You are on page 1of 6

Materials Chemistry and Physics 83 (2004) 129134

Evaluation of the inhibitor effect of l-ascorbic acid


on the corrosion of mild steel
E.S. Ferreira, C. Giacomelli, F.C. Giacomelli, A. Spinelli
Grupo de Estudos de Processos Eletroqumicos e Eletroanalticos (GEPEEA), Departamento de Qumica,
Universidade Federal de Santa Catarina, 88040-900 Florianpolis, SC, Brazil
Received 31 March 2003; received in revised form 10 September 2003; accepted 22 September 2003

Abstract
The influence of l-ascorbic acid (AA) on mild steel corrosion in pH = 26 solutions was investigated using electrochemical and
weight loss techniques. Optical microscopic analyses provided the surface condition. The maximum inhibitor efficiency has been found
in 103 mol dm3 AA, pH = 4, solutions (69%). In these conditions, AA acts as a mixed-type inhibitor. Activation energies of 49.4 and
65.7 kJ mol1 for the corrosion processes were observed in absence and presence of AA, respectively. The AA inhibitor effect could also
be evidenced by surface optical images.
2003 Published by Elsevier B.V.
Keywords: Mild steel; l-Ascorbic acid; Acid inhibition; Organic inhibitors

1. Introduction adsorption bond strength determined by the electron density


and polarizability of the functional group [8].
The main strategy to prevent electrochemical corrosion is l-Ascorbic acid (AA)Vitamin C is an easily obtainable
the metal isolation from corrosive agents in the most effec- and water-soluble compound whose molecule has desirable
tive possible manner. Among the different methods avail- characteristics for a corrosion inhibitor. Nevertheless, a re-
able [13], the use of corrosion inhibitors is usually the view of the literature revealed that over the last two decades
most appropriate way to achieve this objective. Inorganic only a few works concerning the study of corrosion pro-
substances such as phosphate, chromate, dichromate, nitrite, cesses in AA-containing media have been reported [916].
nitrate and sulfide of alkaline metals, salts of cadmium and The AA effect on the corrosion of SS41 alloy was studied
arsenic, which are extensively used as corrosion inhibitors by Sekine et al. [10]. The authors found out a noteworthy
in several media and for different metals and alloys [3], de- corrosion rate decrease in 0.3% NaCl solutions containing
crease considerably the generalized oxidation of metals. On 200 ppm of AA, in comparison to solutions in absence of
the other hand, it is known that some of them (and their AA. The coupons had their surface analyzed by infrared and
derivative compounds) are toxic or pollute the environment ultraviolet spectroscopy. The results allowed the authors to
[4], limiting their application. Therefore, the investigation conclude the AA inhibitor effect is related to its adsorp-
of new non-toxic and green corrosion inhibitors is essential tion onto the electrode surface, but for AA concentrations
to overcome this problem. Among the alternative corrosion higher than 200 ppm complexing reactions producing sol-
inhibitors, organic products containing one or more polar uble Fe(II)-quelates take place. Gonalves and Mello [11]
functions (with N, O and S atoms) have shown to be quite have recently reported the inhibitive effect of AA on the cor-
efficient to prevent corrosion [5], as well as heterocyclic rosion resistance of low-carbon steel in 0.5 mol l1 Na2 SO4
compounds containing polar groups and electrons [6]. The solutions either under illumination or not. The effect is also
inhibiting action of those organic compounds is usually at- believed to be related with the adsorption of ascorbic acid
tributed to interactions with metallic surfaces by adsorption. on the surface active sites, blocking the water adsorption
The polar function is usually regarded as the reaction cen- reaction that is involved in the oxide layer formation.
ter for the adsorption process establishment [7], being the This work describes the inhibitor effect of l-ascorbic acid
on the corrosion of mild steel with 2.34% Cr in acid medium
Corresponding author. Tel.: +55-48-331-9219; fax: +55-48-331-9711. through potentiodynamic polarization curves, weight loss
E-mail address: spin@qmc.ufsc.br (A. Spinelli). and optical microscopy measurements.

0254-0584/$ see front matter 2003 Published by Elsevier B.V.


doi:10.1016/j.matchemphys.2003.09.020
130 E.S. Ferreira et al. / Materials Chemistry and Physics 83 (2004) 129134

Table 1 sonically, degreased with acetone, rinsed with deionized wa-


Chemical composition (wt.%) of the mild steel utilized (Fe is balanced) ter and air-dried. The potentiodynamic polarization curves
C 0.049 were shaped in agreement with the ASTM G5 norm [17].
Mn 0.227 The weight loss experiments were performed according
Cr 2.34 to the ASTM G31 norm [17]. Optical micrographs were
S 0.0005
accomplished using a Neophot 30 microscope.

2. Experimental 3. Results and discussion

The chemical composition of the used mild steel is shown Fig. 1 shows the open circuit potential (OCP) curves for
in Table 1. The chromium content is uncommon. The plate mild steel in pH = 4 solutions with different AA concen-
from where the test coupons and disks used for our experi- trations (103 to 107 mol l1 ). Soon after the electrode
ments were extracted was bought from a small local foundry, immersion, an accentuated displacement of OCP towards
which usually adds variable amounts of Cr in order to in- negative values is observed whether or not AA presence.
crease the corrosion resistance of the final product. This behavior is due to an initial dissolution process of
All the reagents used were of analytical grade acquired electrode oxide film. Afterwards (approximately 20 min),
from Vetec (l-ascorbic acid), Carlos Erba (sodium hydrox- the OCP remains slightly constant for practically all solu-
ide) and Ecibra Reagentes Analticos (98% sulfuric acid). tions, except when the AA concentration is 103 mol l1 . In
All of them were used without previous purification. Dis- this case, a gradual OCP increase occurs, which indicates a
tilled and deionized water was used for all solution prepa- possible film formation onto the electrode surface. Similar
rations. Solutions of 0.01 mol l1 H2 SO4 with pH carefully results were also obtained for pH = 2, 3, 5 and 6 solutions.
adjusted to 2.0, 3.0, 4.0, 5.0 and 6.0 by addition of NaOH Table 2 reports the measured OCP at the end of 60 min
stock solutions were employed as blank, i.e., l-ascorbic acid of immersion for pH = 26 solutions with several AA con-
free. For experiments containing l-ascorbic acid, the solid centrations. The results reproducibility was satisfactory; the
was added to blank solutions to reach final concentrations of deviation is approximately 1015%. In Table 2, it is possible
103 , 104 , 105 , 106 and 107 mol l1 . All experiments
were carried out at room temperature and non-deaerated so- -450
lutions. [AA] (mol l-1)
The electrochemical measurements were carried out with blank
-500
an EG&G-PARC model 263A potentiostat/galvanostat inter- 10-3
faced with a microcomputer using the EG&G-PARC Soft- 10-4
-550
Corr II Model 252/352 software for data acquisition and 10-5
E (mV)

analysis. The electrochemical cell contained five openings: -600 10-6


three of them were used for the electrodes and two for nitro- 10-7
gen bubbling during all the experiments. The counter elec- -650
trode was a graphite rod and the reference electrode was a
saturated calomel electrode (SCE) connected to the cell by a -700
bridge and a LugginHabber capillary. All potentials in the
text are quoted versus this electrode. The working electrode -750
0 600 1200 1800 2400 3000 3600
was a 0.502 cm2 (geometrical surface area) mild steel plate
time (s)
mounted in a glass tube with Araldite epoxy. Prior to the
experiments, the electrode surface was polished with 1200 Fig. 1. OCP curves for mild steel in pH = 4 solutions with 103 to
emery paper. Subsequently, the electrode was cleaned ultra- 107 mol l1 AA.

Table 2
OCP values recorded after 60 min of immersion for mild steel electrodes
[AA] (mol l1 ) OCP/OCPa (mV)

pH = 2 pH = 3 pH = 4 pH = 5 pH = 6

Blank 597 711 668 736 668


103 605/8 727/16 641/+27 732/+4 484/+184
104 655/58 727/16 729/61 712/+24 711/43
105 628/61 734/23 689/21 726/+10 675/7
106 667/70 727/16 731/63 716/+20 676/8
107 631/34 715/4 678/10 727/+9 693/25
a OCP = OCP (inhibitor) OCP (blank).
E.S. Ferreira et al. / Materials Chemistry and Physics 83 (2004) 129134 131

-400 2.2 blank


[AA] = 10-3 mol l-1

1.8
log CR (mdd)

-600
E (mV)

1.4
[AA] (moll-1)
-800 blank 10 -3 1.0
10 -4 10 -5
10 -6 10 -7
0.6
-1000 30 31 32 33 34
10-7 10-6 10-5 10-4 10-3 10-2
j (A cm-2) 104/T (K-1)

Fig. 2. Potentiodynamic polarization curves for mild steel in pH = 4 Fig. 3. Arrhenius plot for mild steel in pH = 4 solution with 103 mol l1
solutions with 103 to 107 mol l1 AA. AA.

solutions and 103 mol l1 AA. The IE of 69% achieved


to notice that the influence of AA content on the OCP is in those conditions was attributed to the adsorption of the
not clear. However, the reached OCP for practically all AA l-dehydroascorbic acid (DHA) (an AA oxidation product,
containing solutions was more negative than in its absence, see hereinafter) onto the mild steel surface.
except for pH = 5 solution. According to Riggs [6], the clas- Fig. 3 shows the Arrhenius plot for the corrosion of mild
sification of a compound as an anodic or cathodic inhibitor steel samples in pH = 4 solutions in the absence and pres-
is feasible when the OCP displacement is at least 85 mV in ence of 103 mol l1 AA. Corrosion rates are expressed in
relation to that one measured for the blank solution. From milligrams per square decimeter per day (mdd). The activa-
the electrochemical and gravimetric studies carried out using tion energy determined from the slope of the Arrhenius plot
pH = 6 NaCl solutions (0.3 and 0.03%), Sekine et al. [10] corresponds to 49.4 kJ mol1 (in the absence of AA) and
proposed that AA is an anodic inhibitor of mild steel cor- 65.7 kJ mol1 (in the presence of AA). The activation en-
rosion, although they did not perform OCP measurements. ergy value obtained in absence of AA is similar to that one
Gonalves and Mello [11] pointed out an positive shift of recorded by other authors in the same conditions [1820].
100 mV in the corrosion potential of low-carbon steel in Since the corrosion process primarily occurs in the surface
0.50 mol l1 Na2 SO4 and 102 mol l1 AA solutions. Our spots free of organic adsorbed molecules [21], the higher
experiments also show a remarkable positive displacement activation energy in presence of AA can be interpreted as
for pH = 6 solutions and relatively high inhibitor concen- being due to the decrease of the surface coverage degree
trations (103 mol l1 ), which is in agreement with the pro- increasing temperature. A similar behavior had been previ-
posals of Sekine et al. [10] and Gonalves and Mello [11]. ously observed in a study of the inhibition of low-carbon
On the other hand, we classify AA as a mixed corrosion in- steel corrosion by propargyl alcohol [22].
hibitor in the moderately low concentration range for the or- Table 3 shows the corrosion rate values and the IE pro-
ganic compound studied. Thus, AA acts simultaneously on moted by a 103 mol l1 AA containing solution of pH = 4
anodic and cathodic reactions. This type of inhibitors does as function of temperature. A proportional corrosion rate in-
not produce OCP displacements (or only very small ones) crease because of the temperature increase from 20 to 60 C
in relation to the blank solution [5]. for pH = 4 solutions either with or without AA is pointed
The potentiodynamic polarization curves for mild steel in out as expected. The IE drastically decreases from 72% at
pH = 4 solutions with different AA concentrations (103 20 C to 39% at 60 C. Fig. 4 shows the composition of AA
to 107 mol l1 ) are shown in Fig. 2. Concerning the an- solutions (molar ratio [23]) as function of pH. As mentioned
odic region in relation to the zero current density poten-
tial, there is no evidence of passive film formation onto Table 3
Temperature effect on the IE of 103 mol l1 AA in pH = 4 solutions
the electrode surface either in the presence or in the ab-
for mild steel
sence of the inhibitor. Curves with a related profile were
also recorded for pH = 2, 3, 5 and 6 solutions. The in- Temperature ( C) Corrosion rate (mdd) IE (%)
hibition efficiency (IE) was calculated using the classical Blank solution 103 mol l1 AA
equation IE = (icorr iinhcorr / icorr ) 100, where icorr and
inh
20.0 17.9 5.1 71.5
icorr are the corrosion current densities measured in pres- 25.0 24.5 7.6 69.0
ence and absence of AA, respectively. The IE promoted by 30.0 25.2 8.1 67.8
AA for mild steel was low, rarely exceeding 30%, even for 40.0 127.9 42.2 67.0
high AA concentrations. The unique experimental condi- 50.0 130.5 76.6 41.8
60.0 152.6 92.5 39.4
tion in which reasonable efficiency was observed is pH = 4
132 E.S. Ferreira et al. / Materials Chemistry and Physics 83 (2004) 129134

600
1.0
weight loss, 2h
0.8
electrochemical
400
molar ratio

0.6 L-ascorbic acid


CR (mdd)

L-monoascorbate anion
0.4 ascorbate anion

200
0.2

0.0
0
0 2 4 6 8 10 12 14
pH 2 3 4 5 6
(a) pH
Fig. 4. Composition of AA solutions as function of pH value.

before, the corrosion inhibition process was attributed to the immersion time / h
60 2
adsorption of DHA onto the electrode surface. Its presence 6
can be understood analyzing the data shown in Fig. 4. For 12
pH = 4, where the highest protection efficiency was per- 24
30
IE (%)

ceived, the solution composition is about 60% l-ascorbic 72


168
acid and 40% l-monoascorbate anion. l-Monoascorbate an-
ions are very reactive species that easily undergo redox re-
0
actions. Fig. 5 illustrates a mechanism for l-monoascorbate
anion oxidation to DHA [2426]. Under the conditions in
which the experiments were carried out, the DHA concen-
-30
tration is certainly higher than the l-monoascorbate anion. 2 3 4 5 6
DHA presents chemical stability in the pH range 46, but
(b) pH
for pH > 7 solutions it undergoes an hydrolysis reaction
producing the l-2,3-diketogluconic acid [25]. Thereby, we Fig. 6. Comparison between corrosion rates (CR) obtained from weight
believe that DHA is the main species responsible for the in- loss and electrochemical experiments (a) and AA IE calculated from
hibition process, although we do not have experimental ev- weight loss measurements (b) for mild steel in pH = 26 solutions
containing 103 mol l1 AA.
idences of DHA presence in the solution. The low IE found
for pH < 4 solution is associated to the higher concentra-
tion of AA, a complexing agent for several metals [10,27], concerning the IE and corrosion rate results from electro-
such as chromium and iron ions, which generates products chemical tests can be satisfactorily applied to the weight
with high stability constants, significantly diminishing the loss ones inasmuch as they matched up. Fig. 6b shows the
inhibitive effect of AA. Dean et al. [5] also reported the cor- IE as function of pH for coupons immersed during several
rosion rate increase of alloys due to the action of complexing periods in 103 mol l1 AA solutions. For pH = 2 and 3
agents. solutions, one observes a mild steel dissolution kinetics in-
Fig. 6a shows the corrosion rate variation as function of crease due to AA presence, reflected by negative IE values,
pH for 103 mol l1 AA solutions obtained from electro- i.e., the corrosion rate is higher for AA containing solutions
chemical and weight loss (immersion time = 2 h) exper- than for blank ones. However, this effect disappears for peri-
iments. One perceives the curves profile similarity, which ods longer than 6 h by reason of oxide film formation, which
indicates an analogous AA inhibition mechanism for both overcomes the AA influence on the system. For shorter pe-
conditions, even though there is a certain data discrepancy riods (2 and 6 h), the potentiality of AA for complexing for-
due to the electrode polarization effects [1]. The comments mation reaction [27] is thought to provoke the corrosion rate

CH2OH CH2OH CH2OH CH2OH CH2OH

HOCH HOCH HOCH HOCH HOCH O O


O O O O O O O O
- H+ - e- - e- - H+
+ H+ + e- + e- + H +
HO OH
-
O OH O. OH HO O
. O O

L-ascorbic acid L-monoascorbate ascorbic L-dehydroascorbic


anion free radicals acid

Fig. 5. Oxidation of l-monoascorbate anion to DHA.


E.S. Ferreira et al. / Materials Chemistry and Physics 83 (2004) 129134 133

4. Conclusions

The results obtained from potentiodynamic polarization


curves, weight loss measurements and optical microscopy
show that AA is a reasonable corrosion inhibitor for mild
steel (with 2.34% Cr). The maximum IE was 69% for
pH = 4 solutions with relatively high AA concentration
(103 mol l1 ). The corrosion process is inhibited by the
adsorption of the DHA (an AA oxidation product) onto
the mild steel surface. DHA adsorption promotes an in-
crease in corrosion activation energies from 49.4 kJ mol1
(blank solution) to 65.7 kJ mol1 (103 mol l1 AA solu-
tion) at pH = 4. The data recorded from OCP curves are
not sufficiently conclusive to clearly characterize AA as
a cathodic or anodic. In most of the experiments carried
out the OCP displacement for different AA concentrations
was not significant, which would characterize it as a mixed
type inhibitor. The beneficial effect of AA on corrosion of
mild steel was also clearly indicated by optical micrographs
taken for coupons immersed in solutions containing AA or
not.

Acknowledgements

The authors thank Conselho Nacional de Desenvolvi-


mento Cientfico e Tecnolgico (CNPq) for financial support
and Grupo de Estudos em Materiais Polimricos (POLI-
MAT) for the microscope facility.

References

[1] H.H. Uhlig, Corrosin y control de corrosin, Urmo, Bilbao, 1975.


[2] D.A. Jones, Principles and Prevention of Corrosion, Macmillan, New
Fig. 7. Optical images of coupons immersed in (a) blank and (b) York, 1992.
103 mol l1 AA solutions of pH = 4. The original obtained images have [3] M.G. Fontana, Corrosion Engineering, 3rd ed., McGraw-Hill,
been posterized and black-white transformed. Singapore, 1986.
[4] S.E. Manahan, Environmental Chemistry, 6th ed., Lewis, Boca Raton,
1994.
increase. Solutions of pH = 4 and 5 with 103 mol l1 AA [5] S.W. Dean Jr., R. Derby, G.T. Vondembussche, Mater. Perform. 20
have presented noteworthy IE values for immersion times (1981) 47.
[6] O.L. Riggs Jr., Corrosion Inhibitors, 2nd ed., C.C. Nathan, Houston,
of 2 and 6 h. Such an effect agrees with the above discussed
TX, 1973.
adsorption inhibition mechanism. Lately, for pH = 6, the [7] M.G. Fontana, K.W. Staehle, Advances in Corrosion Science and
low IE is related to the aggressiveness decrease. Thereby, Technology, vol. 1, Plenum Press, New York, 1970.
the AA has no noticeable influence on corrosion properties [8] S. Sankarapapavinasan, F. Pushpanaden, M.F. Ahmed, B.
of mild steel, despite the DHA existence into the solutions. Electrochem. 5 (1989) 319.
Fig. 7 shows optical images of coupons immersed in (a) [9] A. Rosanoff, G.M. Briggs, B.O. Delumen, J. Agric. Food Chem. 33
(1985) 891.
blank and (b) 103 mol l1 AA solutions of pH = 4. The [10] I. Sekine, Y. Nakahata, H. Tanabe, Corros. Sci. 28 (1988) 987.
original colored images were posterized and black-white [11] R.S. Gonalves, L.D. Mello, Corros. Sci. 43 (2001) 457.
transformed (in this sequence) with the objective of better [12] A.N. Nigan, R.P. Tripathi, M.L. Jangid, K. Dhoot, M.P. Chacharkar,
showing the AA effect. The clear area in the micrographs Corros. Sci. 30 (1990) 201.
comprises rust formation onto the surface, being charac- [13] M. Shailaja, S.V. Narasimhan, J. Nucl. Sci. Technol. 28 (1991) 1107.
[14] A. Albuyaron, A. Feigin, J. Sci. Food Agric. 59 (1992) 101.
terized by its color. The beneficial effect of AA presence
[15] I.H. Farooqi, M.A. Nasir, M.A. Quraishi, Corros. Prevent. Contr. 44
is clearly observable comparing the rust density in Fig. 7a (1997) 129.
(blank solution) and Fig. 7b (103 mol l1 AA solution), [16] G.R. Dey, D.B. Naik, K. Kishore, C.K. Vinayakumar, B. Yuvaraju, G.
which is significantly lower in Fig. 7b than in Fig. 7a. Venkateswaran, P.N. Moorthy, Radiat. Phys. Chem. 51 (1998) 171.
134 E.S. Ferreira et al. / Materials Chemistry and Physics 83 (2004) 129134

[17] American Society for Testing and Materials, Annual Book of ASTM [24] M.B. Davies, J. Austin, D.A. Partridge, Vitamin C: Its Chemistry
Standard, ASTM, Philadelphia, PA, 1978. and Biochemistry, Royal Society of Chemistry, London, 1991.
[18] O.L. Riggs Jr., R.M. Hurd, Corrosion 23 (1967) 252. [25] S. Lewin, Vitamin C: Its Molecular Biology and Medical Potential,
[19] D. Altura, K. Nobe, Corrosion 29 (1973) 433. Academic Press, London, 1976.
[20] J.J. Podesta, A.J. Arvia, Electrochim. Acta 10 (1965) 159. [26] P.A. Seib, B.M. Tolbert, Ascorbic Acid: Chemistry, Metabolism and
[21] A. Sieverts, P. Lueg, Z. Anorg. Chem. 126 (1923) 192. Uses. Advances in Chemistry, Series No. 200, American Chemical
[22] A. Spinelli, R.S. Gonalves, Corros. Sci. 30 (1990) 1235. Society, Washington, DC, 1982.
[23] D.A. Skoog, D.M. West, F.J. Holler, Fundamentals of Analytical [27] A.E. Martell, R.M. Smith, Critical Stability Constants, vol. 3, Plenum
Chemistry, 6th ed., Saunders College Publishing, New York, 1992. Press, New York, 1989.

You might also like