 2πf max 


n 1

the obtained impedance parameters were used to calculate inhibition efficiency (ηe%)
according to the following equation:
R ct(inh)  R ct(0)
ηe %  100%
R ct(inh)

where Rct(inh) and Rct(0) are values of charge transfer resistance with and without
inhibitors, respectively. The calculated λ2 value ranges from 0.00117 to 0.00691, as
shown in Table 4, indicating a good fitting to the equivalent circuit. Theoretically,
lower value of λ2 suggests that the fitted data is in agreement with the experimental
data. It is observed from Table 4 that the Rct values significantly increase while Cdl
values prominently decrease with the addition of inhibitors compared with that in
blank solution. An increase in Rct values is attributable to the formation of protective
layer at the mild steel/solution interface. A reasonable explanation for the decrease in
Cdl values is that the local dielectric constant decreases and/or the thickness of double
layer increases [44]. This behavior indicates that GL and BQAS molecules displace
water molecules and other ions which are originally adsorbed at the metal/solution
interface. Consequently, the protective layer adsorbed at the mild steel surface slows
down the dissolution process of the mild steel. Indeed, the largest effect is obtained
with addition of the mixture which gives a Rct value equal to 39.47 kΩ cm2 and Cdl
value of only 105 μF cm-2. Table 4 also clearly shows that the increase of ‘n’ value
suggests the decrease of surface inhomogeneity, due to the GL and/or BQAS
molecules adsorbed on the active adsorption sites at the mild steel surface.
The inhibition efficiency increases with the addition of the mixed inhibitors, up to
97.9%, which is much larger than that of the monocomponent. This further confirms

2017/6/12 - 11 -
that the inhibitive performance of the mixture for the mild steel in simulated seawater
is much better than that of the monocomponent. Additionally, The inhibition
efficiency from weight loss measurements (ηw%), potentiodynamic polarization
experiments (ηi%), and eletrochemical impedance spectroscopic studies (ηe%) are
shown in Table 5. They are in agreement, suggesting that the three independent
techniques confirm the validity of the results.
3.5 Influence of testing conditions on the inhibition performance
3.5.1 The effect of temperature
The effect of seawater temperature on the corrosion inhibition efficiency for the
mixture (40 mg L-1 GL and 40 mg L-1 BQAS) had been studied. Since seawater
temperature can reach 343 K, temperatures in the range from 298 K to 350 K were
selected for the simulated seawater. Potentiodynamic polarization measurements were
conducted to determine the corrosion inhibition efficiency on the mild steel in the
simulated seawater with and without inhibitors, the results are depicted in Figs. 7 and
8. The polarization parameters, including βa, βc, Ecorr and icorr, are summarized in
Table 6.
As for blank solution, Fig. 7 and Table 6 show that the Ecorr shifts toward more
negative direction and icorr increases with the increase of temperature. This behavior is
ascribed to the increase of the anodic dissolution rate of mild steel with the
temperature. However, in the presence of inhibitors, icorr decreases and Ecorr shifts
significantly in the positive direction with respect to that in the absence of inhibitors
(Fig. 7 and Table 6), indicating effective inhibition at all the selected temperatures.
Fig. 9 shows the corrosion inhibition efficiency determined from the polarization
curves at the selected temperatures. The corrosion inhibition efficiency decreases with
the increase of temperature. The explanation for this phenomenon is that the inhibitors
adsorption/desorption process is temperature dependent. For instance, the adsorption
equilibrium could shift towards the desorption with the increase of temperature,
resulting in a lower coverage and a decrease in formulation molecules surface
conformation order [36,40,45,46]. Nevertheless, corrosion efficiency is sufficiently
high even at 350 K, at around 93.1%.
2017/6/12 - 12 -
To further understand the kinetics of corrosion inhibition process and the
mechanism of the inhibition action, the apparent activation energy values (Ea) was
calculated by the following Arrhenius equation:

log i corr   
Ea
α
2.303R 0T0

where T0, R0 and α are the absolute temperature, gas constant and the Arrhenius
pre-exponetial factor, respectively. Fig. 10 describes the Arrhenius plots calculated
from the Arrhenius equation. The Ea values for the mild steel in the absence and the
presence of inhibitors are about 40.11 kJ mol-1 and 66.34 kJ mol-1, respectively, and
the value of Ea in the presence of inhibitors is greater than that in the absence of
inhibitors, indicating that the corrosion mechanism has been changed by the addition
of inhibitors to the simulated seawater. This is due to that the GL and BQAS
molecules have provided effective inhibition for the mild steel to suppress the charge
and mass transfer process.
3.5.2 The effect of inhibitor concentration
Bode and phase angle plots of the mild steel in the simulated seawater with various
concentrations of inhibitors are illustrated in Fig. 11. The electrochemical impedance
parameters calculated with the Zsimpwin software are listed in Table 7. It is evident
from the Bode curves that the initial addition of inhibitors to seawater leads to an
increase in the interfacial impedance which remains stable with further increase in the
concentrations of inhibitors, indicating decrease in the dissolution of mild steel.
Additionally, the phase angle curves exhibit the single peaks, further indicating that a
single time constant forms in the corrosion process at the mild steel/seawater interface.
The increase in the peak height suggests a more capacitive response at the mild
steel/seawater interface due to the adsorption of more inhibitor molecules on the mild
steel surface. This is in agreement with the changes of inhibition efficiency and
corrosion rate (Fig. 12). The inhibition efficiency values of mild steel increases with
concentrations initially, then stabilizes at about 98.0%.
The effectiveness of the inhibitors should be attributable to the adsorption of GL
and BQAS at the steel mild/solution interface. GL presents in the molecular form and
2017/6/12 - 13 -
BQAS presents in protonated form in the simulated seawater. They can be adsorbed
through the interaction between either the lone pair of electrons on oxygen atoms,
and/or nitrogen atoms or conjugated double bond or both presented in GL and BQAS,
and the vacant d-orbital of iron in the mild steel [47,48]. The inhibition efficiency
increases with concentration of GL and BQAS, due to the availability of larger
number of GL and BQAS in both forms for adsorption at the higher concentration.
3.5.3 The effect of immersion time
To investigate the mixture of GL and BQAS adsorption kinetics and to determine
the time needed for the mixture to reach maximum inhibition efficiency,
electrochemical impedance spectroscopic measurement was performed in the
simulated seawater with the mixture for different immersion time. The results are
described in Fig. 13. The diameter of the Nyquist plots markedly increases with the
increase in the immersion time initially, and then becomes relatively stable (Fig. 13),
indicating a strong inhibition of the dissolution process occurring at the mild steel
surface. The inhibition efficiencies are also shown in Fig. 14. The inhibition
efficiency significantly increases with the immersion time initially, then becomes
relatively stable and reaches 97.9% after an immersion time of 8 h. The equivalent
circuit presented in Fig. 6b is used to simulate the impedance spectra. The relevant
simulated electrochemical parameters of Rct, Cdl, and CPE are presented in Table 8.
As seen from Table 8, the charge transfer resistance Rct increases upon increasing
immersion time within 12 h at a given concentration. A considerable reason for this is
that the adsorption of GL and BQAS molecules at the mild steel surface with the
immersion time form an uniform and increased resistance to the charge transfer
process. Simultaneously, Cdl values decrease with immersion time. Water molecules
presented at the mild steel/solution interface exhibits a higher value of relative
dielectric constant. When water molecules presented at the metal surface are replaced
by GL and BQAS molecules, resulting in the decrease of Cdl value [49-51].
3.6 Adsorption isotherm
To understand the adsorption mechanism of GL and BQAS at the mild steel surface,
the adsorption isotherm was also studied. Among several adsorption isotherms
2017/6/12 - 14 -
including Langmuir, Temkin, Frumkin, Bockris-Swinkeis, and Flory-Huggin
isotherms, the Langmuir isotherm is considered to fit better with the experimental
data. According to the Langmuir adsorption isotherm, the fraction of surface coverage,
θ, is associated to the inhibitor concentration by the following equation:

Cinh 1
  Cinh
θ K ads

where Cinh is the corrosion inhibition concentration, Kads is the adsorptive equilibrium
constant. Generally, the θ value is equal to the inhibition efficiency (η/100) obtained
from the weight loss measurements. A straight line with a correlation coefficient (R2 >
0.99) as described in Fig. 15 is obtained from plotting Cinh/θ against Cinh. The R2 >
0.99 indicates that the adsorption of GL and BQAS fairly obeys the Langmuir
isotherm. In addition, the Kads can be determined by the intercept of straight line. The
adsorptive equilibrium constant Kads is related to the adsorptive standard free energy

(Δ G 0ads ), which can be calculated by the following equation:

ΔG0ads  RTln(1106 K ads )

where R and T are the standard gas constant and thermodynamic temperature,
respectively, 1 × 106 is the concentration of water molecules expressed in mg L-1. The

Kads and Δ G 0ads values are summarized in Table 9. Generally, if the value of Δ G 0ads is

less than -20 kJ mol-1, the adsorption is regarded as the physisorption, in such which,
the adsorption is due to the electrostatic interaction between the inhibitor molecules

and iron atom. However, if the absolute values of Δ G 0ads is 40 kJ mol-1 or higher, the

adsorption can be regarded as the chemisorption in which the covalent bond due to
transfer and/or share of electrons from inhibitors to metal surface [52,53]. Table 9

shows that the calculated values of Δ G 0ads is approximately -33.016 kJ mol-1,

suggesting that the adsorption of GL and BQAS at the mild steel surface involves
both chemical as well as physical interactions. The phenomenon may be explained by
the following reasons: (1) every unit of GL molecules contains partially negatively
charged oxygen atoms, and the negatively charged oxygen atoms can electrostatically

2017/6/12 - 15 -
bind positively charged ions such as Fe3+, Fe2+on the surface of mild steel; (2) the
more electropositive nature of iron, as a Lewis acid, can form coordinative bonds with
nitrogen and oxygen heteroatoms acting as Lewis base that scattered over the BQAS
and GL molecules respectively. Thus, these inter- and intra- molecular forces play an
important role in successfully forming protective layers at the mild steel/seawater
interface, which can effectively shield against aggressive electrolytes, such as Cl-,
SO42-, NO3- and H2O, as illustrated in Fig. 16.
3.7 Mild steel surface analysis
3.7.1 X-ray photoelectron spectroscopy
In order to further understand the chemical nature of mild steel/inhibitors interface
and to demonstrate the adsorption mechanism of GL and BQAS molecules, X-ray
photoelectron spectroscopy (XPS) was carried out. For comparison purpose, the
high-resolution XPS spectra of pure BQAS (C 1s and N 1s), pure GL (C 1s and O 1s)
and the mild steel surface after 4 h of immersion in the simulated seawater with
inhibitors were obtained, as shown in Figs. 17-19.
The C 1s spectrum of pure BQAS (Fig. 17a) exhibits two main peaks. The first
peak appearing at about 286.7 eV due to C-N [54]; the second peak centered at around
285.0 eV which is attributed to -C-C- or C-H [55]. The N 1s spectrum (Fig. 17b)
shows one peak located at approx. 400.4 eV, associated with (-C-N+).
The deconvolution of C 1s spectrum (Fig. 18a) for pure GL gives two peaks at
around 285.0 eV (-C-C- or C-H) and 289.1 eV (-C-O). The O 1s spectrum of pure GL
shows two main peaks, which may be attributable to C-O-C (531.8 eV) and -OH
(533.7 eV) [56].
The deconvolution of the Fe 2p3/2 curves consists of four main peaks (Fig. 19a).
The first peak appeared at low binding energy (706.4 eV) with the high intensity is
attributed to iron nitrides (FexN), this is confirmed by the deconvolution of N 1s (Fig
19d). The second peak locates at 710.7 eV, assigned to Fe3+, which is attributed to
Fe2O3 and FeOOH [57]. The third peak appeared at 715.0 eV is associated with the
statellites of the Fe (II) [58]. The last peak at (about 719.8 eV) is attributable to the
satellites of the Ferric compound [58].
2017/6/12 - 16 -
As for O 1s spectra of mild steel surface after immersion in the simulated seawater
with inhibitors (Fig. 19b), there are two main peaks. The peaks located at 530.1 eV is
attributable to Fe-OH and Fe-O in the iron hydroxides and oxides, respectively [57],
which is in good agreement with the presence of the iron oxide/hydroxide species
revealed in the Fe 2p3/2 spectra. The peak located at 531.8 eV may also has
contribution from oxygen bonded to carbon, which is present in GL molecule. This
behavior demonstrated that GL molecules are chemically and/or physically adsorbed
at the mild steel surface. The obtained-results are consistent to the C 1s spectra (Fig.
19c). The C 1s spectrum shows three obvious peaks. The first and most intense peak
appeared at about 285.0 eV should be attributed to the C-C and/or C-H bonds which
are present in both GL and BQAS molecules. The middle peak at around 286.7 eV is
associated with the presence of C-N groups presented in BQAS molecules. The last
peak at around 289.1 eV should be mainly attributed to the carbon atom of the C-O
which may be result from cabonyl groups of GL molecules.
The detection of N 1s spectrum for the mild steel treated by inhibitors (Fig. 19d)
provides direct evidence that BQAS molecules are adsorbed at the mild steel surface.
The fitting of the N 1s with the two components indicate two chemically distinct
nitrogen coordinations. The peak located at 400.4 eV is attributable to the N+-C bond,
which presents in BQAS molecule. The less intense component located at 398.7 eV is
assigned to N-Fe. This can be explained by that the coordination of the nitrogen atom
in the BQAS molecules with the iron atom of the mild steel surface. The appearance
of N-Fe bond suggests that the complex formed is attributed to the donor acceptor
interaction between N atoms of the BQAS and the vacant of d-orbitals of iron.
These XPS characterization further provide direct evidence of the adsorption of the
GL and BQAS molecules onto the mild steel surface and corroborate the
potentiodynamic polarization and electrochemical impedance studies.
3.7.2 Scanning electronic microscopy
SEM micrographs of the mild steel surface after 1 and 48 h of immersion in the
simulated seawater at 25 ℃, pH=8.0 without inhibitors are presented in Fig. 20. Fig.
20a shows that after 1 h of immersion in the simulated seawater without inhibitors,
2017/6/12 - 17 -
pitting corrosion appears on the mild steel surface. The mild steel surface is
drastically damaged by the corrosive medium with the extended immersion time and
formed corrosion products at the mild steel surface (Fig. 20b). On the contrary, after
48 h of immersion in the simulated seawater in the presence of the mixed inhibitors,
the corrosion of the mild steel is negligible (Fig. 20c-d), due to the formation of
protection layer on the steel surface, which prevents the corrosive medium penetration
to the mild steel surface.
2.7.3 Atomic force microscopy
The surface feature of the mild steel was further characterized by AFM before and
after immersion in the simulated seawater in the absence and presence of the inhibitor
mixture. The surface morphology of the as-polished mild steel and the steel after
immersion in the simulated seawater with and without the inhibitors are shown in Fig.
21. Fig. 21a shows the surface of the mild steel before immersion, revealing the
mechanical grinding marks. The average roughness of the steel surface is about 180
nm (Fig. 21a). As for the mild steel after immersion in the simulated seawater without
the inhibitors, the surface looks relatively uneven and rough (Fig. 21b), and the rough
increases to 1500 nm (Fig. 21b), indicating corrosion of the surface. However, it is
clearly evident in Fig. 21c that the surface of mild steel after immersion in the
seawater containing the mixture of GL and BQAS is relatively smooth compared to
that in Fig 21b. The roughness decreases to about 400 nm (Fig. 21c). The AFM results
further confirm that GL and BQAS molecules absorbed by inter- and intra- molecular
forces at the mild steel surface can effectively protect the mild steel from corrosion.

4. Conclusions
(1) The OCP, weight loss measurement, EIS and potentiodynamic polarization
studies show that the mixture of GL and BQAS can provide effective corrosion
inhibition for mild steel in the simulated seawater, with the inhibition efficiency
up to 98%.
(2) The inhibition performance of the mixture of GL and BQAS is significantly
better than that of the monoconponent due to the synergistic effect. The mxiture

2017/6/12 - 18 -
is a mixed-type inhibitor that can suppress both cathodic reaction and anodic
dissolution of the mild steel in the simulated seawater. The effective inhibition
is achieved by the adsorption of the inhibitors onto the steel surface, which
occurs both chemically as well as physically. The adsorption of the inhibitors
on the mild steel surface obeys Langmuir isotherm.
(3) The potentiodynamic polarization and EIS studies also suggest that the
inhibition efficiency of the mixture increases with the increase of the inhibitor
concentration and immersion time, whereas decreases with the temperature.

Acknowledgments
The research is financially supported by the Major Science and Technology
Projects of Hunan Province, China (2015GK1004), the Research Foundation of
Education Bureau of Guizhou Province, China (Guizhou [2015]402), and Joint
Foundation of Science and Technology Department of Guizhou Province, China
(Guizhou “LH” [2014]7430).

2017/6/12 - 19 -
References
[1] M. Ando, Y. Sayato, Studies on vinyl chloride migrating into drinking water from
polyvinyl chloride pipe and reaction between vinyl chloride and chlorine, Water. Res.
18 (1984) 315-318.
[2] M. Sareni, C. Dehghanian, M.M. Sabet, The effect of molybdate concentration and
hydrodynamic effect on mild steel corrosion inhibition in simulated cooling water,
Corros. Sci. 48 (2006) 1404-1412.
[3] J.A. Leenbeer, Comprehensive approach to preparative isolation and fraction of
dissolved organic carbon from natural water and wastewater, Environ. Sci. Technol.
15 (1981) 578-587.
[4] S. Ghareba, S. Omanovic, Interaction of 12-aminododecanoic acid with a carbon
steel surface: towards the development of ‘green’ corrosion inhibitor, Corros. Sci. 52
(2010) 2104-2113.
[5] R. Touir, N. Dkhireche, M.E. Touhami, M. Lakhrissi, B. Lakhrissi, M. Sfaira,
Corrosion and scale processes and their inhibition in simulated cooling water systems
by monosaccharides derivatives, part I:EIS study, Desalination. 249 (2009) 922-928.
[6] R. Touira, N. Dkhireche, M. Ebn. Touhami, M. Sfaira, O. Senhaji, J.J. Robin, B.
Boutevin, M. Cherkaoui, Study of phosphonate addition and hydrodynamic conditions
on ordinary steel corrosion inhibition in simulated cooling water, Mater. Chem. Phys.
122 (2010) 1-9.
[7] J.B. Sun, G.A. Zhang, W. Liu, M.X. Lu, The formation mechanism of corrosion
scale and electrochemical characteristic of low alloy steel in carbon dioxide-saturated
solution, Corros. Sci. 57 (2012) 131-138.
[8] A.M. Abdel-Gaber, B.A. Abd-EI-Nabey, E. Khamis, D.E. Abd-EI-Khalek, A
natural extract as scale and corrosion inhibitor for steel surface in brine solution,
Desalination. 278 (2009) 37-42.
[9] M.M. Al-Abdallah, A.K. Maayata, M.A. Al-Qudah, N.A.F. Al-Rawashdeh,
Corrosion behavior of copper in chloride media, Open. Corros. J. 2 (2009) 71-76.
[10] M.A. Abu-Dalo, N.A. Al-Rawashdeh, A. Ababneh, Evaluating the performance
2017/6/12 - 20 -
of sulfonated Kraft lignin agent as corrosion inhibitor for iron-based materials in
water distribution systems, Desalination. 313 (2013) 105-114.
[11] Y. Qiang, S. Zhang, S. Xu, L. Yin, The effect of 5-nitroindazole as an inhibitor
for the corrosion of copper in a 3.0% NaCl solution, RSC Adv. 5 (2015) 63866-63873.
[12] B. Hou, The inhibition effect of polyaspartic acid and its mixed inhibitor on mild
steel corrosion in seawater wet/dry cyclic conditions, Int. J. Electrochem. Sci. 11
(2016) 3024-3038.
[13] S. Marcelin, N. Pébére, Synergistic effect between 8-hydroxyquinoline and
bezotriazole for the corrosion protection of 2024 aluminium alloy: A local
electrochemical impedance approach, Corros. Sci. 101 (2015) 66-74.
[14] M. Faustin, A. Maciuk, P. Salvin, C. Roos, M. Lebrini, Corrosion inhibition of
C38 steel by alkaloid extract of Geissospermum laeve in 1M hydrochloric acid:
Electrochemical and phytochemical studies, Corros. Sci. 92 (2015) 287-300.
[15] G. Ji, S. Anjum, S. Sundaram, R. Rrakash, Musa paradisica peel extract as green
corrosion inhibitor for mild steel in HCl solution, Corros. Sci. 90 (2015) 107-117.
[16] C. Zou, X. Yan, Y. Qin, M, Wang, Y. Liu, Inhibiting evaluation of β -
Cyclodextrin - modified acrylamide polymer on alloy steel in sulfuric solution, Corros.
Sci. 85 (2014) 445-454.
[17] N. Ochoa, M. Bello, J. Sancristóbal, V. Balsamo, A. Albornoz, J.L. Brito,
Modified cassava starches as potential corrosion inhibitors for sustainable
development, Mater. Res. 16 (2013) 1209-1219.
[18] B. Qian, J. Wang, M. Zheng, B. Huo, Synergistic effect of polyaspartic acid and
iodide ion on corrosion inhibition of mild steel in H2SO4, Corros. Sci. 75 (2013)
184-192.
[19] A.A. Farag, A.M. Hegazy, Synergistic inhibition effect of potassium iodide and
novel Schiff bases on X65 steel corrosion in 0.5 M H2SO4, Corros. Sci. 74 (2013)
168-177.
[20] S.A. Umoren, Y. Li, F.H. Wang, Influence of iron microstructure on the
performance of polyacrylic acid as corrosion inhibitor in sulfuric acid solution, Corros.
Sci. 53 (2011) 1778-1785.
2017/6/12 - 21 -
[21] K.P. Ansari, M.A. Quraishi, A. Singh, Schiff’s base of pyridyl substituted
triazoles as new and effective corrosion inhibitors for mild steel in hydrochloric acid
solution, Corros. Sci. 79 (2014) 5-15.
[22] V.V. Torres, V.A. Rayol, M. Magalháes, G.M. Viana, L.C.S. Aguiar, S.P. Machado,
E. D’Elia, Stduy of thioureas derivatives synthesized from a green route as corrosion
inhibitors for mild steel in HCl solution, Corros. Sci. 79 (2014) 108-118.
[23] F. Liu, L. Zhang, X. Yan, X. Lu, Y. Gao, C. Zhao, Effect of diesel on corrosion
inhibitors and application of bio-enzyme corrosion inhibitors in the laboratory cooling
water system, Corros. Sci. 93 (2013) 293-300.
[24] S.A. Umoren, I. Obot, A. Madhankumar, Z. Gasem, Performance evaluation of
pectin as ecofriendly corrosion inhibitor for X60 pipeline steel in acid medium:
Experimental and theoretical approaches, Carbohydr. Polym. 124 (2015) 280-291.
[25] J. Zhang, W.W. Song, D.L. Shi, L.W. Niu, M. Du, A dissymmetric bis-quaternary
ammonium salt gemini surfactant as effective inhibitor for Q235 steel in hydrochloric
acid, Prog. Org. Coat. 75 (2012) 284-291.
[26] J. Zhang, F.M. Zhu, W.W. Song, M. Du, Corrosion inhibition mechanism of
imidazoline-based dissymmetric bis-quaternary ammonium salts with different
hydrophobic chain length on Q235 steel in 1 M HCl solution, J. Surfact. Deterg. 16
(2013) 559-569.
[27] F. Lvušić, O. Lahodny-Šarc, H.O. Ćurkovic, V. Alar, Synergistic inhibition of
carbon steel corrosion in seawater by cerium chloride and sodium gluconate, Corros.
Sci. 98 (2015) 88-97.
[28] P. Feng, K. Wan, G. Cai, L. Yang, Y. Li, Synergistic protective effect of
carboxymethyl chitosan and cathodic protection of X70 pipeline steel in seawater,
RSC Adv. 6 (2017) 3419.
[29] R. Ke, K. Chai, J. Wu, Synergistic effect of iron bacteria and vibrio on carbon
steel corrosion in seawater, Int. J. Electrochem. Sci.11 (2016)7461-7474.
[30] ASTM G 31-72 Standard practice for laboratory immersion corrosion testing of
metals ( re-approved 1995 ) 03.02 (1990).
[31] F.W. Quakenbush, S.L. Miller, Composition and analysis of the carotenoids in
2017/6/12 - 22 -
marigold petals, J. Assoc. Off. Anal. Chem. 55 (1972) 617-621.
[32] E. McCafferty, Validation of corrosion rates measured by the Tafel extrapolation
method, Corros. Sci. 47 (2005) 3202-3215.
[33] B.G. Zhang, C.J. He, X. Chen, Z.P. Tian, F.T. Li, The synergistic effect of
polyamidoamine dendrimers and sodium silicate on the corrosion of carbon steel in
soft water, Corros. Sci. 90 (2015) 585-596.
[34] Q.B. Zhang, Y.X . Hua, Corrosion inhibition of mild steel by alkylimidazolium
ionic liquids in hyddrochloric acid, Electrochim. Acta. 54 (2009) 1881-1887.
[35] H. Bentrah, Y. Rahali, A. Chala, Gum arabic as an eco-friendly inhibitor for API
5L X42 pipeline steel in HCl medium, Corros. Sci. 82 (2014) 426-431.
[36] N.Soltani, N. Tavakkoli, M. Khayatkashani, M.R. Jalali, A. Mosavizade, Green
approach to corrosion inhibition of 304 stainless in hydrochloric acid seawater by the
extract of Salvia officinalis leaves, Corros. Sci. 62 ( 2012 ) 122-135.
[37] E. Barsoukov, J.R. Macdonald, Impedance spectroscopy, theory, experimental
and applications, John Wiley & Sons, 2005.
[38] K.C. Emreg ü l, M. Hayvali, Studies on the effect of vanillin and
protocatechualdehyde on corrosion of steel in hydrochloric acid, Mater. Chem. Phys.
83 (2004) 209.
[39] K.C. Emregül, A.A. Aksüt, The behavior of aluminum in alkaline media, Corros.
Sci. 42 (2000) 2051-2067.
[40] X. Wang, H. Yang, F. Wang, An investigation of benzimidazole derivative as
corrosion inhibitor for mild steel in different concentration HCl solution, Corros. Sci.
53 (2011) 113-121.
[41] E.S. Meresht, T.S. Farahani, J. Neshati, 2-Butyne-1,4-diol as a novel corrosion
inhibition for API X65 steel pipeline in carbonate/bicarbonate seawater, Corros. Sci.
54 (2012) 36-44.
[42] E. Gutiérrez, J.A. Rodríguez, J. Cruz-Borbolla, J.G. Alvarado-Rodriguez, P.
Thangarasu, Development of a predictive model for corrosion inhibition of carbon
steel by imidazole and benzimidazole derivatives, Corros. Sci. 108 (2016) 23-35.
[43] M. Punita, B. Sitashree, M.M. Singh, Corrosion inhibition of mild steel in acidic
2017/6/12 - 23 -
seawater by Tagetes erecta ( Marigold flower ) extract as a green inhibitor, Corros. Sci.
85 (2014) 352-362.
[44] P. Bommersbach, C. Alemany-Dumont, J.P. Millet, B. Normand, Hydrodynamic
effect on the behavior of a corrosion inhibitor film: characterization by
electrochemical impedance spectroscopy, Electrochim. Acta 51 (2006) 4011- 4018.
[45] M. Bouklah, B. Hammouti, M. Lagrenée, F. Bentiss, Thermodynamic properties
of 2,5-bis(4-methoxyphenyl)-1,3,4-oxadiazole as a corrosion inhibition for mild steel
in normal sulfuric acid medium, Corros. Sci. 48 (2006) 2831-2842.
[46] A. Popova, M. Christov, A. Vasilev, Inhibitive properties of quaternary
ammonium bromides of N-containing heterocycles on acid mild steel corrosion. PatⅡ:
EIS results, Corros. Sci. 49 ( 2007 ) 3290-3302.
[47] C.B. Verma, M.A. Quraishi, A. Singh, 2-Aminobenzene-1,3-dicarbonitriles as
green corrosion inhibition for mild steel 1 M HCl: Electrochemical, thermodynamic,
surface and quantum chemical investigation, J. Taiwan Inst. Chem. Eng. 49 (2015)
229-239.
[48] X.H. Li, S.D. Deng, H. Fu, X.G. Xie, Synergistic inhibition effects of bamboo
leaf extract/major components and iodide ion on the corrosion of steel in H3PO4
solution, Corros. Sci. 78 (2014) 29-42.
[49] J. Aljourani, K. Raeissi, M.A. Golozar, Benzimidazole and its derivatives as
corrosion inhibitors for mild steel in 1M HCl solution, Corros. Sci. 51 (2009)
1836-1843.
[50] L. Luo, S.T. Zhang, Y. Qiang, Study on 1-allyl-3-butylimidazalium bromine as
corrosion inhibitor for X65 steel in 0.5 M H2SO4 solution, Int. J. Electrochem. Sci. 11
(2016) 8177-8192.
[51] K. Zhang, B. Xu, W. Yang, X. Yin, Y. Liu, Y. Chen, Halogen-subsituted
imidazoline derivatives as corrosion inhibitors for mild steel in hydrochloric acid
seawater, Corros. Sci. 90 (2015) 284-295.
[52] F. Bentiss, M. Lebrini, M. Lagrenee, Thermodynamic characterization of metal
disseawater and inhibitor adsorption processes in mild steel/2,5-bis(n-thienyl)-1, 3,
4-thiadiazoles/hydrochloric acid system, Corros. Sci. 47 (2005) 2915-2931.
2017/6/12 - 24 -
[53] M.A. Pech-Canul, P. Bartolo-Pérez, Inhibition effects of
N-phosphono-methyl-glycine/Zn2+ mixtures on corrosion of steel in neutral chloride
seawater, Surf. Coat. Technol. 184 (2004) 133-140.
[54] G. Moretti, G. Quartarone, A. Tassan, A. Zingales, Inhibition of mild steel
corrosion in 1N sulphuric acid through indole, Mater. Corros. 45 (1994) 64-647.
[55] M. Bouklah, N. Benchat, B. Hammouti, A. Aouniti, S. Kertit, Thermodynamic
characterisation of steel corrosion and inhibitor adsorption of pyridazine compounds
in 0.5 M H2SO4, Mater. Lett. 60 (2006) 1901-1905.
[56] B. Zhang, C. He, C. Wang, P. Sun, F. Li, Y. Lin, Synergistic corrosion inhibition
of environment-friendly inhibitors on the corrosion of carbon steel in soft water,
Corros. Sci. 94 (2015) 6-20.
[57] A.P. Grosvenor, B.A. Kobe, M.C. Bieginger, Investigation of multiplet splitting
of Fe 2p XPS spectra and bonding in iron compounds, Surf. Interface. Anal. 36 (2004)
1564-1574.
[58] F.Z. Bouanis, F. Bentiss, M. Traisne, C. Jama, Enhance corrosion resistance
properties of radio frequency cold plasma nitrided carbon steel: Gravimetric and
electrochemical results, Electrochim. Acta. 54 (2009) 2371- 2378.

2017/6/12 - 25 -
Fig. 1. Chemical structure of the inhibitors: (a) Glucomannan (GL); (b) Bisquaternary
ammonium salt (BQAS)
-0.46

-0.48
Potential (V/SCE)

-0.50 (a)
(b)
-0.52 (c)
(d)
-0.54

-0.56

0 10 20 30 40 50 60
Time (min)
Fig. 2. OCP variations with time of mild steel in the simulated seawater (a) and
containing 80 mg L-1 BQAS (b), 80 mg L-1GL (c) and 40 mg L-1 BQAS + 40 mg L-1
GL (d) at 25 ℃, pH = 8.0.

2017/6/12 - 26 -
100 1.6
(a) (b) GL+BQAS
1.4
95 GL

Corrosion rate (mg cm-2 h-1)


Inhibition efficiency (η%)

1.2 BQAS
90
1.0
85 0.8

80 0.6
0.4
75
GL+BQAS 0.2
70 GL 0.0
BQAS
65 -0.2
0 15 30 45 60-1 75 90 0 15 30 45 60 75 90
Concentration (mg L ) Concentration (mg L-1)

Fig. 3. Inhibition efficiency (a) and corrosion rate (b) curves obtained from the weight
loss measurement after immersion for 4 h in the simulate seawater in the presence of
BQAS ( ), GL ( ) and BQAS + GL (m(GL):m(BQAS)=1:1) ( ) at 25 ℃, pH = 8.0. Error
bars represent the standard deviations.
Current density log i (A cm-2)

-5

-6
1
1: Blank solution 2
-7 -1
3
2: 80 mgL BQAS
-1
3: 80 mgL GL
-8 4: 40 mg L-1 GL+40 mg L-1 BQAS 4
-0.8 -0.7 -0.6 -0.5 -0.4 -0.3
Potential (V/SCE)
Fig. 4. Potentiodynamic polarization curves of mild steel after immersion for 4 h in
the simulated seawater with and without inhibitors at 25 ℃, pH = 8.0.

2017/6/12 - 27 -
800 4000
(a) Experimental data (b) Experimental data
Solid line shows fitted results Solid line shows fitted results

600 3000

|Z"|(Ω cm2)
|Z"|(Ω cm2)

400 2000

200 1000

0 0
0 200 400 600 800 0 1000 2000 3000 4000
Z'(Ω cm2) Z'(Ω cm2)

2
6000 (c) Experimental data (d) Experimental data % (2)
Solid line shows fitted results 40000 Solid line shows fitted results

4500
|Z"|(Ω cm2)
|Z"|(Ω cm2)

30000

3000 20000

1500 10000

0 0
0 1500 3000 4500 6000 0 10000 20000 30000 40000
Z'(Ω cm2) Z'(Ω cm2)

Fig. 5. Nyquist plots of the mild steel in the simulated seawater: (a) blank solution
( ), (b) 80 mg L-1 BQAS ( ), (c) 80 mg L-1 GL ( ), and (d) the mixture of 40 mg L-1
BQAS and 40 mg L-1 GL ( ) at 25 ℃, pH = 8.0.

Fig. 6. The fitted equivalent circuit for the mild steel in the simulated seawater:
without (a) and with inhibitors (b).

2017/6/12 - 28 -
Current density log i (A cm-2)

-4

-5

1: 298 K
2: 318 K
-6 4 3 3: 338 K
2
1 4: 350 K
-1.0 -0.9 -0.8 -0.7 -0.6 -0.5 -0.4
Potential (V/SCE)
Fig. 7. Effect of the seawater temperature on the polarization curves of the mild steel
after immersion for 4 h in the simulated seawater without inhibitors at pH = 8.0.
Current density log i (A cm-2)

-5

-6

-7
1: 298 K
4 2: 318 K
3 3: 338 K
-8
21 4: 350 K
-0.8 -0.7 -0.6 -0.5 -0.4 -0.3 -0.2
Potential (V/SCE)
Fig. 8. Effect of the seawater temperature on the polarization curves of the mild steel
after immersion for 4 h in the simulated seawater in the presence of the mixture of 40
mg L-1 GL and 40 mg L-1 BQAS at pH = 8.0.

2017/6/12 - 29 -
100 0.14
Inhibition efficiency
99 Corrosion rate

Corrosion rate (mg cm-2 h-1)


Inhibition efficiency (η%)

98 0.12

97

96 0.10

95

94 0.08

93

92 0.06
290 300 310 320 330 340 350
Temperature (K)

Fig. 9. Inhibition efficiency ( ) and corrosion rate ( ) of the mild steel, determined
after immersion for 4 h in the simulated seawater in the presence of 40 mg L-1 GL and
40 mg L-1 BQAS at pH = 8.0 at the various temperatures. Error bars represent the
standard deviations.

Blank solution
2 Linear fit
Slope = -2.0955 Slope = -2.0955
2 2
R =0.996 R =0.996
log icorr (μA cm-2)

0 Mixed inhibitors
Linear fit
Slope = -3.4648
2
R = 0.981

-1
2.9 3.0 3.1 3.2 3.3 3.4
1000/T0 (K)
Fig. 10. Arrhenius plots of log icorr versus 1/T0 for the mild steel in the simulated
seawater in the absence ( ) and in the presence ( ) of 40 mg L-1 GL and 40 mg L-1
BQAS at pH = 8.0 for 4 h of immersion time. Error bars represent the standard
deviations (solid lines show linear fit results).

2017/6/12 - 30 -
106 -1 -1 -1
60 mg L 70 mg L 80 mg L
100 mg L
-1
120 mg L
-1 100
5
10

-Phase angle (degree)


80
104
|Z| (Ω cm2)

60
103

40
102

101 20

100 0
10-2 10-1 100 101 102 103 104 105
Frequency (Hz)
Fig. 11. Bode-phase plots of mild steel after 4 h immersion period in the simulated
seawater with 60 mg L-1 ( ), 70 mg L-1 ( ), 80 mg L-1 ( ), 100 mg L-1 ( ) and 120
mg L-1 ( ) mixture of GL and BQAS (m(GL):m(BQAS)=1:1) at 25 ℃,pH = 8.0.
99.0 0.092
Inhibition efficiency
Corrosion rate
98.5 0.090
Corrosion rate (mg cm-2 h-1)
Inhibition efficiency (ηe%)

98.0 0.088

97.5 0.086

97.0 0.084

96.5 0.082

96.0 0.080
60 70 80 90 100 110 120
Concentration of the mixture (m(GL):m(BQAS)=1:1) (mg L-1)

Fig. 12. Inhibition efficiency ( ) and corrosion rate ( ) of the mild steel, determined
after immersion for 4 h in the simulated seawater with various concentrations of the
mixture at 25 ℃,pH = 8.0. Error bars represent the standard deviations.

2017/6/12 - 31 -
40000 1h
2h
3h
30000
4h
|Z''| (Ω cm2)

8h
20000 12 h

10000

0
0 10000 20000 30000 40000
Z' (Ω cm2)
Fig. 13. Nyquist diagrams of mild steel in the simulated seawater with the mixture of
40 mg L-1 GL and 40 mg L-1 BQAS at 25 ℃, pH = 8.0 for 1 h ( ), 2 h ( ), 3 h ( ), 4 h
( ), 8 h ( ) and 12 h ( ).
99 0.090
Inhibition efficiency
Corrosion rate Corrosion rate (mg cm-2 h-1)
Inhibition efficiency (ηe%)

0.088
98

0.086
97
0.084

96
0.082

95 0.080
0 2 4 6 8 10 12
Time (h)
Fig. 14. Inhibition efficiency ( ) and corrosion rate ( ) of the mild steel in the
simulated seawater at 25 ℃, pH = 8.0 with the mixture of 40 mg L-1 GL and 40 mg
L-1 BQAS for various immersion time. Error bars represent the standard deviations.

2017/6/12 - 32 -
130
Mixed inhibitors
120 Linear fit

110
/θ (mg L-1)

100
y=1.005x+1.631
90 R2=0.999
Cin
0

80

70

60
60 70 80 90 100 110 120 130
Cin0 (mg L-1)

Fig. 15. Langmuir adsorption isotherm curves of the mild steel in the simulated
seawater containing the mixture of 40 mg L-1 GL and 40 mg L-1 BQAS ( ) at 25 ℃.
Error bars represent the standard deviations.

Fig. 16. Schematic illustration of the adsorption of GL molecules and BQAS


molecules at the mild steel surface in the simulated seawater.

2017/6/12 - 33 -
(a) C 1s (b) N 1s
+
C-N, 286.7 eV C-N ,400.4 eV

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

C-C, 285.0 eV Fit

Fit

280 282 284 286 288 290 292 294 396 398 400 402 404 406 408
Binding energy (eV) Binding energy (eV)

Fig. 17. Deconvolution of high-resolution XPS spectra of C 1s (a) and N 1s (b) for
BQAS.

(a) C-C, 285.0 eV C 1s (b) 3 C-O-C, 531.8 eV O 1s


4 4
2 3
1 2
Fit Fit
1
Intensity (a.u)

Intensity (a.u)

C-O, 289.1 eV
C-OH, 533.7 eV

276 278 280 282 284 286 288 290 292 294 296 298 526 528 530 532 534 536 538 540
Binding energy (eV) Binding energy (eV)

Fig. 18. Deconvolution of high-resolution XPS spectra of C 1s (a) and O 1s (b) for GL

2017/6/12 - 34 -
(a) Fe 2p (b) Fe-OH, Fe-O, 530.1 eV O 1s
FexN, 706.4 eV
C-O-C, 531.8 eV

Intensity (a.u)
719.8 eV
Intensity (a.u)

Fe3+, 710.7 eV

Fit
Fe(Ⅱ), 715.0 eV

700 705 710 715 720 725 528 529 530 531 532 533 534 535
Binding energy (eV) Binding energy (eV)

(c) C-C, 285.0 eV C 1s (d) C-N+, 400.4 eV N 1s

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

C-N, 286.7 eV

C-O, 289.1eV N-Fe, 398.7 eV

282 284 286 288 290 396 397 398 399 400 401 402 403
Binding energy (eV) Binding energy (eV)

Fig. 19. Deconvolution of high-resolution XPS spectra of Fe 2p (a), O 1s (b), C 1s (c)


and N 1s (d) for the mild steel after immersion for 4 h in the simulated seawater with
BQAS and GL .

2017/6/12 - 35 -
Fig. 20. SEM images of mild steel surface after 1 and 48 h of immersion in the
simulated seawater without inhibitors (a, b) and with the mixture of 40 mg L-1 GL and
40 mg L-1 BQAS (c, d) at 25 ℃, pH = 8.0.

2017/6/12 - 36 -
Fig. 21. 3D morphology of polished mild steel (a), mild steel after immersion for 24 h
in the simulated seawater at 25 ℃, pH = 8.0 without the inhibitors (b) and with the
mixture of 40 mg L-1 GL and 40 mg L-1 BQAS (c). (a’), (b’), (c’) are the height profile
of the corresponding mild steel surface along the marked lines on panel (a), (b), (c).

2017/6/12 - 37 -
Table. 1. Composition of the mild steel
Chemical element C Si Mn Al P Mo S Fe

wt (%) 0.16 0.30 0.52 0.45 0.04 0.01 0.02 balance

Table. 2. Composition of the simulated seawater


Salt NaCl MgCl2 CaCl2 MgSO4 Ca(NO3)2 H 2O

wt (%) 4.12 2.26 0.19 0.056 0.012 balance

Table 3. Potentiodynamic polarization parameters of the mild steel after immersion


for 4 h in the simulated seawater in the absence and in the presence of inhibitors.
Corrosion inhibitor βa (mV dec-1) βc (mV dec-1) Ecorr V/SCE icorr (μA cm-2 ) ηi%

Blank 238 ± 6 -172 ± 9 -0.553 ± 0.004 17.6 ± 0.3 /

BQAS 127 ± 8 -134 ± 7 -0.513 ± 0.001 3.57 ± 0.03 79.7

GL 120 ± 10 -119 ± 11 -0.503 ± 0.005 2.21 ± 0.02 87.4

The mixture of GL
98 ± 6 -101 ± 12 -0.472 ± 0.003 0.164 ± 0.002 99.1
and BQAS

2017/6/12 - 38 -
Table 4. Electrochemical parameters of the mild steel after immersion for 4 h in the
simulated seawater with and without corrosion inhibitors.
Corrosion Rs Rct Cdl CPE
λ2 n ηe%
inhibitor (Ω cm2) (kΩ cm2) (μF cm-2) (μF cm-2)

Blank 0.156 ± 0.003 0.84 ± 0.01 2917 ± 23 0.00132 ± 0.00001 0.12 ± 0.04 361 ± 13 /

BQAS 0.187 ± 0.006 3.09 ± 0.05 384 ± 6 0.00117 ± 0.00004 0.85 ± 0.02 121 ± 11 72.8

GL 0.192 ± 0.001 5.43 ± 0.09 189 ± 11 0.00394 ± 0.00006 0.87 ± 0.01 98 ± 8 84.5

Mixture of

BQAS and 0.176 ± 0.005 39.47 ± 0.01 105 ± 12 0.00691 ± 0.00003 0.89 ± 0.05 78 ± 2 97.9

GL

Table 5. Inhibition efficiency values observed from the weight loss measurements,
potentiodynamic polarization, electrochemical impedance of mild steel after
immersion for 4 h in the simulated seawater with inhibitors at 25 ℃, pH = 8.0.
Inhibition efficiency (%)
Corrosion inhibitor Weight loss Potentiodynamic Electrochemical impedance
measurement polarization spectroscopic

BQAS 77.5 79.7 72.8

GL 85.4 87.4 84.5

Mixture of BQAS and GL 98.8 99.1 97.9

2017/6/12 - 39 -
Table 6. Potentiodynamic polarization parameters of the mild steel after immersion
for 4 h in the simulated seawater in the absence and in the presence of inhibitors at
different temperatures.
Corrosion βa βc Ecorr icorr
T (K)
inhibitor (mV de-1) (mV dec-1) (V/SCE) (μA cm-2)

298 238 ± 6 -172 ± 9 -0.553 ± 0.004 17.6 ± 0.3

318 246 ± 5 -169 ± 8 -0.576 ± 0.003 26.7 ± 0.2


Blank seawater
338 248 ± 8 -178 ± 5 -0.663 ± 0.003 69.3 ± 0.1

350 262 ± 10 -206 ± 7 -0.753 ± 0.002 94.5 ± 0.2

298 98 ± 6 -101 ± 12 -0.472 ± 0.003 0.164 ± 0.002

Mixture of GL 318 108 ± 5 -104 ± 11 -0.481 ± 0.001 0.683 ± 0.006

and BQAS 338 128 ± 9 -111 ± 5 -0.516 ± 0.004 3.34 ± 0.02

350 133 ± 10 -117 ± 7 -0.558 ± 0.003 6.57 ± 0.04

Table 7. Electrochemical impedance parameters of the mild steel after immesion for 4
h in the simulated seawater containing various concentrations of the BQAS and GL.
Concentrations Rs Rct Cdl CPE
λ2 n
(mg L-1) (Ω cm2) (kΩ cm2) (μF cm-2) (μF cm-2)

60 0.136 ± 0.008 24.89 ± 0.10 136 ± 1 0.00312 ± 0.00011 0.82 ± 0.03 93 ± 3

70 0.119 ± 0.005 30.64 ± 0.08 107 ± 5 0.00121 ± 0.00004 0.81 ± 0.02 86 ± 6

80 0.176 ± 0.005 39.47 ± 0.01 105 ± 12 0.00691 ± 0.00003 0.89 ± 0.05 78 ± 2

100 0.196 ± 0.003 42.48 ± 0.13 99 ± 1 0.00532 ± 0.00014 0.91 ± 0.04 75 ± 5

120 0.211 ± 0.005 42.59 ± 0.02 98 ± 2 0.00221 ± 0.00007 0.90 ± 0.07 76 ± 6

2017/6/12 - 40 -
Table 8. Electrochemical impedance parameters of the mild steel in the simulated
seawater containing the mixed inhibitors for various immersion time.
Rs Rct Cdl CPE
Times (h) λ2 n
(Ω cm2) (kΩ cm2) (μF cm-2) (μF cm-2)

1 0.136 ± 0.002 17.83 ± 0.12 134 ± 4 0.00217 ± 0.00011 0.82 ± 0.03 109 ± 3

2 0.119 ± 0.001 34.42 ± 0.09 107 ± 3 0.00181 ± 0.00011 0.93 ± 0.05 86 ± 7

3 0.172 ± 0.002 36.08 ± 0.21 103 ± 5 0.00211 ± 0.00032 0.86 ± 0.11 82 ± 5

4 0.176 ± 0.005 39.47 ± 0.01 105 ± 12 0.00691 ± 0.00003 0.89 ± 0.05 78 ± 2

8 0.211 ± 0.009 40.08 ± 0.06 94 ± 1 0.00231 ± 0.00008 0.90 ± 0.03 78 ± 1

12 0.185 ± 0.004 40.72 ± 0.08 91 ± 5 0.00421 ± 0.00016 0.92 ± 0.03 75 ± 3

Table 9. Equilibrium constant and adsorptive standard free energy at 298 K.

Temperature (K) Kads ( L mg-1) Δ G 0ads (kJ mol-1)

298 0.613 -33.016

2017/6/12 - 41 -

You might also like