Você está na página 1de 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/248209067

Performance of rust converter based in phosphoric and


tannic acids

Article  in  Corrosion Science · June 2004


DOI: 10.1016/j.corsci.2003.09.021

CITATIONS READS

24 1,175

5 authors, including:

Luz Marina Ocampo Carmona Isabel Margarit-Mattos


National University of Colombia Federal University of Rio de Janeiro
10 PUBLICATIONS   79 CITATIONS    48 PUBLICATIONS   781 CITATIONS   

SEE PROFILE SEE PROFILE

O.R. Mattos Susana I. Córdoba de Torresi


Federal University of Rio de Janeiro University of São Paulo
154 PUBLICATIONS   2,425 CITATIONS    231 PUBLICATIONS   4,103 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Laboratory test program to evaluate corrosion resistance of flexible pipe carcass materials at sour service View project

Iron Electrodissolution under Mass transport Control View project

All content following this page was uploaded by Luz Marina Ocampo Carmona on 05 March 2018.

The user has requested enhancement of the downloaded file.


Corrosion Science 46 (2004) 1515–1525
www.elsevier.com/locate/corsci

Performance of rust converter based


in phosphoric and tannic acids
L.M. Ocampo a, I.C.P. Margarit a,b,*, O.R. Mattos a,
ordoba-de-Torresi c, F.L. Fragata d
S.I. C
a
Federal University of Rio de Janeiro, EE/COPPE/PEMM, P.O. Box 68505,
CEP: 21941-972 Rio de Janeiro, Brazil
b
Inor. Proc. Dep., Federal University of Rio de Janeiro, EQ, Rio de Janeiro, Brazil
c
Institute of Chemistry, S~ao Paulo University, P.O. Box 26077, 05513970 S~ao Paulo, Brazil
d
Eletrobras Research Center, CEPEL, P.O. Box 68007, CEP: 21944-970 Rio de Janeiro, Brazil
Received 14 May 2003; accepted 18 September 2003

Abstract

The role of a rust converter with tannic and phosphoric acids is evaluated for painted and
unpainted steel with different corrosion degrees and salts contaminating the rust. The per-
formance of unpainted samples is monitored with impedance and characterized by X-ray,
infrared and Raman spectroscopies. Results show long-term action of the converter. For
painted samples, loss of adhesion and corrosion spread around a scratch are measured after
alternated immersion corrosion test. For rusted samples with high chloride contamination, the
converter is harmful. For rusted samples with lower contamination, the influence of the
converter on the performance of the painting is not detected.
 2003 Elsevier Ltd. All rights reserved.

Keywords: C. Rust; A. Acid solutions; C. Atmospheric corrosion; C. Paint coatings; B. EIS

1. Introduction

Among the surface cleaning techniques, abrasive blasting offers the best level
of oxide removal and leaves the surface with roughness suitable for subsequent

*
Corresponding author. Address: Inor. Proc. Dep., Federal University of Rio de Janeiro, EQ, Rio de
Janeiro, Brazil. Tel.: +55-21-2562-8550; fax: +55-21-2290-6626.
E-mail address: margarit@metalmat.ufrj.br (I.C.P. Margarit).

0010-938X/$ - see front matter  2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2003.09.021
1516 L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525

anticorrosive painting. Nevertheless, this technique cannot be always applied due to


location of the equipment to be treated, its geometry or, by worker health com-
mitment. As an alternative to overcome these difficulties, rust converters are com-
mercially available. The main function of a rust converter is to react with iron oxides
that cannot be completely removed from the surface, leading to a layer where
painting systems can be applied on. Among the rust converters, special attention can
be driven to the ones based on tannins and phosphoric acid. Such products have low
toxicity and tannic acid is obtained from renewable sources [1–4]. The results on
protection efficiency of rust converters are controversial due to several factors as:
tannin type [5], tannin concentration [6], presence of other products [7], pH [8],
reaction time of the converters with the rust [9], application method [10], corrosion
and contamination grades of the rusted steel [11]. These last two aspects are studied
in this work. Actually, this work has two purposes. One of them is to characterize the
importance of washing the converter after application. The other purpose is to study
the influence of corrosion and rust contamination degree with chloride and sulfate
salts on the performance of painted samples and rust conversion itself. These aspects
are scarcely approached in literature.

2. Experimental

Samples of mild steel with 7 · 8 cm2 were degreased and sandblasted to white
metal (ASa3) according to standard ISO 8501-1:1998 (Preparation of steel substrates
before application of paints and related products––visual assessment cleanliness––
Part 1: Rust grades and preparation grades of uncoated steel substrates and steel
substrates after overall removal of previous coating). Then they were pre-corroded
during 50 days in alternated immersion test (AIT). The cycle consisted of 10 min
immersed and 50 min drying at about 35 C. Four solutions were employed in the
AIT: distilled water, Na2 SO4 104 M, NaCl 103 M and NaCl 0.6 M. According to
classification of standard ISO 8501-1:1998, the corrosion of the steel samples in NaCl
0.6 M was D (severe corrosion with many pits). In the other solutions it was C. The
contamination of the rusts with chloride and sulfate was determined by ion chro-
matography following procedure previously employed by Mayne [12]. After the pre-
corrosion stage, the samples were hand-cleaned by wire-brushing and the final
preparation grades were DSt3 for samples withdrawn from NaCl 0.6 M and CSt3 for
the other solutions.
The rust converter employed has the following composition (%w/w): phosphoric
acid 35, tannic acid 5, isopropyl alcohol 12.5, terbutylic acid 12.5, glycerin 10 and
distilled water 25 [1]. The rust converter was applied on the brushed samples in two
layers elapsed by 6 h. A total of 48 samples were used. Twenty-four samples were
washed 10 days later to remove residual converter. Such a long period is not feasible
in the industrial practice, but in this work it is given to guarantee enough time for the
rust conversion. Twelve washed samples and 12 no-washed samples returned to
the AIT during 50 days in order to characterize residual action of the converter on
the composition and morphology of corrosion products.
L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525 1517

The other 24 samples (washed and no-washed) were coated with a surface tolerant
epoxy-mastic paint. The paint was applied by brush in two coats. The total dry
thickness was about 130 lm and a 4 cm scratch was introduced at the bottom part of
each sample. The painted samples were submitted to AIT during 125 days in the
same solutions employed for their pre-corrosion. The cycle for the painted samples
was 16 h immersed and 8 h drying at the room environment. The change on the AIT
cycle was needed because 10 min were not enough to permeate the painting layer,
and 50 min were not sufficient to dry it. The performance of all these samples was
compared with samples without rust converter. At the end of the corrosion test, the
samples were evaluated according to loss of adherence and corrosion spread around
the scratch.
During the AIT, open circuit potential and electrochemical impedance measure-
ments monitored the behaviour of the no-painted samples. The experimental set-up
consisted of a Solartron 1255 connected to an Omnimetra Potentiostat PG09. All the
electrochemical measurements were performed in triplicate samples in a classical
three electrodes cell. Two-identical electrodes in which fewer problems in high fre-
quency range are expected checked the 3-electrodes measured impedances.
The morphology of the converted and no-converted rust layers was evaluated in
cross-cuts of the samples after the AIT. A stereoscopy microscopy Olympus B071
adapted to a digital camera Sony DXC-151A was employed.
At the end of the AIT test, the rust of the samples was removed by deep me-
chanical scraping, homogenized, and analysed by X-ray diffraction (XRD), Fourier
transformed infrared (FTIR) and Raman spectroscopies. XRD measurements were
performed on a HZG4 (Rich Seifert Freiberger Prazisionsmechanik) diffractometer
equipped with Cu(ka) radiation (40 kV/40 mA) and Ni filter. The scans were in the
range of 10–90 ð2hÞ, at 0.05/s and counting until 500 cps. FTIR spectra were
collected in Nicolet Magna-IR 760 spectrometer operating in the transmission mode
between 4000 and 155 cm1 . KBr pellets in a ratio 100/2 were employed. Raman
spectroscopy of cross-cuts of the samples was performed in a Renishaw Raman
imaging microscope (System 3000) according to procedure previously established for
the analysis of iron oxyhydroxides [13].

3. Results and discussion

The surface preparation grades of the samples and rust contamination before
applying the rust converter are shown in Table 1. The underlined values are high
contamination levels for applying painting systems as considered by some authors
[14–16].
The visual aspect of the no-painted samples after the AIT is in Fig. 1. Never mind
the testing solution, the samples with rust converter have more homogeneous aspect
than the no-converted ones.
Micrographies of cross-cuts of the samples show that converted layers are darkish
than no-converted samples, although their thickness are practically the same, Fig. 2.
1518 L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525

Table 1
Surface preparation grades and contamination degrees on the pre-rusted samples before applying the rust
converter
Solution Surface preparation Contamination Degree (lg cm2 )
gradesa Cl SO@
4

Distilled water CSt3 1.2 1.3


Na2 SO4 104 M CSt3 1.6 20:6
NaCl 103 M CSt3 8:9 1.5
NaCl 0.6 M DSt3 3727 0.3
a
According to standard ISO 8501-1:1998.

Fig. 1. Visual aspect of the no-painted samples after 50 days in alternated immersion tests.

For the samples immersed in distilled water, XRD and FTIR spectra in Fig. 3A
and B show that the rust is composed mainly by crystalline a-FeOOH (goethite, G),
c-FeOOH (lepidocrocite, L) and Fe3 O4 (magnetite, M). For the samples tested with
rust converter the a-FeOOH, c-FeOOH, Fe3 O4 and c-Fe2 O3 phases decrease in fa-
vour of the phosphates (ferric phosphate, FP) formation which is shown by the
FTIR measurements, Fig. 3B. These results were observed for all samples tested in
the different solutions, as seen in Figs. 4–6. Such effects become less significant as
the aggressiveness of the solutions increases, i.e., distilled water < Na2 SO4 104
M < NaCl 103 M < NaCl 0.6 M.
Literature points out that the main products of rust conversion are amorphous
iron phosphates and tannates. Lower quantities of tannates than phosphates are
expected due to slower reaction kinetics [17]. Tannates could not be found with the
techniques discussed above. The broad bands of phosphates in the infrared region
hide the tannates bands. The presence of tannin compounds was only detected on the
rust layers by Raman spectroscopy as shown in Fig. 7. Actually, the tannic acid
is fluorescent under Raman radiation (Fig. 7A). However, on Fig. 7B and C it is
L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525 1519

Fig. 2. Micrographies of the samples with and without rust converter, after 50 days of AIT in the sulfate
and chloride solutions.

L L
G L Without converter L L
% Transmittance

M Without converter G G
L G
L+M L L G
M G L L
L+M L M M L LL L
L
CPS

L
L L

FP FP
With Converter FP FP
FP
With converter

20 40 60 80 4000 3200 2400 1600 800


(A) 2θ (B) Wave number (cm-1)

Fig. 3. XRD (A), FTIR (B) of samples tested in distilled water (L: lepdocrocite, M: magnetite, G: goe-
thite, FP: ferric phosphate).
1520 L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525

L Without converter L L
GG
L

% Trasmittance
L G
G L L G
G
L L+M
L Without converter
M G L+M M LL L
CPS

L ML L
L
L L
FP
With converter FP FP
FP
With converter FP

20 40 60 80 4000 3200 2400 1600 800


(A) 2θ (B) Wave number (cm-1)

Fig. 4. XRD (A), FTIR (B) of samples tested in Na2 SO4 104 M (L: lepdocrocite, M: magnetite,
G: goethite, FP: ferric phosphate).

L L G
L Without converter L G
G Without converter G
G L L+M
M M L % Transmitance L
G
L L+M L L
L L L
L
L
L
CPS

FP FP FP
With converter FP FP

With converter

20 40 60 80 4000 3200 2400 1600 800


(A) 2θ (B) Wave number (cm-1)

Fig. 5. XRD (A), FTIR (B) of samples tested in NaCl 103 M (L: lepdocrocite, M: magnetite, G: goethite,
FP: ferric phosphate).

L
Without converter
% Transmittance

M L L
G M L
L M L G
G L+M L+M L L L L
L L
CPS

L L
Without converter L
G G
G FP
With converter M L M
FP
M
M
With converter
20 40 60 80 4000 3200 2400 1600 800
(A) 2θ (B) Wave number (cm-1)

Fig. 6. XRD (A), FTIR (B) of samples tested in NaCl 0.6 M (L: lepdocrocite, M: magnetite, G: goethite,
FP: ferric phosphate).

possible to note at the end of AIT, that compounds derived from the reaction of
tannic acid and iron compounds still remain in the rust layer as attested by the
appearance of bands in a region typical of C–H bonds.
In Fig. 8 there are impedance diagrams for samples tested by AIT in distilled
water with and without rust conversion. The samples without conversion have two
capacitive loops at high frequency range, one loop identified with 40 kHz and an-
other at 272 Hz. At lower frequency range, there is an inductive loop at 4 Hz and
L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525 1521

NaCl 0.6M Phosphate C-H linkings Na2SO4 10-4M Phosphate C-H linkings
(c) (c)
Tannic Acid

(b)
(b)
NaH2PO4.H2O (a)
(a)

500 1000 1500 2000 500 1000 1500 2000 500 1000 1500 2000
(A) Wave number (cm-1) (B) Wave number (cm-1) (C) Wave number (cm-1)

Fig. 7. Raman spectra for: (A) standards of tannic acid and phosphate; (B) rust raised during AIT in
NaCl 0.6 M and (C) rust raised during AIT in Na2 SO4 104 M. The curves identified with low case letters
corresponds to samples: (a) without conversion; (b) just after conversion; (c) after conversion and AIT.

Without converter With converter


2 15 days
15 days
2 0.09Hz
40kHz
1
1 4mHz
0.09Hz 272Hz
40kHz 272Hz 0
0 6Hz
4Hz
-Imag ( )

-1
3 4 5 6 7 8 9 6 7 8 9 10 11 12
9
6 45 days 45 days 4mHz
4mHz
4 6

2 40kHz 272Hz
0.09Hz 3 40kHz
0 272Hz 0.09Hz
4Hz 6Hz
0
3 6 9 12 15 18 3 6 9 12 15 18 21

Real ( )

Fig. 8. Impedance measurements during corrosion test in distilled water.

another capacitive loop in low frequencies without definite limit. Increasing testing
time, the inductive loop diminishes, tending to disappear and the impedances in-
crease about twice. For the converted samples, the main difference can be noted at 15
days testing time. The capacitive loop at lower frequencies has a defined limit at
about 10 X. Along the testing time, it can be seen that converted samples have higher
impedances than no-converted ones.
In Fig. 9 there are impedance diagrams obtained for AIT samples tested in
Na2 SO4 104 M. Once again it can be noted that the main differences on the dia-
grams between converted and no-converted samples are: (i) on the 15th day mea-
surement, converted samples present the low frequency capacitive loop well defined
and (ii) along the testing time, the converted samples have higher impedances than
no-converted ones. These higher impedances suggest a time-standing sealing effect
promoted by the amorphous conversion compounds.
In Fig. 10 there are impedance diagrams obtained for AIT samples tested in NaCl
103 M. The no-converted samples have a cyclic behaviour. At 15 days the diagrams
are characterized by two capacitive loops, followed by an inductive and another
capacitive loops. With 30 days the inductive loop disappears and after 45 days the
1522 L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525

Without converter With converter


8 8
15 days 15 days
4mHz
4 4
40kHz 4mHz
40kHz
0.19Hz
0 0.19Hz 0
27.2Hz 27.2Hz
-Imag (Ω)

0 4 8 12 16 20 0 4 8 12 16 20

45 days 45 days
8 4mHz
4mHz 10
6kHz
4
40kHz
0.19Hz
0.19Hz 0
0
27.2Hz
4 8 12 16 20 24 0 10 20 30 40

Real (Ω)

Fig. 9. Impedance measurements during corrosion test in Na2 SO4 104 M.

Without converter With converter


8 15 days 24 15 days

0.019Hz 16
4
40kHz
8 40kHz
2.7Hz 0.019Hz
0 2.7Hz
0
8 12 16 20 24 28 8 16 24 32 40 48 56
4mHz
30 days 10 30 days
8
-Imag (Ω)

4mHz
4 5 40kHz
40kHz 0.019Hz
0.019Hz 0
0
2 . 7Hz 2.7Hz
4 8 12 16 20 24 0 5 10 15 20 25 30

8 45 days 45 days
20

4 0.019Hz
40kHz 10 40kHz
2.7Hz 0.019Hz
0 2.7Hz
0
8 12 16 20 24 28 10 20 30 40 50 60
Real (Ω)

Fig. 10. Impedance measurements during corrosion test in NaCl 103 M.

diagram is almost the same measured with 15 days suggesting renewal of the reactive
area. In this case, the influence of the rust converter also implies in higher impedances.
However, such influence is not very clearly detected on the shape of the diagrams. In
effect, after 30 days the converted and no-converted samples have very similar dia-
grams. This fact may be related with the higher aggressiveness of the chloride solution
in comparison with distilled water or sulphate solution. This hypothesis is corrobo-
rated by the behaviour of the samples tested in NaCl 0.6 M as shown in Fig. 11.
L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525 1523

3 .0
40kHz 15 days 30 days 40kHz 45 days
1
0.019Hz 2
-Imag ( )

40kHz .5
0.019Hz
0 2.7Hz
1
0.019Hz .0
2.7Hz 2.7Hz
-1 0
8 9 10 11 12 13 9 10 11 12 13 14 15 7.0 7.5 8.0 8.5 9.0 9.5
Real ( )

Fig. 11. Impedance measurements during corrosion test in NaCl 0.6 M.

The reproducibility of the electrochemical measurements among the tested sam-


ples in NaCl 0.6 M was not so good as in the previous solutions. The rust layer was
very thick and fragile. The sequence of diagrams presented in Fig. 11 suggests
constant renewal of the reactive surface. The diagrams are very similar to the ones in
Fig. 10, but the impedance values are still lower in spite of the thicker rust layer (see
Fig. 2). No clear distinction between converted and no-converted samples could be
detected by impedance measurements.
The spectroscopic analysis of the rust layers and electrochemical impedance
measurements performed during AIT show that the kinetic difference between the
two rusts (converted and not converted) tends to disappear with the testing time. The
important point is that the barrier properties improved by converted layers are not
very significant and the total impedance is maintained in the order of 10 X. This
aspect has been already mentioned by others [6,7,18]. It is noteworthy that spec-
troscopic analysis proved that conversion reactions occurred on samples pre-
corroded in all solutions, including NaCl 0.6 M. However, in this last case, the effect
of conversion products on the corrosion process was so meaningless that could not
be clearly detected by impedance. The main question now is to know whether the
rust converter improves or not the performance of painted systems.
The aspects discussed in Figs. 1–11 apply both for washed and no-washed sam-
ples. The importance of washing residual rust converter lives mainly on the per-
formance of a painting layer as shown in Fig. 12.
The loss of adhesion was measured by the tape test on a cross-section at the right
side of the samples. Representative results are presented for the two extreme cases:
distilled water (no contamination of the surface with soluble salts) and NaCl 0.6 M
(the highest contamination with chloride). It can be seen that no-washed samples
have more important adhesion loss than the others. The same happened for samples
tested in Na2 SO4 104 M and NaCl 103 M. Therefore, washing the residual rust
converter would be an important procedure.
In respect to the contamination of the surface, the washed samples tested in NaCl
0.6 M had more important adhesion loss than no-converted samples. In the case
of samples tested in distilled water, as well as Na2 SO4 104 M and NaCl 103 M
(not shown), no significant difference on performance could be detected between
converted + washed samples and samples without conversion. The corrosion advance
around the scratches (at the left side of the samples) could not be evaluated because
the painting system was so adhered that could not be removed.
1524 L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525

Fig. 12. Visual aspect of painted samples after 125 days in AIT.

4. Conclusions

The conversion of rusts formed in alternated immersion tests succeeded with a


formula based on tannic and phosphoric acids as proven by spectroscopic analysis.
In effect, samples corroded until grades C and D in media containing sulfate and
chloride ions in different contents had the iron oxyhydroxides converted mainly to
iron phosphates and tannates. In spite of this, no significant difference on the per-
formance of converted and no-converted painted samples was verified which would
justify the use of a rust converter. With the epoxy-mastic painting, no difference
could be detected even around a scratch. In the case of high chloride contamination
and grade D of corrosion, the presence of the rust converter was even harmful to the
painting performance.

Acknowledgements

To FAPERJ, FUJB, FINEP, CAPES and CNPq for financial support. To Lio-
nice Bezerra from DQI/UFRJ for FTIR measurements. Authors are indebted to the
Lab. of Molecular Spectroscopy (IQ/USP) for Raman facilities and to CBPF to
XRD analysis.

References

[1] C.A. Barrero, L.M. Ocampo, C.E. Arroyave, Corros. Sci. 43 (2001) 1003.
[2] J. Gust, Corros. NACE 47 (6) (1991) 453.
[3] G. Matamala, W. Smeltzer, G. Droguett, Corros. Sci. 42 (2000) 1351.
L.M. Ocampo et al. / Corrosion Science 46 (2004) 1515–1525 1525

[4] A. Raman, B.E. Kuban, A. Razvan, Corros. Sci. 32 (12) (1991) 1295.
[5] L. Sorinas, F. Luzardo, T. Ochoa, E. Caraballo, A. Cabezas, L. Vargas, M. Garcia, Corros~ao e
Protecß~
ao de Materiais 16 (2) (1997) 19.
[6] M. Morcillo, S. Feliu, J. Simancas, J.M. Bastidas, J.C. Galvan, S. Feliu Jr., E.M. Almeida, Corros.
NACE 48 (12) (1992) 1032.
[7] J.C. Galv an, S. Feliu Jr., J. Simancas, M. Morcillo, J.M. Bastidas, E.M. Almeida, S. Feliu,
Electrochim. Acta 37 (11) (1992) 1983.
[8] J. Iwanow, Y.I. Kuznetsov, K. Setkowicz, in: Proceedings of 7th European Symposium on Corrosion
Inhibitors, vol. 5, Ferrara, Italy, 1990, p. 795.
[9] J. Gust, J. Bobrowicz, Corros. NACE 49 (1) (1993) 24.
[10] P.J. Deslauriers, Mater. Perfor. 26 (11) (1987) 35.
[11] A. Raman, in: Proceedings of 10th International Congress on Metallic Corrosion, vol. 1, Karaikudi,
India, 1987, p. 75.
[12] J.E.O. Mayne, J. Appl. Chem. 9 (1959) 673.
[13] D.L.A. de Faria, S.V. Silva, M.T. Oliveira, J. Raman Spectrosc. 28 (1997) 873.
[14] L. Igetoft, Proc. 2nd World Congress: Coatings Systems Bridges, Rolla, USA, 1982.
[15] J. West, P.C. Jackson, UK Corrosion’85, Harrogate, 1985, p. 4.
[16] M. Morcillo, S. Feliu, J.C. Galvan, J. Prot. Coat. Lin. 4 (9) (1987) 38.
[17] J. Gust, Corros. NACE 47 (6) (1991) 453.
[18] M. Favre, D. Landolt, K. Hoffman, M. Stratmann, Corros. Sci. 40 (4/5) (1998) 793.

View publication stats

Você também pode gostar