Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Corrosion Science 208 (2022) 110623

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Integrating experiments, DFT and characterization for comprehensive


corrosion inhibition studies – A case for cinnamaldehyde as an excellent
green inhibitor for steels in acidic media
Dharmendr Kumar *, Venkata Muralidhar K , Vinay Jain , Beena Rai
Physical Sciences Research Area, TCS Research, Tata Consultancy Services Ltd., Hadapsar, Pune 411013, India

A R T I C L E I N F O A B S T R A C T

Keywords: We have investigated the corrosion inhibition mechanisms of TCA for steel in 1 M HCl. Gravimetric, electro­
Steel chemical and AFM studies confirm TCA’s high inhibition efficiencies. Partial desorption is observed in staged
Acid inhibition dilution studies. DFT computations indicate that the geometrically flat structure, π-electrons and unsaturated
Density functional theory
bonds lead to strong TCA-Fe interactions. However, interaction strength weakens with increasing coverage
Adsorption
Interfaces
indicating possible lateral repulsions. A holistic picture of inhibition mechanism comprising a strong, but
inhomogeneous chemisorbed monolayer on steel, coupled with weakly physisorbed layers is proposed.

1. Introduction CI’s for steel, having corrosion inhibition efficiencies exceeding 90 %


[7]. It is found naturally in the bark of cinnamon trees and other species
Acid well stimulation is an extensively used technique for enhancing of the genus Cinnamomum like camphor and cassia [8]. TCA is non-toxic
productivity of oil & gas wells. Generally, 5–28 % HCl is forced under in nature and is used extensively in food and medicines [9]. Several
high pressures to dissolve the rocks in order to enlarge the existing flow papers and patents have been published on TCA elucidating its efficacy
channels and open new ones to the wellbore. Tubular materials & other as CI for steels in acidic media [10–16]. Despite several studies reported
equipment exposed to HCl can undergo severe corrosion under such on TCA as a promising CI, the molecular level understanding of TCA
conditions. A variety of organic molecules including alcohols, alde­ interaction with the steel surface is limited. This is imperative for
hydes, amines, amides, thiourea derivatives, natural extracts, quinolone designing newer, more efficient and greener alternatives as CI’s for
derivatives, triazoles, pyridines etc. play a critical role for reducing the steels.
severity of corrosion in these applications [1–4]. In this work, we report the use of a multipronged experimental and
However, most of these corrosion inhibitors (CI’s) are found to be density functional theory (DFT) computational approach, coupled with
toxic and non-biodegradable in nature [5]. Finšgar and Jackson [6] in analyses of literature reported time-of flight secondary ion mass spec­
their review paper on corrosion inhibitors for steels in acidic media troscopy (ToF-SIMS) data, to comprehensively elucidate the underlying
mention that “Most of the developed corrosion inhibitors or their formula­ mechanisms of TCA inhibition of steel corrosion at room temperature.
tions no longer meet the Oslo Paris Commission requirements because their For instance, on the experimental front, apart from the conventional
primary active ingredients may be harmful if discharged into the environ­ weight loss, polarization, EIS techniques, we have performed staged
ment.” In the wake of increasingly stringent environmental norms, dilution studies in order to understand whether the interaction is via
especially for operations in the North Sea region, there is a need for physisorption or chemisorption. Similarly, on the DFT front, instead of
screening, design and deployment of more environmentally friendly, yet merely relying on simple quantum chemical descriptors, we have per­
efficient inhibitors. formed detailed adsorption and coverage studies in order to bring out
Acetylenic alcohols, such as propargyl alcohol, although highly the likely interaction and orientation of TCA on Fe-surface, and also, the
efficient in their performance, even at high temperatures, tend to be possibility (or lack) of formation of an adherent and cohesive protective
extremely toxic. Trans-cinnamaldehyde (TCA) (Fig. 1), on the contrary, film. Besides, we are able to harmonize the experimental and compu­
is a much greener alternative and has been known to be one of the best tational results in order to establish the nature of the interaction – i.e.

* Corresponding author.
E-mail address: dharmendr.9@tcs.com (D. Kumar).

https://doi.org/10.1016/j.corsci.2022.110623
Received 12 May 2022; Received in revised form 24 August 2022; Accepted 29 August 2022
Available online 31 August 2022
0010-938X/© 2022 Elsevier Ltd. All rights reserved.
D. Kumar et al. Corrosion Science 208 (2022) 110623

2.4. Electrochemical impedance spectroscopy

EIS studies were carried out at OCP with a perturbation voltage of


10 mV between the frequencies 105 and 10− 1 Hz.

2.5. Potential for zero charge measurement


Fig. 1. Chemical structure of trans-cinnamaldehyde.
The potential for zero charge (PZC) was evaluated for each concen­
tration, from capacitance versus potential measurements, taken at an
physisorption/chemisorption/physi+chemisorption. alternating current of frequency 100 Hz, using Mott-Schottky module of
We believe that such comprehensive studies are the need of the hour VersaStudio software of the same instrument as mentioned above.
to help unriddle scientific and industrial challenges pertaining to
corrosion inhibition. 2.6. Chronocoulometry

2. Experimental studies The chronocoulometry technique was carried out for each concen­
tration of TCA in 2 steps. In the first step, the potential was maintained at
Corrosion inhibition characteristics of TCA were evaluated using + 250 mV vs OCP for 100 ms, while in the second step, the potential was
weight-loss tests, potentiodynamic polarization, electrochemical maintained at − 250 mV vs OCP for 100 ms, with no intermediate delay
impedance spectroscopy (EIS), potential for zero charge measurements between the steps. The charge flow at the interface could be captured
(PZC) and chronocoulometry on mild steel in acidic medium. In addi­ using this technique for various concentrations of TCA including that for
tion, staged dilution tests were conducted to understand the nature of the blank solution, at anodic/cathodic potentials.
adsorption of TCA. All the tests were conducted at ~27 ◦ C in ambient
atmosphere. Ametek PARSTAT 1000 AC/DC Potentiostat was employed 2.7. Staged dilution tests
for electrochemical tests on 1 cm2 area of mild steel using Ametek
K0235 flat cell with Pt mesh counter electrode, Ag/AgCl with Vycorr® Initially, the weight-loss test was carried out using 100 ml 1 M HCl
tip as reference electrode in 300 ml electrolyte. In addition, the influ­ solution containing 500 ppm TCA for 2 hrs. After 2 hrs, the 500 ppm
ence of inhibitor on the roughness of the mild steel surface was studied, solution was removed without disturbing the coupon. And then, a
using Atomic Force Microscope (AFM) from Asylum Research with 100 ml 1 M HCl solution with either of 0 and 50 ppm TCA was intro­
Olympus IX51 optical microscope and stand. duced into the same beaker.

2.8. AFM studies


2.1. Materials
Roughness measurements were carried out in tapping or intermittent
TCA (>99 %, CAS: 14371–10–9), procured from Sigma-Aldrich was
mode using silicon tip. Measurements were carried out after tuning of
used as-received without any changes. Various concentrations of in­
the tip, which resulted in the characteristic frequency of ~153.9 kHz.
hibitor were prepared from a stock solution of TCA in 99.9 % ethanol.
1 M HCl was used as blank test solution, serving as a reference for tests
3. Computational methodology
with various concentrations of inhibitor. Mild steel IS2062 with the
composition as mentioned in Table 1, was used to evaluate the inhibitor.
The TCA molecule created in Avogadro-1.2.0 [17] molecular editor
Mild steel coupons were abraded/polished on emery papers of grits 80,
software was fully relaxed by keeping it in a 30×30×30 Bohr3 box using
150, 240, 400 and 600. After polishing, the coupons were cleaned with
DFT. For calculation of frontier molecular orbital energies, the TCA
acetone, wrapped appropriately in paper and stored in a desiccator until
molecule was optimized using NWChem 6.8 [18] computational
further use in electrochemical tests.
chemistry software. B3LYP exchange-correlation functional and the
6–311 + +G** basis set was used during minimization. The TCA ge­
2.2. Weight-loss tests ometry was deemed to be optimized if the force on each atom fell below
0.025 eV/Å.
Weight-loss tests were carried out using 100 ml 1 M HCl containing Plane-Wave, spin-polarized DFT as implemented in the PWscf code
various concentrations of TCA such as 100, 200, 500 and 1000 ppm. The of Quantum Espresso-6.20 [19] was employed to investigate the
metal coupons were dipped in the test solution for 2 h for the study. adsorption mechanism of TCA on Fe (001) surface. Per­
After the weight-loss tests, the samples were carefully removed from test dew–Burke–Ernzerhoff (PBE) [20] parameterization of generalized
solutions and washed with distilled water, followed by air drying. The gradient approximation (GGA) was used for the exchange and correla­
corresponding weights of the coupons were measured before and after tion potential. A kinetic energy cut off of 35 Ry and charge density cut
the test to know the weight loss from each coupon. off of 300 Ry was used for all calculations. The van der Waals in­
teractions were accounted for using Grimme’s semi empirical correla­
tion (DFT-D3) [21] in-built in PWscf. The Fe unit cell was optimized and
2.3. Potentiodynamic polarization studies a 5 × 4 supercell containing 5-layers was prepared and fully optimized
until the force on each atom fell below 0.0025 eV/Å. Owing to the large
Potentiodynamic polarization studies were carried out by sweeping system sizes, Brillouin zone integrations were limited to the gamma
the potential at 1 mV/s scan rate between − 250 mV and + 250 mV point. More information on this section can be found in Kumar et al.
with respect to Open Circuit Potential (OCP). [22].

Table 1
Composition of IS2062 mild steel used for experimental study.
Element C Si Mn P S Cr Mo Ni Cu V Fe

wt % 0.044 0.04 1.15 0.028 0.025 0.051 0.001 0.001 0.052 0.011 Bal.

2
D. Kumar et al. Corrosion Science 208 (2022) 110623

Table 2
Results of weight-loss experiments.
Conc. (ppm) Corrosion Rate (cm/y) Inhibition Efficiency ( %)

Blank 1.24 –
100 0.37 70
200 0.29 77
500 0.13 89
1000 0.11 91

Fig. 3. Polarization curves for various concentrations of TCA.

Table 3
Results of potentiodynamic polarization experiments.
Conc. Ecorr (mV) Icorr (mA/ βc (mV/ βa (mV/ IE ( %)
(ppm) cm2) dec) dec)

Blank -473.755 1.898 193.123 152.761 –


200 -454.663 0.0192 111.627 52.506 98.99
500 -481.579 0.0109 127.397 130.9 99.43
1000 -430.11 0.0120 193.792 76.563 99.37
Fig. 2. Temkin adsorption isotherm plot of TCA on steel surface in 1 M HCl.

ΔGads = -RT ln(55⋅5 Kads)


Different initial geometries of TCA were created on Fe (001) surface
in order to identify the most likely orientation of TCA on steel surface in Where R is the universal gas constant, and T is the absolute temperature
actual conditions. The molecular adsorption was carried out only on one in Kelvin.
side of the Fe (001) surface and a vacuum of 26 Å was introduced be­ From the above expressions, we obtained a ΔGads value of
tween the slabs to avoid self-interaction. − 23.53 kJ/mol. The negative value indicates spontaneous adsorption.
Interaction energies between TCA and the Fe (001) surface were Generally, values > − 20 kJ/mol are considered to represent a relatively
calculated using the following correlation: weaker physisorption of molecules on the metal surface, while values
< − 40 kJ/mol represent stronger chemisorbing molecules arising due to
ΔEint = (Ecomplex - Eslab - N*Emol)/N
sharing or transfer of electrons between the molecule and the metal
where Ecomplex, Eslab and Emol are the total energies of the TCA-Fe (001) surface [23]. Based on the intermediate value of ΔGads so obtained, it
complex, Fe (001) slab system and the TCA molecule respectively, while appears that interaction of TCA on mild steel is via both physical and
N is the number of TCA molecules. chemisorption. It must be pointed out that the inherent assumption for
computation of ΔGads that θ = IE/100 may not be necessarily valid as
4. Results and discussions shown by Walczak et al [24]. Consequently, the accuracy of ΔGads so
obtained is doubtful. Therefore, ΔGads may not be used as a sole and/or
4.1. Experimental results reliable criterion for distinguishing between chemisorption and phys­
isorption, as has been discussed recently by Kokalj [25].
4.1.1. Weight-loss tests
Weight-loss tests were carried out to calculate the inhibition effi­ 4.1.2. Potentiodynamic polarization
ciency (IE) at various concentrations of TCA, according to the following The potentiodynamic polarization experiments were carried out to
formula: find the corrosion currents in various concentrations of the inhibitor,
[ ] which are equivalent to the corrosion rates. The resultant polarization
IE(%) = 1 −
corrosion rate with inhibitor
x 100 plots are given in Fig. 3 and were analyzed using VersaStudio to arrive at
corrosion rate without inhibitor corrosion currents with the help of Tafel regions. The respective results
From Table 2, it can be said that inhibition efficiency increases with shown in Table 3 confirm excellent inhibition efficiencies with the in­
increasing concentration of TCA. During the weight loss tests, it has been hibitor at all the concentrations tested. The corrosion inhibition effi­
observed that the rapidity of bubble evolution decreased with increasing ciency at a particular concentration of inhibitor is calculated using the
concentration. This relates to the hydrogen evolution reaction at the below expression.
cathodic sites. [
Icorr with inhibitor
]
Based on these results, a plot of θ vs ln C (where C is the inhibitor IE(%) = 1 −
Icorr without inhibitor
x 100
concentration and θ is the surface coverage or IE/100) was plotted as
shown in Fig. 2. The plot indicates that the adsorption of TCA on mild The corrosion currents derived from polarization plots show that the
steel in HCl follows the Temkin isotherm: presence of inhibitor reduces the corrosion of mild steel by at least a
hundred times in 1 M HCl. There is a slight change in the cathodic
exp (− 2αθ) = KadsC branch from 200 ppm to 500 ppm while it is more dominant in
1000 ppm. Similarly, such a change could be seen in the anodic branch
where Kads represents the equilibrium constant for adsorption.
as well, suggesting the change in inhibition mechanism. The trends of
From the obtained Kads value, the free energy of adsorption, ΔGads,
polarization curves suggest that, although 200 ppm is sufficient for good
was calculated using the following equation:
inhibition, higher concentrations offer better shielding, evident from

3
D. Kumar et al. Corrosion Science 208 (2022) 110623

Table 4
EIS results for various concentrations of the inhibitor.
Conc. (ppm) Rs (Ω.cm2) Q-Y0 (x10− 4) ( (S-sec)n.cm2) Q-n Rct (Ω.cm2) L (H.cm2) RL (Ω.cm2) IE ( %) Cdl (µF/cm2)

Blank 3.726 6.913 0.897 29.62 441


200 3.472 0.615 0.858 227.5 8 11 86.98 30.3
500 3.512 0.263 0.772 795 11 25 96.27 8.4
1000 2.771 0.314 0.688 1230 35 65 97.59 7.17

Nyquist plots as in Fig. 4(a), i.e. interfacial behavior with the blank and
with the inhibitor respectively. Charge transfer resistance, Rct, was
considered to evaluate the inhibition efficiencies with various concen­
trations of inhibitor according to the following expression.
[ ]
Rct without inhibitor
IE(%) = 1 − x 100
Rct with inhibitor
The impedance due to constant phase element, Q is evaluated using
the following equation [26]:
1
ZQ =
Yo (jω)n
The double layer capacitance (pseudo-capacitance), Cdl, has been
calculated using the following expression [27]:
√ ̅̅̅̅̅̅̅̅̅̅
n
QRct
Cdl =
Rct
The Nyquist plots show a slight change in their appearance with
increasing concentrations of the inhibitor. The size of the inductive
portion in Nyquist plots with inhibitor is reducing with increasing in­
hibitor concentrations (inset in Fig. 4(a)), suggesting that the activity of
adsorption and desorption is decreasing as the concentration is
increased. The increasing Rct values suggest an increasing inhibition,
which is evident at 1000 ppm due to formation of a thick and strong
layer. This further supports the observations from polarization curves.
Also, there is a decrease in double layer capacitance, hinting better
protection or increased double layer thickness.
The Bode plot (Fig. 4(b)) shows increasing impedance with
increasing concentration of TCA, which is a sign of increasing protec­
tion. In addition, peaks of the phase angle curves are moving to higher
frequencies with concentration. For each of the concentrations 200 and
500 ppm, a single peak in phase angle curves is seen at intermediate
frequencies suggesting an incomplete protection [28]. However, with
1000 ppm concentration, two peaks (at 63.10 Hz and 3162.28 Hz) are
seen in phase angle response – the high frequency peak suggesting re­
gions with good protection while the low frequency peak (closer to the
peak of blank at 39.81 Hz) hinting the poorly protected regions i.e.
incomplete coverage. This is in line with the behavior of TCA inferred
and described in other sections i.e. staged dilution tests and the
adsorption isotherm.

4.1.4. Potential for zero charge and surface charge


Fig. 5 shows the plots of absolute capacitance |C| versus potential,
obtained using Mott-Schottky test module of VersaStudio software by
scanning over various potentials at an alternating frequency of 100 Hz.
It can be inferred from Fig. 5(b) that the capacitance at PZC shows a
Fig. 4. (a) Nyquist and (b) Bode plots showing the impedance response of mild larger decrease from 200 ppm to 500 ppm while the decrease from
steel in various concentrations of TCA.
500 ppm to 1000 ppm is not much. From the polarization studies, it is
evident that there is considerable improvement in protection as we in­
anodic branches of 500 ppm and 1000 ppm. The polarization curves crease concentration from 200 ppm to 500 ppm, while no such
show that TCA is a mixed inhibitor for this steel. improvement could be seen from 500 ppm to 1000 ppm. This implies
that 500 ppm concentration provides reasonable monolayer coverage
4.1.3. Electrochemical impedance spectroscopy along with appropriate multilayer masking of uncovered empty spaces,
Electrochemical impedance spectroscopy was carried out to under­ and is close to saturation point. A decrease in capacitance hints an in­
stand the corrosion inhibition by the inhibitor and respective inhibition crease in the double layer thickness. This further supports the inference
efficiencies were calculated and presented in Table 4. The circuits R(QR) drawn from anodic curves of polarization plots, that 200 ppm is not
and R(Q(R(LR))) were used in ZSimpWin software for analyzing the sufficient for better shielding.

4
D. Kumar et al. Corrosion Science 208 (2022) 110623

Fig. 5. Absolute Capacitance |C| vs Potential plots from Mott-Schottky module, revealing PZC for a) blank test solution and b) 1 M HCl with TCA.

adsorption was purely physisorption or chemisorption or a combination


of both. The hypothesis was that the corrosion rates will (not) change
significantly with lowering of inhibition concentration if TCA physisorbs
(chemisorbs) on the steel surface [29], while intermediate values would
obviously imply a combination of both physisorption and
chemisorption.
Three such experiments, each comprising of two stages each of 2 h
duration were performed. Table 5 indicates that with dilution, the
corrosion rate increases, implying that there is desorption to some
extent. The weight loss for the specimen in blank 1 M HCl solution
without any inhibitor is about 1.28 cm/y. However, as the Exp. No. 3 in
Table 5 suggests, after 2 h of such a test, the corrosion rate was merely
0.39 cm/y clearly indicating a combination of physisorption and
chemisorption. Such staged dilution tests, apart from throwing light into
the mechanism of adsorption, also can provide insights into the effects of
Fig. 6. Chronocoulometry results showing charge vs time plot.
varying inhibitor dosage, which is commonly seen in actual field con­
ditions. However, deeper studies are required to know the critical con­
Table 5 centration for sufficient first layer coverage and partitioning effects, that
Results of staged dilution tests. can provide precise control over inhibitor dosage.
Exp. Time per Inhibitor Inhibitor Average
No. stage in Concentration in Concentration in Corrosion 4.1.7. Atomic force microscopy
the Set 1st stage (ppm) 2nd stage (ppm) Rate (cm/y) Fig. 7(a)-(c) show the roughness plots in both 2D and 3D obtained
(hrs) from AFM studies and Table 6 further supports the above findings and
0 2 0 0 1.28 inhibition characteristics of TCA. AFM results show that there is an in­
1 2 500 500 0.13 crease in the roughness of the surface due to the action of 1 M HCl. The
2 2 500 50 0.17 roughness of the sample is almost retained in the case of inhibitor
3 2 500 0 0.40
(1000 ppm) along with 1 M HCl proving the efficient protection by TCA.

4.1.5. Chronocoulometry 4.2. Computational results


The charge versus time plot (Fig. 6) obtained from chronocoulometry
technique, clearly shows that with increasing concentration of TCA, the The DFT optimized structural parameters for TCA (Fig. 8(a)) were
magnitude of charge at the end of each step decreases. The amount of found to match well with gas electron diffraction data of Egawa et al.
charge flow at the interface due to anodic potential seems to match the [30] (cf. Table S1, Supplementary Information). It is evident from Fig. 8
amount of reverse charge flow at the interface due to cathodic potential, (b) and (c) that both the highest occupied molecular orbital (HOMO)
for 200 ppm and 1000 ppm cases. However, with blank solution and and lowest unoccupied molecular orbital (LUMO) are evenly distributed
500 ppm TCA, the reverse charge flow values due to cathodic potential on the entire TCA molecule hinting towards a flat adsorption geometry.
are higher than the charge flow values due to anodic potential. The The computed HOMO, LUMO and HOMO-LUMO gap values using
amount of charge flow (area under the curve) at the interface arrested NWChem 6.8 (Plane-wave/PBE) are − 6.96 eV (− 5.46 eV), − 2.52 eV
increased with increase in inhibitor concentration. (− 2.96 eV) and 4.44 eV (2.50 eV), respectively. However, not much can
be concluded from these plots and values about the nature of the TCA-Fe
4.1.6. Staged dilution tests interactions per se and therefore, we resorted to explicit adsorption
The ambiguity in the values of ΔGads from the weight loss tests studies.
necessitated staged dilution studies, in order to further understand if the Two stable adsorption geometries – flat and vertical, were identified

5
D. Kumar et al. Corrosion Science 208 (2022) 110623

Fig. 7. 2D & 3D Roughness profiles for 5 sq.μm area for sample that is (a) un-corroded/polished, (b) corroded with 1 M HCl without inhibitor, and (c) corroded with
1 M HCl in the presence of 1000 ppm inhibitor. Note – scales are different for (a), (b) and (c).

6
D. Kumar et al. Corrosion Science 208 (2022) 110623

Table 6 additional Fe—C bonds. This explains the significant strengthening in


Roughness values from AFM Analysis. Eint for this configuration with inclusion of DFT-D3. On the contrary, for
Sample Average Roughness, Ra (nm) RMS Roughness, Rq (nm) the vertical configuration, there continues to be only one Fe-O bond
between the Fe (001) surface and the TCA molecule, thus explaining its
Uncorroded 29.095 34.734
Blank 88.771 117.808 weaker Eint values.
Inhibited 32.035 38.729 The reduced distances notwithstanding, it is evident that the spatial
configurations for both flat and vertical adsorption geometries remained
almost unaffected due to dispersion corrections, which is consistent with
in the explicit DFT adsorption studies and are shown in Fig. 9 along with similar findings for other molecules by Bedolla et al. [34] and Kumar
their corresponding adsorption energies. The top panel shows the opti­ et al. [22].
mized DFT structures while the bottom panel shows the dispersion- It must be mentioned that computations for adsorption of protonated
corrected DFT-D3 structures. Two things can be inferred from these TCA (likely under acidic conditions), also indicate similar optimized
results: structures implying that the adsorption mechanism essentially remains
same as that of the neutral molecule (Cf. Supplementary Information
a. The magnitudes of interaction energy values are significantly higher (section B)).
for the flat configuration. (Blue and red regions indicate charge depletion and accumulation
b. With dispersion corrections (DFT-D3), the interaction strengthens respectively). Iso-surface value: ± 0.003 eV/Å3.
drastically for the flat configuration (ΔEint = − 1.96 eV), while not so In order to further understand the nature of the surface-adsorbate
much for the vertical configuration (ΔEint = − 0.23 eV). interactions, the electron density difference and projected density of
states (PDOS) have been plotted in Fig. 10 for the DFT-D3 optimized
These differences in the Eint values between the flat and vertical geometries.
configurations can be explained by analyzing the optimized structures. The computation of the electron density difference (Δρ) was based
At the outset, it is evident that the geometrically flat structure of TCA, on the following expression:
coupled with the presence of π-electrons and unsaturated bonds (C– –C
and C– O) in its structure providing a large number of electrons to the
– Δρ(r) = ρtotal (r)– ρsurface (r) − ρinhibitor(r)
partially filled d-orbitals of Fe, facilitate the strong adsorption of TCA on
Fe surface. Besides, the strong interactions induce the sp2 to sp3 rehy­ where ρtotal, ρsurface and ρinhibitor represent electron densities of Fe-TCA
bridization of the electronic structure of the ring C-atoms as evidenced complex, the isolated Fe (001) surface and the isolated TCA molecule,
by the shifting of all the attached H-atoms above the ring plane. [31]. respectively.
Without DFT-D3, the flat adsorption geometry shows formation of The blue and red regions in the electron density difference plots
several Fe-C/O bonds. The average Fe—C (dFe—C) bond length is ~ (upper panel) show charge depletion and accumulation respectively. On
2.16 Å which is close to the sum of the covalent radii of Fe and C i.e. rcv comparing the flat and vertical adsorption configurations, it is evident
= 1.32 (rFe) + 0.76 (rC) = 2.08 Å [32]. Thus, there is possibly strong that the former shows extensive charge localization between the mole­
coordinate covalent bonding between surface Fe and C-atoms of TCA. cule and the surface, possibly, due to the strong interaction of the
The coordinate covalent bond formation is also possibly present be­ delocalized π-electrons of the aromatic ring as well as the unsaturated
tween Fe and O-atoms in both flat and vertical adsorption geometries. C–– C/C– – O bonds in TCA with the surface Fe-atoms. Whereas, in case of
The computed Fe—O bond lengths (1.98 Å, 1.92 Å) are close to the sum the latter, the charge localization is mostly confined to the interacting
of the covalent radii of Fe and O (rcv = 1.32 (rFe) + 0.66 (rO) = 1.98 Å). surface Fe and O-atoms of the TCA molecule thus indicating compara­
[32,33]. tively weaker interactions.
With dispersion corrections (DFT-D3), the TCA molecule in flat Fig. 10 (lower panels) show the PDOS of TCA/Fe (001) system after
configuration comes further close to the Fe (001) surface, thus reducing the interactions. Only bond-forming atoms were considered for the
the average Fe-C/O bond lengths besides facilitating the formation of plots. A reduced broadening of 0.1 eV was used for these PDOS calcu­
lations. In the flat adsorption geometry, almost all the C and O atoms are

Fig. 8. (a) Optimized molecular structure, (b) HOMO and (c) LUMO plots of TCA (Atom legends: red = O, yellow = C and aqua = H).

7
D. Kumar et al. Corrosion Science 208 (2022) 110623

Fig. 9. Optimized adsorption geometries of TCA on Fe (001) surface along with their adsorption energies. Upper panel: DFT and Bottom Panel: DFT-D3 results. All
indicated bond-lengths are in Å units.

making chemical bonds with Fe (001) surface while only the O-atom is surface. Whereas, the DFT optimized geometries, PDOS, and the large
making chemical bond in case of the vertical adsorption geometry. After perturbation of electronic structure in the electron density difference
adsorption the molecular peaks broaden because of hybridization of the plots, indicate that the molecule is chemisorbed. In order to arrive at a
Fe-3d orbitals with the 2p-orbitals of the interacting C and O-atoms in definitive and consistent explanation of the underlying inhibition
the molecule. In case of the flat adsorption geometry, the spread is in the mechanism, we have performed extensive coverage studies of TCA (at
entire range of Fe-3d states. However, for the vertical adsorption ge­ coverages ranging from 0.042 to 0.125 ml) on Fe (001) surface using
ometry, the spread is not continuous suggesting that a flat orientation of DFT. This is because corrosion inhibition is not only dependent on strong
TCA on the Fe-surface is more favored electronically. inhibitor adsorption, but also on the compactness of packing of the
Bader charges (Table 7) were calculated for the DFT-D3 optimized molecules on the metal surface. [37].
flat configuration using the algorithm developed by Henkelman’s group The optimized structures at each coverage are shown in Fig. 11. The
[35] by generating charge densities with the projector-augmented-wave interaction strength weakens with increasing coverage almost linearly
(PAW) potentials [36] and 800 Ry kinetic energy cutoff for charge by ~0.50 eV across the tested coverage range indicating possible lateral
density. As evident from Table 7, there is an overall charge transfer from repulsions between adjacent molecules (Fig. 12). This validates the
the metal surface to the molecule. The C16 atom of carbonyl (C– – O) Temkin adsorption isotherm model for TCA adsorption, as this model
group receives the maximum electrons i.e., − 0.765e while O atom allows for a linear decrease in the heat of adsorption with coverage [38]:
donates 0.181e electrons possibly indicating their predominant role in
the strong TCA-Fe interactions. ΔHads = ΔH◦ ads - rθ

where ΔH◦ ads is the heat of adsorption at near-zero coverages and r is the
Temkin parameter.
4.3. Harmonizing experimental and computational results
However, as per Kokalj [39], ΔEint is not a good measure for
comparing the stabilities at different coverages. The adsorption surface
From the intermediate magnitude of ΔGads (reported in the weight
free energy γads, i.e., ΔEint normalized to a unit surface area, as a func­
loss section), the Bode plots, and the staged dilution studies, it appears
tion of molecular chemical potential (Fig. 13) is a much better measure.
that TCA is interacting via both physi+chemisorption on the steel

8
D. Kumar et al. Corrosion Science 208 (2022) 110623

Fig. 10. Electron density differences (upper panels) and density of states projected to TCA molecule and Fe (lower panels) for the DFT-D3 optimized flat and vertical
adsorption configurations in Fig. 9.

This γads is given by:


Table 7
Computed Bader charges on isolated (initial) and adsorbed (final) TCA molecule γads ≈ n(ΔEint -μ)/A
for the flat orientation.
Where, n is the number of TCA molecules adsorbed on the supercell
S. No. Atom Initial Final Change having area A. As evident from Fig. 13, in the chemical potential range μ
1 C -0.087 -0.146 -0.059 ϵ [− 3.89, − 3.54], the lowest monolayer coverage (1/24) is thermo­
2 C 0.078 -0.205 -0.282 dynamically most stable among the studied coverages while stability
3 H -0.014 0.058 0.072
decreases with increase in coverage, thus, further confirming the
4 H 0.062 0.079 0.017
5 C -0.056 -0.279 -0.223 intermolecular repulsions between adjacent molecules.
6 H 0.074 0.113 0.039 Interestingly, Kovač and Finšgar [40] using sophisticated ToF-SIMS
7 C -0.055 -0.204 -0.148 imaging have shown that the distribution of TCA molecules on the
8 H 0.064 0.087 0.023 steel surface is not homogeneous (as opposed to the more efficient
9 C -0.006 -0.227 -0.221
10 H 0.062 0.072 0.010
propargyl alcohol), thus leaving some places prone to corrosion attack.
11 C -0.037 -0.077 -0.040 This further substantiates our DFT coverage studies results which indi­
12 C -0.008 -0.234 -0.226 cate the possibility of lateral repulsion by adjacent molecules. Besides,
13 H 0.053 0.086 0.033 this also possibly explains why CO acts as an intensifier for TCA inhi­
14 C -0.063 -0.217 -0.154
bition of steel corrosion in HCl at room temperatures, as reported by
15 H 0.047 0.030 -0.018
16 C 1.645 0.879 -0.765 Cabello et al. [10] as the relatively smaller CO molecules can go and
17 H 0.037 0.087 0.050 chemisorb on the vacant spaces between adjacent TCA molecules and,
18 O -1.795 -1.614 0.181 thus, offer a more compact protective layer against ingress of corrosive

9
D. Kumar et al. Corrosion Science 208 (2022) 110623

Fig. 11. Optimized structures of TCA on Fe (001) surface at various coverages (a) 0.042 (1/24), (b) 0.050 (1/20), (c) 0.062 (1/16), (d) 0.083 (1/12), (e) 0.100 (2/
20), and (f) 0.125 (2/16) ML.

Fig. 12. Effect of TCA coverage (in ML) on the magnitude of interaction en­
ergy, |Eint| (eV).

species.
Additionally, Kovač and Finšgar in the same paper [38] also inves­ Fig. 13. Adsorption surface free energies, γads, as a function of TCA chemical
tigated in detail, using ToF-SIMS, the thermal stability of TCA on steel potential, μ, for the structures shown in Fig. 11.
and showed that the TCA molecules remain stably adsorbed till ~100 ◦ C
and only at higher temperatures do these molecules start to desorb. This adsorption/coverage studies can significantly help in establishing the
clearly indicates that TCA is strongly chemisorbed on the steel surface, underlying adsorption mechanism of a given corrosion inhibitor mole­
thus, substantiating our DFT adsorption studies, as physisorbed mole­ cule and also possibly, expedite the screening/design of novel inhibitors.
cules are supposed to desorb rapidly with increase in temperature. It must be mentioned that the inhibition mechanism of TCA for steels
Thus, based on detailed experiments, DFT calculations, and SIMS in HCl at high temperatures is via a combination of TCA chemisorption
data from literature, a holistic picture (Fig. 14) of the room temperature and film-forming polymerization reactions. However, it has been shown
inhibition mechanism of TCA emerges out suggesting a strong, but by Growcock and Lopp [13], that the former is usually responsible for
inhomogeneous chemisorbed monolayer on steel, coupled with weakly the major inhibiting effect and the polymer film is probably anchored to
physisorbed layers, susceptible to desorption under dynamic conditions. the surface via adsorbed TCA and serves as a diffusion barrier to enhance
Our work thus establishes that an integrated approach combining ex­ the inhibition. Therefore, our work provides important insights into the
periments, characterization techniques (such as ToF-SIMS) and DFT principal mechanism of inhibition at both low as well as high

10
D. Kumar et al. Corrosion Science 208 (2022) 110623

Based on detailed experiments, DFT calculations, and SIMS data


from literature, a holistic picture of the room temperature inhibition
mechanism of TCA emerges out suggesting a strong, but inhomogeneous
chemisorbed monolayer on steel, coupled with weakly physisorbed
layers, susceptible to desorption under dynamic conditions.

CRediT authorship contribution statement

Dharmendr Kumar: Conceptualization, Methodology, Software,


Formal analysis, Investigation, Validation, Data curation, Writing –
original draft preparation, Software, Writing – review & editing, Su­
pervision. Venkata Muralidhar K: Conceptualization, Methodology,
Software, Formal analysis, Investigation, Validation, Data curation,
Writing – original draft preparation, Software, Writing – review &
editing, Supervision. Vinay Jain: Conceptualization, Methodology,
Software, Formal analysis, Investigation, Validation, Data curation,
Writing – original draft preparation, Software, Writing – review &
editing, Supervision. Beena Rai: Project administration, Supervision.
Fig. 14. Schematic of the proposed room temperature inhibition mechanism by
TCA on steel.
Declaration of Competing Interest
temperatures – strong chemisorption of TCA molecules on Fe (001)
surface due to formation of a plethora of Fe-C/O bonds facilitated by a The authors declare that they have no known competing financial
flat adsorption geometry. Owing to the computational challenges asso­ interests or personal relationships that could have appeared to influence
ciated with modeling polymer film-formation via DFT, we have not the work reported in this paper.
taken it into account in this work. However, work is in progress to
capture this via more computationally efficient ReaxFF and classical Data Availability
molecular dynamics techniques.
The raw/processed data required to reproduce these findings cannot
5. Conclusion be shared at this time due to technical or time limitations.

We have comprehensively investigated the mechanisms underlying Acknowledgements


the excellent corrosion inhibition of TCA for mild steel in 1 M HCl using
a combination of experimental and DFT techniques. The following are The authors thank Mr. K. Ananth Krishnan - Chief Technology Officer
some of the salient conclusions from our work: at Tata Consultancy Services (TCS), for the funding and encouragement
towards this research. The help offered by Prof. Suvarna Datar at DIAT,
• Gravimetric, electrochemical and AFM studies confirm the excellent Pune, for the AFM studies is also gratefully acknowledged.
corrosion inhibitive nature of TCA.
• The polarization curves show that TCA is a mixed inhibitor for this
Appendix A. Supplementary information
steel.
• The trends of polarization curves suggest that, although 200 ppm is
Comparison between computed and experimental bond-lengths/
sufficient for good inhibition, higher concentrations offer better
angles, effect of Protonation on TCA adsorption, Bader charge analysis
shielding, evident from anodic branches of 500 ppm and 1000 ppm.
of adsorbed protonated TCA, and Adsorption Surface Free Energies are
This was further corroborated by EIS, Mott-Schottky and chro­
presented in Supplementary Information.
nocoulometry techniques.
• Partial desorption in staged dilution studies performed by reducing
TCA concentration from 500 ppm to 0 ppm, indicate that TCA acts Appendix A. Supporting information
by a combination of physi and chemisorption.
• Two stable adsorption geometries – flat and vertical, were identified Supplementary data associated with this article can be found in the
in the explicit DFT adsorption studies on Fe (001) surface. The online version at doi:10.1016/j.corsci.2022.110623.
significantly stronger interaction energies for flat configuration
suggest this to be the preferred orientation under actual conditions. References
• The geometrically flat structure of TCA, coupled with the presence of
[1] A. Kokalj, N. Kovačević, S. Peljhan, M. Finšgar, A. Lesar, I. Milošev, Triazole,
π-electrons and unsaturated bonds (C– – C and C–– O) in its structure
benzotriazole, and naphthotriazole as copper corrosion inhibitors: I. Molecular
providing a large number of electrons to the partially filled d-orbitals electronic and adsorption properties, Chemphyschem 12 (2011) 3547–3555,
of Fe, facilitate the strong adsorption of TCA on Fe surface as https://doi.org/10.1002/cphc.201100537.
[2] P. Shetty, Corrosion inhibition behaviour of thiourea derivatives in acid media
confirmed by the electron density difference and PDOS plots. against mild steel deterioration: an overview, Surf. Eng. Appl. Electrochem. 53
• Bader charge calculations indicate there is an overall charge transfer (2017) 587–591, https://doi.org/10.3103/s1068375517060126.
from the metal surface to the molecule with the C atom of the [3] Abdolreza Farhadian, Alireza Rahimi, Nehzat Safaei, Alireza Shaabani,
Elaheh Sadeh, Majid Abdouss, Ali Alavi, Exploration of sunflower oil as a
carbonyl (C–– O) group receiving the maximum number of electrons,
renewable biomass source to develop scalable and highly effective corrosion
thus, possibly indicating its predominant role in the strong TCA-Fe inhibitors in a 15 % HCl medium at high temperatures, ACS Appl. Mater. Interfaces
interactions. 13 (2021) 3119–3138.
[4] M.M. Solomon, S.A. Umoren, I.B. Obot, A.A. Sorour, H. Gerengi, Exploration of
• DFT coverage studies of TCA (at coverages ranging from 0.042 to
dextran for application as corrosion inhibitor for steel in strong acid environment:
0.125 ML) on Fe (001) surface show that interaction strength effect of molecular weight, modification, and temperature on efficiency, ACS Appl.
weakens with increasing coverage indicating possible lateral re­ Mater. Interfaces 10 (33) (2018) 28112–28129.
[5] M. Chigondo, F. Chigondo, Recent natural corrosion inhibitors for mild steel: an
pulsions between adjacent molecules.
overview, J. Chem. 2016 (2016) 1–7, https://doi.org/10.1155/2016/6208937.

11
D. Kumar et al. Corrosion Science 208 (2022) 110623

[6] M. Finšgar, J. Jackson, Application of corrosion inhibitors for steels in acidic media [23] E. Khamis, F. Bellucci, R.M. Latanision, E.S.H. El-Ashry, Acid corrosion inhibition
for the oil and gas industry: a review, Corros. Sci. 86 (2014) 17–41, https://doi. of nickel by 2-(triphenosphoranylidene) succinic anhydride, Corrosion 47 (9)
org/10.1016/j.corsci.2014.04.044. (1991) 677–686, https://doi.org/10.5006/1.3585307.
[7] F. Zucchi, G. Trabanelli, G. Brunoro, Iron corrosion inhibition in hot 4 M HCl [24] Monika S. Walczak, Perla Morales-Gil, Robert Lindsay, Determining Gibbs energies
solution by t-cinnamaldehyde and its structure-related compounds, Corros. Sci. 36 of adsorption from corrosion inhibition efficiencies: is it a reliable approach?
(1994) 1683–1690, https://doi.org/10.1016/0010-938x(94)90123-6. Corros. Sci. 155 (2019) 182–185, https://doi.org/10.1016/j.corsci.2019.04.040.
[8] A. Williams, A. Ramsay, T. Hansen, H. Ropiak, H. Mejer, P. Nejsum, et al., [25] A. Kokalj, Corrosion inhibitors: physisorbed or chemisorbed? Corros. Sci. 196
Anthelmintic activity of trans-cinnamaldehyde and A- and B-type (2022), 109939 https://doi.org/10.1016/j.corsci.2021.109939.
proanthocyanidins derived from cinnamon (Cinnamomum verum), Sci. Rep. 5 [26] Orazem, M.E., Tribollet, B., 2008. Electrochemical impedance spectroscopy. New
(2015), https://doi.org/10.1038/srep14791. Jersey, pp.383–389.
[9] M. Friedman, N. Kozukue, L. Harden, Cinnamaldehyde content in foods determined [27] Yeum, B., 2002. EChem Software Tech Note 24-Pseudocapacitance associated with
by gas chromatography− mass spectrometry, J. Agric. Food Chem. 48 (2000) CPE. EChem software: Ann Arbor, MI, USA.
5702–5709, https://doi.org/10.1021/jf000585g. [28] R. Kashkovskiy, K. Strelnikova, A. Fedotova, Application of electrochemical
[10] G. Cabello, G. Funkhouser, J. Cassidy, C. Kiser, J. Lane, A. Cuesta, CO and trans- impedance spectroscopy to study hydrogen sulphide corrosion of steel and its
cinnamaldehyde as corrosion inhibitors of I825, L80-13Cr and N80 alloys in inhibition: a review, Corros. Eng. Sci. Technol. 54 (6) (2019) 493–515.
concentrated HCl solutions at high pressure and temperature, Electrochim. Acta 97 [29] Akshay Umesh Rajopadhye, Understanding the fundamental role of surfactant
(2013) 1–9, https://doi.org/10.1016/j.electacta.2013.03.011. structures in the inhibition of carbon steel corrosion in acidic medium (PhD
[11] Halliburton Energy Services Inc, Methods and Compositions for Inhibiting Metal Thesis), University of Florida, Gainesville, USA, 2018.
Corrosion, US20120035086A1, 2012. [30] T. Egawa, R. Matsumoto, D. Yamamoto, H. Takeuchi, Molecular structure of trans-
[12] Ecolab USA Inc, Acid corrosion inhibitor, US6117364A, 2000. cinnamaldehyde as determined by gas electron diffraction aided by DFT
[13] F. Growcock, V. Lopp, Film formation on steel in cinnamaldehyde-inhibited calculations, J. Mol. Struct. 892 (2008) 158–162, https://doi.org/10.1016/j.
hydrochloric acid, Corrosion 44 (1988) 248–254, https://doi.org/10.5006/ molstruc.2008.05.017.
1.3583933. [31] N. Kovacevic, A. Kokalj, Chemistry of the interaction between azole type corrosion
[14] F. Growcock, W. Frenier, Kinetics of steel corrosion in hydrochloric acid inhibited inhibitor molecules and metal surfaces, Mater. Chem. Phys. 137 (1) (2012)
with trans-cinnamaldehyde, J. Electrochem. Soc. 135 (1988) 817, https://doi.org/ 331–339, https://doi.org/10.1016/j.matchemphys.2012.09.030.
10.1149/1.2095783. [32] Covalent Radius for all the elements in the Periodic Table, Periodictable.Com. (n.
[15] Y. Avdeev, Y. Kuznetsov, A. Buryak, Inhibition of steel corrosion by unsaturated d.). 〈https://periodictable.com/Properties/A/CovalentRadius.v.log.html〉
aldehydes in solutions of mineral acids, Corros. Sci. 69 (2013) 50–60, https://doi. (Accessed 16 January 2020).
org/10.1016/j.corsci.2012.11.016. [33] E. Oguzie, Y. Li, S. Wang, F. Wang, Understanding corrosion inhibition
[16] S. Hossain, A. Al-Shater, S. Kareem, A. Salman, R. Ali, H. Ezuber, et al., mechanisms—experimental and theoretical approach, RSC Adv. 1 (2011) 866,
Cinnamaldehyde as a green inhibitor in mitigating AISI 1015 carbon steel https://doi.org/10.1039/c1ra00148e.
corrosion in HCl, Arab. J. Sci. Eng. 44 (2019) 5489–5499, https://doi.org/ [34] P. Bedolla, G. Feldbauer, M. Wolloch, S. Eder, N. Dörr, P. Mohn, et al., Effects of
10.1007/s13369-019-03793-y. van der Waals interactions in the adsorption of isooctane and ethanol on Fe(100)
[17] M. Hanwell, D. Curtis, D. Lonie, T. Vandermeersch, E. Zurek, G. Hutchison, surfaces, J. Phys. Chem. C. 118 (2014) 17608–17615, https://doi.org/10.1021/
Avogadro: an advanced semantic chemical editor, visualization, and analysis jp503829c.
platform, J. Chemin. 4 (2012), https://doi.org/10.1186/1758-2946-4-17. [35] W. Tang, E. Sanville, G. Henkelman, A grid-based Bader analysis algorithm without
[18] M. Valiev, E. Bylaska, N. Govind, K. Kowalski, T. Straatsma, H. Van Dam, et al., lattice bias, J. Phys.: Condens. Matter 21 (2009), 084204, https://doi.org/
NWChem: a comprehensive and scalable open-source solution for large scale 10.1088/0953-8984/21/8/084204.
molecular simulations, Comput. Phys. Commun. 181 (2010) 1477–1489, https:// [36] P.E. Blöchl, Projector augmented-wave method, Phys. Rev. B 50 (24) (1994)
doi.org/10.1016/j.cpc.2010.04.018. 17953–17979, https://doi.org/10.1103/PhysRevB.50.17953.
[19] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, et al., [37] C.D. Taylor, A. Chandra, J. Vera, N. Sridhar, Multiphysics modelling, quantum
QUANTUM ESPRESSO: a modular and open-source software project for quantum chemistry and risk analysis for corrosion inhibitor design and lifetime prediction,
simulations of materials, J. Phys.: Condens. Matter 21 (2009), 395502, https://doi. Faraday Discuss. 180 (2015) 459–477, https://doi.org/10.1039/C4FD00220B.
org/10.1088/0953-8984/21/39/395502. [38] E. McCafferty, Introduction to Corrosion Science, Springer, NY, 2009, p. 377.
[20] J. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made [39] A. Kokalj, Comment on the A.B. Rocha’s reply to second comment on the paper
simple, Phys. Rev. Lett. 77 (1996) 3865–3868, https://doi.org/10.1103/ ‘‘On the nature of inhibition performance of imidazole on iron surface’, Corros. Sci.
physrevlett.77.3865. 79 (2014) 215, https://doi.org/10.1016/j.corsci.2013.10.034.
[21] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate ab initio [40] J. Kovač, M. Finšgar, Analysis of the thermal stability of very thin surface layers of
parametrization of density functional dispersion correction (DFT-D) for the 94 corrosion inhibitors by time-of-flight secondary ion mass spectrometry, J. Am. Soc.
elements H-Pu, J. Chem. Phys. 132 (2010), 154104, https://doi.org/10.1063/ Mass Spectrom. 29 (2018) 2305–2316, https://doi.org/10.1007/s13361-018-
1.3382344. 2048-1.
[22] D. Kumar, V. Jain, B. Rai, Unravelling the mechanisms of corrosion inhibition of
iron by henna extract: a density functional theory study, Corros. Sci. 142 (2018)
102–109, https://doi.org/10.1016/j.corsci.2018.07.011.

12

You might also like