You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/229225001

The inhibitive effect of hexamethylenetetramine on the acid corrosion of steel

Article  in  Materials Chemistry and Physics · July 2007


DOI: 10.1016/j.matchemphys.2007.02.073

CITATIONS READS

113 1,751

3 authors, including:

Emel Bayol Kadriye Kayakırılmaz


Ömer Halisdemir University Ömer Halisdemir University
33 PUBLICATIONS   491 CITATIONS    37 PUBLICATIONS   637 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Interactions of some Schiff base compounds with mild steel surface in hydrochloric acid solution View project

Bazı Azür Bileşikleri ve Türevlerinin Asidik Ortamda Yumuşak Çeliğin Elektrokimyasal Davranışına Etkilerinin Belirlenmesi View project

All content following this page was uploaded by Emel Bayol on 12 July 2018.

The user has requested enhancement of the downloaded file.


Materials Chemistry and Physics 104 (2007) 74–82

The inhibitive effect of hexamethylenetetramine


on the acid corrosion of steel
E. Bayol a,∗ , K. Kayakırılmaz a , M. Erbil b
aDepartment of Chemistry, Faculty of Science and Art, Nigde University, 51200 Nigde, Turkey
b Department of Chemistry, Faculty of Science and Art, Cukurova University, 01330 Adana, Turkey
Received 10 February 2006; received in revised form 20 February 2007; accepted 22 February 2007

Abstract
The efficiency of hexamethylenetetramine (HMTA), as corrosion inhibitors for steel in de-aerated 0.3 M HCl, 0.1 M H2 SO4 and 0.1 M
H2 SO4 + 1.0 × 10−3 M HCl solutions have been determined by electrochemical studies. It was found that the HMTA acts a good corrosion inhibitor
for steel corrosion in acids solutions. Increase in inhibition efficiency with the increase of concentrations of HMTA shows that inhibition actions
are due to adsorption on the steel surface and adsorption follows the Langmuir isotherm. From the adsorption isotherms values of equilibrium
constant (Kads ), values of free energies of adsorption (G◦ads ), were calculated. The effect of temperature on the corrosion behaviour in the presence
of 1.0 × 10−1 M inhibitors was studied in the temperature range of 293–323 K. The inhibition efficiency of HMTA increased with increasing
temperature up to 323 K. Activation energies (Ea∗ ) were calculated from the obtained corrosion rates at different temperatures. Results obtained
from both potentiodynamic polarisation and AC impedance measurements reveal that the compound is an effective inhibitor for the corrosion of
steel and behave better in HCl than in H2 SO4 . Surface analyses were also carried out to establish the mechanism of corrosion inhibition of steel in
acidic media.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Steel; Corrosion; Hexamethylenetetramine; Electrochemical measurement

1. Introduction as functional group, steric factor, the electronic structure of the


molecule, etc. [7–16]. Lately, it has been found that low molec-
Acid solutions are commonly used in industry, the most ular mass and high water solubility amines give rise to a better
imported field of practice being acid pickling, acid cleaning, adsorption and higher protection against corrosion. Physisorp-
acid decaling and oil well acidizing [1–3]. Because of the gen- tion and chemisorption are the two types of interaction between
eral aggressiveness of acid solutions, organic compound are organic inhibitors and the metal surface. The electrostatic attrac-
widely used as corrosion inhibitors that can be adsorbed on the tion between the charge of hydrophilic groups and the charge
metal surface through donation of electrons such as ␲ electrons of active centres on the metal surface leads to physisorption.
in aromatic rings, or multiple bonds, or unpaired electrons in However, the adsorption of some ions suggests that chemisorp-
the functional groups that contain atoms such as nitrogen, sul- tions involve not only an electrostatic interaction but also a
phur, and oxygen. Adsorption usually takes place through to the partial charge transfer leading to a covalent bond formation.
vacant d-orbitals of the transition metals. Corrosion inhibitors It is believed that the covalent bond would be formed and the
block the active sites and enhance the adsorption process, thus inhibitor chemisorbed on the surface if the adsorption process
decreasing the corrosion rate and extending the life of equipment include charge transfer from the inhibitor to the metal surface.
[4–6]. The nitrogen containing compounds are known to be effi- The adsorption of inhibitors is influenced by several factors such
cient inhibitors, forming covalent bonds with metal surfaces. as the nature and surface charge of the metal, on the type of acid,
Their adsorption depends on physicochemical properties such the distribution of charge in the molecule, the type of interaction
between organic molecules and the metallic surface, which is
ruled by the dipole moment of inhibitors, and the chemical struc-
∗ Corresponding author. Tel.: +90 388 225 20 94; fax: +90 388 225 01 80. ture of organic inhibitors [17–24]. The study of HMTA as acid
E-mail address: emelbayol@nigde.edu.tr (E. Bayol). inhibitors is particularly important since they are non-expensive,

0254-0584/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2007.02.073
E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82 75

soluble in water and contains a pair of unshared electrons for Electrochemical impedance measurements were obtained at instantaneously
donation. The aim of this work is to investigate the effect of hex- measured open circuit potential (Ecorr ) applying 5 mV of amplitude, in frequency
amethylenetetramine on the corrosion of steel in various acids range from 105 to 10−3 Hz. The electrochemical impedance measurements were
carried out with a CHI604A AC electrochemical analyser.
solutions using polarisation and impedance techniques. The morphology of the corroded surface of each specimen was studied using
a Leon 440 scanning electron microscope. All micrographs of the corroded
2. Experimental specimens were taken at a magnification of 1000.

Steel containing 0.097% C, 0.00321% Pb, 0.488% Cu, 0.117% Cr, 0.032% P,
0.099% Si, 0.012% V, 0.004% Nb, 0.054% Mo, 0.07% S, 0.018% Sn, 0.01% W, 3. Results and discussion
0.0042% Co, 0.137% Ni, 0.459% Mn and the remainder iron was used for the
electrochemical measurements. The specimens were embedded in polyester, 3.1. Electrochemical impedance spectroscopy
with a surface area of 0.5 cm2 in contact with the corrosive media. Prior to
each experiment the surface of Fe specimens were mechanically polished with
different grades of emery paper (600–1200), degreased with acetone and rinsed The corrosion behaviour of steel, in 0.3 M HCl, 0.1 M H2 SO4
with distilled water and put into the cell. All the reagents used (HMTA, 37% and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl solutions with and with-
HCl and 98% H2 SO4 ) were of analytical grade acquired from Merck. out additives HMTA in the concentration range of 1.0 × 10−3 to
Polarisation experiments were carried out in a conventional three-electrode
1.0 × 10−1 M, was investigated by EIS at 293 K. Nyquist plots
cell. A platinum counter electrode with 1 cm2 surface area was also used. The
reference electrode was a Ag/AgCl with a Luggin probe positioned near the of steel in acidic solutions with and without of inhibitors dis-
electrode surface. The cell was wear-jacketed and connected to a constant played one capacitive arc which are given in Figs. 1–3 in different
temperature circulator. The test solution is de-aerated with pure nitrogen. Gas solution, respectively.
bubbling is maintained through the experiments. Electrochemical measurements In log f–log Z diagram (Figs. 1b–3b and 4b–6b), film resis-
were carried out using an EG&G scanning potentiostat model 362 potentio- tance (Rfilm ) and accumulation resistance (Ra ) value was
stat/galvanostat. The steel electrode was immersed in the solution for 30 min and
the free corrosion potential, Ecorr , was recorded. The polarisation curves were observed as plateau region, at low frequencies. This diagram
recorded using potentiodynamic technique with a constant scan rate of 1 mV s−1 . gave also another plateau region at highest frequency region,
The cathodic polarisation measurements were recorded first in a potential range which was related to pore resistance, Rpor . The sum of Rfilm
from Ecorr , to higher negative potentials. The anodic measurements were then and Ra values was equal to polarisation resistance (Rp ) value
recorded. The current densities were calculated on the basis of the apparent
[25]. The Rpor , Rfilm , Ra and Rp values were determined from
surface area of the electrode. Potentiostatic experiments were performed to test
the current response to different inhibitor injections. The corrosion rates (icorr ) log f–log Z diagrams and given in Tables 1 and 2.
before and after adding the HMTA were determined using the Tafel extrapolation To obtain the total capacitance of the circuit (C), the fre-
method. quency at which the imaginary component of the impedance is

Fig. 1. (a) Nyquist and (b) Bode plots for steel in 0.3 M HCl + X M HMTA at 298 K. X: 0 M (䊉), 1.0 × 10−3 M (), 1.0 × 10−2 M (), 4.0 × 10−2 M (), 7.0 × 10−2 M
(), 1.0 × 10−1 M (♦).

Fig. 2. (a) Nyquist and (b) Bode plots for steel in 0.1 M H2 SO4 + X M HMTA at 298 K. X: 0 (䊉), 1.0 × 10−3 M (), 1.0 × 10−2 M (), 4.0 × 10−2 M (), 7.0 × 10−2 M
(), 1.0 × 10−1 M (♦).
76 E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82

Fig. 3. (a) Nyquist and (b) Bode plots for steel in 0.1 M H2 SO4 + 1.0 × 10−3 M HCl + X M HMTA at 298 K. X: 0 M (䊉), 1.0 × 10−3 M (), 1.0 × 10−2 M (),
4.0 × 10−2 M (), 7.0 × 10−2 M (), 1.0 × 10−1 M (♦).

maximum (−Zmax  ) is calculated and C values are obtained from It is apparent from these figures that the impedance responses
the following equation [26]: for steel in acidic solutions change significantly with increas-
ing inhibitor concentration. The great effect was observed at
 1 a concentration of 1.0 × 10−1 M, which produce Rp values
f (−Zmax )= (1)
2πCRp of 1342, 1401 and 1307  in 0.3 M HCl, 0.1 M H2 SO4 and
0.1 M H2 SO4 + 1 × 10−3 M HCl, respectively (Figs. 1–3). As
The percent inhibition efficiency, IE% is calculated by polar- the inhibitor concentration increased, the Rp values increased,
isation resistance that is obtained from Nyquist plots, according but the C values tended to decrease. The decrease in C value is
to the equation below: due to the adsorption of inhibitor of the metal surface, which
  led to an increase in IE% [27–29]. The decrease in the C values
Rp − Rpo can result from a decrease in local dielectric constant and/or an
IE% = × 100 (2)
Rp increase in the thickness of the electrical double layer, which
cause the HMTA molecules adsorption on the steel surface.
where Rp and Rpo are the values of the polarisation resistance Figs. 4–6 show Nyquist diagrams for steel at the rest potential
with and without inhibitor, respectively. in the de-aerated test solutions containing 1.0 × 10−1 M HMTA

Fig. 4. (a) Nyquist and (b) Bode plots for steel in 0.3 M HCl and 0.3 M HCl + 1.0 × 10−1 M HMTA. 293 K (䊉), 303 K (), 313 K (), 323 K ().

Fig. 5. Nyquist and (b) Bode plots for steel in 0.1 M H2 SO4 and 0.1 M H2 SO4 + 1.0 × 10−1 M HMTA. 293 K (䊉), 303 K (), 313 K (), 323 K ().
E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82 77

Table 1
Impedance parameters and inhibition efficiency for the corrosion of the steel in 0.3 M HCl, 0.1 M H2 SO4 and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl without and with
addition of various concentrations of HMTA at 293 K
Medium X:[HMTA] (M) C (␮F) Rpor () Rfilm () Ra () Rp () IE%

0.0 100 5 147 89 236 –


1.0 × 10−3 130 4 497 84 581 59.4
1.0 × 10−2 80 4 789 157 946 75.1
0.3 M HCl + X M HMTA
4.0 × 10−2 64 4 994 198 1192 80.2
7.0 × 10−2 60 5 1251 – 1251 81.1
1.0 × 10−1 56 5 1168 174 1342 82.4
0.0 70 6 142 43 185 –
1.0 × 10−3 54 6 483 154 637 70.9
1.0 × 10−2 51 5 624 163 787 76.5
0.1 M H2 SO4 + X M HMTA
4.0 × 10−2 42 6 653 597 1250 85.2
7.0 × 10−2 40 7 785 493 1278 85.5
1.0 × 10−1 36 9 821 580 1401 86.8
0.0 100 13 142 45 187 –
1.0 × 10−3 68 10 436 184 620 69.8
0.1 M H2 SO4 + 1.0 × 10−3 M 1.0 × 10−2 54 7 529 257 786 76.2
HCl + X M HMTA 4.0 × 10−2 42 7 637 583 1220 84.8
7.0 × 10−2 42 10 783 464 1247 85.0
1.0 × 10−1 40 9 989 318 1307 85.7

at different temperatures. In all cases, we note a decrease of Rp H2 SO4 + 1.0 × 10−3 M HCl in absence and presence of HMTA
with increasing temperature, but the values of Rp in the presence under investigation were studied at 293 K. Fig. 7 shows that
of inhibitor are less compared to those obtained with blank assay. the potentiodynamic polarisation curves of steel in 0.3 M HCl
This result confirmed that HMTA acted as an efficient inhibitor in the absence and presence of different amounts (1.0 × 10−3
in the range of temperature studied. to 1.0 × 10−1 M) of HMTA. It is seen in Fig. 7 that the cor-
rosion current density in the region of corrosion potential and
3.2. Polarisation measurements in the range of −0.800 to −0.100 V potential, decreases as the
inhibitor concentration increases. Polarisation curves obtained
Anodic and cathodic potentiodynamic polarisation curves in 0.3 M HCl in the absence and presence of different con-
for steel in de-aerated 0.3 M HCl, 0.1 M H2 SO4 and 0.1 M centration of HMTA have showed similarity with the curves

Table 2
The influence of temperature on the impedance parameters and inhibition efficiency for the corrosion of the steel electrode immersed in 0.3 M HCl (A), 0.1 M H2 SO4
(B) and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl (C) without and with addition of 1.0 × 10−1 M of HMTA
T (K) A A + 0.1 M HMTA

C (␮F) Rpor () Rfilm () Ra () Rp () C (␮F) Rpor () Rfilm () Ra () Rp () IE%

293 100 5 147 89 236 56 5 1168 174 1342 82.4


303 104 3 57 – 57 86 4 597 – 597 90.4
313 148 4 48 – 48 60 4 265 129 394 87.8
323 226 6 22 – 22 76 4 259 – 259 91.5

T (K) B B + 0.1 M HMTA

C (␮F) Rpor () Rfilm () Ra () Rp () C (␮F) Rpor () Rfilm () Ra () Rp () IE%

293 70 6 142 43 185 36 9 821 580 1401 86.8


303 130 6 74 6 80 46 10 353 268 621 87.1
313 142 5 60 – 60 78 9 300 – 300 80.0
323 116 6 48 – 48 66 11 193 – 193 75.1

T (K) C C + 0.1 M HMTA

C (␮F) Rpor () Rfilm () Ra () Rp () C (␮F) Rpor () Rfilm () Ra () Rp () IE%

293 100 13 142 45 187 40 9 989 318 1307 85.7


303 120 5 73 16 89 54 10 479 304 783 88.6
313 138 6 60 – 60 76 10 370 – 370 83.8
323 122 5 30 – 30 80 10 194 – 194 84.5
78 E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82

Fig. 6. (a) Nyquist and (b) Bode plots for steel in 0.1 M H2 SO4 + 1.0 × 10−3 M HCl and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl + 1.0 × 10−1 M HMTA. 293 K (䊉), 303 K
(), 313 K (), 323 K ().

obtained in 0.1 M H2 SO4 and 0.1 M H2 SO4 + 1.0 × 10−3 M 1.0 × 10−1 M of HMTA. The corresponding efficiency values in
HCl. 0.3 M HCl, 0.1 M H2 SO4 and 0.1 M H2 SO4 + 1 × 10−3 M HCl
The inhibition efficiency (IE%) of corrosion of steel is cal- are 88.8%, 82.7% and 87.3%, respectively.
culated by using the following equation: In anodic field, we notice that in the presence of HMTA
  in 0.3 M HCl, there is a reduction of anodic current density
icorr − icorr(inh)
IE% = × 100 (3) (Fig. 7). It is important to note that in anodic domain, for poten-
icorr tial higher than −0.390 V, the presence of HMTA did not change
where icorr(inh) and icorr are the corrosion current density values the current versus potential characteristic. The potential can be
with and without inhibitor, respectively, determined by extrapo- defined as the desorption potential. The behaviour of HMTA
lation of the cathodic Tafel lines to the respective free corrosion at potentials greater than −0.390 V could be associated to the
potential. The values of the corrosion current density (icorr ), the significant steel dissolution leading to desorption of inhibiting
corrosion potential (Ecorr ), cathodic Tafel slopes (βc ) and the film. This dissolution results in the desorption of the adsorbed
percentage inhibition efficient (IE%) obtained as a function of film of inhibitor on the surface of the electrode in 0.3 M HCl
HMTA concentrations are given in Table 3. media. In this case the desorption rate of HMTA is raised more
It is seen in Table 3 that the increase in the concentration than its adsorption. However, HMTA influences anodic reac-
of HMTA slightly shifts the corrosion potential (Ecorr ) to pos- tion at potentials lower than −0.390 V. This fact means that
itive direction and the cathodic Tafel lines (βc ) values show a the inhibition mode of the studied HMTA depends on electrode
light modification when compared with those values obtained potential.
in the reference solutions. This result suggests that the mech- The HMTA is weakly adsorbed on steel, since the metal
anism of hydrogen reduction on the surface of steel is not surface in sulphuric acid is positively charged.
modified by the addition of HMTA [30]. As the concentration of The polarisation curves for steel in 0.3 M HCl in the
HMTA has increased, the value of corrosion current density has absence and presence of 1.0 × 10−1 M HMTA and at various
decreased too and hence the inhibition efficiency has increased, temperatures (293–323 K) are shown in Fig. 8a and b. How-
with inhibitor concentration reaching a maximum value at ever similar results were obtained in 0.1 M H2 SO4 and 0.1 M
H2 SO4 + 1.0 × 10−3 M HCl, their electrochemical parameters
are given in Table 4.
In all cases, we note an increase of corrosion current density
with increasing temperature, but the values of corrosion current
density in the presence of inhibitor are less compared to those
obtained with blank assay. The IE% value in 0.3 M HCl increases
with increasing temperature and reaches 95.2% at 323 K.
The values of activation energies (Ea∗ ) for the dissolution
of steel in all the acids studied, in the absence and presence
of the inhibitors were calculated from the Arrhenius type plot
according to the following equation:
 
E∗
icorr = A exp − a (4)
RT

where icorr is the corrosion rate in the absence and presence of


Fig. 7. Polarisation curves for steel in 0.3 M HCl and 0.3 M HCl + X M HMTA inhibitors, A the Arrhenius constant, Ea∗ the apparent activation
at 298 K. energy, R the gas constant and T is the absolute temperature.
E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82 79

Table 3
Electrochemical polarisation parameters for the steel in 0.3 M HCl, 0.1 M H2 SO4 and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl various concentrations of HMTA at 293 K
Medium X:[HMTA] (M) −Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) IE%

0 505 106 98 –
1.0 × 10−3 495 94 24 75.5
0.3 M
1.0 × 10−2 490 96 19 80.6
HCl + X M
4.0 × 10−2 492 99 15 84.7
HMTA
7.0 × 10−2 493 97 13 86.7
1.0 × 10−1 485 100 11 88.8
0 505 117 98 –
1.0 × 10−3 495 109 27 72.4
0.1 M
1.0 × 10−2 495 98 26 73.5
H2 SO4 + X M
4.0 × 10−2 505 106 24 75.5
HMTA
7.0 × 10−2 510 102 19 80.6
1.0 × 10−1 510 107 17 82.7
0 499 111 110 –
0.1 M 1.0 × 10−3 505 111 29 73.6
H2 SO4 + 1.0 × 10−3 M 1.0 × 10−2 500 98 22 80.0
HCl + X M 4.0 × 10−2 485 106 20 81.8
HMTA 7.0 × 10−2 477 104 17 84.5
1.0 × 10−1 495 107 14 87.3

Table 4
The influence of temperature on the electrochemical parameters for steel electrode immersed in 0.3 M HCl (A), 0.1 M H2 SO4 (B) and 0.1 M H2 SO4 + 1.0 × 10−3 M
HCl (C) without and with addition of 1.0 × 10−1 M of HMTA
T (K) A A + 1.0 × 10−1 M HMTA

−Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) −Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) IE%

293 505 106 98 485 98 11 88.5


303 505 123 277 485 104 26 90.4
313 500 145 446 480 120 35 92.3
323 500 191 1114 545 180 54 95.2

T (K) B B + 1.0 × 10−1 M HMTA

−Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) −Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) IE%

293 505 117 98 500 107 17 84.5


303 512 134 184 505 115 30 84.2
313 509 149 336 540 122 38 88.8
323 505 160 735 595 149 91 92.1

T (K) C C + 1.0 × 10−1 M HMTA


−Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) −Ecorr (mV) −βc (mV dec−1 ) icorr (␮A cm−2 ) IE%

293 499 117 110 495 105 13 86.5


303 510 131 238 520 109 30 86.8
313 510 133 538 560 124 40 90.3
323 510 192 1151 590 185 86 92.7

The plot of ln(icorr ) for steel obtained from the polarisation Table 5
curves as a function of (1/T) gives straight lines with a slope The values of activation parameters Ea∗ for steel in 0.3 M HCl (A), 0.1 M H2 SO4
of −Ea∗ /R (Fig. 9) and the Ea∗ obtained from these graphs are (B) and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl (C) without and with addition of
1.0 × 10−1 M of HMTA
given in Table 5.
The inhibition efficiency of HMTA increases with increas- Compound Ea∗ (kJ mol−1 )
ing temperature and its increase leads to a decrease in the A 61.21
apparent activation energy. The results show that presence of B 52.24
HMTA, gives maximum efficiency, by lowering the activation C 61.87
energy. It is considered to chemisorption on the steel sur- A + 1.0 × 10−1 M HMTA 40.13
B + 1.0 × 10−1 M HMTA 41.32
face [31,32]. The value of apparent activation energy Ea∗ of C + 1.0 × 10−1 M HMTA 46.92
hydrogen evolution reaction in 0.3 M HCl without an inhibitor
80 E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82

pound was discussed in terms of blocking the electrode surface


by adsorption of the molecules through the active centres con-
tained in its structure. The values of surface coverage θ for
different concentrations at 293 K have been used to explain the
best isotherm that determines the adsorption process.
Attempts were made to fit these θ values to various
isotherms including Langmuir, Frumkin, Freundlich and Temkin
isotherms. However the best fit was obtained from the Langmuir
isotherm. This has been also used for other inhibitor systems
[35]. Langmuir isotherm is as follows:
C 1
= +C (5)
θ K
where θ is the degree of coverage on the metal surface, K the
equilibrium constant of the adsorption reaction and C is the
concentration of inhibitor.
The plots of Cinh /θ versus Cinh yield a straight line with corre-
lation coefficient higher than 0.95, showing that the adsorption
of this inhibitor can be fitted to a Langmuir adsorption isotherm
(Fig. 10). This indicates that the inhibitor molecules are adsorbed
on the metal surface and form a barrier, which isolated the metal
surface from the electrolyte. Langmuir adsorption isotherm may
be explained on the basis of interaction between the adsorbed
species on the metal surface [2,36].
The value of the adsorption coefficient (Kads ) for steel in pres-
ence of HMTA is calculated as 909, 556 and 769 in 0.3 M HCl,
0.1 M H2 SO4 and 0.1 M H2 SO4 + 1.0 × 10−3 M HCl, respec-
tively. Results show that HMTA give high values of K in 0.3 M
HCl, indicating that it was adsorbed strongly on the steel sur-
face, thus, the inhibition efficiency increased with increasing
adsorption coefficient.
Fig. 8. Effect of temperature on polarisation curves for steel in (a) 0.3 M HCl The equilibrium constant for adsorption process is related
and (b) 0.3 M HCl + 1.0 × 10−1 M HMTA.
to the free energy of adsorption, Gads , and is expressed by
following equation:
(61.21 kJ mol−1 ) (Table 5) agrees well with the literature data  
on Ea∗ for steel and steel in hydrochloric acid, which are in the 1 Gads
Kads = exp − (6)
range 58–100 kJ mol−1 [33,34]. 55.5 RT
It was found that, the inhibition efficiency of this compound where 55.5 is the molar concentration of water in the solu-
depends on its concentration. The inhibitive action of this com- tion. The value of free energy of adsorption, Gads , for steel

Fig. 9. The relation between ln icorr vs. 1/T for steel in 0.3 M HCl and 0.3 M Fig. 10. Langmuir adsorption isotherm of HMTA on steel in 0.3 M HCl, 0.1 M
HCl solution containing 1.0 × 10−1 M HMTA. H2 SO4 , 0.1 M H2 SO4 + 1.0 × 10−3 M HCl solutions.
E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82 81

Fig. 11. SEM photograph of the surface for steel after 24 h of immersion period (a) before corrosion, (b) 0.3 M HCl, (c) 0.3 M HCl + 1.0 × 10−1 M HMTA, (d) 0.1 M
H2 SO4 , (e) 0.1 M H2 SO4 + 1.0 × 10−1 M HMTA, (f) 0.1 M H2 SO4 + 1.0 × 10−3 M HCl and (g) 0.1 M H2 SO4 + 1.0 × 10−3 M HCl + 1.0 × 10−1 M HMTA.

in presence of HMTA in 0.3 M HCl, 0.1 M H2 SO4 and 0.1 M Gads also suggest the monolayer adsorption of the inhibitor
H2 SO4 + 1 × 10−3 M HCl was calculated as −26.39, −25.19 molecules with the steel surface [37]. Generally, G◦ads val-
and −25.99 kJ mol−1 , respectively. ues up to −20 kJ mol−1 are consistent with the electrostatic
The low and negative value of Gads indicates the spon- interaction between the charged molecules and the charged
taneous adsorption of inhibitors on the surface of steel and metal (physical adsorption), while those that are more nega-
82 E. Bayol et al. / Materials Chemistry and Physics 104 (2007) 74–82

tive than −40 kJ mol−1 involve charge sharing or transfer from 6. The investigated compound was more effective in HCl than
the inhibitor molecules to the metal surface to form a coordinate in H2 SO4 .
type of bond (chemisorption) [38,39].
In the present study, HMTA was found to be effective for References
the inhibition of steel corrosion in 0.3 M HCl. It is well known
that chloride ions have a smaller degree of hydration; they [1] D. Chebabe, Z. Ait Chikh, N. Hajjaji, A. Srhiri, F. Zucchi, Corros. Sci. 45
(2003) 309.
could bring excess negative charges in the vicinity of the inter- [2] K.F. Khaled, N. Hackerman, Mater. Chem. Phys. 82 (2003) 949.
face, which facilitates positively charged ions adsorption onto [3] M.A. Quraishi, R. Sardar, Mater. Chem. Phys. 78 (2002) 425.
the surface. HMTA might be protonated in the acid solution [4] H. Ashassi-Sorkhabi, M.R. Majidi, K. Seyyedi, Appl. Surf. Sci. 225 (2004)
and the protonated HMTA are attached to the steel surface by 176.
means of electrostatic interaction between Cl− and the proto- [5] H. Wang, R. Liu, J. Xin, Corros. Sci. 46 (2004) 2455.
[6] S. Zor, B. Yazıcı, M. Erbil, Corros. Sci. 47 (2005) 2700.
nated HMTA. When HMTA molecules are adsorbed onto the [7] S. Martinez, I. Stern, Appl. Surf. Sci. 199 (2002) 83.
metal surface, coordinate bonds have formed by partial transfer [8] M. Bouklah, A. Attayibat, B. Hammouti, A. Ramdani, S. Radi, M. Benkad-
of electrons from the polar atom of HMTA to the metal surface. dour, Appl. Surf. Sci. 240 (2005) 341.
HMTA is more effective in HCl than in H2 SO4 media, this can [9] M. Düdükçü, B. Yazıcı, M. Erbil, Mater. Chem. Phys. 87 (2004) 138.
be explained by the fact that chloride ions being less hydrated [10] J. Cruz, R. Martinez, J. Genesca, E. Garcia-Ochoa, J. Electranal. Chem.
566 (2004) 111.
than sulphate ions and therefore more strongly adsorbed on the [11] E.S. Ferreria, C. Giacomelli, F.C. Giacomelli, A. Spinelli, Mater. Chem.
metal surface by creating an excess negative charge towards the Phys. 83 (2004) 129.
solution phase, which favours inhibitor adsorption on the steel [12] T. Tüken, B. Yazıcı, M. Erbil, Türk. J. Chem. 26 (2002) 735.
surface. [13] F.I. Halilova, R.M. Aliguliyev, Z.M.O. Rzaev, Anti-Corros. Meth. Mater.
48 (2001) 18.
[14] S. Bilgiç, N. Çalışkan, Appl. Surf. Sci. 152 (1999) 107.
3.3. Morphological examination [15] K. Tebbji, A. Aouniti, M. Benkaddour, H. Oudda, I. Bouabdallah, B. Ham-
mouti, A. Ramdani, Prog. Org. Coat. 54 (2005) 170.
A SEM image of polished steel sample is shown in Fig. 11a. [16] Y. Abboud, A. Abourriche, T. Saffaj, M. Berrada, M. Charrouf, A. Benna-
mara, A. Cherqaoui, D. Takky, Appl. Surf. Sci. 252 (2006) 8178.
The micrograph shows a characteristic inclusion, which was
[17] B. El Mehdi, B. Mernari, M. Traisnel, F. Bentiss, M. Lagrenée, Mater.
probably an oxide inclusion. Fig. 11b, d and f shows SEM image Chem. Phys. 77 (2002) 489.
of the surface of the steel specimens after immersion in acidic [18] H. Luo, Y.O. Guan, K.N. Han, Corrosion 54 (1998) 619.
solutions for 24 h. The micrograph reveals that, the surface was [19] L.M. Vračar, D.M. Dražič, Corros. Sci. 44 (2002) 1669.
strongly damaged in absence of inhibitors. Fig. 11c, e and g [20] F. Bentiss, M. Lagrenee, M. Traisnel, J.C. Hornez, Corros. Sci. Sect. 41
(1999) 789.
shows SEM image of the surface of another steels specimen
[21] A.Y. El-Etre, Corros. Sci. 45 (2003) 2485.
after immersion for the same time interval in all acidic solu- [22] M. Abdallah, Corros. Sci. 46 (2004) 1981.
tions containing 1.0 × 10−1 M HMTA. The micrograph reveals [23] H. Ashassi-Sorkhabi, Z. Ghasemi, D. Seifzadeh, Appl. Surf. Sci. 249 (2005)
that, there is a good protective film present on the steel surface. 408.
This confirms the highest inhibition efficiency of the HMTA at [24] F. Touhami, A. Aouniti, Y. Abed, B. Hammouti, S. Kertit, A. Ramdani, K.
1.0 × 10−1 M concentration. Elkacemi, Corros. Sci. 42 (2000) 929.
[25] T. Tüken, B. Yazıcı, M. Erbil, Appl. Surf. Sci. 252 (2006) 2311.
[26] M. Özcan, İ. Dehri, M. Erbil, Appl. Surf. Sci. 236 (2004) 155.
4. Conclusions [27] S. Rames, S. Rajeswari, Electrochim. Acta 49 (2004) 811.
[28] E. Chaieb, A. Bouyanzer, B. Hammouti, M. Benkaddour, Appl. Surf. Sci.
The main conclusions of this study are given below: 246 (2005) 199.
[29] A. Dadgarnezhad, I. Sheikhshoaie, F. Baghaei, Anti-Corros. Meth. Mater.
51 (2004) 266.
1. The HMTA actually inhibits steel corrosion in the tested acid [30] A. Chetouani, A. Aouniti, B. Hammouti, N. Benchat, T. Benhadda, S.
media. Kertit, Corros. Sci. 45 (2003) 1675.
2. Inhibition efficiency increases with an increase in inhibitor [31] F. Bentiss, M. Traisnel, N. Chaibi, B. Mernari, H. Vezin, M. Lagrenée,
concentration and rise in temperature. Corros. Sci. 44 (2002) 2271.
[32] M.A. Quraishi, F.A. Ansari, D. Jamal, Mater. Chem. Phys. 77 (2002) 687.
3. The HMTA inhibit the corrosion by controlling both the [33] S.A. Ali, A.M. El-Shareef, R.F. Al-Ghamdi, M.T. Saeed, Corros. Sci. 47
anodic and cathodic in 0.1 M H2 SO4 and by controlling (2005) 2659.
anodic in 0.3 M HCl. [34] E.A. Noor, Corros. Sci. 47 (2005) 33.
4. The adsorption of the investigated compound from 0.3 M [35] L. Tang, G. Mu, G. Liu, Corros. Sci. 45 (2003) 2251.
HCl, 0.1 M H2 SO4 and 0.1 M H2 SO4 + 1 × 10−3 M HCl on [36] M.H. Wahdan, A.A. Hermas, M.S. Morad, Mater. Chem. Phys. 76 (2002)
111.
the steel surface obeys a Langmuir adsorption isotherm. [37] H. Ashassi-Sorkhabi, B. Shaabani, D. Seifzadeh, Appl. Surf. Sci. 239
5. The negative values of free energy of adsorption, G◦ads indi- (2005) 154.
cate the spontaneous adsorption of HMTA on the surface of [38] G. Moretti, F. Guidi, G. Grion, Corros. Sci. 46 (2004) 387.
steel. [39] M.S. Morad, A.M. Kamal El-Dean, Corros. Sci. 48 (2006) 3398.

View publication stats

You might also like