You are on page 1of 18

Minerals Engineering 24 (2011) 1379–1396

Contents lists available at SciVerse ScienceDirect

Minerals Engineering
journal homepage: www.elsevier.com/locate/mineng

Acid leaching of metals from deep-sea manganese nodules – A critical review


of fundamentals and applications
G. Senanayake ⇑
Parker Centre, Faculty of Science and Engineering, Murdoch University, WA 6150, Australia

a r t i c l e i n f o a b s t r a c t

Article history: The mass% metal composition of deep sea nodules ranges from 10–28% Mn, 4–16% Fe, 0.3–1.6% Ni, 0.02–
Received 2 March 2010 0.4% Co, and 0.1–1.8% Cu in mixed oxide matrices with alumina and silica. The concentrations of base
Accepted 8 June 2011 metal ions in sea water of pH  8 of the order 103–102 nmol/kg are shown to be dependent on the sol-
Available online 26 August 2011
ubility products (KSP) of carbonate sediments. Cations of higher softness have higher pKSP and lower sol-
ubility. Previously reported leach results of nodules in H2SO4 and HCl under atmospheric pressure and
Keywords: temperatures in the range 30–90 °C and in the absence or presence of SO2, Na2SO3, NaCl and CaCl2 are
Hydrometallurgy
used in the present study to compare, contrast and rationalise the leaching behaviour of metal values.
Leaching
Sea nodules
Leach efficiency of metals increases with increasing acid concentration, and Cu(II) and Zn(II) follow the
Equilibria same trend in HCl. Potential–pH diagrams of base metal oxides show a higher stability of mixed metal
Kinetic models oxides such as ferrites, magnetite and manganite, which causes partial dissolution of high-valent oxides
in the absence of reducing agents. Application of a shrinking core kinetic model in both H2SO4 and HCl
media predicts a proton diffusivity of Dþ H  10
11
cm2 s1 for the dissolution of Ni from nodules. This
value is of the same order as DHþ for the high pressure acid leaching of Ni from limonitic laterite. A linear
correlation between leaching efficiencies of Fe and Ni appears to be a result of co-dissolution of these
metals from NiOFe2O3 or NiFeOOH. The first order dependence of initial dissolution rates of Cu(II) with
respect to H+ concentration in H2SO4, and the beneficial effect of background chloride ions, suggests a dis-
solution mechanism: CuO ? Cu(OH)Clads/aq ? CuSO4. The Cu(II) ions in solution can also affect Ni(II) dis-
solution from oxide by cation exchange mechanism. The presence of SO2 or Na2SO3 as reducing agents
facilitates the acid leaching of high-valent oxides of Mn and Co and other metals incorporated in the
mixed oxide matrix. Whilst Fe(II) ions formed during the reductive leaching of Fe(III)-oxides accelerate
the dissolution of Mn(III)/(IV) oxides, the resultant Mn(II) ions accelerate the dissolution of high-valent
Co-oxides. Leaching efficiency in HCl increases with temperature. As for the SO2/H2O system thermody-
namic calculations predict a decrease in concentration of H+ and HSO 3 at high temperatures, which
retards leaching efficiency. The SO2/H2O/air leach system enhances metal dissolution due to the produc-
tion of H2SO4 via the transition metal catalysed oxidation of SO2 to H2SO4.
Ó 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1380
2. Review of previous studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1381
2.1. Pyro/hydrometallurgical processing options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1381
2.2. Ions in sea water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1382
2.3. Mechanism of nodule formation and dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
2.4. Nodule leaching in different media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
2.4.1. Acid media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1383
2.4.2. Reductive media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1384
2.5. Kinetic models for leaching base metal oxides relevant to nodules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1384
3. Thermodynamics of oxide dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385

⇑ Tel.: +61 89360 2833; fax: +61 89360 6452.


E-mail address: G.Senanayake@murdoch.edu.au

0892-6875/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mineng.2011.06.003
1380 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

3.1. Higher stability of mixed oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385


3.2. Partial or complete dissolution of mixed oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1387
3.3. Potential–pH diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1388
4. Analysis of leach results and rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1388
4.1. Initial rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1388
4.2. Reaction order and mechanism based on initial rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390
4.3. Kinetic models for Ni(II) dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390
4.4. Factors affecting leach efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1391
4.4.1. Chloride ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1391
4.4.2. Copper(II) ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1392
4.4.3. Particle size and temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1392
4.4.4. Reducing agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1393
5. Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1395
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1395
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1395

1. Introduction The minerals constituting nodules are of fine crystallite size of


approximately 200 m2/g surface area (Heimendahl et al., 1976).
Polymetallic nodules, also called manganese nodules, are rock Most nodules are between 5 cm and 10 cm diameter, but can vary
concretions formed of concentric layers of Fe and Mn hydroxides. in size from fine particles detected under microscopes to large
Deep-ocean polymetal resources including sedimentary nodules, pellets of 20 cm diameter (Lenoble, 2000). They can occur at any
sulphide formations and manganese rich volcanic crust are alter- depth, but the highest concentrations have been found between
native sources of base metals for the rapidly growing demand for 4000 and 6000 m (Lenoble, 2000). Avramov (2005) reported the
manganese for steel production, as the corresponding high-grade renewed interest in practical utilisation of manganese bearing
land based ores are being depleted (Han and Fuerstenau, 1975; sea nodules which occur at acceptable depths of 800 m–1000 m.
Monhemius, 1980; Avramov, 2005; Shen et al., 2007; Biswas The desired average abundance for economical processing is
et al., 2009; Sen, 2010). First discovered in the Arctic Ocean of Sibe- 15 kg/m2 over areas of several tenths of a square kilometre. As a re-
ria in 1868 (Lenoble, 2000), the manganese nodules are character- sult of joint exploration between Russia and India in 2000, the
ised by high porosity and surface area, and contain significant industrial mining of deep-water nodules is expected to be started
levels of valuable metals such as nickel, cobalt, copper, and zinc in 2010 (Avramov, 2005, www.isa.org.jm/files/documents/EN/Bro-
(Table 1) in the form of oxides and hydroxides, together with vary- chures/ENG7.pdf).
ing amounts of aluminium silicate and other minerals. Other ele- In an intensive study of different nodule samples by transmis-
ments which are present in manganese and other phases of sion electron microscopy, Heimendahl et al. (1976) identified
nodules include Mo, Ba, Cd, Ca, Mg, Sb, V and Sr (Sen, 2010). The diffraction spot patterns which matched those of sodium birnessite
amount of Pacific Ocean surface deposit alone has been estimated (Na0.24Mn1.06O20.6H2O), todorokite (Na0.2Ca0.05K0.02Mn(IV)4-
to be greater than 100 billion tons (Heimendahl et al., 1976). The Mn(III)2O123H2O), hydrohausmannite (Mn(IV)Mn(II)2O4), and
estimated hypothetical and speculative cobalt resources that exist maghemite (c-Fe2O3). Kanungo and Jena (1988a,b) reported the
in manganese nodules and crusts on the ocean floor are of the mineralogical composition of Indian Ocean nodules as birnessite,
order of millions of tons, compared to the identified world d-MnO2, amorphous hydrated iron oxide, clays, quartz and zeolites.
cobalt resources of 15 million tons (Georgiou and Papangelakis, Both Transition and Scanning Electron Microscopic (TEM and SEM)
2009). analysis revealed that a major portion of hydrous iron oxide occurs

Table 1
Chemical composition of manganese nodules.

Origin Chemical composition (mass%) References


Mn Fe Ni Co Cu Zna
Pacific Ocean 16.8b 12.5 0.60 0.20 0.40 – Heimendahl et al. (1976)
24.5c 11.5 0.70 1.15 0.25 – Heimendahl et al. (1976)
23.5 15.1 0.80 0.20 0.50 – Han and Fuerstenau (1975, 1976a,b)d
31.3 5.62 1.61 0.14 1.75 – Hsiaohong et al. (1992), Hsiaohong (1996)
13.1 15.8 0.59 0.38 0.11 – Niinae et al. (1996)
27.1 4.12 1.23 0.13 1.21 0.12 Vu et al. (2005)
South-West Pacific Basin 16.6 22.8 0.35 0.44 0.21 – Sen (2010)
Samoan Basin 17.3 19.6 0.23 0.23 0.17 – Sen (2010)
Peru Basin 33.1 7.1 1.40 0.09 0.69 – Sen (2010)
Indian Ocean 10.0 11.4 0.26 0.14 0.23 – Kanungo and Jena (1988a,b), Kanungo and Das (1988), Kanungo (1999a,b)
18.3 7.40 0.90 0.13 0.70 0.08 Jana (1993)
20.1 6.29 0.99 0.10 0.10 0.13 Acharya et al. (1999)
24.0 10.0 1.10 0.14 1.2 – Kumari and Natarajan (2002a,b)
South Sea 27.7 8.92 1.62 0.02 0.10 0.08 Shen et al. (2007)
a
Zinc content was not reported by some authors.
b
Depth 4.9 km.
c
Depth 1.27 km.
d
Also contains molybdenum.
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1381

(a) Ni or Co (mass%) in nodules


2.0
2. Review of previous studies
y = 0.1x - 0.2
1.5 2 2.1. Pyro/hydrometallurgical processing options
R = 0.8

1.0
Pyrometallurgical processing of polymetallic nodules involves
Ni classical nickel/copper smelting technology (Lenoble, 2000; Sen,
0.5 Co 2010):

0.0
(i) drying and calcination in a rotary kiln, reduction in a sub-
0 5 10 15 20 25 30 35 merged electric arc furnace, separation of Mn rich slag and
Mn (mass%) in nodules Fe–Cu–Ni–Co rich alloy,
(ii) refining the alloy to remove Mn and Fe by oxidation and
(b) slagging in a converter, and producing a Ni–Cu–Co matte
Ni or Co (mass %) in nodules

2.0
Ni by adding sulphur,
1.5 (iii) treating hot Mn-rich slag in an electric-arc furnace to elimi-
Co
nate P, Ni, Cu, and Co and much of Fe to produce a Fe–Si–Mn
1.0
alloy.

The treatment of Ni–Cu–Co matte involves grinding, selective


0.5
leaching with Cl2 to eliminate sulphur, ion-exchange separation
and electrolysis of Ni. Only high grade manganese containing
0.0
0 5 10 15 20 materials (>40% Mn) are convertible to ferro-manganese alloy
Fe (mass %) in nodules through direct pyrometallurgical processes (Zhang and Cheng,
2007).
(c) There are several factors of high significance in developing a
Ni or Co (mass %) in nodules

2.0 y = 0.7x + 0.4


2
hydrometallurgical flowsheet for the extraction of metals from
R = 0.7 lean ores such as manganese nodules (Biswas et al., 2009):
1.5

(i) ore grade,


1.0
Ni (ii) cost of energy and chemical consumption,
Co (iii) options for byproduct recovery and relative prices of the
0.5
products,
(iv) need for optimisation of the process to maximise metal
0.0 throughput and minimise operating costs.
0.0 0.5 1.0 1.5 2.0
Cu (mass %) in nodules
The increasing interest in hydrometallurgical treatment of me-
Fig. 1. Trends in mass% composition of metals in nodules (Table 1). tal resources is due to the availability of various low cost options
for leaching, separation, purification and recovery of metal values
interlaminated with MnO2, and a substantial amount occurs in the using aqueous processes. Hydrometallurgical processing of man-
form of discrete spherical grains of amorphous iron oxy-hydroxide ganese nodules involves leaching with NH3, HCl or H2SO4 in the
polymers (FeOOHxH2O). Agarwal and Goodrich (2003) reported presence of other reagents (Lenoble, 2000). For example, the Cup-
that polymetallic nodules are composed of mainly MnO2 and rion process involves a reductive ammoniacal leach of fine slurry
Fe2O3 or hydroxides (40–70%), with relatively minor amounts of ground nodules in an agitated tank at low temperature using
(0.1–1%) of Cu, Ni and Co. According to the International Seabed carbon monoxide gas produced from fuel oil. After counter-
Authority (ISA), such ores exhibit the typical composition: current decantation in a series of thickeners, Co is removed by
27–30% Mn, 1.25–1.5% Ni, and 0.2–0.25% Co as noted in Table 1. sulphide precipitation and Ni and Cu are separated using solvent
The mineralogy and technology of mining as well as processing extraction and recovered by electrolysis (Lenoble, 2000). Recovery
options via pyrometallurgy and/or hydrometallurgy influence the of Mn has been proved to be difficult. The technological details of
economic viability of metal extraction from manganese nodules. pre-treatment of nodules by reduction roasting with gas mixtures
The recovery of nickel and cobalt arises as one of the main issues followed by leaching in ammoniacal medium and separation and
in the processing of manganese nodules. General correlations show recovery of metals via solvent extraction have been described by
that higher nickel contents in nodules are associated with higher Sen (2010).
manganese contents (Fig. 1a), lower iron contents (Fig. 1b) and The acid leach process of crushed nodules by H2SO4 at 180 °C
higher copper contents (Fig. 1c). A systematic study of metal leach- and 1.2 MPa steam pressure utilises Mn(II), produced from the
ing from the mixed base metal oxide system of Mn–Fe–Ni–Co–Cu– pre-reduction of nodules as a reducing agent, in order to enhance
Zn in manganese nodules would be advantageous in understanding cobalt dissolution. The pressure leach residue of Fe and Mn is
the leaching behaviour of other mixed base metal oxide systems, dried, calcined and smelted in an electric-arc furnace to remove
such as nickel laterites and spent catalysts. Recent extensive re- P and some Fe as slag. Precipitation of metal ions from leach li-
views on nickel laterites (McDonald and Whittington, 2008a,b) quor with H2S produces CuS and Co–Ni sulphide, the former is
and spent petroleum catalysts (Zeng and Cheng, 2009a,b) highlight roasted to CuO. At the refinery, Cu metal is recovered via acid
the importance of a review on manganese nodules. This paper aims leach of CuO and electrolysis. The Ni–Co sulphide is treated with
to briefly review the previous reports on pyrometallurgical and Cl2/H2O and separated via solvent extraction to produce CoCl2
hydrometallurgical processing options, acid leach kinetics of metal and Ni metal after electrolysis (Lenoble, 2000). The similarities
oxides in manganese nodules, and to rationalize metal oxide disso- in unit operations for metal recovery from the mixed base metal
lution on the basis of reaction equilibria, potential–pH diagrams oxide matrices of manganese nodules, nickel laterites (Sen, 2010)
and heterogeneous kinetic models. and spent petroleum catalysts highlight the importance of briefly
1382 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

Major constituents
(a) 1000 +
Na Cl
-

-1
Cations

Concentration in sea water / mmol kg


2+
100 Mg Anions
2-
SO4
2+
Ca +
K
10
-
HCO3
-
1 Br 3-
BO3

2+
Sr -
0.1 F

0.01

Cations Anions

(b) 100000

10000
-1

Minor constituents
Ion concentration in sea water / nmol kg

1000

100

10

0.1

0.01

0.001
U Be Pb Hg Cd Cs Ag Cr Co Mn Cu Ti As Al V Fe Zn Mo Ni Ba Rb Li Sr

Fig. 2. Concentrations of dissolved constituents in sea water: (a) major (mmol kg1), (b) minor (nmol kg1) Ref: K.K. Turekian: Oceans. 1968. Prentice-Hall and http://
en.wikipedia.org/wiki/Seawater.

reviewing the solubility of metal ions in sea water. This would Fig. 2a and b shows the relative concentrations of anions and
shed more light on the mechanism of nodule formation and base metal cations in seawater. The concentrations of base metals
hydrometallurgical aspects related to the leaching stage on a dissolved in sea water are extremely low as shown in Fig. 2b
comparative basis with other mixed oxides. For example, the con- (<120 nmol kg1) compared to those of Na+, Mg2+ and Ca2+ ions
centrations of metal ions in seawater are those in equilibrium (0.1–470 mmol kg1), shown in Fig. 2a. It is of interest to note that
with their oxides (Han, 1971) or carbonate mineral sediments the solubility of most sparingly soluble divalent metal ions in sea-
(Morse and Arvidson, 2002). water decreases with the increase in pKSP of relevant metal carbon-
ates (Fig. 3a), as expected from the solubility equilibrium:
2.2. Ions in sea water
MCO3 ¼ M2þ þ CO2
3

The seawater pH scale (pHSWS) is based on the total concentra- K SP ¼ aM2þ aCO2
3
tion of H+ expressed by [H+]SWS  {[H+] + [HSO +
4 ] + [HF]} where [H ] K SP ¼ ðcM2þ mM2þ ÞðcCO2 mCO2 Þ
3 3
represents the free proton concentration (Prieto and Millero,
2002). The high pH of seawater (8) is a result of the involvement logðmM2þ  mCO2 Þ ¼ pK SP  logðcM2þ  cCO2 Þ
3 3

of dissolved CO2 in the two equilibria:


where KSP is the solubility product of the carbonate of the relevant
metal ion (pKSP = log KSP), and the terms a, m and c represent the
CO2 þ H2 O ¼ Hþ þ HCO3 ðK 1 Þ activity, concentration (mol/kg) and activity coefficient of each ion,
HCO3 ¼ Hþ þ CO2 respectively, associated with the solubility equilibrium represent-
3 ðK 2 Þ
ing KSP.
Likewise, for the monovalent metal ion carbonates:
depending on the total carbonate, fugacity of CO2, temperature, and
other factors. For example, the concentration of dissolved CO2 in the M2 CO3 ¼ 2Mþ þ CO2
 3  
form of Na2CO3 and NaHCO3 in seawater is 0.26 mmol/kg and log ðmMþ Þ2  mCO2 ¼ pK SP  log ðcMþ Þ2  cCO2
1.77 mmol/kg respectively (Prieto and Millero, 2002). 3 3
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1383

(a) Marcus (1997) as a function pKSP. Cations of higher softness


log {[M ][CO 3 )]} or log{[M ] [CO 3 ]}
0 Ca
2+
2- 2+
Mg correspond to larger values of pKSP (Fig. 3b) leading to lower
2+ solubility (Fig. 3a).
Sr
+ 2

-10 The negative deviations from the linear relationship of slope 1


+
Li
2+
Hg2 in Fig. 3a can be related to the precipitation of less soluble salts
+ 2+
Rb Be compared to carbonates. For example, the large equilibrium con-
-20 + stant of log K = 7.56 for the reaction:
Ag Slope = - 1
BeCO3 þ OH þ H2 O ¼ BeðOHÞ2 þ HCO3
2-

-30 based on the HSC 6.1 data base (Roine, 2002) indicates that the pre-
cipitation of hydroxide is more favourable than that of carbonate.
2+

Metal ions can also form ion-pairs or complexes with various anions
+
-40
Cs (Yn) in sea water such as Cl, SO2  2 
4 , HCO3 , CO3 and F . Table 2 lists
0 5 10 15 20 the equilibrium constants (K1) for the association between Mz+ and
pK SP of MCO 3 or M2 CO3 the anion Yn: Mz+ + Yn = MY(zn)+. The formation constants for
SrCO03 (b1 = 101.3), SrðCO3 Þ2
2 (b2 = 10
1.6
), SrðCO3 Þ4
3 (b3 = 10
1.70
) and
(b)
þ 1.24
1.5 SrHCO3 (b1 = 10 ) have also been reported by Hogfeldt (1964).
Such associations can lower the metal ion activity coefficients and
1.0 enhance the concentration in sea water. Further discussion of the
Softness of metal ions

scattered solubilities of some metal ions in Fig. 3a is beyond the


0.5 scope of the present study, in which the main focus is on the leach-
ing of manganese nodules in acid media.
0.0

2.3. Mechanism of nodule formation and dissolution


-0.5

Han and Fuerstenau (1975, 1976a,b) reviewed the previous


-1.0
studies and noted different views related to origin and mechanism
-1.5
of genesis of manganese nodules via the formation of metal oxides
0 4 8 12 16 20 in the ocean environment:
pK SP of MCO 3 or M2 CO3
(a) Isomorphic substitution of divalent ‘foreign ions’ for Mn2+ in
Fig. 3. Correlation between reported solubility, softness and pKSP of metal the disordered second layer of the double layer structure of
carbonates of M+ and M2+ ions in seawater: (a) solubility, (b) softness; data from
Fig. 2 and HSC 6.1 database (Roine, 2002) softness data from Marcus (1997).
hydrated manganese-manganite.
(b) Physical adsorption of positively charged metal ions in the
ocean on the negatively charged surface of manganese oxi-
des as a scavenging action.
Table 2 (c) Strong interaction and coagulation of minor oxides (Ni, Cu,
Association constants (log K1) of cations and anions in sea water. or Co) with major oxides (Mn or Fe).
Cation Mz+ Anion Yn
Thus, the change in surface charge in the presence of a suffi-
Cl SO2 HCO
3 CO2 F
4 3
ciently high concentration of protons should desorb and dissolve
Li+ 1.32 0.89 0.75 any chemically bound minor oxides of Ni(II), Cu(II), or Co(II) into
Rb+ 1.02 0.89
the bulk solution. The divalent metal ions substituted in the disor-
Mg2+ 0.41 2.19 1.23 3.29 1.32
Sr2+ 0.24 1.18 2.86 0.14 dered layer can also be removed by displacement via an ion-ex-
Ba2+ 0.06 5.85 3.21 0.38 change type metatheses process with other metal ions (Han and
Ca2+ 0.15 2.11 1.25 4.48 0.68 Fuerstenau, 1976a,b). A brief review of previous studies on nodule
Ag+ 3.30 1.09 0.40
leaching in different acid media in the absence or presence of
Hg2+ 7.43 1.63
reducing agents is presented in the next section.
Log K1 at 25 °C for Mz+ + Yn = MY(zn) from the HSC 6.1 database (Roine, 2002)
Hogfeldt (1964), Smith and Martell (1976). 2.4. Nodule leaching in different media

2.4.1. Acid media


n o n o Kanungo and Das (1988) reviewed important facts in relation to
Thus, a plot of log ðmM2þ Þ  mCO2 or log ðmMþ Þ2  mCO2 as a nodule leaching in different acid media:
3 3

function of pKSP in Fig. 3a shows a linear relationship of slope close


to 1 for Pb2+, Cd2+, Co2+, Mn2+, Cu2+, Fe2+, Zn2+, Ni2+, UO2þ 2 and (i) low metal extraction in dilute sulphuric acid solutions at
Ba2+. This supports the proposed role of dissolved carbonate in con- ambient temperatures and pressures compared to high iron
trolling the concentration of metal ions in equilibrium with car- extraction from amorphous oxy-hydroxide of iron,
bonate mineral sediments in sea water (Morse and Arvidson, (ii) low Co extraction of 30–40% unless a reducing agent was
2002). employed,
The different solubilities of metal ions in sea water in Fig. 2a– (iii) good recoveries of Cu, Ni, and Co during high pressure acid
b can be further examined by considering the concept of soft leaching (HPAL) with dilute H2SO4 (or HCl) at 170–250 °C
and hard acids and bases proposed by Pearson (1973). According under oxygen pressure which rendered the separation of
to this concept metal ions prefer ligands of the same kind (soft– insoluble Fe-oxide, and the drawbacks of HPAL due to the
soft or hard–hard) over those of different kind (soft–hard) when expensive equipment and complications due to the regener-
forming coordinated bonds. Fig. 3b plots the softness reported by ation of excess acid,
1384 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

(iv) complete leaching of Mn and Fe from deep-sea nodules in Table 3 (Hsiaohong, 1996). Higher metal extraction was achieved
concentrated HCl (11 mol dm3) at 90–100 °C releasing Cl2, with 2 mol dm3 HCl compared to the results with 1 mol dm3
which can be used to produce HCl or to oxidise MnCl2, and HCl.
subsequent addition of MgO to remove Mn and produce Reductive leaching of manganese nodules by SO2–H2SO4–
MgCl2 which can be pyrohydrolysed to MgO and HCl. (NH4)2SO4 gave high leach efficiency of Cu, Ni, Co, Zn, and Mn in
the range 88.5–99.6%, compared to low leach efficiency of Fe
Charewicz et al. (2001) achieved almost complete dissolution of (2.4–4.9%) due to precipitation of ammonium jarosite: (NH4)Fe3-
all valuable metals from manganese nodules using 4 mol dm3 HCl (SO4)2(OH)6 (Acharya et al., 1999). Stoichiometric amounts of
at 30 °C. Lower acid concentrations yielded similar results at ele- FeSO4–H2SO4 at 90 °C and solid/liquid (S/L) ratio in the range
vated temperatures. Kanungo and Jena (1988b) observed the leach- 1/15–1/5 resulted in greater than 90% leach efficiency of Ni, Cu
ing of only Cu(II), Ni(II), and Co(II) in dilute HCl, without bringing and Mn (Vu et al., 2005). Excess H2SO4 had no influence on metal
much iron into solution, and the need for concentrated HCl and ele- leach efficiency, but 60% excess acid dissolved 85% of metal. Thus,
vated temperature for the leaching of Co associated with Mn. the recommended conditions were: 90 °C, stoichiometric FeSO4,
1.6-fold excess H2SO4, and S/L ratio 1/7.
Some of the results from leaching studies with numerous re-
2.4.2. Reductive media agents published over the past 3–4 decades summarised in this
The reductive leaching of manganese nodules in 1.5 mol dm3 section present an opportunity for a systematic and comparative
HCl is possible with reagents such as pyrite, sodium sulphite, and interpretation of leaching data based on reaction equilibria, speci-
carbon at ambient pressures (Kanungo and Jena, 1988b). Relatively ation, potential–pH relationships, rates and heterogeneous kinetic
high extraction of Ni, Co, Zn, Mn, and Cu compared to Fe was a models involved in the dissolution of mixed base metal oxides in
main advantage in reductive leaching. Zhang et al. (2001) reviewed manganese nodules in different acid media. The analysis of reac-
previous studies on reductive leaching of manganese nodules by tion rates as a function of concentration of reagents, temperature,
various organic and inorganic reagents including HCHO, NH3OHCl, particle size or surface area according to various kinetic models
phenols, NO2, SO2, Na2SO3, H2O2, CH3OH, FeS2, coal, sucrose, glu- would reveal useful information on reaction orders with respect
cose, lignite, sawdust (cellulose) and nickel matte in different med- to relevant reagents, activation energies, rate controlling step(s)
ia such as water, dilute solutions of H2SO4 or HCl, NH3/NH4Cl or and reaction mechanism(s). This in turn may allow the investiga-
(NH4)2CO3. Pyrrhotite can also facilitate the reductive leaching of tion of novel leaching agents in future studies.
manganese nodules in H2SO4 (Shen et al., 2007), where the two
products H2S and FeSO4 released by the dissolution of FeS can
2.5. Kinetic models for leaching base metal oxides relevant to nodules
act as reducing agents. The leaching of metals from nodules in
weakly acidic solutions of pH 4.5 is also facilitated by the presence
Table 4 lists the relevant equations for various heterogeneous
of a fungus – Aspergillus niger. The organic acids such as oxalic and
kinetic models which are applicable for particle leaching (Leven-
citric acids released in this process reduce the host metal oxides/
spiel, 1972). A linear relationship of each function (fX) of fraction
hydroxides and provide the complexing oxalate/citrate ligands to
of metal dissolved (X) in Table 4 has been used to examine the
dissolve about 97% Cu, 98% Ni, 86% Co, 91% Mn and 36% Fe over
validity of a given model. The activation energy (Ea), determined
30 days (Mehta et al., 2010). Electroleaching of manganese nodules
using the Arrhenius equation indicates the chemical or diffusion
in 1 mol dm3 H2SO4 at potentials between 0.6 V and 1.4 V
controlled nature of leaching reaction(s). Han and Fuerstenau
(SCE) causes significant increase in dissolution of Cu, Ni and Co
(Kumari and Natarajan, 2001a,b, 2002a,b). Eletrobioleaching in
the presence of Thiobacillus ferrooxidans and Thiobacillus thiooxi- Table 4
dans at pH 0.5 and 0.6 V (SCE) results over 90% dissolution of Mathematical expressions for rate analysis.
Mn, Fe, Cu, Ni and Co (Kumari and Natarajan, 2001b). Equations Mathematical expressions
The reductive leaching tests for manganese nodules in the pres-    
(1a) t 1=3
¼ bkC
s ¼ 1  ð1  X B Þ qB r 0 t ¼ kapparent t
A
ence of nickel matte in hydrochloric acid produced elemental sul-  
phur and gave leaching efficiencies: 90% Ni, 90% Cu, 100% Mn (1b) 0:5
8:7109 M exp ð27;600=RT ÞðC Fe2þ Þ ðC Hþ Þ
0:5

1  ð1  X B Þ1=3 ¼ qdo t
and >80% Co, in 1.5–2.0 mol dm3 HCl at 95 °C after 80 min (Hsiao-  
(2a) t 2=3 6bDA C A
hong et al., 1992). Jana (1993) also achieved high metal extraction s ¼ 1  3ð1  X B Þ þ 2ð1  X B Þ ¼ q r2 t ¼ ðkapparent Þt
B 0

from nodules even in the presence of dilute HCl by adding NaCl or (2b) 6bD
logðkapparent Þ ¼ log ðC Hþ Þ þ log qrH2 þ
  o

Na2S which caused the evolution of gaseous Cl2 and H2S respec- (3) t 2=3
¼ 2bD A CA
s ¼ 1  ð1  X B Þ qB r 20 t
tively. The evolution of both gases could be reduced, without ad- (4) 1=2
 
t
s ¼ 1  ð1  X B Þ ¼ aCrA2 t
versely affecting the metal extraction, by adding a small amount  2 0
of Na2S. Further studies with nickel matte as a reducing agent (5a) 1  NANðt¼1Þ
A ðtÞ
¼ exp p4L2DA t
showed the formation of a number of mineral phases listed in (5b) ln(1  X) = kapparentt
E 
(6) k ¼ A exp a
RT

Table 3 b = stoichiometric factor of reaction A(aq) + bB(solid) ? products; XB = fraction of B


Solid phases of reductive acid leach residues of manganese nodules with nickel matte reacted after time t; s (s) = time for complete dissolution of B from particle; ro
and HCl. (m) = initial particle radius, do (m) = initial particle diameter; qB (mol m3) = molar
density of B, k (m s1) = intrinsic rate constant of surface reaction or mass transfer
Oxide or sulphide Formula coefficient of A in fluid film; D = diffusivity of A through product layer (m s2) or
pores within the ore particle; kapparent (s1) = apparent rate constant; CA
Single oxides Fe2O3
(mol m3) = concentration of A. Ea = activation energy; T = absolute temperature;
Mixed-oxides NiFe2O4, CuFeO2, CuFe2O4, Cu6Fe3O7
R = gas constant; a = proportionality constant; Eq. (1). Shrinking sphere (without a
Single sulphides NiS, CuS, CuS2, FeS2, Cu9S, Ni3S2, Ni3xS2, Cu7.2S4
solid product on surface); Eq. (2). Shrinking core (diffusion of rectantas/products
Mixed sulphides (CuFeNi)S2, Fe8Ni8S16, (FeNi)3S4, (FeNi)S2,
through a thickening solid product layer on surface); Eq. (3). Small particles; Eq. (4).
Cu5FeS4, (FeNi)S8
Large particles; Eq. (5). Porous particles (diffusion of products within pores); Eq. (6)
Mixed sulphides-oxides Fe6S8O33, Fe4S5O2
Arrhenius equation; References: Levenspiel (1972), Bayramoglu and Tekin (1993)
Other (FeNi), S
(Eq. (1b)), Han and Fuerstenau (1975) (Eq. (5a)), Chiarizia and Horwitz (1991) (Eq.
Based on XRD (Hsiaohong, 1996). (5b)).
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1385

(1976a,b) derived a kinetic model for a first order reaction with re- the dissolution of Ni(II) from manganese nodules in HCl at
spect to the surface concentration (mol cm2) of Cu(II) and Ni(II); 40–75 °C (Kanungo and Jena, 1988a), and Ni(II) and Co(II) from
the dissolution rate was controlled by the surface reaction at pH limonitic laterite in H2SO4 at high pressure and temperature
1 for fine particles of an average diameter of 1 lm or less. In the (230–270 °C) (Georgiou and Papengelakis, 1998, 2009) or in
case of coarse particles, both the surface reaction and pore diffu- H2SO4–SO2 at atmospheric pressure and 90 °C (Senanayake and
sion were rate controlling for Ni(II), whereas pore diffusion alone Das, 2004).
was rate controlling for Cu(II) with a diffusivity A grain size finer than 350 lm has little effect on the leaching of
DCu(II) = 3  1010 cm2 s1. This estimate was based on the assump- metal values from manganese nodules of Indian Ocean origin (Kan-
tion that the pores in the nodules are cylindrical with an average ungo and Jena,1988a). Kinetic studies revealed that the leaching of
pore radius of ro, and average pore length of L, so that most of Mn and Co in HCl followed a parabolic rate law, due to the fact that
the surface area (A) originates from the pores: A = 2proLn = 4/ both birnessite and d-MnO2 occur in the form of thin plate shaped
3pR3qaS, where qa = apparent density of particles, S = specific sur- crystallites in manganese nodules. Thus, the dissolution of Fe as
face area (m2/g), R = radius of a circle having the same area as the well as Ni was controlled by the diffusion of H+ through an insolu-
projected area of the particle, and n = total number of pores in one ble product layer according to a shrinking core kinetic model (Eq.
particle of nodule. The estimated value of ro, independent of the (2) in Table 4). However, the dissolution of Cu and Zn demon-
particle size, was 4 nm with the corresponding values of strated large initial rates and followed a random nucleation model
S = 200 m2/g, qa = 1.5 g/cm3, and porosity = 60%. Average pore given by the equation: ln(1  X) = (kt)m, where X represents the
length turned out to be one-third of the radius of the particle. fraction of metal dissolved after time t, and k and m are constants.
The calculated values of low Cu2+ diffusivity of 3  1010 cm2 s1 These findings warrant further comparative studies to rationalise
through porous media based on Eq. (5a) in Table 4, compared to the thermodynamic stability and acid leaching behaviour of mixed
7  106 cm2 s1 in aqueous media, was consistent with previous base metal oxides such as manganese nodules and laterite ores.
findings (Han and Fuerstenau, 1976a,b).
The differences between rate controlling steps of metal dissolu-
tion expressed by various equations in Table 4 were also reflected 3. Thermodynamics of oxide dissolution
by the observed effect of temperature. For example, Chiarizia and
Horwitz (1991) employed Eq. (5b) to model the Fe(III) dissolution 3.1. Higher stability of mixed oxides
from synthetic goethite and reported kapparent  2 h1 in H2SO4 at
80 °C. This value is larger compared to 0.2–0.7 h1 for the dissolu- Table 5 lists the values of equilibrium constants (K) in
tion of Cu(II) at pH 1–2 and 25 °C from nodules reported by Han ascending order for different types of reactions of various single
and Fuerstenau (1976a,b). The rate of Ni(II) dissolution corre- or mixed base metal oxides. Large values of K for the acid disso-
sponded to a high activation energy, indicating a chemically con- lution according to reactions B4–B9 show that there are no sol-
trolled surface reaction (Han and Fuerstenau, 1976a,b). However, ubility limitations for the acid leaching of most divalent metal
rates of Cu(II) and Co(II) dissolution were independent of particle oxides. The ascending order of K in reactions A1–A19 shows that
size finer than 200 lm and demonstrated a low activation energy; the formation of an oxide of mixed valencies (M(II)/(III)) in the
with rates decreasing with the increase in size over 200 lm. It was form of manganite, cobaltite, magnetite or ferrites in the reverse
concluded that the addition of acid was essential for the dissolu- direction is favoured. Thus, Mn3O4 and Fe3O4 undergo only par-
tion of metal ions chemically bound, rather than physically ad- tial dissolution in acid according to reactions C13–C17 to pro-
sorbed, to the pore surfaces of the two major constituents MnO2 duce M(II) in solution phase and M(III)/(IV) in solid phase,
and FeOOH. Although the addition of CaCl2 and NaCl had no influ- with favourable values of K of the order 1010–1020. In contrast,
ence at pH 6 and 25 °C, the dissolution of Cu(II), Ni(II), Co(II), was partial acid dissolution of Co3O4, CoFe2O4 and MnOFe2O3 accord-
enhanced by these salts at pH 3. This shows the influence of Cl ing to reactions C1–C4 is less favourable due to low values of K
ions on the dissolution of Cu(II) and Co(II) as well as a cation-ex- of the order 104–100.1 (Table 5).
change mechanism for surface reaction, and warrants further The monovalent Cu+ ion is unstable in aqueous media in the ab-
studies. sence of complexing ligands, as shown by the low equilibrium con-
A shrinking particle model (Eq. (1) in Table 4) is applicable for stant for the reaction: Cu2+ + Fe2+ = Cu+ + Fe3+ (K298 K  1010).
the dissolution of metals in a number of different oxide systems: However, it can exist in the solid form of ferrite CuFeO2 as revealed
(i) Cu(II) from a-Al2O3/CuO(10%) in H2SO4 and HCl at 30–50 °C by the potential–pH diagram in Fig. 4 and the favourable equilib-
(Habbache et al., 2009), (ii) Mn(II) from b-MnO2 in acidic FeSO4 rium constants (105) for reverse reactions in A16 and A17. In
solutions at 25 °C–55 °C (Bayramoglu and Tekin, 1993), and (iii) the laboratory, zinc ferrite ZnFe2O4 can be synthesised by heating
Fe(II) from FeOOH by SO2/H2SO4 at 90 °C (Senanayake, 2003a,b, a mixture of ZnO and Fe2O3 to 900 °C (Langova et al., 2009); like-
2004; Senanayake and Das, 2004). Further analysis of results for wise the preparation of CuFeO2, a delafossite-type compound
MnO2 and FeOOH showed the applicability of the mixed potential Cuþ Fe3þ O2
2 , requires the heating of a mixture of Cu2O and Fe2O3
model. For example, the half cell reactions of reduction of MnO2/H+ to 1000 °C (Barnabe et al., 2006). In contrast, the divalent Cu2+ ions
to Mn2+/H2O and oxidation of Fe2+ to Fe3+ lead to the rate expres- tend to exist in the solid forms CuO or Cu(OH)2 along with Fe2O3 or
sion in Eq. (1b) (Table 4) based on the mixed potential model FeOOH, compared to the unstable ferrite form CuOFe2O3, as de-
(Bayramoglu and Tekin, 1993). Miller and Wan (1983) developed scribed by the favourable value of K  101.29–102.65 for reactions
a similar expression for the reductive leaching of electrolytically A21–A25 in Table 5. According to the predicted K values in Table 5,
produced MnO2 by SO2. Kinetic models for the dissolution behav- the relative stabilities of ferrites follow the ascending order
iour of Fe3O4 (Byerley et al., 1979), and FeOOH (Warren and Hay, CuOFe2O3 < CuFeO2  Cu2OFe2O3 < NiFe2O4 < NiOFe2O3. How-
1975; Kumar et al., 1993) under reductive conditions have also ever, the XRD analysis of reductive leach residues of manganese
been reviewed and rationalised on the basis of actual chemical spe- nodules with nickel matte in HCl revealed the presence of both
cies involved in the surface reaction (Senanayake, 2003a,b). CuFeO2, CuFe2O4 as well as NiFe2O4 in the residue (Table 3).
The dissolution of Cu(II) by H2SO4 in porous oxide matrices of The low values of K for reactions A21–A25 in Table 5 also sug-
nodule particles is controlled by the pore diffusion of Cu(II), as gest the relatively low stability of CuO in a mixed-oxide matrix.
found by Han and Fuerstenau (1975, 1976a,b). A shrinking core This generally predicts the possibility of selective dissolution of
with a product layer given by Eq. (2) in Table 4 is applicable for Cu(II) from a mixed oxide matrix with less interference from other
1386 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

Table 5
SO2 represents SO2(aq); data at 298 K and 363 K; values in parentheses are for 523 K; references same as in Table 2.

Reaction categories Log K298 K Log K363 K

A. Relative stability of individual and mixed oxides


A1. Co3O4 + 3H2O = CoO + 2Co(OH)3 17.5 15.6
A2. Mn3O4 = 2MnO + MnO2 15.8 12.8
A3. CoFe2O4 + H2O = CoO + 2FeOOH 13.5 11.4
A4. 2MnOOH = MnO + MnO2 + H2O 11.7 9.12
A5. CoFe2O4 + Fe2+ = Co2+ + Fe3O4 9.93 8.16
A6. Fe2O3 + 3H2O = 2Fe(OH)3 7.24 6.84
A7. Mn3O4 = MnO + Mn2O3 6.87 5.51
A8. Fe3O4 = FeO + Fe2O3 4.48 3.69
A9. Fe3O4 + H2O = FeO + 2FeOOH 4.43 4.02
A10. Mn3O4 + H2O = MnO + 2MnOOH 4.05 3.70
A11. MnOFe2O3 + H2O = MnO + 2FeOOH 3.92 3.65
A12. NiOFe2O3 + H2O = NiO + 2FeOOH 3.79 3.50
A13. NiFe2O4 + H2O = NiO + 2FeOOH 3.51 3.16
A14. ZnMn2O4 = ZnO + Mn2O3 4.63 3.59
A15. Cu2OFe2O3 + H2O = Cu2O + 2FeOOH 5.50 4.86
A16. 2CuFeO2 + H2O = Cu2O + 2FeOOH 5.50 5.50
A17. 2CuFeO2 = Cu2O + Fe2O3 5.55 4.53
A18. ZnFe2O4 + H2O = ZnO + 2FeOOH 2.11 2.31
A19. ZnFe2O4 = ZnO + Fe2O3 2.16 1.98
A20. Fe2O3 + H2O = 2FeOOH 0.05 0.33
A21. CuOFe2O3 = CuO + Fe2O3 1.29 0.90
A22. CuO + H2O = Cu(OH)2 1.31 1.00
A23. CuOFe2O3 + H2O = CuO + 2FeOOH 1.34 0.57
A24. CuOFe2O3 + H2O = Cu(OH)2 + Fe2O3 2.60 1.90
A25. CuOFe2O3 + 2H2O = Cu(OH)2 + 2FeOOH 2.65 1.57
A26. Fe(OH)3 = FeOOH + H2O 3.64 3.26
A27. Fe(OH)3 = 0.5Fe2O3 + 1.5H2O 3.62 3.42
B. Complete dissolution of single oxides
B1. FeOOH + 3H+ = Fe3+ + 2H2O 0.39 1.52
B2. Fe2O3 + 6H+ = 2Fe3+ + 3H2O 0.83 3.38
B3. Fe(OH)3 + 3H+ = Fe3+ + 3H2O 4.03 1.73
B4. CuO + 2H+ = Cu2+ + H2O 7.71 5.69
B5. ZnO + 2H+ = Zn2+ + H2O 11.2 8.45
B6. NiO + 2H+ = Ni2+ + H2O 12.5 9.33
B7. CoO + 2H+ = Co2+ + H2O 13.6 10.2
B8. FeO + 2H+ = Fe2+ + H2O 14.5 11.1
B9. MnO + 2H+ = Mn2+ + H2O 17.9 14.1
C. Partial dissolution of single or mixed oxides
C1. Co3O4 + 2H+ + 3H2O = Co2+ + 2Co(OH)3 + H2O 3.86 5.33
C2. CoFe2O4 + 2H+ = Co2+ + Fe2O3 + H2O 0.06 0.80
C3. CoFe2O4 + 2H+ = Co2+ + 2FeOOH 0.11 1.13
C4. 2MnOOH + 2H+ = Mn2+ + MnO2 + 2H2O 6.16 5.02
C5. CoOFe2O3 + 2H+ = Co2+ + 2FeOOH 7.90 5.19
C6. NiOFe2O3 + 2H+ = Ni2+ + 2FeOOH 8.69 5.83
C7. Mn2O3 + 2H+ = Mn2+ + MnO2 + H2O 8.99 6.83
C8. ZnFe2O4 + 2H+ = Zn2+ + Fe2O3 + H2O 9.06 6.47
C9. ZnFe2O4 + 2H+ = Zn2+ + 2FeOOH 9.10 6.14
C10. CuOFe2O3 + 2H+ = Cu2+ + 2FeOOH 9.05 6.27
C11. CuOFe2O3 + 2H+ = Cu2+ + Fe2O3 + H2O 9.00 6.59
C12. ZnMn2O4 + 2H+ = Zn2+ + 2MnOOH 9.42 6.66
C13. Fe3O4 + 2H+ = Fe2+ + Fe2O3 + H2O 10.0 7.37
C14. Fe3O4 + 2H+ = Fe2+ + 2FeOOH 10.1 7.04
C15. Mn3O4 + 2H+ = Mn2+ + 2MnOOH 13.9 10.4
C16. MnOFe2O3 + 2H+ = Mn2+ + 2FeOOH 14.0 0.49
C17. Mn3O4 + 4H+ = 2Mn2+ + MnO2 + 2H2O 20.0 15.5
D. Complete dissolution of ferrites
D1. CoFe2O4 + 8H+ = Co2+ + 2Fe3+ + 4H2O 0.89 4.18
D2. CuFeO2 + 4H+ = Cu2+ + Fe2+ + 2H2O 7.09 4.43
D3. NiOFe2O3 + 6H+ = Ni2+ + 2Fe3+ + 6H2O 9.47 2.78
D4. NiFe2O4 + 6H+ = Ni2+ + 2Fe3+ + 6H2O 9.76 3.12
D5. CuOFe2O3 + 6H+ = Cu2+ + 2Fe3+ + 6H2O 9.83 3.22
D6. ZnFe2O4 + 8H+ = Zn2+ + 2Fe3+ + 4H2O 9.89 4.43
E. Cation-exchange dissolution
E1. CoFe2O4 + Mn2+ = Co2+ + MnFe2O4 13.9 11.6
E2. Co3O4 + Cu2+ + 3H2O = Co2+ + CuO + 2Co(OH)3 11.6 11.0
E3. Co3O4 + Cu2+ + 4H2O = Co2+ + 2Co(OH)3 + Cu(OH)2 10.3 10.0
E4. CoFe2O4 + Cu2+ = Co2+ + CuOFe2O3 8.94 7.39
E5. CoFe2O4 + Cu2+ + H2O = Co2+ + CuO + 2FeOOH 7.61 6.83
E6. NiOFe2O3 + Cu2+ = Ni2+ + CuOFe2O3 0.36 0.43
E7. NiOFe2O3 + Cu2+ + H2O = Ni2+ + CuO + 2FeOOH 0.97 0.13
E8. Fe3O4 + Cu2+ = Fe2+ + CuOFe2O3 1.05 0.77
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1387

Table 5 (continued)

Reaction categories Log K298 K Log K363 K

E9. Fe3O4 + Cu2+ + 2H2O = Fe2+ + 2FeOOH + Cu(OH)2 3.69 2.34


E10. NiO + Cu2+ = Ni2+ + CuO 4.77 3.64
E11. CoO + Cu2+ = Co2+ + CuO 5.85 4.54
E12. FeO + Cu2+ = Fe2+ + CuO 6.82 5.36
E13. MnO + Cu2+ = Mn2+ + CuO 10.2 8.44
F. Redox dissolution of Cu–Fe-oxides
F1. Cu2+ + Fe2+ = Cu+ + Fe3+ 10.4 8.82
F2. CuFeO2 + 2H+ = Cu2+ + FeO + H2O 7.44 6.62
F3. CuFeO2 + 2H+ = Fe(OH)2 + Cu2+ 5.85 5.71
F4. CuFeO2 + 2H+ = Cu(OH)2 + Fe2+ 0.69 0.26
F5. CuFeO2 + 4H+ = Cu2+ + Fe2+ + 2H2O 7.09 4.43
F6. 2FeOOH + SO2 + 2H+ = 2Fe2+ + SO2 21.5 13.9
4 + 2H2O
F7. Fe2O3 + SO2 + 2H+ = 2Fe2+ + SO2 21.5 13.6
4 + H2O
F8. CuOFe2O3 + SO2 = CuFeO2 + Fe2+ + SO2
4
23.4 15.7
F9. Fe3O4 + SO2 + 4H+ = 3Fe2+ + SO2 31.6 20.9
4 + 2H2O

G. Redox dissolution of Mn–Fe-oxides


G1. MnOFe2O3 + 4H+ = MnO2 + 2Fe2+ + 2H2O 0.36 1.91
G2. 2MnOOH + 2H+ = MnO2 + Mn2+ + 2H2O 6.16 5.02
G3. Mn2O3 + 2H+ = Mn2+ + MnO2 + H2O 8.99 6.83
G4. MnOOH + Fe2+ = Mn2+ + FeOOH 10.3 8.71
G5. MnOFe2O3 + 2H+ = 2FeOOH + Mn2+ 14.0 10.5
G6. MnO2 + 2Fe2+ + 4H+ = Mn2+ + 2Fe3+ + 2H2O 15.2 9.35
G7. Mn3O4 + 2Fe2+ + H2O = 2Mn2+ + MnO + 2FeOOH 16.5 13.7
G8. Mn3O4 + 2Fe2+ = 2Mn2+ + MnOFe2O3 20.4 17.4
G9. Mn2O3 + 2Fe2+ + H2O = 2Mn2+ + 2FeOOH 23.4 19.2
G10. MnO2 + SO2 = Mn2+ + SO2
4
35.8 26.3
G11. MnOFe2O3 + SO2 + 4H+ = Mn2+ + 2Fe2+ + SO2 35.5 24.4
4 + 2H2O
G12. 2MnOOH + 2H+ + SO2 = 2Mn2+ + SO2 42.0 31.3
4 + 2H2O
G13. Mn2O3 + SO2 + 2H+ = 2Mn2+ + SO2 44.9 33.1
4 + H2O
G14. Mn3O4 + 5SO2 + 2H2O = 3Mn2+ + SO2 
4 + 4HSO3
48.8 31.0

H. Redox dissolution of Co–Ni–Fe-oxides


H1. Co3O4 + 3Mn2+ = 3Co2+ + Mn3O4 5.68 5.40 (5.15)
H2. Co3O4 + Mn2+ + 4H+ = 3Co2+ + MnO2 + 2H2O 14.4 10.1 (3.71)
H3. Co3O4 + Mn2+ + 2Fe2+ = 3Co2+ + MnOFe2O3 14.7 11.9 (8.14)
H4. CoFe2O4 + 4H+ + SO2 = Co2+ + 2Fe2+ + SO2 21.6 12.8
4 + 2H2O
H5. NiOOH + Fe2+ = Ni2+ + FeOOH 22.2 18.0
H6. Co3O4 + 2SO2 + 3H+ = 3Co2+ + SO2  48.5 33.7
4 + HSO3 + H2O
H7. Co3O4 + 5SO2 + 2H2O = 3Co2+ + SO2 
4 + 4HSO3
43.15 25.6

J. Redox dissolution of Zn–Fe–Mn-oxides


J1. MnO2 + 4H+ + 2Cl = Mn2+ + Cl2 + 2H2O 6.01 4.40
J2. ZnFe2O4 + 4H+ + SO2 = Zn2+ + 2Fe2+ + SO2 30.6 20.0
4 + 2H2O
J3. 3ZnFe2O4 + 4H+ + SO2 = 3Zn2+ + 2Fe3O4 + SO2
4 + 2H2O
28.6 18.2
J4. ZnMn2O4 + 4H+ + SO2 = Zn2+ + 2Mn2+ + SO2 51.4 38.0
4 + 2H2O

L. Other reactions
L1. 1.5MnO2 + Cr3+ + H2O = 1.5Mn2+ + H2CrO4 7.41 2.98 (2.12)
L2. SO2 + H2O = H+ +HSO3 1.77 2.70
L3. Ni2+ + Cl = NiCl+ 1.01 0.64
L4. Fe2+ + Cl = FeCl+ 0.16 0.14
L5. Zn2+ + Cl = ZnCl+ 0.20 1.76
L6. Mn2+ + Cl = MnCl+ 0.30 1.08
L7. Cu2+ + Cl = CuCl+ 0.38 0.66
0 0.78 0.34
L8. Cu2+ + 2Cl = CuCl2
L9. Co2+ + Cl = CoCl+ 0.57 0.75
L10. Fe3+ + Cl = FeCl2+ 1.48 1.79
L11. Co3O4 + 8H+ + 2Cl = 3Co2+ + Cl2 + 4H2O 9.56 7.59
L12. 2Co(OH)3 + 6H+ + 2Cl = 2Co2+ + Cl2 + 6H2O 13.4 12.9
L13. 2SO2 + O2 + 2H2O = 2H2SO4 56.2 41.2

oxides; the prediction is indeed supported by experimental obser- product layer on the surface retards the reaction as described by a
vations presented in later sections. shrinking core model (Eq. (2) in Table 4). This highlights the need
for a reducing agent to dissolve the Fe–Mn–Co oxides of high
3.2. Partial or complete dissolution of mixed oxides valency M(III)/(IV) and to facilitate the complete dissolution of
other metals associated with manganese nodules in acid. For
The partial dissolution of any of the metals from ferrite in acid is example, the value of K for the acid dissolution of CoFe2O4 in-
likely to produce the insoluble phases FeOOH, Fe2O3, or MnOOH as creases from <101 in the absence of a reducing agent (reactions
shown by some of the reactions (C3–C17) in Table 5. An insoluble C2–C3), to 1022 in the presence of SO2 as a reducing agent (reaction
1388 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

Zn - Cu - Fe - H2O - System at 25.00 C Ni - Fe - Mn - H2O - System at 25.00 C


1.0 1.6 NiOOH
Fe(+3a) FeOOH Cu(OH)2 ZnFe2O4 1.4
0.8
1.2 Fe(+3a) NiO.Fe2O3 MnO4(-a)
Cu(+2a)
0.6 1.0 FeOOH
0.8 Fe(+2a)
0.4 CuFeO2 MnO2
Cu
0.6

Eh (Volts)
ZnO 0.4
Eh (Volts)

0.2
0.2 NiO.Fe2O3
MnOOH
0.0 FeOOH
CuFeO2 0.0 Ni(+2a)
NiO(a) Mn3O4
-0.2 -0.2
Ni
Fe(+2a) -0.4 Fe3O4 MnO.Fe2O3
ZnFe2O4
-0.4 Fe
-0.6
Fe Fe(OH)2
-0.6 Fe3O4
-0.8
Zn(+2a) ZnO
-1.0 Mn(+2a) Mn(OH)2
-0.8 Zn Fe(OH)2
-1.2 Mn
-1.0 -1.4
0 2 4 6 8 10 12 14 -2 0 2 4 6 8 10 12 14
pH C pH

Fig. 4. Potential–pH diagram for Zn–Cu–Fe–H2O system at 25 °C and unit activity Fig. 6. Potential–pH diagram for Ni–Mn–Fe–H2O system at 25 °C and unit activity
(based on HSC 6.1, Roine, 2002, dashed lines represent O2/H2O and H2O/H2 couples). (based on HSC 6.1, Roine, 2002, dashed lines represent O2/H2O and H2O/H2 couples).

region of NiOFe2O3. Trevorite (NiFe2O4) does not appear in Fig. 6,


due to the higher stability of NiOFe2O3, as revealed by the pre-
Mn - Co - Fe - H2O - System at 25.00 C dicted equilibrium constant of K  2 for the conversion Ni-
1.6
Co(OH)3 Fe2O4 ? NiOFe2O3 based on the HSC 6.1 database (Roine, 2002).
1.4
1.2 CoFe2O4
Co3O4 In the case of Co–Mn–Fe mixed oxide system, both FeOOH,
1.0 Fe(+3a) MnO4(-a) MnOOH, as well as the mixed oxides such as Mn3O4, Co3O4 and Co-
FeOOH MnO2
0.8 Fe2O4 appear in the potential–pH diagram in Fig. 5.
0.6 Fe(+2a) The divalent Fe(II) and Mn(II) produced by partial acid dissolu-
MnOOH tion of Fe3O4 and Mn3O4 (reactions C13–C15) can be involved in
Eh (Volts)

0.4
0.2 the reductive leaching of MnO2 and Co3O4 as shown by Fig. 5. For
Mn3O4
0.0 Co(+2a) CoFe2O4 example, the low Eh of the redox couple Fe3+/Fe2+ compared to
-0.2
Co
H+–MnO2 /H2O–Mn2+ leads to a large K value for reaction G6. Like-
-0.4 wise, reactions G4, G6, G7, G8 and H5 have favourable K values.
Fe
-0.6 The potential–pH diagram for Co–Mn–H2O published by Zhang
-0.8
et al. (2002) demonstrates the presence of CoO2, Co3O4, Co(OH)3,
-1.0 Mn(+2a) Mn(OH)2
Co(OH)2, MnO2, Mn2O3, Mn3O4 and Mn(OH)2, and the possibility
-1.2 Mn
of reduction of Co(OH)3 by Mn(II). This view is supported by the
-1.4
-2 0 2 4 6 8 10 12 14 large K values of reactions H2 and H3.
C pH Thus, Figs. 4–6 provide more information than the prelimin-
ary potential–pH diagrams which represented only some of the
Fig. 5. Potential–pH diagram for Mn–Cu–Fe–H2O system at 25 °C and unit activity
(based on HSC 6.1, Roine, 2002, dashed lines represent O2/H2O and H2O/H2 couples). oxides in manganese nodules (e.g. Vu et al., 2005). Mehta
et al. (2010) observed a pH change of 6.8–4.5 and Eh change
of 0.14–0.37 V during a 30 day leach period of nodules by Asper-
H4). Likewise, the complete dissolution of single or mixed oxides gillus niger. These Eh–pH ranges coincide with the stability re-
FeOOH, Fe2O3, Fe3O4, NiFe2O4, MnOOH, MnO2, MnOFe2O3 is also gion of Fe2+ and Mn2+ in Figs. 4 and 5, indicating the reductive
facilitated by SO2 as a reducing agent according to reactions F6, leaching of highvalent Mn-oxides. However, Figs. 4–6 do not
F7, F9, G10–G14, H4, H6, and H7. Although the presence of both show the actual soluble complex species of metal ions formed
SO2 and H+ is required for the dissolution of CoFe2O4 (reaction by the interaction between metal cations and anions such as
H4), the two manganese oxides MnO2 or Mn3O4 need only SO2 SO2  2 2
4 , Cl , SO3 (Table 5) and C2 O4 reported in some of the pre-
for complete dissolution (reactions G10 and G14). Moreover, viously published potential–pH or species distribution diagrams
Fe(II) produced from Fe(III)-oxides can facilitate the complete (Senanayake, 2003a,b; Senanayake and Das, 2004; Chandra
dissolution of MnOOH and MnO2 according to reactions G4 and Jeffrey, 2005). Nevertheless, information from Figs. 4–6
and G6. In the case of Co3O4, the reaction with Mn(II) can lead and Table 5 can be used to rationalise experimental data related
to partial dissolution due to blockage by MnO2 in reaction H2; to the leaching of manganese nodules, described in later
but the presence of SO2 facilitates the complete dissolution of sections.
Co–Mn-oxides in nodules according to reactions G10, H4, H6
and H7.
4. Analysis of leach results and rates

3.3. Potential–pH diagrams 4.1. Initial rates

The potential–pH diagrams in Figs. 4–6 show large stability Fig. 7 shows leach results of nodules in H2SO4 at pH 2 (Fig. 7a)
areas of CuFeO2, ZnFe2O4, CoFe2O4, NiOFe2O3 and MnOFe2O3 at and the effect of pH (Fig. 7b) in the range 0.5–3.0 (Han and Fuerstenau,
the high pH of 8, being close to the pH of sea water. Fig. 6 also 1975). The initial rates at pH 2 based on fraction of metal leached
shows a small area of NiO at low pH compared to the large stability expressed by dXM/dt in Fig. 8a follow the order: Cu (0.25) > Co
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1389

(a) 60 Cu (a) 100


Initial rate (dX/dt)
Efficiency after 400 h

Rate (h ) and effciency (%)


Leaching efficiency (%) Ni
10
40
1
Co
0.1
20

-1
Fe 0.01

0
Mn
0.001
0 200 400 600 800 1000 Mn Fe Ni Co Cu
Time / h
initial rate (dX/dt)
(b) 100 (b) 100 Efficiency (%) after 2 h

Rate (h-1) and efficiency (%)


Leaching efficiency (%)

Cu
80 Ni
after 168 h

60 Fe 10

40
Co
1
20
Mn
0
0.001 0.01 0.1 1 0.1
-3
[H2 SO4] / mol dm Mn Co Ni Fe Cu Zn

Fig. 8. Initial rate and final efficiency of metal ion leaching from manganese
(c) 100 nodules at (a) H2SO4 at pH 2, 25 °C, 100  200 mesh, S/L = 1/50, 400 h (Han and
Fuerstenau, 1975); (b) HCl at 1.5 mol dm3, 80 °C, 100–152 lm, S/L = 1/10, 2 h
Ni
Leaching efficiency

80 (Kanungo and Jena (1988a,b)).


after 3 h (%)

60 Fe

40 Mn 60 Cu
Cu Co
Ni
of Ni, Co, Cu, Mn (%)
Leaching efficiency

20
40
0
0 1 2 3 Co
[HCl] / mol dm-3
20
Fig. 7. Effect of acid concentration on leaching efficiency of metals from manganese y = 3.8x + 0.8 y = 0.1x + 0.4
2 2
nodules (a) H2SO4, pH 2, 25 °C, 100  200 mesh, S/L = 1/50; (b) same conditions as R = 1.0 R = 0.8
(a) with variable H2SO4 concentration representing pH 3, 2, 1, 0.5 and 168 h (Han Mn
and Fuerstenau, 1975); (c) HCl, 80 °C, S/L = 1/5, 110–152 lm, 3 h (Kanungo and Jena 0
0 5 10 15
(1988a,b)).
Leaching efficiency of Fe (%)

Fig. 9. Leaching efficiency of Ni, Co, Cu and Mn at different time intervals in H2SO4
(0.22) > Ni (0.025) > Fe (0.008) > Mn (0.006) in sulphate media. The (pH 2) as a function of leaching efficiency Fe (data from Fig. 7a).
extent of Fe dissolution is lower compared to that of Ni in Figs. 7a–
b and 8a. The low rates and low extent of dissolution of Fe and Mn NiOFe2O3 or NiFe2O4 in solid state as revealed by the high K values
in Figs. 7a and 8a are consistent with the potential–pH diagrams in of the reverse reactions of A12 and A13.
Figs. 4–6 and the values of K in Table 5, which show that acidic pH Favourable values of K for acid dissolution of Fe and Ni are
and reducing agents such as SO2 or Fe(II) are essential for the dis- evident from reactions D3–D4. However, the relatively high rates
solution of Fe(III)/Mn(IV) oxides of high valency. Despite the lower and efficiencies of Ni dissolution compared to those of Fe sup-
dissolution rates of Ni compared to Co, the metal dissolution at pH port different reactions for the dissolution of these two metal
2 after 400 h is higher in Fig. 8a: Cu (59%) > Ni (52%) > Co (32%) > Fe ions shown by reactions B6 and C6 for Ni, and B1 for Fe. More-
(14%) > Mn (1.3%). This indicates the possibility of metatheses type over, the potential–pH diagrams in Figs. 4–6 also show the high
dissolution by metal ion displacement described by Han and Fuer- stability of FeOOH, and the need for strong acid to dissolve FeO-
stenau (1975), as discussed in detail later. OH at very low pH, compared to NiO. The latter requires only
Fig. 9 plots the leach efficiency of Cu, Ni and Mn as a function of mildly acidic conditions because unlike NiO which dissolves
leach efficiency of Fe, in order to highlight the linear correlation of according to the simple reaction B6, the dissolution of FeOOH
Ni–Fe, which is consistent with the leaching results from limonitic in H2SO4 involves Fe(OH)SO4 which reacts further with H2SO4
laterite ores during atmospheric acid leaching (Senanayake and to produce FeSOþ 
4 —FeðSO4 Þ2 (Georgiou and Papangelakis, 1998;
Das, 2004) and reductive bioprocessing (Hallberg et al., 2011). Senanayake and Das, 2004):
The initial leach rate ratio (dXNi/dt)/(dXFe/dt) in Fig. 8a is 0.025/
0.008  3/1, and close to the leach efficiency ratio of (%Ni)/(%Fe) FeOOHðsÞ þ Hþ þ HSO4 ¼ FeðOHÞSO4ðads=aqÞ þ H2 O
after 400 h = 52/14 = 3/1. This suggests the co-existence of Ni in FeðOHÞSO4ðads=aqÞ þ Hþ ¼ FeSOþ4 þ H2 O
the goethite lattice (Fe,NiOOH) in manganese nodules, as in the
FeðOHÞSO4ðads=aqÞ þ HSO4 ¼ FeðSO4 Þ2 þ H2 O
case of laterite ores (Georgiou and Papangelakis, 1998; Senanayake
and Das, 2004), or the existence of the stable mixed oxide FeOOH þ 3H2 SO4 ¼ FeSOþ4 þ FeðSO4 Þ2 þ 2H2 OðoverallÞ
1390 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

In contrast, the Mn dissolution in Fig. 7 remains very low even shows a zero slope at higher acid concentrations. A similar trend
at high acidities, in the absence of a reducing agent, indicating the has been reported for the dissolution of synthetic CuO in HCl based
existence of Mn3O4, MnO2 or MnOOH. on data reported by Majima et al. (1980) according to the reaction
Fig. 8b shows initial leach rates and efficiencies for nodules of mechanism (Majima et al., 1980; Senanayake, 2007):
particle size range 100–152 lm in 1.5 mol dm3 HCl. The leach
efficiency in chloride media after 2 h (Fig. 8b) or 3 h (Fig. 7c) is CuOðsÞ þ HClðaqÞ ¼ CuðOHÞClads=aq þ DhH2 O ðslow equilibrationÞ
much larger than that in sulphate media after 168 h (Fig. 7b) or CuðOHÞClads=aq þ HClðaqÞ ¼ CuCl2 ðaqÞ þ H2 O ðfast reactionÞ
800 h (Fig. 7a). Faster initial rates in chloride media highlights
the role of chloride ions and warrants further discussion presented Relatively higher reaction orders of 1.5–2 for Fe and Co are a re-
in a later section. sult of the involvement of H+, Cl and H2O in the dissolution of
these multi/high valent metal oxides. For example, the rate of dis-
solution of FeOOH follows 2nd order kinetics with respect to HCl
4.2. Reaction order and mechanism based on initial rates concentration, but 1st order kinetics with respect to H+ activity,
suggesting a reaction mechanism via Fe(OH)2Cl (Senanayake,
Fig. 10 shows a log–log plot of initial rate of Cu dissolution 2007):
(dXCu/dt) as a function of acid concentration in order to compare
and contrast the dissolution of Cu(II) from: (i) manganese nodules FeOOHðsÞ þ HClðaqÞ ¼ FeðOHÞ2 Clads=aq þ DhH2 O ðslow equilibrationÞ
in H2SO4 solutions of pH 1, 2 and 3 at 25 °C (Han and Fuerstenau, FeðOHÞ2 Clads=aq þ 2HClðaqÞ ¼ FeCl3 ðaqÞ þ 2H2 O ðfast reactionÞ
1975), (ii) manganese nodules in 1.5 mol dm3 HCl at 40 °C
(Kanungo and Jena, 1988a), and (iii) CuO(10%)/a-Al2O3 in H2SO4 This mechanism and the interaction between Fe3+ and Cl as re-
solutions of concentrations 0.1, 0.5, 1 and 2 mol dm3 at 25 °C vealed by the large value of K in reaction L10 in Table 5 rationalises
(Habbache et al., 2009). The linear relationship of slope close to the increase in Fe dissolution in 0.5–3.0 mol dm3 HCl depicted in
unity suggests a first order dependence of initial dissolution of Fig. 7c. The enhanced leaching of Fe(III)-oxides in concentrated HCl
Cu(II) with respect to C Hþ at low acid concentrations. The two data can also facilitate the leaching of Mn/Co incorporated in ferrite
sets from results reported by Han and Fuerstenau (1975) for man- matrices according to reactions C2, C3, C5 and C16. A reductive role
ganese nodules and Habbache et al. (2009) for CuO(10%)/a-Al2O3 of dissolved Mn(II) (reactions H1–H3) in high-valent Co-oxides
follow the same linear relationship, despite the different starting cannot be ruled out. Although CuCl2 and FeCl3 are considered to
materials. This suggests that the initial dissolution of Cu(II) from be the final products in the reaction mechanisms described above,
the two types of oxides follows the same reaction mechanism, the formation of actual chlorocomplex species of Cu(II) or Fe(III) in
0 þ 0
indicating the possible dissolution of CuO species from manganese chloride solutions (CuCl+, CuCl2 , FeCl2+, FeCl2 , FeCl3 ) is governed by
nodules without much interference from host-oxides (reaction B4). the chloride ion activity of solutions as revealed by the Eh- log aCl
Kanungo and Jena (1988a,b) reported the initial rates (dXM)/dt diagrams (Senanayake and Muir, 2003; Senanayake, 2009).
(h1) of metal dissolution from nodules in 0.5 mol dm3 HCl in
the descending order: Cu (6.31) > Zn (5.37) > Ni (1.70) > Fe 4.3. Kinetic models for Ni(II) dissolution
(0.44) > Mn (0.07) > Co (0.04). Moreover, a log–log plot of initial
rates of metal dissolution as a function of HCl concentration in Fig. 11a shows a good linear relationship of 1  3(1  X)2/3 +
the range 0.5–1.5 mol dm3 at 60 °C revealed the reaction orders 2(1  X) as a function of t, based on Eq. (2a) in Table 4 and
(n) for metal dissolutions with respect to HCl concentration shown the results reported by Han and Fuerstenau (1975) for the dissolu-
in brackets: Co (n = 2) > Fe (n = 1.5) > Ni (n = 0.85) > Mn tion of Ni(II) from nodules in sulphuric acid solutions at pH 2.
(n = 0.8) > Zn (n = 0.22) > Cu (n = 0.16). The value of n  1 indicates Other functions of X described by Eqs. (1), (3) or Eq. (4) do not
first order dependence of Ni and Mn dissolution with respect to show linear relationships. Based on the reported data and Eq.
HCl and the involvement of H+ in the surface reaction. The lower (2a), the slopes of linear relationships, generally described by the
reaction order of n = 0.16 for Cu is consistent with Fig. 10, which apparent rate constants (kapparent in Table 4), decrease with the in-
crease in pH: 2  105 (pH 3), 4  104 (pH 2), 2.4  103 (pH 1).
Thus, a log–log plot of kapparent as a function of C Hþ in Fig. 11b
2
shows a linear relationship which confirms that the diffusion of
H+ through a thickening porous product layer is rate controlling.
The results reported by Kanungo and Jena (1988a,b) for the two
acid concentrations 0.5 mol dm3 HCl and 1.5 mol dm3 HCl plot-
1
ted in Fig 11b also lie on the same line of slope close to unity. In
log{[dXCu /dt]/h-1}

addition, the apparent rate constant for Ni or Fe dissolution from


Slope 1 a limonitic laterite ore in acid solutions (Senanayake and Das,
0 2004) also fit into the same linear relationship, showing that the
diffusion of H+ through a thickening porous product layer is rate
controlling in all cases.
According to the particle size range of 100–200 mesh (150–
-1 a -CuO.Al2O 3 / H2 SO4 (0.1-2 M) 75 lm) for nodules, an average value of ro = 104 m along with
b -Nodules / H 2 SO4 (pH 1-3) b = 0.5 (reaction B6 in Table 5) can be used for the calculation of
c -CuO-Al2 O 3 / HCl (0.5 M) the diffusion coefficient DHþ in Eq. (2b) of Table 4. Based on the fact
d -Nodules / HCl (1.5 M) that manganese nodules consist of predominantly Mn–Fe–oxides
-2 with 17% moisture, the density of nodules is close to 4.9 g cm3.
-4 -3 -2 -1 0 1
This value, along with the nickel grade of 0.8%, leads to a value
log {[H + ] / mol dm-3} of qNi = 4.2  102 mol m3. The substitution of these values in the
Fig. 10. Log–log plot of initial rate of copper dissolution as a function of acid
relationship in Eq. (2b): 6bDHþ =qNi r 2o = 108.17, based on the y-
concentration: (a and c) 10%CuO/a-Al2O3 from Habbache et al., 2009; (b) nodules intercept of Fig. 11b, gives a value of DHþ = 2.4  1011 cm2 s1.
from Han and Fuerstenau, 1975; (d) nodules from Kanungo and Jena (1988a,b)). Table 6 shows that the value of DHþ based on Fig. 11b is in good
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1391

(a) 0.02 f(X) = 1-3(1-X)2/3+2(1-X) (a) 0.6


Cu
Ni

-1
0.4

(dX/dt)/h
Co
f (X)

0.01
0.2
y = 0.0004x - 0.0002
R2 = 0.9997
0.00 0
0 20 40 60 0 1 2 3 4
Time (t) / h pH

(b) -4
1 2
(b) 0.1
Legend

-1
log (k apparent / s ) for Ni or Fe

(dX/dt) / h
3 4 1-2 : Manganese nodules Cu
-5 o
3-8 : Nickel laterite ores at 90 C 0.05
5 6 o
Ni
1. Ni / H2 SO4 (pH 1-3) at 25 C
-6 7 8 o
Co
2. Ni / HCl (0.5, 1.5 M) at 40 C
-1

3. Fe / H 2 SO4 (0.18-0.36 M)-SO2 0


-7 4. Fe / HClO4 (1 M) 0 1 2 3
- -3
y = 1.04x - 8.17 5. Fe / SO2 (pH 2) [Cl ] / mol dm
2
R = 0.98 6. Ni / H 2 SO4 (0.72 M)-SO2
-8
Fig. 12. Effect of pH, NaCl and CaCl2 on initial leaching rates of metals from
7. Fe / H 2 SO4 (0.72 M)-SO2
manganese nodules, 100  200 mesh, 25 °C; (a) in H2SO4, (b) in 1 M NaCl or 1 M
8. Ni / H 2 SO4 (1 M ) CaCl2 at pH 3 (data from Han and Fuerstenau, 1975).
-9
0 1 2 3
+ -3
log [H ] / {mol m }

Fig. 11. (a) Applicability of a shrinking core kinetic model for Ni dissolution from 100 Mn Co Ni Fe Cu Zn
manganese nodules; (b) log–log correlation of apparent rate constant for Ni and Fe
dissolution and H+ concentration. Data from Han and Fuerstenau (1975) at 25 °C
Leaching efficiency (%)

and 75–150 lm, Kanungo and Jena (1988a) at 40 °C and 100–152 lm, Senanayake 80
and Das (2004) at 90 °C and 90–125 lm laterite.
60

Table 6 40
Effect of medium and material on proton diffusion coefficient DHþ .

Medium/material DHþ ( cm2 s1) References 20


Aqueous solutiona 9  105 Bockris and Reddy, 1977
Single-crystal ironb 8  105 Bockris and Reddy, 1977 0
c-MnO2 8  108 Allen et al., 1979 1 2 3 4 5 6 7 8
Product layer on LOc 109 Senanayake and Das, 2004 Test No.
Fe3O4 8  1010 Allen et al., 1979
NiO 1010–1012 Allen et al., 1979 Fig. 13. Leaching efficiency of metals from manganese nodules in HCl under
Product layer on MNd 2  1011 This work different conditions listed in Table 7 (Kanungo and Jena (1988a,b)).
Product layer on LOe 1011 Georgiou and Papangelakis, 1998
a
At infinite dilution.
b
Diffusion of atomic hydrogen into the metal. beneficial effect of increasing chloride follows the order Ni(II) > -
c
Predominantly Fe3O4 produced during reductive leaching of laterite ore with Cu(II) > Co(II) (Fig. 12b). As shown in Fig. 10, the change in acid
SO2/H2SO4 at 80 °C. from 0.5 mol dm3 H2SO4 to 0.5 mol dm3 HCl changes the initial
d
Produced on manganese nodules during acid leaching at 25 °C. rate of Cu(II) dissolution dXCu/dt (10%CuO/a-Al2O3) from 6.76 to
e
Produced on limonitic laterite ore during acid pressure leaching with H2SO4 at
230–270 °C.
50.1. This 7.4 fold increase also confirms the involvement of Cl
ions in the surface reaction. The initial rates dXM/dt in HCl
(Fig. 8b) are generally larger than those in H2SO4 (Fig. 8a) due to
agreement with the other relevant values reported in the litera- high acid and chloride in HCl.
ture. It is of particular interest to note that the value of Fig. 13 compares and contrasts the leaching efficiency of metal
DHþ  1011 cm2 s1 for the leaching of Ni from manganese nod- ions in HCl under different conditions listed in Table 7. Doubling
ules in H2SO4 is of the same order as that for the acid-pressure
leaching of nickel from limonitic laterite reported by Georgiou
Table 7
and Papangelakis (1998).
Test conditions for results in Fig. 13.

Test HCl (M) T (°C) Size (lm) Na2SO3 (%) S/L (%)
4.4. Factors affecting leach efficiency
1 3 80 100–152 0 20
2 1.5 80 100–152 0 20
4.4.1. Chloride ions
3 1.5 40 100–152 0 10
Han and Fuerstenau (1976a) compared and contrasted the ef- 4 1.5 60 100–152 0 10
fect of pH in the range 1–3 and in 1 mol dm3 NaCl or CaCl2 at a 5 1.5 80 100–152 0 10
fixed pH of 3, to examine the beneficial effect of the added cations 6 1.5 80 152–251 0 10
on the dissolution of metals from manganese nodules. Their results 7 1.5 80 152–251 10 10
8 1.5 80 152–251 20 10
plotted in Fig. 12 show that the detrimental effect of increasing pH
follows the order Cu(II) > Co(II) > Ni(II) (Fig. 12a), whilst the Kanungo and Jena (1988a,b), Tests 1, 2 after 3 h, others after 2 h.
1392 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

slopes Table 7) seem to have a negative effect on leach efficiency of Ni,


(a) 100
(a) 1 h Fe, Co and Mn, but a small effect on Cu and Zn. The increase in
Leaching efficiency of Ni (%)

(e) 0.80 temperature from 40 °C to 80 °C without Na2SO3 (Tests 3–5) also


(b) 6 h
80 has a small effect on Cu and Zn compared to the larger beneficial
(c) 48 h
(d) 0.81
effect on other metals. These results from Kanungo and Jena
60 (d) 168 h (1988a,b) are consistent with the findings of Han and Fuerstenau
(e) 800 h (1976a,b) who reported that:
(c) 0.78
40
(i) rates were independent of particle size finer than 200 lm,
(b) 0.27
20 (ii) diffusion of dissolved Cu(II) through pores to bulk solution
was rate controlling for coarser particles and thus Cu disso-
(a) 0.08
0 lution was less sensitive to temperature due to low activa-
0 20 40 60 80 100 tion energy,
Leaching efficiency of Cu (%) (iii) Ni dissolution was controlled by surface reaction and more
sensitive to temperature due to high activation energy.
(b) 60

Ni The reported values of kapparent based on a shrinking sphere ki-


netic model (Eq (1) in Table 4) for the dissolution of Cu(II) and
of Ni, Mn, Co, Fe (%)
Leaching efficiency

Co(II) from mixed oxide MO(10%)/a-Al2O3 show the beneficial ef-


40
fect of Cl; the values of k for both metal ions are higher in HCl
Co than in H2SO4 (Habbache et al., 2009; Boukerche et al., 2010). How-
ever, the magnitude of k for Cu(II) dissolution at 30–50 °C vary
20 from 0.05 to 0.25 min1 compared to the lower values of
Fe k < 0.006 min1 for Co(II) dissolution at 50–85 °C. Despite the ben-
eficial effect of high temperature, the activation energy (Ea) for the
Mn dissolution of Cu from CuO(10%)/a-Al2O3 is 24 kJ mol1 and
0 37 kJ mol1 in HCl and H2SO4, respectively (Habbache et al.,
0 20 40 60 80
2009). These low values of Ea support the observations by Han
Leaching efficeincy of Cu (%)
and Fuerstenau (1976a,b) that the dissolution of Cu is mixed con-
Fig. 14. Leaching efficiency of Ni, Co, Fe, Mn from manganese nodules in H2SO4 at trolled, as noted in Section 2.5.
pH 0.5, 1, 2 and 3 as a function of Cu at different time intervals up to 800 h
(conditions same as in Fig. 7).

the HCl concentration from 1.5 mol dm3 (Test 2) to 3 mol dm3
(a) 100
Mn-SO 2
(Test 1) causes an increase in metal dissolution as shown in
Leaching efficiency (%)

80
Fig. 13: Mn (7–51%) < Co (12–74%) < Fe (34–77%) < Cu (90– Fe-HCl
Co-SO 2
100%)  Ni (92–100%). These changes support a HCl assisted disso-
lution mechanism of Fe(III)-oxides as noted in Section 4.2. 60

4.4.2. Copper(II) ions 40 Co-HCl


The involvement of Cu(II) as a mediator in cation-exchange dis-
solution is possible according to the large values of K in reactions 20
E7–E13, if the precipitated CuO or Cu(OH)2 is rapidly dissolved Mn-HCl
Fe-SO 2
by acid. A reasonably good correlation between Ni and Cu dissolu- 0
tion is shown in Fig. 14a based on the results at pH 0.5–3 after 1 h, 30 40 50 60 70 80 90
6 h, 48 h, 168 h, 800 h. This observation and the comparable Cu and Temperature oC
Ni extractions, despite the low initial rates of Ni compared to Cu in
Fig. 8, shows the involvement of Cu(II) in the dissolution of Ni(II) (b) 100 Ni-HCl
(reactions E7, E10). Again, the dissolution of Fe(II) and Mn(II) re-
mains low in Fig. 14b, without an added reducing agent, even at Cu-HCl
Leaching efficiency (%)

80
high Cu(II) and Ni(II) dissolution. Moreover, the dissolution of Ni
seems to be enhanced after the Co leaching efficiency has reached Zn-HCl
60
a constant value of 33% (Fig. 14b). Direct involvement of Cu2+ in the
dissolution of Co2+ by an ion-exchange type mechanism in reac-
Ni-SO 2
tions E2–E5, and acid dissolution of Co-oxides according to reac- 40
tion C1, are unlikely due to the low values of K. Thus, the linear
Cu-SO 2
relationship between Co–Cu dissolution of slope 1 at low extrac- 20
tions (<30%) in Fig. 14b can be related to the faster rates of acid dis- Zn-SO 2
solution of divalent Cu(II) and Co(II) from oxides as shown in
0
Fig. 8a and reactions B4 and B7. The involvement of reactions C3, 30 40 50 60 70 80 90
C5 and C6 cannot be ruled out. o
Temperature C

Fig. 15. Effect of temperature on leaching efficiency of metals from manganese


4.4.3. Particle size and temperature nodules (a) Fe, Mn, Co; (b) Ni, Cu, Zn; Conditions: SO2 leaching: 2.5% SO2, S/L = 1/12,
As shown in Fig. 13 the increase in particle size in tests carried 15 min (Kanungo and Das, 1988); HCl leaching: 1.5 M HCl, S/L = 1/10, 2 h, 100 lm–
out with 1.5 mol dm3 HCl in the absence of Na2SO3 (Tests 5–6, 152 lm (Kanungo and Jena (1988a,b)).
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1393

100

SO2, H + or HSO3- / mol dm-3


0.25 1.8
SO2

Leaching efficeincy (%)


Concentration of
0.20 80
1.6
0.15
60

pH
pH 1.4
0.10
1.2 40
0.05
HSO3- (or H+)
0.00 1 20 Fe Mn Co
0 50 100 Cu Ni Zn
Temperature oC 0
1 2 3 4 5
Fig. 16. Effect of temperature on S(IV) speciation (calculated for 0.25 mol dm3 SO2 Test No.
using K values from HSC 6.1, Roine, 2002).
Fig. 17. Leaching efficiency of metals from manganese nodules in H2SO4 under
different conditions listed in Table 8 (Acharya et al., 1999).
Fig. 15a–b summarises the effect of temperature on metal
leaching efficiency in 2.5% SO2/H2O (after 15 min) and 1.5 M HCl
(after 2 h). The decrease in leach efficiency of metals in 2.5% SO2 Table 8
at temperatures above 50 °C has been related to the loss of dis- Test conditions for results in Fig. 17.
solved SO2 from solution (Kanungo and Das, 1988). The dissolution
Test H2SO4 (v/v%) SO2 (M) (NH4)2SO4 (M)
kinetics of Fe and Mn oxides by SO2 show the possibility of an elec-
1 2 0 0.15
trochemical surface reaction mechanism which involves H+ and
2 0 0.47 0.15
HSO 3 (Senanayake, 2003a,b, 2004). A decrease in K for the dissoci- 3 1 0.47 0.15
ation reaction L2 (Table 5) at high temperature causes a decrease in 4 2 0.47 0.15
concentration of HSO +
3 and H (Fig. 16). This in turn can cause a de-
5 2 0.47 0
crease in leach efficiency as demonstrated in Fig. 15a–b. Acharya et al. (1999), 90 °C, 10% S/L, 1 h.
The leach efficiency of Fe in 1.5 M HCl is much higher (25–75%,
Fig. 15a) than that in SO2 (<1%) over the temperature range 40–
90 °C, due to the involvement of Cl as a complexing agent for 100 1.5
Fe(III), as noted in Section 4.2. It increases with the increase in tem-
Leaching eficiency (%)

perature; Ni leach efficiency in HCl follows a similar trend in 80 Fe


Mn
Fig. 15b. Whilst the high leach efficiencies of Cu and Zn in HCl
after 15 min

Initial pH
60 Cu
(70–80%) remain relatively unaffected by temperature, those of
1.0 Ni
Mn and Co remain less than 30%, even at high temperatures 40 Co
(Fig. 15a). This indicates the need for a more effective reducing
Zn
agent. For example, the redox reaction of MnO2 with SO2 has a lar- 20
pH
ger value of K  1026 (reaction G10) compared to K  104.4 with
0 0.5
Cl (reaction J1). 0 0.05 0.1
-3
[H 2 SO 4] / mol dm
4.4.4. Reducing agents
4.4.4.1. Sulphite ions or SO2 in HCl. Fig. 13 shows the beneficial ef- Fig. 18. Effect of H2SO4 concentration and initial pH on metal leaching efficiency
from manganese nodules: 3% SO2, 31 °C, S/L 10%, 15 min (Kanungo and Das, 1988).
fect of Na2SO3 on the leaching efficiency of Mn and Co in HCl (Tests
6–8). The predominant S(IV) species in a solution of Na2SO3/HCl is
expected to be SO2. This prediction is based on species distribution
diagrams which show the different pH ranges for the formation of content in nodules. This is consistent with the reductive ability of
various sulphur(IV) species in aqueous media: SO2 (pH < 2), HSO 3 Cu(I) on a limonitic laterite ore (Das et al., 1997). Maximum extrac-
(2 < pH < 7) and SO2 3 (pH > 7) and the low measured pH value tion of metal values from nodules were obtained at low tempera-
3
(<0) of 1.5 mol dm HCl (Senanayake, 2003a,b, 2004). Thus, an in- tures (30–40 °C) and particle size range 150 + 76 lm (Kanungo
crease in concentration of Na2SO3 from 0% to 20% in Tests 6, 7 and 8 and Das, 1988). Vu et al. (2005) reported that 90% metal extraction
in Fig. 13 causes a dramatic increase in Mn and Co leach efficiency from nodules can be achieved in stoichiometric quantities of FeSO4
from 5% to 80% even from coarse nodule particles, showing the role irrespective of the H2SO4 quantities in excess over the stoichiome-
of SO2 as a reducing agent. tric amount at S/L ratios in the range 1/15–1/5. An increase in S/L
The decrease in Fe extraction from 30% to 20% (see Fig. 13) with ratio and a decrease in temperature caused a decrease in metal
the increase in Mn and Co extraction in Tests 6–8 is consistent with extraction.
previous studies on reductive acid leaching of limonitic laterite by Acharya et al. (1999) conducted test work to show the effect of
SO2 (Senanayake and Das, 2004). Here, the Fe(II) formed in a reduc- 2% H2SO4, 0.47 mol dm3 SO2, 0.15 mol dm3 (NH4)2SO4 at 90 °C
ing media can enhance the reductive leaching of Mn, but this in and 10% S/L on metal extraction from manganese nodules after a
turn precipitates Fe(III) in the form of FeOOH or Fe2O3 according leaching period of 1 h. The results summarised in Fig. 17 at differ-
to reactions G4, G7, G8 and G9 causing a low leach efficiency of Fe. ent conditions listed in Table 8 show the effect of: (i) H2SO4 in the
absence of SO2 (Test 1), (ii) SO2 in the absence of H2SO4 (Test 2),
4.4.4.2. SO2 and Fe(II) in H2SO4. Kanungo and Das (1988) conducted (iii) both H2SO4 and SO2 (Test 3), and (iv) increasing acid (Tests
testwork to show the need for aqueous SO2 to rapidly reduce MnO2 3–4). Fig. 18 shows the metal extraction after 15 min of reductive
and leach Mn, Co, Ni, Zn, Fe and Cu. The metal extraction depended leaching by SO2 (3%)/H2SO4 at 31 °C using solutions of different
upon the ratio of SO2 (moles) per unit mass of nodule sample. This acid concentrations. The reductive role of SO2 is also evident from
ratio affects the equilibrium Cu(II) + Fe(II) = Cu(I) + Fe(III) and thus the relatively high extraction of Mn and Co in Fig. 18, even at a low
the selective extraction of Mn, Co and Ni depending on the copper concentration of 0.05 M H2SO4 (pH 0.75): Mn (94%) > Co (90%) > Ni
1394 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

(82%) > Cu (75%) > Zn (54%) > Fe (19%). In contrast, the leach effi- (iv) Ni, Cu and Zn appear to be selectively dissolved from ferrites
ciency of high-valent metals without SO2 in Fig. 7b are only 45% by acid, leaving the high valent Fe and Mn residues behind in
Co and 5% Mn. the solid phase according to reactions C6, C8–C10 with high
The important points to note in Fig. 17 are: values of K;
(v) high leach efficiency of 90% Co in Test 3, compared to 1–6%
(i) leach efficiency of 1% Mn in Test 1 in the absence of SO2 Co in Tests 1 and 2, due to reductive action of SO2/H+
increasing to 75% in Test 2 in the presence of SO2 due to expected from reactions H6–H7;
the fact that MnO2 or Mn3O4 can be dissolved by aqueous (vi) leach efficiency of Fe, Co, Ni and Zn decreased from Test 1 to
SO2 alone in the absence of H2SO4 (reactions G11 and G14); Test 2 and increased from Test 2 to Test 3 due to the need for
(ii) a further increase from 75% to 93% Mn occurred in Test 3 in strong acid, as well as a reducing agent;
the presence of both SO2 and H2SO4 due to reactions G12 (vii) low leach efficiency of iron compared to other metals in
and G13; Tests 3,4 and 5, even under the reducing conditions, is due
(iii) Ni, Cu and Zn show high leach efficiency (>60%) compared to to the precipitation of ammonium jarosite (NH4)Fe3(-
Co, Mn and Fe (0–12%) even in 1 mol dm3 H2SO4 at 90 °C SO4)2(OH)6 (Acharya et al., 1999).
(Test 1), as expected from partial or low dissolution by reac-
tions such as B1, and C2–C7; The precipitation of ammonium jarosite is further exemplified
by the decrease in leaching efficiency from 56.2% Fe to 10.3% Fe
with an increase in (NH4)2SO4 from 0 to 50 g dm3
(0.4 mol dm3) during a leach period of 1 h. The change in leach
100 N2 / 0.5 L/min Open air Air / 0.5 L/min
efficiency of other metals followed the descending order: Co (95.9–
Leaching eficiency (%)

88.8%) > Cu (99.6–93.5%) > Mn (99.8–97.8%) > Zn (99.6–99.2%),


80
compared to the slight increase of Ni (99.2–99.8%) (Acharya
after 20 min

60 et al., 1999).

40 4.4.4.3. Gaseous atmosphere. Reductive leaching in the presence of


SO2 is also sensitive to the nature of the gaseous atmosphere as
20
shown in Fig. 19. The leach efficiency of Fe, Zn and Cu by a solution
of 3.55% SO2 at 31 °C and S/L ratio 1/10 is enhanced when the atmo-
0
Fe Cu Zn Ni Co Mn sphere is changed from N2 flow (0.5 dm3 min1), to open air or air
flow (0.5 dm3 min1). The significance of the beneficial effect of
Fig. 19. Effect of gaseous atmosphere on efficiency of metal leaching from changing atmosphere from N2 to air on Cu dissolution is larger
manganese nodules: 3.55% SO2, 31 °C, S/L 10% (Kanungo and Das, 1988).
(12–72%), compared to Fe (1–2%), and Zn (45–59%). In the absence
of SO2, Cu showed the highest leach efficiency. Kanungo and Das
(1988) related this to the formation of basic Cu(I)–Fe(III)–sulphite
salts which are unstable or insoluble in water. The formation of
100
(a) CuFeO2 is also a possibility according to Fig. 4. The oxidation of
SO2 to H2SO4 by O2 (reaction L13) is catalysed by transition metals
Leaching efficiency of Co (%)

80 such as Cu(II) and Mn(II), which enhance the H2SO4 content. This
(b)
leads to improved leach efficiency in the presence of air shown in
Fig. 19. A similar effect of enhanced H2SO4 concentration is evident
60
(c) in Fig. 17, especially in the case of Ni, Zn and Cu in Tests 2, 3 and 4.

40 4.4.4.4. Manganese(II) ions. Fig. 20 plots Co–Mn leach efficiencies in


(d)
tests conducted under different conditions listed in Table 9 to show
20 linear relationships. A slope close to unity in the presence of
(e) Na2SO3 shows that Mn(II) produced due to the reduction of high
valent Mn-oxides by SO2 acts as a reducing agent for Co-oxides.
0 The relevant chemical equations are G10–G14 and H1–H2 in Ta-
0 20 40 60 80 100
ble 5. Rubisov and Papangelakis (2000) considered the involve-
Leaching efficiency of Mn (%)
ment of reductive leaching reactions L1 and G6 during HPAL of
Fig. 20. Leaching efficiency of Co–Mn under different conditions: (a–c) atmospheric limonitic laterite, as demonstrated by favourable values of
leaching of manganese nodules (d–e) pressure acid leaching of limonitic laterite K  102 for L1 at 250 °C compared to K  103 at 90 °C (Table 5).
(conditions and references in Table 9). Thus, the linear relationship of slope 1 for Co–Mn leaching in

Table 9
Test conditions for results in Fig. 20.

Test Acid (M) T (°C) Size (lm) Na2SO3 (%) S/L Leach time (h) Slope*
#
a 1–1.5 M HCl 80 100–152 0 1/5 3 1.3
b 1.5 HCl 40–60# 100–152 0 1/10 2 1.3
c 1.52 M HCl 80 152–251 0–20%# 1/5 2 1.1
d H2SO4 230 1–2 0 1/5 0–1# 1.1
e H2SO4 250 1–2 0 1/5 0–1# 1.1
*
Slope of linear relationships in Fig. 20.
#
Denotes variable in each test relevant to the different data points in Fig. 20 (data for atmospheric leach from Acharya et al., 1999; HPAL from Georgiou and Papangelakis,
2009).
G. Senanayake / Minerals Engineering 24 (2011) 1379–1396 1395

the HPAL process at 230–250 °C in Fig. 20 based on the results from and HCl media. The involvement of Fe(II) in the reductive leaching
Georgiou and Papangelakis (2009) warrants further studies on of high-valent Mn-oxides causes low Fe dissolution due to precip-
such correlations. itation of Fe(III)-oxides. The linear correlations between leaching
efficiencies of Co and Mn suggest the reductive leaching of high-
valent Co oxides by Mn(II) produced under reducing conditions.
5. Summary and conclusions
Acknowledgement
The concentrations of divalent cations Pb2+, Cd2+, Co2+, Mn2+,
Cu , Fe2+, Zn2+, Ni2+, UO2þ
2+
2 and Ba
2+
in seawater (pH  8) of the or-
Finacial support from the Parker CRC for Integrated Hydromet-
der 103–102 nmol/kg show a linear correlation with solubility
allurgy Solutions and Murdoch University is gratefully
product of metal carbonates. This observation is consistent with
acknowledged.
the previous finding that metal ion solubility in seawater is af-
fected by carbonate sediments (Morse and Arvidson, 2002). The
values of pKSP tend to increase with increasing softness of cations, References
leading to lower concentrations of dissolved cations.
The leaching behaviour of Mn, Fe, Ni, Co, Cu and Zn from man- Acharya, R., Ghosh, M.K., Anand, S., Das, R.P., 1999. Leaching of metals from Indian
ocean nodules in SO2–H2O–H2SO4–(NH4)2SO4 medium. Hydrometallurgy 5,
ganese nodules in acid media (pH 6 4) depends on the nature of 169–175.
the porous oxide matrix. For example, the dissolution of major con- Agarwal, H.P., Goodrich, J.D., 2003. Extraction of copper, nickel, and cobalt from
stituents (Mn, Fe) affects the dissolution of some minor constitu- Indian Ocean Polymetallic Nodules. The Canadian Journal of Chemical
Engineering 81, 303–306.
ents (Ni, Co) due to their coexistence as mixed-oxide matrices Allen, P.R., Hampson, N.A., Bignold, G.J., 1979. The electrodissolution of magnetite.
such as ferrites and magnetite, as revealed by the Eh–pH diagrams. Journal of Electroanalytical Chemistry 99, 299–309.
The estimated equilibrium constants based on the HSC6.1 database Avramov, A., 2005. The alternative-deep-water polymetal nodules in the pacific
(Review). Journal of the University of Chemical Technology and Metallurgy 40,
show a higher thermodynamic stability of mixed valent/metal oxi- 275–287.
des. However, copper appears to be stable in the form of Barnabe, A., Mugnier, E., Presmanes, L., Tailhades, P., 2006. Preparation of delafossite
CuOFe2O3. Thus, Cu(II) dissolution from manganese nodules and CuFeO2 thin films by rf-sputtering on conventional glass substrate. Materials
Letters 60, 3468–3470.
synthetic CuOAl2O3 demonstrates a first order dependence with Bayramoglu, M., Tekin, T., 1993. An electrochemical model for reduction of MnO2
respect to H+ at low concentrations and appears to be generally with Fe2+ ions. Journal of Applied Electrochemistry 23, 1273–1279.
unaffected by host metal oxides. Dissolved Cu(II) ions which slowly Biswas, A., Chakraborti, N., Sen, P.K., 2009. A genetic algorithms based multi-
objective optimization approach applied to a hydrometallurgical circuit for
diffuse out (DCu2þ  1010 cm2 s1) to the solution through the
ocean nodules. Mineral Processing and Extractive Metallurgy Review 30, 163–
pores of nodules can also facilitate Ni(II) dissolution by cation ex- 189.
change type mechanism Cu2+ + NiO = CuO + Ni2+ (K = 105), as re- Bockris, J.O’M., Reddy, A.K.N., 1977. Modern Electrochemistry. Plenum, New York,
vealed by the linear correlation of Cu–Ni leach efficiencies in pp. 1328–1330.
Boukerche, I., Habbache, N., Alane, N., Djerad, S., Tifouti, L., 2010. Dissolution of
H2SO4. Cobalt from CoO/Al2O3 catalyst with mineral acids. Industrial and Engineering
The appearance of NiOFe2O3 in the Eh–pH diagram and the Chemistry Research 49, 6514–6520.
linear relationship of leach efficiencies of Ni(II) and Fe(III) from Byerley, J.J., Rampel, G.I., Garrido, G.F., 1979. Copper catalysed leaching of magnetite
in aqueous sulfur dioxide. Hydrometallurgy 4, 317–336.
nodules in H2SO4 support the view that these two metals Chandra, I., Jeffrey, M.I., 2005. A fundamental study of ferric oxalate for dissolving
dissolve from a mixed oxide matrix such as NiOFe2O3 (or gold in thiosulfate solutions. Hydrometallurgy 77, 191–201.
Fe,NiOOH), as in the case of limonitic laterite ores reported in Charewicz, W.A., Chaoyin, Z., Chmielewski, T., 2001. The leaching behavior of ocean
pollymetallic nodules in chloride solutions. Physicochemical Problems of
the literature (Senanayake and Das, 2004; Hallberg et al., Mineral Proceessing 35, 55–66.
2011). However, the diffusion of H+ through an insoluble product Chiarizia, R., Horwitz, E.P., 1991. New formulations for iron oxides dissolution.
layer appears to be the rate controlling step for Ni(II) and Fe(III) Hydrometallurgy 27, 339–360.
Das, G.K., Anand, S., Das, R.P., Muir, D.M., Senanayake, G., Singh, P., Hefter, G., 1997.
leaching according to a shrinking core kinetic model. This allows Acid leaching of nickel laterites in the presence of sulphur dioxide at
the calculation of a proton diffusivity of DHþ  1011 cm2 s1, atmospheric pressure. In: Cooper, W.C., Mihaylov, I. (Eds.), Hydrometallurgy
which is of the same order as the value of DHþ based on acid and refining of nickel and cobalt, vol. I. Canadian Inst. Min. Metall. & Petroleum,
Montreal, pp. 471–488.
pressure leaching data of Ni from limonitic laterite reported by
Georgiou, D., Papangelakis, V.G., 1998. Sulphuric acid pressure leaching of a
Georgiou and Papangelakis (1998). limonitic laterite: chemistry and kinetics. Hydrometallurgy 49, 23–46.
Metal leaching efficiency increases with increasing acid concen- Georgiou, D., Papangelakis, V.G., 2009. Behavior of cobalt during sulphuric acid
tration and Zn(II) dissolution follows the same trends as Cu(II) in pressure leaching of a limonitic laterite. Hydrometallurgy 100, 35–40.
Habbache, N., Alane, N., Djerad, S., Tifouti, L., 2009. Leaching of copper oxide with
HCl. The presence of Cl ions in H2SO4 accelerates Cu(II) dissolu- different acid solutions. Chemical Engineering Journal 152, 503–508.
tion due to the surface reaction mechanism via Cu(OH)Cl, which Hallberg, K.B., Grail, B.M., du Plessis, C.A., Johnson, D.B., 2011. Reductive dissolution
in turn accelerates Ni(II) dissolution. The chloride assisted rapid of ferric iron minerals: a new approach for bio-processing nickel laterites.
Minerals Engineering 24, 620–624.
leaching of metals from nodules by HCl is also a result of surface Han, K.N., 1971. Geochemistry and extraction of metals from ocean floor manganese
complexation with Cl. Although the increase in temperature en- nodules. PhD thesis, Univerrsity of California, Berkeley, CA, USA.
hances leaching efficiency of Fe(III) and Ni(II), Mn and Co leaching Han, K.N., Fuerstenau, D.W., 1975. Preferential acid leaching of nickel, copper, and
cobalt from ocean-floor manganese nodules. Transactions of the Institution of
efficiencies remain low without SO2 as a reducing agent. Mining & Metallurgy 84 (C), 105–109.
The leaching of all metals with SO2(2.5%)–H2O remain high at Han, K.N., Fuerstenau, D.W., 1976a. Kinetics of the extraction of metals from deep-
low temperatures, but decreases at high temperatures due to low sea manganese nodules: Part I. The pore diffusion controlling case.
Metallurgical Transactions 7B, 679–685.
dissociation according to the reaction: SO2 þ H2 O ¼ Hþ þ HSO 3 at Han, K.N., Fuerstenau, D.W., 1976b. Kinetics of the extraction of metals from deep-
high temperatures. Aeration increases the metal dissolution with sea manganese nodules: Part II. Pore diffusion with chemical reaction.
SO2(2.5%)–H2O due to the formation of H2SO4 caused by the cata- Metallurgical Transactions 7B, 679–692.
Heimendahl, M.V., Hubred, G.L., Fuerstenau, D.W., Thomas, G., 1976. A transmission
lytic effect of transition metal ions on the reaction
electron microscope study of deep-sea manganese nodules. Deep Sea Research
SO2 + 0.5O2 + H2O = H2SO4. 23, 69–79.
The existence of Mn and Co in mixed-valent oxide leads to only Hogfeldt, E., 1964. Stability Constants of Metal–Ion Complexes. Part A: Inorganic
partial dissolution of these two metals from nodules in acids in the Ligands. Pergamon Press, Sydney – Australia.
Hsiaohong, C., Chongyue, F., Di-Ji, Z., 1992. Reduction leaching of manganese
absence of reducing agents. Near complete dissolution of Mn and nodules by nickel matte in hydrochloric acid solution. Hydrometallurgy 28,
Co is possible in the presence of SO2 or Na2SO3 in both H2SO4 269–275.
1396 G. Senanayake / Minerals Engineering 24 (2011) 1379–1396

Hsiaohong, C., 1996. The oxidative leaching of nickel matte with manganese- Pearson, R.G. (Ed.), 1973. Hard and Soft Acids and Bases. Dowden, Hutchinson &
containing oxide ores. Hydrometallurgy 42, 425–434. Ross, Stroudsburg, PA.
Jana, R.K., 1993. Leaching of sea nodules in acidic chloride–sulphide media. Prieto, F.J.M., Millero, F.J., 2002. The values of pK1 + pK2 for the dissolication of
Transactions of the Institution of Mining & Metallurgy 102C, 191–194. carbonic acid in seawater. Geochimica et Cosmochimica Acta 66, 2529–2540.
Kanungo, S.B., Das, R.K., 1988. Extraction of metals from manganese nodules of the Roine, A., 2002. Outokumpu HSC Chemistry Thermochemical Database, ver 6.1.
Indian Ocean by leaching in aqueous solution of sulphur dioxide. Finland: Outokumpu Research Oy.
Hydrometallurgy 20, 135–146. Rubisov, D.H., Papangelakis, V.G., 2000. Sulphuric acid pressure leaching of
Kanungo, S.B., Jena, P.K., 1988a. Studies on the dissolution of metal values in Indian laterites–a comprehensive model of a continuous autoclave. Hydrometallurgy
Ocean origin in dilute hydrochloric acid. Hydrometallurgy 21, 23–39. 58, 89–101.
Kanungo, S.B., Jena, P.K., 1988b. Reduction leaching of manganese nodules of Indian Sen, P.K., 2010. Metals and materials from deep sea nodules: an outlook for the
Ocean origin in dilute hydrochloric acid. Hydrometallurgy 21, 41–58. future. International Materials Reviews 55, 364–390.
Kanungo, S.B., 1999a. Rate process of the reduction leaching of manganese nodules Senanayake, G., 2003a. A surface reaction kinetic model to compare the reductive
in dilute HCl in presence of pyrite: Part I. Dissolution behavior of iron and leaching of iron from goethite, magnetite and limonitic laterite by acidic sulfur
sulphur species during leaching. Hydrometallurgy 52, 313–330. dioxide. Metall. Mater. Trans. 34B, 735–738.
Kanungo, S.B., 1999b. Rate process of the reduction leaching of manganese nodules Senanayake, G., 2003b. The reductive leaching of manganese dioxide: reaction
in dilute HCl in presence of pyrite: Part II. Leaching behavior of manganese. kinetic models and mechanisms. Hydrometallurgy’2003. TMS, Warrendale, pp.
Hydrometallurgy 52, 331–347. 485–498.
Kumar, R., Das, S., Ray, R.K., Biswas, A.K., 1993. Leaching of pure cobalt bearing Senanayake, G., 2004. A mixed surface reaction kinetic model for the reductive
goethites in sulphurous acid: kinetics and mechanisms. Hydrometallurgy 32, leaching of manganese dioxide with acidic sulfur dioxide. Hydrometallurgy 73,
39–59. 215–224.
Kumari, A., Natarajan, K.A., 2001a. Electroleaching of polymetallic ocean nodules to Senanayake, G., 2007. Review of theory and practice of measuring proton activity
recover copper, nickle and cobalt. Minerals Engineering 14, 877–886. and pH in concentrated chloride solutions and application to oxide leaching.
Kumari, A., Natarajan, K.A., 2001b. Electrobioleaching of polymetallic ocean Minerals Engineering 20, 634–645.
nodules. Hydrometallurgy 62, 125–134. Senanayake, G., 2009. A review of chloride assisted copper sulfide leaching by
Kumari, A., Natarajan, K.A., 2002a. Development of a clean bioelectrochemical oxygenated sulfuric acid and mechanistic considerations. Hydrometllurgy 98,
process for leaching of ocean manganese nodules. Minerals Engineering 15, 21–32.
103–106. Senanayake, G., Das, G.K., 2004. A comparative study of leaching kinetics of
Kumari, A., Natarajan, K.A., 2002b. Cathodic reductive dissolution and surface limonitic laterite and synthetic iron oxides in sulfuric acid containing sulphur
adsorption behavior of ocean manganese nodules. Hydrometallurgy 64, 247– dioxide. Hydrometallurgy 72, 59–72.
255. Senanayake, G., Muir, D.M., 2003. Chloride processing of metal sulphides: review of
Langova, S., Lesko, J., Matysek, D., 2009. Selective leaching of zinc from zinc ferrite fundamentals and applications. Hydrometallurgy’2003. TMS, Warrendale, pp.
with hydrochloric acid. Hydrometallurgy 95, 179–182. 517–531.
Levenspiel, O., 1972. Chemical Reaction Engineering. Willey, New York. Shen, Y., Xue, W., Li, W., Li, S., Liu, X., 2007. Recovery of Mn2+, Co2+, and Ni2+ from
Lenoble, J.P., 2000. Polymetallic Nodules. International Seabed Authority. pp. 8. manganese nodules by redox leaching and solvent extraction. Transactions of
Majima, H., Awakura, Y., Yazaki, T., Chiakamori, Y., 1980. Acid disolution of cupric the Nonferrous Metals Society of China 17, 1105–1111.
oxide. Metall. Trans. 11B, 209–214. Smith, R.M., Martell, A.E., 1976. Critical Stability Constants,, .. Inorganic Complexes,
Marcus, Y., 1997. Ion Properties. New York, Marcel Dekker. first ed., vol. 4. Plenum Press, New York – USA.
McDonald, R.G., Whittington, B.I., 2008a. Atmospheric acid leaching of nickel Vu, H., Jandova, J., Lisa, K., Vranka, F., 2005. Leaching of manganese deep ocean
laterites review Part I. Sulphuric acid technologies. Hydrometallurgy 91, 35–55. nodules in FeSO4–H2SO4–H2O solutions. Hydrometallurgy 77, 147–153.
McDonald, R.G., Whittington, B.I., 2008b. Atmospheric acid leaching of nickel Warren, I.H., Hay, M.G., 1975. Leaching of iron oxide with aqueous solution of
laterites review Part II. Chloride and bio-technologies. Hydrometallurgy 91, 56– sulphur dioxide. Transactions of the Institution of Mining & Metallurgy, London
69. 84, C49–C53.
Mehta, K.D., Das, C., Pandey, B.D., 2010. Leaching of copper, nickel and cobalt from Zhang, Y., Liu, Q., Sun, C., 2001. Sulfuric acid leaching of ocean manganese nodules
Indian Ocean manganese nodules by Aspergillus niger. Hydrometallurgy 105, using phenols as reducing agents. Minerals Engineering 14, 525–537.
89–95. Zhang, W., Singh, P., Muir, D.M., 2002. Oxidative precipitation of manganese with
Miller, J.D., Wan, R., 1983. Reaction kinetics for the leaching of MnO2 by sulfur SO2/O2 and separation from cobalt and nickel. Hydrometallurgy 63, 127–135.
dioxide. Hydrometallurgy 10, 219–242. Zhang, W., Cheng, C.Y., 2007. Manganese metallurgy review: Part I: Leaching of
Monhemius, A.J., 1980. In: Burkin, A.R. (Ed.), The Extractive Metallurgy of Deep Sea ores/secondary materials and recovery of electrolytic/chemical manganese
Nodules: Topics in Non-ferrous Extractive Metallurgy. Blackwell SCI, London, dioxide. Hydrometallurgy 89, 137–159.
pp. 42–69. Zeng, L., Cheng, C.Y., 2009a. A literature review of the recovery of molybdenum and
Morse, J.W., Arvidson, R.S., 2002. The dissolution kinetics of major sedimentary vanadium from spent hydrodesulphurisation catalysts; Part I: metallurgical
carbonate minerals. Earth-Science Reviews 58, 51–84. processes. Hydrometallurgy 98, 1–9.
Niinae, M., Komatsu, N., Nakahiro, Y., Wakamatsu, Y., Shibata, J., 1996. Preferential Zeng, L., Cheng, C.Y., 2009b. A literature review of the recovery of molybdenum and
leaching of cobalt, nickel, and copper from cobalt-rich ferromanganese crusts vanadium from spent hydrodesulphurisation catalysts; Part II: separation and
with ammoniacal solutions using ammonium thiosulfate and ammonium purification. Hydrometallurgy 98, 10–20.
sulfite as reducing agents. Hydrometallurgy 40, 111–121.

You might also like