Inhitor Coco MEA 103molecules
Inhitor Coco MEA 103molecules
Inhitor Coco MEA 103molecules
net/publication/368308790
CITATIONS READS
2 139
6 authors, including:
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Omar Dagdag on 07 February 2023.
1 Department of Chemistry, School of Chemical Engineering and Physical Sciences, Lovely Professional
University, Phagwara 144402, India
2 Centre for Materials Science, College of Science, Engineering and Technology, University of South Africa,
Johannesburg 1710, South Africa
3 Department of Chemistry, Faculty of Natural and Mathematics Science, University of Prishtina,
10000 Prishtina, Kosovo
4 NCE, Department of Science and Technology, Government of Bihar, Patna 803108, India
5 Interdisciplinary Research Center for Advanced Materials, King Fahd University of Petroleum and Minerals,
Dhahran 31261, Saudi Arabia
* Correspondence: [email protected] (E.E.E.); [email protected] (A.K.)
Abstract: Recent studies indicate that surfactants are a relatively new and effective class of corrosion
inhibitors that almost entirely meet the criteria for a chemical to be used as an aqueous phase
corrosion inhibitor. They possess the ideal hydrophilicity to hydrophobicity ratio, which is crucial for
effective interfacial interactions. In this study, a coconut-based non-ionic surfactant, namely, coco
monoethanolamide (CMEA), was investigated for corrosion inhibition behaviour against mild steel
(MS) in 1 M HCl employing the experimental and computational techniques. The surface morphology
was studied employing the scanning electron microscope (SEM), atomic force microscope (AFM),
and contact measurements. The critical micelle concentration (CMC) was evaluated to be 0.556 mM
Citation: Ganjoo, R.; Sharma, S.;
and the surface tension corresponding to the CMC was 65.28 mN/m. CMEA manifests the best
Sharma, P.K.; Dagdag, O.; Berisha, A.;
inhibition efficiency (η%) of 99.01% at 0.6163 mM (at 60 ◦ C). CMEA performs as a mixed-type inhibitor
Ebenso, E.E.; Kumar, A.; Verma, C.
and its adsorption at the MS/1 M HCl interface followed the Langmuir isotherm. The theoretical
Coco Monoethanolamide Surfactant
findings from density functional theory (DFT), Monte Carlo (MC), and molecular dynamics (MD)
as a Sustainable Corrosion Inhibitor
for Mild Steel: Theoretical and
simulations accorded with the experimental findings. The MC simulation’s assessment of CMEA’s
Experimental Investigations. high adsorption energy (−185 Kcal/mol) proved that the CMEA efficiently and spontaneously
Molecules 2023, 28, 1581. adsorbs at the interface.
https://doi.org/10.3390/
molecules28041581 Keywords: non-ionic surfactant; coco mono ethanol amide; corrosion inhibition; electrochemical and
theoretical techniques
Academic Editor: Changjian Lin
sector. These initiatives aim to cut the cost of corrosion by 15 to 35% which is equivalent to
USD 375 billion to USD 875 billion, respectively [8,9]. Due to growing urbanization and
industrialisation, the negative effects of corrosion, such as safety and financial penalties,
are anticipated to increase. Therefore, depending on the type of metal and electrolytes,
corrosion experts have devised numerous attempts. Every effort has advantages and disad-
vantages unique to it. Nevertheless, it has been proven that one of the best and most widely
used techniques in aqueous electrolytes is the employment of organic molecules [10–13].
Numerous organic compound families, particularly heterocycle ones, are widely used in
corrosion prevention [14–16]. But they face a problem with low solubility, particularly for
those with aromatic rings (s). Surfactants, on the other hand, are a relatively recent class
of corrosion inhibitors that have nearly all the necessary characteristics for a material to
be used as an aqueous phase corrosion inhibitor [17–20]. They have the proper ratio of
hydrophilicity (polar functional groups) to hydrophobicity (hydrocarbon chains), which is
primarily required for powerful interfacial interactions [17–20].
Surfactants have an amphiphilic structure consisting of a polar head and a nonpolar
tail. They could be bola form, Gemini-type, zwitterionic, anionic, cationic, nonionic, or
anionic, depending on the type of head. Surfactants are commonly used for lubrication,
anti-wear, degreasing, anti-scaling, and acid picking of iron and steel, as well as mineral ore
processing and acidification of oil wells. They are also extensively employed to avoid the
corrosion of metals and alloys [19,21–24]. Surfactants are a particular family of organic in-
hibitors that adsorb on metal surfaces, offering effective metal corrosion protection [25–27].
The study and synthesis of surfactants (non-ionic, anionic, cationic, and zwitterionic) as
inhibitors in diverse media have been the focus of many researchers [28–36]. In addition to
being ecologically friendly and sustainable, non-ionic surfactants also have the potential to
reduce corrosion. They have centres with a high electron density (N, S, O, P, π electrons
and aromatic ring), enabling them to adsorb on metallic surfaces and protecting them from
corrosive environments [37–40]. Some green corrosion inhibitors, according to reports, are,
however, costly or only work well when administered in high concentrations. Non-ionic
surfactants are reported to exhibit significant inhibitory efficiency for the corrosion of
metals in a variety of corrosive conditions and to have noticeably lower critical micelle con-
centrations (CMCs) than equivalent ionic surfactants [19,38,41,42]. Undoubtedly, exploring
ecological and inexpensive corrosion inhibitors for metal corrosion inhibition is significant.
Coco mono ethanol amide (CMEA) is a coconut-based surfactant, derived from the fatty
acids from coconut oil and monoethanolamine (MEA) and it is widely used in cosmetic
products and causes no harm to marine life or human safety. The structure of CMEA shows
that the nonpolar hydrophobic moiety is a long carbon chain, and the polar hydrophilic
moiety of the compound comprises readily protonated nitrogen and oxygen atoms that
provide the surfactant with a remarkable anti-corrosion potential. This research aims to
examine the mechanism of CMEA’s inhibition and its effectiveness for MS in 1 M HCl. The
findings of weight loss (WL), electrochemical impedance spectroscope (EIS), and poten-
tiodynamic polarization (PDP) to investigate the surfactant’s inhibitory performance are
presented in Sections 3.2 and 3.3, respectively, and to examine the surface characteristic of
mild steel (MS), scanning electron microscope (SEM), atomic force microscope (AFM), and
contact angle examination were carried out; the discussions are provided in Section 3.4. To
validate the experimental results and to comprehend the connection between the inhibitor
structure and MS, theoretical studies are also performed.
Figure
Figure1.1.The
Theplot of Surface
plot tension
of Surface vs. concentration.
tension vs. concentration.
2.2.WL
2.2. WLStudies
Studies
2.2.1.Impact
2.2.1. Impactof of Temperature
Temperature andand Concentration
Concentration
The
TheWL WL evaluations
evaluationswereweredone at different
done concentrations
at different at 30–60at°C30–60
concentrations ◦ C to comprehend
to comprehend
the
themode
mode of of
adsorption
adsorption(physical or chemical).
(physical The results
or chemical). Theofresults
the WLof experiments are ex‐
the WL experiments are
hibited in Table 1. It can be concluded that the values for inhibitory effectiveness
exhibited in Table 1. It can be concluded that the values for inhibitory effectiveness (WL%) (WL%)
and
andsurface
surfacecoverage
coverage (θ)(θ)
riserise
along withwith
along the concentration
the concentration of CMEA. The highest
of CMEA. The η% WL
highest η%WL
was obtained at a dose of 0.6163 mM. With the rise in temperature from 30 ℃ to 60 ◦℃, the ◦
was obtained at a dose of 0.6163 mM. With the rise in temperature from 30 C to 60 C, the
inhibition effectiveness increased and was highest at 60 ℃ and◦0.6163 mM concentration,
inhibition effectiveness increased and was highest at 60 C and 0.6163 mM concentration,
with an inhibition efficiency of 99.01%, and no increase in efficiency was seen beyond this
with an inhibition
concentration, which isefficiency
accreditedofto99.01%,
the MS’sand noareas
active increase
beingin efficiency
fully was
saturated. Theseen
in‐ beyond
this concentration, which is accredited to the MS’s active areas being
crease in the inhibition effectiveness at higher temperatures is ascribed to enhancement in fully saturated. The
increase in the inhibition effectiveness at higher temperatures is
the energy of the adsorbate molecules to cross the activation energy barrier easily and ascribed to enhancement
in thetoenergy
leads of the adsorbate
chemisorption. The corrosion rate, ν (mg.cm
molecules to cross 2 hthe activation
−1), is dependentenergy
on bothbarrier
the tem‐easily and
leads toand
perature concentration. The corrosion rate, ν (mg.cm2 h−1 ), is dependent on both the
chemisorption.
temperature and concentration.
Table 1. WL parameters for mild steel corrosion in 1 M HCl at different doses of CMEA at 30–
60℃.
Table 1. WL parameters for mild steel corrosion in 1 M HCl at different doses of CMEA at 30–60◦ C.
Temperature (°C) Conc. (mM) Weight Loss (mg) 𝝂 (mg cm−2 h−1) Surface Coverage (θ) ηWL%
Temperature (◦ C) Conc. (mM) Weight Loss (mg) Surface Coverage (θ)
Blank 55.944 0.518ν ‐ ‐ ηWL %
Blank 55.944 0.518 - -
0.2054
0.2054
23.112
23.112
0.214
0.214
0.5854
0.5854
58.54 58.54
◦C 0.3081
0.3081 16.956
16.956 0.157
0.157 0.6956
0.6956 69.56 69.56
3030 °C 0.4109 10.692 0.099 0.8075
0.4109
0.5136
10.6927.776
0.099
0.072
0.8075
0.8598
80.75 80.75
85.98
0.5136
0.6163 7.776 4.32 0.072
0.040 0.8598
0.9223 85.98 92.23
0.6163
Blank 4.32 70.95 0.040
0.657 0.9223
- 92.23 -
0.2054 28.188 0.261 0.6015 60.15
Blank
0.3081
70.95 19.44
0.657
0.180
‐
0.7253
‐ 72.53
40 ◦ C
0.2054
0.4109 28.188
11.664 0.261
0.108 0.6015
0.8345 60.15 83.45
0.5136
40 °C 0.3081
0.6163
19.447.884
3.996
0.073
0.180
0.037
0.8887
0.7253
0.9426
72.53 88.87
94.26
0.4109
Blank
11.664
80.244
0.108
0.743
0.8345
-
83.45 -
0.5136
0.2054 7.88426.028 0.073
0.241 0.8887
0.6748 88.87 67.48
0.3081 12.636 0.117 0.7617 76.17
50 ◦ C 0.4109 11.556 0.107 0.8547 85.47
0.5136 8.1 0.075 0.8987 89.87
0.6163 2.7 0.025 0.9653 96.53
Blank 88.128 0.816 - -
0.2054 26.892 0.249 0.6946 69.46
0.3081 18.684 0.173 0.7876 78.76
60 ◦ C 0.4109 11.124 0.103 0.8735 87.35
0.5136 7.236 0.067 0.9169 91.69
0.6163 0.864 0.008 0.9895 99.01
Blank 88.128 0.816 ‐ ‐
0.2054 26.892 0.249 0.6946 69.46
0.3081 18.684 0.173 0.7876 78.76
60 °C
0.4109 11.124 0.103 0.8735 87.35
0.5136 7.236 0.067 0.9169 91.69
Molecules 2023, 28, 1581 4 of 24
0.6163 0.864 0.008 0.9895 99.01
The WL experiment results were used to plot the graphs between η%WL vs. tempera‐
The corrosion
ture and WL experiment
rate vs. results were used
concentration to plot
(Figure thegraphs
2). The graphsdemonstrate
between η%howWL vs. tem-
temper‐
perature and corrosion
ature influences rateefficiency,
inhibition vs. concentration (Figure
and also the 2). The
necessity graphs
of the demonstrate
corrosion rate. how
temperature influences inhibition efficiency, and also the necessity of the corrosion rate.
Figure 2. The
Figure 2. The impact
impact of
of concentration
concentration (Left)
(Left) and
and temperature
temperature (Right)
(Right) on
on the
the corrosion
corrosion inhibition
inhibition
potential
potential of
of CMEA
CMEA for
for MS
MS in
in 11 M
M HCl.
HCl.
2.2.2. Activation Parameters Studies
2.2.2. Activation Parameters Studies
The activation parameters like activation entropy (∆S*), activation enthalpy (∆H*), and
The activation parameters like activation entropy (∆S*), activation enthalpy (∆H*),
activation energy (Ea ) were studied for the optimum concentration of 0.6163 mM at 30, 40,
and activation energy (Ea) were studied for the optimum concentration of 0.6163 mM at
50, and 60 ◦ C to further examine the inhibition properties of CMEA and their dependence
30, 40, corrosion
on the 50, and 60rate
°C (ν
to)further examineparameters
. The activation the inhibition properties
were ofby
calculated CMEA and theirand
the Arrhenius de‐
pendence on the corrosion rate (𝜈 . The
Eyring equations given below [33,34,43,44]: activation parameters were calculated by the Ar‐
rhenius and Eyring equations given below [33,34,43,44]:
log(ν) = log A − Ea /2.303RT (1)
log 𝜈 log 𝐴 𝐸 /2.303𝑅𝑇 (1)
∗
∆S ∗ ∆H ∗
RT
ν = 𝑅𝑇exp Δ𝑆 exp − Δ𝐻 ∗ (2)
𝜈 Nh exp R exp RT (2)
𝑁ℎ 𝑅 𝑅𝑇
where v, T, h, N, A, and Ea denote the corrosion rate, temperature, Planck constant, Avo-
gadro 𝑣, T, h, Arrhenius
wherenumber, N, A, and pre-exponential
Ea denote the corrosion rate,
factor, and temperature,
corrosion Planck
activation constant,
energy, Avo‐
respectively.
gadroThenumber, Arrhenius
log v vs. pre‐exponential
1/T (Equation factor, and
(5)) and log(v/T) vs.corrosion activation
1/T (Equation energy,
(6)) were respec‐
plotted to
attain Ea and ∆S* and ∆H* values. The activation parameters obtained for 0.6163 mM CMEA
tively.
are displayed in Table 2. The ∆H* value is positive, indicating endothermic adsorption,
which means the adsorption increases with a temperature increase. The negative value of
∆S* designates a decrease in the entropy attributed to the CMEA adsorption. The Arrhenius
and Eyring plots for 0.6163 mM concentration of CMEA are exhibited in Figure 3.
Cinh 1
= Cinh + (3)
θ Kads
where the equilibrium adsorption constant is denoted by Kads and whose values were
derived from the straight line’s intercept, which was obtained by plotting Cinh /θ against
Cinh . The LAI at different temperatures is displayed in Figure 4.
Molecules 2023, 28, x FOR PEER REVIEW 5 of 26
Molecules 2023, 28, 1581 The log v vs. 1/T (Equation (5)) and log(𝑣/T) vs. 1/T (Equation (6)) were plotted 5 of 24
to
attain Ea and ∆S* and ∆H* values. The activation parameters obtained for 0.6163 mM
CMEA are displayed in Table 2. The ∆H* value is positive, indicating endothermic ad‐
2. Activation
sorption,
Table parameters
which means obtained for MS
the adsorption corrosion
increases in 1aMtemperature
with HCl solution in the 0.6163
increase. mM
The CMEA.
negative
value of ∆S* designates a decrease in −the entropy attributed to the CMEA adsorption. The
Concentration (mM) Ea KJ mol 1 ∆H* KJ mol−1 ∆S* KJ.mol−1 K−1
Arrhenius and Eyring plots for 0.6163 mM concentration of CMEA are exhibited in Figure
3. 0.6163 40.83 43.85 −0.099
Table 2. Activation parameters obtained for MS corrosion in 1 M HCl solution in the 0.6163 mM
CMEA.
Table 3. Thermodynamic adsorption parameters in 1 M HCl for MS at various CMEA doses between
30 and 60 °C.
Table 3. Thermodynamic adsorption parameters in 1 M HCl for MS at various CMEA doses between
30 and 60 ◦ C.
Temp. (◦ C) Kads × 104 (M−1 ) ∆G0ads KJ·mol−1 ∆H0ads KJ·mol−1 ∆S0ads KJ·mol−1 ·K−1
30 0.504 −31.59 0.140
40 0.534 −32.78 0.139
10.96
50 0.685 −34.50 0.131
60 0.720 −35.70 0.140
The ∆Gads
0 values were derived from the K
ads using Equation (8) [47,48]:
−∆G◦ ads
1
Kads = exp (4)
55.5 RT
The physical symbols have their standard meaning [45,46]. From Table 3, it can be
seen that the ∆Gads
0 were negative, suggesting the spontaneous reaction of the inhibitors
being adsorbed on the MS, and the values are between −30 and −40 KJ mol−1 and hence,
signify mixed adsorption, i.e., both physical (around −20 kJ mol−1 ) and chemical ad-
sorption (>−40 kJ mol−1 ); and the adsorption tends to go towards chemisorption with
increasing temperature.
The heat of the adsorption (∆Hads 0 ) value obtained from the Van ’t Hoff equation
(Equation (9)) displayed in Table 3 was positive, indicating an endothermic reaction [49,50]:
−∆H ◦ ads
log Kads = + constant (5)
2.303RT
∆Hads
0 was obtained by the slope of a straight line by plotting log k
ads vs. 1/T. The
∆S0ads values were obtained by substituting the other thermodynamic values in the basic
thermodynamic equation given below [51,52]:
◦ ◦
∆Gads = ∆Hads − T∆S◦ads (6)
The adsorption entropy (∆S0ads ) values decreased with the temperature, indicating an
enhancement in the adsorption of the inhibitor. The ∆S0ads values for the CMEA adsorption
were calculated to be positive with no discernible change at various temperatures.
Table 4. PDP parameters for MS corrosion in 1 M HCl in the absence and presence of various doses
of CMEA at 30 °C.
Corrosion
Conc. jcorr (μA ‒bc Polarization
Ecorr (V) ba (V/dec) Rate ηPDP%
(mM) cm−2) (V/dec) Resistance (Ω)
(mm/yr.)
Blank −0.427 1028 0.121 0.158 28.981 11.945 ‐
0.2054 −0.462 410.24 0.133 0.150 63.608 5.6158 60.09
0.3081 −0.453 273.62 0.098 0.117 98.047 2.7611 73.38
0.4109 −0.467 142.56 0.092 0.140 169.83 1.6565 86.13
0.5136 −0.462 125.96 0.106 0.134 204.71 1.4636 87.74
. 0.6163 −0.464
OCP vs. Time plot for29.632 0.096
MS in 1 M HCl solution0.135 828.11varying doses
with and without 0.344of CMEA.
97.11
Figure 5. OCP vs. Time plot for MS in 1 M HCl solution with and without varying doses of CMEA.
Figure 6.
Figure 6. Tafel
Tafelplots
plotsfor
formild
mildsteel
steelfor
forblank and
blank varying
and doses
varying of CMEA
doses at 30
of CMEA at°C
30in
◦ C1 in
M 1HCl.
M HCl.
4. PDP
TableAs may parameters
be noticed for MSthe
from corrosion
findingsindisplayed
1 M HCl ininthe absence
Table andthe
4, with presence
rise inofthe
various
inhib‐doses
of CMEA
itor 30 j◦corr
dose,atthe C. values dropped from 1028 to 26.632 μA cm−2 for blank to 0.6163 mM
CMEA because an adsorbed layer of CMEA developed on the MS surface. The ba and bc
both Polarization Corrosion
Conc. (mM) Ecorr (V) jcorr showed
(µA cm−2 a ) small
ba change
(V/dec) in their
−bvalues,
c (V/dec) with increasing concentration demonstrating ηPDP %
Resistance (Ω) Rate (mm/yr.)
that the corrosion kinetics was not affected by the inhibitor’s presence, but the change was
Blank −0.427 1028
predominant in the b0.121 0.158 28.981 11.945
c, suggesting the inhibition phenomena to be cathodic controlled,
-
0.2054 −0.462 410.24 0.133 0.150 63.608 5.6158 60.09
0.3081 −0.453 which273.62
was further supported
0.098 by the0.117
Ecorr values being shifted towards
98.047 2.7611a more negative
73.38
0.4109 −0.467 direction.
142.56 0.092 0.140 169.83 1.6565 86.13
0.5136 −0.462 125.96
The 0.106
Ecorr values showed a shift of0.134 204.71CMEA to be
<85 mv, signifying 1.4636 87.74
a mixed‐type inhib‐
0.6163 −0.464 29.632 0.096 0.135 828.11 0.344 97.11
itor, and it controlled both the hydrogen evolution (cathodic) and metal dissolution (an‐
odic) reaction. The inhibition efficiency ηPDP% improved from 60.09 to 97.11%. The highest
inhibition
As may efficiency was from
be noticed obtained for 0.6163
the findings mM closeintoTable
displayed the CMC of the surfactant,
4, with and
rise in the inhibitor
beyond
dose, thethat,
jcorrno moredropped
values increase from
in the1028
efficiency wasµA
to 26.632 seen, − 2
cm owing to thetosurfactant’s
for blank 0.6163 mMad‐ CMEA
sorbed layer being saturated.
because an adsorbed layer of CMEA developed on the MS surface. The ba and bc both
showed a small change in their values, with increasing concentration demonstrating that
the corrosion kinetics was not affected by the inhibitor’s presence, but the change was pre-
dominant in the bc , suggesting the inhibition phenomena to be cathodic controlled, which
was further supported by the Ecorr values being shifted towards a more negative direction.
The Ecorr values showed a shift of <85 mv, signifying CMEA to be a mixed-type
inhibitor, and it controlled both the hydrogen evolution (cathodic) and metal dissolution
(anodic) reaction. The inhibition efficiency ηPDP % improved from 60.09 to 97.11%. The
highest inhibition efficiency was obtained for 0.6163 mM close to the CMC of the surfactant,
and beyond that, no more increase in the efficiency was seen, owing to the surfactant’s
adsorbed layer being saturated.
Molecules 2023, 28, x FOR PEER REVIEW
2.3.3. Electrochemical Impedance Spectroscopy (EIS) Studies 9 of 26
The EIS experiments were executed at 30 ◦ C to comprehend the kinetic and mechanistic
behaviour of the electrochemical system under consideration. The Nyquist plots were
phase angle using
constructed and frequency plots (Figure
the information 8) for
from the EISblank
findings and (Figure
inhibited 7) test
andsolutions.
the Bode It was
phase
evident
angle and from the Nyquist
frequency curves that
plots (Figure the
8) for width
blank andofinhibited
the capacitive loops expanded
test solutions. as the
It was evident
inhibitor doses rose,
from the Nyquist attributed
curves that thetowidth
an improvement
of the capacitive in theloops
surface’s
expandedabsorption
as theof the in‐
inhibitor
hibitor film.attributed
doses rose, The depression in the loops was
to an improvement in thebecause
surface’s of the non‐uniformity
absorption and coarse‐
of the inhibitor film.
ness depression
The of the metalin surface and frequency
the loops was because dispersion, and also suggested
of the non-uniformity and that the MS corro‐
coarseness of the
metalwas
sion surface and frequency
controlled by the chargedispersion, and alsoThe
transfer process. suggested
shape ofthat
the the MS corrosion
impedance curveswas
did
controlled
not change,by the charge
implying that transfer
there was process. The shape
no modification of the
to the impedance
corrosion curveswith
mechanism did not
the
change, implying
inhibitor addition. that there was
Employing the no
novamodification
2.1.4 software, to the thecorrosion mechanism
most appropriate with the
electrochem‐
inhibitor
ical addition.
equivalent Employing
circuit was used thetonova 2.1.4 software,
simulate the mostasappropriate
the EIS findings revealed inelectrochemical
Figure 7, con‐
equivalent
sisting of acircuit
serieswas used to simulate
connection betweenthe theEIS findings
solution as revealed
resistance (Rsin Figure
) and 7, consisting
these impedance of
a series connection between the solution resistance
components, together with a parallel connection between (R s ) and these impedance components,
the constant phase element
together
(CPE) and with
the apolarisation
parallel connection
resistancebetween
(Rp). The theelectrochemical
constant phasecharacteristics
element (CPE) andwere
that the
polarisation
evaluated from resistance (Rp ). The
the equivalent electrochemical
circuit are outlinedcharacteristics
in Table 5. that were evaluated from
the equivalent circuit are outlined in Table 5.
Figure8.8. Bode
Figure Bode phase
phaseangle
angle(Left)
(Left)and
andfrequency
frequency(Right)
(Right)plots
plotsofofmild
mildsteel
steelfor
foraablank
blankand
andinhibited
inhibited
solution ◦ C in 1 M HCl.
solutionhaving
havingvarying
varyingconcentrations
concentrationsof
ofCMEA
CMEAatat30
30 °C in 1 M HCl.
1
Cdl = (8)
2π f maxR p
where f max is the frequency at which the impedance’s imaginary component has the
highest value.
Table 5 demonstrates that Rs values were quite low as compared to the charge transfer
resistance (Rct ) values, which demonstrated that the corrosion mechanism was primarily
regulated by the transfer of electrons between the MS and the defensive layer and the
resistance of the specimen to oxidation during the application of an external potential.
The acquired values of Rs indicate that the solution conductivity is decreased by the
introduction of the CMEA, which relates to increased blockage at the metal surface’s active
sites. The obtained values of Rs are greater for the solution that contains an inhibitor. It
is evident that Rct enhanced as the inhibitor dosage increased, which validated that the
inhibitor’s adsorption layer had greater corrosion resistance. The Rct values increased from
Molecules 2023, 28, 1581 10 of 24
21.45 ohm.cm2 to 758.53 ohm.cm2 for blank and 0.6163 mM inhibitor concentration. In
addition, the reduction in Cdl value was seen because of the enhancement in the width of
the double layer or drop in the local dielectric constant as a result of inhibitor adsorption
on the MS surface. This is steady with the Helmholtz model, as shown by the subsequent
equation [57,58]:
εε 0
Cdl = S (9)
δ
where the protective layer’s dielectric constant and the permittivity of free space
(8.854 × 10−14 F cm−1 ) is denoted by ε and ε 0 , δ represents the protective layer’s width and
S stands for the electrode’s surface area. In order to anticipate the dissolution process, phase
shift (n) was considered as an indicator. The steady n values demonstrated that the charge
transfer mechanism controlled the dissolution without and with various concentrations
of CMEA.
The Bode plots obtained by plotting the EIS results are presented in Figure 8 (Bode
phase angle plots) and Figure 8 (Bode impedance plot). In the Bode phase angle plot, the
phase angle was plotted against frequency to recognize the impact of frequency on the
corrosion protection and at the surface/solution interface, and it was observed that there
is just one time constant associated with the creation of an electric double layer. A basic
understanding of the inhibitory activity of the inhibitors may be acquired from the phase
angle at high frequencies. The magnitude of the capacitive electrochemical behaviour
increases with increasing negative phase angle value. The ideal capacitor has a maximum
phase angle of 90◦ at the intermediate frequency. The phase angle value is lower in the
absence of CMEA than it is in its presence (Figure 8). The rise in the maximum phase angle
with increasing inhibitor doses facilitates the additional inhibitor molecules’ adsorption on
the MS surface, hence decreasing the metal dissolution rate and enhancing its protective
qualities. The Bode impedance graphs (Figure 8) demonstrate that absolute impedance
values increased at low frequencies with an increase in inhibitor doses. This specifies that
the defensive abilities of the inhibitor increase as its concentration increases.
surface with roughness and irregularities in the topography. In the presence of an inhibi‐
tor in an
Figure 9. acidic
SEM medium, the least surface roughness afterwas obtained
in 11and a much smoother
Figure 9. SEM photomicrographs (a) polished
photomicrographs (a) polished surface, after
surface, immersion
immersion in M HCl
M HCl for
for 66 h.
h. (b)
(b) Blank.
Blank.
surface
(c) CMEA.was obtained (Figure 10e,f).
(c) CMEA.
Figure10.
Figure 2-Dand
10.2‐D and3‐D
3-DAFM
AFMmicrographs
micrographsfor
forMild
Mildsteel
steel(a,b).
(a,b). Polished
Polishedsample.
sample.(c,d)
(c,d)Blank
Blanksample
sample
and(e,f)
and (e,f)sample
sample++0.6163
0.6163mM
mMCMEA
CMEAin in11M
MHCl.
HCl.
TheTheaverage roughness
average roughnessvalues andand
values rootroot
mean square
mean values
square were
values calculated
were with
calculated thethe
with
help of nano
help scope
of nano analysis;
scope 1.51.5software
analysis; softwarewas wasused.
used.The
Theinhibited
inhibited MS sample exhibited
exhibitedthe
theleast
leastaverage
averagesurface
surfaceroughness
roughnessofof20.420.4nm,
nm, which
which was
was very
very low
low as as compared
compared to the
to the blank
blank
as aasconsequence
a consequence of the
of the defensive
defensive filmfilm of CMEA
of CMEA formed
formed on the
on the MS.MS.
base and Lewis acid correspondingly. Hence, higher values of ELUMO enhance the donor
ability and lower values of ELUMO indicate higher accepter ability. The calculated EHOMO
and ELUMO values of CMEA are −7.203 and 0.414 eV, correspondingly, which indicate the
facile electron transfer from HOMO of the inhibitor to LUMO of the Fe atom and support
the chemisorption. The effectiveness of the electron transfer, chemical stability, easier
polarization, and effective polarization of the molecule was further supported by the
Molecules 2023, 28, x FOR PEER REVIEW 14 ofless
26
energy gap (∆E) as seen in Table 7.
Figure 12. HOMO, LUMO, and ESP images of CMEA with optimised structure.
Figure 12. HOMO, LUMO, and ESP images of CMEA with optimised structure.
Figure
Figure 13.
13. MC
MC and
and MD
MD attained
attained from
from the
the adsorption
adsorption configurations
configurations of
of the
the CMEA
CMEA in
in the simulated
the simulated
corrosive media on the MS surface.
corrosive media on the MS surface.
CMEA
CMEA hadhad aa rigid
rigidstructure
structurewith
witha a–C=O
–C=Oπ π bond,
bond, andand electrons
electrons forfor
thethe vacant
vacant ironiron
“d”
“d” orbital would come from heteroatoms containing lone pairs of electrons
orbital would come from heteroatoms containing lone pairs of electrons to make coordinateto make co‐
ordinate bonds. Consequently,
bonds. Consequently, the molecule
the molecule adsorbs onadsorbs
the Feon the surface
(110) Fe (110)in surface in a planner
a planner manner
mannercould
which which could maximize
maximize the interaction
the interaction between CMEAbetween CMEA
and the MS and the MSassurface,
surface, displayedas
displayed in Figure 13.
in Figure 13.
The following Equation (14) is used to determine the adsorption energy (Eads) for the
investigated inhibitors on the surface of Fe (110) [74]:
h i
Eads = Etotal − Esur f ace+water + Einhibitor (10)
where Etotal is the system’s overall energy as a consequence of the CMEA and the metal
surface interaction; Esurface + water is the energy of the Fe (110) surface in combination
Molecules 2023, 28, 1581 16 of 24
with water molecules before adsorption; and EInhibitor is the free energy of the CMEA,
respectively.
In MC simulation, a huge number of arbitrarily chosen mixtures of molecules and
ions are generated in a simulation box. The several configurations are examined until the
system reaches its energy equilibrium, which is shown by a smoothing of the mean average
energy profile. Figure 14a shows a characteristic energy profile for CMEA adsorption on Fe
(110) in a vacuum, which comprises average total energy, total energy, electrostatic, van
der Waals, and intramolecular energy. The results of MC simulations show that the CMEA
Molecules 2023, 28, x FOR PEER REVIEW interacts effectively with the Fe surface (Figure 14b), with −185.85 kcal/mol adsorption
17 of 26
energy (selected from the max value of P[E]) [75,76].
The The
following
lengthEquation (14)between
of the bond is used to determine
the Iron and the
the adsorption energywas
atoms of CMEA (Eads) for
calculated
the investigated inhibitors
using the radial on thefunction
distribution surface of Fe (110)
(RDF) [74]: of the MD trajectory. By evaluating
analysis
bond length values, the various bond types that were formed were recognized [77,78].
𝐸 𝐸 𝐸 𝐸 (1)
where Etotal is the system’s overall energy as a consequence of the CMEA and the metal
surface interaction; Esurface + water is the energy of the Fe (110) surface in combination with
water molecules before adsorption; and EInhibitor is the free energy of the CMEA, respec‐
CMEA interacts effectively with the Fe surface (Figure 14b), with −185.85 kcal/mol adsorp‐
tion energy (selected from the max value of P[E]) [75,76].
Molecules 2023, 28, 1581 The length of the bond between the Iron and the atoms of CMEA was calculated17usingof 24
the radial distribution function (RDF) analysis of the MD trajectory. By evaluating bond
length values, the various bond types that were formed were recognized [77,78]. Peaks in
the RDF graph that arise at certain distances from the metal surface provide information
Peaks in the RDF graph that arise at certain distances from the metal surface provide
about the type of adsorption activity occurring on the metal [67,79,80]. When the peak is
information about the type of adsorption activity occurring on the metal [67,79,80]. When
present between 1 and 3.5 Å, it depicts the chemisorption mechanism, and at distances
the peak is present between 1 and 3.5 Å, it depicts the chemisorption mechanism, and
higher than 3.5 Å, it signifies physisorption. The O and A atom’s RDF peak values are shown
at distances higher than 3.5 Å, it signifies physisorption. The O and A atom’s RDF peak
in Figure 15, and the inhibitors are less than 3.5 Å away from the Fe surface (Figure 15),
values are shown in Figure 15, and the inhibitors are less than 3.5 Å away from the Fe
indicating that the interaction between Fe (110) and inhibitor is mostly chemisorption
surface (Figure 15), indicating that the interaction between Fe (110) and inhibitor is mostly
[65,81].
chemisorption [65,81].
Figure 15.RDF
Figure15. RDFOOand
andNNatoms
atomsofofthe
theCMEA
CMEAininthe
thesimulated
simulatedcorrosive
corrosivesolution
solutiononto
ontoFe
Fesurface
surface
attainedby
attained byMDMDsimulation.
simulation.
2.6.
2.6.The
TheInhibition
InhibitionMechanism
Mechanism
It is well documented that organic compounds including surfactants become effective
It is well documented that organic compounds including surfactants become effec‐
by getting adsorbed on the metal surface. In the present investigation, the outcomes of
tive by getting adsorbed on the metal surface. In the present investigation, the outcomes
different analyses suggest that CMEA adsorbs on the metal surface and builds a defensive
of different analyses suggest that CMEA adsorbs on the metal surface and builds a defen‐
layer. The adsorption of CMEA on the MS surface in 1 M HCl can be described by
sive layer. The adsorption of CMEA on the MS surface in 1 M HCl can be described by
employing the physiochemisorption mode, just as with conventional organic corrosion
employing the physiochemisorption mode, just as with conventional organic corrosion
inhibitors. As seen in Figure 16b, protonation of the heteroatoms of CMEA in the forms
inhibitors. As seen in Figure 16b, protonation of the heteroatoms of CMEA in the forms of
of >C=O (carbonyl), -OH (hydroxyl), and >NH (2◦ -amine) can occur easily in an acidic
>C=O (carbonyl), ‐OH (hydroxyl), and >NH (2°‐amine) can occur easily in an acidic solu‐
solution. Therefore, the CMEA can exit in its mon-, di-, and/or tri-protonated (cationic)
tion. Therefore, the CMEA can exit in its mon‐, di‐, and/or tri‐protonated (cationic) forms
forms in 1 M HCl medium. As opposed to that, the build-up of counterions causes
in 1 M HCl medium. As opposed to that, the build‐up of counterions causes the metallic
the metallic surface to develop a negative charge (hydroxide and chloride ions) [82,83].
surface to develop a negative charge (hydroxide and chloride ions) [82,83]. Through phy‐
Through physisorption, these diametrically opposed charged moieties are attracted to one
sisorption, these diametrically
another. Therefore, opposed CMEA
interactions between charged andmoieties
the MSare attracted
surface in an to one solution
acidic another.
Therefore, interactions
may begin with physisorption.between CMEA and the MS surface in an acidic solution may begin
withHowever,
physisorption.
the outcomes of the present study suggest that chemisorption is the true
However,
mechanism of CMEA the outcomes of the
adsorption. Thispresent studyfrom
could result suggest that chemisorption
heteroatoms is the
deprotonating true
as they
mechanism of CMEA adsorption. This could result from heteroatoms deprotonating
approach a metallic surface by taking in electrons. Therefore, through a process known as as
they approach a metallic surface by taking in electrons. Therefore, through
donation or transfer, heteroatoms (N and O) move their unshared electron pairs to metallic a process
d-orbitals. Additionally, because metals (in this example, iron) are already electron-rich
species, this form of charge transfer results in an interelectronic repulsion state [82,83]. This
causes the iron to donate its additional electron through a mechanism known as retro- or
back-donation to the empty p-orbitals of C, O, and N. There is a noticeable relationship
between the extent of the donation and the extent of the retrodonation, and this relationship
known as donation or transfer, heteroatoms (N and O) move their unshared electron pairs
to metallic d‐orbitals. Additionally, because metals (in this example, iron) are already elec‐
tron‐rich species, this form of charge transfer results in an interelectronic repulsion state
Molecules 2023, 28, 1581 18 of 24
[82,83]. This causes the iron to donate its additional electron through a mechanism known
as retro‐ or back‐donation to the empty p‐orbitals of C, O, and N. There is a noticeable
relationship between the extent of the donation and the extent of the retrodonation, and
is known
this as synergism
relationship [82,83].
is known The mechanism
as synergism of corrosion
[82,83]. inhibition
The mechanism of and modesinhibition
corrosion of CMEA
adsorption on the MS surface in 1 M HCl are presented in Figure 16.
and modes of CMEA adsorption on the MS surface in 1 M HCl are presented in Figure 16.
Figure 16. (a) The mechanism of corrosion inhibition. (b) Physisorption and chemisorption
chemisorption and
and (c)
Protonation of different functional groups in 1 M HCl.
Protonation of different functional groups in 1 M HCl.
3. Experimental
3.1. Specimen, Reagents
3.1. Specimen, Reagents and
and Materials
Materials
The
The MS
MS coupons
coupons were
were acquired
acquired from
from JKJK steel
steel companies (located in
companies (located in Punjab,
Punjab, India)
India)
with
with chemical composition C‐2.83%, Si‐0.38%, P‐0.02%, V‐0.27%, Mn‐0.47%, and the rest
chemical composition C-2.83%, Si-0.38%, P-0.02%, V-0.27%, Mn-0.47%, and the rest
was Fe. Coco mono ethanol amide (Figure 17) was purchased from BLD
was Fe. Coco mono ethanol amide (Figure 17) was purchased from BLD pharma, India. pharma, India.
Coco monoethanolamide (CMEA) is a non-ionic and waxy surfactant. CMEA contains the
Coco monoethanolamide (CMEA) is a non‐ionic and waxy surfactant. CMEA contains the
minimum amount of free amine and is typically used in formulations that are pH-sensitive.
minimum amount of free amine and is typically used in formulations that are pH‐sensi‐
CMEA is well recognized for its well-known moisturizer effect on finished goods. This
tive. CMEA is well recognized for its well‐known moisturizer effect on finished goods.
product is also known to contain and make excellent use of a superb emulsifier, thickening,
This product is also known to contain and make excellent use of a superb emulsifier, thick‐
and wetting ingredient. Additionally, it has significant oil solubility and good oil emulsifier
ening, and wetting ingredient. Additionally, it has significant oil solubility and good oil
qualities. CMEA can also be used as a stabilizer and foam booster. CMEA modifies the
emulsifier qualities. CMEA can also be used as a stabilizer and foam booster. CMEA mod‐
structure of foam to produce richer, denser foam. It is excellent for shampoo, shaving
ifies the structure of foam to produce richer, denser foam. It is excellent for shampoo,
cream, and liquid soap formulations when used as a viscosity controller. The electrolyte
shaving cream, and liquid soap formulations when used as a viscosity controller. The elec‐
(HCl 37%), acetone, and ethanol were procured from Sigma Aldrich. The coupons to be
trolyte (HCl 37%), acetone, and ethanol were procured from Sigma Aldrich. The coupons
used for weight loss measurements were carved into dimensions (4.5 cm × 4 cm × 0.2 cm)
to be used for weight loss measurements were carved into dimensions (4.5 cm × 4 cm × 0.2
and were grazed with emery paper of diverse grades (180 to 2200) to attain a smooth and
cm) and
glass wereand
surface grazed withwith
cleaned emery paperwater
purified of diverse grades
and dried (180undergoing
before to 2200) to attain a smooth
the experiment.
and glass surface and cleaned with purified water and dried before 2 undergoing
For the electrochemical experiment, a 5.5 cm long MS rod having a 1 cm uncovered area the
was applied.
Molecules 2023, 28, x FOR PEER REVIEW 20 of 26
Figure
Figure 17.
17. Molecular
Molecular structure
structure of
of coco
coco mono
mono ethanol amide (CMEA).
ethanol amide (CMEA).
3.3. Weight
3.3. Weight Loss
Loss Investigations
Investigations
The WL
WL study
study was
was done
done atat four ◦ C) on MS coupons
The four different
different temperatures
temperatures (30–60
(30–60 °C) on MS coupons
having an 2
having an 18
18cmcm2area.
area.The
Thecoupons
couponswere
weregrazed
grazed with emery
with emerypapers to remove
papers any any
to remove dirt dirt
and
scales present and then splashed with distilled water, acetone, and ethanol
and scales present and then splashed with distilled water, acetone, and ethanol and and weighed
after drying.
weighed afterThe samples
drying. Theafter weighing
samples afterwere submerged
weighing in the test solution
were submerged contained
in the test solutionin
six beakers, i.e., 150 mL of 1 M HCl (Blank) and five with different inhibitor concentrations
contained in six beakers, i.e., 150 mL of 1 M HCl (Blank) and five with different inhibitor
0.2054 mM, 0.3081 mM, 0.4109 mM, 0.5136 mM, 0.6163 mM added to the test solution for
concentrations 0.2054 mM, 0.3081 mM, 0.4109 mM, 0.5136 mM, 0.6163 mM added to the
6 h at 30–60 ◦ C temperature. The temperature was regulated with the help of an oven.
test solution for 6 h at 30–60 °C temperature. The temperature was regulated with the help
After immersion in the different experimental solutions for 6 h, the MS specimens were
of an oven. After immersion in the different experimental solutions for 6 h, the MS speci‐
taken out of the beakers and then cleaned and washed with water acetone and dried with a
mens were taken out of the beakers and then cleaned and washed with water acetone and
hot air blower. The weight loss of the coupons was measured accurately thereafter. The
dried with a hot air blower. The weight loss of the coupons was measured accurately
corrosion rate (ν, mg cm2 h−1 ) and inhibition effectiveness (ηwL %) was evaluated by the
thereafter. The corrosion rate (𝜈, mg cm2 h−1) and inhibition effectiveness (𝜂 % was eval‐
Equations (1) and (2), respectively [58,84]:
uated by the Equations (1) and (2), respectively [58,84]:
W0 − W𝑊
ν= 𝑊 (11)
𝜈 At (12)
𝐴𝑡
%) = 𝜈
ν0 − ν 𝜈
× 100
𝜂ηWL (% ν0 100 (12)
(12)
𝜈
where W0 and W denote the WL earlier and after dipping in a corrosive medium, respec-
tively; νW
where and W denote the WL earlier and after dipping in a corrosive medium, re‐
0 and ν signify the corrosion rate of the blank specimen and the corrosion rate in
spectively;
the inhibitor’s and 𝜈 signify
𝜈 presence, the corrosion rate of the blank specimen and the corrosion
correspondingly.
rate in the inhibitor’s presence, correspondingly.
3.4. Electrochemical Techniques
A typical three-electrode 1 L corrosion cell of Autolab PGSTAT204 equipment, 302N
model with a FRA32M impedance analyser, was employed to perform the PDP and EIS
experiments. An inbuilt nova software 2.1.4 version was utilized to calculate the electro-
chemical parameters. The test solution having an MS rod (1 cm2 exposed area) dipped
in it as a working electrode was kept for 1 h initially to attain steady-state equilibrium
potential, i.e., open circuit potential (OCP). The unstirred test solution’s OCP was then
scrutinized as a function of time till a straight line parallel to the X-axis was attained that
validated the achievement of steady-state potential. After the OCP was attained, the PDP
experiment was run under a potential range of −0.25 to + 0.25 V at a scan rate of 0.001 V/s.
The inhibition efficacy (ηPDP %) of the PDP experiment was evaluated by the Equation (3)
as follows [85]:
j0 − jcorr
ηPDP (%) = corr 0 × 100 (13)
jcorr
where the corrosion current densities in the absence and presence of inhibitor are denoted
0
by jcorr and jcorr , correspondingly.
Molecules 2023, 28, 1581 20 of 24
4. Conclusions
The corrosion inhibition behaviour of coco monoethanolamide (CMEA), a non-ionic
surfactant based on coconut, against MS in 1 M HCl was examined utilizing both experi-
mental and computational methodologies. The CMEA displayed an excellent inhibition
efficacy of 97.17% at 0.6163 mM concentration of CMEA, which was close to the CMC
value (0.556 mM). No further increase in the inhibition efficiency was seen beyond this
concentration attributable to the saturation of the adsorbed inhibitor layer on the MS.
The weight loss results were in accordance with the EIS and PDP results. The impact of
temperature was examined by a WL experiment and an improvement in the inhibition
effectiveness was seen with the increase in temperature depicting chemisorption. At 60 ◦ C,
the maximum inhibition effectiveness of the order of 99% was obtained at 0.6163 mM. The
adsorption isotherm that fitted was the Langmuir isotherm. From the PDP outcomes, it was
evident that the CMEA was a mixed inhibitor and predominantly controlled the hydrogen
evolution (cathodic) reaction, as was clear from the shift in Ecorr values (<85 mV). The Rct
values increased from 21.45 to 758.53 ohm cm2 owing to the development of an adsorbed
layer of surfactant that controlled the charge transfer process. The creation of a defensive
layer was validated by the SEM and AFM micrographs that depicted smooth surfaces
for inhibited coupons. From the DFT studies, the energy gap value between EHOMO and
ELUMO was evaluated to be 7.617 eV, which supported the facile electron transfer from
EHOMO of the inhibitor to ELUMO of the MS. From the MD trajectory’s RDF results, the bond
length between the inhibitor heteroatoms and Fe was less than 3.5 Å, which confirmed
the inhibitor’s chemisorption behaviour. Hence, CMEA was confirmed to be a potent
corrosion inhibitor.
Molecules 2023, 28, 1581 21 of 24
Author Contributions: Writing—original draft preparation, R.G.; writing—review and editing, A.K.
and C.V.; methodology, R.G., S.S. and P.K.S.; software, O.D., A.B., E.E.E.; supervision, A.K. and C.V.
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: The authors confirm that the data supporting the findings of this study
are available within the article.
Conflicts of Interest: The authors declare that they are clear of any financial or personal conflicts that
may have seemed to affect the research described in this paper.
Sample Availability: Samples of the compounds are available from the authors.
References
1. Fontana, M.G.; Greene, N.D. Corrosion Engineering; McGraw-Hill: New York, NY, USA, 1967.
2. Sastri, V.S. Green Corrosion Inhibitors: Theory and Practice; John Wiley & Sons: Hoboken, NJ, USA, 2012.
3. Sastri, V.S. Corrosion Inhibitors: Principles and Applications; Wiley: New York, NY, USA, 1998.
4. Migahed, M.; Al-Sabagh, A. Beneficial role of surfactants as corrosion inhibitors in petroleum industry: A review article. Chem.
Eng. Commun. 2009, 196, 1054–1075. [CrossRef]
5. Prabha, S.S.; Rathish, R.J.; Dorothy, R.; Brindha, G.; Pandiarajan, M.; Al-Hashem, A.; Rajendran, S. Corrosion problems in
petroleum industry and their solution. Eur. Chem. Bull. 2014, 3, 300–307.
6. Solomon, M.M.; Uzoma, I.E.; Olugbuyiro, J.A.; Ademosun, O.T. A censorious appraisal of the oil well acidizing corrosion
inhibitors. J. Pet. Sci. Eng. 2022, 215, 110711. [CrossRef]
7. Ansari, K.R.; Chauhan, D.S.; Singh, A.; Saji, V.S.; Quraishi, M.A. Corrosion inhibitors for acidizing process in oil and gas sectors.
In Corrosion Inhibitors in the Oil and Gas Industry; Wiley: Hoboken, NJ, USA, 2020; pp. 151–176.
8. Koch, G. Cost of corrosion. In Trends in Oil and Gas Corrosion Research and Technologies; Woodhead Publishing: Sawston, UK, 2017;
pp. 3–30.
9. Song, G.-L.; Mishra, A. Selection and Characterization for Corrosion Control. Acta Metall. Sin. 2017, 4. Available online:
https://www.cnki.com.cn/Article/CJFDTotal-JSXY201704001.htm (accessed on 26 December 2022).
10. Goyal, M.; Kumar, S.; Bahadur, I.; Verma, C.; Ebenso, E.E. Organic corrosion inhibitors for industrial cleaning of ferrous and
non-ferrous metals in acidic solutions: A review. J. Mol. Liq. 2018, 256, 565–573. [CrossRef]
11. Sanyal, B. Organic compounds as corrosion inhibitors in different environments—A review. Prog. Org. Coat. 1981, 9, 165–236.
[CrossRef]
12. Xhanari, K.; Finšgar, M. Organic corrosion inhibitors for aluminum and its alloys in chloride and alkaline solutions: A review.
Arab. J. Chem. 2019, 12, 4646–4663. [CrossRef]
13. Ganjoo, R.; Verma, C.; Kumar, A.; Quraishi, M. Colloidal and interface aqueous chemistry of dyes: Past, present and future
scenarios in corrosion mitigation. Adv. Colloid Interface Sci. 2022, 311, 102832. [CrossRef]
14. Verma, C.; Abdellattif, M.H.; Alfantazi, A.; Quraishi, M. N-heterocycle compounds as aqueous phase corrosion inhibitors: A
robust, effective and economic substitute. J. Mol. Liq. 2021, 340, 117211. [CrossRef]
15. Quraishi, M.A.; Chauhan, D.S.; Saji, V.S. Heterocyclic biomolecules as green corrosion inhibitors. J. Mol. Liq. 2021, 341, 117265.
[CrossRef]
16. Verma, C.; Quraishi, M.A.; Rhee, K.Y. Corrosion inhibition relevance of semicarbazides: Electronic structure, reactivity and
coordination chemistry. Rev. Chem. Eng. 2022. [CrossRef]
17. Zhu, Y.; Free, M.L.; Woollam, R.; Durnie, W. A review of surfactants as corrosion inhibitors and associated modeling. Prog. Mater.
Sci. 2017, 90, 159–223. [CrossRef]
18. Verma, C.; Hussain, C.M.; Quraishi, M.; Alfantazi, A. Green surfactants for corrosion control: Design, performance and
applications. Adv. Colloid Interface Sci. 2022, 311, 102822. [CrossRef]
19. Verma, C.; Quraishi, M.; Rhee, K. Hydrophilicity and hydrophobicity consideration of organic surfactant compounds: Effect of
alkyl chain length on corrosion protection. Adv. Colloid Interface Sci. 2022, 306, 102723. [CrossRef]
20. Aslam, R.; Mobin, M.; Aslam, J.; Aslam, A.; Zehra, S.; Masroor, S. Application of surfactants as anticorrosive materials: A
comprehensive review. Adv. Colloid Interface Sci. 2021, 295, 102481. [CrossRef]
21. Migahed, M.; Attya, M.; Rashwan, S.; Abd El-Raouf, M.; Al-Sabagh, A. Synthesis of some novel non ionic surfactants based on
tolyltriazole and evaluation their performance as corrosion inhibitors for carbon steel. Egypt. J. Pet. 2013, 22, 149–160. [CrossRef]
22. Mobin, M.; Zehra, S.; Parveen, M. L-Cysteine as corrosion inhibitor for mild steel in 1 M HCl and synergistic effect of anionic,
cationic and non-ionic surfactants. J. Mol. Liq. 2016, 216, 598–607. [CrossRef]
23. Ganjoo, R.; Kumar, A. Current Trends in Anti-corrosion Studies of Surfactants on Metals and Alloys. J. Bio-Tribo-Corros. 2022, 8, 2.
[CrossRef]
Molecules 2023, 28, 1581 22 of 24
24. Verma, C.; Yaagoob, I.Y.; Goni, L.K.; Mazumder, M.A.J.; Ali, S.A.; Quraishi, M.A. Synthesis of Polymeric Surfactant Containing
Bis-cationic Motifs as a Highly Efficient acid Corrosion Inhibitor for C1018 Carbon steel. New J. Chem. 2023. [CrossRef]
25. Negm, N.; Al Sabagh, A.; Migahed, M.; Bary, H.A.; El Din, H. Effectiveness of some diquaternary ammonium surfactants as
corrosion inhibitors for carbon steel in 0.5 M HCl solution. Corros. Sci. 2010, 52, 2122–2132. [CrossRef]
26. Malik, M.A.; Hashim, M.A.; Nabi, F.; Al-Thabaiti, S.A.; Khan, Z. Anti-corrosion ability of surfactants: A review. Int. J. Electrochem.
Sci. 2011, 6, 1927–1948.
27. Zhang, T.; Pan, Z.; Gao, H. Novel synthesized gemini surfactant as corrosion inhibitor for carbon steel in HCl solution.
J. Surfactants Deterg. 2015, 18, 1003–1009. [CrossRef]
28. Asefi, D.; Mahmoodi, N.M.; Arami, M. Effect of nonionic co-surfactants on corrosion inhibition effect of cationic gemini surfactant.
Colloids Surf. A Physicochem. Eng. Asp. 2010, 355, 183–186. [CrossRef]
29. Kaczerewska, O.; Leiva-Garcia, R.; Akid, R.; Brycki, B.; Kowalczyk, I.; Pospieszny, T. Effectiveness of O-bridged cationic gemini
surfactants as corrosion inhibitors for stainless steel in 3 M HCl: Experimental and theoretical studies. J. Mol. Liq. 2018, 249,
1113–1124. [CrossRef]
30. Shaban, S.M.; Abd-Elaal, A.A.; Tawfik, S.M. Gravimetric and electrochemical evaluation of three nonionic dithiol surfactants as
corrosion inhibitors for mild steel in 1 M HCl solution. J. Mol. Liq. 2016, 216, 392–400. [CrossRef]
31. Bedair, M.; Soliman, S.; Metwally, M. Synthesis and characterization of some nonionic surfactants as corrosion inhibitors for steel
in 1.0 M HCl (Experimental and computational study). J. Ind. Eng. Chem. 2016, 41, 10–22. [CrossRef]
32. Aslam, R.; Mobin, M.; Zehra, S.; Obot, I.B.; Ebenso, E.E. N, N0 -Dialkylcystine Gemini and Monomeric N-Alkyl Cysteine
Surfactants as Corrosion inhibitors on mild steel corrosion in 1 M HCl solution: A Comparative study. ACS Omega 2017, 2,
5691–5707. [CrossRef]
33. Zhang, Y.; Pan, Y.; Li, P.; Zeng, X.; Guo, B.; Pan, J.; Hou, L.; Yin, X. Novel Schiff base-based cationic Gemini surfactants as corrosion
inhibitors for Q235 carbon steel and printed circuit boards. Colloids Surf. A Physicochem. Eng. Asp. 2021, 623, 126717. [CrossRef]
34. Shaban, S.M.; Elsharif, A.M.; Elged, A.H.; Eluskkary, M.; Aiad, I.; Soliman, E. Some new phospho-zwitterionic Gemini surfactants
as corrosion inhibitors for carbon steel in 1.0 M HCl solution. Environ. Technol. Innov. 2021, 24, 102051. [CrossRef]
35. Mobin, M.; Noori, S. Adsorption and corrosion inhibition behaviour of zwitterionic gemini surfactant for mild steel in 0.5 M HCl.
Tenside Surfactants Deterg. 2016, 53, 357–367. [CrossRef]
36. Javadian, S.; Yousefi, A.; Neshati, J. Synergistic effect of mixed cationic and anionic surfactants on the corrosion inhibitor behavior
of mild steel in 3.5% NaCl. Appl. Surf. Sci. 2013, 285, 674–681. [CrossRef]
37. Bedir, A.G.; Abd El-raouf, M.; Abdel-Mawgoud, S.; Negm, N.A.; El Basiony, N. Corrosion Inhibition of Carbon Steel in
Hydrochloric Acid Solution Using Ethoxylated Nonionic Surfactants Based on Schiff Base: Electrochemical and Computational
Investigations. ACS Omega 2021, 6, 4300–4312. [CrossRef]
38. Gebril, M.A.; Bedair, M.A.; Soliman, S.A.; Bakr, M.F.; Mohamed, M.B.I. Experimental and computational studies of the influence
of non-ionic surfactants with coumarin moiety as corrosion inhibitors for carbon steel in 1.0 M HCl. J. Mol. Liq. 2022, 349, 118445.
[CrossRef]
39. Li, X.; Deng, S.; Du, G. Nonionic surfactant of coconut diethanolamide as a novel corrosion inhibitor for cold rolled steel in both
HCl and H2SO4 solutions. J. Taiwan Inst. Chem. Eng. 2022, 131, 104171. [CrossRef]
40. Hegazy, M.; El-Tabei, A.; Bedair, A.; Sadeq, M. An investigation of three novel nonionic surfactants as corrosion inhibitor for
carbon steel in 0.5 M H2SO4. Corros. Sci. 2012, 54, 219–230. [CrossRef]
41. Pal, A.; Sarkar, R.; Karmakar, K.; Mondal, M.H.; Saha, B. Surfactant as an anti-corrosive agent: A review. Tenside Surfactants Deterg.
2022, 59, 363–372. [CrossRef]
42. Migahed, M.A.; Abd-El-Raouf, M.; Al-Sabagh, A.M.; Abd-El-Bary, H.M. Effectiveness of some non ionic surfactants as corrosion
inhibitors for carbon steel pipelines in oil fields. Electrochim. Acta 2005, 50, 4683–4689. [CrossRef]
43. Ahmady, A.R.; Hosseinzadeh, P.; Solouk, A.; Akbari, S.; Szulc, A.M.; Brycki, B.E. Cationic gemini surfactant properties, its
potential as a promising bioapplication candidate, and strategies for improving its biocompatibility: A review. Adv. Colloid
Interface Sci. 2022, 299, 102581. [CrossRef]
44. Jin, X.; Wang, J.; Zheng, S.; Li, J.; Ma, X.; Feng, L.; Zhu, H.; Hu, Z. The study of surface activity and anti-corrosion of novel
surfactants for carbon steel in 1 M HCl. J. Mol. Liq. 2022, 353, 118747. [CrossRef]
45. Abdallah, M.; Hegazy, M.; Ahmed, H.; Al-Gorair, A.S.; Hawsawi, H.; Morad, M.; Benhiba, F.; Warad, I.; Zarrouk, A. Appraisal of
synthetic cationic Gemini surfactants as highly efficient inhibitors for carbon steel in the acidization of oil and gas wells: An
experimental and computational approach. RSC Adv. 2022, 12, 17050–17064. [CrossRef]
46. Cheng, Y.; Wang, C.; Wang, S.; Zeng, N.; Lei, S. Comparison of anionic surfactants dodecylbenzene sulfonic acid and 1, 2,
4-triazole for inhibition of Co corrosion and study of the mechanism for passivation of the Co surface by dodecylbenzene sulfonic
acid. J. Mol. Liq. 2022, 353, 118792. [CrossRef]
47. Bhardwaj, N.; Sharma, P.; Berisha, A.; Mehmeti, V.; Dagdag, O.; Kumar, V. Monte Carlo simulation, molecular dynamic simulation,
quantum chemical calculation and anti-corrosive behaviour of Citrus limetta pulp waste extract for stainless steel (SS-410) in
acidic medium. Mater. Chem. Phys. 2022, 284, 126052. [CrossRef]
48. Dagdag, O.; Berisha, A.; Mehmeti, V.; Haldhar, R.; Berdimurodov, E.; Hamed, O.; Jodeh, S.; Lgaz, H.; Sherif, E.-S.M.; Ebenso,
E.E. Epoxy coating as effective anti-corrosive polymeric material for aluminum alloys: Formulation, electrochemical and
computational approaches. J. Mol. Liq. 2022, 346, 117886. [CrossRef]
Molecules 2023, 28, 1581 23 of 24
49. Abdellattif, M.H.; Alrefaee, S.H.; Dagdag, O.; Verma, C.; Quraishi, M. Calotropis procera extract as an environmental friendly
corrosion Inhibitor: Computational demonstrations. J. Mol. Liq. 2021, 337, 116954. [CrossRef]
50. Pakiet, M.; Tedim, J.; Kowalczyk, I.; Brycki, B. Functionalised novel gemini surfactants as corrosion inhibitors for mild steel in 50
mM NaCl: Experimental and theoretical insights. Colloids Surf. A Physicochem. Eng. Asp. 2019, 580, 123699. [CrossRef]
51. Han, P.; Chen, C.; Li, W.; Yu, H.; Xu, Y.; Ma, L.; Zheng, Y. Synergistic effect of mixing cationic and nonionic surfactants on
corrosion inhibition of mild steel in HCl: Experimental and theoretical investigations. J. Colloid Interface Sci. 2018, 516, 398–406.
[CrossRef]
52. El-Monem, M.A.; Shaban, M.M.; Migahed, M.A.; Khalil, M.M. Synthesis, characterization, and computational chemical study
of aliphatic tricationic surfactants as corrosion inhibitors for metallic equipment in oil fields. ACS Omega 2020, 5, 26626–26639.
[CrossRef] [PubMed]
53. Aslam, R.; Mobin, M.; Aslam, J.; Lgaz, H. Sugar based N, N0 -didodecyl-N, N0 digluconamideethylenediamine gemini surfactant
as corrosion inhibitor for mild steel in 3.5% NaCl solution-effect of synergistic KI additive. Sci. Rep. 2018, 8, 3690. [CrossRef]
54. Wang, D.; Li, Y.; Chen, B.; Zhang, L. Novel surfactants as green corrosion inhibitors for mild steel in 15% HCl: Experimental and
theoretical studies. Chem. Eng. J. 2020, 402, 126219. [CrossRef]
55. Lyu, B.; Liu, H.; Li, P.; Gao, D.; Ma, J. Preparation and properties of polymeric surfactants: A potential corrosion inhibitor of
carbon steel in acidic medium. J. Ind. Eng. Chem. 2019, 80, 411–424. [CrossRef]
56. El-Maksoud, A.; El-Dossoki, F.; Migahed, M.; Gouda, M.; El-Gharkawy, E.-S.R. New Imidazol-1-ium Bromide Derivative
Surfactants as Corrosion Inhibitors for Carbon Steel in 1 M HCl Solutions: Experimental and Theoretical Studies. J. Bio-Tribo-
Corros. 2021, 7, 156. [CrossRef]
57. Zhao, R.; Xu, W.; Yu, Q.; Niu, L. Synergistic effect of SAMs of S-containing amino acids and surfactant on corrosion inhibition of
316L stainless steel in 0.5 M NaCl solution. J. Mol. Liq. 2020, 318, 114322. [CrossRef]
58. Mobin, M.; Aslam, R.; Aslam, J. Synergistic effect of cationic gemini surfactants and butanol on the corrosion inhibition
performance of mild steel in acid solution. Mater. Chem. Phys. 2019, 223, 623–633. [CrossRef]
59. El-Tabei, A.; Hegazy, M.; Bedair, A.; El Basiony, N.; Sadeq, M. Novel macrocyclic cationic surfactants: Synthesis, experimental and
theoretical studies of their corrosion inhibition activity for carbon steel and their antimicrobial activities. J. Mol. Liq. 2022, 345,
116990. [CrossRef]
60. Mobin, M.; Parveen, M.; Aslam, R. Effect of different additives, temperature, and immersion time on the inhibition behavior of
L-valine for mild steel corrosion in 5% HCl solution. J. Phys. Chem. Solids 2022, 161, 110422. [CrossRef]
61. Wang, X.; Yang, J.; Chen, X.; Liu, C.; Zhao, J. Synergistic inhibition properties and microstructures of self-assembled imidazoline
and phosphate ester mixture for carbon steel corrosion in the CO2 brine solution. J. Mol. Liq. 2022, 357, 119140. [CrossRef]
62. Liu, J.; Zheng, T.; Wang, J.; Jia, G. The inhibition performance of novel amino acid-based amphiprotic surfactants on aluminum
alloys in sodium chloride solutions: Experimental and theoretical studies. Mater. Chem. Phys. 2022, 278, 125622. [CrossRef]
63. Odewunmi, N.A.; Mazumder, M.A.; Ali, S.A. Tipping Effect of Tetra-alkylammonium on the Potency of N-(6-(1H-benzo [d]
imidazol-1-yl) hexyl)-N, N-dimethyldodecan-1-aminium bromide (BIDAB) as Corrosion Inhibitor of Austenitic 304L Stainless
Steel in Oil and Gas Acidization: Experimental and DFT Approach. J. Mol. Liq. 2022, 360, 119431.
64. Palimi, M.; Tang, Y.; Alvarez, V.; Kuru, E.; Li, D. Green corrosion inhibitors for drilling operation: New derivatives of fatty acid-
based inhibitors in drilling fluids for 1018 carbon steel in CO2-saturated KCl environments. Mater. Chem. Phys. 2022, 288, 126406.
[CrossRef]
65. Dagdag, O.; Safi, Z.; Erramli, H.; Wazzan, N.; Guo, L.; Verma, C.; Ebenso, E.; Kaya, S.; El Harfi, A. Epoxy prepolymer as a novel
anti-corrosive material for carbon steel in acidic solution: Electrochemical, surface and computational studies. Mater. Today
Commun. 2020, 22, 100800. [CrossRef]
66. Dagdag, O.; Haldhar, R.; Kim, S.-C.; Guo, L.; Gouri, M.E.; Berdimurodov, E.; Hamed, O.; Jodeh, S.; Akpan, E.D.; Ebenso, E.E.
Recent progress in epoxy resins as corrosion inhibitors: Design and performance. J. Adhes. Sci. Technol. 2022, 1–22. [CrossRef]
67. Berdimurodov, E.; Kholikov, A.; Akbarov, K.; Guo, L.; Kaya, S.; Katin, K.P.; Verma, D.K.; Rbaa, M.; Dagdag, O. Novel cucurbit [6]
uril-based [3] rotaxane supramolecular ionic liquid as a green and excellent corrosion inhibitor for the chemical industry. Colloids
Surf. A Physicochem. Eng. Asp. 2022, 633, 127837. [CrossRef]
68. Prasad, D.; Dagdag, O.; Safi, Z.; Wazzan, N.; Guo, L. Cinnamoum tamala leaves extract highly efficient corrosion bio-inhibitor for
low carbon steel: Applying computational and experimental studies. J. Mol. Liq. 2022, 347, 118218. [CrossRef]
69. Berdimurodov, E.; Kholikov, A.; Akbarov, K.; Guo, L.; Kaya, S.; Katin, K.P.; Verma, D.K.; Rbaa, M.; Dagdag, O.; Haldhar, R. Novel
gossypol–indole modification as a green corrosion inhibitor for low–carbon steel in aggressive alkaline–saline solution. Colloids
Surf. A Physicochem. Eng. Asp. 2022, 637, 128207. [CrossRef]
70. Dagdag, O.; El Harfi, A.; El Gana, L.; Safi, Z.S.; Guo, L.; Berisha, A.; Verma, C.; Ebenso, E.E.; Wazzan, N.; El Gouri, M. Designing
of phosphorous based highly functional dendrimeric macromolecular resin as an effective coating material for carbon steel in
NaCl: Computational and experimental studies. J. Appl. Polym. Sci. 2021, 138, 49673. [CrossRef]
71. Erramli, H.; Dagdag, O.; Safi, Z.; Wazzan, N.; Guo, L.; Abbout, S.; Ebenso, E.; Verma, C.; Haldhar, R.; El Gouri, M. Trifunctional
epoxy resin as anticorrosive material for carbon steel in 1 M HCl: Experimental and computational studies. Surf. Interfaces
2020, 21, 100707. [CrossRef]
Molecules 2023, 28, 1581 24 of 24
72. Dagdag, O.; El Harfi, A.; Safi, Z.; Guo, L.; Kaya, S.; Verma, C.; Ebenso, E.; Wazzan, N.; Quraishi, M.; El Bachiri, A. Cyclotriphosp-
hazene based dendrimeric epoxy resin as an anti-corrosive material for copper in 3% NaCl: Experimental and computational
demonstrations. J. Mol. Liq. 2020, 308, 113020. [CrossRef]
73. Dagdag, O.; Safi, Z.; Qiang, Y.; Erramli, H.; Guo, L.; Verma, C.; Ebenso, E.E.; Kabir, A.; Wazzan, N.; El Harfi, A. Synthesis of
macromolecular aromatic epoxy resins as anticorrosive materials: Computational modeling reinforced experimental studies. ACS
Omega 2020, 5, 3151–3164. [CrossRef]
74. Damej, M.; Hsissou, R.; Berisha, A.; Azgaou, K.; Sadiku, M.; Benmessaoud, M.; Labjar, N. New epoxy resin as a corrosion inhibitor
for the protection of carbon steel C38 in 1M HCl. experimental and theoretical studies (DFT, MC, and MD). J. Mol. Struct. 2022,
1254, 132425. [CrossRef]
75. Eddy, N.O.; Ita, B.I. QSAR, DFT and quantum chemical studies on the inhibition potentials of some carbozones for the corrosion
of mild steel in HCl. J. Mol. Model. 2011, 17, 359–376. [CrossRef]
76. El Ashry, E.S.H.; El Nemr, A.; Esawy, S.A.; Ragab, S. Corrosion inhibitors-Part II: Quantum chemical studies on the corrosion
inhibitions of steel in acidic medium by some triazole, oxadiazole and thiadiazole derivatives. Electrochim. Acta 2006, 51,
3957–3968. [CrossRef]
77. Fergachi, O.; Benhiba, F.; Rbaa, M.; Ouakki, M.; Galai, M.; Touir, R.; Lakhrissi, B.; Oudda, H.; Touhami, M.E. Corrosion inhibition
of ordinary steel in 5.0 M HCl medium by benzimidazole derivatives: Electrochemical, UV–visible spectrometry, and DFT
calculations. J. Bio-Tribo-Corros. 2019, 5, 21. [CrossRef]
78. Lukovits, I.; Kalman, E.; Zucchi, F. Corrosion inhibitors—Correlation between electronic structure and efficiency. Corrosion
2001, 57, 3–8. [CrossRef]
79. Mazlan, N.; Jumbri, K.; Kassim, M.A.; Wahab, R.A.; Rahman, M.B.A. Density functional theory and molecular dynamics
simulation studies of bio-based fatty hydrazide-corrosion inhibitors on Fe (1 1 0) in acidic media. J. Mol. Liq. 2022, 347, 118321.
[CrossRef]
80. Berisha, A. Experimental, Monte Carlo and molecular dynamic study on corrosion inhibition of mild steel by pyridine derivatives
in aqueous perchloric acid. Electrochem 2020, 1, 188–199. [CrossRef]
81. Hsissou, R.; Abbout, S.; Seghiri, R.; Rehioui, M.; Berisha, A.; Erramli, H.; Assouag, M.; Elharfi, A. Evaluation of corrosion
inhibition performance of phosphorus polymer for carbon steel in [1 M] HCl: Computational studies (DFT, MC and MD
simulations). J. Mater. Res. Technol. 2020, 9, 2691–2703. [CrossRef]
82. Haldhar, R.; Prasad, D.; Mandal, N.; Benhiba, F.; Bahadur, I.; Dagdag, O. Anticorrosive properties of a green and sustainable
inhibitor from leaves extract of Cannabis sativa plant: Experimental and theoretical approach. Colloids Surf. A Physicochem. Eng.
Asp. 2021, 614, 126211. [CrossRef]
83. Berdimurodov, E.; Kholikov, A.; Akbarov, K.; Guo, L.; Kaya, S.; Verma, D.K.; Rbaa, M.; Dagdag, O. Novel glycoluril pharma-
ceutically active compound as a green corrosion inhibitor for the oil and gas industry. J. Electroanal. Chem. 2022, 907, 116055.
[CrossRef]
84. Bhardwaj, N.; Sharma, P.; Guo, L.; Dagdag, O.; Kumar, V. Molecular dynamic simulation and Quantum chemical calculation of
phytochemicals present in Beta vulgaris and electrochemical behaviour of Beta vulgaris peel extract as green corrosion inhibitor
for stainless steel (SS-410) in acidic medium. Colloids Surf. A Physicochem. Eng. Asp. 2022, 632, 127707. [CrossRef]
85. Dagdag, O.; Hsissou, R.; El Harfi, A.; El Gana, L.; Safi, Z.; Guo, L.; Verma, C.; Ebenso, E.E.; El Gouri, M. Development and
Anti-corrosion Performance of Polymeric Epoxy Resin and their Zinc Phosphate Composite on 15CDV6 Steel in 3wt% NaCl:
Experimental and Computational Studies. J. Bio-Tribo-Corros. 2020, 6, 112. [CrossRef]
86. Yadav, D.K.; Quraishi, M.A. Application of some condensed uracils as corrosion inhibitors for mild steel: Gravimetric, elec-
trochemical, surface morphological, UV–visible, and theoretical investigations. Ind. Eng. Chem. Res. 2012, 51, 14966–14979.
[CrossRef]
87. Espinoza-Vázquez, A.; Rodríguez-Gómez, F.J.; Martínez-Cruz, I.K.; Ángeles-Beltrán, D.; Negrón-Silva, G.E.; Palomar-Pardavé, M.;
Romero, L.L.; Pérez-Martínez, D.; Navarrete-López, A.M. Adsorption and corrosion inhibition behaviour of new theophylline–
triazole-based derivatives for steel in acidic medium. R. Soc. Open Sci. 2019, 6, 181738. [CrossRef] [PubMed]
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.