You are on page 1of 21

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/273317586

Research on corrosion inhibitors for acid


stimulation

Conference Paper in NACE - International Corrosion Conference Series · March 2012

CITATIONS READS

11 962

4 authors, including:

E.B. Barmatov Tamera Hughes


Schlumberger Limited University of Cincinnati
117 PUBLICATIONS 798 CITATIONS 11 PUBLICATIONS 146 CITATIONS

SEE PROFILE SEE PROFILE

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

Flow induced corrosion View project

All content following this page was uploaded by E.B. Barmatov on 09 March 2015.

The user has requested enhancement of the downloaded file.


C2012-0001573

Research on Corrosion Inhibitors for Acid Stimulation

Evgeny Barmatov Jill Geddes


Schlumberger Cambridge Research Schlumberger Cambridge Research
High Cross, Madingley Road High Cross, Madingley Road
Cambridge, CB3 OEL, UK Cambridge, CB3 OEL, UK

Trevor Hughes Michaela Nagl


Schlumberger Cambridge Research Schlumberger Cambridge Research
High Cross, Madingley Road High Cross, Madingley Road
Cambridge, CB3 OEL, UK Cambridge, CB3 OEL, UK

ABSTRACT
In the Oil & Gas industry, formation treatment with acids is an established method to stimulate
carbonates or dissolve fines. Either organic acids such as citric or acetic acid or mineral acids such as
HCl or HF are injected into the well at high concentrations. In the absence of corrosion inhibitors, the
generalized corrosion rate increases exponentially with acid concentration and temperature. This can
be reduced to an acceptable level by adding appropriate concentrations of inhibitor products in order to
protect and prolong the useful lifetime of the hardware present in the well. Currently, inhibitor packages
demonstrate high efficiency with carbon steels and at temperatures below 150°C. An overview of
commonly applied acid corrosion inhibitors will be presented and a comparison is made between
published data and data measured in-house. Some common pitfalls associated with the evaluation of
acid corrosion inhibitors are highlighted. Weight loss analysis is a widely used method for assessing the
efficiency of corrosion inhibitors. The effects of surface roughness and acid volume to metal surface
area ratio are reviewed for gravimetric experiments in solutions of 14 and 28 wt. % hydrochloric acid at
78°C. Results on low carbon steel Coiled Tubing (CT) coupons that were either glass bead blasted,
pickled or polished are reported. This paper also investigates the inhibiting effect of different classes of
corrosion inhibitors in 14 wt. % HCl at 78°C using electrochemical methods (linear polarization
resistance and Tafel extrapolation method) and evaluates the synergetic behavior of acetylenic
alcohols with quaternary ammonium cations (QUATs).
Key words: corrosion, inhibitors, hydrochloric acid, Oil & Gas industry

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
INTRODUCTION
Acid stimulation of oil and gas wells is a widely established technique to increase hydrocarbon
production. In the absence of corrosion inhibitors, matrix acidizing treatment fluids would induce severe
corrosion of downhole equipment. Since the 1930s, corrosion inhibitors have been used in matrix
acidizing fluid to protect the exposed metal assets. The earliest inhibitors were inorganic salts or acids
such as arsenate or arsenic acid but by the mid-1970s these had been replaced by the use of organic
molecules. The latter generally contain a heteroatom such as N, O, P or S and an aromatic or
unsaturated functionality. Although today nitrogen-containing compounds constitute the largest class of
inhibitors for hydrochloric acid formulations, acetylenic alcohols such as propargyl alcohol or 1-octyn-3-
ol are widely used in non-oxidizing and oxidizing acids providing good inhibition performance in high
temperature applications. Both nitrogen-containing molecules and acetylenic alcohols are claimed to
form a film on the metal surface and can retard the metal-dissolution process (anodic reaction) as well
as the hydrogen-evolution (cathodic reaction).
All organic inhibitors work by adsorbing on the metal surface. The mode of adsorption depends on the
chemical structure of the molecule(s), the composition of the solution, the composition/properties of the
metal surface and the electrochemical potential at the metal/solution interface. Depending on the type
of metal-inhibitor interaction, the inhibitors can be classified into physi-sorbed, chemi-sorbed, and
polymerizable species. Physi-sorption is the result of electrostatic attractive forces between inhibiting
organic ions or dipoles and the electrically charged surface of the metal.
Unsaturated molecules such as alkynols adsorb on the metal surface via chemi-sorption of the oxygen
lone pair and the π-electrons of the triple bond. Babic-Samardzija et al. evaluated the inhibitive
properties of 2-butyn-1-ol, 3-butyn-1-ol, 3-pentyn-1-ol and 4-pentyn-1-ol on iron in 1M HCl at ambient
temperatures.1 They observed mixed-type inhibition and a linear correlation between inhibitor efficiency
and both immersion time and inhibitor concentration. Furthermore, longer chains combined with the
terminal alkyne functionality protected the metal more from corrosion compared to shorter molecules
with an internal triple bond. X-ray photoelectron spectroscopy (XPS) studies confirmed the formation of
an organic carbon coating on the surface, which had been suggested earlier by several authors.2
Complexes formed only by interaction with the π-electrons of acetylene or substituted alkynes are
relatively unstable while the presence of a hydroxyl group next to the triple bond enhances the initial
interaction with the metal surface. Under acid conditions, α-hydroxy alkynyls are protonated resulting in
enhanced adsorption and the initiation of dehydration and rearrangements to form a polymeric film.
While propargyl alcohol is soluble in acid, the solubility of α-hydroxy alkynyls decreases with increasing
carbon chain length. As reported by Tedeschi,3 the solubility of such alkynols increases when combined
with a quaternary ammonium surfactants.
Longer chain alkynols may form a thicker hydrophobic layer on the metal surface provided they can be
dispersed in concentrated acid solutions. In this paper, the long chain alkynols 1-octyn-3-ol and 4-ethyl-
1-octyn-3-ol are investigated and their performance is compared with propargyl alcohol. Furthermore,
the effect of combining these alkynols with cationic surfactants was studied.
The present work describes the use of physi-sorbed, chemi-sorbed and polymerizable corrosion
inhibitors on low carbon steel in 14 and 28 wt. % HCl solutions at different temperatures. The behavior
of several series of corrosion inhibitors was evaluated by gravimetric (weight loss) experiments, linear
polarization resistance and Tafel extrapolation methods.

EXPERIMENTAL PROCEDURE
Weight loss analysis
The coiled tubing (CT) pipe was cut into coupons of 29±3 cm2 surface area using a water-cooled band
saw to minimize any changes in the properties of the metal that may result from heating. After cutting,
the coupons were stamped for identification, glass bead blasted (BB) and cleaned in acetone to remove
any residue from the machining process. The coupons were then dried, placed in zip lock bags and
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
stored in airtight jars filled with water adsorbent, until they were used. Coupons are stored in airtight
containers to prevent contact with humidity or airborne contaminants which could affect the corrosion
rate during testing or cause premature corrosion during storage. Corrosion rates were evaluated by
exposing HS80 or HS110 steel coupons to 14 wt. % and 28 wt. % HCl for 3 hours at atmospheric
pressure. The temperature was maintained at 78±1°C. In each test, one coupon was placed in a 400 ml
glass jar containing 200 ml preheated acid. Note that the surface area of each test coupon was
determined by accurate measurement of its dimensions. Roughly 7 ml acid solution was used per
square centimeter of coupon surface (200 ml acid per coupon). Corrosion inhibitor A was used for
evaluation of inhibited acid corrosion rates. For un-inhibited acid tests, the hydrochloric acid basefluid
was used without additives. Since the tests were designed to simulate oilwell stimulation, no effort was
made to remove dissolved oxygen from the acid. An argon blanket was used to stop additional intrusion
of oxygen during the 3 hour exposure period.

Description Pitting Index


None 0
Minor edge corrosion 1
Pitting on edge only 2
Pin point pits on surface <25 3
Pin point pits on surface >25 4
Table 1
Pitting index
Following each test, the coupons were rinsed in acetone and scrubbed with soap and water to remove
the residual inhibitor film and corrosion deposits. A final rinse in acetone was completed prior to re-
weighing the metal coupons to calculate the weight loss and corrosion rate. The coupons were
examined visually for localized corrosion, observed as pits or edge attack. The maximum allowable
corrosion rate for these tests, as recommended by internal procedures, is ≤0.05 lb/ft2 (0.244 kg/m2)
weight loss during the test period and a pitting index (P.I.) ≤2 (and preferably zero) as described in
Table 1.
Electrochemical analysis
All the electrochemical measurements were carried out in the potentiostatic mode using a three-
electrode cell arrangement. An Autolab(1) PGSTAT 302N Potentiostat combined with multiplexer was
used. The setup comprises a 0.5 L glass container with a thermostatic water-jacket and a glass top
designed to be fitted with Ag/AgCl (3M KCl) reference electrode, a graphite counter electrode, a
thermometer and a gas two-way purge tube. A fine Luggin capillary was placed close to the working
electrode to minimize ohmic resistance effects. The experiments were carried out under static working
electrode conditions in de-aerated solutions saturated with nitrogen. The temperature was maintained
at 78±1°C. Magnetic stirring was used to give an even temperature distribution and good mixing of the
inhibitor. Nitrogen saturation was initiated 30 minutes prior to the tests and was continued throughout
the experiment.
A 1.4 mm thick, 16 mm diameter, flat, circular steel disc with exposed surface area of 1.0 cm2 mounted
in holder was used as the working electrode. No crevice corrosion was observed under the Teflon O-
ring used to mount the disk in the holder. Before placing the electrode into the test solution (14 wt. %
HCl containing various concentrations of corrosion inhibitor), the metal sample disks were ground with
600 grit SiC paper, washed in DI water, rinsed with acetone and water and dried at room temperature.
Potentiodynamic polarization studies

(1)
Autolab products are available from Metrohm Autolab UK.
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
Potentiodynamic polarization tests were carried out to evaluate the corrosion current (icorr), open circuit
potential (OCP or Ecorr), and Tafel slopes of cathodic (bc) and anodic (ba) curves. The potentiodynamic
polarization was carried out after the linear polarization test. Tafel curves were obtained by changing
the electrode potential automatically from OCP-200 mV to OCP+400 mV at a scan rate of 1 mV/s. The
corrosion current was conventionally determined by extrapolation of the cathodic Tafel slope to Ecorr.
The corrosion rate (CR) and the inhibitor efficiency was calculated using the following equations:

CR = icorr × WE / t × A × d, (1)

I.E. (%) = (1 - icorr / i*corr) × 100, (2)

where i*corr and icorr are the corrosion current in the absence and presence of corrosion inhibitors,
respectively, and t, W E, A, d are time, equivalent weight of iron, sample area and density, respectively.
Linear polarization studies
Linear polarization measurements were carried out in a potential range ±5 mV with respect to the OCP
at a scan rate of 0.2 mV/s. Polarization resistance (Rp) was determined from the slope of the potential
versus the current. The Tafel slopes determined from the potentiodynamic polarization tests and the
polarization resistance was used to calculate the corrosion current using the Stern-Geary equation:4

Icorr = ((βa × βc) / (2.303 (βa + βc)) × 1 / Rp. (3)

The corrosion current can be converted to weight loss using Faraday’s law. The inhibitor efficiency
(I.E.) was calculated from the following formula:

I.E. (%) = (1 - R*p / Rp) × 100, (4)

were R*p and Rp represent polarization resistance values for uninhibited and inhibited solutions,
respectively. Typically, a time interval of 60 minutes was used to reach the steady-state OCP.
Materials
Two types of Tenaris coiled tubing with outer diameter 1.5 in and wall thickness 0.156 in were used to
prepare the coupons for weight loss analysis and the disks for electrochemical testing. HS80 contains
0.1-0.15% C, 0.6-0.9% Mn, <0.03% P, <0.005% S, 0.3-0.5% Si, 0.45-0.7 Cr, <0.4 Cu and <0.25% Ni.
HS110 contains 0.1-0.15% C, 0.6-0.9% Mn, <0.025% P, <0.005% S, 0.25-0.4% Si, 0.55-0.7% Cr,
<0.4% Cu, 0.14-0.3% Ni and 0.25-0.45% Mo. In each case, Fe makes up the remainder. HS80 and
HS110 are low carbon steels with a ferrite/pearlite microstructure. Ferrite appears as the dark phase in
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
the Scanning Electron Microscope (SEM) images (Figure 1) while pearlite (composed of alternating
ferrite/cementite layers) appears as the light phase. The HS80 grain size is in the range 5-10 µm. In
contrast, the HS110 grain size is <5 µm. The two steels have a different chemical composition; notably,
HS110 contains 0.25-0.45 wt. % molybdenum.

Figure 1: SEM images of HS80 and HS110 microstructures after chemical etching in 2% nital.
Analytical reagent grade 14 wt. % HCl (1.069 g/cm3 at 20°C) was supplied from Aldrich. 28 wt. % HCl
solution was prepared from an analytical reagent grade of HCl (37 wt. %) and Millipore DI water. All
inhibitors, including propargyl alcohol (PA), 4-ethyl-1-octyn-3-ol (EOCT), 1-octyn-3-ol (OCT), n-
dodecylpyridinium chloride (DPC), benzyldimethylhexadecylammonium chloride, 2,2’-biquinoline,
tripropargyl amine, 3-butyn-1-ol and 3-octyn-1-ol, were purchased from Aldrich and used without further
purification.
Acid corrosion inhibitor A is designed for corrosion inhibition in most hydrochloric acid systems (5 to 28
wt. %) and in “Mud Acids” (mixtures of hydrochloric and hydrofluoric acid) when conventional carbon
steels (N80, J55, L80 casing and HS80, HS90, HS110 coiled tubing) and alloys such as 13Cr are
encountered at temperatures up to 149°C. The chemistry of corrosion inhibitor A includes acetylenic
alcohols and surfactants.
Surface examination studies
A Mitutoyo Surftest SJ-400 contact profilometer was used to determine surface roughness parameters.
The roughness of the inner or outer faces of weight loss coupons was assessed by multiple 12.5 mm
line measurements taken along the longitudinal axis of the parent tubing. The stylus velocity was 0.1
mm/s. Each 12.5 mm line consists of 10000 x-z data points, i.e. on average, the value of z is recorded
after each 1.25 µm movement of the stylus. Typically, between five and nine 12.5 mm line
measurements were recorded on each face of the test coupon. The parameter Ra (the arithmetic mean
of the deviation of the profile from the mean line) was used to quantify surface roughness.
RESULTS AND DISCUSSION
Weight loss analysis
Several significant parameters should be considered closely when developing a coupon weight loss
method. Firstly, we need to consider the ratio between the acid volume and the coupon surface area
(V/A). Most published experimental procedures perform weight loss corrosion experiments in a static
environment with a limited fluid volume. In the oilfield, the volume of acid per coupon surface area is
usually standardized at a value matching the V/A ratio for the pipe used in the operation.5,6 For
example, a V/A ratio of 21 ml/in2 (3.4 ml/cm2) would simulate acidizing through a 5.5 in diameter pipe.
The latter V/A ratio relates to the internal volume of the acid in contact with the inner diameter of the
pipe but this ratio does not take into account the real case wherein the inner surface experiences
contact with freshly pumped acid in a dynamic environment.
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
The V/A ratio can affect the coupon weight loss results. Firstly, the availability of oxygen may depend
on the acid volume. If the corrosion rate is high and/or the V/A ratio is too low, the concentration of
hydrochloric acid may change during the experiment due to acid consumption. This can be a significant
problem when the metal coupon is tested in uninhibited acid or when only small amounts of corrosion
inhibitor are present. In such cases, acid consumption leads to a lower cumulative weight loss.
An analysis of weight loss data for a HS80 (BB) coupon immersed in uninhibited hydrochloric acid
(initial concentration 28 wt. %) showed that the acid concentration was reduced to 19 wt. % by the end
of a 3 hour test. Since the corrosion rate increases exponentially with HCl concentration,7-9 the
cumulative weight loss ~1.2 lb/ft2 (5.859 kg/m2) determined in this test is grossly underestimated due to
acid consumption. In contrast, when the corrosion rate is ≤0.05 lb/ft2/3 hrs (0.244 kg/m2), only a very
small change in the hydrochloric acid concentration occurs (from 28 wt. % to ≤27.6 wt. %). This is
considered to be a quasi-stationary state and so the resultant cumulative weight loss results are
reliable. Thus, all corrosion rates exceeding 0.05 lb/ft2 (0.244 kg/m2) per test period may not be
accurate and should be used for rough estimates. An important consequence is that the high corrosion
rates of coupons tested in uninhibited acid solutions should not be used as a baseline to calculate
inhibitor efficiency. In extreme cases, the inhibitor efficiency may be strongly overestimated. This is a
serious limitation of the weight loss method.

{3}
Weight loss, lb/ft /3 hs

{3}
0.4
2

HS110 BB
HS110 pickled
{7-8}

0.2

{7}
{6-7}

{3} {3} {3}


{0}
{4} {0}
0.0 {3} {1} {0-2}

0.00 0.05 0.10 0.15


Inhibitor A, wt. %
Figure 2: Effect of inhibitor A concentration and surface texture/finish of HS110 (BB; “as received” (AR)
and pickled in 14 wt. % HCl for 15 min at 78°C) on coupon weight loss and pitting index (number in
brackets) tested in 28 wt. % HCl at 78°C. Dotted line shows maximum weight loss from specification.

ASTM G31 recommends the use of large enough volumes of the test solution to avoid any appreciable
change in its corrosivity during the test, either by exhaustion of corrosive constituents or by
accumulation of corrosion products. ASTM recommended practice is to use up to 0.4 L/cm2, which
would correspond to about 12 L acid solution for a typical weight loss coupon used in this study. This is
not a practical volume for routine laboratory testing.
The important effect of the presence and concentration of oxygen on the rate of metal dissolution under
acidic conditions has been reported in the literature. In certain cases, the corrosion rate was observed
to increase in the presence of oxygen.10 Our tests indicate that, in the absence of oxygen, the acid
volume to metal surface area ratio (varied in the range 3.4 – 20.7 ml/cm2) has a negligible effect on the
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
cumulative weight loss. However, in the presence of oxygen, the cumulative weight loss increases
when the acid volume to metal surface ratio is increased through the same range. In the latter case, the
oxygen to metal surface area ratio increases with acid volume. An additional effect may be observed in
certain inhibited acid solutions wherein the stability and/or efficiency of the inhibitor molecules is
affected by the presence and concentration of oxygen.
Some typical results of our corrosion tests are shown in Figures 2 and 3. Figure 2 compares the weight
loss and pitting behavior of several bead blasted and as received and pickled HS110 coupons after
immersion in 28 wt. % hydrochloric acid at 78°C in the absence and presence of corrosion inhibitor A.
Figure 3 shows the effect of acid concentration and inhibitor concentration on the corrosion rate of bead
blasted HS80 coupons. The ordinate of each graph shows the cumulative weight loss per unit surface
area for a 3 hour test. Error bars on selected data points show the variation of replicate tests. The
spread in the weight loss values for a given testing condition does not exceed ±10% of the average
when the pitting index is ≤3.
A systematic decrease in the corrosion rate of HS110 (BB) is observed with increasing inhibitor A
concentration (Figure 2); these data show that ≥0.035 wt. % inhibitor A is required to ensure a weight
loss <0.05 lb/ft2 (0.244 kg/m2). When this critical inhibitor concentration is exceeded, the cumulative
weight approaches the x-axis asymptotically indicating that a saturation concentration is achieved. Data
such as these are used to determine the concentration of inhibitor A which is required for a 3 hour
acidizing treatment (28 wt. % HCl) of a reservoir with temperature 78°C. Several other parameters such
as the fluid heat-up profile are also considered.

1.0
Weight loss, lb/ft /3 hrs

{0}
2

28% HCl
15% HCl
0.5
{9}
{2}

{4-7}
{4} {3} {1} {0}
0.0
{0}
0.00 0.07 0.14
Inhibitor A, wt. %
Figure 3: Effect of acid concentration (14 and 28 wt. % HCl) and inhibitor concentration on weight loss
and pitting index (number in brackets) for HS80 BB coupons. Dotted line shows maximum weight loss
from specification.
Note that the surface texture and preparation of the specimen has a significant influence on the weight
loss results (Figure 2). In our initial development studies we chose four different surface finishes to
assess the effects of surface texture and roughness on the weight loss and pitting index results. The
following surface texture/roughness conditions were examined in inhibited 14 or 28 wt. % HCl:
1) Glass bead blasted coupons: As mentioned previously, test coupons are cut from oilfield tubing and
are prepared by standard machining operations. After this step, the roughness of the edges and
faces of the coupon may be different. The cut edges could become sites of preferential attack.
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
Therefore, glass bead blasting is used to obtain a uniform surface roughness of all the coupon
surfaces. We measured similar roughness of the outer and inner surfaces of an HS80 BB coupon
(outer: Ra = 5.55±9% µm and inner: Ra = 3.13±9% µm). HS80 BB and HS110 BB coupons have very
similar texture and roughness allowing us to compare the performance of inhibitors on the two metal
types.
2) Pickled coupons: These were prepared by cutting to the same dimensions but they were not glass
bead blasted. Prior to use, these coupons were pickled in 14 wt. % HCl at 78°C for 15 min to remove
surface corrosion products. Measurements showed that the HS110 BB coupons are significantly
smoother than the pickled AR HS110 coupons. Average values of Ra are 7.9±9% µm (outer
diameter) for the pickled AR HS110 coupons.
3) Coupons with 240 grit surface finish: These were prepared from AR coupons which were flattened
and then the surface was ground with a 240 grit SiC abrasive paper. Average Ra values were
0.15±4% µm.
4) Coupons with mirror polished finish: Again, AR coupons were flattened and then the surface was
ground with 600 and 1200 grit abrasive SiC papers followed by polishing with silica spheres (6 µm
diameter). Average values of Ra were 0.045±6% µm.
The surface texture and roughness of the test coupon may affect the corrosion rate, the pitting
behaviour and the adsorption of inhibitor due to differences in the surface area of the specimen and
differences in surface stress. We found that, especially at low concentrations of inhibitor A, the surface
texture and roughness has a strong influence on the corrosion rate (Figure 2). Glass bead blasted
coupons are more reactive than the pickled coupons. Glass bead blasting introduces stresses, plastic
deformation and microstrains and changes in the heterogeneity of the surfaces may occur via
fragmentation of grains. Reddy et. al. concluded that the residual stress stored within the material after
bead blasting acts as a source of activation energy and thus enhance corrosion rate.11 Our results
confirm the important effect of residual stress due to glass bead blasting. The samples with the highest
Ra range (pickled coupon, Ra=7.9 µm) show a lower corrosion rate than the glass bead blasted
samples (Ra=5.55 µm). However, comparable weight loss data were obtained for the glass bead
blasted and acid pickled coupons when corrosion inhibitor A concentrations exceeded 0.14 wt. %
(Figure 2). Thus, when there is a sufficient concentration of inhibitor molecules available, the corrosion
is very effectively inhibited and no significant effect of the surface texture/roughness is observed.

(b)
BB

(a)
0.008
Bead Blasted 0.075
Weight loss, lb/ft /3 hr

{7}
Pickled
2
Corrosion rate, lb/ft

{0}
2

0.006 Grit 240


Mirror polished

Mirror polished 0.050

0.004
240 rgit
Pickled

{0}
0.025
{0}
0.002 {3}
{0}
{0} {0}

0.000
0.000
0.00 0.02 0.04 0.06
Inhibitor A, 0.05 wt. %
Inhibitor A, wt. %

Figure 4: Effect of inhibitor A concentration and surface finish of HS80 (BB; “as received” and pickled
with 14 wt. % HCl for 15 min at 78°C; 240 grit and mirror polished) on weight loss tested in 14 wt. %
HCl (a) and 28 wt. % HCl (b) at 78°C. Pitting index given in brackets.
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
Figure 4 (b) shows that, when 0.05 wt. % corrosion inhibitor A is present, the corrosion rate in 28 wt. %
HCl follows the order: BB > Pickled ≅ 240 grit ≅ Mirror polished. This indicates that, under these
conditions, the effect of surface stress is more important than surface roughness. The BB and Pickled
coupons attain respective pitting indices of 7 and 3 but no pits are formed on the 240 grit and Mirror
polished samples. Figure 4(a) shows the results obtained for 14 wt. % HCl solutions. In this case,
throughout the inhibitor concentration range 0.01-0.05 wt. %, the cumulative weight loss follows the
order: BB > 240 grit > Pickled > Mirror polished and no pits are formed on any of these samples. This
indicates that both surface stress and roughness play a role when the basefluid is 14 wt. % HCl.
Electrochemical analysis
Figure 5 illustrates the procedure selected to measure the effectiveness of inhibition as a function of
inhibitor concentration by the linear polarization method. The experiment begins with a 30-60 min.
treatment in 14 wt. % HCl without corrosion inhibitors in order to attain a stable value of OCP. During
this period, Ecorr changes most rapidly at the beginning and stabilizes after 60 minutes (Figure 5). We
followed the recommendation of Kelly et. al. to wait for a change of less than 5 mV in Ecorr over a 10
minute period before initiating LPR experiments.10 At this point the uninhibited corrosion rate is just over
1000 mm/yr, which was then used as a baseline for the calculation of inhibitor efficiency. The next step
in our procedure is the “inhibited stage”. After each successive addition of inhibitor, Ecorr is monitored
and a LPR measurement is performed every 10 min. This allows measurement of the instantaneous
corrosion rate at several successive inhibitor concentrations within a single experiment.

Acid stage Inhibited stage


100
0.01% 0.03% 0.05% 0.09% 0.15% 0.2%

-0.30
80
Ecorr, V

-0.33

icorr, mA
60

-0.36
40
-0.39
20
-0.42

0
0 60 120 180
Time, min
Figure 5: Multi-step procedure used to measure the instantaneous corrosion rate and inhibitor efficiency
by LPR. Both the corrosion current icorr and the corrosion potential Ecorr are monitored for each
successive concentration of inhibitor cocktail A. The data pertain to HS80 in 14 wt. % HCl.

As shown in Figure 5, as the inhibitor concentration increases, both Ecorr and Rp increase and the
corrosion current icorr decreases. This indicates that the coverage of the inhibitor on the sample
increases successively with increasing inhibitor concentration. The inhibitor efficiency (Figure 6)
increases within a narrow concentration range and at 0.05 wt. % the efficiency reaches its maximum
~96-98%. At 0.2 wt. % inhibitor A, the corrosion rate of HS80 is 4.6 mm/yr (equivalent to a cumulative
weight loss 0.0025 lb/ft2 (0.0122 kg/m2) in 3 hours). Corrosion inhibitor A is effective for both metal types
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
HS80 and HS100. However, HS110 requires less inhibitor to reach the maximum efficiency. A similar
trend was observed by weight loss analysis.
100

75

I.E., % 50 HS110
HS80
25

0
0.00 0.05 0.10 0.15 0.20

Inhibitor A, wt. %
Figure 6: Inhibitor efficiency of inhibitor cocktail A on HS80 and HS110 in 14 wt. % HCl at 78°C.

1 2
1
3
0.1
Current, A

0.01
Eu
1E-3

1E-4

-0.6 -0.4 -0.2 0.0


Potential applied, V

Figure 7: The polarization curves for HS80 in uninhibited 14 wt. % HCl solutions (1) and in the presence
of 0.053 wt. % (2) and 0.219 wt. % (3) of corrosion inhibitor A at 78°C.
Potentiodynamic polarization studies were carried out to evaluate the corrosion mechanism for inhibitor
cocktail A (this cocktail is a multicomponent blend containing polymerizable acetylenic alcohols,
surfactants and a solvent package). Figure 6 shows the polarization curves of HS80 in uninhibited 14
wt. % HCl and in the presence of two different concentrations of corrosion inhibitor A. As the inhibitor
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
concentration increases, both the anodic and cathodic currents are reduced. This suggests that
inhibitor A reduces anodic dissolution and also retards hydrogen evolution, i.e. a mixed type inhibition
mechanism. However, when the inhibitor A concentration is ≥0.01 wt. %, we note that the OCP shifts to
a more noble value which indicates a predominant effect on the anodic reaction.
Detailed analysis of the anodic region in the Tafel plot shows that within a certain anodic potential
range, the current increases steeply (Figure 7). The potential at the intersection of the two linear slopes
is termed Eu, the potential of unpolarizability.12 Eu is associated with the onset of desorption of
absorbed species. When E>Eu, the inhibitor coverage on the sample decreases sharply. For the case
studied, the rapid increase of the anodic current at the second polarization region may be due to
desorption of the polymeric film.
As mentioned previously, film forming organic inhibitors can be classified by their predominant metal-
inhibitor interaction: physi-sorbed inhibitors, chemi-sorbed inhibitors and polymerizable inhibitors. The
inhibitor film restricts diffusion of ions or molecules to or from the metal surface and thus retards the
rate of corrosion reactions.13 More detailed mechanisms consider blocking of cathodic or anodic
reaction sites, participation in electrode reactions, e.g. the hydrogenation reaction of acetylenic
alcohols,14 and/or alternation of the electric double layer that forms at the metal/solution interface.15
Surfactants are a major class of film-forming corrosion inhibitors.16,17 Surfactants adsorb as a result of
electrostatic attractive forces between inhibiting organic ions or dipoles and the electrically charged
surface of the metal. Physi-sorption is influenced by the surface charge of the electrical double layer,
the chemical structure of the surfactants, the composition of the electrolyte and temperature.
100

75
I.E., %

~1 mM
50

25 N+

Cl-

0 3 6 9
n-dodecylpyridinium chloride, mM
Figure 8: Inhibitor efficiency for HS80 in 14 wt. % HCl containing different concentrations of n-
dodecylpyridinium chloride at 78°C. The CMC of DPC is 1.3±0.2 mM (as determined by Dynamic Light
Scattering in 14 wt. % HCl at 25°C).
Figures 8 and 9 show the behavior of a typical cationic surfactant, n-dodecylpyridinium chloride (DPC),
on HS80 in 14 wt. % hydrochloric acid at 78°C. The data reveal linear trends with different slopes below
and above the critical micelle concentration (CMC) of the surfactant. The critical concentration between
these two linear regions is ~1.0 mM which is similar to the CMC (1.3±0.2 mM) determined by Dynamic
Light Scattering (DLS). Below this critical concentration, the cationic surfactant molecules physi-sorb on
the oppositely charged steel surface providing a low inhibitor efficiency due to a surface coverage
below ~50%. As the concentration of surfactant increases, the charge at the surface becomes
neutralized. At the critical concentration, the hydrophobicity reaches a maximum as the hydrophobic
groups orientate away from the polar surface into the bulk solution.18 When the surface charge is
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
completely neutralized, the system lowers the Gibbs free energy by creating a double layer on the
surface. This double layer is more hydrophilic as the hydrophilic head groups now form the interface
with the bulk solution. A decrease in hydrophobicity has been attributed to bilayer adsorption of DPC on
the substrate.19 Generally, the adsorption of cationic surfactants results in a more positive corrosion
potential Ecorr (anodic type inhibitor) and retards the discharge of positively charged hydrogen ions.

100
1
75 2
I.E., %

3
50

25

0.00 0.08 0.16


Benzyldimethylhexadecylammonium chloride, wt. %

Figure 9: Inhibitor efficiency for HS80 in 14 wt. % HCl containing different concentrations of
benzyldimethylhexadecylammonium chloride at 40°C (1), 60°C (2) and 78°C (3).
The inhibitor efficiency of the cationic surfactant benzyldimethylhexadecylammonium chloride at various
temperatures is shown in Figure 9. The inhibitor efficiency increases with inhibitor concentration but
decreases with increasing temperature. Figure 10 plots the logarithm of the corrosion rate in the
absence and presence of benzyldimethylhexadecylammonium chloride as a function of reciprocal
temperature. The data for uninhibited and inhibited solutions are characterized by a linear dependence
of the corrosion rate on reciprocal temperature and hence can be described by an Arrhenius equation:
CR = A× exp (Ea / RT),
where CR is corrosion rate, A is the pre-exponential frequency factor, Ea is the apparent activation
energy, R is the molar gas constant and T is absolute temperature. The apparent activation energy Ea
is 37±6 kcal/mol (157±26 kJ/mol) for uninhibited 14 wt. % hydrochloric acid and 56±5 kcal/mol (237±21
kJ/mol) for the acid solution inhibited with benzyldimethylhexadecylammonium chloride at a
concentration above its CMC. The dependence of the corrosion rate on temperature is similar to that
reported for a low carbon steel (0.14% C) in 14 wt. % HCl at 30-110°C by Sathiya Priya et al..20 The
behavior of 5.02 mM benzyldimethylhexadecylammonium chloride in 14 wt. % HCl is characteristic of a
physi-sorbed inhibitor21 as the value of the apparent Ea is greater than that obtained in the uninhibited
solution.

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
8

lnCR, mm/yr
4
1

2
0
0.0028 0.0030 0.0032
-1
1/T, K
Figure 10: Arrhenius plot for HS80 in 14 wt. % HCl in the absence (1) and in the presence (2) of
benzyldimethylhexadecylammonium chloride (5.02 mM).

Chemi-sorbed molecules constitute the second class of inhibitors studied. In this case, the adsorption
process involves physi-sorption driven by van der Waals forces and further stabilisation of the adsorbed
molecules by chemi-sorption. Chemi-sorption takes place more slowly than electrostatic adsorption and
with a higher activation energy. When the binding process is dominated by chemi-sorption, higher
degrees of inhibition should be expected at higher temperatures. Chemi-sorption involves electron
transfer from electron-rich sites within the structure of the inhibitor molecule(s) to vacant low energy
orbitals in the metal. Typically, such electron-rich sites are heteroatoms with lone pair(s) of electrons,
multiple bonds and aromatic rings (π-electrons). The strength of the adsorption bond formed by a
heteroatom is determined by its electron density and polarisability. The inhibition efficiency of a
homologous series of organic molecules differing only in heteroatom type should follow the order:
P>Se>S>N>O due to the increase in electronegativity and decrease in polarisability through this series.
Figure 11 compare the efficiency of some commercially available inhibitor molecules which are capable
of chemi-sorption. These are water soluble 2,2’-biquinoline, tripropargyl amine and 3-butyn-1-ol and the
oil soluble 3-octyn-1-ol. Potentiodynamic polarization studies show that these compounds suppress
both the anodic and cathodic process, i.e., they are mixed-type inhibitors. All these inhibitors show
increased efficiency with increasing concentration and inhibitor efficiency decreases in the order 2,2’-
biquinoline > tripropargyl amine >3-butyn-1-ol > 3-octyn-1-ol.
The high efficiency of 2,2’-biquinoline is probably due to the aromatic rings (π-electrons) and the
presence of heteroatoms (nitrogen). The maximum efficiency of the water soluble acetylenic
compounds, tripropargyl amine and 3-butyn-1-ol, is in range 81-85% compared to 91% for 2,2’-
biquinoline. 2,2’-biquinoline and 3-butyn-1-ol are soluble in acid forming onium cations –OH2+ and
=NH+- that can be electrostatically adsorbed. Adsorbed onium ions are discharged on the negative
charged surface forming neutral molecules that can be adsorbed by donating their lone electron pair.
This results in an increase of the d-character of iron.22 The inhibition efficiency of acetylenic alcohols
depends on the position of the triple bond. The low efficiency of 3-octyn-1-ol (70% at 31.7 mM) is
attributed to a lower electron density around the adsorption center due to the C2–C3 position of the

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
triple bond1 and to the poor solubility of 3-octyn-1-ol in acid. Overall, the efficiency of tripropargyl amine
and 3-butyn-1-ol is lower than that of the polymerizable inhibitors, as discussed in paper.23
2,2'-Biquinoline
100
N
N

75
Tripropargyl amine
I.E., %

50
N

3-Butyn-1-ol
25
HO

3-Octyn-1-ol
0
OH
0 10 20 30
C.I., mM
Figure 11: Inhibitor efficiency of chemi-sorbed inhibitors on HS80 in 14 wt. % HCl at 78°C.

For some acetylenic alcohols, α-alkenylphenones and α,β-unsaturated aldehydes, surface initiated
polymerization is possible during the corrosion inhibition process. Many studies have shown that there
is no alternative to polymerizable inhibitors for protection of oilwell equipment during acid stimulation.
By comparison, physi-sorbed cationic surfactants readily desorb from the metal surfaces at
temperatures above 40 to 80°C.
The inhibition mechanism of polymerizable acetylenic compounds has been discussed extensively for
more than 50 years. The current concept of the inhibitive effect of acetylenic compounds on corrosion
can be divided into two steps, namely, chemi-sorption of acetylenic derivatives on the metal surface
followed by the formation of a protective film by polymerization on the active metal surface14,24-29 (Figure
12). Such a reaction/polymerization is surface-catalyzed.
The polymer films formed by acetylenic alcohols on carbon steel surfaces were studied by ex-situ
ellipsometry in the dry condition. Propargyl alcohol (0.2 wt. %) forms a hydrophobic film on HS80 with a
thickness of 11.3±0.9 Å after submersion in 14 wt. % HCl at 78°C for 2 hours. This corresponds to an
equivalent thickness of nearly 3 monolayers of the inhibitor molecules. Similar results were obtained in
the literature for propargyl alcohol and other polymerizable acetylenic alcohols:30 film thicknesses range
from a monolayer formed in a few minutes to films up to 50Å thick after 24 hours. A film thickness ≈50Å
is reported to be the maximum film thickness possible.
XPS was used to investigate the chemical composition of HS80 surfaces. The composition of the
surface without inhibitor treatment was compared with that of the surface after treatment with propargyl
alcohol and after treatment with a combination of propargyl alcohol and cationic surfactants. Higher
carbon was found on the surface treated with inhibitor(s). Since in XPS only the top 10 nm of the
surface is analyzed, the high carbon signal from the inhibitor layer obscures the signal from the iron
within the steel. The iron signal increases after successive etching of the inhibitor film by Ar+ sputtering.
This allows us to obtain a depth profile of the inhibitor film composition.

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
Direct electroinitiated polymerization of
An anion radical was considered to be generated by
PA on the surface 24,27 the cathodic process of PA adsorbed on the surface,
followed by growth of the PA .
“Ads” denotes an adsorbed species at the surface.
OH Polymer
ads

e-
Electrochemical (cathodic ) hydrogenation of PA26,29
HC CCH2OH + 2H+ + 2e- Polymer
OH OH
Allyl alcohol

Hydrogenation: Allyl alcohol is formed by cathodic reduction of PA


on the Fe surface in the acid solutions

The surface polymerization of a species derived from PA25,26

H 2O
HC CCH2OH
Formation of allene as an intermediate of PA to polymerize
HO OH
on the steel surface in HCl through hydrogenation and
dehydration processes is observed.

- H2O

O OH O Polymer
Acrolein

Figure 12: Mechanisms of chemical transformation and polymerization of propargyl alcohol.

100

75
I.E., %

OH

50
1-Octyn-3-ol
OH

4-ethyl-1-octyn-3-ol
25
Propargyl alcohol
HO

0 10 20 30

C.I., mM
Figure 13: Inhibitor efficiency of polymerizable acetylenic alcohols on HS80 in 14 wt. % HCl at 78°C.

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
Figure 13 shows the inhibitor efficiency of some commercially available polymerizable acetylenic
alcohols: oil soluble 1-octyn-3-ol and 4-ethyl-1-octyn-3-ol and water soluble propargyl alcohol. The
strong inhibition efficiency of polymerizable acetylenic alcohols is observed. The results show that
inhibitor efficiency increases with increasing chain length of the polymerizable acetylenic alcohol. The
inhibitor efficiency of 2.4 mM OCT and of 8.4 mM EOCT is >99 %. Propargyl alcohol is less
hydrophobic and shows a maximum inhibitor efficiency of ~98 % at 36 mM (0.2 wt. %).
Podobaev et. al. report that chemi-sorbed, polymerizable acetylenic alcohol inhibitors can give
acceptable corrosion control in the temperature range up to 100-140°C in 4M hydrochloric acid and,
evidently, inhibitor efficiency increases with increasing temperature in the range 25-100°C.31
It is very rare that a single inhibitor is used to control corrosion in aggressive media. Usually, a variety
of inhibitors are combined to achieve a more efficient performance. The most useful inhibitor products
and, in particular those with good temperature tolerance, make use of a combination of (i) synergistic
effects, (ii) reactions at the metal surface to form “secondary” inhibitors and polymeric films and (iii)
intensifiers or inhibitor aids which show no inhibition when used alone but which enhance the
performance of inhibitors and their mixtures. For example, the combination of acetylenic alcohols with
quinoline-based quaternary ammonium compounds, a surfactant and formic acid gives acceptable
corrosion control at 104-177°C (220-350°F).32,33 Thus synergistic blends are produced by mixing
several components.
To evaluate the synergistic effect between the two classes of inhibitors, e.g. polymerizable acetylenic
alcohols and a surfactant, the synergism parameter introduced by Aramaki and Hackerman34 can be
used. The synergistic parameter S is defined as the ratio between the theoretical decrease (CRth) and
the actual (CR*) decrease of the corrosion rate:

S = (CRth /CR0) / (CR*/CR0) = (1-θth) / (1-θ*) = (1 - (θ1 + θ2) + θ1 × θ2) / (1-θ*), (5)

where θi = I.E.i /100. The synergistic parameter S approaches unity when the two components either do
not interact with each other or the interactions lead to opposite mechanisms. For the case S>1, the
synergistic effect is positive and the corrosion rate is decreased by a factor of S. In the case of 0<S<1,
an antagonistic interaction predominates. This measurement of the synergistic parameter cannot be
related to kinetic or thermodynamic values of the system as its definition is formal. Indeed, an accurate
estimation of the theoretical decrease of the corrosion rate without interaction between the compounds
would need a theoretical and reliable model of the mechanisms of adsorptions of the two compounds.
The synergistic effect is not directly related to the inhibitor efficiency. Figure 14 shows that a blend of
two molecules, EOCT and DPC, at a total inhibitor concentration of 1.4 mM, attains an inhibition
efficiency 92%. The corresponding synergetic parameter is 5.6. In contrast, the blend at total inhibitor
concentration 5.7 mM shows I.E.~97 % and S=1.8. Thus, a strong synergistic effect does not
necessarily correspond to a higher inhibitor efficiency.
It was reported in the mid-1980s that cationic surfactants improve the inhibitive performance of
acetylenic alcohols. A simple model proposed by Frenier and Growcock28 can be used to predict this
synergy. This model is based on (i) solubilization of the oil-soluble acetylenic alcohol in acid solution,
and (ii) better surface coverage due to co-adsorption of surfactant and inhibitor on an iron interface.
The solubility of some polymerizable corrosion inhibitors is limited to the low millimole range. Generally,
as the molecular weight (or, correspondingly, the hydrophobicity) of polymerizable acetylenic alcohols
increases, their solubility in aqueous systems decreases. Therefore, their inhibition efficiency is
enhanced by the presence of an appropriate solubilizer. Two different types of polymerizable inhibitors
were chosen in order to examine the role of surfactant – the water soluble propargyl alcohol and the
©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
predominantly oil soluble 4-ethyl-1-octyn-3-ol. The solubility of the latter is less than 0.05% (0.32 mM)
in 14 wt. % HCl at room temperature.35
100 10

3 8
75
I.E., % 1

Synergy, S
6
50 2
4

25
2
4
0 0
0.0 1.5 3.0 4.5 6.0
Total concentration, mM
Figure 14: Inhibition efficiency of EOCT (1), DPC (2) and the blend EOCT-DPC containing 35 mol. %
DPC (3) and the synergistic parameter of the blend (4). HS80 in 14 wt. % HCl at 78°C

Prior to electrochemistry experiments, DLS measurements were made on 14 wt. % HCl solutions
containing a range of concentrations of EOCT (0.1-2 mM) and two concentrations of DPC (one below
and one above the CMC) at 25°C. At a concentration of 0.36 mM DPC (below the CMC), the blends
form an oil-in-water emulsion composed of large droplets (700±100 nm diameter) for the entire range of
EOCT concentration. In this case, the surfactant is partly adsorbed on the EOCT(oil)/water interface. At
a concentration of 1.42 mM DPC (above the CMC), the oil soluble EOCT is very effectively solubilized
in micelles. The average diameter of these micelles increases with EOCT concentration: from 470 nm
at [EOCT]=0.64 mM to 890 nm at [EOCT]=3.89 mM.
Figure 15 shows synergistic parameter maps for blends of 4-ethyl-1-octyn-3-ol (Figure 15a) or
propargyl alcohol (Figure 15b) with DPC. The DPC surfactant greatly improves the performance of both
acetylenic alcohols. For the EOCT-DPC blends, strong synergy (S>3) occurs at a concentration range
of 0.3-0.6 mM DPC. The maximum synergistic parameter, Smax~6 occurs at a EOCT:DPC ratio 1:0.5
(mM:mM). For the PA-DPC blends, a significant synergistic effect (Smax=1.8 at 1.4 mM DPC) was found
despite the fact that propargyl alcohol is completely soluble in the acid. This indicates that the
solubilisation effect does not provide a complete explanation of the synergy in the studied inhibitor-
surfactant blends. Nevertheless, the solubilization of oil-soluble inhibitors is required to obtain
homogeneous inhibitor cocktails and may affect other aspects of the inhibition process, e.g. the inhibitor
adsorption kinetics, the inhibitor film persistence, stability of inhibitor in solution, etc.
Another aspect to be considered is the co-adsorption of surfactant and inhibitor on the metal surface.
As already mentioned, surfactants, such as DPC, increase the hydrophobicity of steel by physi-sorption
on the negatively charged chloride-covered metal surface. Such a hydrophobic surface can attract the
inhibitor and enhance its adsorption.7,16 The maximum in the synergistic parameter corresponds to a
composition which can maximize the hydrophobicity of the surface. For the two cases studied, this
composition corresponds to a surfactant concentration below the CMC (EOCT/DPC blends) or close to
the CMC (PA/DPC blends). For a constant concentration of the acetylenic alcohol, the synergistic effect
increases with increasing DPC concentration but then decreases when the DPC concentration is above

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
the CMC. Thus, at DPC concentrations above the CMC, the polymerizable inhibitor competes with the
surfactant for adsorption sites on the metal surface.

6 10 0.7
(a) 1.0 1.1 (b)
1.0
4-ethyl-1-octyn-3-ol, mM

0.8

Propargyl alcohol, mM
8
1.0
1.2 0.8
0.9
4
1.5
6
1.1
2.0
1.0
1.3
2.5 4 1.0
1.7
2
3.0 1.6
3.5
2
1.4
6.0 1.5
4.0 1.2

0 1.1
0
0 1 2 0 2 4
1-dodecylpyridinium chloride, mM 1-dodecylpyridinium chloride, mM

Figure 15: Synergistic parameter for EOCT:DPC blends (a) and PA:DPC blends (b). Dotted line shows
S=1. Filled circles are experimental data.

SUMMARY
This paper reviews current corrosion testing methods employed in the Schlumberger Cambridge
Research corrosion laboratory. These include gravimetric methods to determine the cumulative weight
loss and pitting behavior of coupons immersed in uninhibited and inhibited acid solutions such as 14-28
wt. % hydrochloric acid and electrochemical methods to determine linear polarization resistance, Tafel
behavior and inhibition efficiency.
In the development of our gravimetric methods, we investigated the effects of (1) coupon surface
texture and roughness, (2) acid volume to metal surface area ratio (V/A), (3) absence/presence and
concentration of oxygen, (4) acid concentration and (5) inhibitor concentration on cumulative weight
loss and pitting behavior. For the V/A ratio chosen in our experiments, cumulative weight loss
measurements are reliable when the weight loss during the test period is ≤0.05 lb/ft2 (0.244 kg/m2).
Regarding the surface texture and roughness of the test coupons, we observe that roughness seems to
influence the corrosion rate less than does the stress introduced by glass bead blasting.
Electrochemical methods were used to compare the inhibition performance of inhibitor products
(complex cocktails), pure inhibitors and synergistic blends in 14 wt. % HCl at 40-78°C. The pure
inhibitors were classified into physi-sorbed, chemi-sorbed and polymerizable types. Amongst these
inhibitor types, polymerizable inhibitors provide superior performance in strong hydrochloric acid
solutions at high temperature. Long-chain octynols perform better than propargyl alcohol. Blends of 4-
ethyl-1-octyn-3-ol (EOCT) and n-dodecylpyridinium chloride (DPC) yield synergistic parameters, S>3,
with a maximum Smax~6 at a EOCT:DPC ratio around 1:0.5 (mM:mM). By comparison, blends of DPC
with propargyl alcohol (PA) yield a maximum synergistic parameter Smax=1.8. The cationic surfactant
DPC solubilizes EOCT but the solubilisation effect alone does not provide a complete explanation of
the synergy in the studied inhibitor-surfactant blends. For the two cases studied, the maximum synergy
is obtained at a surfactant concentration below its CMC for the EOCT/DPC blends or close to its CMC

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
for the PA/DPC blends. For a constant acetylenic alcohol concentration, the synergistic effect increases
with increasing DPC concentration but then decreases when the DPC concentration is above the CMC.
ACKNOWLEDGMENTS
The authors wish to thank Prof. Anne Neville (School of Mechanical Engineering , Leeds University,
UK) for valuable discussions and interns Eleonore Noussan and Theo Chevallier (ParisTech Institute of
Technology, France), and Lewis Neve (Nottingham University, UK) for their strong contributions.
REFERENCES
1. K. Babic-Samardzija, C. Lupu, N. Hackerman, A. R. Barron, “Inhibitive properties, adsorption and
surface study of butyn-1-ol and pentyn-1-ol alcohols as corrosion inhibitors for iron in HCl”, J. Mater.
Chem., 2005, 15, 1908.
2. W.W. Frenier, D.G. Hill, R. Jasinski, “Corrosion inhibitors for acid jobs”, Oilfield Review, 1989, July
1989.
3. R.J. Tedeschi, “Acetylenic corrosion inhibitors”, Corrosion, 1975, 31, 4, 130.
4. M.Stern, A.L. Geary, J. Electrochem. Soc., 1965, 112, 1025.
5. C.F. Smith, F.E. Dollarhide, N.J. Byth, “Acid corrosion inhibitors - are we getting what we need?”,
SPE 5644, presented at the SPE AIME 50th Annual Fall Technical Conference and Exhibition, Dallas,
Texas, USA Sept. 28-Oct. 1, 1975.; SPE Journal of Petroleum Technology, 1978, 737.
6. W.E. Billings, D. Morris, “Effect of acid volume and inhibitor quantity on corrosion of steel oil field
tubing in hydrochloric acid”, Corrosion – NACE, 1960, 17, 208.
7. F. B. Growcock, W. W. Frenier, “Kinetics of steel corrosion in hydrochloric acid trans-
cinnamaldehyde”, J. Electrochem. Soc.: Electrochemical Science and Technology, 1988, 135, 4, 817.
8. F. B. Growcock, R. J. Jasinski, “Time-resolved impedance spectroscopy of mild steel in
concentrated hydrochloric acid”, J. Electrochem. Soc.,1989, 136, 8, 2310.
9. A. A. Khadom, A. S. Yaro, A. A. H. Kadum, A. S. AlTaie, A.Y. Musa, “The Effect of temperature and
acid concentration on corrosion of low carbon steel in hydrochloric acid media”, American Journal of
Applied Sciences, 2009, 6 (7), 1403.
10. R.G. Kelly, J.R. Scully, D.W. Shoesmith, R.G. Buchneit, “Electrochemical techniques in corrosion
science and engineering”, Marcel Dekker Inc., New York, 2003, 54.
11. B.S.K. Reddy, B. Ramamoorthy, P.K. Nair, “Surface integrity aspects and their influence on
corrosion behavior of ground surfaces”, IE (I) Journal-PR, 2005, 86, 35.
12. K.E. Heusler, G.H. Cartledge, J. Electrochem. Soc. 108, 1961, 732.
13. “Handbook of corrosion engineering”, Ed. by Pierre R. Roberge, McGraw-Hill Professional, 2000,
842.
14. N. I. Podobaev, Ya. G. Avdeev, “A review of acetylene compounds as inhibitors of acid corrosion of
iron”, Protection of Metals, 2004, 40, 1, 7.
15. A. Raman, R. Quinones, L. Barriger, R. Eastman, A. Parsi, E. S. Gawalt, “Understanding organic
film behavior on alloy and metal oxides”, Langmuir, 2010, 26 (3), 1747.
16. G. Trabanelli, “Inhibitors – an old remedy for a new challenge”, NACE, Corrosion, Cincinnati, Ohio,
March 1991, 410.
17. M.A. Migahed, A.M. Al-Sabagh, “Beneficial role of surfactant as corrosion inhibitors in petroleum
industry: A review article”, Chem. Eng. Comm., 2009, 196, 1054.
18. A. Fan, P. Somasundaran, N. J. Turro, “Adsorption of alkyltrimethylammonium bromides on
negatively charged alumina”, Langmuir, 1997, 13, 506.

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.
19. P. Somasundaran, L. Zhang, “Adsorption of surfactants on minerals for wettability control in
improved oil recovery processes”, Journal of Petroleum Science and Engineering, 2006, 52, 198.
20. A.R. Sathiya Priya, V.S. Muralidharan, A. Subramania, “Development of novel acidizing inhibitors
for carbon steel corrosion in 15% boiling hydrochloric acid”, Corrosion, 2008, 64, 6, 541.
21. S.A Umoren, M.M. Solomon, I.I. Udosoro, A.P. Udoh, “Synergistic and antagonistic effects between
halide ions and carboxymethyl cellulose for corrosion inhibition of mild steel in sulfuric acid solution”,
Cellulose, 2010, 17, 635.
22. K. Kobayashi, K. Shimizu and M. Iida, “Structural effects of organic compounds as corrosion
inhibitors for hydrogen entry into iron in sulphuric acid”, Corros. Sci., 1993, 35, 1431.
23. B. Yang, N. G. Smart and J. O’M. Bockris, “Ellipsometric investigation of the adsorption of
acetylenic alcohols on iron”, Electrochim. Acta, 1992, 37, 317.
24. K. Aramaki, E. Fujioka, “Spectroscopic investigations on the inhibition mechanism of propargyl
alcohol for iron corrosion in hydrochloric acid at elevated temperatures”, Corrosion, 1997, 53, 4, 319.
25. F.B. Growcock, V.R. Lopp, Corros. Sci. 28, 1988, 379.
26. F.B. Growcock, V.R. Lopp, R.J. Jasinski, J. Electrochem. Soc. 135 (1988): p. 823.
27. Faraponov V.V., Grovu M., Simionescu C.I., “Electrochemical polymerization of acetylenic
derivatives. I. Anionic polymerization of phenylacetylene and diphenyldiacetylene”, J. Polym. Sci. 1977,
15, 2041.
28. W.W Fernier, F.B Growcock, “Mechanisms of corrosion inhibitors used in acidizing wells”, SPE
Production Engineering, 1988, 584.
29. Ya. G. Avdeev, N. I. Podobaev, “The role of acrolein in the inhibition of the acid corrosion of iron
with propargyl alcohol”, Protection of Metals, 2005, 41, 6, 592.
30. G.W. Poling, “Infrared studies of protective films formed by acetylenic corrosion inhibitors”, J.
Electrochem. Soc.: Electrochemical science, 1967, 114, 12, 1209.
31. Balezin, S.A., Podobaev, N.I., Voskresenskii, A.G., Vasil'ev, V.V., “Mechanism of the protective
action of acetylene compounds against corrosion of steel in hydrochloric acid”, Proc. III Int. Congr. on
Metal Corrosion, Moscow: Mir, 1966, 2, 7.
32. Frenier, W.W., Grannan, S.E., Hill, D.E., “Process and composition for inhibiting iron and steel
corrosion”, European Patent 0276879, 1988.
33. Jasinski, R.L., Frenier, W.W., “Process and composition for protecting chrome steel”, European
Patent 0471400 A1, 1992.
34. K.Aramaki, N. Hackermann, “Inhibition mechanism of medium-sized polymethyleneimine”, J.
Electrochem. Soc., 1969, 568.
35. J. Funkhouser, “Acid corrosion inhibition with secondary acetylenic alcohols”, NACE, Tulsa,
Oklahoma, 1960

©2012 by NACE International. Requests for permission to publish this manuscript in any form, in part or in whole, must be in writing to NACE International,
Publications Division, 1440 South Creek Drive, Houston, Texas 77084. The material presented and the views expressed in this paper are solely those of the
author(s) and are not necessarily endorsed by the Association.

View publication stats

You might also like