Sie sind auf Seite 1von 22

Res Chem Intermed (2013) 39:1927–1948

DOI 10.1007/s11164-012-0726-3

Experimental, quantum chemical calculations,


and molecular dynamic simulations insight
into the corrosion inhibition properties
of 2-(6-methylpyridin-2-yl)oxazolo
[5,4-f][1,10]phenanthroline on mild steel

I. B. Obot • N. O. Obi-Egbedi • E. E. Ebenso •

A. S. Afolabi • E. E Oguzie
Received: 21 March 2012 / Accepted: 10 July 2012 / Published online: 24 July 2012
Ó Springer Science+Business Media B.V. 2012

Abstract 2-(6-Methylpyridin-2-yl)oxazolo[5,4-f][1,10]phenanthroline (MOP) was


synthesized and characterized by elemental analysis and Fourier-transform infrared
(FT-IR), 1H nuclear magnetic resonance (NMR), and 13C NMR spectra. MOP was
evaluated as a corrosion inhibitor for carbon steel in 0.5 M H2SO4 solution using the
standard gravimetric technique at 303–333 K. Quantum chemical calculations and
molecular dynamic (MD) simulations were applied to analyze the experimental data
and elucidate the adsorption behavior and inhibition mechanism of MOP. Results
obtained show that MOP is an efficient inhibitor for mild steel in H2SO4 solution.
The inhibition efficiency was found to increase with increase in MOP concentration
but decreased with temperature. Activation parameters and Gibbs free energy for
the adsorption process using statistical physics were calculated and discussed. The
adsorption of MOP was found to involve both physical and chemical adsorption
mechanisms. Density functional theory (DFT) calculations suggest that nitrogen and

I. B. Obot (&)
Department of Chemistry, Faculty of Science, University of Uyo, Uyo PMB 1017, Akwa Ibom
State, Nigeria
e-mail: proffoime@yahoo.com

N. O. Obi-Egbedi
Department of Chemistry, University of Ibadan, Ibadan, Nigeria

E. E. Ebenso
Department of Chemistry, School of Mathematical and Physical Sciences, North West University,
Mafikeng Campus, Private Bag X2046, Mmabatho 2735, South Africa

A. S. Afolabi
Department of Civil and Chemical Engineering, University of South Africa, Florida Campus,
Johannesburg, South Africa

E. E Oguzie
Electrochemistry and Material Science Research Laboratory, Department of Chemistry, Federal
University of Technology, Owerri PMB 1526, Nigeria

123
1928 I. B. Obot et al.

oxygen atoms present in the MOP structure were the active reaction sites for the
inhibitor adsorption on mild steel surface via donor–acceptor interactions between
the lone pairs on nitrogen and oxygen atoms together with the p-electrons of the
heterocyclic and the vacant d-orbital of iron atoms. The adsorption of MOP on Fe
(1 1 0) surface was parallel to the surface so as to maximize contact, as shown in the
MD simulations. The experiments together with DFT and MD simulations provide
further insight into the mechanism of interaction between MOP and mild steel.

Keywords Mild steel  Corrosion inhibitor  Quantum chemical calculation 


Molecular dynamic simulation

Introduction

Corrosion prevention is of great importance in industrial application of materials.


Among numerous corrosion prevention measures, corrosion inhibitors, bearing
advantages of economy, high efficiency, and wide applicability, have been applied
in various fields such as petroleum extraction and refining, iron and steel, electric
power, construction, etc. [1]. In recent years, with the recognition of environmental
protection, development of more effective and environmentally friendly corrosion
inhibitors has drawn more attention [2–4].
Most of the efficient inhibitors used in industry are organic compounds which
mainly contain oxygen, sulfur, and nitrogen atoms and multiple bonds in the
molecule through which they are adsorbed onto the metal surface [5]. Moreover,
many N-heterocyclic compounds have been used as effective inhibitors of corrosion
of metals and alloys in aqueous media [6–10]. It has been reported that many
organic inhibitors usually promote the formation of a chelate on the metal surface,
which includes the transfer of electrons from the organic compounds to the metal,
forming coordinate covalent bonding during such chemical adsorption process [11].
In this way, the metal acts as an electrophile, whereas the nucleophile centers of the
inhibitor molecule are normally heteroatoms with free electron pairs which are
readily available for sharing, to form a bond.
Theoretical chemistry has been used recently to explain the mechanism of
corrosion inhibition, such as quantum chemical calculations, which have been
proved to be a very powerful tool for studying the mechanism [12–14]. Quantum
chemical studies using density functional theory (DFT) have been successfully
performed by our research group to link the corrosion inhibition efficiency with the
molecular/electronic properties of organic molecules to explain the electron transfer
mechanism between the inhibitor and the steel surface [15–25].
Although various experimental techniques and theoretical methods have been
developed to study the relation between the efficiency and structural properties of
inhibitor molecules, little is known about the interactions between the adsorbed
molecules and metal surfaces [26], which is critical to understand the corrosion
inhibition phenomena. With suitable models, computer simulation can be a practical
way to study these complex processes. Computer simulation has been used to
investigate the formation of adsorption layers on metals and provided some valuable

123
Experimental, quantum calculations, and dynamic simulations 1929

information about the microstructures of the surfaces [27]. Recently, molecular


dynamics (MD) simulations have been used to study the interaction between several
corrosion inhibitors and metal surfaces [28–30]. Results from those studies
indicated that MD simulations can provide better insights into the design of
inhibitor systems with superior properties, and the adsorption energy between
organic molecule and metal surface given by MD simulations can be used to
interpret the difference in inhibition efficiency between organic inhibitors.
Phenanthroline derivatives are of interest due to their potential activity against
cancer and bacterial, viral, and fungal infections. Recently, Roy et al. [31] reported
on the use of phenanthroline derivatives with improved selectivity as DNA-
targeting anticancer or antimicrobial drugs. Imidazo[4,5-f][1,10]phenanthrolines are
versatile ligands because they can form stable complexes with various d-block
transition metals [32]. Metal complexes of phenanthroline derivative [for example,
Mn(II) complex of 2H-5-hydroxy-1,2,5-oxadiazo[3,4-f][1,10]phenanthroline] has
been reported to show antitumor activity [33]. The choice of 2-(6-methylpyridin-2-
yl)oxazolo[5,4-f][1,10]phenanthroline (MOP) for the present investigation was
based on the following considerations: (i) it can be synthesized easily from
relatively cheap materials; (ii) it contains three heterocyclic moieties in one
compound, i.e., oxazole, phenanthroline, and pyridine rings with several p-electrons
and aromatic systems containing four N and one O atoms, which can induce greater
adsorption of the inhibitor molecule onto the surface of mild steel compared with
compounds containing only phenanthroline ring already used as corrosion inhibitors
[34, 35]; and (iii) the compound has higher molecular weight than most
N-heterocyclic compounds already studied as corrosion inhibitors, such as
piperazine derivatives [36]. Moreover, there is no report in the literature on
experimental, quantum chemical, and MD simulations studies of 2-(6-methylpyri-
din-2-yl)oxazolo[5,4-f][1,10]phenanthroline as a corrosion inhibitor for mild steel in
sulfuric acid.
In the present work, the inhibition effect of 2-(6-methylpyridin-2-yl)oxazolo[5,4-
f][1,10]phenanthroline (MOP) on corrosion of mild steel in 0.5 M H2SO4 at
303–333 K was studied by the standard weight loss method. Also, quantum
chemical calculations using DFT and MD simulations studies were performed to
simulate the adsorption of MOP onto the iron surface.

Experimental

General considerations

All starting materials were obtained commercially as reagent grade and used
without further purification. Phenanthroline-5,6-dione was synthesized according to
literature procedure [37]. All air- or moisture-sensitive manipulations were carried
out under nitrogen atmosphere using standard Schlenk techniques. All purifications
were done on silica gel column to exclude impurities, and thin-layer chromatog-
raphy (TLC) glass slides were routinely employed to monitor extent of reaction as
well as the progress of silica gel column chromatography.

123
1930 I. B. Obot et al.

CH 3
N N
N

O
N

2-(6-methylpyridin-2-yl)oxazolo[5,4-f][1,10]phenanthroline
Fig. 1 Name and molecular structure of the inhibitor

Synthesis of 2-(6-methylpyridin-2-yl)oxazolo[5,4-f][1,10]phenanthroline

The inhibitor (MOP) was synthesized as reported in the literature [37]. Phenan-
throline-5,6-dione (3.00 g, 0.014 mol), ammonium acetate (22 g, 0.29 mol), and
6-methylpyridine-2-carboxaldehyde (2.18 g, 18.01 mmol, 1.25 equivalent) were
added, and the mixture was refluxed for 3 h. After cooling, the reaction mixture was
neutralized by concentrated aqueous ammonia and the volume was reduced under
reduced pressure. The organic components were extracted using chloroform
(50 mL, thrice), and the crude mixture was purified on silica gel column using
ethanol/ethylacetate/petroleum ether (1:10:10) as eluent. The portions that con-
tained pure products were concentrated, and the precipitates were filtered, washed
with small amount of ethanol, and dried to afford isolated product MOP (0.37 g,
8.2 %) as purple microcrystals. Figure 1 depicts the molecular structure of MOP.
Analytical data for inhibitor MOP: Yield, 0.37 g, 8.2 %. M.p. 266–268 °C.
Selected IR peaks (KBr disc, cm-1): v 3016 s, 1587 s, 1558 s, 1459 vs, 1062 s, 739 vs.
1
H NMR (400 MHz, TMS, CDCl3); d 9.27 (dd, J = 2.0, 5.7 Hz, 2H); 9.05 (dd,
J = 2.1, 10.8 Hz, 1H); 8.02 (dd, J = 2.1, 10.9 Hz, 1H); 8.28 (d, J = 10.3 Hz, 1H);
7.80 (m, 3H); 7.38 (d, J = 10.3 Hz, 1H); 2.79 (s, 3H). 13C NMR (100 MHz, CDCl3):
162.50, 159.85, 150.03, 149.75, 145.24, 145.09, 144.07, 137.40, 134.87, 131.20,
129.21, 125.54, 123.65, 123.40, 123.07, 120.62, 118.09, 124.79. Anal. Calc. for
C19H12N4O: C, 73.07; H, 3.87; N, 17.94 %. Found: C, 72.98; H, 3.85; N, 17.87 %.

Instrumentation

C, H, and N analyses were carried out on a Flash EA 1112 microanalyzer. 1H and 13C
NMR spectra were obtained with a Bruker ARX-400 MHz spectrometer using CDCl3
as solvent and tetramethylsilane (TMS) as internal standard. Fourier-transform
infrared (FT-IR) spectra were recorded in KBr pellets using a Shimadzu 8740 FT-IR
spectrometer as KBr discs in the range of 4,000 to 400 cm-1. The melting point was
determined on a digital melting point instrument (Electrothermal model 9200).

123
Experimental, quantum calculations, and dynamic simulations 1931

Material

Tests were performed on a freshly prepared sheet of mild steel of the following
composition (wt%): 0.13 % C, 0.18 % Si, 0.39 % Mn, 0.40 % P, 0.04 % S, 0.025 % Cu,
and bal Fe. Specimens used in the weight loss experiment were mechanically cut into
5.0 9 4.0 9 0.8 cm3 dimensions, then abraded with SiC abrasive papers of 320, 400,
and 600 grit, respectively, washed in absolute ethanol and acetone, dried at room
temperature, and stored in a moisture-free desiccator before use in corrosion studies [22].

Solutions

The aggressive solutions, 0.5 M H2SO4 were prepared by dilution of analytical-


grade 98 % H2SO4 with distilled water. Stock solution of MOP was made in 10:1
water:methanol mixture to ensure solubility [23]. This stock solution was used for
all experimental purposes. The concentration range of MOP prepared and used in
this study was 2–10 lM.

Gravimetric measurements

Weight loss measurements were conducted under total immersion using 250-mL-
capacity beakers containing 200 mL test solution at 303–333 K maintained in a
thermostated water bath. The mild steel coupons were weighed and suspended in the
beaker with the help of rod and hook. The coupons were retrieved at 2-h intervals
progressively for 10 h, washed thoroughly in 20 % NaOH solution containing
200 g/L zinc dust [16] with bristle brush, rinsed several times in deionized water,
cleaned, dried in acetone, and reweighed. Weight loss, in grams, was taken as the
difference in the weight of the mild steel coupons before and after immersion in
different test solutions as determined using a LP 120 digital balance with sensitivity
of ±0.1 mg. Then, the tests were repeated at different temperatures. To obtain good
reproducibility, experiments were carried out in triplicate. In this present study,
standard deviation values among parallel triplicate experiments were found to be
smaller than ±2 %, indicating good reproducibility.
The corrosion rate (q) in mg cm-2 h-1 was calculated from the following basic
equation [38]:
DW
q¼ ; ð1Þ
At
where W is the average weight loss of three mild steel sheets, A is the total area of
one mild steel specimen, and t is the immersion time (10 h). With the calculated
corrosion rate, the inhibition efficiency (%I) was calculated as follows:
 
q1  q2
%I ¼  100; ð2Þ
q1
where q1 and q2 are the corrosion rates of the mild steel coupons in the absence and
presence of MOP, respectively.

123
1932 I. B. Obot et al.

Computational details

B3LYP, a version of the DFT method that uses Becke’s three-parameter functional (B3)
and includes a mixture of Hartree–Fock (HF) with DFT exchange terms associated with
the gradient-corrected correlation functional of Lee, Yang, and Parr (LYP), was used in
this work to carry out quantum calculations. Then, full geometry optimization together
with vibrational analysis of the optimized structures of the inhibitor was carried out at
the B3LYP/6-31G(d) level of theory using the Spartan’06 V112 program package [39]
to determine whether they correspond to a minimum in the potential energy curve. The
quantum chemical parameters were calculated for molecules in neutral as well as
protonated form for comparison. It is well known that the phenomenon of
electrochemical corrosion occurs in liquid phase. As a result, it was necessary to
include solvent effects in the computational calculations. In the Spartan’06 V112
program, self-consistent reaction field (SCRF) methods were used to perform
calculations in aqueous solution. These methods model the solvent as a continuum of
uniform dielectric constant, and the solute is placed in the cavity within it.
There is no doubt that the recent progress in DFT has provided a very useful tool
for understanding molecular properties and for describing the behavior of atoms in
molecules. DFT methods have become very popular in the last decade due to their
accuracy and shorter computational time. DFT has been found to be successful in
providing insights into chemical reactivity and selectivity, in terms of global
parameters such as electronegativity (v), hardness (g), and softness (r), and local
ones such as the Fukui function f(r) and local softness s(r). Thus, for an N-electron
system with total electronic energy E and an external potential v(r), the chemical
potential l, known as the negative of the electronegativity v, has been defined as the
first derivative of E with respect to N at constant v(r) [40]:
 
oE
v ¼ l ¼  : ð3Þ
oN mðrÞ
Hardness (g) has been defined within DFT as the second derivative of E with
respect to N at constant v(r) [41]:
 2   
o E ol
g¼ 2
¼ : ð4Þ
oN mðrÞ oN mðrÞ
The number of electrons transferred (DN) from the inhibitor molecule to the
metal surface can be calculated by using the following equation [42]:
vFe  vinh
DN ¼ ; ð5Þ
½2ðgFe þ gFe Þ
where vFe and vinh denote the absolute electronegativity of iron and the inhibitor
molecule, respectively, and gFe and ginh denote the absolute hardness of iron and the
inhibitor molecule, respectively.
I and A are related in turn to the energy of the highest occupied molecular orbital
(EHOMO) and of the lowest unoccupied molecular orbital (ELUMO) using the
equations below [43]:

123
Experimental, quantum calculations, and dynamic simulations 1933

I ¼ EHOMO ; ð6Þ
A ¼ ELUMO : ð7Þ
These quantities are related to the electron affinity (A) and ionization potential (I)
using the equations below:
IþA ELUMO þ EHOMO
v¼ ; v¼ ; ð8Þ
2 2
IA ELUMO  EHOMO
g¼ ; g¼ : ð9Þ
2 2
Global softness can also be defined as [12]
1
r¼ : ð10Þ
g
The local reactivity of the MOP molecule was analyzed through evaluation of the
Fukui indices [44]. The Fukui indices are measures of chemical reactivity, as well as
an indicative of the reactive regions and the nucleophilic and electrophilic behavior of
the molecule. Regions of a molecule where the Fukui function is large are chemically
softer than regions where the Fukui function is small, and by invoking the hard and
soft acids and bases (HSAB) principle in a local sense, one may establish the behavior
of the different sites with respect to hard or soft reagents. The Fukui function f(r) is
defined as the first derivative of the electronic density q(r) with respect to the number
of electrons N at constant external potential v(r). Thus, using a scheme of finite-
difference approximations from Mulliken population analysis of atoms in MOP and
depending on the direction of electron transfer, we have [44]
fkþ ¼ qk ðN þ 1Þ  qk ðNÞ ðfor nucleophilic attackÞ; ð11Þ
fk ¼ qk ðNÞ  qk ðN  1Þ ðfor electrophilic attackÞ; ð12Þ
qk ðN þ 1Þ  qk ðN  1Þ
fko ¼ ðfor radical attackÞ; ð13Þ
2
where qk is the gross charge of atom k in the molecule, i.e., the electron density at a
point r in space around the molecule. N corresponds to the number of electrons in
the molecule. N ? 1 corresponds to an anion, with an electron added to the LUMO
of the neutral molecule; N - 1 corresponds to the cation with an electron removed
from the HOMO of the neutral molecule. All calculations were done at the ground-
state geometry. These functions can be condensed to the nuclei by using an atomic
charge partitioning scheme, such as Mulliken population analysis in Eqs. (11–13).
MD simulation of the interaction between a single MOP molecule and the Fe
surface was performed using Forcite quench MD in Material Studio (MS)
Modelling 4.0 software to sample many different low-energy minima and to
determine the global energy minimum [45]. Calculations were carried out in a
14 9 12 supercell using the condensed-phase optimized molecular potentials for
atomistic simulation studies (COMPASS) force field and the Smart algorithm. Of the
many kinds of Fe surfaces, Fe (1 1 0) is the most densely packed and also the most
stable [46]. The Fe crystal was cleaved along the (1 1 0) plane. The Fe slab built for

123
1934 I. B. Obot et al.

the docking process was significantly larger than the inhibitor molecules in order to
avoid edge effects during docking. Temperature was fixed at 350 K, with NVE
ensemble, with time step of 1 fs and simulation time of 5 ps. The system was
quenched every 250 steps with the Fe (1 1 0) surface atoms constrained. Optimized
structure of the inhibitor molecule was used for the simulation.
Adsorption of the MOP molecule onto the Fe (1 1 0) surface provides access to
the adsorption energetics and its effect on the inhibition efficiency of MOP. Thus,
the adsorption energy, Eads, between MOP molecule and Fe (1 1 0) surface was
calculated using Eq. (14) [46]:
Eads ¼ Ecomplex  ðEinh þ EFe Þ; ð14Þ
where Ecomplex is the total energy of the surface and inhibitor, EFe is the energy of
the Fe surface without the inhibitor, and Einh is the energy of the inhibitor without
the Fe surface.

Results and discussion

Weight loss measurements

The corrosion inhibition performance of organic compounds can be evaluated using


electrochemical and chemical techniques. For the chemical methods, weight loss
measurement is ideally suited for long-term immersion testing. Corroborative
results between weight loss and other techniques have been reported [47–51].
The anodic dissolution of iron in acidic media and the corresponding cathodic
reaction have been reported to proceed as follows [23]:
Fe ! Fe2þ þ 2e ð15Þ
þ 
2H þ 2e ! 2Hads ! H2 ð16Þ
As a result of these reactions, including the high solubility of the corrosion
products, the metal loses weight in the solution. Figure 2 shows plots of weight loss
versus time for mild steel in 0.5 M H2SO4 without and with different concentrations
of MOP at 303 K. From these plots, it is evident that the weight loss of mild steel in
the different test solutions increases with time. Further inspection of the plots
reveals that the weight loss of mild steel was reduced in the presence of MOP
compared with in the free acid solution: an indication of the inhibiting effect of acid
corrosion of mild steel.
The calculated values of corrosion rate (q) and inhibition efficiency (%I) obtained
from weight loss measurements for different concentrations of MOP in 0.5 M H2SO4
after 10 h immersion at 303–333 K are listed in Table 1. It is evident from this table
and Fig. 3 that the corrosion rate decreased with increasing inhibitor concentration
but increased with rise in temperature. Table 1 also shows that the inhibition
efficiency (%I) increased with increasing inhibitor concentration, reaching a
maximum of 90.0 %. This may be due to adsorption of MOP molecules onto the
mild steel surface through nonbonding electron pairs of the four nitrogen and one

123
Experimental, quantum calculations, and dynamic simulations 1935

Fig. 2 Variation of weight loss against time for mild steel corrosion in 0.5 M H2SO4 in the presence of
different concentrations of MOP at 303 K

Table 1 Calculated values of corrosion rate and inhibition efficiency for mild steel corrosion in 0.5 M
H2SO4 in absence and presence of MOP at 303–333 K
System/concentration Corrosion rate (mg cm-2 h-1) Inhibition efficiency (%I)

303 K 313 K 323 K 333 K 303 K 313 K 323 K 333 K

Blank 1.20 2.53 4.89 6.00 – – – –


2 9 10-6 M 0.59 1.39 2.76 3.45 51 45 44 21
4 9 10-6 M 0.44 1.11 2.35 3.20 53 56 51 30
6 9 10-6 M 0.26 0.76 1.76 3.00 78 70 64 43
8 9 10-6 M 0.20 0.63 1.49 2.50 83 75 70 53
10 9 10-6 M 0.12 0.38 1.03 2.00 90 85 79 62

oxygen atoms as well as the p-electrons of the aromatic rings. Moreover, the high
molecular weight of MOP ensures effective surface coverage of the inhibitor on the
steel surface. This isolates the steel from the aggressive acid solution, thus inhibiting
its dissolution. Similar observation has been documented [22].
Figure 4 shows the variation of percentage inhibition efficiency with tempera-
ture. It is clear from the figure that the percentage inhibition efficiency increases
with concentration but decreases with temperature. The increase in percentage
inhibition efficiency of MOP with concentration may be due to adsorption of its
molecules onto the mild steel surface. The decrease in inhibition efficiency with
increase in temperature may probably be due to decreasing strength of adsorption
(shifting the adsorption–desorption equilibrium towards desorption) and roughening
of the electrode surface which results from enhanced corrosion [51].

123
1936 I. B. Obot et al.

Fig. 3 Relationship between corrosion rate and MOP concentration at different temperatures

Fig. 4 Variation of inhibition efficiency of MOP with temperature at different concentrations

Effect of temperature

A kinetic model was employed to further explain the inhibition properties of the
inhibitor. The apparent activation energy for the corrosion process was calculated
from the Arrhenius equation [20–22]:
Ea
log q ¼ log A  ; ð17Þ
2:303RT
where q is the corrosion rate Ea is the apparent activation energy, R is the molar gas
constant (8.314 J K-1 mol-1), T is the absolute temperature, and A is the frequency

123
Experimental, quantum calculations, and dynamic simulations 1937

Fig. 5 Arrhenius plot for mild steel corrosion in 0.5 M H2SO4 in absence and presence of different
concentrations of MOP

factor. The plot of log q against 1/T for mild steel corrosion in 0.5 M H2SO4 in the
absence and presence of different concentrations of MOP is presented in Fig. 5. The
activation parameters are presented in Table 2.
The activation energy increased in the presence of MOP, which indicated
physical adsorption (electrostatic) in the first stage [22]. Nevertheless, the
adsorption of an organic molecule is not only a physical or chemical adsorption
phenomenon. A wide spectrum of conditions ranging from the dominance of
chemisorption or electrostatic effects can be seen in other adsorption experimental
data [52]. The activation energy rose with increasing inhibitor concentration,
suggesting strong adsorption of inhibitor molecules at the metal surface. Szauer and
Brand [53] explained that the increase in activation energy can be attributed to an
appreciable decrease in the adsorption of the inhibitor on the mild steel surface with
increase in temperature. However, Vracar and Drazic [54] opined that the criteria of
adsorption type obtained from the change of activation energy cannot be taken as

Table 2 Activation parameters of mild steel dissolution in 0.5 M H2SO4 in absence and presence of
different concentrations of MOP
C (M) Ea (kJ mol-1) DH* (kJ mol-1) DS* (J mol-1 K-1)

Blank 46.03 193.38 259.09


-6
2 9 10 50.28 208.70 291.26
4 9 10-6 56.02 227.65 346.40
6 9 10-6 68.01 272.07 477.18
8 9 10-6 70.12 281.46 501.49
10 9 10-6 78.78 308.45 525.78

123
1938 I. B. Obot et al.

decisive due to competition adsorption with water, whose removal from the metal
surface also requires some activation energy. In other words, the so-called physical
adsorption process may contain chemisorption process simultaneously and vice
versa.
The experimental corrosion rate values obtained from weight loss measurements
for mild steel in 0.5 M H2SO4 in the absence and presence of MOP were used to
further gain insight into the change of enthalpy (DH*) and entropy (DS*) of
activation for the formation of the activation complex in the transition state using
the transition equation [24]
     
RT DS DH 
q¼ exp exp ; ð18Þ
Nh R RT
where q is the corrosion rate, h is Planck’s constant (6.626176 9 10-34 J s), N is
Avogadro’s number (6.02252 9 1023 mol-1), R is the universal gas constant, and
T is the absolute temperature. Figure 6 shows the plot of log q/T versus 1/T for mild
steel corrosion in 0.5 M H2SO4 in the absence and presence of different concen-
trations of MOP. Straight lines were obtained with slope of (DH*/2.303R) and
intercept of [log(R/Nh) ? (DS*/2.303R)], from which the values of DH* and DS*
were computed as listed in Table 2.
It can be seen from Table 2 that the obtained DH* for mild steel in blank solution is
lower than that in the inhibited solution, indicating that the corrosion reaction for the
mild steel in the blank solution needs lower energy to occur when compared with the
inhibited solution [55]. The positive values of DH* indicate that the dissolution of
mild steel is an endothermic process. The entropy of activation DS* was also positive
in the absence and presence of MOP, implying that the rate-determining step for the

Fig. 6 Transition-state plot for mild steel corrosion in 0.5 M H2SO4 in the absence and presence of
different concentrations of MOP

123
Experimental, quantum calculations, and dynamic simulations 1939

activated complex is dissociation rather than association. In other words, the


adsorption process is accompanied by an increase in entropy, which is the driving
force for the adsorption of inhibitor onto the mild steel surface [22, 56].

Thermodynamic consideration using the statistical physics model

The adsorption of an inhibitor species, I, on a metal surface, M, can be represented


by a simplified equation:
M þ I $ M ðI Þads ð19Þ
Let M of the above reaction be the system in the ensemble, and the solvent
containing inhibitor molecules as donor particles be the medium. The complex-
forming process can be regarded as the course of distribution of donor particles to
the system. So, it is justifiable to extend the model of variable number of particles in
statistical physics for the inhibiting process.
According to statistical physics [57], the change of free energy of adsorption
DGoads can be calculated from Eq. (20) as follows [58]:
 
1h DGoads RT ln C
ln ¼  ; ð20Þ
h / /
where C is the concentration of inhibitor particles, u is the distribution modulus, and
h is the surface coverage.
The curve-fitting of data in Table 1 to the statistical model at 303–333 K is
presented in Fig. 7. Good correlation coefficient (R2 [ 0.98) was obtained. / and
DGoads were calculated from the slope and intercept of Eq. (20). All the calculated
parameters are given in Table 3.
It is well known that absolute values of DGoads of the order of 20 kJ mol-1 or
lower indicate physisorption; those of the order of 40 kJ mol-1 or higher involve
charge sharing or a transfer from the inhibitor molecules to the metal surface to

Fig. 7 Application of the statistical physics model to the corrosion protection behavior of MOP

123
1940 I. B. Obot et al.

Table 3 Some parameters from the statistical physics model for mild steel in 0.5 M H2SO4 in the
presence of MOP
Temperature (K) R2 / DGoads (kJ mol-1)

303 0.990 1.90 9 103 -32.75


313 0.987 2.20 9 103 -33.32
323 0.999 2.80 9 103 -34.08
333 0.993 2.40 9 103 -32.75

form a coordinate type of bond [59, 60]. Accordingly, the values of DGoads obtained
in the present study indicate that the adsorption mechanism of MOP on mild steel
involves two types of interaction: chemisorption and physisorption. Indeed, due to
the strong adsorption of water molecules on the surface of mild steel, one may
assume that adsorption occurs first due to the physical forces. The removal of water
molecules from the surface is accompanied by chemical interaction between the
metal surface and the adsorbate, which leads to chemisorption.

Quantum chemical calculations

The major driving force of quantum chemical research is to understand and explain
the inhibitory effects of compounds in molecular terms. Among the quantum
chemical methods for evaluation of corrosion inhibitors, the DFT method has shown
significant promise and appears to be adequate for pointing out the changes in
electronic structures responsible for inhibitory action [61, 62].
Figures 8, 9, 10 and 11 show the optimized geometry, the HOMO density
distribution, the LUMO density distribution, and the Mulliken charge population
analysis plots for MOP (neutral molecule) in aqueous phase obtained with DFT at
the B3LYP/6-31G(d) level of theory. As can be seen from Figs. 9 and 10, the
HOMO and the LUMO are distributed over the entire MOP molecule due to the
presence of lone electron pairs in the four N and one O atoms and the delocalization
of p-electrons in the phenanthroline, oxazole, and pyridine rings in the MOP
molecule. The presence of these adsorption centers can cause flat orientation of
MOP molecules on the surface of steel, thus high degree of surface coverage and
inhibition efficiency is expected for MOP from the theoretical point of view. These
results suggest that the four N and one O atoms are the probable reactive sites for
adsorption on MOP on the metal surface. It has been found that the adsorption of an
inhibitor on metal surface can occur on the basis of donor–acceptor interactions
between the lone electron pairs in heteroatoms together with the p-electrons of
heterocyclic compound and the vacant d-orbital of the metal surface atoms [63].
Figure 11 shows the Mulliken atomic charges calculated for MOP. It has been
reported that, the more negative the atomic charges of the adsorbed center, the more
easily the atom donates its electron to the unoccupied orbital of the metal [64]. It is
clear from Fig. 11 that nitrogen and oxygen as well as some carbons atoms carry
negative charge centers which could offer electrons to the mild steel surface to form
a coordinate bond.

123
Experimental, quantum calculations, and dynamic simulations 1941

Fig. 8 Optimized structure of MOP (ball and stick model)

Fig. 9 The highest occupied molecular orbital (HOMO) density of MOP using DFT at the B3LYP/6-
31G(d) basis set level

123
1942 I. B. Obot et al.

Fig. 10 The lowest unoccupied molecular orbital (LUMO) density of MOP using DFT at the B3LYP/6-
31G(d) basis set level

Fig. 11 Mulliken charge population analysis of MOP using DFT at the B3LYP/6-31G(d) basis set level

123
Experimental, quantum calculations, and dynamic simulations 1943

Table 4 Some molecular


Quantum chemical properties Neutral Protonated
properties of MOP calculated
form form
using DFT at the B3LYP/6-
31G(d) basis set level
Total energy (au) -1025.53 -1025.95
in aqueous phase
EHOMO (eV) -5.86 -9.65
ELUMO (eV) -1.80 -6.09
DE (eV) 4.06 3.56
Dipole moment (D) 4.62 6.84
Molecular weight (amu) 312.33 313.33
Molecular area (Å2) 312.33 313.34
Molecular volume (Å3) 308.62 311.21
Ionization potential (I) (eV) 5.86 9.65
Electron affinity (A) (eV) 1.80 6.09
Electronegativity (v) 3.83 7.87
Hardness (g) 2.03 1.78
Softness (r) 0.49 0.56
Fraction of electrons 0.78 -0.25
transferred (DN)

Higher values of EHOMO are likely to indicate a tendency of the molecule to


donate electrons to appropriate acceptor molecules with low energy or empty
electron orbital. It is evident from Table 4 that MOP has higher EHOMO in the
neutral form than in the protonated form. This means that the electron-donating
ability of MOP is greater in the neutral form. The LUMO energy is directly related
to the electron affinity and characterizes the susceptibility of the molecule towards
attack by nucleophiles. The lower the value of ELUMO, the stronger the electron-
accepting ability of the molecule. It is clear from Table 4 that the protonated form
of MOP exhibits the lowest EHOMO, making the neutral form the most likely form
for the interaction of mild steel with MOP molecule. Low values of the energy gap
(DE) will provide good inhibition efficiencies, because the excitation energy to
remove an electron from the last occupied orbital will be low [65]. A molecule with
a narrow energy gap is more polarizable and is generally associated with high
chemical reactivity and low kinetic stability, being termed a soft molecule [66].
According to Wang et al. [67], adsorption of inhibitor onto a metallic surface occurs
at the part of the molecule which has the greatest softness and lowest hardness. The
result from Table 4 shows that MOP in the protonated form has the lowest energy
gap and lowest hardness. For the dipole moment, higher values of dipole moment
will favor enhancement of corrosion inhibition [68]. It is also clear from Table 4
that the protonated form has higher dipole moment than the neutral form of MOP. If
the organic molecules offer electrons to unoccupied d-orbital of metals and the
electrons are accepted in the d-orbital of metals by using antibonding orbital to form
covalent bond, the adsorption process is chemisorption, and there will be an obvious
correlation between the quantum chemical parameters and the inhibitive effect of
MOP in the neutral or molecular form only. However, it is evident from the
theoretical calculations that no direct correlation can be obtained in the neutral or

123
1944 I. B. Obot et al.

Table 5 Calculated Mulliken atomic charges, Fukui functions, and philicity indices for heteroatoms of
MOP using DFT at the B3LYP/6-31G(d) basis set level
Atom qN qNþ1 qN1 fkþ fk

N1 -0.475 -0.457 -0.489 0.018 0.014


N2 -0.477 -0.461 -0.495 0.016 0.018
N3 -0.558 -0.535 -0.605 0.023 0.047
N4 -0.533 -0.520 -0.569 0.013 0.036
O1 -0.502 -0.477 -0.548 0.025 0.046

protonated form alone. This difficulty indicates the complex nature of the
interactions involved in the corrosion inhibition process [69]. In other words, both
physisorption and chemisorption might take place in the corrosion process, which is
in agreement with the result of the free energy of the adsorption process, DGoads ,
obtained experimentally.
The fraction of electrons transferred from inhibitor to the mild steel surface, DN,
was calculated both in the neutral and in the protonated forms of MOP, as listed in
Table 4. According to Lukovits, if DN \ 3.6, the inhibition efficiency increased
with increasing electron-donating ability at the metal surface [70]. In this study, the
value of DN for MOP was less than 3.6 for both neutral and protonated forms. This
shows that the increase in inhibition efficiency was due solely to the electron-
donating ability of MOP in both forms.
The Fukui indices calculated for the charged species (N ? 1 and N - 1) as well
as the neutral species (N) are presented in Table 5. For simplicity, only the charges
and Fukui functions over the Nitrogen (N) and Oxygen (O) atoms are presented. For
a finite system such as an inhibitor molecule, when the molecule is accepting
electrons, one has fkþ , the index for nucleophilic attack; when the molecule is
donating electrons, one has fk , the index for electrophilic attack. It is possible to
observe from Table 5 that atoms N3, O1, and N4 are the most susceptible sites for
electrophilic attack. These sites present the highest values of fk , being 0.047 for N3,
0.046 for O1, and 0.036 for N4. In the same vein, O1 and N3 are the most
susceptible sites for nucleophilic attack. These sites have the highest values of fkþ ,
being 0.025 for O1 and 0.023 for N3. Similar conclusions have been reported
recently by Feng et al. [71], based on the experimental and theoretical study of
inhibition performance of imidazoline derivative for carbon steel.

MD simulations

To obtain more information on the adsorption behavior of MOP, MD simulation


studies were performed to simulate the adsorption structure of MOP molecule on
mild steel and to further analyze the interactions between MOP molecule and Fe
(1 1 0) surface.
Figure 12 show top and side views of the lowest-energy adsorption model for
single molecules of MOP on the Fe (1 1 0) surface. Solvent and charge effects have
been neglected. As expected from its large molecular size, the inhibitor molecules

123
Experimental, quantum calculations, and dynamic simulations 1945

Fig. 12 a Top and b side views of the lowest-energy adsorption model for a single molecule of 2-(6-
methylpyridin-2-yl)oxazolo[5,4-f][1,10]phenanthroline (MOP) using MD simulation

are preferentially oriented parallel on the Fe surface in order to maximize contact


and increase the surface coverage. This adsorption mode can be attributed to the
strong interaction between the three rings making up MOP, namely phenanthroline,
oxazole, and pyridine rings, and the metal surface. As stated above, N and O atoms
in MOP can offer electrons to the unoccupied d-orbital of iron to form coordinate
bonds, and antibonding orbital of p-electrons in the three rings can accept electrons
from the d-orbital of iron to form feedback bonds [72]. In addition to the chemical
interaction just described, the physical interaction between the inhibitor molecules
and the metal surfaces driven by van der Waals dispersion forces may also
contribute to the net molecule surface attraction, since the strength of physical
interaction generally scales with the size of the molecule [73]. The high negative
adsorption energy of -720.48 kJ mol-1 calculated for the adsorption of MOP on Fe
surface implies that the interaction between MOP molecules and Fe surface is
spontaneous, strong, and stable. Similar reports have been previously documented

123
1946 I. B. Obot et al.

[74, 75]. Thus, the results obtained from MD simulations and quantum chemical
calculations support those obtained experimentally.

Conclusions

1. The investigated inhibitor (MOP) shows excellent performance for inhibition of


mild steel corrosion in 0.5 M H2SO4.
2. The inhibition efficiency increases with increase in the concentration of the
studied inhibitor and with decrease in temperature.
3. The adsorption of MOP on the steel surface obeys the statistical physics
adsorption model. Furthermore, the obtained values of DGoads indicate that the
adsorption of MOP molecules onto the steel surface is a spontaneous process
involving both physisorption and chemisorption mechanisms.
4. Quantum chemical calculation results reveal that the nitrogen and oxygen
atoms of the MOP molecule are the active sites by which the inhibitor
molecules can directly adsorb onto mild steel surface via sharing of electrons
with iron atoms.
5. The MD simulation suggests that MOP molecules adsorb on Fe (1 1 0) surface
in a planar manner with a high negative adsorption energy.
6. Data obtained from experiments along with quantum chemical calculations and
MD simulation results can provide a more complete scheme of the interactions
between MOP and the steel surface.

Acknowledgments The authors are grateful to Dr. A.O. Eseola of the Chemical Sciences Department,
Redeemer’s University, Redemption City, Km. 46 Lagos—Ibadan Expressway, Nigeria for the synthesis
and characterization of the organic inhibitor used in this work.

References

1. J. Liu, W. Yu, J. Zhang, S. Hu, L. You, G. Qjao, Appl. Surf. Sci. 256, 4729 (2010)
2. I.M. Rodriguez-Valdez, A. Martinez-Villfane, D. Mitnik, J. Mol. Struct. Theochem 713, 65 (2005)
3. M.A. Amin, K.F. Khaled, Q. Mohsen, H.A. Arida, Corros. Sci. 52, 1684 (2010)
4. M.M. Solomon, S.A. Umoren, I.I. Udousoro, A.P. Udoh, Corros. Sci. 52, 1317 (2010)
5. S.E. Nataraja, T.V. Venkatesha, K. Manjunatha, B. Poojary, M.K. Pavithra, H.C. Tandon, Corros.
Sci. 53, 2651 (2011)
6. X. Li, S. Deng, H. Fu, T. Li, Electrochim. Acta 54, 4089 (2009)
7. E. Machnikova, K.H. Whitemire, N. Hackerman, Electrochim. Acta 53, 6024 (2008)
8. F. Bentiss, M. Lagrenee, J. Mater. Environ. Sci. 2(1), 13 (2011)
9. A. Kokalj, Electrochim. Acta 56, 745 (2010)
10. S. Ghareba, S. Omanovic, Corros. Sci. 52, 2104 (2010)
11. A.K. Singh, M.A. Quraishi, Corros. Sci. 52, 1373 (2010)
12. W. Yang, R.G. Parr, Proc. Natl. Acad. Sci. USA 82, 6723 (1985)
13. D. Li, S. Zhang, H. Bian, Int. J. Quant. Chem. 110, 1772 (2010)
14. N.O. Eddy, B.I. Ita, Int. J. Quant. Chem. 111, 3456 (2011)
15. I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52, 923 (2010)
16. I.B. Obot, N.O. Obi-Egbedi, Mater. Chem. Phys. 122, 325 (2010)
17. I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52, 282 (2010)
18. I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52, 198 (2010)

123
Experimental, quantum calculations, and dynamic simulations 1947

19. I.B. Obot, N.O. Obi-Egbedi, Colloids Surf. A. Physicochem. Asp. 330, 207 (2008)
20. I.B. Obot, N.O. Obi-Egbedi, Curr. Appl. Phys. 11, 382 (2011)
21. I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, Corros. Sci. 51, 1868 (2009)
22. I.B. Obot, N.O. Obi-Egbedi, A.O. Eseola, Ind. Eng. Chem. Res. 50, 2098 (2011)
23. N.O. Obi-Egbedi, I.B. Obot, Corros. Sci. 53, 282 (2011)
24. I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52, 657 (2010)
25. N.O. Obi-Egbedi, I.B. Obot, M.I. El-Khaiary, S.A. Umoren, E.E. Ebenso, Int. J. Electrochem. Sci. 6,
5649 (2011)
26. K.F. Khaled, Electrochim. Acta 53, 3484 (2008)
27. D. Fischer, A. Curioni, W. Andreoni, Langmuir 19, 3567 (2003)
28. K.F. Khaled, Electrochim. Acta 54, 4345 (2009)
29. S. Xia, M. Qiu, L. Yu, F. Liu, H. Zhao, Corros. Sci. 50, 2021 (2008)
30. L. Tang, L. Yao, C. Kong, W. Yang, Y. Chen, Corros. Sci. 53, 2046 (2011)
31. S. Roy, K.D. Hagen, P.U. Maheswari, M. Lutz, A.L. Spek, J. Reedijk, G.P. Van-Wezel, Chem-
MedChem 3, 1427 (2008)
32. T. Cardinaels, J. Ramaekers, P. Nockemann, K. Driesen, K. Van Hecke, L. Van Meervelt, S. Lei, S.
De Feyter, D. Guillon, B. Donnio, K. Binnemans, Chem. Mater. 20, 1278 (2008)
33. Z. Zu, H. Liu, S. Xiao, M. Yang, X. Bu, J. Inorg. Biochem. 90, 79 (2002)
34. G.N. Mu, X.M. Li, F. Li, Mater. Chem. Phys. 86, 59 (2004)
35. X.M. Li, L.B. Tang, L. Li, G.N. Mu, G.H. Liu, Corros. Sci. 48, 308 (2006)
36. G. Bereket, C. Ogretir, C. Ozsahin, J. Mol. Struct. (Theochem) 663, 39 (2003)
37. Eseola AO Photoluminescent and catalytic properties of azole-substituted pyridine and phenol
ligands and their NiII and ZnII complexes. PhD thesis. University of Ibadan, Ibadan, p. 222 (2010)
38. I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, E.E. Ebenso, Chem. Eng. Commun. 198, 711 (2011)
39. Spartan,06 Wavefunction, Inc. Irvine, CA: Y. Shao, L.F. Molnar, Y. Jung, J. Kussmann, C. Och-
senfeld, S.T. Brown, A.T.B. Gilbert, L.V. Slipehenko, S.V. Levehenko, D.P. O’Neill, R.A. DiStasio
Jr., R.C. Lochan, T.Wang, G.J.O. Beran, N.A. Besley, J.M. Herbert, C.Y. Lin, T. Van Voorhis, S.H.
Chien, A. Sodt, R.P. Steele, V.A. Rassolov, P.E. Maslen, P.P. Korambath, R.D. Adamson, B. Austin,
J. Baker, E.F.C. Byrd, H. Dachsel, R.J. Doerksen, A. Dreuw, B.D. Dunietz, A.D. Dutoi, T.R. Furlani,
S.R. Gwaltney, A. Heyden, S. Hirata, C.P. Hsu, G. Kedziora, R.Z. Khalliulin, P.Klunzinger, A.M.
Lee, M.S. Lee, W.Z. Liang, I.Lotan, N. Nair, B.Peters, E.I. Proynov, P.A. Pieniazek, Y.M. Rhee, J.
Ritchie, E. Rosta, C.D. Sherrill, A.C. Simmonett, J.E. Subotnik, H.L. Woodcock III, W.Zhang, A.T.
Bell, A.K. Chakraborty, D.M. Chipman, F.J. Keil, A. Warshel, W.J. Hehre, H.F. Schaefer, J. Kong,
A.I. krylov, P.M.W. Gill and M. Head-Gordon. Phys. Chem. Chem. Phys. 8, 3172 (2006)
40. R.G. Parr, R.A. Donnelly, M. Levy, W.E. Palke, J. Chem. Phys. 68, 3801 (1978)
41. R.G. Parr, R.G. Pearson, J. Am. Chem. Soc. 105, 7512 (1983)
42. R.G. Pearson, Inorg. Chem. 27, 734 (1988)
43. T.A. Koopmans, Physica 1, 104 (1933)
44. W. Yang, W.J. Mortier, J. Am. Chem. Soc. 108, 5708 (1986)
45. K.F. Khaled, M.A. Amin, Corros. Sci. 51, 2098 (2009)
46. G. Bereket, C. Ogretir, A. Yurt, J. Mol. Struct. (Theochem) 571, 13 (2001)
47. M.M. El-Naggar, Corros. Sci. 49, 2226 (2007)
48. S.A. Umoren, E.E. Ebenso, Mater. Chem. Phys. 106, 387 (2007)
49. A.Y. El-Etre, Corros. Sci. 45, 2485 (2003)
50. M. Lebrini, F. Bentiss, H. Vezin, M. Lagrenee, Corros. Sci. 48, 1291 (2006)
51. S.S. Abd El Rehim, S.M. Sayyah, M.M. El-Deed, S.M. Kamal, R.E. Azooz, Mater. Chem. Phys. 123,
20 (2010)
52. I. Ahamad, R. Prasad, M.A. Quraishi, Mater. Chem. Phys. 124, 1155 (2010)
53. T. Szauer, A. Brand, Electrochim. Acta 26, 1219 (1981)
54. L.M. Vracar, D.M. Drazic, Corros. Sci. 44, 1669 (2002)
55. A.Y. Musa, A.H. Kadhum, A.B. Mohamad, M.F. Takriff, E.P. Chee, Curr. Appl. Phys. 12, 325 (2012)
56. S.A. Umoren, I.B. Obot, I.O. Igwe, Open Corros. J. 2, 1 (2009)
57. L.D. Landau, E.M. Lifshitz, Statistical Physics (Pergamon, Oxford, 1980), p. 107
58. H.L. Wang, H.B. Fan, J.S. Zheng, Mater. Chem. Phys. 77, 655 (2002)
59. S.A. Umoren, I.B. Obot, Surf. Rev. Lett. 15(3), 277 (2008)
60. S.A. Umoren, I.B. Obot, E.E. Ebenso, P.C. Okafor, O. Ogbobe, E.E. Oguzie, Anti-Corros. Methods
Mater. 53(5), 277 (2006)

123
1948 I. B. Obot et al.

61. S. John, B. Joseph, K.V. Balakrishnan, K.K. Aravindakshan, A. Joseph, Mater. Chem. Phys. 123, 218
(2010)
62. H. Ju, Y. Li, Corros. Sci. 49, 4185 (2007)
63. N. Khalil, Electrochim. Acta 48, 2635 (2003)
64. N.O. Obi-Egbedi, I.B. Obot, M.I. El-Khaiary, J. Mol. Struct. 1002, 86 (2011)
65. G. Gece, Corros. Sci. 50, 2981 (2008)
66. A. Dwivedi, N. Misra, Der Pharma Chem. 2(2), 58 (2010)
67. H. Wang, X. Wang, H. Wang, L. Wang, A. Liu, J. Mol. Model. 13, 147 (2007)
68. W. Chen, H.Q. Luo, N.B. Li, Corros. Sci. 53, 3356 (2011)
69. W. Li, Q. He, C. Pei, B. Hou, Electrochim. Acta 52, 6386 (2007)
70. I. Lukovits, E. Lalman, F. Zucchi, Corrosion 57, 3 (2001)
71. L. Feng, H. Yang, F. Wang, Electrochim. Acta (2011). doi:10.1016/j.electacta.2011.09.063
72. H. Ashassi-Sorkhabi, B. Shaabani, D. Seifzadeh, Electrochim. Acta 50, 3446 (2005)
73. D.J. Lavrich, S.M. Wetterer, G. Scoles, J. Phys. Chem. B 102, 3456 (1998)
74. M.K. Awad, M.R. Mustafa, M.M. Abo-Elnga, J. Mol. Struct. (Theochem) 959, 66 (2010)
75. J. Fu, S. Li, L. Cao, Y. Wang, L. Yan, L. Lu, J. Mater. Sci. 45, 979 (2010)

123

Das könnte Ihnen auch gefallen