CORROSION Y PROTECCION, June 13 2023

Inhibition of X52 Corrosion in CO2-Saturated Brine by a Dialkyl-Diamide from Coffee Bagasse Oil

N. B. Gomez-Guzman 1, Jorge Canto 2, L. M. Martinez-de-la-Escalera 2, Adrián Neri 2 and J. Porcayo-Calderon 3,*

1

2 3

* Correspondence: jporcayoc@gmail.com

Abstract: This work reports the performance of a green corrosion inhibitor with double hydrocarbon chain. The evaluated inhibitor was a dialkyl-diamide from coffee bagasse oil and its electrochemical behavior was evaluated on an API-X52 steel in CO2-saturated brine at 60 ◦C. The electrochemical behavior was determined by measurements of open circuit potential, polarization resistance, and electrochemical impedance spectroscopy. In addition, the thermodynamic parameters of the corrosion process were obtained in the temperature range from 40 ◦C to 80 ◦C. Electrochemical studies showed that the inhibitor is capable of suppressing metal dissolution by up to 99% at 25 ppm. On the other hand, the thermodynamic parameters indicate that when adding the inhibitor, there is a strong increase in both Ea and ∆H◦ values, and that as time increases, they decrease until reaching similar values to those observed in the absence of the inhibitor. Furthermore, ∆S◦ values tend to become more negative with immersion time because of the formation of a stable film on the metal surface.

Keywords: API X-52; sweet corrosion; dialkyl-diamide; coffee bagasse oil; electrochemical techniques

1. Introduction

Amide- and imidazoline-type compounds of fatty acids of vegetable origin (am- pholytic compounds) have been found to be excellent corrosion inhibitors because in their structure they have nitrogenous groups and aliphatic chains that act as a metallophilic and hydrophobic group, respectively [1–8]. It has been recognized that the most efficient in- hibitors are those with aliphatic chains with 16–18 carbon atoms [9], and since in vegetable oils the largest proportion of their fatty acids correspond precisely to those with 18C chains, it is for this reason that they are used for the synthesis of corrosion inhibitors and the high inhibition efficiencies obtained.

In the hydrocarbon production and transportation industry, one of the main causes of material degradation has been attributed to the presence of dissolved CO2 [10–12], and its inhibition process has generally been addressed by injecting organic compounds into the corrosive fluid. These molecules are made up of a hydrophilic group (polar group) and a hydrophobic group (hydrocarbon chain). The adsorption of these molecules to the metal surface generates a protective layer by reducing its wettability caused by the hydrophobic group [13,14].

The use of oils obtained from agro-industrial waste has proven to be a sustainable alternative for the development and synthesis of green corrosion inhibitors, and its excellent performance as sweet corrosion inhibitors has been demonstrated [6,15–18]. Its oils have a content of fatty acids such as stearic, oleic, linoleic, and palmitic, suitable for the synthesis of amide- and imidazoline-type inhibitors, and in particular those inhibitors obtained from coffee bagasse oil have proven to be efficient corrosion inhibitors [4–7].

CIICAp, Universidad Autonoma del Estado de Morelos, Avenida Universidad 1001, Cuernavaca 62209, Morelos, MexicoCorrosion y Proteccion (CyP), Buffon 46, Mexico City 11590, MexicoDepartment of Chemical Engineering and Metallurgy, University of Sonora, Hermosillo 83000, Sonora, Mexico

  

Citation: Gomez-Guzman, N.B.; Canto, J.; Martinez-de-la-Escalera, L.M.; Neri, A.; Porcayo-Calderon, J. Inhibition of X52 Corrosion in CO2-Saturated Brine by a Dialkyl-Diamide from Coffee Bagasse Oil.Molecules2023,28,763. https:// doi.org/10.3390/molecules28020763

Academic Editor: Sourav Kr. Saha

Received: 29 November 2022 Revised: 24 December 2022 Accepted: 26 December 2022 Published: 12 January 2023

Copyright: © 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https:// creativecommons.org/licenses/by/ 4.0/).

Molecules 2023, 28, 763. https://doi.org/10.3390/molecules28020763 https://www.mdpi.com/journal/molecules

Molecules 2023, 28, x FOR PEER REVIEW 2 of 23

Molecules 2023, 28, 763

2 of 21

On the other hand, it has been shown that the presence of more than one hydrophobic group attached to the polar group can increase the inhibition efficiency [13,19]. Therefore,

On the other hand, it has been shown that the presence of more than one hydrophobic

the idea of this work is to determine the effect of the presence of two hydrocarbon chains

group attached to the polar group can increase the inhibition efficiency [13,19]. Therefore, on the performance of an inhibitor obtained from coffee bagasse oil as an alternative green

the idea of this work is to determine the effect of the presence of two hydrocarbon chains

inhibitor for sweet corrosion. Its performance has been evaluated on an API-X52 steel in

on the performance of an inhibitor obtained from coffee bagasse oil as an alternative green binrhinibei(t3o%r fobrysweiegthctoNrraoCsilo)nsa.tIutsrapterdfoinrmCaOn2ceath6a0s b°CeebnyemvaelaunasteodfonpeancAirPcIu-iXt5p2ostetenetliainl,

mpoelanriszaotfiopnotrenstiisotadnycnea,manicdpeolleacrtirzoacthioenmmicaelasimurpeemdeantcse, tmheatshueremmoednytnsa. mInicapdadriatimone,tebrys omfetahnescofrrpoostieontipordoycneassmhicavpeolbaereiznadtieotnermeianseudreinmaentetsm, tpheerathtueremroadnygneafmroimc p4a0ratom8e0te°rCs .of the corrosion process have been determined in a temperature range from 40 to 80 ◦C.2. Results and Discussion

pbroilnaeri(z3a%tiobnyrweseiisgtahntcNe,aCanl)dsaetluecrtartoecdhienmCicOal iamt 6p0edCanbcye meansusroefmoepnetns.cIinrcuaditdpitoitoen,tibayl, 2

2. Results and Discussion

2.1. Open Circuit Potential Measurements

2.1. Open Circuit Potential Measurements

Figure 1 shows the variation in OCP values as a function of time for API X52 steel in

Figure 1 shows the variation in OCP values as a function of time for API X52 steel in a a 3% NaCl solution saturated with CO2 at 60 °C with and without the addition of an in-

3% NaCl solution saturated with CO2 at 60 ◦C with and without the addition of an inhibitor.

hibitor. In this type of study, the observed trend is indicative of the thermodynamic sta-

In this type of study, the observed trend is indicative of the thermodynamic stability of

bility of the metal surface in the electrolyte where it is immersed [20]. It is observed that,

the metal surface in the electrolyte where it is immersed [20]. It is observed that, in the

in the absence of the inhibitor (0 ppm), the OCP values tend to slowly increase as a func-

absence of the inhibitor (0 ppm), the OCP values tend to slowly increase as a function of

tion of time from −740 mV to −725 mV throughout the assay. The observed trend indicates

time from −740 mV to −725 mV throughout the assay. The observed trend indicates that

that a layer of corrosion products with protective characteristics is possibly being formed

a layer of corrosion products with protective characteristics is possibly being formed on

on the steel surface since their OCP values change in the noble direction. On the other

the steel surface since their OCP values change in the noble direction. On the other hand,

hand, in the presence of the inhibitor, a rapid increase in OCP values is observed in all

in the presence of the inhibitor, a rapid increase in OCP values is observed in all cases;

cases; this trend suggests that in the presence of the inhibitor, the steel surface showed a

this trend suggests that in the presence of the inhibitor, the steel surface showed a more

more noble behavior due to the adsorption of inhibitor molecules. In general, the trend of

noble behavior due to the adsorption of inhibitor molecules. In general, the trend of OCP

OCP values is a function of the inhibitor concentration. With the lowest inhibitor concen-

values is a function of the inhibitor concentration. With the lowest inhibitor concentration

tration (5 ppm), the OCP values tend to decrease after 2 h, this could be because this con-

(5 ppm), the OCP values tend to decrease after 2 h, this could be because this concentration

centration is insufficient to form a protective film on the steel surface. At 10, 50, and 100

is insufficient to form a protective film on the steel surface. At 10, 50, and 100 ppm, the OCP

ppm, the OCP values are practically stable after 2 h of immersion, and with 25 ppm the

values are practically stable after 2 h of immersion, and with 25 ppm the noblest values

noblest values were obtained with a tendency to increase as a function of time.

were obtained with a tendency to increase as a function of time.

Figure 1..OOCPCPvavriartiaotniofnorfAorPIA-XP5I2-Xs5te2elstineeCl Oin2-sCaOtur-astaetdurbartiende abtr6in0e°Ca,t w60ith Can, dwwitihthaonudt inwhiitbhiotourt 2

addition.

inhibitor addition.

2.2. Linear Polarization Resistance Measurements

2.2. Linear Polarization Resistance Measurements

Figure 2a shows the variation in the Rp values as a function of time for the API X52 Figure 2a shows the variation in the Rp values as a function of time for the API X52

steelina3%NaClsolutionsaturatedwithCO at60◦Cwithandwithouttheadditionof steel in a 3% NaCl solution saturated with CO2 at 60 °C with and without the addition of

the inhibitor. Since the value of the resistance to polarization is inversely proportional to the

the inhibitor. Since the value of the resistance to polarization is inversely proportional to

corrosion rate, it is common to interpret that the trend of these values indicates whether the

the corrosion rate, it is common to interpret that the trend of these values indicates

material is undergoing an active corrosion process (continuous decrease in the Rp value),

ules 2023, 28, x FOR PEER REVIEW 3 of 23 Molecules 2023, 28, 763

whether the material is undergoing an active corrosion process (continuous decrease in

the Rp value), the material remains immune, or a stable protective film is present (constant

3 of 21

the material remains immune, or a stable protective film is present (constant Rp values), or

Rp values), or a protective film is developing on the surface (continuous increase in Rp values).

(a)

a protective film is developing on the surface (continuous increase in Rp values).

 

(b)Figure 2. (a) Rp variation for API-X52 steel in CO2-saturated brine at 60 °C, with and without inhib-

BasedontheBasbeodveo,nithcaenaboevees,taitbclaisnhebdeethstaatb,liinshtheedathbaset,nicnetohfetahbesienhcieboitfotrh,ethienhRipbitor,theRpvaluesremavinalaulmesorsetmcoainstaalmntoisnttchoenfsitrasntt6ihnothfeimfimrster6shioonf,bimutmseurbssioenq,ubeuntlsyutbhseyquinecnrtelyasteheyincreaseslightly untisllitghhetleyndunotfilttheetenstd. Tofhitshemtaeyst.beThinisdimcatyivbeeoifntdhiecaftoivrme oatfiothneofforsmtaabtlieonanodf stable andprotectivecoprrootesicotinveprcodrruocstisonbapsreoddmucatisnblyasoendimroaninclayrboonniarotens.caTrhbiosnisatceos.nsTishtiesnitswcoitnhsistentwith

what has been reported in the literature, where it has been indicated that from 60 ◦C, iron what has been reported in the literature, where it has been indicated that from 60 °C, iron

carbonate with protective properties begins to precipitate on the steel surface [21]. In the carbonate with protective properties begins to precipitate on the steel surface [21]. In the

Figure 2. (a) Rp variation for API-X52 steel in CO2-saturated brine at 60 C, with and without itor addition. (b) Inhibition efficiency as a function of inhibitor concentration.

inhibitor addition. (b) Inhibition efficiency as a function of inhibitor concentration.

presence of the inhibitor, the Rp values indicate that a protective film is developing on the

presence of the inhibitor, the Rp values indicate that a protective film is developing on the

metal surface due to the adsorption of the inhibitor molecules, forming a barrier that limits

metal surface due to the adsorption of the inhibitor molecules, forming a barrier that limits

the entry of aggressive species. The protective capacity of the inhibitor increases when

the entry of aggressive species. The protective capacity of the inhibitor increases when its

its concentration increases from 5 ppm to 25 ppm, and above this concentration (50 and

concentration increases from 5 ppm to 25 ppm, and above this concentration (50 and 100

100 ppm) the resistance values decrease. The decrease in Rp values is possibly due to its

ppm) the resistance values decrease. The decrease in Rp values is possibly due to its con-

concentration being too high. This causes repulsive forces that limit the free access of the

centration being too high. This causes repulsive forces that limit the free access of the mol-

molecules to the metal surface or cause the detachment of already adsorbed molecules. At

ecules to the metal surface or cause the detachment of already adsorbed molecules. At this

this temperature, the optimal dose of the inhibitor is 25 ppm. The Rp values obtained are

temperature, the optimal dose of the inhibitor is 25 ppm. The Rp values obtained are con-

considerably larger than those reported for similar inhibitors [17,19]. siderably larger than those reported for similar inhibitors [17,19].

c

Molecules 2023, 28, 763

4 of 21

In Figure 2b, the inhibition efficiency derived from the data in Figure 2a is shown. The inhibition efficiency values were determined according to the following relationship:

E(%) = Rpi −Rpb ×100 (1) Rpi

where Rpb is the value of Rp in the absence of the inhibitor and Rpi is the value of Rp in the presence of the inhibitor. From the figure in the absence of the inhibitor after an initial corrosion process, the formation or precipitation of corrosion products on the surface of the steel caused a decrease in its corrosion rate, observing a reduction close to 20% at the end of the test. However, with the addition of 5 and 10 ppm of the inhibitor, an inhibition efficiency of the order of 50–70% was obtained, but at higher concentrations the efficiency increased above 95%. The trend indicates that the maximum protection efficiency is reached with 25 ppm of the inhibitor with an inhibition efficiency of 99%. At this concentration, the greatest surface coverage is achieved, which limits the free contact of the electrolyte with the metal surface and makes diffusion through the protective layer the limiting step in the metal dissolution process [22].

The inhibition efficiencies are higher than those obtained with similar inhibitors ob- tained from other vegetable oils [17,19] and imidazolines with more than one hydrocarbon chain [13]. The higher inhibition efficiencies may be due to the presence of unsatura- tions in the hydrocarbon chains that favor the adsorption of the inhibitor to the metal surface [14,23,24].

Figure 3 shows the morphological aspects of the clean surface of the evaluated steel in the presence of the different inhibitor concentrations. It is evident that the greatest damage observed corresponds to the condition in the absence of the inhibitor. At low concentrations of the inhibitor (5 and 10 ppm), the surface showed strongly adhered corrosion product residues, and at high concentrations (50 and 100 ppm) the concentration of residual corrosion products was lower. It is clear that at 25 ppm the surface of the steel is practically free of surface defects due to the corrosive action of the electrolyte.

2.3. Electrochemical Impedance Spectroscopy

Figure 4 shows the Nyquist and Bode diagrams for API X52 steel in a 3% NaCl solution saturated with CO2 at 60 ◦C with and without the addition of the inhibitor after 24 h of immersion. According to the Nyquist diagram, in all cases, the apparent presence of a single capacitive semicircle is observed, the diameter of which increases as the inhibitor dose increases up to 25 ppm, and at higher concentrations it tends to decrease. Since the diameter of a capacitive semicircle is equal to Rct + Rs, the observed trend is consistent with the Rp values observed at 24 h (Figure 2a). According to the Bode plot in its impedance modulus format, in the absence of the inhibitor the presence of both high and low frequency plateaus is observed, and a linear relationship in the intermediate frequency region. However, in the presence of the inhibitor, the high-frequency plateau is formed at frequencies greater than 10 kHz, and at lower frequencies the presence of two linear relationships (log |Z| − log f ) is observed. In the low frequency region, the low frequency plateau is observed, whose value increases with increasing inhibitor concentration up to 25 ppm and at higher concentrations it tends to decrease. The trend observed in the maximum value of the low-frequency plateau coincides with the trend of the Rp values (Figure 2a). On the other hand, according to the Bode diagram in its phase angle format, in the absence of an inhibitor, the formation of a time constant is detected, whose maximum is located around 10 Hz with a value of 63◦. In the presence of an inhibitor, two-time constants are observed, where the first is in the high frequency region with a maximum phase angle between 45–50◦ for concentrations greater than 10 ppm, and between 1–10 Hz the presence of the second time constant is noted, with a maximum phase angle around 67◦ with the addition of 25 ppm of the inhibitor. Due to the location of the first time constant (high frequency region) it can be associated with the formation of a thin layer of the inhibitor molecules onto the steel surface, and because in the low frequency region the angle of

Molecules 2023, 28, 763 5 of 21

Molecules 2023, 28, x FOR PEER REVIEW

◦ 5 of 23 phase tends to 0 , it can be suggested that the inhibitor film is dense, which significantly

reduced the diffusion of aggressive species of the electrolyte towards the metal surface.

 

0 ppm

5 ppm

 

10 ppm 25 ppm

 

50 ppm 100 ppmFigure 3. Morphological aspects of the API-X52 steel surface after the corrosion test in CO2-saturated

Figure 3. Morphological aspects of the API-X52 steel surface after the corrosion test in CO2-saturated brine at 60 °C. ◦

brine at 60 C.

2.3. Electrochemical Impedance Spectroscopy

Figure 4 shows the Nyquist and Bode diagrams for API X52 steel in a 3% NaCl solu- tion saturated with CO2 at 60 °C with and without the addition of the inhibitor after 24 h of immersion. According to the Nyquist diagram, in all cases, the apparent presence of a single capacitive semicircle is observed, the diameter of which increases as the inhibitor

lecules 2023, 28, x FOR PEER REVIEW 7 of 23 Molecules 2023, 28, 763

(a)

(b)

(c)

6 of 21

 

Figure4.NyFqiguuisrte(a4.)aNndyqBuoidste(ab),ca)npdloBtsofdoerA(bP,cI)-Xp5l2otsstefeolrinACPOI-X2-5sa2tustreaetledinbrCinOe-asfateturr2a4tehdimbrmineer-after24h 2

sion at 60 °C, with and without inhibitor addition.

immersion at 60 C, with and without inhibitor addition.

Molecules 2023, 28, x FOR PEER REVIEW 8 of 23

Molecules 2023, 28, 763

7 of 21

Figure 5 shows the evolution of the Nyquist and Bode diagrams for the steel in the

Figure 5 shows the evolution of the Nyquist and Bode diagrams for the steel in the absence of the inhibitor (Figure 5a) and with 25 ppm of the inhibitor (Figure 5b), which,

absence of the inhibitor (Figure 5a) and with 25 ppm of the inhibitor (Figure 5b), which, according to the results, showed the best inhibition efficiency.

according to the results, showed the best inhibition efficiency.

    

(a) 0 ppm (b) 25 ppmFigure 5. (a) Nyquist and Bode plots for API-X52 steel in CO2-saturated brine a 60 °◦C. (b) Nyquist

Figure 5. (a) Nyquist and Bode plots for API-X52 steel in CO2-saturated brine a 60 C. (b) Nyquist and Bode plots for API-X52 steel in CO2-saturated brine a 60 °C◦ and 25 ppm of inhibitor.

and Bode plots for API-X52 steel in CO2-saturated brine a 60 C and 25 ppm of inhibitor.

InItnhtehaebasbesnecnecoefotfhteheinihnihbiibtoitror(F(iFgiugurere5a5)a,)t, htheeNNyyqquuisitstddiaiaggrraamsshhoowssttheeeevolluttiion ofoafcaacpapcaitciivteivseesmemicirccirleclwe whohsoesedidaimametetrerapapparaernentltylyininccrereaasseesswitithhimimmeerrssiionttiime..Theincinrecaresaesieninddiaimameteeterrmaaybeassociatedwiitthththeefoformrmataiotinonanadn/do/roprrperceipciptaitiaotnioonfocofrcrorsrion-

products on the steel surface, which reduced the reaction area. The Bode plot in its

sion products on the steel surface, which reduced the reaction area. The Bode plot in its

impedance modulus format shows the typical behavior of a corrosion process with a single

impedance modulus format shows the typical behavior of a corrosion process with a sin-

time constant, that is, formation of the high-frequency plateau starting at 1 kHz, a linear

gle time constant, that is, formation of the high-frequency plateau starting at 1 kHz, a

relationship (log |Z| − log f ) in the intermediate frequency region, and the presence of the linear relationship (log |Z| − log f) in the intermediate frequency region, and the presence

low-frequency plateau. According to the value of the impedance module, |Z|, the increase of the low-frequency plateau. According to the value of the impedance module, |Z|, the

in the value of Rct is negligible as the immersion time elapses. This suggests a low protective

increase in the value of Rct is negligible as the immersion time elapses. This suggests a

capacity of the corrosion products formed on the steel surface. The Bode plot in its phase

low protective capacity of the corrosion products formed on the steel surface. The Bode

angle format shows only one time constant, which agrees with the formation of a single

plot in its phase angle format shows only one time constant, which agrees with the for-

capacitive semicircle and a single log |Z| − log f linear relationship in the intermediate mation of a single capacitive semicircle and a single log |Z| − log f linear relationship in frequency region. The maximum of the phase angle tends to increase (52 → 63◦) and move the intermediate frequency region. The maximum of the phase angle tends to increase (52

to lower frequencies (50 → 10 Hz) with the immersion time. This behavior can be associated → 63°) and move to lower frequencies (50 → 10 Hz) with the immersion time. This behav-

with the formation of a layer of corrosion products on the steel surface.

ior can be associated with the formation of a layer of corrosion products on the steel sur-

On the other hand, with the addition of 25 ppm of the inhibitor (Figure 5b), the Nyquist diagram shows the apparent presence of a single capacitive semicircle whose

face.

On the other hand, with the addition of 25 ppm of the inhibitor (Figure 5b), the

diameter increases significantly with immersion time. The magnitude of the increase in

Nyquist diagram shows the apparent presence of a single capacitive semicircle whose

Molecules 2023, 28, x FOR PEER REVIEW 9 of 23

Molecules 2023, 28, 763

8 of 21

diameter increases significantly with immersion time. The magnitude of the increase in

the diameter of the capacitive semicircle is associated with the adsorption of the inhibitor

on the steel surface, thereby forming a protective layer that isolates the metal surface from

the diameter of the capacitive semicircle is associated with the adsorption of the inhibitor

the aggressive electrolyte. From the Bode diagram in its impedance modulus format, after

on the steel surface, thereby forming a protective layer that isolates the metal surface from

ththeeagadgrdeistsioivne oelfetchtreoliynthei.bFirtomr, the Bhoigdhe-dfrieaqguraemnciyn pitlsaitmeapuedbaengciensmtodufolurms foartmfraet,qaufetenrcies

tghreeadtedritihoano1f0tkhHe zinahnibditinort,htahterhegigiohn-fraelqinueanrcyreplaltaitoenasuhbipe,gliongs |toZ|fo−rmlogaft, firsefqouremnecdie.sThe

garpeaptearrtahnacne1o0fktHhizs alindeainr rtehlattiroegnisohnipa ilnintehairsrrealantgioenosfhfirpe,qlougen|cZie|s−islaogchf,airsafcotremrisetdic. Tsihgen of

atphpeeadrasnocreptoifonthoisfloinrgearnriecliantihoinbsithoiprsiwn tihthisarahnygderopf hfroebqiucegnrcoieuspisfoarcmhaerdacbtyeraishtiycdsrigoncaorfbon

the adsorption of organic inhibitors with a hydrophobic group formed by a hydrocarbon

chain in the structure of their molecule [4,6,15,23]. In the intermediate frequency region,

chain in the structure of their molecule [4,6,15,23]. In the intermediate frequency region, the presence of another linear relationship is observed, log |Z| − log f, whose extension

the presence of another linear relationship is observed, log |Z| − log f, whose extension increases with immersion time from 10 Hz to approximately 0.1 Hz. In the low-frequency

increases with immersion time from 10 Hz to approximately 0.1 Hz. In the low-frequency

region, the low-frequency plateau is observed, whose magnitude of the impedance mod-

region, the low-frequency plateau is observed, whose magnitude of the impedance modulus

ulus increases with immersion time up to approximately two orders of magnitude at the

increases with immersion time up to approximately two orders of magnitude at the end

end of the test. This increase is due to the increase in resistance to load transfer due to the

of the test. This increase is due to the increase in resistance to load transfer due to the

adsorption and formation of a protective film on the steel surface. From the Bode diagram

adsorption and formation of a protective film on the steel surface. From the Bode diagram

in its phase angle format, in the high frequency region the formation and evolution of the

in its phase angle format, in the high frequency region the formation and evolution of the

first-time constant associated with the protective layer of the inhibitor adsorbed on the

first-time constant associated with the protective layer of the inhibitor adsorbed on the

steel surface is observed. Its location is around 2 kHz, reaching a maximum phase angle

steel surface is observed. Its location is around 2 kHz, reaching a maximum phase angle

of 42◦° after approximately 24 h of immersion. The presence of this time constant is also a of 42 after approximately 24 h of immersion. The presence of this time constant is also a

characteristic of organic inhibitors with a hydrophobic group formed by a hydrocarbon

characteristic of organic inhibitors with a hydrophobic group formed by a hydrocarbon

chain [4,6,15,23], however, its position and magnitude of maximum phase angle is lower

chain [4,6,15,23], however, its position and magnitude of maximum phase angle is lower than reported. In the intermediate frequency region, the presence and evolution of the

than reported. In the intermediate frequency region, the presence and evolution of the

second time constant is observed, which corresponds to the capacitive response of the

second time constant is observed, which corresponds to the capacitive response of themeetatallilcicssuurrfafaccee..IIttiisobservedtthattthiistiimeconsttanttmovesffrom40Hztto2Hz,,and its

◦◦

itms amxaimxiummumphpahsaeseanagnlgeleinicnrceraesaessesfrforomm5500° ttoo68° att the end of the tteesstt.. Thiissbbeehhaavvioiorr is

ischcahrarcatcetreirsitsictiocfoofrogragnaicnicnhinibhiitboirtsorwshwerhemreumchucohf tohfetmheolmecoulleecuadlesoardbssoirnbsa ipnlaanaprlafansahrion

f[a1s4h,i2o5n].[T14h,e25c]h.aTrhacetcehriasrtaicsteorfistthicestwofot-htiemtwe coo-tnimstaenctosnssutagngtesstsuthgagteastlathrgateapalarrtgoefpthaertinofhib-

the inhibitor molecule was adsorbed in a planar manner and only part of the length of the

itor molecule was adsorbed in a planar manner and only part of the length of the hydro-

hydrocarbon chains was directed towards the electrolyte.

carbon chains was directed towards the electrolyte.

Taking into account the structure of the inhibitor (Figure 6), it is possible that its ad- Taking into account the structure of the inhibitor (Figure 6), it is possible that its ad-

sorption to the metal surface was carried out from the N-rich group (diaminodiethylamine

sorption to the metal surface was carried out from the N-rich group (diaminodiethylamine

group) to the unsaturations present in the hydrocarbon chains [1,14,23,24]. group) to the unsaturations present in the hydrocarbon chains [1,14,23,24].

Figure 6. Chemical structure of the dialkyl-diamide of coffee bagasse oil (R = alkyl chains). Figure 6. Chemical structure of the dialkyl-diamide of coffee bagasse oil (R = alkyl chains).

ItIthhaassbbeenenrerpeoprotretdedthtahtathtehienihnihbitboirtosrwsiwthitthwtowhoyhdyrodprhopobhiocbcihcacihnasinshsoswhohwigheigrherinhibition efficiencies than those observed with the inhibitors of one and three hydrophobic

inhibition efficiencies than those observed with the inhibitors of one and three hydropho-

chains [13]. In addition, the high content of hydrocarbon chains with unsaturations (≈57%) bic chains [13]. In addition, the high content of hydrocarbon chains with unsaturations

and short hydrocarbon chains (≈36%) could favor the phase angles observed in both time (≈57%) and short hydrocarbon chains (≈36%) could favor the phase angles observed in

constants, and their adsorption mode could occur as suggested in Figure 7.both time constants, and their adsorption mode could occur as suggested in Figure 7.

The proposed scheme is consistent with the findings reported by M.A.J. Mazumder et al. [13] and J. Porcayo-Calderon et al. [14]. On the one hand, despite the fact that the presence of NaCl and CO2 affects the optimal concentration of inhibition, the inhibition mechanism is based on the displacement of the water molecules adsorbed to the metal surface due to the coordinate bond existing between the empty sites of the d-orbitals of Fe, and the π-electrons from the N centers, and on the other hand, the surface coverage is increased by the presence of the existing double bonds in the hydrocarbon chains.

Molecules 2023, 28, 763 9 of 21 Molecules 2023, 28, x FOR PEER REVIEW 10 of 23

Figure 7. Schematic of the inhibitor adsorption mode onto metal surface.

The proposed scheme is consistent with the findings reported by M.A. J. Mazumder et al. [13] and J. Porcayo-Calderon et al. [14]. On the one hand, despite the fact that the presence of NaCl and CO2 affects the optimal concentration of inhibition, the inhibition mechanism is based on the displacement of the water molecules adsorbed to the metal surface due to the coordinate bond existing between the empty sites of the d-orbitals of

Fe, and the π-electrons from the N centers, and on the other hand, the surface coverage is Figure 7. Schematic of the inhibitor adsorption mode onto metal surface.

Figure 7. Schematic of the inhibitor adsorption mode onto metal surface.increased by the presence of the existing double bonds in the hydrocarbon chains.

Based on the behavior observed in the evolution of the impedance spectra, Figure 8 The proposed scheme is consistent with the findings reported by M.A. J. Mazumder

Based on the behavior observed in the evolution of the impedance spectra, Figure 8

shows the equivalent circuits proposed to model the electrochemical behavior of the X52 et al. [13] and J. Porcayo-Calderon et al. [14]. On the one hand, despite the fact that the

shows the equivalent circuits proposed to model the electrochemical behavior of the

steel, both in the absence and in the presence of the corrosion inhibitor. In this case, Rs Xp5r2essetneceel,obfoNthaCinl atnhde CabOs2enacffecatnsdthienotphteimparlesceoncenotfratthioencorfrionshioibnitinonh,ibtihteori.nhIinbithioisn

represents the electrolyte resistance, Rct the charge transfer resistance, Rf the resistance of thecmaseec,hRansirsempriessbenastsedthoenetlhecetrdoilsyptleacremsisetnatnocfe,thRectwtahterchmaorglecutrlaesnsafdersorrebseisdtatnocteh,eRmf tehtael

adsorbed inhibitor film, ZCPEdl the impedance of the constant phase element (CPE) of the dou-rseusrisfatacnecdeuoefttohtehaedcsoorrbdeidnaitnehbibointdorefixilsmti,nZg betwetehne tihmepemdapntycesiotefsthoef tchoendst-aonrbt iptahlassoef

ble layer, and ZCPEf the impedance of the constant phase element of the adsorbed inhibitor eFlem, aenndt t(hCePEπ-)eolef ctthreondsoufrbolme lathyerN, acnednZters, anthde oimn ptheedaonthcerohf athned,ctohnestsaunrtfapcheacsoeveeleramgenits

CPEfoinf cthreaasdedsobrbyetdheinphriebsietonrcefilomf t.he existing double bonds in the hydrocarbon chains.

film.

CPEdl

Based on the behavior observed in the evolution of the impedance spectra, Figure 8 shows the equivalent circuits proposed to model the electrochemical behavior of the X52 steel, both in the absence and in the presence of the corrosion inhibitor. In this case, Rs represents the electrolyte resistance, Rct the charge transfer resistance, Rf the resistance of the adsorbed inhibitor film, ZCPEdl the impedance of the constant phase element (CPE) of the dou- ble layer, and ZCPEf the impedance of the constant phase element of the adsorbed inhibitor film.

 

(a) (b) Figure 8. Equivalent circuits: (a) without inhibitor; (b) with inhibitor.

Figure 8. Equivalent circuits: (a) without inhibitor; (b) with inhibitor.

InInggeenneerraal,l,iittiisscommon to use the CPEiinsstteeaadooffaacacappaacictiotor rtotococmompepnesnastaetfeofrosrusrufarc-efaircregirurelagruitliaersitaiensdanodnn-uoni-fuonrmifodrimstrdibisutrtiobnutoiofnthoefcthaercgheatrgaenstrfearn.sGfern.eGraelnlye,rainllyth,einsethcaeses,cathsesc,atphaecciatipvaecsiteimveicsiercmleisciarcplpesearpdpeparedsseepdre, sasned,thaendetghreedeeogfrdeeporfedsseipornesdseiponendespoendthseopnhtahsepohfatsheoCfPtEhe[2C6P].ET[h2e6]i.mTphedimanpcedoafntcheeoCfPthEeisCdPeEpiesndepnetnodnenthteofnretqhueefnrecqyuaendcyisadnedfiinseddefbinyetdhebeyxtphrese(sxai)porne:ssion: (b)

1Figure 8. Equivalent circuits: (a) withouZt inhib=itor; (b)􏰃with, inhibitor. (2)

Z

=Y (iω) , (2) 􏰀􏰁􏰂 0 􏰉

CPE n

ffa=cefrierqruegenuclayr,itaiensdand=nsolno-puenoifforthme driesltartiibountilongo|f Zth|e −chalorgef.traIfnsnfe=r. 1G, eCnPerEalrleyp, irnestehnetsse quency, and n = slope of the relation log |Z| − log f. If n = 1, CPE represents an ideal

acnasiedse,athlecacpapacaictiotirvwesheemreicYirocle=scaappaecairtadnecper,easnseddi,fand=th0e.5d,eCgPreEeroefpdrepsernestsioanWdaerpbeunrdgscapacitor where Yo = capacitance, and if n = 0.5, CPE represents a Warburg impedance

iomnptehdeapnhceasweiothf tdhieffuCsPiEon[a2l6c].hTahraecitmer.peTdhaenvcaeluofesthoef CthPeEcaispdaceiptaencdenotf othnetdhoeufrbelqeulaeynecry with diffusional character. The values of the capacitance of the double layer and of the

andoisftdhefinehdibbityotrhfielmexpcarnesbsieoonb:tainedfromthevalueofYoaccordingto[27]: inhibitor film can be obtained from the value of Yo according to [27]:

􏰄􏰅 􏰆􏰇􏰈

IIn tgheinsecrasl,ei,t Yisoco=mpmropnotrotiuosnealtihtye CfaPcEtoirn;stiea=d­o−f 1a, cωapa=ci2toπrf t=o caonmgupleanrsafrteqfuorensucyr,- In this case, Yo = proportionality factor; i = √−1, ω = 2πf = angular frequency, f = fre-

􏰃

Z =(1−n)n, (2) 􏰀􏰁􏰂 i 􏰉i

􏰊􏰋1

Ci = Y0iRi 􏰄􏰅􏰆􏰇􏰈 , (3) In this case, Yo = proportionality factor; i = √−1, ω = 2πf = angular frequency, f = fre-

Based on the proposed models, the fit of the impedance spectra showed the results

quency, and n = slope of the relation log |Z| − log f. If n = 1, CPE represents an ideal

of Figures 9 and 10. Figure 9a shows a high coincidence of the calculated Rct values capacitor where Yo = capacitance, and if n = 0.5, CPE represents a Warburg impedance

with those Rp values obtained from the measurements of resistance to linear polarization

with diffusional character. The values of the capacitance of the double layer and of the

(Figure 2a). This suggests that the proposed equivalent circuits correctly represent the inhibitor film can be obtained from the value of Yo according to [27]:

electrode processes occurring on the steel surface. The Rct values are higher than those reported with similar inhibitors obtained from other vegetable oils [19] and imidazolines

𝐶􏰎 =􏰏𝑌􏰐􏰎𝑅􏰎 􏰔 ,

Based on the proposed models, the fit of the impedance spectra showed the results of Figures 9 and 10. Figure 9a shows a high coincidence of the calculated Rct values with those Rp values obtained from the measurements of resistance to linear polarization (Fig-

Molecules 2023, 28, 763ure 2a). This suggests that the proposed equivalent circuits correctly represent the elec- trode processes occurring on the steel surface. The Rct values are higher than those re- ported with similar inhibitors obtained from other vegetable oils [19] and imidazolines

with more than one hydrocarbon chain [13]. On the other hand, from Figure 9b the Rf

10 of 21

with more than one hydrocarbon chain [13]. On the other hand, from Figure 9b the Rf values are lower than those of Rct, and its value is higher at concentrations greater than

values are lower than those of Rct, and its value is higher at concentrations greater than

10 ppm, reaching maximum values between 60–80 ohms-cm2. 2 10 ppm, reaching maximum values between 60–80 ohms-cm .

(a) Rct vs. time

 

(b) Rf vs. timeFigure 9. Rct and Rf values as a function of time and inhibitor concentration.

Figure 9. Rct and Rf values as a function of time and inhibitor concentration.

Regarding the capacitance values, according to Figure 10a, in the absence of the inhibitor, the Cdl values show an increase with immersion time, however, for concentrations of 5 and 10 ppm, the Cdl values show approximately the same values as in the absence of the inhibitor. At concentrations greater than 10 ppm, the Cdl values are lower and tend to decrease with immersion time until reaching approximately the same value. The trend of these values is inversely proportional to the values of Rct, that is, as the values of Rct increase, the values of Cdl tend to decrease. Furthermore, according to the Cdl definition:

Cdl = εεoA (4) d

If the terms εo (vacuum dielectric constant) and A (reaction area) are constant, the only terms that favor a decrease in capacitance are a decrease in the dielectric constant (ε) of the adsorbed film and/or an increase in the thickness of the double layer (d). Therefore, since the inhibitor molecules (low dielectric constant) have replaced the water molecules

Molecules 2023, 28, 763

11 of 21

es 2023, 28, x FOR PEER REVIEW

(high dielectric constant) on the steel surface, this favored the decrease in Cdl values, as

well as an increase in the double layer thickness [14,28–30]. On the other hand, according

to Figure 10b, the Cf values are lower than the Cdl and show a tendency to decrease

with immersion time. For inhibitor concentrations greater than 10 ppm, their values are

the lowest and very close to each other, according to the definition of capacitance, this

suggests that the thickness of the inhibitor film adsorbed to the metal surface is greater for these concentrations.

(a)

12 of 23

 

(b)Figure 10. Cdl and Cf values as a function of time and inhibitor concentration.

Figure 10. Cdl and Cf values as a function of time and inhibitor concentration. 2.4. Potentiodynamic Polarization

Regarding the capacitance values, according to Figure 10a, in the absence of the in- hibitor,theCdlvalOuensceshthoewopantiminaclrienahsiebiwtiiotnhcimonmcenrtsriaotniotnimwea,shdoewtervmerin,efodr,pcontecnetniotrday-namicpolariza-tionsof5andt1i0onppcumrv,ethsewCedrelpvearlufoersmsehdowatadpifpferroexnimtimatemlyertshioensatimesvalnudesteamspinertahteuraebs-.Inallcases,forsence of the inehaicbhitporo.laArtizcaotniocnencturravtieoansnegwreawteorkthinagne1l0ecptrpomde, twheasCudsledv.alues are lower

andtendtodecreasFeigwuirteh1im1amsehroswiosntthimepeoulnatrilzraetaiocnhicnugravpespraotx4i0maCteliynthesaabmsenvcaeluofe.theinhibitor;itThe trend of thiessoebvsaelruveesdistihnavterresgealyrdplreosps orftitohneailmtomtehresvioanluteims oef, Ralclt,ththeactuisr,vaestshheovwaluthees same behavior. of Rct increaseT, thheeavnaolduiecsborfaCncdhl tsehnodwtsotdhaectrtehaesset.eFeulrutnhderemrwoerne,tacncoarcdtiivnegctorrtohseioCndplrdoecfe-ss,thatis,large

inition:

increases in current density by slightly increasing the potential. This behavior is characteris- tic of materials that are not capable of developing a protective oxide. A displacement of the

C =εε􏰘𝐴polarization curves to nobler corrosion potentials is observed as time elapses until 18 h to

𝑑

(4)

subsequently decrease. This can be associated with the inability of the material to develop a stable protective oxide. On the other hand, the corrosion current density values only show

􏰖􏰗

If the terms εo (vacuum dielectric constant) and A (reaction area) are constant, theslight oscillations and are within the same order of magnitude. This confirms that at 40 ◦C

only terms that favor a decrease in capacitance are a decrease in the dielectric constant (ε)the steel undergoes an active corrosion process and is incapable of developing a stable

of the adsorbed film and/or an increase in the thickness of the double layer (d). Therefore,passive layer. At 60 ◦C (Figure 11b), the polarization curves also show a similar behavior

since the inhibitor molecules (low dielectric constant) have replaced the water molecules (high dielectric constant) on the steel surface, this favored the decrease in Cdl values, as well as an increase in the double layer thickness [14,28–30]. On the other hand, according to Figure 10b, the Cf values are lower than the Cdl and show a tendency to decrease with immersion time. For inhibitor concentrations greater than 10 ppm, their values are the

l

lowest and very close to each other, according to the definition of capacitance, this sug-

Molecules 2023, 28, 763

12 of 21

regardless of the immersion time. However, after 3 h of immersion, a shift in the corrosion potential is observed towards nobler values with a slight shift to lower current densities. The anodic branch shows an active corrosion process, although an increase in its slope is observed with increasing immersion time. This may be associated with the precipitation of protective layers on the surface of the working electrode. It is known that from 60 ◦C the formation of protective layers based on iron carbonate become more adherent and protective, thereby reducing the corrosion rate [31]. At 80 ◦C (Figure 11c), the polarization curves show a slight shift to lower current densities but with oscillations in its corrosion potential, being more active at 3 and 12 h. Until 12 h, the anodic branch shows an active

Molecules 2023, 28, x FOR PEER REVIcEoWrrosion process and after 18 h an increase in the slope of the anodic branch is obs14erovfe2d3, which may correspond to the formation of protective corrosion products on the surface of

the working electrode.

(a) (d)

(b) (e)

    

(c) (f)Figure 11. Potentiodynamic polarization curves, as a function of time and temperature, for API-X52

Figure 11. Potentiodynamic polarization curves, as a function of time and temperature, for API-X52 steel in CO2-saturated brine, without and with 25 ppm of inhibitor. (a) 40 °C, 0 ppm. (b) 60 °C, 0

steel in CO2-saturated brine, without and with 25 ppm of inhibitor. (a) 40 ◦C, 0 ppm. (b) 60 ◦C, 0 ppm. ppm. (c) 80 °C, 0 ppm. (d) 40 °C, 25 ppm. (e) 60 °C, 25 ppm. (f) 80 °C, 25 ppm.

(c) 80 ◦C, 0 ppm. (d) 40 ◦C, 25 ppm. (e) 60 ◦C, 25 ppm. (f) 80 ◦C, 25 ppm.

However, in the presence of the inhibitor at 40 °C (Figure 11d), it is observed that after 3 h of immersion, the polarization curves have shifted to more noble potentials (around 250 to 300 mV with respect to Ecorr) and at lower current densities (about an order of magnitude lower). These changes show that the inhibitor has adsorbed on the

Molecules 2023, 28, 763

13 of 21

However, in the presence of the inhibitor at 40 ◦C (Figure 11d), it is observed that after 3 h of immersion, the polarization curves have shifted to more noble potentials (around 250 to 300 mV with respect to Ecorr) and at lower current densities (about an order of magnitude lower). These changes show that the inhibitor has adsorbed on the surface of the working electrode acting as a barrier between the electrolyte and the metal surface, since both the anodic and cathodic currents have decreased. However, even though the anodic branch shows an increase in its slope, it is observed that with an increase of around 100 mV

Molecules 2023, 28, x FOR PEER REVIwEWith respect to the corrosion potential, large increases in current density are gen1e5raotfe2d3.

This is possibly due to detachment of the inhibitor due to low adsorption forces. On the

other hand, at 60 ◦C (Figure 11e), once again it is observed that after 3 h of immersion,

the polarization curves show a shift to more noble potentials (around 100 to 150 mV

orders of magnitude). These displacements indicate inhibitor adsorption and increased

with respect to Ecorr), and at lower current densities (about two orders of magnitude).

protection of the working electrode. The anodic branch shows a greater increase in its

These displacements indicate inhibitor adsorption and increased protection of the working

slope than that observed at 40 °C, however around 150 mV above Ecorr a large increase

electrode. The anodic branch shows a greater increase in its slope than that observed at

in current density is observed, indicating desorption of the inhibitor film. At 80 °C (Figure

40 ◦C, however around 150 mV above Ecorr a large increase in current density is observed, 11f), it is observed that after 3 h of immersion, a displacement of the polarization curves indicating desorption of the inhibitor film. At 80 ◦C (Figure 11f), it is observed that after occurs to nobler potentials and lower current densities (at least by an order of magnitude). 3 h of immersion, a displacement of the polarization curves occurs to nobler potentials and However, it is observed that the corrosion potentials keep oscillating, possibly due to ad- lower current densities (at least by an order of magnitude). However, it is observed that sorption–desorption processes of the inhibitor molecules. It is possible that at this temper- the corrosion potentials keep oscillating, possibly due to adsorption–desorption processes ature the concentration of the inhibitor is not sufficient to guarantee the protection of the of the inhibitor molecules. It is possible that at this temperature the concentration of working electrode. As at 60 °C, the anodic branch shows an increase in its slope, being the inhibitor is not sufficient to guarantee the protection of the working electrode. As at more noticeable after 9 h of immersion. Similarly to what was observed at 60 °C, around 60 ◦C, the anodic branch shows an increase in its slope, being more noticeable after 9 h of 150 mV above Ecorr there is an increase in current density, which may correspond to the immersion. Similarly to what was observed at 60 ◦C, around 150 mV above Ecorr there is desorption of the inhibitor film.

an increase in current density, which may correspond to the desorption of the inhibitor film.

In general, with the addition of the inhibitor, a shift to lower current densities was

In general, with the addition of the inhibitor, a shift to lower current densities was

observed, which in turn caused a change in both slopes (anodic and cathodic), indicating

observed, which in turn caused a change in both slopes (anodic and cathodic), indicating a

a decrease in both the anodic reaction rate (metallic dissolution) and the cathodic reaction

decrease in both the anodic reaction rate (metallic dissolution) and the cathodic reaction

rate (H2 evolution) [32,33].

rate (H2 evolution) [32,33].Figure 12 shows the variation in the Ecorr values obtained from the potentiodynamic

Figure 12 shows the variation in the Ecorr values obtained from the potentiodynamic polarization curves. In the absence of the inhibitor, it is observed that the Ecorr values are

polarization curves. In the absence of the inhibitor, it is observed that the Ecorr values are

more active with increasing temperature, this indicates that the material becomes more

more active with increasing temperature, this indicates that the material becomes more

susceptible to corrosion. However, in the presence of the inhibitor, Ecorr values tend to

susceptible to corrosion. However, in the presence of the inhibitor, Ecorr values tend

be more noble, indicating a tendency to protect. This is more evident at 40 °C and◦ 60 °C, to be more noble, indicating a tendency to protect. This is more evident at 40 C and

however at 80 °C oscillations in the Ecorr values are observed, possibly due to adsorption–

60 ◦C, however at 80 ◦C oscillations in the Ecorr values are observed, possibly due to desorption processes of the inhibitor molecules.

adsorption–desorption processes of the inhibitor molecules.

Figure 12. Corrosion potential variation for API-X52 steel in 3% NaCl solution saturated with CO2, Figure 12. Corrosion potential variation for API-X52 steel in 3% NaCl solution saturated with CO2,

determined from potentiodynamic polarization curves.

determined from potentiodynamic polarization curves.

Figure 13 shows the variation in the Icorr values, obtained from the potentiodynamic polarization curves, as a function of time and temperature. From the graph in the absence of the inhibitor the steel shows Icorr values of the same order of magnitude regardless of the test temperature, with a tendency to decrease with immersion time and increase with

Molecules 2023, 28, 763

14 of 21

Figure 13 shows the variation in the Icorr values, obtained from the potentiodynamic

polarization curves, as a function of time and temperature. From the graph in the absence

of the inhibitor the steel shows Icorr values of the same order of magnitude regardless of

the test temperature, with a tendency to decrease with immersion time and increase with

temperature. On the other hand, in the presence of the inhibitor, it is observed that the Icorr

Molecules 2023, 28, x FOR PEER REVIEW 16 of 23 values tend to decrease by more than one order of magnitude in the first 3 h of immersion

and subsequently remain in the same order of magnitude regardless of the temperature. Since the values of Icorr are inversely proportional to the resistance to polarization (Rp), it

observed that the trend in the values of Icorr adjusts to that observed with the values of Rp

is observed that the trend in the values of Icorr adjusts to that observed with the values

(Figure 2a).

of Rp (Figure 2a).

Figure13.CorrosiionccurrreennttddeennsistiytyvvarairaitaiotinonfofroArPAIP-XI-5X25s2tesetleeinli3n%3N%aNCalCsollsuotilountiosantusaratuterdatwedithwith CO2,, determined frompotteenntitoioddyynnaammicicppoloalrairziaztaiotinoncucruvrevs.es.

2The effect of temperature on the corrosion process can be explained with the use of

The effect of temperature on the corrosion process can be explained with the use of

the Arrhenius equation, where the activation parameters of both the anodic branch and the

the Arrhenius equation, where the activation parameters of both the anodic branch and

cathodic branch can be calculated by means of the following equation [34,35]:the cathodic branch can be calculated by means of the following equation [34,35]:

(5)whereEa((JJmooll)is)tihsethaectaivcatitvioantioeneerngeyrogfythoefcthoerrcoosrioronspiornocpersos;cRestsh;eRutnhieveurnsaivlegrasaclognastant

Icorr = k exp − Icorr = kexp􏰙−

􏰛

(5) (J mol−1 K−1); k a constant; and Icorr is the corrosion current density (A cm−2). The activation

RT RTconstant (J mol−1 K−1); k a constant; and Icorr is the corrosion current density (A cm−2).

−1 −1

The activation energy in the absence and presence of the inhibitor was determined by

energy in the absence and presence of the inhibitor was determined by the slope of the graph

the slope of the graph of Rln(Icorr) against 1/T (Figure 14), and the results obtained are of Rln(Icorr) against 1/T (Figure 14), and the results obtained are shown in Table 1.

shown in Table 1.The enthalpy (∆Ha◦) and entropy (∆Sa◦) change can be obtained by the transition

state equation, an alternative form of the Arrhenius equation [22,34,36]: RT 􏰌 ∆S◦ 􏰍 􏰌 ∆H◦ 􏰍

􏰌 Ea􏰍E􏰚

I = Nh exp R exp − RT (6)

where h is Planck’s constant (6.626176 × 10−34 J s); N Avogadro’s number (6.023 × 1023 mol−1). Plotting ln(Icorr/T) against 1/T yields a straight line whose slope is, (−∆H◦/R), with in- tercept of (ln(R/Nh) + ∆S◦/R) (Figure 15). Table 1 reports the values of Ea, ∆H◦, and ∆S◦, as a function of the immersion time of the corrosion process without the addition of the inhibitor and with the addition of 25 ppm of the inhibitor.

ules 2023, 28, x FOR PEER REVIEW Molecules 2023, 28, 763

17 of 23

15 of 21

Time (h)

(J mol−1) (J mol−1)

(J mol−1)(J mol−1K−1)

11,−029184

(J mol−1K−1) (J mol−1)

(J mol−1) (J mol−1)

(J mol−1)(J mol−1K−1)

3 13,853

11,091

−285

−294

33,542

19,925 −285

96 131,2 87

11,590

8828 10,−127688

22,687 215,6,6060

TimEea(h) (J mol−1)

Ea

Ea

18

12,672

9910

−294 22,687

17,098

14,337 −308 −285

0 ppm

0 ppm ∆H◦

∆S◦ Ea

25 ppm

25 ppm ∆H◦

∆S◦(J mol−1K−1)

ΔH°

ΔS°

ΔH°

ΔS°

102,953286

−2289,1422

18,84−5263 −291

(a)

(b)Figure 14. ArFrihgeunrieu1s4p. lAotr,rhfoernituhse pclaolct,ufloartitohne ocfalEcua,latsioan foufnEcati,oans oaffuimncmtieornsiofnimtime:rs(aio)nwtitmhoe:u(ta) without

inhibitor, (b) with 25 ppm inhibitor.inhibitor, (b) with 25 ppm inhibitor.

Table 1. ActivTatbiolen1p.aAracmtiveatteirosnopfatrhaemceotrerrossoiofnthpercoocrersossoiofnApPrIo-cXe5s2s sotfeAelPiIn-Xth5e2 astbeseelnincethaendabpsreensceenacned presence ofinhibitor. ofinhibitor.

00 13,3237

13,237 103,845735

10,475

−284

13,237 3160,3,40475

10,475 −284

33,54−284 −232

6

13,287

10,526

−288 36,304

28,422

25,660 −263 −232

9 11,590

24

12 12,93818 12,672 9910 −294 17,098 14,337 −308

8828

−294

−296 21,606

19,925

10,643 −323 −291

12,027

9266

13,404

10,176

−291

18,845

−1238,5237

According to the calculated values, it is observed that in the absence of the inhibitor,

24 12,027 9266 −296 13,404 10,643 −323

the activation energy oscillated between 12 and 14 kJ/mol during the entire test. The magnitude of Ea represents the energy barrier necessary for metal dissolution to take

The enthalpy (ΔHa°) and entropy (ΔSa°) change can be obtained by the transition

place. Ea values around 60–80 kJ/mol have been reported for the dissolution of Fe

state equation, an alternative form of the Arrhenius equation [22,34,36]:

in H2SO4 solutions [22,33,36,37], and about 40 kJ/mol in HCl solutions [38]. In elec-

ules 2023, 28, x FOR PEER REVIEW Molecules 2023, 28, 763

18 of 23

I = RT exp 􏰙∆S°􏰛 exp 􏰙− ∆H°􏰛trolytes such as thoseNehvaluatedRhere, a wide RraTnge of Ea values is reported, ranging

from 8 to 23 kJ/mol [39–45] up to 30–33 kJ/mol [2,3,24], depending on the NaCl content where h is Planck’s constant (6.626176 × 10−34 J s); N Avogadro’s number (6.023 × 1023

of the brine, in general being that the lowest values are reported for concentrations below

mol−1). Plotting ln(Icorr/T) against 1/T yields a straight line whose slope is, (−∆H°/R), with

3.5% and the highest for concentrations of 5%. The low activation energy determined here

intercept of (ln(R/Nh) + ΔS°/R) (Figure 15). Table 1 reports the values of Ea, ΔH°, and ΔS°,

suggests a greater aggressiveness of the CO2-saturated brine electrolyte, possibly due to a

as a function of the immersion time of the corrosion process without the addition of the

greater conductivity due to the ionic species present.

inhibitor and with the addition of 25 ppm of the inhibitor.

(a)

(b)Figure 15. ArFrihgeunrieu1s5p.lAotr,rhfoerntihues pcalolct,ufloartitohneocfalΔcuHla°taionndoΔfS∆°H, asaanfdu∆ncSti,oansoaffiumnmctieornsiofnimtime:r(saio)n time: (a)

without inhibitor, (b) with 25 ppm inhibitor.without inhibitor, (b) with 25 ppm inhibitor.

16 of 21

(6)

 

◦◦

However, in the presence of the inhibitor, the activation energy increased to 36.3 kJ/mol

According to the calculated values, it is observed that in the absence of the inhibitor,

two hours after the inhibitor had been added, and subsequently its value tended to de-

the activation energy oscillated between 12 and 14 kJ/mol during the entire test. The mag-

crease until it reached a value similar to that obtained in the absence of the inhibitor. The

nitude of Ea represents the energy barrier necessary for metal dissolution to take place.

increase in the value of Ea has been associated with an initial stage of physical adsorption

Ea values around 60–80 kJ/mol have been reported for the dissolution of Fe in H2SO4 so-

of the inhibitor, and its decrease or low variation to a chemisorption process, considering

lutions [22,33,36,37], and about 40 kJ/mol in HCl solutions [38]. In electrolytes such as

that part of the energy is used for the chemical reaction [22,33,37]. Other studies suggest those evaluated here, a wide range of Ea values is reported, ranging from 8 to 23 kJ/mol [39–

that the decrease in the value of Ea is due to a slow adsorption of the inhibitor or to a

45] up to 30–33 kJ/mol [2,3,24], depending on the NaCl content of the brine, in general being

change in the net corrosion reaction between the unprotected part with respect to the

that the lowest values are reported for concentrations below 3.5% and the highest for concen-

protected one, that is, the reaction speed in the protected area is substantially less than in

trations of 5%. The low activation energy determined here suggests a greater aggressiveness

the unprotected area [22,38].

Molecules 2023, 28, 763

17 of 21

However, it should be noted that many of the studies report the calculation of Ea from measurements made after adding the inhibitor to the electrolyte and its magnitude with respect to the value of Ea without the inhibitor is associated with the type of ad- sorption of the inhibitor (physical or chemical). For example, for the dissolution of Fe in acid media, increases in the value of Ea are reported and it is associated with an in- crease in the energy barrier for corrosion to take place [33,36–38], and others report a decrease and associate it with a chemisorption process [22]. In media similar to the one used here, increases (30–50 kJ/mol) have been reported, and these have been associated with a physisorption process, an increase in the energy barrier, or an increase in the thick- ness of the double layer [24,39,40,43,45], and decreases (5–20 kJ/mol), to a chemisorption process [2,3,24,41,42,44].

It has also been observed [24] that the Ea value initially increases and subsequently decreases with the increase in inhibitor concentration, and this has been associated with a change in the adsorption mode, from physical to chemical adsorption. At low concen- trations, there is a blockage by physisorption of the active sites because the three N atoms (positive charge) interact with the metallic surface (negative charge), in addition to the fact that the double bonds of the hydrocarbon chain can also interact with the surface through its π electrons, and with an increasing concentration the electron-rich amine groups block the active sites by chemisorption.

Then, based on the above, it can be said that the observed increase in Ea values is due to the fact that the inhibitor is initially adsorbed by a physical adsorption process and the subsequent decrease is due to a change in the type of adsorption, that is say from physical adsorption to chemical adsorption.

Regarding the activation enthalpy, in the absence of the inhibitor, values between 9–11 kJ/mol were obtained during the test period. The positive values obtained indicate that the metal dissolution process is endothermic [22], ∆H◦ values around 60 kJ/mol have been reported for the dissolution of Fe in H2SO4 solutions [22,36], and 40 kJ/mol in HCl solutions [38]. In media similar to those evaluated here, a wide range of ∆H◦ values is reported, ranging between 5–15 kJ/mol [42–44]. The low enthalpy value determined here also suggests that the dissolution of Fe in CO2-saturated brine is an endothermic process.

On the other hand, in the presence of the inhibitor, ∆H◦ increases up to 33.5 kJ/mol after two hours of the inhibitor having been added and subsequently its value tends to decrease until it reaches a value similar to that obtained in the absence of the inhibitor. ∆H◦ values for the dissolution of Fe in acidic media are reported shortly after adding the inhibitor in the order of 50–90 kJ/mol [22,36,38]. In corrosive media similar to the one used here, ∆H◦ values of the order of 5–33 kJ/mol have been reported [42–44]. In general, in both cases (with and without inhibitor) it is observed that the Ea and ∆H◦ values show the same trend as well as very similar values. This agrees with the proposal that for a chemical reaction in solution both values should theoretically be equal [44].

Finally, with respect to entropy, in the absence of the inhibitor, ∆S◦ values ranging between −280 and −300 J/mol-K were obtained throughout the test. Positive values have been associated with an increase in the order of the activated complex and that this is the rate-limiting step and that it represents an association step [22,34,36,38]. ∆S◦ values between −50 and −120 J/mol-K have been reported for the dissolution of Fe in H2SO4solutions [22,36], and about −180 J/mol-K in HCl solutions [38]. In media similar to the one used here, ∆S◦ values of the order of −150 to −170 J/mol-K have been reported [42,44].

In the presence of the inhibitor, ∆S◦ increases to −232 J/mol-K after two hours of being added and subsequently its value tends to decrease constantly up to a value of −323 J/mol-K at the end of the test. ∆S◦ values for the dissolution of Fe in acid media are reported shortly after adding the inhibitor between −50 and −120 J/mol-K [22,36,38]. In media similar to the one used here, ∆S◦ values of the order of −160 and −221 J/mol-K have been reported [42,44]. The negative values of ∆S◦, with and without the presence of the inhibitor, suggest that the complex activated in the rate-limiting step represents an associative rather than a dissociative step, and this implies an increase in order as the

Molecules 2023, 28, 763

18 of 21

reaction goes from reagents to the activated complex [42,44]. It has also been reported that when the concentration of the inhibitor increases, the ∆S◦ values decrease due to a higher order of the Fe-inhibitor complex, and that when its value is more negative it is because the metallic surface is more homogeneous. This because of a more orderly and stably adsorbed inhibitor film [44]. Then, based on the above, the trend observed in this study suggests that the inhibitor adsorption increases with immersion time, forming a highly ordered and stable film on the steel surface.

3. Materials and Methods

The tests were carried out with cylindrical working electrodes (reaction area 4.53 cm2) made of API-X52 steel. Prior to each test, the working electrodes were roughened with emery paper up to 600 grits, and later washed with distilled water, and finally cleaned with acetone and dried.

For the electrochemical tests, an electrochemical cell with three electrodes was used, where the reference electrode was one of saturated calomel (SCE), a graphite rod was used as the auxiliary electrode, and API-X52 steel was used as the working electrode (WE). The corrosion tests were carried out at 60 ◦C in a sodium chloride solution (3% by weight) saturated with CO2.

The evaluated inhibitor was a dialkyl-diamide from coffee bagasse oil that, according to its synthesis process [17,19,46–48], has a general structure as shown in Figure 6. The diaminodiethylamine group of the molecule is the electron-rich hydrophilic section that promotes its adsorption to the metal surface and hydrocarbon chains is the hydrophobic region that decreases the wettability of the steel surface [13].

The inhibitor is a mixture of dialkyl-diamide molecules where 49% corresponds to linoleic acid, 36% to palmitic acid, 8% to oleic acid, and 7% to stearic acid. The proportion of hydrocarbon chains is consistent with the composition of fatty acids reported for similar inhibitors in previous studies [4–7]. The concentration of inhibitor used in the corrosion tests was 5, 10, 25, 50, and 100 ppm.

Inhibitor performance was determined by open circuit potential (OCP), linear polar- ization resistance (LPR), and electrochemical impedance spectroscopy (EIS) measurements. The OCP measurements were made at one-hour intervals for 24 h, the LPR measurements, at one-hour intervals for 24 h, for which the working electrode was polarized ±10 mV with respect to its open circuit potential at a scan rate of 10 mV/min, and EIS studies applying a 10 mV peak-to-peak signal with respect to the OCP value in the frequency range of 100 kHz to 0.01 Hz.

Finally, because it has been observed that the amide and imidazoline inhibitors derived from fatty acids reach their maximum surface coverage between 4 and 12 h after their addi- tion, depending on the origin of the fatty acids, type of fatty acids, and proportion of fatty acids [5,8,23,25,49], once the optimal dose of inhibition was determined, potentiodynamic polarization curves were performed, at different immersion times of the working electrode, at a scan rate of 1 mV/s from −400 mV to +600 mV with respect to the corrosion potential. Each polarization curve was obtained with new working electrodes at different immersion times (0, 1, 3, 6, 9, 12, 18, and 24 h). The electrochemical parameters (Ecorr, Icorr, anodic, and cathodic slopes) were determined according to the Tafel extrapolation procedure.

4. Conclusions

A green dialkyl-diamide inhibitor from coffee bagasse oil was evaluated as a sweet rust inhibitor. From the OCP measurements, a rapid increase in OCP values was observed at the time of its addition, thereby causing a more noble behavior of the metal surface. LPR measurements showed a large increase in Rp values indicating that a protective film is developing on the metal surface due to the adsorption of inhibitor molecules, and that the optimal inhibition concentration is 25 ppm with an inhibition efficiency of 99%. In addition, by means of EIS measurements, the optimal concentration of inhibition was confirmed, and the Rct values obtained showed the same behavior as the Rp values. The evolution of the

Molecules 2023, 28, 763

19 of 21

References

impedance spectra suggests that the adsorption of the inhibitor to the metal surface occurs from the diaminodiethylamine group to the unsaturations present in the hydrocarbon chains. The potentiodynamic polarization tests carried out at different immersion times have shown that when the inhibitor is added, the values of Ea and ∆H increase, but as the immersion time elapses, and with it the surface coverage, these tend to decrease to values similar to those observed in the absence of the inhibitor, in addition to the fact that the values of ∆S tend to be more negative because of the formation of a more stable and uniform protective layer on the metallic surface.

Author Contributions: Data curation, N.B.G.-G.; Formal analysis, N.B.G.-G.; Investigation, N.B.G.-G., A.N. and J.P.-C.; Methodology, J.C., L.M.M.-d.-l.-E., A.N. and J.P.-C.; Resources, J.C. and L.M.M.-d.-l.-E.; Supervision, A.N. and J.P.-C.; Validation, L.M.M.-d.-l.-E.; Writing—original draft, J.P.-C.; Writing—review and editing, J.C. and J.P.-C. All authors have read and agreed to the published version of the manuscript.

Funding: This research received no external funding. Institutional Review Board Statement: Not applicable. Informed Consent Statement: Not applicable.Data Availability Statement: Not applicable.

Conflicts of Interest: The authors declare no conflict of interest.Sample Availability: Samples of the compounds are not available from the authors.

Yoo, S.-H.; Kim, Y.-W.; Chung, K.; Baik, S.-Y.; Kim, J.-S. Synthesis and corrosion inhibition behavior of imidazoline derivatives based on vegetable oil. Corros. Sci. 2012, 59, 42–54. [CrossRef]

Jevremovi, I.; Singer, M.; Nesic, S.; Miskovic-Stankovic, V. Electrochemistry of carbon dioxide corrosion mitigation using tall oil diethylenetriamine imidazoline as corrosion inhibitor for mild Steel. Corros. Mater. 2016, 67, 756–768. [CrossRef]

Abbasov, V.M.; El-Lateef, H.M.A.; Aliyeva, L.I.; Qasimov, E.E.; Ismayilov, I.T.; Khalaf Abbasov, M.M. A study of the corrosion inhibition of mild steel C1018 in CO2-saturated brine using some novel surfactants based on corn oil. Egypt. J. Pet. 2013, 22, 451–470. [CrossRef]

Porcayo-Calderon, J.; Martínez de la Escalera, L.M.; Canto, J.; Casales-Diaz, M. Imidazoline Derivatives Based on Coffee Oil as CO2 Corrosion Inhibitor. Int. J. Electrochem. Sci. 2015, 10, 3160–3176.

Velazquez-Torres, N.; Martinez, H.; Porcayo-Calderon, J.; Vazquez-Velez, E.; Gonzalez-Rodriguez, J.G.; Martinez-Gomez, L. Use of an amide-type corrosion inhibitor synthesized from the coffee bagasse oil on the corrosion of Cu in NaCl. Green Chem. Lett. Rev. 2018, 11, 1–11. [CrossRef]

Gomez-Guzman, N.B.; Martinez de la Escalera, D.M.; Porcayo-Calderon, J.; Gonzalez-Rodriguez, J.G.; Martinez-Gomez, L. Performance of an Amide-Based Inhibitor Derived from Coffee Bagasse Oil as Corrosion Inhibitor for X70 Steel in CO2-Saturated Brine. Green Chem. Lett. Rev. 2019, 12, 49–61. [CrossRef]

Gonzalez-Rodriguez, J.G.; Gomez-Guzman, N.B.; Porcayo-Calderon, J. Corrosion Inhibition of X70 Pipeline Steel Under Hydro- dynamic Conditions of CO2 with Amide Extraction from Coffee Bagasse. J. Bio-Tribo-Corros. 2021, 7, 86. [CrossRef]

Reyes-Dorantes, E.; Zúñiga-Díaz, J.; Quinto-Hernández, A.; Porcayo-Calderon, J.; Gonzalez-Rodriguez, J.G.; Pedraza-Basulto, G.K.; Martínez-Gomez, L. Rice Bran as Source for the Synthesis of Imidazoline-type Inhibitors: Synthesis and Corrosion Performance. Int. J. Electrochem. Sci. 2018, 13, 101–118. [CrossRef]

de Damborenea, J.; Bastidas, J.M.; Vaquez, A.J. Adsorption and inhibitive properties of four primary aliphatic amines on mild steel in 2 M hydrochloric acid. Electrochim. Acta 1997, 42, 455. [CrossRef]

Kahyarian, A.; Brown, B.; Nesic, S. The Unified Mechanism of Corrosion in Aqueous Weak Acids Solutions: A Review of the Recent Developments in Mechanistic Understandings of Mild Steel Corrosion in the Presence of Carboxylic Acids, Carbon Dioxide, and Hydrogen Sulfide. Corrosion 2020, 6, 268–278. [CrossRef] [PubMed]

Kahyarian, A.; Nesic, S. On the mechanism of carbon dioxide corrosion of mild steel: Experimental investigation and mathematical modeling at elevated pressures and nonideal solutions. Corros. Sci. 2020, 173, 108719. [CrossRef]

Wright, R.F.; Brand, E.R.; Ziomek-Moroz, M.; Tylczak, J.H.; Ohodnicki, P.R., Jr. Effect of HCO-3 on electrochemical kinetics of carbon steel corrosion in CO2-saturated brines. Electrochim. Acta 2018, 290, 626–638. [CrossRef]

Mazumder, M.A.; Nazal, M.K.; Faiz, M.; Ali, S.A. Imidazolines containing single-, twin-and triple-tailed hydrophobes and hydrophilic pendants (CH2CH2NH)nH as inhibitors of mild steel corrosion in CO2-0.5 M NaCl. RSC Adv. 2016, 6, 12348–12362. [CrossRef]

Molecules 2023, 28, 763 20 of 21

Porcayo-Calderon, J.; Regla, I.; Vazquez-Velez, E.; Martinez de la Escalera, L.M.; Canto, J.; Casales-Diaz, M. Effect of the unsaturation of the hydrocarbon chain of fatty-amides on the CO2-corrosion of carbon steel using EIS and real-time corrosion measurement. J. Spectrosc. 2015, 2015, 184140. [CrossRef]

Salinas-Solano, G.; Porcayo-Calderon, J.; Martínez de la Escalera, L.M.; Canto, J.; Casales-Diaz, M.; Sotelo-Mazon, O.; Henao, J.; Martinez-Gomez, L. Development and evaluation of a green corrosion inhibitor based on rice bran oil obtained from agro- industrial waste. Ind. Crops Prod. 2018, 119, 111–124. [CrossRef]

Sotelo-Mazon, O.; Valdez-Rodriguez, S.; Porcayo-Calderon, J.; Casales-Diaz, M.; Henao, J.; Salinas-Solano, G.; Valenzuela-Lagarda, J.L.; Martinez-Gomez, L. Corrosion protection of 1018 carbon steel using an avocado oil-based inhibitor. Green Chem. Lett. Rev. 2019, 12, 255–270. [CrossRef]

Sanchez-Salazar, E.; Vazquez-Velez, E.; Uruchurtu, J.; Porcayo-Calderon, J.; Casales, M.; Rosales-Cadena, I.; Lopes-Cecenes, R.; Gonzalez-Rodriguez, J.G. Use of a Gemini-surfactant synthesized from the mango seed oil as a CO2-corrosion inhibitor for X-120 steel. Materials 2021, 14, 4206. [CrossRef] [PubMed]

Salinas-Solano, G.; Porcayo-Calderon, J.; Larios-Galvez, A.K.; Gonzalez-Rodriguez, J.G. Pouteria sapota as green CO2-corrosion inhibition of carbon steel. J. Electrochem. Sci. Eng. 2022, 12, 383–398. [CrossRef]

Cruz-Zabalegui, A.; Vazquez-Velez, E.; Galicia-Aguilar, G.; Casales-Diaz, M.; Lopez-Sesenes, R.; Gonzalez-Rodriguez, J.G.; Martinez-Gomez, L. Use of a non-ionic gemini-surfactant synthesized from the wasted avocado oil as a CO2- corrosion inhibitor for X-52 steel. Ind. Crops Prod. 2019, 133, 203–211. [CrossRef]

Gurappa, I. Characterization of different materials for corrosion resistance under simulated body fluid conditions. Mater. Charact. 2002, 49, 73–79. [CrossRef]

Barker, R.; Burkle, D.; Charpentier, T.; Thompson, H.; Neville, A. A review of iron carbonate (FeCO3) formation in the oil and gas industry. Corros. Sci. 2018, 142, 312–341. [CrossRef]

Bouklah, M.; Hammouti, B.; Lagrene, M.; Bentiss, F. Thermodynamic properties of 2,5-bis(4-methoxyphenyl)-1,3,4-oxadiazole as a corrosion inhibitor for mild steel in normal sulfuric acid medium. Corros. Sci. 2006, 48, 2831–2842. [CrossRef]

Porcayo-Calderon, J.; Rivera-Muñoz, E.M.; Peza-Ledesma, C.; Casales-Diaz, M.; Martínez de la Escalera, L.M.; Canto, J.; Martinez-Gomez, L. Sustainable Development of Palm Oil: Synthesis and Electrochemical Performance of Corrosion Inhibitors. J. Electrochem. Sci. Technol. 2017, 8, 133–145. [CrossRef]

Desimone, M.P.; Gordillo, G.; Simison, S.N. The effect of temperature and concentration on the corrosion inhibition mechanism of an amphiphilic amido-amine in CO2 saturated solution. Corros. Sci. 2011, 53, 4033–4043. [CrossRef]

Godavarthi, S.; Porcayo-Calderon, J.; Casales-Diaz, M.; Vazquez-Velez, E.; Neri, A.; Martinez-Gomez, L. Electrochemical Analysis and Quantum Chemistry of Castor Oil-Based Corrosion Inhibitors. Curr. Anal. Chem. 2016, 12, 476–488. [CrossRef]

Ferrer, J.E.; Victori, L. Oxygen evolution reaction on the iridium electrode in basic medium studied by electrochemical impedance spectroscopy. Electrochim. Acta 1994, 39, 581–588. [CrossRef]

Ma, H.; Cheng, X.; Li, G.; Chen, S.; Quan, Z.; Zhao, S.; Niu, L. The influence of hydrogen sulfide on corrosion of iron under different conditions. Corros. Sci. 2000, 42, 1669–1683. [CrossRef]

Musa, A.Y.; Jalgham, T.T.R.; Mohamad, A.B. Molecular dynamic and quantum chemical calculations for phthalazine derivatives as corrosion inhibitors of mild steel in 1 M HCl. Corros. Sci. 2012, 56, 176–183. [CrossRef]

Solmaz, R. Investigation of the inhibition effect of 5-((E)-4-phenylbuta-1, 3-dienylideneamino)-1, 3, 4-thiadiazole-2-thiol Schiff base on mild steel corrosion in hydrochloric acid. Corros. Sci. 2010, 52, 3321–3330. [CrossRef]

Ahamad, I.; Quraishi, M.A. Bis(benzimidazol-2-yl) disulphide: An efficient water soluble inhibitor for corrosion of mild steel in acid media. Corros. Sci. 2009, 51, 2006–2013. [CrossRef]

Hernandez, J.; Muñoz, A.; Genesca, J. Formation of iron-carbonate scale-layer and corrosion mechanism of API X70 pipeline steel in carbon dioxide-saturated 3% sodium chloride. Afinidad 2012, 69, 251–258.

Zhang, H.-H.; Pang, X.; Zhou, M.; Liu, C.; Wei, L.; Gao, K. The behavior of pre-corrosion effect on the performance of imidazoline- based inhibitor in 3 wt.% NaCl solution saturated with CO2. Appl. Surf. Sci. 2015, 356, 63–72. [CrossRef]

Vracˇar, L.M.; Dražic ́, D.M. Adsorption and corrosion inhibitive properties of some organic molecules on iron electrode in sulfuric acid. Corros. Sci. 2002, 44, 1669–1680. [CrossRef]

El Rehim, S.S.A.; Hassan, H.H.; Amin, M.A. Corrosion inhibition of aluminym by 1,1 (laury amido) propyl ammonium chloride in HCl solution. Mater. Chem. Phys. 2001, 70, 64–72. [CrossRef]

Porcayo-Calderon, J.; Martinez de la Escalera, L.M.; Canto, J.; Casales-Diaz, M.; Salinas-Bravo, V.M. Effect of the Temperature on the CO2-Corrosion of Ni3Al. Int. J. Electrochem. Sci. 2015, 10, 3136–3151.

Gholami, M.; Danaee, I.; Maddaht, M.H.; Rashvand Avei, M. Corralated ab Initio and Electroanalytical Study on inhibition Behavior of 2-Mercaptobenzothiazole and Its Thiole-Thione Tautomerism Effect for the Corrosion of Steel (API 5L X52) in Sulphuric Acid Solution. Ind. Eng. Chem. Res. 2013, 52, 14875–14889. [CrossRef]

Oguzie, E.E.; Onuoha, G.N.; Onuchukwu, A.I. Inhibitory mechanism of mild steel corrosion in 2M sulphuric acid solution by methylene blue dye. Mater. Chem. Phys. 2005, 89, 305–311. [CrossRef]

Noor, E.A.; Al-Moubaraki, A.H. Thermodynamic study of metal corrosion and inhibitor adsorption processes in mild steel/1-methyl-4[4À(-X)-styryl pyridinium iodides/hydrochloric acid systems. Mater. Chem. Phys. 2008, 110, 145–154. [CrossRef]

Ibrahim, T.; Gomes, E.; Obot, I.B.; Khamis, M.; Zour, M.A. Corrosion inhibition of mild steel by Calotropis procera leaves extract

in a CO2 saturated sodium chloride solution. J. Adhes. Sci. Techno. 2016, 30, 2523–2543. [CrossRef]

Molecules 2023, 28, 763 21 of 21

Zhang, G.; Chen, C.; Lu, M.; Chai, C.; Wu, Y. Evaluation of inhibition efficiency of an imidazoline derivative in CO2-containing aqueous solution. Mater. Chem. Phys. 2007, 105, 331–340. [CrossRef]

Zhang, H.-H.; Gao, K.; Yan, L.; Pang, X. Inhibition of the corrosion of X70 and Q235 steel in CO2-saturated brine by imidazoline- based inhibitor. J. Electroanal. Chem. 2017, 791, 83–94. [CrossRef]

Usman, B.J.; Gasem, Z.-M.; Umoren, S.A.; Solomon, M.M. Eco-friendly 2-Thiobarbituric acid as a corrosion inhibitor for API 5L X60 steel in simulated sweet oilfield environment: Electrochemical and surface analysis studies. Sci. Rep. 2019, 9, 1–17. [CrossRef] [PubMed]

Peimania, A.; Nasr-Esfahani, M. Application of anise extract for corrosion inhibition of carbon steel in CO2 saturated 3.0% NaCl solution. Prot. Met. Phys. Chem. 2018, 54, 122–134. [CrossRef]

Zhang, H.-H.; Pang, X.; Gao, K. Effect of surface roughness on the performance of thioureido imidozaline inhibitor in CO2- saturated brine. Corros. Sci. 2019, 157, 189–204. [CrossRef]

Wang, B.; Du, M.; Zhang, J.; Li, C.; Liu, J.; Liu, H.; Lia, R.; Li, Z. Corrosion inhibition of mild steel by the hydrolysate of an imidazoline-based inhibitor in CO2-saturated solution. RSC Adv. 2019, 9, 36546–36557. [CrossRef]

Bistline, R.G.; Hampson, J.W.; Lin Field, W.M. Synthesis and properties of fatty imidazolines and their N-(2-aminoethyl) derivatives. J. Am. Oil Chem. Soc. 1983, 60, 823–828. [CrossRef]

Wu, Y.; Herrington, P.R. Thermal reactions of fatty acids with diethylene triamine. J. Am. Oil Chem. Soc. 1997, 74, 61–64. [CrossRef]

Duda, Y.; Govea-Rueda, R.; Galicia, M.; Beltrán, H.I.; Zamudio-Rivera, L.S. Corrosion Inhibitors: Design, Performance, and

Computer Simulation. J. Phys. Chem. B 2005, 109, 22674–22684. [CrossRef]

Reyes-Dorantes, E.; Zuñiga-Díaz, J.; Quinto-Hernandez, A.; Porcayo-Calderon, J.; Gonzalez-Rodriguez, J.G.; Martinez-Gomez, L.

Fatty Amides from Crude Rice Bran Oil as Green Corrosion Inhibitors. J. Chem. 2017, 2017, 2871034. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Written by

CORROSION Y PROTECCION

Tags

Previous Sweet Corrosion Inhibition by CO2 Capture
Next Internal Corrosion Solution for Gathering Production Gas Pipelines Involving Palm Oil Amide Based Corrosion Inhibitors.