You are on page 1of 5

Corrosion Science 50 (2008) 15461550

Contents lists available at ScienceDirect

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

Inhibitive action of mangrove tannins and phosphoric acid on pre-rusted steel


via electrochemical methods
Adah A. Rahim a,*, E. Rocca b, J. Steinmetz b, M. Jain Kassim a
a
b

School of Chemical Sciences, University Sains Malaysia, 11800 Penang, Malaysia


Laboratoire de Chimie du Solide Mineral, Universite Henri Poincare, Nancy I, B.P. 239, 54506 Vandoeuvre Les Nancy, France

a r t i c l e

i n f o

Article history:
Received 2 January 2007
Accepted 11 February 2008
Available online 7 March 2008

Keywords:
I. Inhibition efciency
P. Plant extracts
P. Phosphoric acid
T. Tannins

a b s t r a c t
The inhibitive action of mangrove tannins, extracted from mangrove barks and phosphoric acid, on prerusted steel in a 3.5% NaCl solution was evaluated and the inhibitive efciency was compared with that of
mimosa tannins. From the electrochemical studies, the inhibition efciency of solutions containing
3.0 g L1 tannins depended upon the concentration of phosphoric acid added and the pH of the solution.
At pH 0.5 and pH 2.0, inhibition was greatest with mangrove and mimosa tannins alone, while at pH 5.5
the addition of phosphoric acid alone gave the highest inhibition.
2008 Elsevier Ltd. All rights reserved.

1. Introduction
The known hazardous effects of most synthetic corrosion inhibitors are reasons to search for safer and environmentally friendly
natural products. Plant extracts, for example, contain diverse compounds that often can be extracted at low cost. Naturally occurring
substances such as vanillin [1], Opuntia extracts [2], lawsonia extract [3], natural honey [4] and extracts of chamomile, halfabar,
black cumin and kidney bean [5] are plant compounds recently
evaluated as corrosion inhibitors for different metals in various
environments while other studies of plant extracts included the
various tannin extracts [613].
Condensed tannins are a group of phenolic polymers which are
widely distributed in the plant kingdom particularly with woody
growth habit. These compounds consist of avan-3-ol units linked
together through C4C6 or C4C8 bonds. These polyavonoids
present phloroglucinol or resorsinol A-rings and catechol or pyrogallol B-rings (Fig. 1). Mangrove (Rhizophora apiculata) tannins
consist primarily of condensed tannins or proanthocyanidins. HPLC
analyses of condensed tannins from the mangrove following depolymerisation in phloroglucinol and ethanol identied four avonoid monomers namely catechin, epicatechin, epigallocatechin
and epicatechin gallate [14]. Mimosa tannins are also condensed
tannins and according to a comparative C13 NMR study of poly-

* Corresponding author. Fax: +6046574854.


E-mail address: adah@usm.my (A.A. Rahim).
0010-938X,/$ - see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2008.02.013

avonoids, mimosa tannins were found to be predominantly prosenidin/prorobinetidin-type tannins [15].


Tannins as corrosion inhibitors are applied both in solvent and
waterborne pre-treatment formulations. These formulations could
be applied on partially rusted substrates, reducing the effort
needed for cleaning the surface by methods which proved to be
expensive and are not applicable in many situations. The transformation of rusty iron into a blue-black coating layer has been attributed to the interaction of polyphenolic moieties from the tannins
with iron oxides and oxyhydroxides, thereby forming ferrictannate complexes as the major product. Gust and Suwalski [12] have
reported that via Mossbauer spectroscopy, a mixture of mono and
bis complexes were formed as a result of a reaction between rustphase components and oak tannins in aqueous solution. Similar
mixture of mono and bis-type complexes was also observed by
Jaen et al. [16] when reacting several plant extracts of Panama with
ferrous and ferric salts. The view that ferrictannates of dark blue
colour are highly insoluble and acts as a barrier layer are shared by
several authors [11,13,17]. However, the ability of the ferrictannates to protect against further corrosion has not been resolved.
Tannins are thought to be more effective when used in conjunction
with phosphoric acid [11,18], but some investigators think that
even the efciency of this type of pre-treatment is inadequate
[13,19]. The action of tannins and phosphoric acid was reported
to be mainly supercial [19] and dependent on the exposure
conditions [10,1921] and recommended concentrations differ
from one work to another [20,21]. Given the inconsistent results,
electrochemical studies of pre-rusted steel in the presence of

1547

A.A. Rahim et al. / Corrosion Science 50 (2008) 15461550

5
6'

HO 7 8

1
9 O

5 10 4
OH

OH
4'

1' 2' 3' OH

OH

Fig. 1. Phloroglucinol or resorcinol A-ring and catechol or pyrogallol B-ring.

mangrove tannins and phosphoric acid are conducted and compared with that of mimosa tannins.
2. Experimental details
2.1. Tannin isolation
Mangrove bark samples from 15 year old trees were obtained
from the Matang Forest, Malaysia. The barks were dried and grinded to 250 mesh followed by further drying until a constant
weight was obtained. Tannins from mangrove barks were extracted using 70% acetone as described elsewhere [14]. Commercial mimosa tannin powder from bark extracts were obtained
from SILVACHIMICA, Italy and was used without further
purication.

B 117 [22], were immersed in solutions containing 15% (w/v) phosphoric acid, 0.5% (w/v) mangrove tannins and 15% (w/v) phosphoric acid and 0.5% (w/v) mimosa tannins and 15% (w/v) phosphoric
acid at pH 4.0 and for 24 h. The surface morphology was evaluated
by SEM analysis with a SEM S-2500 Hitachi Thermo NORAN
equipped with an energy-dispersive X-ray spectrometer. A Goniometre C diffractometer, incorporating a cobalt radiation

Table 1
Percentage inhibition of 3.0 g L1 mangrove tannins and mimosa tannins with the
addition of 15%, 30% and 50% phosphoric acid in 3.5% NaCl solution
Solution composition

Percentage inhibition (0.5%)

Mangrove tannins
Mangrove tannins + 15% H3PO4
Mangrove tannins + 30% H3PO4
Mangrove tannins + 50% H3PO4
Mimosa tannins
Mimosa tannins + 15% H3PO4
Mimosa tannins + 30% H3PO4
Mimosa tannins + 50% H3PO4
15% H3PO4
30% H3PO4
50% H3PO4
*

pH 0.5

pH 2.0

pH 5.5

77.3
79.0
56.4
59.2
77.4
64.3
50.5
58.5
0.0

57.6
30.7
24.5
26.8
24.1
12.7
0.0
0.0
18.2

38.7
33.1
73.1
54.4
0.0
10.4
60.3
20.4
66.9
90.1
52.6

Denotes negative percentage inhibition.

2.2. Electrochemical studies


Electrochemical tests were carried out in a three-electrode electrochemical cell connected to an EGG Princeton 273A potentiostat. A
circular and horizontal working electrode (2.8 cm2) consisting of a
pre-rusted sample prepared by subjecting the mild steel plates to
the salt spray chamber according to the ASTM B 117 [22] standard
procedure [6 h exposure in 5% (w/v) NaCl at 98% humidity,
1.0 mL h1 spray rate, 1.0 kg cm3 of pressure and followed by drying in an oven at 40 C] was placed at the bottom of the cell. Data
acquisition and the calculation of electrochemical parameters were
made with a 352 Soft Corr software. The reference electrode was a
KCl-saturated calomel electrode (E = 0.241 V/SHE), and all working
electrode potentials were measured against this reference electrode.
The following experiment sequence was used and done in
duplicates:

Fig. 2. Potentiodynamic curves of pre-rusted steel containing mangrove tannins


and phosphoric acid in 3.5% NaCl solution at pH 0.5.

(i) Measurements of the corrosion potential (Ecorr) and the


polarisation resistance (Rp), were performed every 90 min
for 18 h, with a scan rate of 0.166 mV s1 for a range
(Ecorr 20 mV).
(ii) Recording the potentiodynamic curve, i = f(E), from 300 to
300 mV versus Ecorr with a sweep rate of 1 mV s1 after 18 h
of sample immersion.

15% phosphoric acid


250

30% phosphoric acid


50% phosphoric acid

200

Rp /ohms.cm 2

The standard solution consisted of 3.5% (w/v) NaCl and the test
solutions contained 3.0 g L1 mangrove tannins, 3.0 g L1 mimosa
tannins and 15% (w/v), 30% (w/v) and 50% (w/v) phosphoric acid.
In addition, 15% (w/v), 30% (w/v) and 50% (w/v) phosphoric acid
were added to the 3 g L1 tannin solutions. All solutions were prepared in 3.5% (w/v) NaCl solution. The pH of the solution was adjusted to pH 0.5, 2.0 and 5.5 using 2 M H2SO4 and 3 M NaOH
solutions. All chemicals were AR grade.

NaCl

150

100

50

2.3. Surface analysis


Pre-rusted coupons (2.5 cm  3.5 cm) that had been prepared
by subjecting them to a salt spray chamber according to the ASTM

10

15

20

t /hr
Fig. 3. Variation of polarisation resistance, Rp with time of immersion containing
15%, 30% and 50% phosphoric acid in 3.5% NaCl solution at pH 2.0.

1548

A.A. Rahim et al. / Corrosion Science 50 (2008) 15461550

(k = 1.78892 ) was used to obtain the X-ray diffraction patterns of


the deposits.
3. Results
3.1. Electrochemical studies
The percentage inhibition from the polarisation resistance measurements (Table 1) was calculated as

Fig. 4. Potentiodynamic curves of pre-rusted steel containing 3.0 g L1 mangrove


tannins and phosphoric acid in 3.5% NaCl at pH 5.5.

Fig. 5. Potentiodynamic curves of pre-rusted steel containing 15%, 30%, and 50%
phosphoric acid in 3.5% NaCl solution at pH 5.5.

IE%

Rp  Rp0
 100
Rp

where Rp 0 and Rp are the uninhibited and inhibited polarisation values, respectively, assuming that bA and bc are unmodied by inhibition, performed at pH 0.5, 2.0 and 5.5 on pre-rusted steels in a 3.5%
NaCl solution. The inhibition efciency of mangrove and mimosa
tannins on pre-rusted steels was found to decrease as the pH increased, a result similar to the trend previously observed for the
inhibition performance of tannins on clean steels [23]. Conversely,
the inhibition efciency increased with increasing pH when phosphoric acid was used alone.
At pH 0.5, an inhibition of almost 80% was achieved when mangrove and mimosa tannins were used. The corrosion protection
provided by both tannins however did not increase with the addition of 15% phosphoric acid. Inhibition efciency was further reduced with increasing phosphoric acid concentrations. From the
potentiodynamic curves of the pre-rusted steels, modications in
the anodic and cathodic curves were observed and the Ecorr also
shifted to more positive values with either the addition of mangrove tannins alone or the addition of mangrove tannins and phosphoric acid (Fig. 2).
At pH 2.0, the mangrove tannins provided a better protection
(57% inhibition) than the mimosa tannins (24% inhibition). The
addition of 15%, 30% and 50% phosphoric acid to the mangrove tannins reduced the inhibition efciency by almost 50%. The addition
of 15% phosphoric acid to mimosa tannin also reduced the inhibition efciency further. The increase in phosphoric acid concentrations to 30% and 50% resulted in essentially no inhibition. This
change was not a surprise since the Rp values decreased with increases in the concentration of phosphoric acid and may have
accelerated corrosion after 18 h of immersion when 30% and 50%
phosphoric acid were used (Fig. 3).
When the inhibition studies were conducted at pH 5.5, which is
the natural pH of a 3.5% NaCl solution, Ecorr initially became more
negative in the presence of mangrove tannins. With the addition
of 15%, 30% and 50% phosphoric acid to the mangrove tannins the
Ecorr gradually shifted to more positive values (Fig. 4). The addition
of both mangrove tannins and phosphoric acid had the effect of
lowering both the cathodic and anodic curves. The addition of
30% phosphoric acid had a greater effect on the anodic reaction than
did the addition of 50% phosphoric acid. This observation is consistent with the results obtained from the potentiodynamic curves of
steels containing phosphoric acid alone (Fig. 5). Anodic shifts were
seen evident for all phosphoric acid concentrations, with the addition of 30% phosphoric acid resulting in the greatest anodic shift to
lower density current and consequently the greatest inhibition. The
use of phosphoric acid alone is favoured at pH 5.5, since the use of
30% phosphoric acid resulted in 90% inhibition (Table 1). Similarly,
the addition of 30% phosphoric acid to mimosa tannin increased the

Fig. 6. XRD pattern of bare rust surface immersed in 0.5% (w/v) mangrove tannins and 15% (w/v) phosphoric acid solutions: L lepidocrocite, M magnetite, V vivianite.

A.A. Rahim et al. / Corrosion Science 50 (2008) 15461550

1549

amount of inhibition relative to the use of mimosa tannin alone. No


inhibition was observed for mimosa tannin alone at pH 5.5, indicating that mangrove tannins are preferred at this pH.
3.2. Surface analysis
Surface observations of pre-rusted samples after polarisation
revealed a blue-black deposit typical of ferrictannates if mangrove tannins and phosphoric acid were present at pH 2.0, whereas
a mixture of blue-black, bluish-white and yellowish deposits were
observed at pH 5.5. Similar deposits also were observed on prerusted samples immersed in tannin and phosphoric solutions.
The XRD of the pre-rusted sample revealed lepidocrocite and magnetite as its main components, the blue black ferrictannates were
amorphous and the white-bluish colour of the deposit, characteristic of iron phosphate was identied as vivianite, Fe3(PO4)2  8H2O
(Fig. 6). The yellowish ferric-phosphates were amorphous, consistent with the observations of Kandori et al. [24]. SEM micrographs
of tannins and phosphoric acid treated samples had some interesting features (Fig. 7). Samples treated in phosphoric acid had stacks
of well-ordered facetted akes of phosphates amidst irregularly
shaped cracks. When phosphoric acid was added to mangrove tannins, the akes were replaced by a blend of tannates and phosphates with the cracks still visible. In solutions with mimosa
tannins and phosphoric acid, the morphology was transformed
into spectacular structures of sea-shells scattered amidst the
cracked structures of tannates.

4. Discussion
In NaCl solution, ferrous chloride and sodium hydroxide are initially deposited at the anode and cathode, respectively. The ferrous
chloride formed often is transformed into ferrous hydroxide, which
in turn combines with oxygen to form rust according to the following reactions [25]:

Fig. 7. SEM micrographs of pre-rusted plates containing: (a) 15% H3PO4; (b) 15%
H3PO4 and 0.5% mangrove tannins and (c) 15% H3PO4 and 0.5% mimosa tannins.

FeCl2 2NaOH ! FeOH2 2NaCl

4FeOH2 O2 ! 4FeOOH 2H2 O

At pH 0.5 the inhibition mechanism probably is associated with the


chemisorption of tannin molecules onto the pre-rusted surface as
shown in our previous studies [23]. Meanwhile at pH 2.0, the inhibition was achieved through the formation of ferrictannates as a
result of the reaction between tannins and the Fe ions available,
which is discussed in more detail elsewhere [23]. Consequently
the formation of rust in reaction Eq. (3) is avoided.
The successful use of phosphoric acid alone depended on the
concentration of phosphoric acid and the pH of the solution. The
efciency increased with increasing pH for 15% phosphoric acid,
while inhibition occurred only at pH 5.5 for 30% and 50% phosphoric acid. Using 30% phosphoric acid alone is certainly favoured at
pH 5.5 when 90% inhibition was achieved. According to Almeida
et al. [20], the point of initiation of insoluble phosphates when
the metal is iron occurs at pH 56. Thus in this study, phosphate
formation was maximum with 30% phosphoric acid as observed
from the very thick deposits after the polarisation process. A passivation plateau, which was induced by the reaction between phosphoric acid and rust forming phosphates, subsequently retards the
anodic dissolution reaction. Increasing the phosphoric acid to 50%
may have resulted in the formation of soluble complexes, leading
to the reduction in the inhibition efciency.
The mechanisms of reactions of phosphoric acid with the most
common iron phases forming phosphates have been reported in
several works[10,1921,26]. Infrared spectroscopy has shown that
lepidocrocite reacted most rapidly with phosphoric acid followed
by magnetite and nally goethite [10,19,21] while the formation

1550

A.A. Rahim et al. / Corrosion Science 50 (2008) 15461550

of ferric phosphates was found to be dependent on the aggressiveness of the solution used [26]. It has also been shown that different
phases of phosphates are formed depending on the surface conditions, exposure time and concentrations of phosphoric acid. Solutions containing 15 and 33% phosphoric acid seem to lead to the
formation of vivianite on rust layers rich in lepidocrocite between
1 and 10 months exposure at natural atmosphere and after 24
months when the main components were lepidocrocite and goethite, precipitation of phosphates was not observed [20].
Based on the data of the surface analysis of the deposits via XRD
and SEM, the phosphoric acid initially dissolves the rust and the
following phosphate precipitation reaction is proposed:
H3 PO4 FeOOH ! FePO4 2H2 O
and
2H3 PO4 3Fe3 O4 5H2 O ! Fe3 PO4 2  8H2 O 3Fe2 O3
3Fe2 O3 3H2 O ! 6FeOOH
2H3 PO4 3Fe3 O4 8H2 O ! Fe3 PO4 2  8H2 Ovivianite 6FeOOH
The joint effect of both the tannins and the phosphoric acid was
unfavourable at this pH due to the competing reactions between
mangrove tannins and the phosphoric acid with the pre-rusted
steel. Gust [10] and Nasrazadani [21] reported that the reaction
kinetics of tannins on the various iron phases was slower to that
of phosphoric acid and consequently more phosphates are formed
as compared to ferrictannates. It is proposed that the phosphoric
acid initially dissolves the rust and ferrictannates are formed
according the mechanism proposed in our previous work [23]. In
addition to the phosphates formation as described above, the reducing ability of mangrove tannins to reduce Fe3+ to Fe2+ ions [14] facilitates the precipitation of phosphates. However, the inhibition
could not exceed that of phosphoric acid alone.
5. Conclusion
The inhibition efciency of pre-rusted steel in 3.5% NaCl solution containing 3 g L1 mangrove tannins depended on the concen-

tration of phosphoric acid and the pH of solution. At pH 0.5 and pH


2.0, inhibition was best with mangrove tannins alone. At pH 5.5,
the inhibition efciency increased when the phosphoric acid was
added to the mangrove tannins. In fact the use of phosphoric acid
alone is favoured at this pH.
References
[1] A.Y. El-Etre, Corrosion Science 43 (2001) 1031.
[2] A.Y. El-Etre, Corrosion Science 45 (2003) 2485.
[3] A.Y. El-Etre, M. Abdullah, Z.E. El-Tantawy, Corrosion Science 47 (2005)
385.
[4] A.M. Abdel-Gaber, B.A. Abd-El-Nabey, I.M. Sidahmed, A.M. El-Zayady, M.
Saadawy, Corrosion Science 48 (9) (2006) 2765.
[5] M. Morcillo, S. Feliu, J. Simancas, J.M. Bastidas, J.C. Galvan, S. Feliu Jr., E.M.
Almeida, Corrosion NACE 48 (12) (1992) 1032.
[6] S. Martinez, I. Stern, Journal of Applied Electrochemistry 31 (2001) 973.
[7] S. Martinez, I. Stern, Applied Surface Science 199 (2002) 83.
[8] S. Martinez, Materials Chemistry and Physics 77 (2002) 97.
[9] T.K. Ross, R.A. Francis, Corrosion Science 18 (1978) 351.
[10] J. Gust, Corrosion 47 (1991) 453.
[11] J. Gust, J. Bobrowicz, Corrosion 49 (1993) 24.
[12] J. Gust, J. Suwalski, Corrosion Science 50 (5) (1994) 355.
[13] O.R. Pardini, J.I. Amalvy, A.R. Di Sarli, R. Romagnoli, V.F. Vetere, Journal of
Coating and Technology 73 (2001) 99.
[14] Adah A. Rahim, E. Rocca, J. Steinmetz, M.J. Kassim, M. Sani Ibrahim, H. Osman,
Food Chemistry 107 (2008) 200.
[15] T. Pizzi, A. Stephanou, Applied Polymer Science 50 (1993) 2105.
[16] J.A. Jaen, E. Garcia de Saldana, C. Hernandez, Hyperne Interactions 122 (1999)
139.
[17] A.J. Seavell, Journal of Oil Colour Chemistry Association 8 (1992) 293.
[18] G. Matamala, W. Smeltzer, G. Droguett, Corrosion NACE 50 (1994) 270.
[19] C.A. Barrero, L.M. Ocampo, C.E. Arroyave, Corrosion Science 43 (2001)
1003.
[20] E. Almeida, D. Pereira, M.O. Figueiredo, V.M.M. Lobo, M. Morcillo, Corrosion
Science 39 (9) (1997) 1561.
[21] S. Nasrazadani, Corrosion Science 39 (1011) (1997) 1845.
[22] ASTM B 117. Salt spray resistance test. American Society for testing and
materials, Philadelphia, 1973.
[23] Adah A. Rahim, E. Rocca, J. Steinmetz, M.J. Kassim, R. Adnan, M. Sani Ibrahim,
Corrosion Science 49 (2007) 402.
[24] K. Kondari, T. Kuwae, T. Ishikawa, Journal of Colloid and Interface Science 300
(2006) 225.
[25] L.L. Shreir (Ed.), in: Corrosion, Vol. 2, Corrosion Control, second ed., NewnesButterworth, London, 1978.
[26] L.M. Ocampo, I.C.P. Margarit, O.R. Mattos, S.I. Cordoba-de-Torresi, F.L. Fragata,
Corrosion Science 46 (2004) 1515.

You might also like