Sie sind auf Seite 1von 9

Electrochimica Acta 54 (2009) 21712179

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Effects of surfactants and their mixtures on inhibition of the corrosion


process of ferritic stainless steel
R. Fuchs-Godec
Faculty of Chemistry and Chemical Engineering, University of Maribor, Smetanova 17, 2000 Maribor, Slovenia

a r t i c l e

i n f o

Article history:
Received 21 July 2008
Received in revised form 10 October 2008
Accepted 10 October 2008
Available online 18 October 2008
Keywords:
Corrosion inhibitor
Mixtures
Cationic surfactant
Zwitterionic surfactant
FloryHuggins isotherm
Sulphuric acid

a b s t r a c t
The corrosion inhibition characteristics regarding mixtures of cationic/zwitterionic types of surfactant
(Myristyltrimethylammonium bromide/Palmitylsulfo-betaineas), and non-ionic surfactant TRITON-X405 mixed with 1 mM of KBr, as corrosion inhibitors for stainless steel (SS) (type X4Cr13) in aqueous
solutions of 2 M H2 SO4 were investigated using potentiodynamic polarisation measurements. The polarisation data showed that mixtures of the surfactants used in this study acted as mixed-type inhibitors,
adsorbing on the stainless steel surface in agreement with the FloryHuggins adsorption isotherm. The
tensiometric results of this study suggest the existence of a second state of aggregation for zwitterionic/cationic surfactant mixtures. From these values of the free energy of adsorption, which in both
mixtures decreased with respect to a single surfactant, we concluded that the adsorption in mixtures
was stronger. The mixtures studied here showed good inhibition properties for ferritic stainless steel type
X4Cr13 in 2 M H2 SO4 solution.
2008 Elsevier Ltd. All rights reserved.

1. Introduction
Surfactants are compounds that can be found in a multitude of
domains, from industrial settings to research laboratories and are
the part of our daily lives. Due to their unique structure they can
drastically modify the interfacial properties. This effect is important for industrial processes such as otation, the cosmetic and food
industries, drugs delivery, emulsication, chemical mechanical polishing, as also for corrosion inhibition. The latter topic is also the
subject of this paper.
It is well known that corrosion never stops but its scope and
severity can be lessened. The addition of surfactants into aggressive media such as acid solutions is one of the methods for achieving
this goal. Several studies suggested that most organic inhibitors are
adsorbed on the metals surface, displacing water molecules from
the surface and forming a compact lm as a barrier [1,2]. In other
words, the adsorbed corrosion inhibitor may be sterically blocking the surface, thereby, either restricting the access of corrosion
species to the surface or transferring the corrosion product from it.
The organization and packing of the inhibitor molecules on the surface may be important, as well as the coverage fraction. The ability
of a surfactant molecule to adsorb on the metal surface was found
to be responsible for the corrosion inhibition of the metal surface.

Fax: +386 2 2527 774.


E-mail address: fuchs@uni-mb.si.
0013-4686/$ see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2008.10.014

In order to be effective, the inhibitor molecule must displace water


from the metal surface, interact with anodic or cathodic reaction
sites to retard oxidation and reduction corrosion reaction, and prevent the transport of water and other corrosion-active species to
the surface. Several studies have been presented regarding the corrosion of various types of steel and their inhibition by different
types of organic inhibitors in acid solution [39]. Both ionic and
non-ionic surfactants have been reported to inhibit the corrosion
of metals such as copper, aluminium, and mild-steel [1014].
If a single surfactant molecule possesses both anionic and
cationic functional groups, it is called amphoteric or zwitterionic.
In most cases it is the pH that determines, which group dominates,
thus favouring one ionisation or another. This can be anionic at
alkaline pH and cationic at acid pH values, with an amphoteric
behaviour at intermediate pH values [15,16]. Mixtures of surfactants have found many applications in technology. These mixtures
often show synergistic effects, which are evidence of non-ideal
behaviour and are the reason for their extensive use in industry. An
objective of the present work is to study the applicability of nonionic surfactant mixtures and the zwitterionic type of surfactants,
as corrosion inhibitors for stainless steel (SS), type X4Cr13 in aqueous solutions of 2 mol L1 H2 SO4 (mol L1 = M). In the case of the
non-ionic surfactant mixture, we used the ethoxylated octyl phenyl
alcohol (under the trade name TRITON-X series), known as TRITONX-405 with the chemical structure C8 H17 C6 H4 (OCH2 CH2 )40 OH
with the addition of 1 mM of KBr. Part of the present study is
actually an extension of our previous work [17], in which we inves-

2172

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

tigated the inhibition abilities of two non-ionic surfactants from


the TRITON-X series, namely TRITON-X-405 and TRITON-X-100.
Conclusions obtained from these electrochemical measurements
[17] indicated the difculties in organization, and the packing of
inhibitor molecules of type TRITON-X-405 on the metal surface.
One reason for these difculties is probably the surfactant size,
the number of polar ethylene oxide groups of TRITON-X-100 being
between 9 and 10, and for TRITON-X-405 around 40. Consequently,
the inhibiting lm formed during cathodic polarisation was probably more porous and of lower grade, as was in the case when
presenting TRITON-X-100 in the solution. Therefore the dynamics of the anodic process could have a more destructive effect
on the self-assembling layer. For zwitterionic surfactant mixtures,
we studied the system 3-N-dimethylpalmitylammoniopropane
sulfonate or palmitylsulfo-betaineas (zwitterionic surfactant,
ZWITTER C16 ), and Myristyltrimethylammonium bromide (MTABr)
as the ionic component in solutions of sulphuric acid. The (3-Ndimethylpalmitylammoniopropane sulfonate) sulfobetaines were
chosen since these surfactants are insensitive to the pH of the system. They remain as real zwitterionic surfactant for all pH values
of the solution. This is the reason why they are the most commonly
used class of amphoteric surfactants; they can be found in textile
softeners, hair rinse formulas and as corrosion inhibition additives.

2. Experimental
2.1. Materials
The zwitterionic (SigmaAldrich, CAS Number 2281-11-0) and
the cationic surfactant (Fluka, CAS Number 1119-97-7) used in this
study were of pure quality (>97%) and used without purication.
The non-ionic surfactant TRITON-X-405 (with activity of 70 wt%)
was a Fluka product (CAS Number 9002-93-2) and also used without purication. The surfactants had the following structures:

All the solutions were prepared using water obtained from a


Millipore Super-Q system. The experimental concentrations were
within the range 7.0 107 to 3.5 104 mol L1 for TRITON-X-405,
and in mixtures where 1 mM of KBr was added, the concentration
range being between 7.0 107 and 7.0 104 mol L1 . In the case

Table 1
The composition of ferritic stainless steel type X4Cr13 in wt%.
wt%
C
Si
S
Cr
Ni
Cu

0.04
0.471
0.02
13.20
0.307
0.213

of zwitterionic/cationic surfactant mixtures we studied two systems, where the concentration of zwitterionic surfactant was xed
at c = 1.0 105 and at 1.0 104 mol L1 , while the concentration
of added MTABr was changing from 1.0 105 to 1.0 103 mol L1
Cylindrically shaped specimens were made from a rod of ferritic
stainless steel of type X4Cr13. The composition of used steel in wt%
is shown in Table 1
2.2. Surface tension measurements
The surface tension () in 2 mol L1 H2 SO4 was measured by the
Wilhelmy-plate method, using a Krss-K12 processor tensiometer
at 25 C. The temperature was kept constant (0.1 C) by circulating the water through a jacketed vessel containing the solution.
The solutions concentration was varied by adding aliquots of stock
solution of known concentration, to the known volume of the solution in the vessel. The plate was cleaned over a methanbutan
ame.
2.3. Electrochemical measurements
Electrochemical experiments were performed with a conventional three-electrode conguration. All the potentials were
measured against a saturated calomel electrode (SCE) and the

counter electrode was made from Pt. In all experiments, electrochemical polarisation was started 30 min after the working
electrode was immersed in solution, to allow the stabilization of
the stationary potential. Before each measurement, the sample was
cathodically polarized at 1.0 V (SCE) for 10 min (electrochemical-

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

2173

Fig. 1. Variations in surface tension regarding concentrations for (a) TRITON-X-405 in water, (b) TRITON-X-405 in 2.0 M H2 SO4 and (c) TRITON-X-405 + 1 mM KBr in 2.0 M
H2 SO4 at 25 C.

cathodic cleaning by H2 evolution. In this way the hydrogen bubbles


mechanically remove the remainders of the hand polishing treatment) and then allowed to reach a stable open-circuit potential,
which was attained after about 30 min. The potentiodynamic current potential curves were recorded by automatically changing the
electrode potential from 0.7 V to 0.9 V (SCE) at a scanning rate
of 2 mV s1 . All the experiments were performed at 25 C 1 C. A
SOLATRON 1287 Electrochemical Interface was used to apply and
control the potential. The data were collected using CorrWare and
interpreted using the CorrView software. This software was developed by Scribner Associates, Inc. The working electrode was ferritic
stainless steel of type X4Cr13. The test specimens were xed in a
PTFE holder, and the geometric area of the electrode exposed to
electrolyte was 0.785 cm2 . The metal surface was hand polished
successively with emery papers of grades 400, 600, 800, 1000 and
1200. Next, the specimen was ne polished with diamond paste
to obtain a mirror-like shined surface. After polishing, the working electrode was washed with ethanol, rinsed several times with
distilled water, and nally dried using hot air.
3. Results and discussions
3.1. Determination of the critical micelle concentration of
surfactants and mixtures in 2 M H2 SO4
The results for surface tension were plotted as a function of
the surfactant concentrations logarithm and the critical micelle
concentration (cmc) was estimated from the break-point in the
resulting curve. Representative plots of surface tension vs. the
decade logarithm of surfactant concentration log10 C are shown in
Fig. 1ac for TRITON-X-405 with and without the addition of 1 mM
of KBr. Fig. 2ad represents the same relationship valid for zwitte-

rionic/cationic surfactant mixtures and, for a better illustration of


the effect, the corresponding plots for single surfactants are also
shown. At least three independent measurements were done for
each system, to check for the reproducibility of the surface tension
curves.
The addition of electrolytes, organic compounds or different
types of surfactants, is known to modify the properties of surfactants aqueous solutions; this includes their solubility, the cmc and
aggregation numbers. Therefore the additives may equally have an
effect on adsorption at the solid/liquid interface. Using the additives, the inhibition properties of non-ionic surfactants could be
improved or worsened. These phenomena are related to the famous
Hofmeister series, which is an empirical measure of ions hydration
degree. The Hofmeister series orders ions with increased salting-in
potency from left to right, and is as follows:
SO4 2 , HPO4 2 , OH , F , HCOO , CH3 COO , Cl , Br , NO3 , I ,
SCN , ClO4
Ions on the left of Cl represent in some way a borderline. They
are called salting-out ions (structure-makers) or cosmotropic ions.
The opposite holds true for the ions on the right of Cl , which
are known as salting-in ions (structure-breakers) or chaotropic
ions. This effect usually leads to an increase in the solubility of
amphiphiles [1823]. The salting-out effect is produced when the
salt, competing with the surfactant for hydration water, reduces the
amount of water available in the micelles for polar chain hydration. Thus, micelle formation will be produced at lower surfactant
concentration. The salting-in effect (micelle hydration) is produced
because the salt ions break the water structure (intermolecular interactions), making the water molecules more accessible for
hydration of the surfactant molecules [24].

2174

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

Fig. 2. Variations in surface tension regarding concentrations for (a) palmitylsulfo-betaineas (ZWITTER C16 ), (b) MTABr, (c) ZWITTER C16 c = 1.0 105 M + MTABr and (d)
ZWITTER C16 c = 1.0 104 M + MTABr all in 2.0 M H2 SO4 at 25 C.

Since the interaction between the ionic species and non-ionic


surfactants is weak, the effect of inorganic salts on the properties
of non-ionic surfactants is usually small. However, a large amount
of inorganic salt would affect the solutions properties through their
inuence on the water properties. It is well known that the salts,
added in high concentration, either depress or elevate the cloud
points and cmcs of the non-ionic surfactants, depending on the
salting-out or salting-in nature of the salt species [20].
In the case of TRITON-X-405 in 2 M H2 SO4 , the salt effect
induced by an addition of KBr has been interpreted in terms of
salt-induced reduction or the enhancement of the ethylene oxide
units hydration regarding the surfactant molecules in aqueous
media. Notice that the number of these units is around 40 for
TRITON-X-405. Cosmotropic anions as SO4 2 , producing high electric elds at short distances, bind their water molecules strongly
and compete efciently for water with the ethylene oxide units. This
phenomenon leads to the dehydration of the surfactant molecules
hydrophilic part, an effect which depresses the cmc value of
the non-ionic and increases attractive interactions between these
micelles.
The opposite effect can be achieved with chaotropic ions Br
(TRITON-X-405 + 1 mM KBr in 2.0 M H2 SO4 ), which disrupt the
icebergs of water molecules in bulk solution and increase the
concentration of free water molecules, which in turn may form
hydrogen bonds with the ethylene oxide units of the non-ionic surfactant. Thus, the chaotropic ions act to promote the hydration of
the hydrophilic part of the TRITON-X-405, which possibly becomes
one of the contributions to the elevation of cmc value, caused by
the addition of the 1 mM KBr. Fig. 1 clearly shows the phenomenon
mentioned above. For TRITON-X-405 in 2 M H2 SO4 mixtures, up to

one order of magnitude higher values for the cmc were observed,
with respect to the KBr salt free solution (Fig. 1b and c).
The next subject of our investigation was the system containing zwitterionic/cationic surfactant mixture. The concentration of
the zwitterionic surfactant was kept constant while we varied the
concentration of cationic surfactant. It can be noticed in Fig. 2, that
the cmc of a surfactant when alone is signicantly lower (at least
one order of magnitude) compared to the situation when it is in a
mixture. On the other hand, it is evident from Fig. 2c and d that,
in addition to a normal break-point, another break appeared in
the curve. Two different stages of aggregation have been observed
before [2529].
Velazquez and co-workers [30] reported that aqueous solutions
of the cationic DTAB and the zwitterionic surfactants, mix ideally
and that any interaction is reduced as the hydrocarbon length of the
zwitterionic surfactant increases. They emphasized their observation [28] that, in addition to electrostatic attraction, ion-specic
interactions can also be important. The differences in interaction
between anionic surfactants with zwitterionic and cationic surfactants can be interpreted on the basis location of the charged groups
[31]. In zwitterionic sulfobetaines, the positive charge is close to the
micelle core, therefore, in an anionic/zwitterionic mixed micelle,
the negative charge of the anionic surfactant is close to the positive charge of the dipole. In orders to achieve the closest proximity
of opposite charges in the cationic/zwitterionic micelles, the tails
of the cationic surfactants would have to protrude from the core,
making this way the micellization process less favourable.
The tensiometric results discussed in this study seem to suggest the existence of a second aggregation state in the mixed
surfactant systems. The cmcs of mixtures composed from zwit-

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

2175

Fig. 3. Potentiodynamic polarisation curves (2 mV s1 ) for TRITON-X-405 with and without the addition of 1 mM of KBr on SS type X4Cr13 in 2.0 M H2 SO4 .

terionic surfactant C16 of xed concentration; c = 1.0 105


M + MTABr were determined to be at a concentration of
added MTABr equal to: cmc1 = 2.0 0.2 104 mol L1 and
cmc2 = 4.6 0.2 104 mol L1 of added MTABr. In the case
where the concentration of zwitterionic surfactant C16 was one
order of magnitude higher c = 1.0 104 M + MTABr, the cmcs
of the mixture were be cmc1 = 1.2 0.02 104 mol L1 and
cmc2 = 1.6 0.02 104 mol L1 of added MTABr.
An interpretation of the micellization process is possible as follows. In both surfactants the hydrophilic head is a quaternized
nitrogen atom having, a positive charge. We assume, therefore, that
the driving-force for the micellization in the zwitterionic/cationic
mixtures is competition between the repulsion of the hydrophilic
part and the hydrophobic (via water) interaction of alkyl chains as
an attractive interaction between the constituent surfactants in the
mixed micelle.
When the concentration of alkyl chains located inside the
mixed micelle is sufciently high, attractive hydrophobic interaction predominated and normal spherical mixed micelles were
formed. This was conrmed by the fact that cmc1 had, in
both cases, approximately the same values for the total concentrations of both surfactants (zwitterionic + cationic) in the
solution (i.e. 2.12.2 104 mol L1 ). A noticeable difference was
observed at cmc2 , where the total concentration of surfactants was approximately twice as high, in contrast to the case
where the xed concentration of zwitterionic surfactant was at
c = 1.0 104 mol L1 . This phenomenon could be interpreted as
higher concentrations for ionic surfactants being needed, in order
to overcome electrostatic repulsion between ionic head groups
during the aggregation process. Further, more highly ordered structures of the adsorbed aggregates could be achieved by further
adsorption from the solution (the step from cmc1 to cmc2 ), as are
co-adsorption structural transformations, or rearrangement, inside
the formerly existing micelles.
A study of some relevance for our work has recently been
published [32]. In their molecular dynamics study of a short
polyelectrolyte with quaternized ammonium ion (ionene) Vlachy
and coworkers [32], divided the counter ions to adsorbed if
found within certain arbitrary distance from the nitrogen atom,
and free otherwise. The results of this analysis indicated that,
when approaching an ionene counterions such as Cl , Br and
I might loose one or two water molecules. In our case, when a
mixture with a xed concentration of zwitterionic surfactant of
c = 1.0 105 mol L1 was used, the micelles at cmc1 were mostly
composed of the cationic surfactant MTABr. We can speculate,
therefore, that something similar can happen, as documented
in [32]. At a sufciently high concentration of MTABr, the bromide counterion can approach the quaternized nitrogen atom of

the zwitterionic surfactant close enough to release some water.


This rearrangement may yield further adsorption from the solution (co-adsorption), which is eventually concluded at cmc2 ,
leading to more-ordered structures. The shapes of the micelles
formed this way cannot be completely spherical anymore; but
perhaps somewhere between spherical and cylindrical. In the
other case, when the concentration of zwitterionic surfactant was
c = 1.0 104 mol L1 , the amount for both surfactants at cmc1 was
approximately 1:1. In the micelle core, therefore, an alternative
arrangement of surfactants A and B is expected (pattern ABAB. . .).
Electrostatic repulsion between the ionic head groups of MTABr
was lower because of a higher concentration of the zwitterionic
surfactant being present, which thus reduced ionic charge interaction. For this reason, the cmc2 of such a mixture is attained at
a considerably lower concentration of added cationic surfactant
(cmc2 = 1.6 104 mol L1 of added MTABr). The alky chain of the
zwitterionic surfactant was slightly longer than the alky chain of
the cationic surfactant. Thus, it could be suggested that the mixed
micellar aggregates formed in this system, have hydrodynamic radii
higher than those formed in the rst mixture. The SO3 H group
in the zwitterionic surfactant is believed to be strongly hydrated,
therefore, we assumed that the bromide counter-ion would loose
more water than in the rst case, when the xed concentration of
zwitterionic surfactant was equal to c = 1.0 105 mol L1 .
3.2. Electrochemical results
The effect of TRITON-X-405 (in presence and absence of 1 mM
of KBr) on the currentpotential characteristics displayed by the
polarisation curves of the ferritic stainless steel type X4Cr13 in
2 M H2 SO4 , is presented in Fig. 3. Voltammogram exhibited a
signicant decrease in anodic current peaks in 2.0 M H2 SO4 . Consequently, the charge under the oxidation curves was progressively
reduced, even at very low concentrations of added surfactant. This
phenomenon was observed at c = 3.5 106 mol L1 . Moreover, a
substantial decrease in the cathodic current density was observed
at this low concentration of added surfactant. It is necessary to
mention that further increase in surfactant concentrations (up to
two orders of magnitude higher than the initial value) had little
effect on inhibition. Further, decreasing of anodic current peaks
was observed with the addition of KBr in a concentration of 1 mM.
It is necessary to mention that with addition of the surfactant, the proles on voltammogram (Fig. 3) obtain at proles.
The at portion extends to about 100 mV in both directions, with
regard to the corrosion potential of the 2 M H2 SO4 solution in the
absence of inhibitors (0.461 V vs. SCE). The best inhibition efciency is expected within the potential region where the current is
at minimum. The at portion was additionally extended by about

2176

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

50 mV, but only in regard to the anodic direction, with the addition of bromide ions (KBr) to TRITON-X-405 in 2 M H2 SO4 solution.
As mentioned previously, the TRITON-X-405 has difculty when
organizing and packing on the metal surface (because of its size,
probably). In our previous study it was conrmed, that desorption
of the protective layer formed in the presence of TRITON-X-405
started sooner, as in the case of TRITON-X-100 [17]. The free
species, which appeared on the metal surface between monomers
or micelles, also, represent those active points which initialised the
corrosion process with a destructive effect on the self-assembling
layer formed during cathodic polarisation.
Therefore the improvement in inhibition efciency (IE) was
achieved, with the addition of bromide ions (KBr) in the chosen
solution. We can assume that the bromide ions additionally inhibit
the corrosion by getting adsorbed electrostatically on the positively
charged surface in the anodic region, or with the help of synergistic
effect between KBr and TRITON-X-405 (with displacement of more
water molecules from the metal surface than in the case, when
TRITON-X-405 was individually presented in the solution).
An increase in IE was also obtained in the cases of mixtures
composed of zwitterionic surfactant and cationic surfactant in 2 M
H2 SO4 solutions in comparison with the example where zwitterionic surfactants was alone in sulphuric acid solutions. Fig. 4
illustrates those of inhibitions properties on corrosion parameters
at various concentrations of both surfactants in mixtures. The same
effect occurred in the case of TRITON-X-405 + 1 mM KBr mixture i.e.
extension of the at portion in the anodic direction but only, when
only the zwitterionic surfactant was present in the acid solution.
This phenomenon was observed at exactly the same concentration
of added bromide ions (cMTABr = 1.0 103 M, cKBr = 1.0 103 M).
It could be speculated that bromide ions are responsible for the
positive synergism and for improvement of inhibition efciency.
The electrochemical parameters obtained from polarisation curves,
corrosion potential (Ecorr ), corrosion current density (icorr ), the
polarisation resistance (Rp ) and the inhibition efciency, are shown
in Table 2. The polarisation resistance was obtained by linear polarisation within the potential range of 10 mV with respect to Ecorr .
Extrapolation of the Tafel line allowed for calculation of the corrosion current density (icorr ). These parameters were determined
simultaneously (CorrView software).
It is clear, from the obtained corrosion data, that Tafel lines
shifted towards more negative and more positive potentials during the anodic and cathodic processes, respectively, relative to the
blank curve. This means that the selected compound acts as mixedtype inhibitor. Next, a shift of Ecorr was observed towards a more
noble value, and to an order of magnitude for lower corrosion
current density icorr was attained. Moreover, to an order and half
of magnitude lower icorr for chosen mixtures with respect to the
inhibitor-free solution was observed (Figs. 5 and 6).

Fig. 4. Potentiodynamic polarisation curves (2 mV s1 ) for palmitylsulfo-betaineas


(ZWITTER C16 ) with and without the addition of x M MTABr on SS type X4Cr13 in
2.0 M H2 SO4 .

3.3. Adsorption isotherms


The surface coverage, , as well as, the polarisation resistance
Rp , and the corrosion current density icorr , were calculated from
the kinetic parameters measured in corrosion processes by using
the following equations:
 =1


icorr
icorr

IE (%) = 1
 =1

Rp
Rp

(1)

icorr
icorr

100

(2)

(3)

Fig. 5. Inuence of added TRITON-X-405 with and without the addition of 1 mM of


KBr on the cathodic and anodic behaviour of stainless steel type X4Cr13 in 2.0 M
H2 SO4 .

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

2177

Table 2
Kinetic parameters for corrosion of stainless steel type X4Cr13 obtained from
potentiodynamic polarisation curves in 2.0 M H2 SO4 for TRITON-X-405, TRITON-X405 + 1 mM of KBr, for palmitylsulfo-betaineas with and without added x M MTABr
at 25 C.
Ecorr (V vs. SCE)

Rp ( cm2 )

icorr

Rp

x M TRITON-X-405
0
9.96 104
1.55 104
7.0 107
1.36 104
3.5 106
1.29 104
7.0 106
1.10 104
3.5 105
1.02 104
7.0 105

0.461
0.438
0.429
0.430
0.428
0.420

15.60
97.10
115.09
118.10
141.32
143.61

0.844
0.863
0.870
0.889
0.897

0.839
0.864
0.868
0.890
0.891

x M TRITON-X-405 + 103 M KBr


0
9.96 104
8.59 105
7.0 107
7.81 105
3.5 106
7.45 105
7.0 106
3.5 105
6.71 105
7.0 105
6.34 105
5.67 105
3.5 104
5.00 105
7.0 104

0.461
0.429
0.424
0.421
0.416
0.405
0.408
0.405

15.60
152.24
173.77
183.13
198.66
208.02
233.15
264.73

0.913
0.922
0.925
0.932
0.936
0.943
0.949

0.897
0.910
0.915
0.921
0.925
0.933
0.941

x M ZWITTER C16
0
9.96 104
8.66 104
1.0 106
7.53 104
3.0 106
5.11 104
1.0 105
3.59 104
2.5 105
2.35 104
5.0 105
1.41 104
1.0 104
8.49 105
5.0 104

0.461
0.459
0.457
0.455
0.452
0.445
0.444
0.436

17.531
19.832
29.800
39.800
63.600
104.860
175.051

0.140
0.223
0.487
0.630
0.764
0.858
0.915

0.110
0.213
0.477
0.608
0.755
0.851
0.911

105 M ZWITTER C16 + x M MTABr


0.461
0
9.96 104
5.33 104
0.451
1.0 105
5
4
3.70 10
0.450
3.0 10
2.15 104
0.443
1.0 104
1.48 104
0.440
2.0 104
9.28 105
0.433
4.5 104
8.75 105
0.431
5.0 104
5.83 105
0.419
1.0 103

31.99
39.80
83.35
103.31
184.05
219.25
253.93

0.464
0.628
0.784
0.852
0.907
0.912
0.941

0.512
0.608
0.812
0.849
0.915
0.928
0.938

104 M ZWITTER C16 + x M MTABr


0
9.96 104
1.0 105
2.12 104
0.447
1.65 104
0.446
3.0 105
1.17 104
0.440
1.0 104
4
4
1.06 10
0.440
1.2 10
4
5
9.71 10
0.441
1.6 10
6.60 105
0.427
5.0 104
5.15 105
0.417
1.0 103

77.74
113.86
135.42
140.54
155.14
230.04
304.23

0.786
0.834
0.883
0.893
0.902
0.930
0.949

0.799
0.863
0.885
0.890
0.900
0.932
0.949

2.0 M H2 SO4 +


IE (%) = 1

icorr (A cm2 )

Rp
Rp


100

(4)

The notation icorr , and Rp was used for measurements without



added surfactant, while the primed quantities icorr
and Rp applied
when the surfactant was added to the solution of 2.0 M H2 SO4 .
The adsorption of TRITON-X-405 mixed with 1 mM KBr in 2 M
H2 SO4 was in good agreement with the FloryHuggins isotherm in
the form:

log


cinh

= log(xK) + x log[1 ]

(5)

The same behaviour was shown for cationiczwitterionic surfactant mixture. The adsorption isotherms plotted in Figs. 7 and 8,
are for both investigated systems. According to the FloryHuggins
model, plots of log[/C] vs. log[1 ] produce straight lines of slope
(x) and intercept log (xK), as shown in Fig. 7, which is valid for the
used non-ionic surfactant mixture (TRITON-X-405, KBr). The data
indicate that the values of x were approximately 10 without added
KBr and around 13 when KBr was in the mixture. This suggests

Fig. 6. Inuence of added palmitylsulfo-betaineas (ZWITTER C16 ) with and without


the addition of x M MTABr on the cathodic and anodic behaviour of stainless steel
type X4Cr13 in 2.0 M H2 SO4 .

that one molecule of TRITON-X-405 adsorbed on the metal surface


replaces 10 water molecules, and increases to 13 with the help of a
synergistic effect between KBr and TRITON-X-405.
At a low concentration of Palmityl sulfobetaine, c = 1.0 105 M,
the addition of MTABr had no noticeable effect on the number of
replaced water molecules by one mole of Palmityl sulfobetaine. It
was approximately the same as in the case where MTABr was not
present in the mixture. When the concentration increased by an
order of magnitude (c = 1.0 104 M), the number of replaced water
molecules increased from 1.6 to around 3.0.
The same behaviour was obtained in the case where  was evaluated from the polarisation resistance. K is the modied adsorption
equilibrium constant, which can be related to the free energy of
adsorption Gads given by Eq. (6):
K=

1
csolvent

exp

 G

ads

RT


(6)

2178

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

Fig. 7. FloryHuggins adsorption isotherm for TRITON-X-405 with and without the addition of 1 mM of KBr on SS type X4Cr13 in 2.0 M H2 SO4 obtained from the corrosion
current density (icorr ).

Fig. 8. FloryHuggins adsorption isotherm for palmitylsulfo-betaineas (ZWITTER C16 ) with and without added x M MTABr on SS type X4Cr13 in 2.0 M H2 SO4 obtained from
the corrosion current density (icorr ).

Here csolvent is the molar concentration of the solvent, which, in


the case of water, is 55.5 mol L1 . Gads values calculated from
Figs. 7 and 8 are negative in accordance with Eq. (6), suggesting
the spontaneity of the adsorption process (Table 3).
Values for the free energy of adsorption Gads , up to
20 kJ mol1 are generally consistent with the strength of electrostatic interaction between charged molecules and charged metal
(physical adsorption), while those more negative than 40 kJ mol1
involve charge sharing or its transfer from the inhibitor molecules
to the metal surface to form a co-ordinate type of bond (chemisorption). The calculated values of Gads for all investigated mixtures
decreased with respect to a single surfactant. It was reduced from
87.5 kJ mol1 to around 115.66 kJ mol1 for TRITON-X-405, in
case when 1 mM of KBr was added in to the solution of 2 M H2 SO4 ,
so chemisorption is assumed. Triton has a large number of oxygen
with unpaired electrons. It is for this reason that TRITON-X-100 is
able to adsorb on the metal surface through lone pairs of electrons
on O atoms. In the instance of cationic/zwitterionic surfactant mixture (MTABr/Palmityl sulfobetaine) these values decreased with the
addition of MTABr from38.15 to 46.77 kJ mol1 . On the basis of

the obtained values for the free energy of adsorption, it could be


concluded that adsorption increased. It can be speculated that the
bromide ion, (which is added as KBr and mixed with non-ionic
surfactant TRITON-X-405 or it is presented as counterion in the
ionic surfactant MTABr and mixed with zwitterionic surfactant),
improved the inhibition properties of the used mixtures, as corrosion inhibitors, for ferritic stainless steel type X4Cr13 in 2 M H2 SO4
solution.
Fig. 9 shows the inhibitors efciency for both zwitterionic/cationic surfactant mixtures as a function of the concentration
(the concentration of the zwitterionic surfactant was at a xed

Table 3
Calculated values of Gads (from the FloryHuggins adsorption isotherm,
where  was obtained from icorr ) for TRITON-X-405, TRITON-X-405 + 1 mM KBr,
palmitylsulfo-betaineas (ZWITTER C16 ), 1.0 105 M ZWITTER C16 + x M MTABr and
1.0 104 M ZWITTER C16 + x M MTABr all in 2.0 M H2 SO4 at 25 C.
Surfactant

Gads /kJ mol1

TRITON-X-405
TRITON-X-405 + 1 mM KBr
palmitylsulfo-betaineas (ZWITTER C16 )
1.0 105 M ZWITTER C16 + x M MTABr
1.0 104 M ZWITTER C16 + x M MTABr

87.77
115.66
38.16
37.57
46.77

Fig. 9. The inhibition efciency as a function of concentration regarding added


1.0 105 M ZWITTER C16 + xM MTABr and 1.0 104 M ZWITTER C16 + x M MTABr
all in 2.0 M H2 SO4 at 25 C ( obtained from icorr )(palmitylsulfo-betaineasZWITTER
C16 ).

R. Fuchs-Godec / Electrochimica Acta 54 (2009) 21712179

value, while the concentration of the cationic surfactant MTABr varied. A curve with three slopes, was found in both cases. The slope
changes at a certain concentration of added cationic surfactant,
namely at the same values obtained regarding the surface tension
measurements.
The rst slope was below the rst cmc (cmc1 ); another trend
with some linearity and a reduced slope was found between
the rst and the second cmc (cmc2 ); and the third above the
cmc2 , i.e. cmc1 = 2.0 104 mol L1 and cmc2 = 4.6 104 mol L1
for the mixture of the 1.0 105 M ZWITTER C16 + x M MTABr.
For the mixture composed from 1.0 104 M ZWITTER C16 + x
M MTABr the cmcs values were cmc1 = 1.2 104 mol L1 and
cmc2 = 1.6 104 mol L1 all in 2.0 M H2 SO4 at 25 C. On the basis
of good agreement between the surface tension and the electrochemical measurements, it can be assumed, that either (i) then is
the existence of a second aggregation state, and/or (ii) the structural changes of the original micelles, that is transformation into
another geometry.
Further to this point, Ducker et al. [33] showed that
cationic and zwitterionic surfactant mixtures form surfaceaggregates with a structure intermediate between the structures
formed by each of the pure surfactants. A mixture of cylinderforming (cationic surfactanttetradecyltrimethylammonium bromide forms cylinders on mica) and sphere-forming surfactants
(zwitterionic surfactantdodecyldimethylammoniopropane sulfonate, a hemisphere-forming surfactant on mica) produces shapes
ranging from cylinders to truncated cylinders or spheres. The most
obvious change is the length of the aggregates, depending on the
mole fraction of the two surfactants. This one is difcult to quantify because it is non-linear with the mole fraction. Only about 20%
of the surfactant molecules in bulk need be cylinder-forming to
obtain the cylindrical aggregates Finally, it is important to include
those forces which concentrate one surfactant over another at the
metal surface (electrostatic interactions between surfactant headgroups and the charges at the solid substrate, and the hydrogen
bonds (or lack of) between the solvent and the solid substrate) [33].
More experimental data are needed to obtain fuller information
about the accurate sizes and shapes of the formed mixed micelles.
On the basis of those kinetic parameters measured during the corrosion processes, it was found that both chosen mixtures had the
same amounts of surfactants total concentrations in the solution
(i.e. 2.1 to 2.2 104 mol L1 ) at the cmc1 , but a slightly different
value for the achieved inhibition efciency at that concentration
(IE = 85, 90%). In regard to this, and from the adsorption isotherm
results, it could be speculated that the mixed micelles formed in
the solution with a higher concentration of the zwitterionic surfactant, are probably larger (due to longer alky chain), highly charged
(due to the loss of more water molecules of the bromide counterions), and due to the stronger electrostatic attraction, they are also
more strongly adsorbed on the metal surface: notice that calculated
Gads values decrease from 37.57to 46.77 kJ mol1 .
4. Conclusion
From the obtained corrosion data, it is clear that the Tafel lines
shifted towards more negative and more positive potentials for
the anodic and cathodic processes, respectively relative to the
blank curve. This means that the selected compound in the mixtures acts as a mixed-type inhibitor.
The tensiometric results in this study could suggest the existence
of a second state of aggregation for the zwitterionic/cationic surfactant mixtures.

2179

Adsorption of the used surfactants and theirs mixtures agreed


with FloryHuggins adsorption isotherm.
In the case of TRITON-X-405, about 10 to 11 molecules of water
seem to be replaced by one molecule of surfactant. By adding the
1 mM of KBr, the value of replaced molecules of water increased
to 13.
In the case of ZWITTER C16 palmitylsulfo-betaineas, the number
of replaced water molecules increased from 1.6 to 3, when the
concentration of the ZWITTER C16 in the chosen mixture was not
lower than 1.0 104 M.
The Gads values, calculated using the FloryHuggins adsorption
model, suggest the chemisorption mechanism for TRITONX-405 and also the same for the mixture of the TRITONX-405 + 1 mM of KBr. The calculated values of Gads were
between 37.57 kJ mol1 and 46.77 kJ mol1 , for the zwitterionic/cationic surfactant mixtures. This indicates a very strong
electrostatic attraction between the solvent (monomers or
micelles) and the solid substrate (metal surface).
The polarisation curves show an order of magnitude lower for
corrosion current density icorr and to an order and a half of magnitude lower icorr for the chosen mixtures with respect to the
inhibitor-free solution. From this reason, all of the selected mixtures indicated good inhibition properties for ferritic stainless
steel type X4Cr13 in 2 M H2 SO4 solution.
References
[1] R. Zhang, P. Somasundaran, Adv. Colloid Interface Sci. 213 (2006) 123.
[2] R. Atkin, V.S.J. Craig, E.J. Walness, S. Biggs, J. Colloid Interface Sci. 266 (2003)
236.
[3] M.A. Migahed, Mater. Chem Phys. 93 (2005) 48.
[4] F. Bentiss, M. Traisnel, M. Lagrene, Corros. Sci. 42 (2000) 127.
[5] M. Lagrene, B. Mernaru, N. Chaibi, M. Traisnel, H. Vezin, F. Bentiss, Corros. Sci.
43 (2001) 951.
[6] A. Popova, M. Christov, S. Raicheva, E. Sokolova, Corros. Sci. 46 (2004)
1333.
[7] M. Hosseini, F.L. Mertens, M.R. Arshadi, Corros. Sci. 45 (2003) 1473.
[8] L. Xueming, T. Libin, L. Lin, M. Guannan, L. Guangheng, Corros. Sci. 48 (2006)
308.
[9] F. Bentiss, M. Lebrini, M. Lagrene, Corros. Sci. 47 (2005 2915).
[10] R. FuchsGodec, Colloids Surf. A 280 (2006) 130.
[11] R. FuchsGodec, V. Dolecek, Colloids Surf. A 244 (2004) 73.
[12] V. Branzoi, F. Golgovici, F. Branzoi, Mater. Chem. Phys. 78 (2002) 122.
[13] S.T. Keera, M.A. Deyab, Colloids Surf. A 266 (2005) 129.
[14] D.F. Evans, H. Wennerstrm, The Colloidal Domain, 2nd ed., WileyVCH, New
York, 1999.
[15] T.F. Tadros, Applied Surfactants, WileyVCH Verlag, GmbH & Co KgaA, Winheim,
2005.
[16] S. Paria, K.C. Khilar, Adv. Colloid Interface Sci. 110 (2004) 75.
[17] R. Fuchs-Godec, Electrochim. Acta 52 (2007) 4974.
[18] S. Miyagishi, K. Okada, T. Asakawa, J. Colloid Interface Sci. 238 (2001) 91.
[19] K.S. Sharma, S.R. Patil, A.K. Rakshit, Colloids Surf. A 219 (2006) 67.
[20] E. Leontidis, Curr. Opin. Colloid Interface Sci. 7 (2002) 81.
[21] H. Schot, J. Colloid Interface Sci. 189 (1997) 117.
[22] T. Gu, P.A.G. Gomez, Colloids Surf. A 147 (1999) 365.
[23] B. Hribar, N.T. Southall, V. Vlachy, K.A. Dill, J. Am. Chem. Soc. 124 (2002) 12302.
[24] D.M. Nevskaia, A. Guerrero-Ruz, J.D. LpezGonzlez, J. Colloid Interface Sci.
205 (1998) 97.
[25] J.-H. Mu, G.-Z. Li, W.-C. Zhang, Z.-W. Wang, Colloids Surf. A 194 (2001) 1.
[26] D. Lopez-Diaz, I. Garcia-Mateos, M.M. Velazquez, J. Colloid Interface Sci. 299
(2006) 858.
[27] P.A. Gonzalez, J.L. Czapkiewicz, J.L.D. Castilo, J.R. Rodriguez, Colloids Surf. A 193
(2001) 129.
[28] K.S. Sharma, S.R. Patil, A.K. Rakshit, et al., J. Phys. Chem. B 108 (2004) 355.
[29] D.J.F. Taylor, R.K. Thomas, J. Penfold, Adv. Colloid Interface Sci. 132 (2007)
69.
[30] D. LopezDiaz, I. GarciaMateos, M.M. Velazquez, Colloids Surf. A 270 (2005)
153.
[31] A. Shiloach, D. Blankschtein, Langmuir 13 (1997) 3968.
[32] M. Druchok, B. Hribar-Lee, H. Krienke, V. Vlachy, Chem. Phys. Lett. 450 (2008)
281.
[33] W.A. Ducker, E.J. Wanless, Langmuir 12 (1996) 5915.

Das könnte Ihnen auch gefallen