You are on page 1of 42

Accepted Manuscript

Title: Amino acids modified konjac glucomannan as green


corrosion inhibitors for mild steel in HCl solution

Authors: Kegui Zhang, Wenzhong Yang, Xiaoshuang Yin,


Yun Chen, Ying Liu, Jinxun Le, Bin Xu

PII: S0144-8617(17)31225-0
DOI: https://doi.org/10.1016/j.carbpol.2017.10.069
Reference: CARP 12918

To appear in:

Received date: 11-8-2017


Revised date: 16-9-2017
Accepted date: 21-10-2017

Please cite this article as: Zhang, Kegui., Yang, Wenzhong., Yin, Xiaoshuang., Chen,
Yun., Liu, Ying., Le, Jinxun., & Xu, Bin., Amino acids modified konjac glucomannan
as green corrosion inhibitors for mild steel in HCl solution.Carbohydrate Polymers
https://doi.org/10.1016/j.carbpol.2017.10.069

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Amino acids modified konjac glucomannan as green corrosion

inhibitors for mild steel in HCl solution


Kegui Zhang,a Wenzhong Yang,a,* yangwz@njtech.edu.cn , Xiaoshuang Yin, a, Yun
Chen, a, Ying Liu, a Jinxun Le,a Bin Xub

aCollege of Chemistry and Molecular Engineering, Nanjing Tech University, Nanjing

210009, China

bNanjing Institute of Environmental Sciences, Ministry of Environmental Protection,

Nanjing 210042, China

*Corresponding author.: Telephone: +8625 83172359

Highlights
 1. Konjac glucomannan is modified with amino acids to prepare polysaccharide esters.
 2. Synthetic polymers act as mixed-type inhibitors for steel in hydrochloric acid.
 3. The adsorption of inhibitor molecules on metal surface follows Langmuir isotherm.
 4. UV-vis spectra confirm the formation of inhibitor-Fe2+ complex in HCl solution.

ABSTRACT
Konjac glucomannan (KGM) was modified with amino acids to synthesize

polysaccharide esters (KGMA and KGMH) which were evaluated as corrosion inhibitor

for mild steel in 0.5 M HCl solution by weight loss tests, Tafel curves, electrochemical

impedance spectroscopy (EIS) and scanning electron microscopy (SEM). The synthetic

polymers were found to have the lower water absorbency and the higher water solubility

than KGM. Gravimetric measurements showed the maximum efficiencies of KGMA and

KGMH for decreasing the corrosion rate of metal at 2000 ppm are up to 89.9% and

92.4%, respectively. Polarization curves indicated polysaccharide esters could retard

both hydrogen evolution reaction and metal dissolution reaction and that the inhibitory

effect was concentration dependent. Besides, EIS studies demonstrated the corrosion

1
resistance of mild steel in hydrochloric acid was enhanced by polymer additives. The

observations on mild steel surface by SEM suggested the metallic substrate obtained

good protection against corrosion in inhibitor-containing HCl solution. The adsorptive

behavior of polymer molecules including physisorption and chemisorption obeyed

Langmuir isotherm and UV-vis spectra revealed the formation of inhibitor-ion complex.

The better performance of KGMH on corrosion inhibition was attributed to the easier

transfer of π-electron or n-electron, which was confirmed by quantum chemical

calculations.

KeyWords: Adsorption; Corrosion inhibition; Mild steel; Electrochemical technique;

konjac glucomannan.
1. Introduction
Nowadays, despite of the great challenge from advanced alloys such as aluminum

and titanium, mild steel is still an indispensable material in industry due to high cost-

effectiveness. Nevertheless, it often suffers terrible corrosion, especially in acid

solutions which have broad applications in acid picking, descaling and scale cleaning.

As an effective countermeasure, inhibitors have been widely employed to control

metallic corrosion. Up to now, many inorganic salts with oxidizability and organic

compounds containing aromatic ring and heteroatoms (N, O and S) have been employed

to protect iron in aggressive medium (Dohare, Ansari, Quraishi, & Obot, 2017; Elemike,

Nwankwo, Onwudiwe, & Hosten, 2017; Khan, Basirun, Kazi, Ahmed, Magaji, Ahmed,

Khan, & Rehman, 2017; Koh & Park, 2017). In order to retard the corrosion process

occurring on metal surface in different ways, inorganic additives react with mild steel to

generate a passivated oxide film while organic compounds cover the exposed substrate

2
to form a water-resisting layer (Cui, Shi, & Pei, 2017; Kumar, Yadav, & Singh, 2017).

However, for the sake of the sustainable development of environment, some toxic

inhibitors like chromates should be prohibited and replaced by efficient and eco-friendly

alternatives, e.g., caffeic acid, amino acids and plant extracts (Awad, Saad, Shaaban,

Jahdaly, & Hazazi, 2017; Oguzie, Oguzie, Akalezi, Udeze, Ogbulie, & Njoku, 2013;

Souza & Spinelli, 2009).

Carbohydrate substances including konjac glucomannan (KGM), cellulose and

pectin are abundant renewable resources in the nature. As chemically stable,

biodegradable and nontoxic polymers, they have been widely used in food and

pharmaceutical industry (Sinha & Kumria, 2001; Wang, Chen, Zhou, Nirasawa, Tatsumi,

Li, & Cheng, 2017). Recently, considering high affinity and good adsorption ability on

metal surface, researchers from corrosion field are paying much attention to

polysaccharides to seek potential corrosion inhibitors for metallic materials in various

media (Umoren & Eduok, 2016). In the published literature, hydroxypropyl

methylcellulose was found to have synergistic effect with KI on the acid corrosion

inhibition of mild steel (Arukalam, 2014). Polysaccharide from plantago was reported

to act as a green corrosion inhibitor for carbon steel in hydrochloric acid (Mobin & Rizvi,

2017). Broadly speaking, other common carbohydrate polymers of chitosan, starch and

pectin were also tested as effective inhibitors for iron corrosion (Brindha, Mallika, &

Sathyanarayana, 2015; Fekry & Mohamed, 2010; Umoren & Ekanem, 2010).

Surprisingly, we notice that in previous publications there are few reports about the

inhibitive performance of KGM even though it has outstanding performance of film-

3
forming. Konjac glucomannan is comprised of β-1, 4 linked D-mannose and D-glucose

residues with probable ratio of 1.6:1 (Duan, Lu, Chen, Duan, Wang, Gao, & Pan, 2010).

A small number of acetyl groups which contribute to water solubility are attached

randomly to C-6 position of the saccharide units along the molecular backbone

approximately 1 per 19 sugar residues. However, vast hydrogen bonds in or between

molecules may strongly increase the water absorbency of KGM and thus result in the

formation of insoluble agglomeration in solution, which may be the disadvantage for

KGM to perform as an effective inhibitor for metal (Chen, Li, & Li, 2011). We deduce

that KGM would better dissolve in acid solution and adhere to steel surface if the amount

of hydroxyl group was reduced by esterification reaction. The criterion for selecting

applicable reagents to esterify KGM is comprehensive. First, the selected materials

should be eco-friendly because the synthetic esters may hydrolyze gradually in the

environment. Secondly, their molecular structures should contain specific groups to

improve the properties of water solubility and corrosion inhibition when the functional

groups of OH- in KGM were partly removed. As mentioned above, some amino acids

have been examined as green and low-cost corrosion inhibitors for various metals and –

NH2 groups play a key role in solubility and corrosion control (Barouni et al., 2008;

Gece & Bilgiç, 2010; Yadav, Sarkar, & Purkait, 2015). As a result, amino acid

compounds can be good candidates for the chemical modification of polysaccharide.

The further selection of two amino acids is aimed at studying the accurate structure-

activity relationship of inhibitors, i.e., they should have similar molecular structures but

different properties like polarity, aromaticity and heteroatom amount. After a careful

4
evaluation on 20 known amino acids, we decide to modify KGM with L-arginine and L-

histidine since their structures can be distinguished by different N-containing groups. In

this way, the hydroxyl groups in polysaccharide molecules will be partly replaced and

the electron-rich groups like imidazole ring and amidogen will be grafted to molecular

backbone. In other words, new konjac glucomannan esters with enhanced water

solubility and more active sites to favor the adsorption on metal are designed and

synthesized for the first time.

In our study, water-soluble konjac glucomannan L-arginine ester (KGMA) and

konjac glucomannan L-histidine ester (KGMH) were synthesized in lab and then tested

as corrosion inhibitors for mild steel in 0.5 M HCl solution by using a few experimental

techniques of weight loss, potentiodynamic polarization curves, electrochemical

impedance spectroscopy, scanning electron microscopy and UV-visible spectra.

Moreover, quantum chemical calculations were performed to investigate the structure-

property relationship of new polysaccharide esters regarding corrosion inhibition.


2. Experimental methods
2.1 Materials

Commercial konjac powder (purity 98%) was purchased from Yizhi Konjac

Factory, Hubei Province, China. L-histidine and L-arginine were purchased from

Shanghai Aladdin biological technology Co., Ltd. Other chemicals used in our

experiments were all of analytical grade and the ultrapure water was prepared by a Milli-

Q instrument.

2.2 Preparation and characterization of polysaccharide esters

2.2.1 General synthetic procedure

5
Amino acid modified carbohydrate polymers, KGMA and KGMH were

synthesized according to the esterification reaction shown in Fig. 1. First of all, a mixture

of 0.1 g p-toluenesulfonic acid (TsOH, catalyst), 1.0 g KGM and 2.0 g amino acid were

added to a 250 mL three-necked flask which contained 100 mL dimethyl sulfoxide

(DMSO). Subsequently, the solution was heated to 393 K with magnetic stirring and

refluxing. About 24 h later, the residue obtained by immediate filtration was washed by

ethanol (50 mL × 3) to produce target products which were used as corrosion inhibitors

in the following experiments.

2.2.2 Molecular weight measurement

According to the literature, the molecular weight distributions of KGM, KGMH

and KGMA were measured by a gel permeation chromatography instrument (PL-

GPC50) with a MLN-online analysis software (Chen, Zhao, Liu, Li, Liu, Wu, Shi,

Norde, & Li, 2016). The injected volume of the sample dissolved in 0.1 mol/L sodium

nitrate was 100 μL and the column temperature was maintained at 298 K with a flow

rate of 0.5 mL/min.

2.2.3 Degree of substitution

The degree of substitution (DS) of aforesaid esterification reaction was calculated

via using the following equation (Klemm, Philipp, Heinze, Heinze, & Wagenknecht,

1998):

n(KOH) M r (RU)
DS  (1)
ms  M r (RCO)

where n(KOH) is the total alkali in mol consumed in saponification reaction, Mr(RU) is

the molar mass in g/mol of the repeated unit of polysaccharide, ms is the weight of sample

6
in g and Mr(RCO) is the molar mass in g/mol of substituent group.

2.2.4 Water absorption and solubility

The water absorbency and solubility of KGM and its esters were measured based

on previous publications (Xiao, Dai, Wang, Ni, Yan, Fang, Corke, & Jiang, 2015). In

brief, the dry sample (m1) was fully dispersed in 50 mL ultrapure water at 298 K by 24

h of magnetic stirring. After that, the suspension was centrifuged at 6000 rpm for 30 min

and the precipitate was then weighed (m2). Finally, the supernatant was dried at 393K in

a vacuum oven and the residual weight was measured as m3. For those polymers, the

water absorption (Wa) per g dry sample and the solubility (S) in 100 mL water can be

calculated as following equations:

m2   m1  m3 
Wa  (2)
m1  m3

S  2m3 (3)

2.2.5 FT-IR analysis

The chemical structures of KGM and its esters were identified by Fourier transform

infrared spectroscopy. 100 mg KBr powder and 10 mg sample were well mixed and then

pressed to obtain a transparent slice. The infrared spectroscopy from 400 – 4000 cm-1

was recorded by a Thermo Nicoletis-5 infrared spectrophotometer.

2.3 Solutions and samples

To prepare the corrosive solution of 0.5 M hydrochloric acid, an analytical grade

HCl (37 wt.%) was diluted by ultrapure water. The concentrations of added

polysaccharide esters in test solutions are 100, 200, 500, 1000 and 2000 ppm. The

specimens of mild steel studied in the work have the following elemental composition:

7
wt. %, C: 0.17; Mn: 0.37; Si: 0.20; S: 0.03; P: 0.01 and balance Fe. For gravimetric

measurements, iron material was mechanically processed to produce a cuboid sample of

5.00 cm × 2.50 cm × 0.20 cm. In electrochemical section, a cylinder of mild steel was

embedded into epoxy resin to prepare the working electrode (WE) with a bare surface

of 0.5 cm2 on the bottom. Prior to use, the electrode surface was abraded with SIC paper

(500#, 1000#, 2000#), rinsed with ethanol and dried by cold air.

2.4 Weight loss measurements

Weight loss tests can show the intuitive inhibitory effect of additives and provide

reliable data of the corrosion process with or without inhibitors. Before testing, iron

samples were washed with ethanol and water, dried in cold air and weighed by an

analytical balance. Mild steel cells were hung in glass beakers which contained 300 mL

of 0.5 M HCl solution in the absence and presence of different concentrations of

inhibitors. The system temperature was maintained at 298±1 K by using a thermostat

water bath. After 8 h of immersion, the tested samples were washed with acetone and

ultrapure water, dried in cold air and weighed. The corrosion rate of mild steel (Δv) was

calculated using the following formula:


W
v  (4)
St
where W is weight loss in g, S is superficial area of sample in m2 and t is immersion time

in h. The inhibitive efficiency (η) and surface coverage (θ) of studied polymers are

evaluated according to equations (Zhang, Yang, Xu, Yin, Liu, & Chen, 2015):

v 0  v
η 100% (5)
v 0

8

 (6)
100
where Δv0 and Δv are the corrosion rates of mild steel samples in 0.5 M HCl solution

without and with inhibitors, respectively.

2.5 Electrochemical tests

Electrochemical experiments to investigate the corrosion behavior of mild steel in

0.5 M HCl in the absence and presence of inhibitors were performed using a Gamry

Reference 4000 electrochemical workstation with ESA410 analysis software. All the

electrochemical tests were executed at 298 K with a common three-electrode system

which was composed of an iron electrode (working electrode), a platinum electrode

(counter electrode) and a saturated calomel electrode (SCE) connected to a Luggin

capillary (reference electrode). Before testing, the iron electrode was immersed in

unaerated electrolyte for 1 h to obtain a steady open circuit potential (Eocp). For

polarization curve measurements, a potential range of ± 300 mV (vs. SCE) versus open

circuit potential was dynamically scanned on the surface of mild steel with a rate of 1

mV/s. Besides, electrochemical impedance spectroscopy was recorded from 100 kHz to

100 mHz with a perturbation of 10 mV alternating voltage.

2.6 Scanning electron microscopy

Under 8 kV extra high tension and high vacuum, A ZEISS MERLIN SEM

instrument was used to observe the surface morphologies of the steel samples before and

after 24 h of immersion in 0.5 M hydrochloric acid in the absence and presence of

inhibitors at 298 K.

2.7 UV-visible spectrophotometric measurements

9
UV-visible spectrum was recorded to investigate the chemical interaction between

polysaccharide esters and mild steel substrate in 0.5 M HCl solution. The spectra from

200 nm – 700 nm were measured for inhibitor-containing solution before and after the

immersion of mild steel at 298 K for 7 days using an Agilent 8693 Spectrophotometer.

Under each condition, all the experimental measurements were performed in triplicate

to guarantee the reproducibility of results. .

2.8 Quantum chemical calculations

As a powerful theoretical tool, quantum chemistry was used to study the internal

relationship between the structures of KGM esters and their effect on corrosion

inhibition for mild steel in hydrochloric acid. In gas phase, the stereo structures of the

monomeric units of KGMA and KGMH were optimized by running a Gaussian 09W

software which was supported by density functional theory (DFT) and B3LYP function

as well as 6–31g basis set. Moreover, frequency calculation was also introduced to obtain

the steady configuration of inhibitor molecules with the lowest energy.

3. Results and discussion

3.1 Characterization of carbohydrate polymers

3.1.1 Physical properties

The physical parameters of KGM, KGMA and KGMH are summarized in Table S1

(Supplementary Information), including appearance, weight-average molecular weight

(Mw), number-average molecular weight (Mn), polydispersity coefficient (PC), degree of

substitution (DS), water adsorption (Wa) and solubility (S). As we can see, the average

molecular weights of KGM and its derivatives are all up to 105 in magnitude.

10
Polydispersity coefficients for KGM, KGMA and KGMH are 1.13, 1.10 and 1.10,

successively. That indicates the molecular weight distributions of those polymers are

relatively concentrated. The DS values show the hydroxyl groups in KGM molecules are

partly replaced by amino acids and that L-histidine has a slightly better ability to esterify

KGM than L-arginine. The water absorbency is 32.5 g for KGM, 12.4 g for KGMA and

15.8 g for KGMH, attributed to the decrease in hydroxyl amount which then results in

the reduction of both intramolecular and intermolecular hydrogen bonds. As suggested

by the data in Table S1, the lower water absorbency and the polar amino group in

molecules are deemed to enhance the water solubility of KGMA and KGMH.

3.1.2 Chemical structure

FT-IR spectra to characterize the functional groups of polysaccharide polymers are

presented in Fig. 2. We can observe that the strong peak at 3430 cm-1 in the spectrum of

KGM indicates the stretching vibration of the hydroxyl in molecule. Moreover, two

small peaks located at 2976 cm-1 and 2895 cm-1 stand for the stretching vibration of C–

H bond. In the spectra of KGMA and KGMH, the characteristic peaks of –OH obviously

become weak, ascribed to the consumption of hydroxyl groups by esterification reaction.

In addition, the absorption peaks at 1735 cm-1 for KGMA and 1740 cm-1 for KGMH

suggest the typical stretching vibration of –COO– groups in the structures of

carbohydrate esters. The sharp peak at 1627 cm-1 for KGMA is an indication of C=N

bond from L-arginine, and the adsorption band of quadruple peaks from 2887 cm-1 to

3134 cm-1 for KGMH is the “fingerprint” of the imidazole ring in L-histidine (Hegazy,

Hasan, Emara, Bakr, & Youssef, 2012; Yoshida, & Ishida, 1984). Consequently, the FT-

11
IR spectra confirm the occurrence of the esterification reaction between KGM and

selected amino acids.

3.2 Gravimetric measurements

3.2.1 Effect of inhibitor concentration

The inhibitory effects of KGMA and KGMH on the corrosion of mild steel in 0.5

M HCl solution at 298 K were evaluated by weight loss method. Corrosion data for

metallic samples obtained after 8 h of immersion in test electrolyte are listed in Table 1.

It is evident that the corrosion rate of mild steel in blank solution (15.7 g/hm2) is greatly

reduced by polysaccharide esters. For example, the dissolution rates of iron with the

addition of 2000 ppm inhibitors are 1.59 g/hm2 for KGMA and 1.20 g/hm2 for KGMH.

As shown in Table 1, the inhibition efficiencies of polymers are concentration dependent

and KGMH always shows the better performance than KGMA at same concentration.

Such difference in inhibiting effect may be attributed to a larger coverage of the former

on bare surface of mild steel in aggressive solution.

12
3.2.2 Adsorption isotherm

It is well recognized that most organic molecules retard metal corrosion in acid

solution by displacing the pre-adsorbed water molecules on substrate surface and

forming a water-blocking barrier. To study the adsorptive behavior of inhibitors on metal

surface, some adsorption isotherms including Temkin, Frumkin, Flory–Huggins and

Langmuir have been examined to explore the potential connection between inhibitor

concentration (C) and the surface coverage (θ) shown in Table 1. It has been found that

the adsorption of both KGMA and KGMH molecules follows Langmuir isotherm

(Zhang, Xu, Yang, Yin, Liu & Chen, 2015):

C 1
 C (7)
 K ads
where Kads is the equilibrium constant of adsorption. The fitting lines in Fig. 3 show the

plots of C/θ have a linear relation with C. Meanwhile the slopes and linear correlation

coefficients (R2) of curves are all about 1. Those results indicate the formation of

inhibitor monolayer at solid/liquid interface (Ramachandran, Tsai, Blanco, Chen, Tang,

& Goddard, 1996). The values of Kads are calculated from the intercept of the lines and

presented in Table 1. Usually, a higher Kads suggests a stronger interaction between

adsorbate and adsorbent. As we can see, the obtained values of Kads for KGMA and

KGMH are 0.041 and 0.141, respectively. This confirms the better inhibition

performance of the latter on the acid corrosion of mild steel.

The standard free energy of adsorption of the system including inhibitors and mild steel

(ΔGads) is obtained by using the following formula (Umoren, Obot, Madhankumar, &

Gasem, 2015):

Gads   RT ln 1106 Kads  (8)


where R is the ideal gas constant in J/mol·K, T is the absolute temperature in K and 1 ×

106 is the concentration of water in ppm. As shown in Table 1, the calculated ΔGads

13
values for KGMA and KGMH are –26.36 kJ/mol and –29.38 kJ/mol, respectively,

suggesting the spontaneous and strong adsorption of inhibitor molecules towards metal

surface. According to literature, those ΔGads values in the range from –20 kJ/mol to –40

kJ/mol indicate a mixed-type adsorption including a physical electrostatic interaction

and a chemical coordination bond (Dandia, Gupta, Singh, & Quraishi, 2013). Therefore,

the higher ΔGads value suggests it is more favorable for KGMH to form an adsorptive

barrier on mild steel surface in hydrochloric acid.

3.3 Electrochemical tests

3.3.1 Potentiodynamic polarization curves

Tafel lines for mild steel electrode immersed in 0.5 M HCl solution with and

without different concentrations of inhibitors are presented in Fig. 4. It is shown that the

addition of polymers results in a noticeable shift of polarization curves towards a lower

corrosion current density, i.e., the electrochemical corrosion of metal in acid solution is

controlled by KGMA and KGMH (Dutta, Saha, Adhikari, Banerjee, & Sukul, 2017). In

the presence of inhibitors, the shapes of polarization curves have no distinct changes

compared with the blank one, indicating that polysaccharide esters block the corrosion

process of iron by slowing down the kinetics of electrode reaction rather than altering

the reaction mechanism. We can observe the corrosion potentials in inhibitor-containing

solution move negatively with respect to that in blank solution and that the shift extents

are concentration dependent. Even so, the studied polymers can not be arbitrarily

regarded as cathodic-type inhibitors since the corrosion current density of anode reaction

is reduced as well. In addition, the displacements of corrosion potential less than 85 mV

(vs SCE) in inhibited solution are reported to be an indication of mixed-type inhibitory

actions (Kowsari, Arman, Shahini, Zandi, Ehsani, Naderi, Hanza, & Mehdipour, 2016).

Last but not least, the relatively separate lines in the cathodic area show the inhibitory

14
effect of inhibitor on hydrogen evolution reaction is predominant.

Table S2 (Supplementary Information) lists relevant parameters of polarization

curves such as corrosion potential (Ecorr), cathodic and anodic Tafel slopes (βc and βa),

corrosion current density (Icorr) and the inhibition efficiency (η) calculated based on the

equation as follows (Mobin & Rizvi, 2016):

0
I corr  I corr
 0
 100% (9)
I corr
0
where I corr and I corr are the values of corrosion current density for mild steel in the

absence and presence of inhibitors, respectively. Results from Table S2 reveal the

corrosion current density flowing through steel electrode (1.920 mA/cm2) is decreased

by the introduction of polymer additives and that such reduction effect depends on

inhibitor concentrations. For instance, in presence of 2000 ppm inhibitors, the lowest

current densities for KGMA-containing and KGMH-containing solution are 0.254

mA/cm2 and 0.159 mA/cm2, respectively. The variations of βc and βa obtained in

inhibited solution confirm the mixed inhibitory effect of polymers on both iron

dissolution reaction and hydrogen production reaction, compared to those obtained in

free solution. Data of inhibition efficiency show KGMH can better control the acid-

induced corrosion of mild steel than KGMA, which is in good agreement with the results

of gravimetric measurements.

3.3.2 Impedance studies

In this work, electrochemical impedance spectroscopy was recorded on working

electrode to investigate the corrosion performance occurring at the interface between

mild steel and hydrochloric acid solution. The Nyquist plots for steel electrode immersed

in 0.5 M HCl in the absence and presence of inhibitors are presented in Fig. 5. As it can

be seen, the Nyquist diagrams in both uninhibited and inhibited solution present irregular

15
semicircles, suggesting the occurrence of non-ideal electrochemical behavior on the

heterogeneous surface of mild steel (Xu, Yang, Liu, Yin, Gong, & Chen, 2014). It is clear

that there is only one capacitive semicircle in the studied frequency range from 100 mHz

to 100 kHz, i.e., the interaction between inhibitor molecules and steel substrate in acid

solution can be characterized by the typical model of double layer capacitance (Cdl). In

addition, the diameters of those capacitive loops increasing with inhibitor concentrations

show the corrosion resistance of metal surface is continuously enhanced by the

adsorption of more polymer molecules.

Electrochemical impedance data in the form of Bode are presented in Fig. 6. As

evident from the graphs, two impedance plateaus at the high frequency and low

frequency are indications of solution resistance (Rs) and charge transfer resistance (Rct),

respectively (Ansari, Quraishi, & Singh, 2014). As the concentration of polysaccharide

esters increases, the values of Rct keep increasing while the values of Rs have small

changes compared with the blank solution. That indicates the added polymers are fully

adsorbed on metal surface to form an effective protective layer against corrosion. The

curves of phase angle for mild steel in 0.5 M HCl solution with and without inhibitors

are drawn in Fig. 7. It can be observed that all the phase angles in the figure are less than

90o,

16
indicating the rough surface of metallic electrode and complicated electrochemical

processes. With the addition of polymer inhibitors, phase angle moves upward to larger

values and the shift trend is concentration dependent. This phenomenon demonstrates

more inhibitor molecules adhere to the exposed surface of mild steel and make it more

homogeneous. The solo peak of phase angle located at middle frequency reveals the

single time constant related to double layer capacitance.

Above-mentioned analysis makes it reasonable to fit the impedance data with a

Randles equivalent circuit (Fig. S1, Supplementary Information) composed of solution

resistance (Rs), charge transfer resistance (Rct) and constant phase element (CPE). The

physical property of CPE can be defined by the following formula (Yilmaz, Fitoz, Ergun,

& Emregul, 2016):

ZCPE  Y01  j 
n
(10)
where Y0 is the magnitude of CPE, w is the angular frequency, j2 = –1 is an imaginary

number and n is the exponent of CPE. For n = 1, CPE can be regarded as a pure capacitor.

For n = –1 or 0, CPE stands for an inductor or a pure resistance, accordingly.

Besides, the values of double layer capacitance (Cdl) and inhibition efficiency (η) were

obtained by using the equations below:

Cdl  Y0  j 2f 
n 1
(11)
Rct  Rct0
 100% (12)
Rct
where f is the frequency at which the imaginary component of the impedance has a

maximum, Rct and Rct0 are the charge transfer resistance in the presence and absence

of inhibitors, respectively. Corresponding impedance parameters obtained by low-error

fitting and further calculation are provided in Table S3 (Supplementary Information). It

can be seen that the values of Rs in both inhibitor-free solution and inhibitor-containing

solution are very close (1 Ωcm2 < Rs < 2 Ωcm2). For Rct in the blank solution (4.06

17
Ωcm2), it keeps increasing as the dosage of inhibitors increases. For example, values of

Rct for the acid solution containing 2000 ppm KGMA and 2000 ppm KGMH are 36.08

Ωcm2 and 46.59 Ωcm2, respectively. The results of inhibition efficiency show the highest

effectiveness of inhibitor at 2000 ppm is 88.7% for KGMA and 91.3% for KGMH and

that the latter always has a better inhibitory effect than the former. In order to study the

influence of inhibitors on the double layer capacitance at metal/solution interface,

Helmholtz model is employed as follows:

 0
Cdl  S (13)
d
where d is the thickness of the double layer capacitor, S is the exposed area of electrode,

ε0 and ε are the dielectric constant in vacuum and test solution, respectively (Srivastava,

Tiwari, Srivastava, Prakash, & Ji, 2017). Table S3 shows the value of Cdl in 0.5 M HCl

is reduced by the introduction of inhibitors and that such decrease depends on the

concentrations of added polymers. According to Helmholtz formula, the decrease in Cdl

may be attributed to the decrease of local dielectric constant or the increase in the

thickness of double layer capacitor. In other words, the pre-adsorbed water molecules or

hydrated ions are expelled by added inhibitor molecules which can cover the active sites

on metal surface and retard the diffusion of corrosive substance. Furthermore, the

declining Cdl values may suggest the adsorptive layer of polymer molecules formed on

substrate surface become more compact and thicker when more inhibitors are added to

corrosive medium.

As shown in Table S4 (Supplementary Information), the inhibition efficiencies of

KGMA and KGMH for mild steel corroding in 0.5 M HCl solution have been compared

with the reported effectiveness of some other natural products (Abdel-Gaber et al, 2011;

Bello et al, 2010; EL-Haddad, 2014; Fares, Maayta, & Al-Mustafa, 2012; Ji, Anjum,

Sundaram, & Prakash, 2015; Li, Deng, Fu, & Xie, 2014; Sangeetha, Meenakshi, &

18
Sundaram, 2016; Singh, Ahamad, & Quraishi, 2016). The data from Table S4 suggest,

among those green inhibitors konjac glucomannan esters do not show the best inhibition

performance. However, the efficiencies of 90% and 92% are acceptable for restraining

the acid corrosion of iron. Therefore KGMA and KGMH can be classified into good

polysaccharide inhibitors.

3.4 Surface morphological characterization

The surface morphologies of mild steel samples before and after 24 h of immersion

in 0.5 M HCl solution in the absence and presence of 2000 ppm inhibitor were examined

by SEM and the obtained images are presented in Fig. S2 (Supplementary Information).

The observation on Fig. S2 (a) shows the iron coupon without immersion has the

smoothest surface on which there are some distinct traces of polishing as well as

scattered defects. As shown in Fig. S2 (b), the sample of mild steel immersed in

hydrochloric acid is badly corroded by aggressive solution and huge corrosion products

and cracks can be found on substrate surface. In inhibitor-containing solutions, the much

smoother surfaces without obvious corrosion pits displayed in Fig. S2 (c) and (d)

demonstrate the acid corrosion of iron is greatly inhibited by KGMA and KGMH. By

careful comparison of those two images, we can notice that the surface smoothness of

the sample in KGMH-added solution is higher than the one in KGMH-added solution

due to clearer lines of polishing and less particles of corrosion product in Fig. S2 (d).

That distinction confirms the better inhibiting effect of KHMH for mild steel corroding

in acid environment.

3.5 UV-visible spectroscopic measurements

The UV-visible spectra for 0.5 M HCl solution with 2000 ppm inhibitors before and

after 7 days of immersion of mild steel samples were recorded to confirm the adsorption

of polymer molecules and further formation of inhibitor-Fe (II) complex (Fig. S3,

19
Supplementary Information). First of all, we should point out that, as shown in Fig. S3

(a) and (b), the absorption spectra mainly distributed from 200 nm to 400 nm in the

presence of KGMA and KGMH are almost the same, indicating the active sites to absorb

ultraviolet lights are consistent for different inhibitor molecules. Before the immersion

of mild steel, the adsorption band located at about 200 nm is attributed to the π – π*

transition of –COO– group and the absorbance values are 1.52 for KGMA and 2.08 for

KGMH (Fan, Smuts, Bai, Walsh, Armstrong, & Schug, 2016). The greater absorbance

of the latter may result from its less molecular weight. A weak peak of adsorption can be

observed at 250 nm with an absorbance of 0.12 for KGMA and 0.19 for KGMH,

suggesting the n – α* transition of N or O atoms. After 7 days of immersion, the

adsorption bands at 200 nm move to 209 nm for KGMA (adsorption 3.0) and 213 nm

for KGMH (adsorption 3.5). Such phenomenon of bathochromic shift is ascribed to the

decrease in the energy gap of π – π* transition caused by the formation of a complex

between metal ions and inhibitor molecules in acid solution (Yadav, & Quraishi, 2012).

We can see the ultraviolet adsorption band at 250 nm is not affected by the

organometallic compound although it is shielded by the broad adsorption band of π – π*

transition. Interestingly, new and compressed bands appear at 340 nm for KGMA and

338 nm for KGMH, probably attributed to the charge transfer transition between the iron

substrate with empty d-orbital and the electron-rich groups in inhibitor molecules like

imidazole ring and N atoms as well as O atoms (Rondi, Rodriguez, Feurer, & Cannizzo,

2015). As a result, UV-visible spectra confirm the adsorption of polysaccharide esters

on mild steel surface and the formation of metal–inhibitor complex.

3.6 Quantum chemical calculations

To better understand the interaction of inhibitor with mild steel and the inhibitory

ability of polymers determined by molecular structures at a molecular level, theoretical

20
calculations were performed using the methodology of density functional theory (DFT).

Herein, considering the enormous size of polysaccharide esters, we choose a repeated

unit of KGMA and KGMH as the representative molecules for quantum chemical

studies. Fig. S4 (Supplementary Information) presents the optimized geometrical

structures of inhibitor molecules with the lowest energy. As can be seen from Fig. S4,

KGMA (a) has a larger size of stable configuration than KGMH (b) due to the linear

stretching of long-chain substituent groups. We may deduce that it is easier for KGMH

to be adsorbed on substrate surface because of a weaker steric effect.

According to Frontier Molecular Orbital Theory, the corrosion inhibition reactivity

of compounds is closely related to the highest occupied molecular orbital (HOMO) and

the lowest unoccupied molecular orbital (LUMO). The distributions of HOMO and

LUMO for inhibitor molecules are shown in Fig. S5 and corresponding parameters are

provided in Table S5 (Supplementary Information). Inspection of Fig. S5 shows the

HOMO of KGMA is mainly located at the N-containing groups in the branched chain

at C-3 position and that the LUMO is primarily distributed at the ester group as well as

amidogen in the substituted chain at C-2 position. For KGMH, the HOMO nearly covers

the whole substituted chain at C-2 position and the LUMO unexpectedly appears at the

glucoside ring. Such inconstant distributions of molecular orbital possess polymer

inhibitors of more active sites to donate electrons to the vacant d-orbital of Fe atoms or

accept free electron from metal using their anti-band orbital. As we know, the energy of

HOMO (EHOMO) suggests the electron donating ability of molecules whereas the energy

of LUMO (ELUMO) indicates the ability of molecules to accept electrons. That is to say,

values of higher EHOMO or lower ELUMO reveal inhibitor molecules can share electrons

with metal more easily and thereby possess better inhibitive performance on corrosion.

Moreover, the difference between EHOMO and ELUMO (ΔE) means the energy gap of the

21
charge transition process, i.e., a smaller value of ΔE shows a stronger interaction of the

inhibitor-iron system and hence a higher inhibition efficiency (Gece,2008; Tian, Li, Cao,

& Hou, 2013; Hegazy, Badawi, Abd El Rehim, & Kamel, 2013). The data from Table

S5 manifest KGMH has a higher value of EHOMO (–4.54 eV) and lower values of ELUMO

(–0.85 eV) and ΔE (3.69 eV) than KGMA, which theoretically accounts for the better

ability of the former to restrain corrosion for mild steel in HCl solution.

Dipole moment (μ) is another important parameter to evaluate the adsorptive ability

of an organic molecule (Abd El-Lateef, Soliman, & Tantawy, 2017). However, there are

different opinions about the influence of μ on corrosion inhibition in previous reports.

Some authors have demonstrated that a lower value of μ will enhance the accumulation

of inhibitor molecules on metal surface (Kabanda, Murulana, Ozcan, Karadag, Dehri,

Obot, & Ebenso, 2012). In contrast, some other researches have shown that a higher

value of μ will result in the increase of inhibition efficiency due to an enhanced dipole-

dipole interaction between molecules and metallic substrate (Khalil, 2003). In our work,

the values of μ for KGMA and KGMH are 11.93 D and 6.24 D, respectively. Given the

better performance of KGMH on corrosion inhibition, the first viewpoint about dipole

moment is deemed reasonable for this study.

The ionization potential (I) and the electron affinity (A) related to the energy of

molecular orbital are defined as I = –EHOMO and A = –ELUMO, respectively. In addition,

electronegativity (χ) and global hardness (γ) of the studied molecule are also quoted as

follows (Abdullach, 2002):

IA
 (14)
2
IA
 (15)
2
The fraction of electrons transferred from inhibitor molecule to metal surface (ΔN) is

calculated by the quantum chemical equation (Kovacevic, & Kokalj, 2013):

22
 Fe  inh
N  (16)
2   Fe   inh 
where χFE ≈ 7 eV and γFe = 0 eV are the theoretical values of the electronegativity and

global hardness of iron, respectively. General speaking, ΔN suggests the intensity and

direction of electron transfer. As proposed by Lukovits et al., a greater value of ΔN

implies a better electron donating ability and hence a higher inhibition efficiency if ΔN

is less than 3.6 (Lukovits, Kálmán, & Zucchi, 2001). Meanwhile, values of ΔN exhibit

the electron transfer from the molecule to metal surface in the case of ΔN > 0 and an

opposite action in the case of ΔN < 0 (Ju, Kai, & Li, 2008). As shown in Table S5, the

calculated ΔN values are 0.84 for KGMA and 1.17 for KGMH, indicating those inhibitor

molecules mainly donate electrons to iron surface and that KGMH has a stronger

electron-sharing trend with metal than KGMA. The total energy (Et) of studied

molecules is –2139.88 Ha for KGMA and –2276.62 Ha for KGMH, i.e., the later is more

active to interact with Fe substrate because of a higher energy. This again indicates the

better effect of KGMH molecules on corrosion control.

3.7 Mechanism of corrosion inhibition

In this study, experimental results show the acid corrosion of mild steel is retarded

by the mixed-type adsorption of KGM esters which involves both physisorption and

chemisorption. It has been reported that the surface of iron in hydrochloric acid is

positively charged (Yüce & Kardas, 2012). Afterwards, a number of chloride ions (Cl-)

may be attracted to metal surface due to specific adsorption, which will charge the

substrate surface with negative charge. We can assume that the inhibitor molecule could

be protonated because of the electrophilic attack of H+ on the N atoms with lone pairs of

electrons. As a result, the polymer molecule with positive charge will be adsorbed on

mild steel surface by a physical electrostatic interaction. At the same time, the inhibitor

molecule may donate free π-electrons from unsaturated bonds (–C=O– and –C=N–) or

23
n-electrons from N and O atoms to the empty d-orbital of Fe to form a coordinate bond

and accept electrons from iron with anti-bonding orbital to yield a feedback bond. In a

word, those processes will favor the adsorption of polysaccharide esters which can

effectively isolate the aggressive medium and protect mild steel against acid corrosion.

4. Conclusions

In this work, konjac glucomannan is modified by esterification reaction with L-

histidine and L-arginine to prepared potential corrosion inhibitors, KGMA and KGMH.

The results of gravimetric measurements reveal both KGMA and KGMH are effective

inhibitors for mild steel corroding in 0.5 M HCl solution and that their inhibition

efficiencies are concentration dependent. Potentiodynamic polarization curves indicate

those polysaccharide esters perform as mixed-type corrosion inhibitors with a dominant

cathodic effect. EIS studies suggest the charge transfer resistance of corrosion process is

enhanced by polymer additives and that KGMH has a better inhibitory effectiveness than

KGMA. The adsorption of inhibitor molecules on iron surface is spontaneous and

follows Langmuir isotherm. SEM analysis proves that mild steel in inhibitor-containing

hydrochloric acid receives good protection compared with the uninhibited sample.

Quantum chemical calculations provide useful information about the electronic property

of polymer molecules and clarify the difference in inhibition efficiency. UV-vis

spectroscopic measurements confirm the interaction between polysaccharide derivatives

and mild steel surface and the formation of inhibitor-Fe2+ complex in HCl solution.

Acknowledgments

This work was supported by the National Science & Technology Pillar Program

during the Twelfth Five-year Plan Period for Seawater Desalination Technology (Grant

No. 2015BAB08B00), Natural Science Foundation for Young Scholars of Jiangsu

Province, China (Grant No. BK20160983), National Key R&D Plan (Grant No.

24
2017YFC0404100) and National Natural Science Foundation of China (Grant No.

21605084).

References

Abdel-Gaber, A. M., Abd-El-Nabey, B. A., Khamis, E., & Abd-El-Khalek, D. E.

(2011). A natural extract as scale and corrosion inhibitor for steel surface in brine

solution. Desalination, 278, 337–342.

Abd El-Lateef, H. M., Soliman, K. A., & Tantawy, A. H. (2017). Novel synthesized

Schiff Base-based cationic gemini surfactants: Electrochemical investigation,

theoretical modeling and applicability as biodegradable inhibitors for mild steel

against acidic corrosion. Journal of Molecular Liquids, 232, 478–498.

Abdullach, M. (2002). Rhodanine azosulpha drugs as corrosion inhibitors for

corrosion of 304 stainless steel in hydrochloric acid solution. Corrosion Science, 44,

717–728.

Ansari, K. R., Quraishi, M. A., & Singh, A. (2014). Schiff’s base of pyridyl substituted

triazoles as new and effective corrosion inhibitors for mild steel in hydrochloric acid

solution. Corrosion Science, 79, 5–15.

Arukalam, I. O. (2014). Durability and synergistic effects of KI on the acid corrosion

inhibition of mild steel by hydroxypropyl methylcellulose. Carbohydrate Polymers,

112, 291–299.

Awad, M. I., Saad, A. F., Shaaban, M. R., Jahdaly, B. A., & Hazazi, O. A. (2017).

New insight into the mechanism of the inhibition of corrosion of mild steel by some

amino acids. International Journal of Electrochemical Science, 12, 1657–1669.

Barouni, K., Bazzi, L., Salghi, R., Mihit, M., Hammouti, B., Albourine, A., &

Issami, S. (2008). Some amino acids as corrosion inhibitors for copper in nitric acid

25
solution. Materials Letters, 62, 3325–3327.

Bello, M., Ochoa, N., Balsamo, V., López-Carrasquero, F., Coll, S., Monsalve, A., &

González, G. (2010). Modified cassava starches as corrosion inhibitors of carbon

steel: An electrochemical and morphological approach. Carbohydrate Polymers, 82,

561–568.

Brindha, T., Mallika, J., & Sathyanarayana, M. V. (2015). Synergistic effect

betweenstarch and substituted piperidin-4-one on the corrosion inhibition of mild

steelin acidic medium. Journal of Materials and Environmental Science, 6, 191–200.

Chen, J., Li, J., & Li, B. (2011). Identification of molecular driving forces involved in the

gelation of konjac glucomannan: Effect of degree of deacetylation on hydrophobic

association. Carbohydrate Polymers, 86, 865–871.

Chen, Y. Y., Zhao, H. Y., Liu, X. W., Li, Z. S., Liu, B., Wu, J. D., Shi, M. X., Norde, W., &

Li, Y. (2016). TEMPO-oxidized Konjac glucomannan as appliance for the preparation

of hard capsules. Carbohydrate Polymers, 143, 262–269.

Cui, J., Shi, R., & Pei, Y. (2017). Novel inorganic solid controlled-release inhibitor for

Q235-b anticorrosion treatment in 1 M HCl. Applied Surface Science, 416, 213–224.

Dandia, A., Gupta, S. L., Singh, P., & Quraishi, M. A. (2013). Ultrasound-assisted

synthesis of pyrazolo[3,4‑b]pyridines as potential corrosion inhibitors for mild steel

in 1.0 M HCl. ACS Sustainable Chemistry & Engineering, 1, 1303–1310.

Dohare, P., Ansari, K. R., Quraishi, M. A., & Obot, I. B. (2017). Pyranpyrazole

derivatives as novel corrosion inhibitors for mild steel useful for industrial pickling

process: Experimental and Quantum Chemical study. Journal of Industrial and

Engineering Chemistry, 52, 197–210.

26
Duan, J. C., Lu, Q., Chen, R. W., Duan, Y. Q., Wang, L. F., Gao, L., & Pan, S. Y. (2010).

Synthesis of a novel flocculant on the basis of crosslinked Konjac glucomannan-

graft-polyacrylamide-co-sodium xanthate and its application in removal of Cu2+ ion.

Carbohydrate Polymers, 80, 436–441.

Dutta, A., Saha, S. K., Adhikari, U., Banerjee, P., & Sukul, D. (2017). Effect of

substitution on corrosion inhibition properties of 2-(substituted phenyl)

benzimidazole derivatives on mild steel in 1 M HCl solution: A combined

experimental and theoretical approach. Corrosion Science, 123, 256–266.

Elemike, E. E., Nwankwo, H. U., Onwudiwe, D. C., & Hosten, E. C. (2017).

Synthesis, structures, spectral properties and DFT quantum chemical calculations of

(E)-4-(((4-propylphenyl)imino)methyl)phenol and (E)-4-((2-

tolylimino)methyl)phenol; their corrosion inhibition studies of mild steel in aqueous

HCl. Journal of Molecular Structure, 1141, 12–22.

El-Haddad, M. N. (2014). Hydroxyethylcellulose used as an eco-friendly inhibitor for

1018c-steel corrosion in 3.5% NaCl solution. Carbohydrate Polymers, 112, 595–602.

Fan, H., Smuts, J., Bai, L., Walsh, P., Armstrong, D. W., & Schug, K. A. (2016). Gas

chromatography–vacuum ultraviolet spectroscopy for analysis of fatty acid methyl

esters. Food Chemistry, 194, 265–271.

Fares, M. M., Maayta, A. K., & Al-Mustafa, J. A. (2012). Corrosion inhibition of iota-

carrageenan natural polymer on aluminum in presence of zwitterion mediator in

HCl media. Corrosion Science, 65, 223–230.

27
Fekry, A. M., & Mohamed, R. R. (2010). Acetyl thiourea chitosan as an eco-friendly

inhibitor for mild steel in sulphuric acid medium. Electrochimica Acta, 55, 1933–

1939.

Gece, G. (2008). The use of quantum chemical methods in corrosion inhibitor studies.

Corrosion Science, 50, 2981–2992.

Gece, G., & Bilgiç, S. (2010). A theoretical study on the inhibition efficiencies of some

amino acids as corrosion inhibitors of nickel. Corrosion Science, 52, 3435–3443.

Hegazy, M.A., Badawi, A.M., Abd El Rehim, S. S., & Kamel, W.M. (2013). Corrosion

inhibition of carbon steel using novel N-(2-(2-mercaptoacetoxy)ethyl)-N,N-

dimethyl) dodecan-1-aminium bromide during acid pickling. Corrosion Science, 69,

110–122.

Hegazy, M. A., Hasan, A. M., Emara, M. M. M., Bakr, M. F., & Youssef, A. H. (2012).

Evaluating four synthesized Schiff bases as corrosion inhibitors on the carbon steel

in 1 M hydrochloric acid. Corrosion Science, 65, 67–76.

Ji, G., Anjum, S., Sundaram, S., & Prakash, R. (2015). Musa paradisica peel extract as

green corrosion inhibitor for mild steel in HCl solution. Corrosion Science, 90, 107–

117.

Ju, H., Kai, Z. P., & Li, Y. (2008). Aminic nitrogen-bearing polydentate Schiff base

compounds as corrosion inhibitors for iron in acidic media: a quantum chemical

calculation. Corrosion Science, 50, 865–871.

Kabanda, M. M., Murulana, L. C., Ozcan, M., Karadag, F., Dehri, I., Obot, I. B., &

Ebenso, E. E. (2012). Quantum chemical studies on the corrosion inhibition of mild

28
steel by some triazoles and benzimidazole derivatives in acidic medium.

International Journal of Electrochemical Science, 7, 5035–5056.

Khalil, N. (2003). Quantum chemical approach of corrosion inhibition. Electrochimica

Acta, 48, 2635–2640.

Khan, G., Basirun, W. J., Kazi, S. N., Ahmed, P., Magaji, L., Ahmed, S. M., Khan,

G. M., & Rehman, M. A. (2017). Electrochemical investigation on the corrosion

inhibition of mild steel by quinazoline Schiff base compounds in hydrochloric acid

solution. Journal of Colloid and Interface Science, 502, 134–145.

Klemm, D., Philipp, B., Heinze, T., Heinze, U., & Wagenknecht, W. (1998).

Comprehensive Cellulose Chemistry. New York: Wiley VCH, Vol. II.

Koh, E., & Park, S. (2017). Self-anticorrosion performance efficiency of renewable

dimer-acid-based polyol microcapsules containing corrosion inhibitors with two

triazole groups. Progress in Organic Coatings, 109, 61–69.

Kovacevic, N., & Kokalj, A. (2013). The relation between adsorption bonding and

corrosion inhibition of azole molecules on copper. Corrosion Science, 73, 7–17.

Kowsari, E., Arman, S. Y., Shahini, M. H., Zandi, H., Ehsani, A., Naderi, R., Hanza, A.

P., & Mehdipour, M. (2016). In situ synthesis, electrochemical and quantum

chemical analysis of an amino acid-derived ionic liquid inhibitor for corrosion

protection of mild steel in 1M HCl solution. Corrosion Science, 112, 73–85.

Kumar, R., Yadav, O. S., & Singh, G. (2017). Electrochemical and surface

characterization of a new eco-friendly corrosion inhibitor for mild steel in acidic

media: A cumulative study. Journal of Molecular Liquids, 237, 413–427.

29
Li, X. H., Deng, S. D., Fu, H., & Xie, X. G. (2014). Synergistic inhibition effects of

bamboo leaf extract/major components and iodide ion on the corrosion of steel in

H3PO4 solution. Corrosion Science, 78, 29–42.

Lukovits, I., Kálmán, E., & Zucchi, F. (2001). Corrosion inhibitors-correlation between

electronic structure and efficiency. Corrosion, 57, 3–8.

Mobin, M., & Rizvi, M. (2016). Inhibitory effect of xanthan gum and synergistic

surfactant additives for mild steel corrosion in 1 M HCl. Carbohydrate Polymers,

136, 384–393.

Mobin, M., & Rizvi, M. (2017). Polysaccharide from Plantago as a green corrosion

inhibitor for carbonsteel in 1 M HCl solution. Carbohydrate Polymers, 160, 172–183.

Oguzie, E. E., Oguzie, K. L., Akalezi, C. O., Udeze, I. O., Ogbulie, J. N., & Njoku, V. O.

(2013). Natural products for materials protection: corrosion and microbial growth

inhibition using capsicum frutescens biomass extracts. ACS Sustainable Chemistry &

Engineering, 1, 214–225.

Ramachandran, S., Tsai, B. L., Blanco, M., Chen, H., Tang, Y. C., & Goddard, W. A.

(1996). Self-assembled monolayer mechanism for corrosion inhibition of iron by

imidazolines. Langmiur, 12, 6419–6428.

Rondi, A., Rodriguez, Y., Feurer, T., & Cannizzo, A. (2015). Solvation-driven charge

transfer and localization in metal complexes. Accounts of Chemical Research, 48,

1432–1440.

30
Sangeetha, Y., Meenakshi, S., & Sundaram, C. S. (2016). Corrosion inhibition of

aminated hydroxyl ethyl cellulose on mild steel in acidic condition. Carbohydrate

Polymers, 150, 13–20.

Singh, A., Ahamad, I., & Quraishi, M. A. (2016). Piper longum extract as green

corrosion inhibitor for aluminium in NaOH solution. Arabian Journal of Chemistry,

9, S1584–S1589.

Sinha, V. R., & Kumria, R. (2001). Polysaccharides in colon-specific drug delivery.

International Journal of Pharmaceutics, 224, 19–38.

Souza, F. S., & Spinelli, A. (2009). Caffeic acid as a green corrosion inhibitor for mild

steel. Corrosion Science, 51, 642–649.

Srivastava, M., Tiwari, P., Srivastava, S. K., Prakash, R., & Ji, G. (2017).

Electrochemical investigation of Irbesartan drug molecules as an inhibitor of mild

steel corrosion in 1 M HCl and 0.5 M H2SO4 solutions. Journal of Molecular Liquids,

236, 184–197.

Tian, H., Li, W., Cao, K., & Hou, B. (2013). Potent inhibition of copper corrosion in

neutral chloride media by novel thiadiazole derivatives. Corrosion Science, 73, 281–

291.

Umoren, S. A., & Eduok, U. M. (2016). Application of carbohydrate polymers as

corrosion inhibitors for metal substrates in different media: A review. Carbohydrate

Polymers, 140, 314–341.

Umoren, S. A., & Ekanem, U. F. (2010). Inhibition of mild steel corrosion in H2SO4

using exudate gum from Pachylobus edulis and synergistic potassium halide

additives. Chemical Engineering Communications, 197, 1339–1356.

31
Umoren, S. A., Obot, I. B., Madhankumar, A., & Gasem, Z. M. (2015). Performance

evaluation of pectin as ecofriendly corrosion inhibitor for X60 pipeline steel in acid

medium: Experimental and theoretical approaches. Carbohydrate Polymers, 124,

280–291.

Wang, Y., Chen, Y. H., Zhou, Y., Nirasawa, S., Tatsumi, E., Li, X. T., & Cheng, Y. Q.

(2017). Effects of konjac glucomannan on heat-induced changes of wheat gluten

structure. Food Chemistry, 229, 409–416.

Xiao, M., Dai, S. H., Wang, L., Ni, X. W., Yan, W. L., Fang, Y. P., Corke, H., & Jiang, F.

T. (2015). Carboxymethyl modification of konjac glucomannan affects water

binding properties. Carbohydrate Polymers, 130, 1–8.

Xu, B., Yang, W. Z., Liu, Y., Yin, X. S., Gong, W. N., & Chen, Y. Z. (2014). Experimental

and theoretical evaluation of two pyridinecarboxaldehyde thiosemicarbazone

compounds as corrosion inhibitors for mild steel in hydrochloric acid solution.

Corrosion Science, 78, 260–268.

Yadav, D. K., & Quraishi, M. A. (2012). Application of some condensed uracils

ascorrosion inhibitor for mild steel: Gravimetric, electrochemical,

surfacemorphological, UV–visible and theoretical investigations. Industrial &

Engineering Chemistry and Research, 51, 14966–14979.

Yadav, M., Sarkar, T. K., & Purkait, T. (2015). Amino acid compounds as eco-friendly

corrosion inhibitor for N80 steel in HCl solution: Electrochemical and theoretical

approaches. Journal of Molecular Liquids, 212, 731–738.

32
Yilmaz, N., Fitoz, A., Ergun, Y., & Emregul, K. C. (2016). A combined electrochemical

and theoretical study into the effect of 2-((thiazole-2-ylimino)methyl)phenol as a

corrosion inhibitor for mild steel in a highly acidic environment. Corrosion Science,

111, 110–120.

Yoshida, S., & Ishida, H. (1984). A FT-IR reflection-absorption spectroscopic study of

an epoxy coating on imidazole-treated copper. Journal of Adhesion, 16, 217–232.

Yüce, A. O., & Kardas, G. (2012). Adsorption and inhibition effect of 2-

thiohydantoinon mild steel corrosion in 0.1 M HCl. Corrosion Science, 58, 86–94.

Zhang, K. G., Xu, B., Yang, W. Z., Yin, X. S., Liu, Y., & Chen, Y. Z. (2015). Halogen-

substituted imidazoline derivatives as corrosion inhibitors for mild steel in

hydrochloric acid solution. Corrosion Science, 90, 284–295.

Zhang, K. G., Yang, W. Z., Xu, B., Yin, X. S., Liu, Y., & Chen, Y. Z. (2015).

Corrosion inhibition of mild steel by bromide-substituted imidazoline in hydrochloric

acid. Journal of the Taiwan Institute of Chemical Engineers, 57, 167–174.

Figure captions

Fig. 1 The synthetic route to prepare konjac glucomannan esters by using amino acids.

Fig. 2 FT-IR spectroscopy of KGM, KGMA and KGMH.

Fig. 3 Langmuir isotherm plots for the adsorption of polymer inhibitors on mild steel

surface in 0.5 M HCl at 298K.

Fig. 4 Potentiodynamic polarization curves recorded on iron electrodes immersed in 0.5

M HCl solution in the absence and presence of different concentrations of inhibitor at

298 K: (a), KGMA; (b), KGMH.

Fig. 5 Nyquist plots for mild steel samples measured in 0.5 M HCl solution without and

33
with various concentrations of KGMA (a) and KGMH (b) at 298 K.

Fig. 6 Bode curves of mild steel specimens determined in 0.5 M hydrochloric acid

without and with the addition of different concentrations of inhibitor at 298 K: (a),

KGMA; (b), KGMH.

Fig. 7 Variations of the phase angle for mild steel electrode exposed to 0.5 M HCl in the

absence and presence of various concentrations of polysaccharide derivatives at 298 K:

(a), KGMA; (b), KGMH.

34
35
Fig. 1

Fig. 2

36
Fig. 3

37
Fig. 4

38
Fig. 5

39
Fig. 6

40
Fig. 7

41

You might also like