Anda di halaman 1dari 8

CORROSION SCIENCE SECTION

576 CORROSIONJULY 2006


Submitted for publication May 2005; in revised form, October
2005.

Corresponding author. E-mail: g.montesperelli@univpm.it.


* Dipartimento di Scienze e Tecnologie Chimiche, Universit di
Roma Tor Vergata, Via della Ricerca Scientica, 00133 Rome,
Italy.
** Nalco Italiana S.r.l., Via Ninna II, 04012 Cisterna di Latina (Lt),
Italy.
*** Dipartimento di Fisica ed Ingegneria dei Materiali e del Territorio,
Universit Politecnica delle Marche, Via Brecce Bianche, 60131
Ancona, Italy.
Study of the Inhibition Mechanism of Imidazolines
by Electrochemical Impedance Spectroscopy
G. Gusmano,* P. Labella,** G. Montesperelli,
,
*** A. Privitera,** and S. Tassinari**
ABSTRACT
In this paper, the results of evaluating the activity of imidazo-
line-based corrosion inhibitors by means of electrochemical
impedance spectroscopy (EIS) and other traditional electro-
chemical methods are reported. Tests were carried out on AISI
1040 (UNS G10400) mild steel specimens immersed in solu-
tion simulating the composition of an aqueous phase during
oil production and transportation: high salinity, presence of
carbon dioxide (CO
2
), anaerobic environment, 40C tempera-
ture. Two different inhibitor formulations were tested in the
presence of different salting agents (acetic acid [CH
3
COOH]
and thioglycolic acid [HSCH
2
COOH or TGA]). EIS data were
analyzed using equivalent circuits. The electrochemical tests
indicated evidence of different mechanisms and activity
depending on inhibitor formulation.
KEY WORDS: carbon dioxide corrosion, carbon steel, corrosion
inhibitors, electrochemical impedance spectroscopy
INTRODUCTION
Corrosion is one of the most serious problems in the
oil and gas production and transportation industry.
The environmental conditions in oileld systems can
be very different, but they can be generally summa-
rized as follows:
anaerobic conditions,
a high concentration of carbon dioxide (CO
2
),
high salinity, and
a variable concentration of suldes.
1-2
Depending on environmental composition, different
forms of corrosion can take place in oil and gas wells
and ow lines.
3
Among them, CO
2
corrosion, usually
called sweet corrosion, is one of the most severe be-
cause the effect of CO
2
and/or carbonic acid (H
2
CO
3
)
can enhance the electrochemical cathodic reaction.
4-6
The use of corrosion inhibitors is currently used
to protect against corrosion in all petrochemical facili-
ties in the world, because it is cost-effective and ex-
ible.
7-8
The total consumption of corrosion inhibitors in
the United States has doubled from approximately
$600 million in 1982 to nearly $1.1 billion in 1998.
9
It also has been reported that the U.S. market for
corrosion inhibitors is estimated to be up to $1.6 bil-
lion per year by the year 2006.
10
Some of the most effective corrosion inhibitors
for oil and gas pipeline application are the fatty acid
imidazolines. The general chemical structure of im-
idazoline inhibitors is shown in Figure 1. It consists
of three parts: a ve-atom ring with two nitrogen ele-
ments (part A), a pendant side chain with a functional
group (part B), and a long hydrocarbon chain (part C).
The inhibition mechanism of imidazoline inhibitors
is poorly understood and some discrepancies can be
observed in the main literature.
11-17
It can be depicted
as follows:
0010-9312/06/000109/$5.00+$0.50/0
2006, NACE International
CORROSION SCIENCE SECTION
CORROSIONVol. 62, No. 7 577
The head group (part A) promotes a strong
bonding of the molecule to the surface of the
metal and forms an ordered lm of one or more
layers.
The lone pair on the CH
2
CH
2
NH
2
pendant
chain group (part B) could also help to im-
prove the molecular adsorption to the surface,
thereby improving the corrosion inhibition
performance.
The hydrocarbon tail (part C) promotes a hydro-
phobic surface covering of the metal surface.
The hydrocarbon tail and pendant group must have
a favorable partition coefcient between hydrocarbon
and water phase, to permit the lm to build a hydro-
phobic barrier that blocks water from the surface.
Recent works, based on surface binding energies
calculations, demonstrated a favorable adsorption of
imidazolines on iron oxide (Fe
3
O
4
)
18
and iron carbon-
ate (FeCO
3
)
19
that could form a rst layer on the metal
surface, but no experimental evidences are conrmed.
Moreover, imidazolines have very low solubility in
a water phase, and to increase solubility, these com-
pounds are usually converted to a salt with acids. The
most common acid of these types of formulations is
acetic acid (CH
3
COOH). It has been demonstrated that
carboxylates can inuence the orientation of organic
inhibitors on the steel surface.
20

In practice, the application of inhibitors can be
very problematic because of the variation of the envi-
ronment composition. For these reasons, the use of
some monitoring techniques able to evaluate inhibitor
efciency and to determine critical concentration,
depending on the environmental condition, is essen-
tial. There are many techniques that can be used for
inhibitor testing. A few of them are able to give an im-
proved fundamental knowledge of corrosion processes
and the mechanisms of corrosion inhibitors, to ac-
quire better control methods in relation to the envi-
ronmental corrosiveness.
Electrochemical impedance spectroscopy (EIS)
has often been used in various corrosion elds such
as coatings, passive layer buildup and breakdown,
and inhibitor evaluation.
21-23
EIS is well known as a
powerful tool able to provide information on corrosion
rate, both on corrosion and protection kinetics and
mechanism.
In this work, EIS, together with other traditional
electrochemical techniques, were used to study the
imidazoline-based inhibitor lm formation and evalu-
ation. In particular, the effect of thioglycolic acid
(HSCH
2
COOH [TGA]) in conjunction with acetic acid
as the salting agent in the inhibitor formulation was
studied.
EXPERIMENTAL PROCEDURES
Tests were performed on AISI 1040
(1)
(UNS
G10400)
(2)
mild steel samples. Test electrodes were
rods, 4 mm in diameter. The surface preparation con-
sisted of abrading the specimen with silicon carbide
(SiC) 500-grit paper, washing in water then in ethanol
(C
2
H
5
OH), and drying in air. The surface electrode
was delimitated by enamel and the exposed area was
10.2 cm
2
.
Tests were carried out in an unstirred solution
of 30 g/L sodium chloride (NaCl) in which a constant
CO
2
bubbling was performed throughout the tests.
A sealed cell was used with a vent to prevent over-
pressure. The cell was thermostat-controlled at 40C
and oxygen concentration was measured using an
Orbisphere Laboratories model 2713

probe and kept


under 10 ppb during all the tests. The pH value mea-
sured throughout the tests was in the range from 3.8
to 4.0.
Inhibitors were a commercial mixture of imidazo-
lines with an aliphatic chain varying between 12 and
18 carbon atoms, as shown in Figure 1. Two different
commercial inhibitors (Nalco Italiana [Rome, Italy]),
at concentrations of 25 ppm or 50 ppm were used
and their compositions are reported in Table 1. Blank
tests were also performed.
EIS measurements were performed using a Solar-
tron 1260

frequency response analyzer and a Solar-


tron 1287

electrochemical interface. An alternating


current (AC) signal, 5 mV in amplitude, was applied
at the open-circuit potential (OCP) in the frequency
range of 5 mHz to 500 kHz. Data were collected and

(1)
American Iron and Steel Institute (AISI), 1140 Connecticut Ave.,
Ste. 705, Washington, DC 20036.

(2)
UNS numbers are listed in Metals and Alloys in the Unied Num-
bering System, published by the Society of Automotive Engineers
(SAE International) and cosponsored by ASTM International.

Trade name.
FIGURE 1. General chemical structure of imidazolines used in this
work.
TABLE 1
Composition of Inhibitors
Components A (wt%) B (wt%)
Imidazolines 25 25
Acetic acid (CH
3
COOH) 6 3
Thioglycolic acid (HSCH
2
COOH) 3
Isopropanol ([CH
3
]
2
CHOH) 10 10
Water 59 59
CORROSION SCIENCE SECTION
578 CORROSIONJULY 2006
RESULTS AND DISCUSSION
OCP trends as a function of immersion time for
tests at an inhibitor concentration of 50 ppm are re-
ported in Figure 2, in which the curve for a blank test
is also reported. Curves for inhibited solutions show
a similar trend, with a very strong potential ennoble-
ment with respect to the blank test, already detectable
in the rst hours of immersion. After 4 h, OCP kept
a stable value of 650 mV vs. silver/silver chloride
(Ag/AgCl) to 640 mV
Ag/AgCl
in spite of the 710 mV
Ag/AgCl

measured in the blank test.
Polarization curves were carried out in the same
conditions. Figure 3 reports polarization curves ac-
quired after 0, 2, and 6 h of immersion for a test with
inhibitor A at 50 ppm. Curves conrmed an OCP
shifted toward anodic values and evidenced a marked
variation with time of both anodic (
a
) and cathodic
(
c
) Tafel slopes. In the anodic segment of polariza-
tion curves collected after 2 h and 6 h of immersion,
a slope variation could be detected, as shown with
arrows in Figure 3. This potential, usually called
desorption potential (E
d
),
24
reveals the potential at
which the lm breakdown begins. The lm formation
and stabilization was conrmed by the increasing of
E
d
from 584 mV to 557 mV, with immersion time
increasing from 2 h to 6 h. Data from the polariza-
tion curves after 6 h of immersion are summarized in
Table 2 in which
a
,
c
, and E
d
for tests on both inhibi-
tors and a blank test are reported.
A comparison of data coming from inhibitors A
and B revealed a strong decrease of the cathodic Tafel
slope of inhibitor B with respect to inhibitor A and a
clear lm stabilization. The anodic part remained un-
affected. Polarization curves collected in the solution
of inhibitor B presumably reveal the effect of TGA. It
has been reported that TGA can have an inhibition or
activation effect in aqueous environments containing
CO
2
, depending on pH, temperature, and its concen-
tration. It has also been reported that TGA can have
inuence on the cathodic Tafel slope, without any ap-
preciable inuence on the anodic part.
25
A corrosion rate calculation by LPR, shown in Fig-
ure 4, conrmed this nding. After about 12 h of im-
mersion, the corrosion rate for inhibitor A at 25 ppm
settled at 0.15 mm/y, while the curve related to the
same inhibitor at 50 ppm settled at 0.10 mm/y. The
same tests for inhibitor B showed values lower than
0.02 mm/y after only 5 h for both concentrations.
Blank tests evidenced an average corrosion rate of
1.40 mm/y.
Figures 5(a) and (b) show an EIS plot in Bode
format recorded on a test with 25 ppm of inhibitor A
at different immersion times. The increasing of im-
pedance magnitude during the test indicates that the
presence of the inhibitor strongly reduced the rate
of electrochemical reactions on the electrode sur-
face. Total resistance increased with immersion time,
FIGURE 2. OCP trend as a function of time for inhibited solutions
(50 ppm) and for a blank test.
FIGURE 3. Polarization curves for a test with inhibitor A (50 ppm) at
different immersion times.
TABLE 2

a
,
c
, and E
d
After 6 h of Immersion
Blank Inhibitor A Inhibitor B

a
(mV) 93 130 129

c
(mV) 318 217 89
E
d
(mV) 557 439
analyzed using Zplot-Zview software. Polarization
curves were carried out using a Solartron model
1285

potentiostat-galvanostat, with a scan rate of


4 mV/s.
Corrosion rate measurements were performed
using linear polarization resistance (LPR) with a corro-
simeter using a three-electrode corrosion probe.
Although tests were run in triplicate, data are
presented as individual experiments. The calculated
error was within 10 % for each test.
CORROSION SCIENCE SECTION
CORROSIONVol. 62, No. 7 579
reaching a maximum value after 22 h, thus reaching
a steady state. The vs. log plot recorded of elec-
trodes after immersion, and shown in Figure 5(b),
shows a single phase angle shift at 100 Hz that can
be ascribed to charge-transfer and double-layer char-
acteristics.
15
After 2 h from the immersion, the phase
plot showed a new phase angle shift at a higher fre-
quency range (>1 kHz), which was more pronounced
with increasing time. This new peak means that the
structure of the electrode interface has changed as a
consequence of lm buildup. At the same time, the
charge-transfer peak shifted at lower frequencies
(10 Hz and lower). The presence of the two-phase
angle shifts in the vs. log plot suggests that there
are two main kinetic processes on the electrode sur-
face. Consistent with other authors,
15,26-27
the high-
frequency loop can be ascribed to the inhibitor lm
because of its small time constant that causes a
phase shift in the high-frequency limit.
Figures 5(a) and (b) also show an EIS plot for a
blank test after 4 h of immersion. As expected, a very
low total resistance and only one time constant were
detected.
Usually, impedance data are analyzed by us-
ing an equivalent circuit, taking into account the
contribution of each phenomenon such as polariza-
tion resistance, double layer, lm formation, etc. The
nonlinear least-square (NLLS) method is usually used
to simultaneously and iteratively adjust the circuit
element values to obtain the best match between
the experimental and simulated data. The statisti-
cal relevance of the simulation is usually given by
the chi-squared (
2
) test that easily gives the good-
ness of a t by calculating the square of the standard
deviation between the experimental data and the
calculated spectrum. Another signicant parameter
is the weighted sum of squares (WSOS), which is pro-
portional to the average percentage error between the
experimental data and the calculated values. The best
equivalent circuit is that which can minimize
2
and
WSOS.
Moreover, all the impedance spectra acquired
had a form of a depressed semicircle in the Nyquist
representation, as shown in Figure 6, for a test with
25 ppm of inhibitor A. This may have different ex-
planations because of the roughness of a solid metal
electrode,
28
the inhomogeneous reaction rate on the
electrode surface,
29
or the varying thickness or com-
position of a coating.
30
In these cases, constant phase
elements (CPE) are generally used instead of a real
capacitance. The CPE parameters were converted into
capacitance values following the reported procedure.
31
Figure 7 reports the equivalent circuits used
in this paper. The rst EIS spectra acquired a few
minutes after the immersion were analyzed through
a simple equivalent circuit for the corrosion system
under charge-transfer control reported in Figure 7(a),
with R
s
being the solution resistance, R
ct
the charge-
transfer resistance, and C
dl
the double-layer capaci-
tance. No mass-transport phenomena were detected,
but, in spite of a very low oxygen concentration, it
FIGURE 4. Corrosion rate by LPR as a function of immersion time.
FIGURE 5. Bode plot of EIS data during lming process with 25 ppm
of inhibitor A.
(a)
(b)
CORROSION SCIENCE SECTION
580 CORROSIONJULY 2006
may not be excluded that mass phenomena may ap-
pear at frequencies lower than 5 mHz.
EIS data of electrodes immersed in inhibitor A
solution were analyzed by different equivalent cir-
cuits. Among them, the circuit in Figure 7(b) gave the
best results in terms of
2
and WSOS. The circuit in
Figure 7(b) is often reported in the literature
32-33
to
simulate the behavior of coated metals, where R
f
is
the pore resistance that is due to the formation of an
ionically conducting path across the lm and C
f
is the
capacitance of the lm. The use of this circuit gave
satisfactory results as shown in Figure 8 for an EIS
acquisition in a test with 25 ppm of inhibitor A.
The
2
test and WSOS values were 4.2 10
4
and
3.6 10
2
for circuit 6b, thus conrming a good t.
The use of circuit 6b permitted the analysis of all the
spectra acquired after 2 h from immersion into the
solution containing inhibitor A.
The use of the equivalent circuit with the NLLS
t permitted the calculation of the charge-transfer
and lm parameters during the test. Figure 9 reports
the trend of R
ct
and R
f
for a typical test in inhibitor A
at 50 ppm. In the rst hours of immersion, R
ct
and R
f

have very similar trends characterized by a quick
increase, conrming that the lm formation is fast.
After 10 h, R
ct
settled at a constant value of about
1,350 cm
2
, whereas R
f
increased slightly. Figure
10 reports the trend of C
dl
and C
f
for the same test
illustrated in Figure 9. C
dl
and C
f
show an inverse
trend, with a strong decrease in C
dl
in the very rst
part of the test and a slight increase in C
f
during the
rst 20 h of the test. After that, both capacitances
come to a constant and very close value of about
870 F/cm
2
.
Considering that the electrode surface (A) is con-
stant throughout the test, capacitance values are
directly proportional to the dielectric constant of the
lm (
f
) and inversely proportional to the thickness of
the layer (d) following the relationship:

C
A
d
f
f
=
0

(1)
where
0
is the permittivity of free space.
Concerning the evolution of C
f
with time, the
growing of an inhibitor lm may, in principle, cause
an increase in the thickness and density of the lm.
FIGURE 6. Nyquist plot of EIS data during lming process with 25 ppm of inhibitor A.
FIGURE 7. Equivalent circuits used in this paper.
FIGURE 8. Fitting of a typical Bode plot for a test with inhibitor A,
using the equivalent circuit of Figure 7(b).
CORROSION SCIENCE SECTION
CORROSIONVol. 62, No. 7 581
Theoretical calculation demonstrated that imidazoline
lm presumably forms a bilayer on an inner lm of
FeCO
3
and/or Fe
3
O
4
and takes place very quickly.
18,34
Moreover, optical polarization studies have con-
rmed that the imidazoline molecules are at on the
surface.
35
In this hypothesis, the lm thickness could
be assumed constant after a few hours from im-
mersion. The growing lm becomes a more compact
structure, closing the porosity and thus increasing
density. The increase in the compactness generates
an increase in the dielectric constant and therefore in
capacitance. On the other hand, the decrease in C
dl

can be explained by the decrease in charge density
within the double layer as the imidazoline lm grows.
EIS tests on inhibitor B were performed and ana-
lyzed with the same test procedure on inhibitor A.
Figures 11(a) and (b) show some EIS spectra in Bode
format, for a test with 25 ppm of inhibitor B at differ-
ent immersion times. The magnitude vs. frequency
plot shows a fast increase in total resistance. It
arrived at a constant value after about 9 h, which is
much faster than inhibitor A.
The phase angle plot shows a very large angle
shift that increases in width with time, which actually
is formed from at least two or more different peaks.
This large phase shift suggests a more complicated
lm-metal interface structure with respect to inhibitor
A.
The role of TGA is not well understood at the
moment. It has been reported that, under anaerobic
conditions, a complex anion [Fe
(II)
(S-CH
2
-COO)
2
]
2
is
formed
36-37
and can be oxidized to form iron disulde
(FeS
2
) that precipitates on the metal surface.
38
An XPS
investigation on electrodes immersed in the same con-
dition of corrosion tests has been undertaken. Prelim-
inary results seem to conrm the presence of an iron
sulde and/or disulde layer on the metal surface.
For this reason, EIS data on inhibitor B were t-
ted with the equivalent circuit of Figure 7(c), in which
FIGURE 9. R
ct
and R
f
trend as a function of time for a typical test in
inhibitor A at 50 ppm.
FIGURE 10. C
dl
and C
f
trend as a function of time for a typical test in
inhibitor A at 50 ppm.
FIGURE 11. Bode plot of EIS data during the lming process with
25 ppm of inhibitor B.
(b)
(a)
CORROSION SCIENCE SECTION
582 CORROSIONJULY 2006
solution and acts as an inner layer. The very low R
z

value indicates that it is not protective.
Capacitance values are summarized in Figure 14.
The C
dl
showed a trend very comparable to that of test
results with inhibitor A, with a decrease in the rst
part of the test followed by reaching a constant value
of about 500 F/cm
2
. This value is lower than that
calculated for the tests in inhibitor A, as predicted
from the higher level of protection of inhibitor B. The
trend of C
f
is strictly connected to the evolution of
lm structure. The inhibitor adsorption on the metal
surface is faster with respect to inhibitor A, probably
because of the presence of TGA. It can be assumed
that the adsorption stage takes approximately 2 h, in
which C
f
decreases as a consequence of the lm thick-
ness increasing. After that, a trend comparable to that
of inhibitor A has been detected with a slight increase
of C
f
due to the buildup of a more compact struc-
ture. The C
z
contribution is approximately constant
throughout the test and its value settled at about
50 F/cm
2
. This lead to the deduction that the effect
of TGA occurred before imidazoline lm growth. In
any case, its contribution took place in the very rst
part of the test.
CONCLUSIONS
An electrochemical analysis by means of different
techniques, using two imidazoline-based inhibitors
with two different formulations, was performed. Tests
were carried out in 30 g/L NaCl at 40C, with con-
stant CO
2
bubbling. Some blank tests, without the
presence of an inhibitor, were also performed.
All the electrochemical tests were consistent and
demonstrated than inhibitor B has an activity higher
than that of inhibitor A.
OCP measurements evidenced a strong ennoble-
ment of both the inhibitors with respect to the blank
tests.
FIGURE 12. Fitting of a typical Bode plot for a test with inhibitor B,
using the equivalent circuit in Figure 7(c).
FIGURE 13. R
ct
, R
f
, and R
z
trend as a function of time for a typical
test in inhibitor B at 50 ppm.
FIGURE 14. C
dl
, C
f
, and C
z
trend as a function of time for a typical
test in inhibitor B at 50 ppm.
R
z
and C
z
have been added to the previous circuit, in
accordance with early experimental ndings of an iron
sulde/disulde layer. At the moment, additional XPS
investigations are in progress and nal results will be
published soon.
A comparison between experimental and calcu-
lated EIS spectra is shown in Figure 12. The use of
circuit 6c provided a tenfold decrease of the
2
value
from 2.3 10
3
to 3.2 10
4
, thus justifying the intro-
duction of the new element.
Resistance values coming from EIS data tting
are shown in Figure 13. R
ct
and R
f
have the same
trend since both of them are related to the lm for-
mation. R
f
is about one order of magnitude higher
with respect to that of inhibitor A. As a consequence
of this, R
ct
reached a steady value of 3,500 cm
2
. R
z

gave a poor contribution to the total resistance (about
15 cm
2
) and is essentially unaffected by time. As
a rst assumption, TGA seems to be inaccessible by
CORROSION SCIENCE SECTION
CORROSIONVol. 62, No. 7 583
Polarization curves revealed an increasing of anodic
Tafel slopes for both inhibitors. The cathodic segment
revealed a very strong depolarization effect for inhibi-
tor B and a milder one for inhibitor A. The presence of
a desorption potential revealed the growth of a lm on
the electrode surface, more stable on inhibitor B with
respect to inhibitor A.
LPR evidenced very slow corrosion rates (0.10 mm/y
and 0.02 mm/y for inhibitors A and B, respectively),
with respect to blank tests (1.40 mm/y).
EIS revealed two different protection mechanisms
of the inhibitors evidenced by the use of two equiva-
lent circuits to simulated EIS data.
The presence of TGA within inhibitor B introduced
an additional RC element in the equivalent circuit
used to t the EIS data.
Presumably, the TGA contribution is faster than
imidazolines adsorption, and assists the growth of
a thicker and more compact lm structure, as con-
rmed from the evolution of lm resistance and
capacitance as a function of immersion time.
Some surface investigations are ongoing to deter-
mine the role of TGA on the protection mechanism of
inhibitor B. Preliminary results seem to indicate the
presence of iron suldes on the metal surface.
REFERENCES
1. A.A. El Hosary, R.M. Saleh, A.O. Abdel Fattah, Corrosion Inhibi-
tors in Petroleum ProductionEffect of Water Salinity and Con-
tent, 8th European Symp. in Corrosion Inhibitors (8SEIC),
Suppl. no. 10, Sez. V (Ferrara, Italy: Ann. Univ. Ferrara, N.S.,
1995), p. 1,269.
2. H.J. Choi, R.L. Cepulis, J.B. Lee, Corrosion 45, 11 (1989): p. 943.
3. ASM Handbook, vol. 13, Corrosion, 9th ed. (Materials Park, OH:
ASM International, 1987), p. 1,232.
4. C. de Waard, D.E. Milliams, Corrosion 31, 5 (1975): p. 177.
5. A. Ikeda, M. Ueda, S. Mukai, CO
2
Corrosion Behaviour and
Mechanism of Carbon Steel and Alloy Steel, CORROSION/83,
paper no. 45 (Houston, TX: NACE International, 1983).
6. G. Schmitt, B. Rothman, Werkst. Korros. (Mater. Corros.) 28, 12
(1977): p. 816.
7. S. Webster, Corrosion Inhibitor Selection for Oileld Pipeline,
CORROSION/93, paper no. 109 (Houston, TX: NACE, 1993).
8. A. Nestle, Corrosion Inhibitors in Petroleum Production Primary
Recovery, Corrosion Inhibitors, ed. C.C. Nathan (Houston, TX:
NACE, 1973), p. 61-65.
9. Corrosion Inhibitors, report by Publications Resource Group,
Business Communications Company, July 1999.
10. Freedonia Industry Study, Corrosion Inhibitors to 2006, Study
no. 1570, June 2002, http://www.freedoniagroup.com.
11. S. Ramachandran, V. Jovancicevic, Corrosion 55, 3 (1999): p. 259.
12. A. Edwards, C. Osborne, S. Webster, D. Klenerman, M. Joseph,
P. Ostovar, M. Doyle, Corros. Sci. 36 (1993): p. 315.
13. V. Jovancicevic, S. Ramachandran, P. Prince, Corrosion 55, 5
(1999): p. 449.
14. I. Esih, T. Soric, Z. Pavlinic, Br. Corros. J. 33, 4 (1998): p. 309.
15. Y.L. Tan, S. Bailey, B. Kinsella, Corros. Sci. 38, 9 (1996): p. 1,545.
16. D. Wang, S. Li, Y. Ying, M. Wang, H. Xiao, Z. Chen, Corros. Sci.
41, 10 (1999): p. 1,911.
17. A.J. Szyprowski, Hydrogen Sulphide Corrosion of SteelMecha-
nism of Action of Imidazoline Inhibitors, 8th European Symp. in
Corrosion Inhibitors (8SEIC), Sez. V, Suppl. no. 10 (Ferrara, Italy:
Ann. Univ. Ferrara, N.S., 1995), p. 1,229.
18. S. Ramachandran, V. Jovancicevic, Molecular Modeling of the
Inhibitor of Mild Steel CO
2
by Imidazolines, CORROSION/98,
paper no. 17 (Houston, TX: NACE, 1998).
19. B.A. Alink, G.G. Duggan, V. Jovancicevic, J.A. McMahon, S.
Ramachandran, Baker Petrolite, a supplement to Mater. Perform.
3 (1999): p. 20.
20. G.A. Salensky, M.G. Cobb, D.S. Everhart, Ind. Eng. Chem., Prod.
Res. Dev. 25, 2 (1986): p. 133.
21. L. Domingues, C. Oliveira, J.C.S. Fernandes, M.G.S. Ferreira,
Electrochim. Acta 47, 13-14 (2002): p. 2,253.
22. C. Monticelli, A. Frignani, G. Trabanelli, J. Appl. Electrochem.
32, 5 (2002): p. 527.
23. R. Subramanian, V. Lakshminarayanan, Corros. Sci. 44, 3
(2002): p. 535.
24. F. Bentiss, M. Lagrene, M. Traisnel, Corrosion 56, 7 (2000): p. 733.
25. K. Bilkova, N. Hackerman, M. Bartos, Corrosion 60, 10 (2004):
p. 965.
26. C.H. Tsai, F. Mansfeld, Corrosion 49, 9 (1993): p. 726.
27. I. Thompson, D. Campbell, Corros. Sci. 36, 1 (1994): p. 187.
28. W.H. Mulder, J.H. Sluytes, T. Pajkossy, I. Nyikos, J. Electroanal.
Chem. 285, 1-2 (1990): p. 103.
29. C.H. Kim, S.I. Pyun, J.H. Kim, Electrochim. Acta 48, 23 (2003):
p. 3,455.
30. C.A. Schiller, W. Strunz, Electrochim. Acta 46, 24-25 (2001):
p. 3,619.
31. C.H. Hsu, F. Mansfeld, Corrosion 57, 9 (2001): p. 747.
32. L. Beaunier, I. Epelboin, J.C. Lestrade, H. Takenouti, Surf.
Technol. 4 (1976): p. 237.
33. F.B. Mansfeld, C.H. Tsai, Corrosion 47, 12 (1991): p. 958.
34. K. Sapre, S. Seal, P. Jepson, H.B. Wang, Z. Rahman, T. Smith,
Corros. Sci. 45, 1 (2003): p. 59.
35. A.J. McMahon, Colloids Surf. 59, 2 (1991): p. 187.
36. F.P. Dwyer, Chelating Agents and Metal Chelates, eds. F.P.
Dwyer, D.P. Mellor (New York, NY: Academic Press, 1964), p. 374.
37. J.C. Bailar, Jr., D. Busch, eds., The Chemistry of the Coordina-
tion Compounds (New York, NY: Reinhold Publishing Corpora-
tion, 1956), p. 52.
38. V. Jovancicevic, Y.S. Ahn, J. Dougherty, B. Alink, CO
2
Corrosion
Inhibition by Sulfur-Containing Organic Compounds, CORRO-
SION/2000, paper no. 7 (Houston, TX: NACE, 2000), p. 39.

Anda mungkin juga menyukai