You are on page 1of 16

NACE Paper No.

MECCOCT18-12310

Phosphate Esters as Oilfield Corrosion Inhibition Boosters with


Combined Asphaltene Mitigation Effect

Per-Erik Hellberg, Hans Oskarsson


AkzoNobel Surface Chemistry AB
RD&I Lab. Hamnvägen 2
SE-44485 Stenungsund
Sweden

Monika Walczak, Rob Lindsay


Corrosion & Protection Centre
School of Materials
The University of Manchester
Sackville Street
Manchester
United Kingdom

ABSTRACT

In the present study, some structure-property relationships of phosphate esters connected to


different aspects of use in oilfield applications, such as formulability and performance under
shifting conditions, as well as the use as single inhibitors and performance boosters in
formulation with other types of oilfield corrosion inhibitors, will be discussed. In specific, their
co-functionality with imidazolines on steel surfaces under sweet corrosion conditions is
highlighted. Inhibition performance is demonstrated by standard test methods, while the
actual adsorption on various surfaces was shown by Quarz Crystal Microbalance with
Dissipation (QCM-D) testing. The preferential adsorption of phosphate ester versus
imidazoline on iron, as well as the high likelihood of phosphate esters attaching to iron
carbonate free substrate was concluded using X-ray photoelectron spectroscopy (XPS).
Additionally, asphaltene dispersion data imply a double functionality of phosphate ester
surfactants as corrosion- and asphaltene inhibitors.

Key words: CO2 corrosion, phosphate ester, imidazoline, inhibition, QCM-D, X-ray
photoelectron spectroscopy, XPS, asphaltene

INTRODUCTION

Background

One of the most abundant methods to mitigate corrosion in oil-and gasfield installations
is the use of film-forming organic corrosion inhibitors. No doubt, this is one of the most
important classes of oilfield production chemicals as ranked by volumes used globally. The
most common types of surfactants utilized for this application today are cationic,
either with a permanent charge or as induced by low pH. These include e.g. fatty
amines, alkoxylated fatty amines, imidazolines, pyridinium quats and quarternary ammonium
compounds like alkyl benzyl quats1-2. However, also anionic surfactants can be very useful in
this context. Phosphate esters (PE) are anionic surfactants finding use in many diverse
applications, like dispersion, lubrication, solubilisation, cleaning and wetting, to mention a
few. The adsorption of alkyl phosphate esters on various iron-based surfaces and their
potential corrosion inhibiting properties has been described in literature3-6. In addition, they
have also been well recognized as oilfield corrosion inhibitors7-8. On the other hand, literature
around their co-inhibitive properties in formulation with other surfactant classes, and
especially the mechanism behind this under oilfield-relevant conditions is not as abundant.

All of the inhibitors mentioned above are surface active compounds, a property which is
inherently required by the application where correct distribution between oil- and water
phases as well as good attachment to metal surfaces and the creation of a water-repelling
film are key elements. However, having this nature may make some surfactants complex to
formulate with other types of surface active corrosion inhibitors as well as with with other
types of production chemicals. Additionally, oilfield corrosion inhibitors will have to be
compatible with a wide range of conditions, including heavy brines, high temperatures and
varying water-to-oil ratios. The potential for PE to meet some of these challenges is
discussed in this paper.

Commonly, a “synergist” or “enhancer” is many times added to enhance the performance of


the primary inhibitor in a formulation. These are usually sulfur-containing organic
compounds. We are demonstrating that PE may work together with workhorse inhibitors like
imidazolines to improve their functionality.

Flow assurance challenges due to precipitation and adsorption/deposition of asphaltenes


during hydrocarbon production is a well-known problem. Asphaltenes represent the most
polar fractions of crude oil, and they consist of a polyaromatic core with pending alkyl side
chains. Molecules that can inhibit these processes or remove already formed deposits are of
high interest to keep oil-and gas production running smoothly9. Substances that further to this
demonstrate secondary functionalities are even more highly valued in the oilfield industry. In
this paper we show experimental results indicating certain PE work well in keeping
asphaltene dispersed, at the same time as this class of surfactants perform excellently as
corrosion inhibitors.

In summary, the paper will demonstrate why phosphate ester surfactants may be valuable
components in oilfield corrosion inhibition formulations, especially in combo with
imidazolines, and with a potential dual effect in asphaltene problem mitigation.

Chemistry of Phosphate Esters

Phosphate esters are anionic surfactants. Depending on method of synthesis, commercially


available surfactant type PE are either mainly mono-esters with one hydrophobic tail or a
blend of mono- and di-esters with one or two hydrophobic tails (Figure 1). Further, PE may
be used in their acid form or neutralized to a salt with different counter ions (Figure 2). Very
commonly, an ethylene oxide spacer is inserted between the alkyl chain and the phosphate
head-group. Introduction of this spacer enables further possibilities to tune the functionality.

Unlike most ester based chemistries, the PE are very stable to alkali as well as under
moderate acidic conditions which makes them insensitive for hydrolysis in a wide pH range.
Figure 1: Mono- and di-phosphate ester

Figure 2: Salt of a surface active phosphate ester

EXPERIMENTAL PROCEDURE

Chemicals

Pure model surfactants

Mono-N-Dodecyl phosphate (dodecyl phosphate ester, DDPE) was purchased from Merck,
while (Z)-2-2(2-(octadec-9-en-1-yl)-4,5-dihydro-1H-imidazole-1-yl)ethanamine (model
imidazoline – OMID) was synthesised at AkzoNobel Surface Chemistry, Stenungsund
(Sweden). The structures are displayed in Figure 3. More in detail, the OMID was
synthesized via a condensation reaction between di-ethylene tri-amine (DETA, tech. quality)
and oleic acid (Fisher), with a large (4:1) molar excess of DETA in order to only get the
mono-alkyl adduct. Excess DETA was removed together with water due to the high
temperature and vacuum used to complete the reaction. 1H-NMR was used to control the
purity of the product, and the analysis also confirmed <1% free DETA being present in the
final product.

Figure 3: Model surfactants DDPE (left) and OMID (right)

Technical surfactants

A variety of surface active PE with different structure specifics were investigated (Table 1).

Table 1.

Technical phosphate esters and their structure elements

Phosphate Hydrophobe Mono (M) or Mono/ Ethylene oxide (EO)


ester (br =branched) di-ester blend (M/D) – spacer size
PE1 long/linear M/D medium
PE2 short/linear M -
PE3 medium, br M/D medium
PE4 medium, br M/D no
PE5 medium, br M short
PE6 medium, br M/D short
Hydrophobe size for most PE presented in Table 1 is in the range of C12-C18. In the case of
mono- and di-ester mixtures the typical ratio is around 50/50 mono- and di-esters. All
surfactants are in the acid form except for PE2, which is a potassium salt. The codings PE1-
PE6 will be used throughout this paper.
A N-(Oleyl-vegetable alkyl) trimethylene diamine + 3 EO (acetate) was used in the QCM-D
tests while a Tall Oil Fatty Acid (TOFA)-imidazoline was used in the LPR tests.

In the solubility studies, a diamine alkoxylate as well as an imidazoline were used. The
diamine is based on a tallow fatty acid type hydrophobe while the imidazoline is the same
that was utilized in the LPR corrosion tests. Likewise, hydrotrope I and II were used in the
solubility experiments. These are sugar-based nonionic surfactants with a straight (I) or
branched (II) alkyl chain.

All technical surfactants were received from from AkzoNobel Surface Chemistry,
Stenungsund (Sweden) and used without any purification.

Crude oil

The crude oil used for the asphaltene dispersion testing originated from the Norwegian
sector of the North Sea and contains 4% asphaltenes according to SARA analysis.

Equipment

Corrosion inhibition by Linear Polarization Resistance (LPR) for Technical surfactants

Experiments were carried out using 1 L jacketed cells (Multiport Cell, Gamry), potentiostat
Interface 1000E and multiplexer ECM8 (both from Gamry). LPR Sweep was from -10 mV to
+10 mV, sweep rate 0.1 mV/s.

Working electrodes were made of C1018 Carbon Steel (UNS G10180), h = ½” and d = 3/8”
(purchased from Metal Samples). The specimens were polished with SiC paper and rinsed
with acetone prior to use. Reference electrode was an external Ag/AgCl3 in saturated KCl,
via a Luggin capillary and counter electrode was a graphite rod. Electrolyte was 3% w/w
NaCl as standard (unless otherwise mentioned), stirred at 300 rpm by a magnetic bar.

The electrolyte was de-oxygenated by sparging with CO2 for 2h, followed by pre-corrosion for
2h before the inhibitor was added. Sparging with CO2 continued for the entire test period.

Corrosion inhibition by Linear Polarization Resistance (LPR) for Model surfactants

Sample Preparation
Polycrystalline disc samples were mounted in epoxy resin with only one flat side uncoated
(exposed area 1.13 cm2) and then grounded with a series of SiC papers (250 grit, 400 grit,
600 grit, 800 grit, 1200 grit). The surface was washed with ethanol and de-ionized water
afore immersion. Samples did not undergo any pre-corrosion treatment and were immersed
directly into the electrolyte with the inhibitor.

Experimental Method
Data were acquired in a 1 L jacketed glass cell, using a typical three-electrode arrangement
comprising a working electrode (polycrystalline iron disc), a platinum counter electrode, and
a platinum electrode as a reference electrode. During the experiment the water-jacketed cell
was connected to a chiller (Julabo F250 recirculating cooler) to maintain the temperature
constant (24°C). For CO2-purged experiments the electrochemistry setup was placed inside
a nitrogen-purged glove box (Inert system). The electrochemical experiments were carried
out in a solution of 10 ppb or less concentration of oxygen (O2). A computer-controlled
potentiostat (Gamry Reference 600+, Gamry Instruments) was used to acquire the LPR
data, employing a sweep rate of 0.125 mV/s-1 over the range -10 mV to 10 mV with respect
to open circuit potential (OCP).

Quarz Crystal Microbalance with Dissipation (QCM-D)

The QCM-D data were recorded on a Q-Sense E4 system from Q-Sense (Biolin Scientific).
Temperature was +25ºC and the flowrate 50 µL/min. Sensor material was stainless steel
(SS2343), iron, corroded iron (sensor kept in 3% NaCl overnight prior to test) or Fe2O3.
Sensors were used as received from Biolin Scientific except for the corroded iron.
Flow sequence: DI water – tap water – inhibitor (in tap water) – tap water. Tap water was
from Stenungsund, Sweden (soft, 3.8 ºdH).

X-ray Photoelectron Spectroscopy (XPS)

Sample Preparation
High purity polycrystalline iron (99.99+%, Goodfellow) in cylindrical form was employed for
XPS studies. Samples were grinded with SiC papers (600 grit, 800 grit, 1200 grit, 2400 grit,
4000 grit), and then polished with diamond paste (3μm and 1 μm) until a mirror finish was
obtained. Subsequently, they were washed with acetone, ethanol and de-ionized water, and
dried.

Experimental Method
XPS was performed in a Kratos Axis Ultra facility, equipped with a glove box/load lock
system for sample introduction. Samples for XPS analysis were immersed in solution inside
a N2-purged glove box, transferred to the glove box of XPS system in an air-tight portable cell
build from vacuum compatible metal tubing and subsequently removed and inserted directly
into the XPS instrument. This approach circumvents exposure to the ambient laboratory
atmosphere, avoiding likely post immersion substrate oxidation. Upon removal of a sample
from solution, it was blown-dry with a stream of nitrogen to prevent evaporation and
subsequent physical deposition of solution components onto the sample surface.

For acquisition of XPS data, monochromated Al Kα X-rays (hν = 1486.6 eV, Dhν = 0.6 eV)
were employed as the photon source. Emitted photoelectrons were collected using a 165
mm hemispherical energy analyser incorporating a delay line detection system using hybrid
mode. Data were acquired at an analyser pass energy of 20 eV. The angle subtended by the
X-ray beam and the entrance lens of the analyser was 60°. Binding Energies (BEs) were
calibrated by assigning a BE value of 706.7 eV to the Fe 2p 3/2 metallic iron component. The
BE of C 1s hydrocarbon component of a reference powder sample was allocated to BE of
285 eV. The intensity of each presented HR spectra was normalized to the signal of bulk
iron.

Asphaltene dispersion testing

For these tests, a Turbiscan Lab Expert instrument (Formulaction, France) equipped with a
near-infrared light source and detection systems for transmission as well as light scattering
(backscattering) was used. Temperature was +25ºC or +55ºC.

Turbiscan settings
Each test was scanned every 2 min in 14 min, then every 5 min in 90 min, thus total time for
one screening test was 1h 46min. The transmission was recorded at 40 mm tube height.

Inhibitor preparation
The inhibitor was dissolved in and diluted to 10% activity in toluene. For tests with lower
concentration, the inhibitor was further diluted to 2% or 0.2%.

Test procedure
First, 20 ml heptane was added to a Turbiscan tube. Secondly, 1000 ppm of surfactant
sample with 10% activity was added to the tube, followed by 150 µl crude oil*. After the
additions the tube was shaken horizontally at 400 rpm in 60 s before screening in the
Turbiscan. If 1000 ppm showed good result with regard to transmission, a lower conc. such
as 250 or 50 ppm was tested. For screenings at 55°C, the tube with heptane was pre-heated
in a 56°C water bath during 40 minutes before addition of inhibitor and crude.

RESULTS AND DISCUSSION

Physicochemical and applicational properties of Phosphate Esters

Water solubility

Neutralized PE are well water soluble, and ethoxylation further improve the water solubility
as well as adding hard water tolerance10. Thus, ethoxylated alkyl phosphates have a much
better solubility in water than alkyl phosphates, and the more ethylene oxide the better
solubility. An alkyl phosphate without any ethylenoxy structure element is quite sensitive to
hard water and may form insoluble calcium phosphates. A shorter hydrophobe and higher
ethoxylation give a good solubility also in high caustic solutions. Mono-esters are more water
soluble than di-esters.

All solubility and compatibility studies were performed by visual inspection (neked eye). In
Figure 4, the water solubility of a typical surfactant phosphate ester under different conditions
is exemplified. As can be seen, neutralization with alkali improves the water solubility (Figure
4, a-b). Completely clear and long-term stable solutions can be achieved by formulating with
a hydrotrope (Figure 4, c-d), as well as by blending by other film-forming inhibitor surfactants
(Figure 4, e-f).

Figure 4: Water solubility of PE1 under various conditions


a. 5% in water ”as is”. pH = 2.0; b. 5% in 5% NaOH ”as is”. pH = 13.3; c. 5% in water,
6% hydrotrope I added; d. 5% in water, 3.5% hydrotrope II added; e. 5% in water-based
formulation with 10% imidazoline and 5% diamine alkoxylate, solvent etc; f. Same as e)
– diluted 1:10 with water

Solvent and brine compatibility

As previously mentioned, brine compatibility and the potential for formulation in a wide range
of solvents are useful prerequisites for a potential oilfield corrosion inhibitor. Figure 5 displays
the CaCl2 compatibility for some PE. PE1 is less soluble in this medium than in water (Figure
5, a-b), but still no immediate precipitations or phase separations are noted. Again, with
addition of a suitable hydrotrope excellent compatibility can be accomplished (Figure 5, c).
If structure is changed from PE1 to PE2, the effect of removing to EO spacer - less salt
tolerance - is obvious. When we, on the other hand, go from PE1 to PE3 (Figure 5, b-d), it is
clear that a shorter hydrophobe and longer spacer improve the compatibility.

Figure 5: Brine compatibility of phosphate esters under various conditions


a. 5% in water ”as is”. pH = 2.0; b-e: 3% surfactant in 18% w/w CaCl2 :
b. PE1 ; c. PE1 with 3% (act.) hydrotrope II added.; d. PE2 (K+ salt); e. PE3

When it comes to formulability in various solvents (Table 2) no doubt PE1, as a typical


example, shows to blend well with a range of common solvent types.

Table 2.

Solubility of PE1 (5% w/w, 20ºC) in various solvents

Water Dispersible (w/o pH adjustment)


Methanol Soluble
Ethanol Soluble
Isopropanol Soluble
MEG Soluble/Dispersible
Propylene glycol Soluble
BDG Soluble
Xylene Soluble
Alifatic solvent Soluble
Diesel Soluble
Soy Bean Oil Soluble

Adsorption of Phosphate Esters by QCM-D

In order to reach a basic understanding of adsorption properties of PE on oilfield-relevant


metallic surfaces, especially in the context of their negative charge as a contrast to the large
majority of film-forming corrosion inhibitors being positively charged, some QCM-D
measurements were carried out. With QCM-D, the kinetics of mass changes as well as
structural changes of an adsorbed film, such as water inclusion, are obtained simultaneously.
Thus, this technique offers a way to investigate how well corrosion inhibitors adsorb to
different types of surfaces as well as the properties of the adsorbed layer11-13. This could then
serve as a screening tool or to gain deeper understanding on how both commercially well
established and novel film-forming corrosion inhibitors, e.g. intended for the oil-and gas
industry, carry out their task. In order to first illustrate the adsorption profile of a typical
amine-based film-forming surfactant, the result for a N-(Oleyl-vegetable alkyl) trimethylene
diamine + 3 EO (acetate) is shown in Figure 6.
Figure 6: Adsorption of N-(Oleyl-vegetable alkyl) trimethylene diamine + 3 EO (acetate)
to stainless steel

Blue lines show frequency, which is correlated with mass change (adsorption), while red
lines show dissipation (“damping”), a measure of the rigidity/viscoelasticity of the film. Further
details of the technique are given in the relevant references11-13. In this case a rigid film, quite
insensitive to water rinse and seemingly fully developed already at a 10 ppm dosage, is
formed. The measured adsorption of 300 ng/cm2 corresponds to a calculated film thickness
of around 3 nm. When it comes to PE, the adsorption of PE3 on four different surfaces was
investigated; stainless steel, iron, corroded iron (sensor in 3% NaCl overnight prior to test)
and Fe2O3. In Figure 7 and 8, the results for two of these, stainless steel and Fe2O3, are
shown. In both cases a rapid initial adsorption takes place even at low surfactant
concentrations, while it may require additional phosphate ester, up to around 100 ppm,
before the sensor surface is seemingly saturated. Also, partial desorption takes place at the
onset of pure water rinse, irrespectively of surfactant concentration. The same pattern as in
Figure 7 and 8 is also seen with the two remaining surfaces used (data not shown). Typical
film thicknesses (as calculated from mass changes in the different experiments) are in the
range of 2-3 nm. When PE1 was tested instead of PE3 on the stainless steel surface,
adsorpion behaviour was again similar (data not shown). To conclude, PE seem to adsorb
well on various iron-based surfaces, even though the formed films may not be as rigid and
resistant to rinse as for nitrogen-based inhibitors carrying a positive charge.

Figure 7: Adsorption of PE3 to stainless steel


Figure 8: Adsorption of PE3 to Fe2O3

Corrosion inhibition testing

First, the sweet corrosion inhibition performance for a range of PE with varying structure was
investigated under typical conditions (Figure 9). Inhibition results turn out to be good
considering a single component inhibitor system (as contrast to an optimized blend), with up
to > 95% reduction in corrosion rate. In principle, and within certain limits, there are no major
performance differences related to changing hydrophobe, going from mono- to or
mono/diester, or changing size of the EO spacer14.

Figure 9: One-phase LPR CO2-bubble test results for PE3-PE6. Fluid is 3 % NaCl,
T=+60ºC and active inhibitor conc. 20 ppm. Inhibition %-age is given relative to
baseline
Imidazolines are well-known inhibitors of sweet corrosion. They form flow resistant and self-
healing films at low dosages15. At very low concentrations, it may however take some time for
the film to develop fully, and full corrosion protection in some cases is not reached until after
one day or more. Thus, other components such as e.g. sulfur-containing synergists are
sometimes used to top-up the performance of a formulation.

A typical test result without synergist is displayed in Figure 10, where an industry standard
TOFA imidazoline is used at 10 ppm dosage. Attempting to combine the good features of PE
and imidazolines, different molecules from these product classes were formulated together in
various blends and tested under miscellanous conditions. One example of this is shown in
Figure 10, with PE2 being used in a 1:2 weight-% ratio with the imidazoline earlier
mentioned. Total concentration of the two inhibitors added is 10 ppm. With this combination,
decrease of corrosion rate takes place much more rapidly than with the imidazoline only.
This indicates a potential to achieve excellent short- as well as long-term protection at very
low total inhibitor concentrations. Results of further testing of the same imidazoline with
another phosphate ester, PE1, varying molar ratios, at different temperatures and during
longer time are displayed in Figure 11. This shows that these combinations work well also at
low additions of phosphate ester, e.g. 1:7 w/w PE1:imidazoline. The formulations gives low
corrosion rate at temperatures from +40ºC to at least +80ºC, also during extended (>60h)
tests. With time and more challenging conditions the combinations with higher ratio of
imidazoline develops their efficiency in a positive direction even from initial good inhibition of
>90%. Notably, this is at half the total inhibitor dosage vs the single component phosphate
ester system (20 ppm PE1).

Figure 10: One-phase LPR CO2-bubble test results for TOFA Imidazoline as single
component and in blend with PE2. Fluid is 3 % NaCl, T=+60ºC and total active inhibitor
conc. 10 ppm
Figure 11: One-phase LPR CO2-bubble test results for PE1 as single component and in
various ratios with a TOFA imidazoline. Fluid is 3 % NaCl , T=+40ºC - +80ºC and total
active inhibitor conc. 10 or 20 ppm

Adsorption of Phosphate Ester/Imidazoline blends by XPS

The corrosion rate values of metal during the 6 hours immersion in CO2-purged 3% NaCl/IPA
in the presence of 0.02 mM DDPE, 0.02 mM OMID, and 0.005/0.02 mM and 0.02/0.02 mM
OMID/DDPE mixtures are presented in Figure 12 (a). The decrease of corrosion rate under
CO2-purged conditions is evident in all presented systems given the inhibition efficiency
above 90%. Analogously to its commercial counterpart, the OMID inhibition action advances
as a function of time and it is being improved by DDPE addition.

To gain an insight into the chemical composition of the adsorbed layer of inhibitor(s) as well
as determine the substrate termination, the X-ray Photoelectron Spectroscopy method has
been employed. Focusing on the high-resolution C 1s core level spectra for uninhibited iron
in CO2-purged 3% NaCl/IPA solution, Figure 12 (b) bottom spectrum, the corrosion product
formation is apparent due to the feature at B.E. ~290.1 eV. Such signal is associated with
carbonates like FeCO3, Na2CO3, and NaHCO3 developing in a carbonic acid environment16.
The intensity of discussed feature is entirely quenched in the presence of 0.02 mM DDPE
and diminished in the presence of 0.02 mM OMID demonstrating that both inhibitors supress
the carbonates formation. The DDPE is more effective and therefore this is likely the reason
for its good properties as an intensifier.
(a) 6h (b) C 1s 6h
10 CO2-purged 3% NaCl/IPA
0 mM DATA
0.02 mM DDPE
0.02 mM OMID component
0.02 mM OMID 0.02 mM DDPE background
0.005 mM OMID 0.02 mM DDPE FIT
Corrosion Rate (mm.year )
-1

CO3

Intensity (Arb. Units)


1
0.02 mM OMID

IE = 91%

0.1
IE = 95%
0.02 mM DDPE
IE = 97%

0.01 IE = 99%
x 0.35

3% NaCl/IPA

1 2 3 4 5 6 292 290 288 286 284


Time (hours) Binding Energy (eV)

Figure 12: Corrosion rates in CO2-purged 3% NaCl/IPA alone and in the presence of
0.02 mM DDPE, 0.02 mM OMID, and 0.005/0.02 mM and 0.02/0.02 mM OMID/DDPE
mixtures as a function of immersion time (a) HR C 1s XPS spectra of iron substrate
uninhibited and inhibited with 0.02 mM DDPE and 0.02 mM OMID in CO2-purged 3%
NaCl/IPA (b) The C 1s signal related to uninhibited substrate is scaled of a factor of
0.35 due to the high signal intensity. All experiments have been performed in room
temperature

The organic film formation due to the adsorption of DDPE and OMID species are evident
from the P 2p and N 1s features presence respectively, Figure 13 (a-b). Apparently, the B.E.
~133.8 eV of 2p3/2 signal (P’’) related to inhibited substrate differs from the B.E. ~134.7 eV of
2p3/2 signal (P’) of DDPE in a powder form, Figure 13 (b). This strongly suggest the DDPE
molecule being adsorbed via PO4 group (or more specifically -OH, =O) interaction with
cationic iron species. The 2p3/2 feature at B.E. ~133.8 eV corresponds to B.E. of P–O-Fe type
of bondage reported for FePO417. This binding energy has been also found for alkyl
phosphate adsorption on titanium18. The main features in N 1s spectrum for OMID have been
identified at B.E. ~399.2 eV (N’), B.E. ~400.3 eV (N’’), B.E. ~401.3 eV (N’’’), and associated
to the NH2, C=N and C-N, and protonated nitrogen species respectively. The similar N 1s
signal for imidazoline based inhibitor has been reported previously19 in 1 M HCl. No specific
feature has been assigned to the adsorbed OMID molecule as it is believed to adsorb atop
the layer of counterions20. Given the presence of both N1s and P 2p signal for DDPE and
OMID blends the inhibitors most likely co-adsorb.
(a) 6h
3% NaCl/IPA
(b) P' P'' 6h
3% NaCl/IPA
N 1s P 2p
N''' N'' N'

0.005 mM OMID 0.02 mM DDPE


0.005 mM OMID 0.02 mM DDPE
Intensity (Arb. Units)

0.02 mM OMID 0.02 mM DDPE 0.02 mM OMID 0.02 mM DDPE

0.02 mM OMID

0.02 mM OMID 0.02 mM DDPE

0.02 mM DDPE

DATA
component
background
FIT
DDPE powder

404 402 400 398 138 136 134 132 130


Binding Energy (eV) Binding Energy (eV)

Figure 13: XPS HR N 1s (a) and P 2p (b) spectra of iron substrate inhibited with 0.02
mM DDPE, 0.02 mM OMID, 0.005/0.02 mM OMID/DDPE, and 0.02/0.02 mM OMID/DDPE
in CO2-purged 3% NaCl/IPA solution. The reference P 2p spectrum of DDPE inhibitor in
a powder form is also presented. All experiments have been performed in room
temperature

Interestingly, there is a change in N’’/N’’’ intensity ratio in favour to protonated nitrogen


species in OMID signal when co-adsorbed with DDPE in two different molar ratios. This
could be related to the adsorbed OMID molecule itself as well as to the residue protonated
nitrogen present in traces of a corrosion product. Assuming the OMID and DDPE co-
adsorbed simultaneously in one layer atop the metal substrate and there are no nitrogen
impurities, the simple atomic ratio analysis has been implemented to estimate the number of
OMID vs number of DDPE adsorbed molecules. The same P/N ~ 3 has been found for both,
an equimolar and a 4:1 mole/mole DDPE : OMID examined blends, which means there is
one adsorbed molecule of OMID per 10 molecules of DDPE. DDPE dominates in the
adsorbed layer of protective organic film regardless the concentration in the solution for the
examined systems. This preferential adsorption naturally gives a good clue to why even
minor aliquotes of PE in an imidazoline/PE formulation improve the total corrosion inhibition
performance as earlier discussed. We are planning additional experiments including looking
at a blend with a large molar surplus of OMID to even better mimic some of the inhibition
experiments. Then, further investigations are needed to fully understand the driving force
behind this, since either bulk causes like different solubility profiles, or surface interactions
such as exact binding/packing/co-adsorption mechanisms could potentially be the main
reason. We are presently looking into avenues to additional understanding of these
processes.
Asphaltene dispersion testing

In addition to the corrosion inhibition tests one of the PE giving good results in this respect,
PE1, was also evaluated in a well-recognized lab screening procedure for asphaltene
inhibition21. The method involves adding an asphaltene-rich crude to hexane or heptane to
induce precipitation and aggregation of the asphaltene molecules. Without any inhibitor
present, asphaltenes will precipitate and form larger aggregates, which will then settle at
bottom of the test tube, meaning light transmission higher up in the tube will increase.
If, on the other hand, an efficient inhibitor is added, it will keep the precipitated
(nanoaggregated) asphaltenes dispersed, preventing them from forming larger aggegates.
They do not settle and light transmission is kept constant. In a real life application scenario,
settling of large aggregates may result in hard-treated deposits which could cause serious
flow assurance problems, while dispersed asphaltenes follow the liquid flow without creating
issues. In Figure 14, the results with PE1 as inhibitor are displayed. While the transmission
without inhibitor more or less instantly start to increase from the initial value, PE1 keeps
asphaltenes dispersed at a concentration of 50 ppm or lower at +25ºC, and at 250 ppm or
lower at +55ºC. Normally, these tests are more challenging at higher temperatures. Tests at
even lower inhibitor concentrations and with additional crude oils are presently pending.

Figure 14: NIR light transmission @40 mm height vs time for different additions of PE1
@+25ºC or +55ºC

CONCLUSIONS

- Phosphate ester surfactants display several properties making them worthwhile considering
as inhibitors for different scenarios of Oilfield corrosion, including compatibility.

- Although equipped with a different head group charge than”standard” amine based
inhibitors, they are shown to adsorb to various relevant metal surfaces.
- Within a certain space of structure variation, sweet corrosion inhibition performance seem
robust, although fine tuning can give minor improvements.
- They may be especially valuable as component in blends with other types of inhibitors
rather than used as single components.

- Phosphate esters suppress Fe(CO)3 formation and co-adsorb preferentially to imidazolines


on a metal surface.

- Finally, positive asphaltene inhibition results indicate this class of surfactants may be
employed as combination inhibitors for corrosion and asphaltene in oil-and gasfield
applications, provided optimized structures are employed.

ACKNOWLEDGEMENTS

The authors would like to thank Fredrik Andersson and Thomas Ljungdahl, both AkzoNobel
Surface Chemistry, Stenungsund, Sweden for QCM-D measurements and synthesis of the
OMID model substance, respectively.

REFERENCES

1. M. Gregg, S. Ramachandran, “Review of Corrosion Inhibitor Developments and testing


for Offshore Oil and Gas Production Systems”, NACE 04422, CORROSION 2004, New
Orleans, La., 28 March -1 April, 2004.

2. M. A. Kelland, Production Chemicals for the Oil and Gas Industry, CRC press (2009),
p. 200-205

3. M. Foss, E. Gulbrandsen, J. Sjöblom, “Adsorption of Corrosion Inhibitors onto Iron


Carbonate (FeCO3) Studied by Zeta Potential Measurements”, Journal of Dispersion Science
and Technology, 31:2 (2010), p. 200-208

4. X. Gao, S. Liu, H. Lu, F. Gao, H. Ma, “Corrosion Inhibition of Iron in Acidic Solutions by
Monoalkyl Phosphate Esters with Different Chain Lengths”, Ind. Eng. Chem. Res. 54 (2015),
p.1941−1952

5. J. Kim, W. Li, B. L. Philip, C. P. Grey, “Phosphate adsorption on the iron oxyhydroxides


goethite (α-FeOOH), akaganeite (β-FeOOH), and lepidocrocite (γ-FeOOH):
a 31P NMR Study”, Energy Environ. Sci. 4 (2011), p.4298

6. C. Zhao, X. Gao, H. Lu, R. Yan, C. Wang, H. Ma, “Spontaneous formation of mono-n-


butyl phosphate and mono-n-hexyl phosphate thin films on the iron surface in aqueous
solution and their corrosion protection property”, RSC Adv. 5 (2015), p.54420

7. B. Alink, B. Outlaw, V. Jovancicevic, S. Ramachandran, S. Campbell, “Mechanism of


CO2 corrosion inhibition by phosphate esters”, CORROSION/99, paper no.37
8. O. Yepez, N. Obeyesekere, J Wylde, “Development of Novel Phosphate Based Inhibitors
Effective For Oxygen Corrosion”, SPE173723, Presented at the SPE International
Symposium on Oilfield Chemistry held in The Woodlands, Texas, USA, 13–15 April 2015

9. S. Subramanian, L. Buscetti, S. Simon, M. Sacré, J. Sjöblom, “Influence of Fatty-


Alkylamine Amphiphile on the Asphaltene Adsorption/Deposition at the Solid/Liquid Interface
under Precipitating Conditions”, Energy Fuels, 32, 4 (2018), p.4772–4782
10. W.M. Linfield, Anionic surfactants part II, Surfactant science series, Vol 7, Marcel Dekker
(1976), p.508-511
11. M. Knag, J. Sjöblom, G. Oye, E. Gulbrandsen, “A quartz crystal microbalance study of
the adsorption of quaternary ammonium derivatives on iron and cementite”, Colloids and
Surfaces A: Physicochem Eng. Aspects 250 (2004), p.269

12. P-E. Hellberg, “Environmentally Acceptable Polymeric Corrosion Inhibitors”, SPE104780,


Presented at the SPE International Symposium on Oilfield Chemistry held in The Woodlands,
Texas, USA, 11–13 April 2011

13. Y. Ding, B. Brown, D. Young, S. Nesic, M. Singer, " Effect of Temperature on Adsorption
Behavior and Corrosion Inhibition Performance of Imidazoline-Type Inhibitor",
CORROSION/17, paper no. 9350 (New Orleans LA: NACE International, 2017)

14. P-E. Hellberg, “Phosphate Esters as Inhibitors of Oilfield Corrosion”, Presented at the
Eurocorr 2017 conference held in Prague, Czech Republic, 3-7 September 2017

15. Y. Xiong, B. Brown, B. Kinsella, S. Nesic, M. Singer, E. Pailleret, "AFM Studies of the
Adhesion Properties of Surfactant Corrosion Inhibitor Films ", CORROSION/13, paper no.
2521 (NACE International, 2013)

16. K. Dideriksen, C. Frandsen, N. Bovet, A.F. Wallace, O. Sel, T. Arbour, A. Navrotsky, J.J.
De Yoreo, J.F. Banfield, “Formation and transformation of a short range ordered iron
carbonate precursor”, Geochimica et Cosmochimica Acta 164 (2015), p.94-109

17. Y. Barbaux, M. Dekiouk, D. Le Maguer, L. Gengembre, D. Huchette, J. Grimblot, “Bulk


and surface analysis of a Fe-P-O oxydehydrogenation catalyst”, Applied Catalysis A: General
90 (1992), p.51-60

18. D.M. Spori, N.V. Venkataraman, S.G.P. Tosatti, F. Durmaz, N.D. Spencer, S. Zürcher,
“Influence of Alkyl Chain Length on Phosphate Self-Assembled Monolayers”, Langmuir 23
(2007), p.8053-8060

19. O. Olivares‐Xometl, N.V. Likhanova, R. Martínez‐Palou, M.A. Domínguez‐Aguilar,


“Electrochemistry and XPS study of an imidazoline as corrosion inhibitor of mild steel in an
acidic environment”, Materials and Corrosion 60 (2009), p.14-21

20. B.J. Usman, S.A. Ali, “Carbon Dioxide Corrosion Inhibitors: A review”, Arabian Journal
for Science and Engineering 43 (2018), p.1-22

21. M. Wang, J. Kaufman, X. Chen, C Sungail, “Development and Evaluation of Non-Ionic


Polymeric Surfactants as Asphaltene Inhibitors”, SPE173720, Presented at the SPE
International Symposium on Oilfield Chemistry held in The Woodlands, Texas, USA,
13–15 April 2015

You might also like