You are on page 1of 9

hydrometallurgy

ELSEVIER Hydrometallurgy42 (1996) 257-265

Mechanisms of dissolution of iron oxides in


aqueous oxalic acid solutions
D. Panias, M. Taxiarchou, I. Paspaliaris, A. Kontopoulos *
Laboratory of Metallurgy, National Technical University of Athens, P.O. Box 64056, GR-157 80 Zographos,
Greece
Received 20 April 1995; accepted 30 October 1995

Abstract

The dissolution of pure iron oxides by organic acids has been extensively reviewed. The
mechanism of dissolution comprises three distinct steps: (1) adsorption of organic ligands on the
iron oxide surface; (2) non-reductive dissolution; and (3) reductive dissolution. Reductive dissolu-
tion involves two stages: an induction period and an autocatalytic period. The overall dissolution
process is affected by the pH of the initial solution, temperature, the exposure of solution to UV
radiation and the addition of bivalent iron in the initial solution.

1. Introduction

The dissolution of metal oxides is an important process in several fields, such as


hydrometallurgy, the passivity of metals and cleaning of metal surfaces. Iron oxides are
the most studied of the various metal oxides, due to their wide occurrence in natural
systems. Many experimental studies have been reported in the literature concerning the
dissolution of pure iron oxides by means of inorganic or organic acids. All studies aimed
to identify:
1. the most promising acids (organic and inorganic) as solvent agents;
2. the conditions that intensify their action;
3. the probable mechanisms prevailing during the dissolution process.
In hydrometallurgical processes, bacterially produced organic acids have been pro-
posed as an alternative, and probably less expensive, leaching agent. Organic acids such
as oxalic, citric, ascorbic, acetic, fumaric and tartaric acid, have been screened for their

* Corresponding author. E-mail: kontop@metal.ntua.gr.

0304-386X/96/$15.00 Copyright © 1996 Elsevier Science B.V. All rights reserved.


SSDI 0304-386X(95)00104-2
258 D. Panias et al. / Hydrometallurgy 42 (1996) 257-265

ability to solubilise iron and other metal oxides. Of the above organic acids, oxalic
[1-3], citric [4] and ascorbic [5] are the most used carboxylic acids, due to their
effectiveness as solvent reagents. Iminodiacetic acid (IDA), N-hydroxyethyl-iminoacetic
acid (HIDA) and ethylenediaminotetraacetic acid (EDTA) have also been studied [6,7]
as solvent reagents in aqueous solutions with satisfactory results.
The present review aims to collect literature data concerning iron oxide dissolution in
organic solvents and is especially focused on the mechanisms prevailing in this process.

2. Mechanism of dissolution of iron oxides in organic acids

Numerous kinetic studies have been performed on the dissolution of iron oxides with
organic acids in order to reveal the reaction mechanism. Most of them have been
conducted with pure iron oxides or synthetic ferrites [8-12] rather than with iron-bearing
minerals. In such systems interference by other metallic compounds on the dissolution
process is excluded and attention to the real chemical pathways is given.
The dissolution mechanism involves three different processes taking place simultane-
ously:
1. adsorption of organic ligands from the solution on the system interface;
2. non-reductive dissolution [6,10,13];
3. reductive dissolution [1-3,5-7,10,14].

2.1. Adsorption o f organic ligands on the system interface

It is generally accepted that the first step of the dissolution reaction is the adsorption
of carboxylic acid on the surface of iron oxides. When an iron oxide particle is
suspended in an acidic solution an electrical double layer [15-20] is established on the
system interface:
H , L ~ nH+ + L "- Acid ionisation (1)
)Fem-O + H +~ )Fem-O... H + Protonation of oxygen (2)
The surface of the oxide can behave as a Lewis base (electron pair donor) according
to Eq. (2) and, consequently, the system interface appears positively charged. At a given
pH value of the initial solution, the higher the ionic strength of the solution, the higher
the surface charge associated with adsorbed protons [19]. The above pathway plays an
important role in oxide dissolution as the metal-oxygen bond is loosened due to the
protonation process. The surface hydroxyl groups (-OH) become active sites for the
subsequent adsorption of organic ligands [1-4,13,19,21-24], as described by:
)Fern-OH + + L"- + H + ~ [)FellI-L] -(n-2) q- H20 Surface complexation (3)
The above is a competitive reaction between active surface hydroxyl groups and
anions from the acidic solution, such as oxalates, citrates and ascorbates.
The greater the stability of [FenI-L] -(~ -2) complexes, the greater the yield of Eq. (3).
Generally, as the chelating capacity of a ligand increases, the stability of [Fern-L] -~" -2)
complexes also increases and the above reaction shifts to the right.
D. Panias et al. /Hydrometallurgy 42 (1996) 257-265 259

100% 100"/o

[#1 2] 2-
0
0 6O% "~ eO*/,
o o
4O% t~ 40*/0

20% 1~ 20%

0% 0%
0 2 4 6 2 3 4

pH pH
Fig. 1. Speciation in bulk solution as a function of pH for (left) oxalic acid and (right) Fe 2+/oxalic acid [1].

2.1.1. Adsorption of oxalates on the system interface


A pH-dependent equilibrium is established between the two different surface
> Fem-oxalate complexes [2,3,24]:

o OH

IIC ~ C'--
I 0
ife-- O~ ~-'-- 0 ~Pe-- 0 /
+H + + HC:Oi
OH
e-- 0 / C---- 0
c--c__--o
II
o

The speciation of oxalate solutions as a function of pH is shown in Fig. 1. This figure


also shows the speciation of Fe2+/oxalic acid solution.

2.2. Non-reductive dissolution

The non-reductive dissolution pathway is a simple desorption process. It involves the


desorption of adsorbed surface ferric complex ions and their transfer to the acidic
solution:
[)FeIII_L] -(n-2) + H + ~ - [Fe 3+-LJ(,q)
1-(n-3) + )H
-

(4)
The non-reductive dissolution mechanism removes only the more reactive sites of the
oxide surface. The number of the latter increases with decreasing pH and increasing
temperature [6,13]. As observed in the potential energy-reaction coordinate diagram
(Fig. 2), the desorption process is characterised by a high activation energy. As a result,
non-reductive dissolution at low temperature is not an operative pathway. As the
temperature is raised, the above pathway may become increasingly important and,
eventually, override the reductive dissolution process as the most important pathway
[13].
260 D. Panias et al. / Hydrometallurgy 42 (1996) 257-265

activation
energy of
lesorption
)rocees
=4 Fe"l(~
[Fe'"L]~){,}
~__t/
[Fe"lL]~-2~o
Reaction Coordinate
Fig. 2. Reaction coordinate diagram for non-reductive dissolution process in acidic solution.

2.3. Reductive dissolution

Reductive dissolution can be divided in two different stages interacting with each
other:
1. induction period (first stage);
2. autocatalytic dissolution period (second stage).

2.3.1. Induction period


During the induction period the generation of ferrous ions in the solution takes place.
In the presence of lattice Fe n , as in magnetite, ferrous ions are dissolved [2,3,5] and their
concentration in the solution is slowly increased, as described by the following equation:

[)Fen L ] - (n- 2) --o tr e~ e 2+ -L.](aq


T 1 - ( n) - 2 ) (5)

This observation is consistent with theories concerning the dissolution of metal


oxides [19]. In general, the thermodynamic solubility, the ionic character of the M - O
bonds [2,25] and the rate of dissolution decrease as the charge of the cation increases
and its radius decreases. Iron as a transition element appears in two oxidation states. Fe II
can be transferred to the solution more readily than Fe III because of the greater lability
(kinetic instability) of the F e l l - o bond as compared to the FetII-o bond.
In the absence of lattice Fe II ions, as in hematite, the generation of Fe 2+ ions in the
solution is a very slow process. It involves electron transfer [2,7,10,13] from the
adsorbed complex to surface Fe llI ions:

DFeIII L ._]-(n-2)~_~ [)Feli L(n_I)_]-(n-2)

Electron transfer from the ligand to F e I11 (6)

DFen_L(n - 1)- ]-{n-2)._, Fe~aq) + products of ligand oxidation Dissolution (7)

2+ + L n - ~ [Fe
Fe(aq) " 2+ -L](aq)
_ ~-(n-2) Complexation (8)
D. Panias et a l . / Hydrometallurgy 42 (1996) 257-265 261

At this stage, the dissolution of iron takes place in the form of Fe 2+ rather than in the
form of Fe 3+.

2.3.1.1. Induction period in the Fetn-oxalates system. When oxalic acid is used as
solvent reagent, the above mechanism can be described by the following equations [10]:
[)Fem-C20~ -] ~ [ ) F e n - C 2 0 ; ] Electron transfer (9)
2 [ ) F e ' I - c 2 0 4 ] + 2H +---> 2Fe(2aq)+ 2CO 2 + C 2 0 ~- + 2)H Dissolution as Fe 2 ÷
(10)
At this stage, the factors affecting the rate of dissolution are:
Temperature: It is reported [2,7,10] that, at high temperature (150°C), complete
dissolution is rapidly achieved; the solution contains only Fe 2+ species generated by the
reduction of Fe nl ions with oxalate ions, through a partially or totally heterogeneous
process.
Visible light impacts: Photochemical electron transfer on surface > FeIn-oxalate
complexes provides an additional pathway for the start of dissolution. Photolysis of
[Fe(C204)3] 3- takes place over a wide range of wavelengths (visible and ultraviolet
region) according to the following equation [1-3]:

2[ Fe3* ( C204 )3 ]~aq, "~ 2[ Fez * (C204) ] (aq) + 3C2 O2- + 2CO2 ( 11 )

2.3.2. Autocatalytic dissolution period


When a sufficient amount of ferrous oxalate ions has been formed, the secondary
reductive dissolution step becomes operative and the whole process is accelerated. This
pathway is described by the following equations:
[ ) F e m - L ] - ( n - 2 ) + [We2+ -Ll(aq)
- -,-(n-2) ~ DFelU-L] -(n-2) "" [Fe2+-L] -(n-2)

Adsorption of complex to surface (12)


[)FeIll L ] - ( n - 2 ) . . . [We2+_L]-(n-2)¢¢..~ [)Well L ] - ( n - l ) . . . [We3+ L] -(n-3)
Electron transfer (13)
DFe,,_L]-(~-1) . . . [Fe3 + _ L ] - ( ~ - 3 ) . [)FeII_L]-(.-,) + [Fe3 +_L]-tn-3)
Desorption (14)
" 2+ - -~-(n-2)
H +'-" tFe -t.l aq + )H Dissolution (15)
The above reaction scheme can be divided into three elementary steps:
1. adsorption of aqueous solution ferrous complexes on the surface ferric complexes;
2. a fast outer-sphere or inner-sphere electron transfer [18,20,26] and formation of Fe II
on the system interface; electron microscopy has shown that Fe n generated on the
surface by electron transfer is more reactive than normal lattice Fe n [3], as a result,
its transfer to the solution by Eq. (15) is easier in relation to Eq. (5) and Eq. (6), (7)
and Eq. (8);
3. desorption of ferric complexes and transfer of trivalent iron in the solution.
262 D. Panias et al. / HydrometaUurgy 42 (1996) 257-265

60

50.

40

E 30

20

10 ¸

0
0 0.5
[ Fe(ll). ]l[Fe(ll)~d I
I1 concentration,expressed as its ratio respective to the
Fig. 3. Induction period, "r, as a function of Fe(aq)
concentrationof FeII originatedby total dissolutionof the solid (pH 4.1, temperature30°C, magnetite/oxalic
acid solution) [2].

2.3.2.1. Autocatalytic dissolution period in iron-oxalate system. In the case of dissolu-


tion in oxalic acid the above reductive dissolution pathway can be illustrated by the
following equations [2,3].
[ ) F e " ' - o x ] + [Fe2+-oX]Caq)~ ) F e " I - o x . . . Fe2+-ox (16)
) F e " I - o x . . . Fe 2+-ox ~ ) F e " - o x . . . Fe 3÷-ox (17)
)FeI'-ox... Fe 3+-ox --, ) F e l [ - o x q- [Fe 3+ OX](aq) (18)
) F e n - o x --~ [Fe2+-oX](aq) (19)
where ox denotes any species derived from oxalic acid:
[ox] = [H2C204] + [HC204] + [C2042- ]

2.3.3. Factors affecting reductive pathway


The rate of reductive dissolution is affected by two factors:
Addition of ferrous ions in the initial acidic solution: Addition of bivalent iron in the
initial solution exerts a beneficial influence on the overall dissolution rate because it
reduces the time-consuming induction period (Fig. 3). As can be understood by the
previous reductive dissolution mechanism, the presence of Fe E+ ions in the initial
solution favours the autocatalytic dissolution period according to Eqs. (6), (7) and (12).
Illumination of iron oxides suspensions: The illumination of metal oxide suspensions
with UV radiation gives exactly the same result [1,4,27]. As has been previously
reported, UV radiation assists the reduction of surface ferric ions and accelerates the
overall reductive pathway.

3. Factors affecting the dissolution m e c h a n i s m

The dissolution process is affected by several parameters. The most important are:
pH of the initial aqueous solution: pH values in the range 2 - 3 give satisfactory
D. Panias et al. / Hydrometallurgy 42 (1996) 257-265 263

2.5

m 2
E 1.5

g 1
"- 0.5

0 I I I
2 3 4

pH

Fig. 4. Influenceof pH on the initial dissolution rate of magnetite, R0 ([Oxalic acid]= 0.1 M, temperature
30oc) [3].

results, with the optimum value being 2.6 [1,5,10,17,19]. The influence of pH on the
dissolution rate of magnetite in the system oxalic acid-magnetite is plotted in Fig. 4.
Temperature: It is generally accepted that high temperatures, above 90°C, enhance
the dissolution of iron oxides [1,5,16,17,19].
Chelating capacity of polycarboxylic acids: Chelating capacity is the ability of
polycarboxylic acids to form chelated compounds (chelates) with metal ions in the
solution. It is well known that the dissolution process is assisted by chelating agents
such as EDTA and oxalic acid [4-7,14,16,19,22,28,29].
Photochemical effects: The dissolution of iron oxides suspended in organic acid
solutions is promoted by ultraviolet and visible light radiation [1,4,16,27,30]. This
process involves photochemical reduction of trivalent iron to bivalent, exerting a
beneficial influence on the dissolution process.

4. C o n c l u s i o n s

The dissolution of iron oxides with organic acids comprises three distinct steps:
1. Adsorption of organic ligands on the iron oxide surface. This is a chemisorption
rather than a physical sorption process. The Lewis acid-base properties of surface
iron oxides must be taken into account for the better understanding of this step.
2. Non-reductive dissolution. This is a simple desorption process of surface iron-ligand
complexes. It is characterised by high activation energy and, therefore, it is an
important dissolution pathway only at high temperatures.
3. Reductive dissolution. This is the main mechanism of iron oxides dissolution with
organic acids. It consists of two different stages:
3.1. induction period;
3.2. autocatalytic period.
The overall dissolution process is affected by:
1. pH of the initial solution;
2. temperature;
3. illumination of the solution with UV radiation;
4. the addition of bivalent iron to the initial solution.
264 D. Panh~s et al. / Hydrometallurgy 42 (1996) 257-265

5. L i s t o f symbols
L n- : a n y organic ligand with oxidation n u m b e r n, such as oxalate (L n - =
C2042- or H C 2 0 4 ) , citrate ( L n - = C 6 H 5 0 7 3 - or C6H6072- or
C6H70 7)
): particle surface
)FelIl: trivalent lattice iron on the particle surface
)FelI: b i v a l e n t lattice iron on the particle surface
[)Fe-L]: surface c o m p l e x
a d s o r b e d species on the particle surface
II, HI: oxidation n u m b e r of surface lattice iron
n+,n-: v a l e n c e o f aqueous species

Acknowledgements

The financial support of the E u r o p e a n C o m m i s s i o n w i t h i n the framework o f the


B r i t e - E u r a m II P r o g r a m (Contract No. B R E 2 - C T 9 2 - 0 2 1 5 ) is gratefully acknowledged.

References
[1] Comell, R.M. and Schindler, P.W., Photochemical dissolution of goethite in acid/oxalate solution. Clays
Clay Miner., 35(5) (1987): 347-352.
[2] Baumgartner, E.C., Blesa, M.A., Marinovich, H.A. and Maroto, A.J.G., Heterogeneous electron transfer
as a pathway in the dissolution of magnetite in oxalic acid solutions. Inorg. Chem., 22 (1983):
2224-2226.
[3] Blesa, M.A., Marinovich, H.A., Baumgartner, E.C. and Maroto, A.J.G., Mechanism of dissolution of
magnetite by oxalic acid-ferrous ion solutions. Inorg. Chem., 26 (1987): 3713-3717.
[4] Waite, T.D. and Morel, F.M.M., Photoreduction dissolution of colloidal iron oxide: effect of citrate. J.
Colloid Interface Sci., 102(1) (1984): 121-137.
[5] Afonso, M.S., Morando, P.J., Blesa, M.A., Banwart, S. and Stumm, W., The reductive dissolution of iron
oxides by ascorbate. J. Colloid Interface Sci., 138(1) (1990): 74-82.
[6] Tongs, R., Blesa, M.A. and Matijevic, E., Interactions of metal hydrous oxides with chelating agents.
VII. Dissolution of hematite. J. Colloid Interface Sci., 131(2) (1989): 567-579.
[7] Tones, R., Blesa, M.A. and Matijevic, E., Interactions of metal hydrous oxides with chelating agents. IX.
Reductive dissolution of hematite and magnetite by aminocarboxylic acids. J. Colloid Interface Sci.,
134(2) (1990): 475-485.
[8] Valverde, N., Investigations on the rate of dissolution of ternary oxide systems in acidic solutions. Ber.
Bunsenges. Phys. Chem., 81(4) (1977): 380-384.
[9] Regazzoni, A.E. and Matijevic, E., Interactions of metal hydrous oxides with chelating agents. VI.
Dissolution of a nickel ferrite in EDTA solutions. Natl. Assoc. Corrosion Eng., 40(5) (1984): 257-261.
[10] Sellers, R.M. and Williams, W.J., High-temperature dissolution of nickel chromium ferrites by oxalic
acid and nitrilotriacetic acid. Faraday Discuss. Chem. Soc., 77 (1984): 265-274.
[11] Biesa, M.A, Maroto, A.J.G. and Morando, P.J., Dissolution of cobalt ferrites by thioglycolic acid. J.
Chem. Soc. Faraday Trans. I, 82 (1986): 2345-2352.
[12] Segal, M.G. and Sellers, R.M., Kinetics of metal oxide dissolution: reductive dissolution of nickel ferrite
by tris(picolinato)vanadium(II). J. Chem. Soc. Faraday Trans. I, 78 (1982): 1149-1164.
[13] Rucda, E.H. and Grassi, R.L., Adsorption and dissolution in the system goethite/aqueous EDTA. J.
Colloid Interface Sci., 106(1) (1985): 243-246.
D. Panias et al. / Hydrometallurgy 42 (1996) 257-265 265

[14] Blesa, M.A, Borghi, E.B., Maroto, J.G. and Regazzoni, A.E., Adsorption of EDTA and iron-EDTA
complexes of dissolution of magnetite by EDTA. J. Colloid Interface Sci., 98(2) (1984): 295-305.
[15] Godchev, I.G. and Kipriyanov, N.A., Regular kinetic features of the dissolution of metal oxides in acidic
media. Russ. Chem. Rev., 53(11)(1984): 1039-1061.
[16] Valverde, N. and Wagner, C., Considerations on the kinetics and the mechanism of the dissolution of
metal oxides in acidic solutions. Ber. Bunsenges. Phys. Chem., 80(4) (1976): 330-333.
[17] Gorichev, I.G. and Kipriyanov, N.A., Kinetics of the dissolution of oxide phases in acids. Russ. J. Phys.
Chem., 55(11) (1981): 1558-1568.
[18] Vermilyea, D.A., The dissolution of ionic compounds in aqueous media. J. Electrochem. Soc., 113(10)
(1966): 1067-1070.
[19] Blesa, M.A. and Maroto, AJ.G., Dissolution of metal oxides. J. Chim. Phys., 83(11/12) (1986):
757-764.
[20] Pankow, Aquatic Chemistry Concepts. Lewis, Michigan (1991), pp. 603-641.
[21] Comell, R.M. and Schindler, P.W., Infrared study of the adsorption of hydroxycarboxylic acids on
ct-FeOOH and amorphous Fe(lII) hydroxide. Colloid Polymer Sci., 258 (1980): 1171-1175.
[22] Rubio, J. and Matijevic, E., Interactions of metal hydrous oxides with chelating agents. I. fl-FeOOH-
EDTA. J. Colloid Interface Sci., 68(3) (1979): 408-421.
[23] Grauer, R. and Stumm, W., Die Koordinationschemie oxidischer Grenzflachen and ihre Auswirkung auf
die Auflosungskinetik oxidischer Festphasen in wal3dgen Losungen. Colloid Polymer Sci., 260 (1982):
959-970.
[24] Buckland, A.D., Rochester, C.H. and Topham, S.A., Infrared study of the adsorption of carboxylic acids
on hematite and goethite immersed in carbon tetrachloride. J. Chem. Soc. Faraday Trans. I, 76 (1980):
302-313.
[25] Valverde, N., Investigations on the rate of dissolution of metal oxides in acidic solutions with additions of
redox couples and complexing agents. Ber. Bunsenges. Phys. Chem., 80(4) (1976): 333-340.
[26] Marcus, R.A., Electron, proton and related transfers. Faraday Discuss. Chem. Soc., 74 (1982): 7-15.
[27] Buxton, G.V., Rhodes, T. and Sellers, R.M., Radiation-induced dissolution of colloidal hematite. Nature,
295 (1982): 583-585.
[28] Baumgartner, E.C., Blesa, M.A. and Maroto, A.J.G., Kinetics of the dissolution of magnetite in
thioglycolic acid solutions. J. Chem. Soc. Dalton Trans., (1982): 1649-1654.
[29] Hidalgo, M.d.V., Katz, N.E., Maroto, A.J.G. and Blesa, M.A., The dissolution of magnetite by
nitrilotriacetatoferrate(II). J. Chem. Soc. Faraday Trans. I, 84(1) (1988): 9-18.
[30] Litter, M.I. and Blesa, M.A., Photodissolution of iron oxides. I. Magnetite in EDTA solutions. J. Colloid
Interface Sci., 125(2) (1988): 679-687.

You might also like