You are on page 1of 6

Materials Chemistry and Physics 126 (2011) 983988

Contents lists available at ScienceDirect

Materials Chemistry and Physics


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

The inhibition effect of some pyrimidine derivatives on austenitic stainless steel


in acidic media
Necla Caliskan , Esvet Akbas
Chemistry Department, Faculty of Sciences and Letters, University of Yuzuncu Yil, 65080 Van, Turkey

a r t i c l e

i n f o

Article history:
Received 10 March 2010
Received in revised form
18 September 2010
Accepted 29 November 2010
Keywords:
Corrosion
Austenitic stainless steel
Polarization
Pyrimidine derivatives

a b s t r a c t
5-Benzoyl-4-(substitutedphenyl)-6-phenyl-3,4-dihydropyrimidine-2(1H)-(thio)ones (DHPMs) (I and II)
were synthesized using the Biginelli three component cyclocondensation reaction of an appropriate
-diketone, arylaldehyde, and (thio) urea. The effect of these corrosion inhibitors on the corrosion of
austenitic stainless steel in 0.5 M H2 SO4 has been studied by electrochemical methods using Tafel plot,
linear polarization and electrochemical impedance spectroscopy at 298 K. The inhibition efciencies
obtained from all the methods employed are in good agreement. The adsorption of the DHPMs onto the
stainless steel surface was found to follow Langmuir and DubininRadushkevich adsorption isotherm
models. Negative values of Gads in the acidic media ensured the spontaneity of the adsorption process. Results show DHPM I to be the best inhibitor with a mean efciency of 91% at 2 103 M additive
concentration.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Stainless steel has a wide scope of applications in different
industries. This type of stainless steel is covered with a protective
lm rich in chromium (oxides/hydroxides) that imparts corrosion
resistance to its surface. Amount of chromium prevents the formation of rust in unpolluted atmospheric environments. However,
acidic solutions are aggressive to this lm layer and results in
severe pitting formation [1,2]. Acid solutions are widely used in
industries for pickling, acid cleaning of boilers, descaling and oil
well acidizing [3,4]. Sulfuric acid is generally the choice in the
steel surface treatment basically due to its lower cost, minimal
fumes and non-corrosive nature of the SO4 2 ion [5]. Corrosion
inhibitors are needed to reduce the corrosion rates of metallic
materials in this area. Most of the well-known acid inhibitors are
heterocyclic compounds containing bonds, heteroatom phosphorus, sulfur, oxygen and nitrogen [6] as well as aromatic rings
in their structure which are the major adsorption centers [7].
The compounds containing both nitrogen and sulfur can provide
excellent inhibition, compared with compounds containing only
nitrogen or sulfur [4,8]. Generally, a strong interaction causes
higher inhibition efciency, the inhibition increases in the sequence
O < N < S [9,10].
Recently, some studies have been presented on heterocyclic
compounds, such as pyrimidine [9,11], pyridazine [10], bicycloisox-

azolidine [12], Schiff base [13,14], and pyrazole derivatives [15] as


effective corrosion inhibitors for metals in acidic media.
The aim of this study is to investigate the corrosion of
stainless steel in 0.5 M H2 SO4 solution in the presence of two
pyrimidine compounds namely 5-benzoyl-4,6-diphenyl-1,2,3,4tetrahydro-2-thiopyrimidine (DHPM I) and 5-benzoyl-6-phenyl4-p-tolyl-3,4-dihydropyrimidine-2(1H)-one (DHPM II) that these
compounds contain NHCONH and NHCSNH groups
as active centers (Scheme 1). Moreover, these molecules can
be easily synthesized from relatively cheap materials. Pyrimidines, however, have hardly been studied in spite of their
corrosion inhibition properties. They nd diverse applications
in pharmaceutical applications such as analgesic, antipyretic,
antihypertensive, anti-inammatory drugs, pesticides, herbicides,
plant growth regulators, and organic calcium channel modulators
[1624].
DHPMs were prepared via the general method of Biginelli
cyclocondensation reaction in acetic acid. The investigation is performed using potentiodynamic polarization and electrochemical
impedance spectroscopy techniques in order to explain the mechanism of the inhibition action. Adsorption isotherms were tested for
their relevance to describe the adsorption behavior of the studied
compounds.
2. Experimental
2.1. Preparation of electrode

Corresponding author. Tel.: +90 432 2251024/2283.


E-mail address: ncaliskan7@hotmail.com (N. Caliskan).
0254-0584/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2010.11.051

The working electrode was mounted in polyester with following chemical composition (wt.%) C: 0.0425; Si: 0.421; Mn: 2.13; P: 0.0133; S: 0.113; Cr: 18.51; Mo:

984

N. Caliskan, E. Akbas / Materials Chemistry and Physics 126 (2011) 983988

Scheme 1. The molecular structure and name of inhibitor used 5-benzoyl-4,6diphenyl-1,2,3,4-tetrahydro-2-thiopyrimidin (DHPM I); 5-Benzoyl-6-phenyl-4-ptolyl-3,4-dihydropyrimidin-2(1H)-one (DHPM II).

0.563; Ni: 8.34; Al: 0.0334; Co: 0.0901; Cu: 0.358; Fe: balance. The steel electrode was coated with polyester except its bottom surface with surface area of
0.19625 cm2 .

Fig. 1. Polarization curves for steel in 0.5 M H2 SO4 in the absence and presence of
different concentrations of DHPM I at 298 K.

calculated by:

2.2. Materials
A mixture of 1,3-diphenyl-1,3-propanedione (1.6 mmol), aryl aldehyde
(1.1 mmol), (thio)urea (1.1 mmol) and 20 ml of glacial acetic acid containing a few
drops concentrated hydrochloric acid was heated under reux for 8 h. After the
precipitate were ltered off, the solution was held at the room temperature 25 C
for one night and recrystallized from suitable solvents. The synthesis method of
DHPM II was not able to be reached, in spite of it is available the theoretical calculations [25a]. Therefore, it has been thought that the synthesis method applied in the
present study is different [25b]. DHPM I synthesized according to Ref. [26]. DHPMs
were characterized on the basis of their spectral data and elementary analyses. The
molecular structures of DHPM I and II are shown in Scheme 1.
2.3. Test solutions
Solvents and material synthesis chemicals are commercially available (Aldrich,
Fluka, Merck) and were used without further purication. 0.5 mol dm3 H2 SO4 and
inhibitor solutions were prepared using double distilled water. Inhibitor concentrations were chosen as 1 104 , 5 104 , 1 103 and 2 103 mol dm3 . Solutions
were mixed with a magnetic stirrer. For each experiment, a freshly prepared solution
was used.

Rp = A

dE
di

where A is the surface area of the electrode. The Rp values were used to calculate the
inhibition efciencies (Rp , %), using the following equation. Rp and Rp are the polarization resistances in the presence and absence of the organic additive, respectively.

Rp (%) =

 R

Rp


100

Rp

The EIS experiments were conducted in the frequency range of 40 000 Hz to


0.01 Hz at corrosion potential by applying 5 mV sine wave AC voltage. The double
layer capacitance (Cdl ) and the polarization resistance (Rp ) were calculated from
Nyquist plots [27,28].
Since Rp is inversely proportional to the corrosion current it can be used to
calculate the inhibitor efciency (EIS , %), from the relation:
EIS (%) =

 R

Rp
Rp

100

where Rp and Rp are uninhibited and inhibited polarization resistances, respectively.

2.4. Electrochemical measurements

3. Results and discussion


The potentiodynamic polarization curves, linear polarization resistance (LPR)
and electrochemical impedance spectroscopy (EIS) measurements were carried out
using a PC controlled Iviumstat PGZ 301 system with an electrochemical system
frequency response analyser (FRA). A double-wall one-compartment cell with a
three-electrode conguration was used. Before each experiment, the electrode was
polished with 1200 grit emery paper, washed thoroughly with bidistilled water
and put into the cell. A platinum rod was used as the counter electrode and a
Ag/AgCl electrode served as a reference electrode. All the potentials reported here
were referred to the Ag/AgCl. The cell was a water-jacketed version, connected to a
constant temperature circulator at 298 K.
The working electrode was immersed in test solution for 30 min to establish
steady state open circuit potential (Eocp ). After measuring the Eocp , the electrochemical measurements were performed. The potential sweep rate was 1 mV s1 . The
potential was scanned in the negative direction from Ecorr and subsequently in the
positive direction. Inhibition efciencies, pol (%), can be calculated from corrosion
currents determined using the Tafel extrapolation method [27]. The percentage inhibition efciency, , and the surface coverage, , were calculated from the following
equations:

pol (%) =

=

i i
i

3.1. Polarization measurements


Potentiodynamic polarization curves of steel in 0.5 M H2 SO4 in
the absence and presence of various concentrations of DHPMs at
298 K are shown in Figs. 1 and 2. Anodic and cathodic currents were
inhibited more effectively with an increase in the concentration of

100

i i
i

where i and i are the corrosion current densities without and with inhibitor. The
corrosion currents were determined from the intersection of the Tafel plots for each
case.
LPR measurements were carried out by recording the electrode potential
10 mV around open circuit potential with 0.1 mV s1 scan rate [27]. Polarization
resistance values were obtained from the slope of the potentialcurrent lines and

Fig. 2. Polarization curves for steel in 0.5 M H2 SO4 in the absence and presence of
different concentrations of DHPM II at 298 K.

N. Caliskan, E. Akbas / Materials Chemistry and Physics 126 (2011) 983988

985

Table 1
Corrosion parameters for steel in 0.5 M H2 SO4 in the absence and presence of DHPM I and II.
DHPMs

Inhibitor concentration (mol dm3 )

Ecorr (V)

icorr (mA cm2 )

a (mV dec1 )

c (mV dec1 )

pol (%)

a

None
1 104
5 104
1 103
2 103

0.393
0.339
0.320
0.296
0.259

0.0465
0.0124
0.0062
0.0057
0.0046

93
62
68
49
42

85
89
101
136
176

73
87
88
90

0.73
0.87
0.88
0.90

II

None
1 104
5 104
1 103
2 103

0.393
0.356
0.343
0.341
0.340

0.0465
0.0234
0.0189
0.0102
0.0081

93
51
55
57
58

85
110
105
103
100

50
59
78
83

0.50
0.59
0.78
0.83

Degree of surface coverage.

the inhibitors, but the reduction in the anodic current was greater
than that of the cathodic current.
These results showed that the addition of DHPMs reduces anodic
dissolution and also retards the hydrogen evolution reaction.
The corresponding electrochemical parameters, i.e., corrosion
current density (icorr ), Tafel slope constants (a and c ), corrosion
potential (Ecorr ), inhibition efciency,  (%) and surface coverage
degrees () values were calculated from these curves and are given
in Table 1. The inhibition efciency follows order, DHPM I > DHPM
II.
In acidic solutions, the anodic reaction of corrosion is the passage of metal ions from the metal surface into the solution, and
the cathodic reaction is the discharge of hydrogen ions to produce
hydrogen gas or to reduce oxygen. The inhibitor may affect either
the anodic or the cathodic reaction, or both [29,30]. The values of
the anodic Tafel constant, a , calculated for inhibited solutions are
smaller than obtained for the uninhibited in H2 SO4 concentration;
this indicates that the inhibitors affected anodic reactions. The values of the cathodic Tafel constant, c , showed an opposite trend.
Also corrosion potentials shifted to the positive direction. A compound can be classied as an anodic- or a cathodic-type inhibitor
when the change in the Ecorr value is larger than 85 mV [30,31].
Since the largest displacement exhibited by DHPM I was 134 mV
(Table 1), it may be concluded that this molecule should be considered as an anodic-type inhibitor, meaning that the addition of
compounds to a 0.5 M H2 SO4 solution reduces the anodic dissolution of steel. The change of Ecorr was observed 53 mV; therefore the
DHPM II is a mixed-type inhibitor (Table 1), meaning that the addition of compounds to a 0.5 M H2 SO4 solution reduces the anodic
dissolution of steel and also retards the cathodic hydrogen evolution reaction. The corrosion inhibition increased with inhibitor
concentration. Particularly, the DHPM I exhibited higher inhibition effect than DHPM II. The process of inhibition is inuenced
by the nature and the charge of the metal, chemical structure of
the organic inhibitor and the type of electrolyte [32].
The surface charge of the metal is due to the electrical eld which
emerges at the interface by immersion in the electrolyte. This surface charge can be determined by comparing the potential of zero
charge (PZC) and the stationary potential of the metal in the corresponding medium [4]. As PZC corresponds to the potential at which
the surface of electrode is charge-free, at the corrosion potential
the metal surface can be positively or negatively charged. The surface charge of immersed iron in the H2 SO4 solution is positive or
slightly negative compared with PZC [4,32]. Corrosion mechanism
of the acid dissolution of stainless steel is dependent not only on
the hydrogen ion concentration but also on the counter ion of the
acid. The majority of corrosion inhibition studies of steel in H2 SO4
media using different organic compounds showed less inhibition
efciency than in HCl medium [33]. Therefore some studies are also
continuing in our laboratory on pyrimidine compounds on stainless
steel in hydrochloric acid.

It was reported [34] that SO4 2 has a low tendency to adsorb


on the steel surface. The steel surface is more positively charged
at corrosion potential. Hence the concentration of surface complex formed is not sufcient to cover the electrode surface from
the solution or the complex is not so stable that it is desorbed from the surface. In both cases, the corrosion rate is high
[34].
Benzoyl group may not be effective as adsorption center because
of the tautomerisation benzoyl group and pyrimidine ring system
in the compounds (Scheme 2). Moreover, the ligands might not be
deprotonize in acid solutions, it means that only three regions of
C O (pyr), C S and NH are responsible for interaction (Scheme 1).
The high inhibition effect of the DHPM I could be explained by
the presence of the sulfur atom. The functional groups, such as C S,
NH in the inhibitor molecules are superior as the adsorption center
for DHPM I. Due to the high nucleophilic character of S atom the
C S region is convenient for adsorption. The functional groups of
DHPM II in which the nucleophilic character of N(3) atom is higher
than O atom, are the regions of C O (pyr), NH and the last one
might interact with steel. As explained above the compounds are
adsorbed through strong interactions between negatively charged
atoms of sulfur, oxygen or nitrogen and positively charged steel
surface.
3.2. Linear polarization measurements
Polarization resistance and inhibitor efciency increases with
increasing inhibitor concentration. Polarization resistance and the
inhibitor efciency Rp (%) values are given in Table 2 indicated
that the adsorption of the inhibitor on the steel surface blocked
the active sites and inhibit corrosion [27]. The inhibition efciency
follows order, DHPM I > DHPM II.
3.3. Electrochemical impedance spectroscopy measurements (EIS)
EIS provides a rapid and convenient way to evaluate the performance of the organic-coated metals and has been widely used for
investigation of protective properties of organic inhibitors on metals [35]. Nyquist plots of steel in 0.5 M H2 SO4 solution in absence

Scheme 2. Schematic representation of studied pyrimidine compounds belong to


tautomerisation.

986

N. Caliskan, E. Akbas / Materials Chemistry and Physics 126 (2011) 983988

Table 2
Polarization resistances and inhibition efciencies obtained from the linear polarization method for steel in 0.5 M H2 SO4 .
DHPMs

Inhibitor concentration (mol dm3 )

Ecorr (V)

Rp ( cm2 )

Rp

None
1 104
5 104
1 103
2 103

0.396
0.340
0.320
0.300
0.266

87.52
244.92
607.30
750.69
858.12

64
86
88
90

II

None
1 104
5 104
1 103
2 103

0.396
0.363
0.351
0.345
0.340

87.52
149.26
185.54
344.43
458.22

41
53
75
81

Fig. 5. Nyquist diagrams for steel in 0.5 M H2 SO4 in the absence and presence of
different concentrations of DHPM II at 298 K.

eters such as charge transfer resistance of the corrosion reaction


(Rp ) and the capacity of the double layer (Cdl ) can be deduced from
the analysis of the Nyquist diagram [27,38]. The charge transfer
resistance, Rp , values could be calculated from the difference in
impedance at lower and higher frequencies. Cdl can be determined
at maximum frequency (fmax ), at which the imaginary component
of the Nyquist plot is maximum, and calculated using the follow
relation:
Cdl =
Fig. 3. Nyquist diagrams for steel in 0.5 M H2 SO4 of DHPMs in 2 103 mol dm3
concentration at 298 K.

and presence of various concentrations of pyrimidine derivatives


are given in Figs. 35.
It could be noticed from the data of Figs. 35 that the impedance
semicircle size depends on the type and concentration of the
inhibitor used.
The diameter of semicircle increased after the addition of
inhibitors to the aggressive solution. This increase was more pronounced with increasing inhibitor concentration which indicates
adsorption of inhibitor molecules on the metal surface [36] The
semicircles are observed to be depressed into the Z (reel axis) of
the Nyquist plot as a result of the roughness of the electrode surface
and other inhomogeneities of the solid surface [37,38]. The param-

1
1

Rp
2fmax

The calculated impedance parameters are given in Table 3.


The Nyquist plots are not perfect semicircles as expected from
the theory of EIS. The deviation from ideal semicircle is generally
attributed to the frequency dispersion as well as to the inhomogeneities of surface and mass transport resistant [36]. On the metal
side, electrons control the charge distribution whereas on the solution side it is controlled by ions. Since ions are much larger than the
electrons, the equivalent ions to the charge on the metal will occupy
quite a large volume on the solution side of the double layer [39,40].
It can be obtained from Table 3 that, the charge transfer resistance
(Rp ) increases as the concentration of the inhibitor increases for the
two inhibitors studied. The increase in Rp values is attributed to
the formation of protective lm on the electrode/solution interface
[32]. The thickness of this protective layer increases with increase
in inhibitor concentration.
Also, the capacitance of electrical double layer (Cdl ) decreases
in the presence of inhibitors. This decrease in Cdl results from a
decrease in local dielectric constant and/or an increase in the thickness of the double layer [41], suggesting that DHPMs inhibit the iron
Table 3
Polarization resistance and inhibitor efciencies for steel in 0.5 M H2 SO4 obtained
using the impedance method.

Fig. 4. Nyquist diagrams for steel in 0.5 M H2 SO4 in the absence and presence of
different concentrations of DHPM I at 298 K.

DHPMs

Inhibitor
concentration
(mol dm3 )

Rp ( cm22 )

Cdl (F cm2 )

fmax

EIS (%)

None
1 104
5 104
1 103
2 103

65.78
252.82
693.44
721.94
812.21

235.0
67.0
31.4
18.1
16.1

10.3
9.4
7.3
12.2
12.2

74
91
91
92

II

None
1 104
5 104
1 103
2 103

65.78
121.95
179.03
344.49
445.22

235.0
113.5
90.75
53.74
45.27

10.3
11.5
9.8
8.6
7.9

46
63
81
85

N. Caliskan, E. Akbas / Materials Chemistry and Physics 126 (2011) 983988

987

corrosion by adsorption at steel/acid interface. A low capacitance


may result if water molecules at the electrode interface are largely
replaced by organic inhibitor molecules through adsorption.
The inhibition efciency follows the order: DHPM I > DHPM II.
3.4. Adsorption isotherm
The adsorption process consists of the replacement of water
molecules at a corroding interface according to following process.
Org(sol) + nH2 O(ads) Org(ads) + nH2 O(sol)
where Org(sol) and Org(ads) are the organic molecules in the solution and adsorbed on the metal surface, respectively, and n is the
number of water molecules replaced by the organic molecules [36].
It is generally assumed that the adsorption of the inhibitor at the
electrode/solution interface is the rst step in the action mechanism of the inhibitors in aggressive acid media. Four types of
adsorption may take place in the inhibiting phenomena involving
organic molecules at the metal/solution interface: (1) electrostatic
attraction between charged molecules and the charged metal, (2)
interaction of unshared electron pairs in the molecule with the
metal, (3) interaction of electrons with the metal and (4) a combination of the above [42]. Adsorption results from the polar or
charged nature of the organic molecule/ionic species rst establishing a physisorbed surface lm (through electrostatic or van der
Waals forces) that may stabilize further through chemisorption to
form a donor type bond. Other mechanisms are concerned with
precipitation, complexation, etc. [43].
It is essential to know the mode of adsorption and the adsorption
isotherm that can give important information on the interaction of
inhibitor and metal surface [27]. Several adsorption isotherms were
tested in order to nd the best suitable adsorption isotherm for
adsorption of DHPMs on the stainless steel surface in 0.5 M H2 SO4
solution. Langmuir adsorption isotherms were found more suitable
with an average correlation coefcient of 0.9999 for DHPMs (Fig. 6).
Langmuir adsorption isotherm can be expressed by the following
equation:
C(inh)


1
+ C(inh)
K(ads)

where  is the degree of coverage on the metal surface, C(inh)


is the inhibitor concentration in the electrolyte and K(ads) is the
equilibrium constant for the adsorptiondesorption process. A
representative Langmuir adsorption isotherm using the potentiodynamic polarization data is given in Fig. 6.
The values of equilibrium constant, Kads calculated from the
reciprocal of the intercept of isotherm line. The values of K are
found as 10 104 and 5 104 dm3 mol1 for compound DHPM I
and DHPM II successively.
The standard free energy of adsorption of inhibitors (Gads ) on
steel surface can be evaluated with the following equation:

Fig. 6. Langmuir adsorption plot for steel in 0.5 M H2 SO4 containing different concentrations of DHPM I and DHPM II.

negative values of Gads showed that the adsorption of inhibitor


molecules on the metal surface is spontaneous [27,39]. The value of
Gads is less than 40 kJ mol1 , indicating electrostatic interaction
between the charged metal surface, i.e., physical adsorption [44].
Experimental results obtained in this study were further tted
into DubininRadushkevich isotherm model (DR). This model was
initially used to distinguish between physical and chemical adsorption for removal of some pollutants from aqueous solutions by
adsorption on various adsorbents [45]. Recently, Solomon et al. [46]
have applied this model in explaining the mechanism of adsorption
of corrosion inhibitor onto a metal surface in acidic medium.
The DR isotherm is more general than the Langmuir isotherm,
because it does not assume a homogeneous surface or constant
sorption potential. The DR equation is [43]:
ln  = ln max K  2
where  max is the maximum surface coverage, K is the constant
related to adsorption energy (mol2 kJ2 ), is Polanyi potential. The
Polanyi potential (kJ mol1 ) is written as:

= RT  ln 1 +

Cinh

In this equation, Cinh is the concentration of the inhibitor. R is gas


constant (kJ K1 mol1 ) and T is temperature (K).
The application of the DR isotherm model is depicted in Fig. 7,
which shows that a plots of ln  versus 2 . Application of the DR

Gads = RT ln K
The standard free energy of adsorption values of 20 kJ mol1
or less negative are associated with an electrostatic interaction
between charged molecules and charged metal surface (physical adsorption); those of 40 kJ mol1 or more negative involves
charge sharing or transfer from the inhibitor molecules to the metal
surface to form a co-ordinate covalent bond (chemical adsorption)
[37].
The calculated standard free energy of adsorption values were
found 28.81 and 26.81 kJ mol1 for DHPM I and DHPM II,
respectively. (Thermodynamic parameters were determined from
the potentiodynamic polarization measurements.) The decreasing
value of Gads reects the increasing adsorption capability. The

Fig. 7. DubininRadushkevich adsorption isotherm model for stainless steel in 0.5 M


H2 SO4 containing DHPMs.

988

N. Caliskan, E. Akbas / Materials Chemistry and Physics 126 (2011) 983988

adsorption isotherm model gave adsorption energies of 3.25 and


3.22 kJ mol1 , respectively, for DHPM I and DHPM II at 298 K. The
adsorption energy values provide information about the adsorption mechanism, i.e. whether it involves chemical or physical
adsorption. Thus, if the value of adsorption energy is between
8 kJ mol1 and 16 kJ mol1 , the adsorption process corresponds to
ion exchange. Conversely, if the adsorption energy is less than
8 kJ mol1 , the adsorption process is physical in nature [42,43].
The mean free energy of the adsorption E is
E = (2K  )

0.5

In the present work were less than those expected for chemical
adsorption process, thereby suggesting that the adsorption mechanism may be a combination of electrostatic interaction and physical
sorption.
4. Conclusions
The inhibition and adsorption effect of DHPMs on the corrosion
behavior of the stainless steel in 0.5 M H2 SO4 was studied using
different techniques. The following points can be emphasized:

[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]

[22]

1. The pyrimidine compounds studied are good inhibitors for stainless steel in 0.5 M H2 SO4 .
2. The inhibition efciency increases with inhibitor concentration.
3. The addition of DHPM I and DHPM II induces a decrease in both
anodic and cathodic currents. The corrosion potential (Ecorr ) of
DHPM I was observed to shift towards more noble potentials
with increasing additive concentration, indicating the inhibitors
to be of anodic character and formation of a surface lm. Since
the largest displacement exhibited by DHPM II was 53 mV, it
may be concluded that this inhibitor should be considered as a
mixed-type inhibitor. The presence of the sulfur atom increases
the inhibition efciency.
4. The negative values of Gads show that the adsorption of
inhibitor molecules on the metal surface is spontaneous.
5. The adsorption DHPMs on stainless steel surface can be approximated by Langmuir and DubininRadushkevich isotherm
models. DubininRadushkevich model suggests that DHPMs are
adsorbed on stainless steel surface by physical adsorption mechanism.
Acknowledgement
This study was supported by the Research Fund of Yznc Yl
University, Van, Turkey.
References
[1] A. Galal, N.F. Atta, M.H.S. Al-Hassan, Mater. Chem. Phys. 89 (2005) 38.
[2] A.S. Fouda, M. Abdallah, S.M. Al-Ashrey, A.A. Abdel-Fattah, Desalination 250
(2010) 538.

[23]
[24]
[25]

[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]

S.A. Abd El-Maksoud, A.S. Fouda, Mater. Chem. Phys. 93 (2005) 84.
J. Aljourani, M.A. Golozar, K. Raeissi, Mater. Chem. Phys. 121 (2010) 320.
R. Hasanov, S. Bilge, S. Bilgic, G. Gece, Z. Klc, Corros. Sci. 52 (2010) 984.
S.A. Abd El-Maksoud, Appl. Surf. Sci. 206 (2003) 129.
F. Xu, J. Duan, S. Zhang, B. Hou, Mater. Lett. 62 (2008) 4072.
Y. Abboud, A. Abourriche, T. Saffaj, M. Berrada, M. Charrouf, A. Bennamara, N.
Al Himidi, H. Hannache, Mater. Chem. Phys. 105 (2007) 1.
H.S. Awad, S.A. Gawad, Anti-Corros. Method Mater. 52 (2005) 328.
A. Chetouani, A. Aouniti, B. Hammouti, N. Benchat, T. Benhadda, S. Kertit, Corros.
Sci. 45 (2003) 1675.
G. Bereket, C. gretir, M. Yaman, E. Hr, J. Mol. Struct. (Theochem) 625 (2003)
31.
S.A. Ali, H.A. Al-Muallem, M.T. Saeed, S.U. Rahman, Corros. Sci. 50 (2008)
664.
S. Bilgic, N. Calskan, J. Appl. Electrochem. 31 (2001) 79.
R. Hasanov, M. Sadkoglu, S. Bilgic, Appl. Surf. Sci. 253 (2007) 3913.
M.A. Elmorsi, A.M. Hassanein, Corros. Sci. 41 (1999) 2337.
N.A. Hassan, Molecules 5 (2000) 827.
M. Pemmsin, C. Lnu-Due, F. Hoguet, C. Gaultier, J. Narcisse, Eur. J. Chem. 23
(1988) 543.
P.A.S. Smith, R.O. Kan, J. Org. Chem. 29 (1964) 2261.
A. Cannito, M. Pemmsin, C. Lnu-Due, F. Hoguet, C. Gaultier, J. Narcisse, Eur. J.
Chem. 25 (1990) 635.
S. Nega, J. Aionso, A. Diazj, F. Junquere, J. Heterocycl. Chem. 27 (1990)
269.
S. Tetsuo, T. Mikio, H. Hidetoshi, H. Daijiro, I. Akira, Jpn. Kokai Tokyo Koho JP
62,132, 884 (1987);
S. Tetsuo, T. Mikio, Chem. Abstr. 107 (1987) 198350h.
P.K. Chakaravorty, W.J. Grelnlee, K. Dooseap, N.B. Mantlo, A.A. Patchett, A.P.C.T.
Int. Appl. WO 92.20.687.156 (1992);
P.K. Chakaravorty, W.J. Grelnlee, K. Dooseap, N.B. Mantlo, A.A. Patchett, Chem.
Abstr. 118 (1993) 213104d.
C.J. Shishoo, K.S. Jain, J. Heterocycl. Chem. 29 (1992) 883.
C.O. Kappe, Molecules 3 (1998) 1.
(a) L.Y. Ivanovskaya, Z.D. Dubovenko, V.P. Mamaev, Inst. Org. Khim. Novosibirsk,
USSR (Izvestiya Sibirskogo Otdeleniya Akademii Nauk SSSR, Seriya Khimicheskikh Nauk) 6 (1969) 132;
(b) F. Aslanoglu, Master Thesis, Yuzuncu Yil University, FBE Van, Turkey, 2007.
F. Aslanoglu, E. Akbas, M. Snmez, B. Anl, Phosphorus Sulfur Silicon Relat. Elem.
182 (2007) 1589.
K.C. Emregul, E. Duzgun, O. Atakol, Corros. Sci. 48 (2006) 3243.
A.V. Shanbhag, T.V. Venkatesha, R.A. Prabhu, R.G. Kalkhambkar, G.M. Kulkarni,
J. Appl. Electrochem. 38 (2008) 279.
S.V. Ramesh, A.V. Adhikari, Corros. Sci. 50 (2008) 55.
A.Y. Musa, A.A.H. Kadhum, A.B. Mohamad, M.S. Takriff, A.R. Daud, S.K.
Kamarudin, Corros. Sci. 52 (2010) 526.
W. Li, Q. He, S. Zhang, C. Pei, B. Hou, J. Appl. Electrochem. 38 (2008) 289.
S.K. Shukla, M.A. Quraishi, Mater. Chem. Phys. 120 (2010) 142.
M.K. Pavithra, T.V. Venkatesha, K. Vathsala, K.O. Nayana, Corros. Sci. 52 (2010)
3811.
A.A. Hermas, M.S. Morad, Corros. Sci. 50 (2008) 2710.
R. Solmaz, G. Kardas, M. Culha, B. Yazc, M. Erbil, Electrochim. Acta 53 (2008)
5941.
R. Solmaz, Corros. Sci. 52 (2010) 3321.
S.V. Ramesh, A.V. Adhikari, Mater. Chem. Phys. 115 (2009) 618.
S.A. Abd El-Maksoud, Int. J. Electrochem. Sci. 3 (2008) 528.
H. Ashassi-Sorkhabi, B. Shaabani, D. Seifzadeha, Electrochim. Acta 50 (2005)
3446.
M. Ozcan, I. Dehri, M. Erbil, Appl. Surf. Sci. 236 (2004) 155.
M. Behpour, S.M. Ghoreishi, N. Soltani, M. Salavati-Niasari, Corros. Sci. 51 (2009)
1073.
S. Bilgic, N. Caliskan, Appl. Surf. Sci. 152 (1999) 107.
S.A. Abd El-Maksoud, Mater. Corros. 54 (2003) 106.
I.B. Obot, N.O. Obi-Egbedi, N.W. Odozi, Corros. Sci. 52 (2010) 923.
A.R. Kul, N. Caliskan, Adsorpt. Sci. Technol. 27 (2009) 85.
M.M. Solomon, S.A. Umoren, I.I. Udosoro, A.P. Udoh, Corros. Sci. 52 (2010)
1317.

You might also like