You are on page 1of 13

Introduction and the Significance of Electrometallurgy$

JW Evans, University of California, Berkeley, CA, USA


r 2016 Elsevier Inc. All rights reserved.

Except for iron alloys and recycled metals, electrochemical reactions are used in producing much of the world’s metals. Copper and
aluminum are examples. World production of aluminum in electrolytic cells totaled 49.3 million metric tons in 2014 (Bray, 2015)
with China, Russia, and Canada being the major producers. Measured in tons, aluminum’s use exceeds that of all other metals
except iron and its alloys.
Worldwide copper production was 18.7 million metric tons with most of this production being electrorefined or electrowon
(Brininstool, 2015). Chile is the biggest producer of copper (31% of global production).
Other metals that are produced partially by electrolytic processes include cadmium, chromium, cobalt, lithium, manganese,
magnesium, nickel, and zinc.
This contribution describes the fundamental electrochemistry, thermodynamics, kinetics, and engineering phenomena that
govern electrometallurgy.

Fundamental Electrochemical Phenomena

Electronic and Ionic Conductors


Materials can be classified as nonconductors (of electricity), semiconductors, and conductors. Conduction is the movement of
electrical charges which are either electrons or ions. Metals and graphite display electronic conduction while conduction in
aqueous solutions or molten salts is ionic.

Electrochemical Reactions
Electrochemical reactions entail the transfer of electrons to or from a molecule, atom, or ion at an interface between an electronic
conductor, the electrode (through which the electrons reach or leave the interface), and an ionic conductor (through which the
ions travel). For example, copper can be deposited from aqueous solutions by the reaction

Cu2þ þ 2e ¼ Cu ½1

As illustrated on the left of Figure 1, addition of electrons to a species is reduction in the chemist’s terminology and reaction [1]
is termed a cathodic one; while this reduction is occurring at an electrode we term it a cathode. Reaction [1] can be reversed by
raising the potential of the electrode on which the copper has been deposited, whereupon oxidation (removal of electrons from
copper atoms to produce copper ions) occurs and this is then an anodic reaction occurring at an anode.

Current Density
Electrochemical reactions occur at interfaces; they are inherently heterogeneous and an appropriate measure of reaction rate is the
number of ions reduced (or produced) per unit interface area per unit time. From Faraday’s law (see below) this is directly linked
to the current per unit area, i.e., the current density, expressed in mA cm2 or A m2. It will be seen subsequently that this
heterogeneous nature is the bane of electrometallurgy; to produce a metal by an electrochemical route a very large electrode surface
area must be provided.

Electrochemical Equilibrium and Standard Electrode Potentials


Let us consider an electrode of a metal, say copper, inserted into an aqueous solution containing the ion of that metal, i.e., the
cupric ion Cu2 þ , a cation (identified as such by its positive charge). The aqueous solution will contain other ions besides the
cations, notably an anion (SO24 , for example). To an extremely good approximation the charges on the cations total to the charges
on the anions so that, in the bulk of the solution, the solution is electrically neutral. Consider now an electrical potential that we
might measure using a voltmeter. We should not expect the piece of copper to be at the same potential as the electrolyte; they are
chemically very different. There will be a difference in potential that will depend on parameters such as temperature, composition
of the electrolyte, and on whether any electrochemical reaction is occurring. Let us restrict our attention to a circumstance where


Change History: April 2015. J.W. Evans updated the production statistics in the first two paragraphs (along with references). New figure (Figure 6) is included.
‘Energy consumption’ and ‘magnesium production’ have been updated. New reference is included and minor editorial changes made in the article.

Reference Module in Materials Science and Materials Engineering doi:10.1016/B978-0-12-803581-8.03594-3 1


2 Introduction and the Significance of Electrometallurgy

External
circuit
2ε- 2ε-

Oxygen
bubble

Electrolyte

Cu2++ 2ε-= Cu

H2O = 2H++ 1/2O2+ 2ε-

Cathode “Inert” anode

Figure 1 Schematic diagram of an electrolytic cell in which copper is being deposited at a cathode and oxygen generated at an anode.

there is no net reaction so that equilibrium exists. We expect the solution to have the same potential everywhere if its composition
and temperature are uniform, except very close to the electrode where its presence would disturb this potential. The potential
difference between electrode and solution will still depend on the electrolyte and composition but, if the temperature is 25˚C and
the only cation present is copper at a concentration of one mole per kilogram of water (a one molal solution), then the potential
difference is fixed at 0.34 V (metal potential  electrolyte potential). This is the ‘standard half cell potential,’ or ‘standard electrode
potential,’ for copper and values for other metals in contact with one molal solutions are given in Table 1.

Table 1 Standard electrode potentials

E0 (V)
3þ 
Au þ 3e ¼Au 1.50
O2 þ 4H þ þ 4e  ¼2H2O 1.23
Hg2 þ þ 2e  ¼Hg 0.80
Ag þ þ e  ¼ Ag 0.80
Cu þ þ e ¼Cu 0.52
O2 þ 2H2O þ 4e ¼40H 0.40
Cu2 þ þ 2e ¼Cu 0.34
H þ þ e  ¼1/2H2 0.000
Fe3 þ þ 3e ¼Fe  0.04
Pb2 þ þ 2e ¼Pb  0.13
Sn2 þ þ 2e ¼Sn  0.14
Ni2 þ þ 2e ¼Ni  0.23
Co2 þ þ 2e ¼Co  0.27
Cd2 þ þ 2e ¼Cd  0.40
Fe2 þ þ 2e ¼Fe  0.44
Cr3 þ þ 3e ¼Cr  0.74
Zn2 þ þ 2e ¼Zn  0.76
Cr2 þ þ 2e ¼Cr  0.91
Mn2 þ þ 2e ¼Mn  1.05
Al3 þ þ 3e ¼Al  1.66
Ti2 þ þ 2e ¼Ti  1.75
Mg2 þ þ 2e ¼Mg -2.38
Na þ þ e ¼Na  2.71
Ca2 þ þ 2e ¼Ca  2.84
K þ þ e ¼K  2.92
Li þ þ e ¼Li  3.01

Source: Adapted from Bard et al. (1980).


Introduction and the Significance of Electrometallurgy 3

Note that in Table 1 there are nonmetallic electrodes. Oxygen can be reduced in an aqueous solution when the oxygen is
bubbled across the surface of an electronic conductor. Usually this conductor is catalytic (e.g., a strip of platinum) for the
reduction reaction and also serves as an electrical connection between the bubbles and the outside world. Similarly the hydrogen
ion can be reduced to produce hydrogen gas on a strip of platinum. More commonly this reaction is used in reverse with hydrogen
bubbled across the surface of the platinum. The platinum catalyzes

H2 ¼ 2Hþ þ 2e ½2

If the solution is at 25 1C and one molal in hydrogen ions, with hydrogen bubbles at atmospheric pressure, then we have a
standard hydrogen electrode (SHE) and the platinum has the potential in Table 1. This last potential is exactly zero! This results
from an arbitrary choice. It is impossible to use a voltmeter to measure the potential between a metal and an electrolyte without
immersing a second electronic conductor into the electrolyte. Therefore only differences in half cell potentials can be measured.
This difficulty is resolved by arbitrarily making the half cell potential of the SHE zero, whereupon all the entries in Table 1 are
potential differences between the metal electrode and the SHE.

Nernst’s Equation
What of the potential of a metal electrode under conditions where the solution is not one molal in the corresponding ion? The
answer is provided by Nernst’s equation
 
RT aO
E ¼ E0 þ ln ½3
nF aR

Here E is the potential of the electrode (electrode potential – solution potential), E0 is the standard electrode potential, R the
gas constant (8.3144 J mol1 K1), T absolute temperature, n the number of electrons in the reduction reaction (n ¼ 2 for reaction
[1], for example), F is Faraday’s constant (96 484 C mol1), aO is the activity of the oxidized species (e.g., activity of cupric ions in
the solution in the case of reaction [1]), and aR the activity of the reduced species. In the case of the deposition of metals aR is the
activity of the solid or liquid metal and, if the metal is close to pure this activity can be taken as unity.
For the present purposes it will be sufficient to regard the activity as equivalent to a concentration.
For reactions that are not the simple deposition of a metal, more complicated forms of Nernst’s equation apply. For example,
for the oxidation of water to produce oxygen

2H2 O ¼ 4Hþ þ O2 þ 4e ½4

 
RT pO2 a4Hþ
E ¼ E0 þ ln ½5
4F aH2 O

Here oxygen activity is taken as its partial pressure in atmospheres (pO2 ). The activity of water, aH2 O can be approximated as
unity provided the aqueous solution is not too concentrated.

An Electrochemical Cell
Consider the system sketched in Figure 1. The solution is one molal in Cu2 þ ions and H þ ions; oxygen bubbles are in contact with
an inert metal (say lead) at one atmosphere pressure. The copper and inert metal are connected via an external circuit that could be
a simple resistor or a DC power supply. If that circuit is merely an open one (e.g., a voltmeter with a very high internal resistance),
then no current flows and both electrodes are in equilibrium with the solution and at their standard electrode potentials.
Consequently

Oxygen electrode potential  Copper electrode potential ¼ 1:23 V  0:34 V ¼ 0:89 V ½6

Clearly the oxygen electrode is at a higher voltage than the copper and, if we replaced the open external circuit with a resistor,
current flows from the oxygen electrode to the copper electrode, or if the external circuit is of electronic conductors, electrons pass
from the copper to the oxygen electrode. These electrons originate from

Cu ¼ Cu2þ þ 2e ½7


4 Introduction and the Significance of Electrometallurgy

at the copper–solution interface and are consumed at the oxygen electrode by

1=2O2 þ 2Hþ þ 2e ¼ H2 O ½8

This couple of reactions is consuming rather than producing copper. If we wish to produce metallic copper from a copper
bearing solution we need to reverse them. A DC power supply must be used in the circuit to reverse the current flow so that
electrons pass from the oxygen electrode to the copper. The DC power supply must achieve at least 0.89 V between its terminals
(more if the solution is more dilute in Cu2 þ than 1 molal or if it is more acid than 1 molal in H þ ions). It is now apparent why
the production of metals by electrometallurgy consumes significant electricity. Each cell must be supplied with current at a voltage
exceeding that of the difference of two half cell potentials. The rate (Watts) at which the electrical energy is consumed is the
product of that voltage and the current.

Reaction Kinetics
So far this treatment has been largely thermodynamic, concerned with equilibrium potentials between an electrode and a solution.
Practical electrochemical reactions are seldom close to equilibrium and consideration of electrode kinetics is appropriate. An
important parameter in kinetics is the ‘overpotential,’ Z, the amount by which the potential of an electrode exceeds the equilibrium
potential. For many electrochemical reactions the rate can be expressed by the Butler–Volmer equation:
    
anF ð1  aÞnF
i ¼ i0 exp Z  exp Z ½9
RT RT

where i is the current density at the electrode, i0 is a (concentration dependent) constant, known as the ‘exchange current density,’ a
is a second constant (the ‘transfer coefficient’), and n is the number of electrons in the reaction. Cathodic overpotentials are
negative (electrode potential less than equilibrium value) and the current density at a cathode is positive.
At large (positive or negative) overpotentials one of the exponential terms on the right of this equation is insignificant and [9]
reduces to the Tafel equation

Z ¼ a þ blni ½10

These equations are usually satisfactory provided the transfer of an ion to the electrode provides no limitation to reaction.
However, it is impossible to react an ion more rapidly than it can be transferred and this imposes a ‘limiting current’ on the
electrode. Some of these features are illustrated in Figure 2 which is for two reactions: the cathodic deposition of a metal and a
second reaction, generation of hydrogen, which is frequently a competing reaction. The negative of the electrode potential is used
as the abscissa and EH4EM, the former being the equilibrium potential for hydrogen generation and the latter for metal
deposition. The corresponding rates (current densities) are iH (broken line) and iM (solid line). Both reaction rates accelerate as the
potential is moved more negative than E. The rate of metal deposition accelerates more rapidly than hydrogen evolution but
eventually reaches a limiting current (density), iLM. The current density for hydrogen increases without limit in the range sketched.

iH
Current (density)

iM

EH EM
Negative of potential
Figure 2 Sketch of the dependence of current density on potential for two cathodic reactions: the generation of hydrogen and the deposition of
metal. EH and EM are the equilibrium potentials for those reactions and iH and iM are the respective current densities. Note the limiting current
displayed by the metal deposition reaction.
Introduction and the Significance of Electrometallurgy 5

It is apparent from this figure that there is an optimum potential where the fraction of the current used to deposit metal reaches a
maximum. This fraction is increased by increasing the mass transfer of the cation to the electrode.

Transport of Ions and Supporting Electrolytes


In dilute solutions the ions move according to (Newman, 1973)

N i ¼  zi ui Fci ∇f  Di ∇ci þ ci v ½11

where
Ni is the flux of ion i
Zi is the charge on ion i (negative for anions)
ui is the mobility of ion i
F is Faraday’s constant
ci is the concentration of ion i (moles per unit volume)
is the potential
Di is the diffusivity of ion i
v is the solution velocity.

The first term on the right in eqn [11] is the migration of ion i driven by a gradient in potential, f; the second is diffusion driven
by concentration gradients; and the third is the convection of the ion due to the movement of the electrolyte. In concentrated
solutions additional terms arise.
A similar equation to [11] describes the flux of other ions and a significant fact is that the fluxes are linked, even in the case of
dilute solutions, by the equations. One notable effect of this linkage is that of the ‘supporting electrolyte.’ This is a chemical
species, for example, sulfuric acid solution in the case of aqueous electrolytes, which might not participate in either reactions but
serves to lower the cell potential. For example, in electrorefining of copper the electrolyte contains sulfuric acid. Figure 3 shows the
calculated potential and concentration gradients for a cell with planar parallel electrodes; the horizontal axis is distance from the
cathode. Parameters have been chosen to be representative of copper electrodeposition at low current density and the right hand
figure shows distributions with no supporting electrolyte (H2SO4) present. Even though the sulfuric acid is not participating in the
reaction at either electrode (and there is zero flux of H þ or SO42 ions), the acid lowers the potential difference across the cell by
its effect on the first term on the right in eqn [11].

With supporting electrolyte Without supporting electrolyte


10 40
C1 C1
C2 C2
9 Pot. Pot.
C3 35 C3
Ionic concs(millimole/cc) and potential(mV)

Ionic concs(millimole/cc) and potential(mV)

8
30
7

25
6

5 20

4
15

3
10
2

5
1

0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Distance from cathode, cms Distance from cathode, cms
Figure 3 Potential and concentration distributions calculated from dilute solution theory for a cell in which copper is anodically dissolving at one
electrode (on the right) and depositing on the cathode 2 cm away. C3 is the concentration of copper ion.
6 Introduction and the Significance of Electrometallurgy

Faraday’s Law and Current Efficiency


Let us again consider the reduction of cupric ion in aqueous solution (eqn [1]). From stoichiometry it is seen that the
passage of two electrons results in the deposition of one copper atom. As the change on an electron is 1.60219  1019 C
and there are 6.02205  1023 atoms per molar mass of copper (or anything else), we see that the passage of
2  1.60219  1019  6.02205  1023 C ¼ 2  96 484 C results in depositing one molar mass of copper. For the case where n
electrons are needed in the electrochemical reaction, for each atom (molecule) produced we conclude that n  96 484 C results in
one molar mass of electrochemistry, which is Faraday’s law. This law sets a limit on the amount of deposition occurring. In many
cases the amount of electrochemistry anticipated from Faraday’s law is not achieved for one of the following reasons:

(a) an electrochemical side reaction, such as the reduction of H þ ions from aqueous solution;
(b) an electronic conduction path between electrodes, for example, when two electrodes are on a support with a nonzero
conductivity; and
(c) chemical destruction of the metal deposited at the cathode, for example, by a chemical species.

The ratio of the amount of metal produced to that anticipated from Faraday’s law is known as the current efficiency or Faradaic
efficiency of the cathodic reaction (usually expressed as a percentage). The productivity of a cell is maximized when its current
efficiency is 100% and in the electrowinning of metals, particularly zinc electrowinning where hydrogen production occurs at the
cathode, much attention is paid to this measure of performance. For the common case sketched in Figure 4 the current efficiency
passes through a maximum as the cell current is raised (see Figure 2).
Another measure of cell performance is the energy efficiency, the ratio of the electrical energy thermodynamically required for
the desired reaction to the actual energy consumption. This efficiency is dependent on the cell voltage, which increases as cell
current is increased and, as sketched in Figure 4, the energy efficiency typically shows a maximum, although not at the same current
as the current efficiency maximum.

Aspects of the Electrodeposition of Metals

In many electrochemical reactions, for example, oxidation of organic species in solution to other species in solution, the electrode
is unchanging. In the case of metal deposition however, the very cathode surface is the metal deposited. Even starting with a
perfectly planar electrode, it would be unusual for a planar surface to be sustained as deposition proceeds. The development of
surface irregularities has both theoretical and practical significance. A uniform deposition implies a uniform current density and
(local) current efficiency; neither is likely.

Energy
consumed per
unit of
production

Voltage

Current
efficiency

Cell current (density)


Figure 4 Typical performance of cell performance parameters on current density in electrowinning of metals.
Introduction and the Significance of Electrometallurgy 7

Nucleation and Growth of Electrodeposits


Let us acknowledge that a surface that is smooth to our macroscopic senses (or under an optical microscope), is actually irregular
on an atomic scale. It may be polycrystalline and, even if it is a single crystal, it is likely to have steps at which there is a wall of
atoms as sketched in Figure 5 from Bockris and Reddy (1970). In the sketch the steps have ‘kinks.’ Deposition is thought to proceed
as follows. Cations find their way from the bulk of the electrolyte to the surface and attach to the surface. That attachment will be
accompanied by the loss of molecules solvating the ion and may entail the reduction of the ion by electrons from the metal. This
newly formed ‘adion’ (or ‘adatom’) undergoes surface diffusion until it reaches a step. Now it migrates to a kink and becomes
bound on three sides. In this way kinks move along steps, advancing the steps by one atom, and steps sweep across surfaces.
Occasionally atoms adsorbed on the top surface cluster to form a nucleus from which a new layer of atoms can grow by the
advancement of a step in several directions. This nucleation proceeds more rapidly at higher overpotentials because diffusion of
adions to kinks cannot keep up with the arrival of ions at the surface.
Figure 5 has a step that is the height of one atom and would not be visible in an optical microscope. With many metal deposits,
steps are visible in the optical microscope and these ‘macrosteps’ are thought to form (to the height of many hundreds of atoms)
by microsteps catching up to one another.
This simple picture becomes more complicated on a polycrystalline surface where the crystals are not oriented and surfaces of
different orientation may grow at different rates. Atoms must be incorporated into grain boundaries which may migrate and other
complexities of phase transformations occur.

Deposition on Heterogeneous Surfaces and Underpotential Deposition


The situation is also more complicated in the case of the initial deposition of metal onto an electrode of a different material,
copper onto platinum for example. The energetics of the deposition of the first layer of atoms (perhaps the next one or two) are
likely to be different than those of the subsequent layers. Equation [3] shows us that the activity of the reduced phase aR affects the
equilibrium potential. In the case of homogeneous deposition (say copper onto copper) that activity is one. However, in the case
of deposition onto a different metal, the depositing metal may be present only as a fraction of a monolayer and its activity will be
less than one. Consequently the equilibrium potential is shifted upward and the first monolayer of the metal will deposit at a
potential that is less cathodic than required for deposition of subsequent layers. This is known as ‘underpotential’ deposition.

Development of Electrodeposit Morphology


The distribution of current across an electrode determines the distribution of the metal deposition. The electrochemist speaks of
‘primary,’ ‘secondary,’ and ‘tertiary’ current distribution. The first is the distribution calculated when kinetic effects and mass
transfer effects at the electrodes are ignored. Then each electrode surface can be treated as being at uniform potential, assuming that
the electrode material itself has negligible resistivity. Furthermore, conduction in the electrolyte obeys Laplace’s equation
(assuming uniform electrolyte conductivity) and solutions to this equation from mathematical physics can be exploited. For
example, the primary current distribution for planar parallel electrodes, immersed in an electrolyte extending beyond their edges,
has an infinite current density (infinite deposition rate) at the edges. Under these conditions where a uniform deposit might have
been expected the opposite results!

Figure 5 Sketch of the deposition of a solvated ion from solution onto a growing metal deposit. Arrival of the adion at the surface is followed by
diffusion across the surface to a step and thence to a kink site (after Bockris and Reddy, 1970).
8 Introduction and the Significance of Electrometallurgy

A secondary distribution is one allowing also for electrode kinetics. In the case of the plane parallel electrodes discussed above
this allowance brings the edge current density to a finite value but one still greater than in the center of the electrode, unless the
reaction kinetics are slow.
The tertiary distribution is one with an additional allowance for mass transfer to the electrode. If mass transfer to the surface is
uniform, then this limitation provides a leveling influence on the deposition. However, in many cases, for example, near a leading
edge, there are significant variations in mass transfer with position and the tertiary distribution is nonuniform. At high current
densities, near the limiting current density, nonuniform deposits are the norm.
We have treated deposit morphologies at the nanoscale and at the length scale of the electrode. Now consider what happens on
the intermediate scale of deposit surface roughness. A protuberance on the surface of a deposit will grow preferentially when the
primary current distribution holds. The current lines can converge on the tip of the protuberance, increasing the current density
and the deposition rate there. Preferential growth also occurs when convective mass transfer controls the deposition. Then a
concentration boundary layer exits near the electrode surface and protuberances, sitting higher in the boundary layer, experience
higher concentrations of the depositing ion. This suggests that operation near the limiting current results in rough deposits or the
occurrence of dendrites (such as shown in Figure 6). Preferential deposition on certain crystal faces can also contribute to dendrite
formation.
Nucleation can have a major impact on deposit morphology. At high overpotential the electrode will experience a high
nucleation rate, compared to the subsequent growth rate of the nuclei. The deposit will be fine-grained but, if the high over-
potential, near the limiting current, is maintained then the irregular deposits described above will occur. Consequently there has
been much use of conditions where the overpotential is varied (pulsed deposition and periodic current reversal), particularly in
electroplating. In industrial electrodeposition of metals, such as copper, organic additives (animal glues for example) are used as
‘leveling agents’; probably the mechanism of action of these agents is one retarding the kinetics of the growth.

Electrodeposition of Metals from Aqueous Solutions

Electrometallurgy distinguishes between electrowinning and electrorefining. In electrowinning, the chemical species yielding the
metal enters the cell as a compound, for example, in a solution pumped into the cell. In contrast, in electrorefining, the metal
enters the cell as its element (in impure form) and forms the anodes within the cells.

Copper Electrowinning
Typical of electrowinning from aqueous solutions is copper electrowinning. The electrolyte entering the cells contains 25–60 g l1
of copper, 50–180 g l1 of sulfuric acid, and 5–10 g l1 of iron salts (Habashi, 1998). It is produced by leaching of copper ores,
with solution purification by solvent extraction.
The desired cathodic reaction is reaction [1] and the anodic reaction is [4] so that both oxygen and acid are generated at the
anode and electrolysis results in the replacement of copper ions by hydrogen ions. A cathodic side reaction is hydrogen generation
by

2Hþ þ 2e ¼ H2 ½12

This reaction is thermodynamically unfavorable compared to copper deposition, as indicated by Table 1 and, in terms of
Figure 2, EM would lie well to the left of EH. Consequently little side reaction occurs and copper electrowinning cells operate at

Figure 6 Zinc dendrites produced by electrodeposition from an aqueous potassium Zincate electrolyte (from Keist, 2013).
Introduction and the Significance of Electrometallurgy 9

Figure 7 Sketch of a copper electrowinning cell from Evans and DeJonghe (2002). The cell is seen from the side and in section with
interspersed anodes and cathodes suspended into the electrolyte.

close to 100% current efficiency. The anodes are of a lead alloy that is intended to be inert although traces of lead are typically
present in the copper cathodes. In recent years titanium anodes with a platinum group coating that is catalytic for oxygen
evolution have been introduced at many plants (see Section Emerging Technologies). A sketch of a typical copper electrowinning
cell appears as Figure 7. The anodes and cathodes (seen edge on in Figure 7) are approximately 1 to 2 m2 and are interleaved as
shown. The cathodes in a cell are electrically in parallel, as are the anodes. Each cell is in series with adjacent cells with the cathodes
of one cell connected to the anodes of the next cell. The cathodes are placed in the cell as ‘starter sheets,’ i.e., thin sheets of copper.
These sheets grow until they are of the order of 1 cm thick, whereupon they are harvested by an overhead crane and replaced with
new starter sheets. Starter sheets are made in separate cells by depositing copper onto stainless steel cathodes from which it can be
pealed. This has led to an alternative technology without starter sheets where the electrowinning cells have ‘cathode blanks’ of
stainless steel. When the copper deposit on the cathode blank is thick enough the blank and copper are withdrawn from the cell,
the two separated mechanically and the blank returned to the cells. Cathodes are subsequently melted and cast into ‘wirebars’ or
coiled rod for sale.
One major difficulty of copper electrowinning is the low current densities at the electrodes (typically 200–300 A m2). This,
and Faraday’s law, results in the need for enormous electrode surface areas, large numbers of cells and large buildings (‘tank-
houses’) with resulting large capital costs. The result is that the electrowinning step is the most capital intensive in many copper
plants.
A second difficulty is the mechanical harvesting of the cells and placement of starter sheets (or cathode blanks) adding
significantly to capital and operating costs.
A third difficulty is the ‘acid mist’ above the cells generated from droplets of electrolyte ejected by bursting gas bubbles. This is
detrimental to workers and forces significant expenditures on ventilation.
Finally, uneven cathode deposits can grow to the point of shorting cells and a significant amount of labor is involved in tending
the cells to avoid them.

Zinc Electrowinning and the Importance of Reaction Kinetics


Section Reaction Kinetics, in particular Figure 2, discussed the difficulty of electrodepositing a metal when the generation of
hydrogen is a thermodynamically more favorable alternative. This is the case for zinc. Table 1 reveals that EM in Figure 2 lies well to
the right of EH for all practical concentrations and we should expect copious hydrogen evolution with little, if any, zinc deposition
on electrolyzing an acidified zinc bearing solution. Fortunately the kinetics of hydrogen evolution on a pure zinc surface are very
slow; in terms of Figure 2 the climb of the broken line is gentle. Because of this kinetic limitation, it is possible to run zinc
electrowinning cells at current efficiencies of around 90%. The caveat is that the zinc sulfate electrolyte must be low in impurities
and not concentrated in acid. Some impurities, for example, antimony and arsenic, must be removed to levels of parts per billion
because they are catalytic for hydrogen evolution. Current densities, temperature, and the presence of organic additives all affect
current efficiency. Even the morphology of the deposited zinc has an effect, perhaps as a consequence of rough deposits providing
nucleation sites for hydrogen bubbles.
Zinc electrowinning cells are similar to copper electrowinning cells (except that aluminum cathode blanks are used rather than
stainless steel). Cell voltages are typically 3–4 V with current densities of 200–500 A m2. Electrolytes (from leaching and solution
10 Introduction and the Significance of Electrometallurgy

purification) contain a few tens of grams per liter of zinc (as zinc sulfate) and similar amounts of acid. Within the cells roughly half
of the zinc is removed and the sulfuric acid concentration increases to, perhaps, 200 g l1.
Other metals produced by routes entailing electrowinning from aqueous solution include cobalt, nickel, cadmium, chromium,
gold, silver, and manganese.

Electrorefining of Copper
The common route for turning copper sulfide ores into metal is high temperature oxidation of sulfides to produce an impure
liquid copper that is cast into slab-like anodes. Principal impurities are iron and precious metals (gold and silver). The anodes are
electrolytically refined in cells sketched in Figure 8. There is a resemblance to the cell of Figure 7. Both are constructed of acid
resistant concrete and contain cathodes (connected in parallel) interleaved with anodes (also connected in parallel). The principal
difference lies in what occurs at the anodes: oxygen evolution at the electrowinning anodes, dissolution of copper – the reverse of
reaction [1], at the electrorefining anodes.
Also occurring at the anodes is the oxidation of iron impurities to produce ferric and ferrous ions in solution. The anodes are
dissolved to the point at which they start to fall apart. The residuum is removed from the cell and re-melted/cast into fresh anodes
and fresh anodes are placed in the cell. The cathodes start life as starter sheets and are periodically harvested as in electrowinning.
Alternatively stainless steel cathode blanks are used as described above under Copper Electrowinning. A glance at Table 1 shows
that there is little tendency for the iron in the electrolyte to co-deposit with the copper; copper deposition is thermodynamically
much more favorable. Precious metal impurities in the anode are not oxidized, at the anode potentials employed, but fall to the
bottom of the cell as an ‘anode slime’ which is occasionally harvested from the cells and sent for recovery of precious metals.
Adjacent cells are connected in series and several hundred cells would be contained in a typical electrorefining tankhouse.
Typical operating conditions for copper electrorefining cells are (Habashi, 1998): 100–200 A m2 at a cell voltage of 0.3 Vwith an
electrolyte of 40 g l1 of copper and 150–200 g l1 of sulfuric acid.

Other Electrorefining Operations


In one technology for producing nickel, electrorefining of impure nickel sulfide (‘nickel matte’) is used. The cell is divided by a
porous partition allowing passage of current with minimal intermixing of the electrolytes between anode and cathode com-
partments. Sulfur, other particulate and dissolved impurities produced in the former chamber by the dissolution of the matte are
pumped out of the cell with the electrolyte and removed. The electrolyte is then pumped back to the cell.
Precious metals are also electrorefined, silver using nitrate electrolytes and stainless steel cathodes, gold using chloride elec-
trolytes and thin gold ribbon cathodes (Habashi, 1998).

Figure 8 Sketch of a copper electrorefining cell from Evans and DeJonghe (2002). Interspersed copper anodes and cathodes are suspended in an
electrolyte.
Introduction and the Significance of Electrometallurgy 11

Electrodeposition from Molten Salts

The Table 1 of standard electrode potentials shows increasing difficulty of electrowinning of metals from aqueous solutions.
Referring to Figure 2, the point EM moves to the right as we move down the table. Copper is easily electrowon, cobalt with a little
difficulty, zinc with considerable difficulty. At aluminum, electrodeposition from aqueous solution becomes impossible; hydrogen
is the only thing generated at the cathode (by reaction [12]). The solution to this difficulty is to electrodeposit the metal from a
solution containing no hydrogen ions and this is employed in the primary technology for producing aluminum, the ‘Bayer–Hall
process.’

The Hall–Héroult Cell for Producing Aluminum


Pure aluminum oxide, produced from the ore ‘bauxite’ in the Bayer process is dissolved in a molten salt electrolyte at about 960 1C
within a ‘Hall–Héroult’ cell depicted in Figure 9.
Hall–Héroult cells are a few meters across (the horizontal dimension in Figure 9) and 10–20 m long. They are of a steel shell
lined with refractory material and fitted with a carbon bottom through which steel ‘collector bars’ pass. Above the bottom is a pool
of molten aluminum and above that molten salt electrolyte (‘bath’) containing sodium fluoride, aluminum fluoride, and other
fluorides in lesser amounts, along with dissolved Al2O3 (a few percent). Dipping into the bath are carbon anodes suspended from
a superstructure permitting the anodes to be raised or lowered and bringing current to the cell. A cell contains 20–30 anodes. The
anodic reaction is the generation of CO2 by the discharge of oxide and oxy-fluoride anions which, with the carbon anode, form
this gas. The cathodic reaction, at the interface between the bath and aluminum, is the discharge of cations to form aluminum
metal, which joins the liquid pool. Every few hours, aluminum is siphoned from the cell and anodes are lowered and periodically
replaced as they are consumed. The current, of a few hundred kiloamps, passes down through the aluminum, through the bottom
to the collector bars and then to the superstructure of the next cell. A plant might have a hundred cells in series (a ‘potline’) with
two or four potlines producing a total of 100–300 000 t per year of metal.
The current efficiency of a typical Hall–Héroult cell is well above 90% but the energy efficiency seldom reaches 50%. Because
primary aluminum production is a significant consumer of electrical energy and because of the charge for electricity is for power,
rather than current, and is increasing in many developed countries, the latter percentage has long been a cause for concern for over
a century. Over the twentieth century the energy expenditure dropped from approximately 27 kWh per kg of aluminum to
approximately 15 kWh per kg at the end of the century (Evans and DeJonghe, 2002) and to 14.6 kWh per kg in 2013 (Interna-
tional Aluminium Institute). Much of the electrical energy is consumed in passing current through the resistance of the electrolyte

Figure 9 Sketch of a Hall–Héroult cell for the production of aluminum from Evans and DeJonghe (2002). Carbon anodes are suspended in a
molten salt electrolyte above a pool of liquid aluminum contained in a carbon lined cell. Current flows from the cathode bus to the anodes of the
next cell.
12 Introduction and the Significance of Electrometallurgy

between the anodes and the aluminum (the anode–cathode distance or ‘ACD’). This ACD is approximately 4 cm and its reduction,
say to 2 cm, would save considerable electrical energy. Unfortunately such a reduction is precluded by the necessity of avoiding
shorting aluminum to the anodes. An ACD of 4 cm provides a margin of safety for the uneven surface of the anode (from uneven
carbon consumption) and for deformation of the bath–metal interface under magnetohydrodynamic forces. Additionally, the
thermal energy generated in the ACD is then sufficient to compensate for heat losses as well as the thermal energy required for the
reactions.
Reduction of the ACD requires the positions of the bottoms of the anodes to be well defined and also the bath–metal interface.
The former might be achieved using an anode of an inert anode material. This anode would generate oxygen, rather than CO2,
which could be an additional environmental advantage. Limited success has been achieved, in laboratory and pilot-scale cells, with
ceramic, metallic, and cermet anodes (Xiao et al., 1996).
The bath–aluminum interface can be stabilized by making the aluminum a thin film on a wetted surface, the film draining to a
sump to be siphoned off. One of the few electronic conductors that is wetted by aluminum, but insoluble in it, is titanium
diboride. This material or TiB2–carbon composites have met with some success in laboratory and pilot-scale cells (Tabereaux et al.,
1998).

The Production of Magnesium


Magnesium is produced by one of two routes: nonelectrolytic ones (the ‘Pidgeon process’ or ‘Magnetherm process’) and by the
electrolysis of magnesium chloride dissolved in a mixture of molten chlorides. Steel cathodes and graphite anodes are used in
the latter. In one variant anhydrous magnesium chloride is fed to the cell, the only anodic reaction is chlorine generation and the
anodes are therefore essentially inert. In a second variant the cost of completely dehydrating the MgCl2 is avoided and partly
dehydrated salt is fed to the cell. Now both chlorine and oxides of carbon are generated at the graphite anodes, which must be
periodically replaced. Steel has no solubility in liquid magnesium and the magnesium floats off the cathodes and is removed from
the cell top.
In recent years there has been a tendency for the electrolytic production of magnesium to be replaced by the older Pigeon
process, particularly in China.

Emerging Technologies

Improvements in electrometallurgy fall into three categories: improved materials, novel electrochemistry, and advances in engi-
neering. Section ‘The Hall–Héroult Cell for Producing Aluminum’ treated the improvements in the Hall–Héroult cell which could
result from inert anodes and cathodes wetted by aluminum. Another material that may find widespread use in electrowinning
metals from aqueous solution is DSAs. This ‘dimensionally stable anode’ consists of an expanded titanium mesh with a pro-
prietary coating that catalyzes oxygen evolution. The mesh is designed so that oxygen bubbles evolve from the back of the mesh
(rather than the side facing the cathode) minimizing their interference with the current. Lower cell voltages are reported with DSA
and the contamination of the cathode with lead from anodes is avoided.
One example of novel electrochemistry is the use of low temperature, aprotic solvents for electrowinning and electrorefining of
aluminum. Reddy and colleagues (Wu et al., 2000, 2001; Kamavaram and Reddy, 2002) have used ionic liquids as electrolytes. An
impure aluminum anode was electrorefined using a 1-butyl-3-methylimidazolium chloride and anhydrous AlCl3 electrolyte with a
copper cathode. The same electrolyte was used to electrowin aluminum from AlCl3 using a graphite anode.
Another novel electrochemistry of note if that of Fray and colleagues (Fray, 2000; Chen et al., 2000) for the production of
titanium and other metals that cannot be electrowon from aqueous solution. This is an alternative to the cumbersome and
expensive Kroll process (Newton, 1959) used, with little change, since WWII for producing this metal. Their approach is an
electrochemical deoxidation of pellets of TiO2 in a molten salt electrolyte.
Finally there is scope for improved engineering of the cells depicted in Figures 7 and 8. These cells result in huge and expensive
tankhouses because of the need for the large electrode area imposed by current density limitations. However, the slab-like
electrodes in these figures have a very low surface area per unit volume. The necessary surface area can be accommodated in much
smaller equipment if the slabs are replaced with beds of particles. In electrodeposition the particles must be kept moving and work
has been done on fluidized and spouted bed electrodes for metal deposition (Masterson and Evans, 1982; Salas-Morales et al.,
1997a,b; Evans and Jiricny, 2011).
Improvements of this kind and the evolutionary improvement of existing technology will ensure that electrometallurgy will
play a significant role in metals production in years to come.

References

Bard, A.J., Faulkner, L.R., 1980. Electrochemical Methods Fundamentals and Applications. New York, NY: John Wiley & Sons, Inc.
Introduction and the Significance of Electrometallurgy 13

Bockris, J.O.M., Reddy, A.K.N., 1970. Modern Electrochemistry Vol. 2: An Introduction to an Interdisciplinary Area. New York NY: Plenum Press.
Bray, E.L., 2015. U.S. Geological Survey, Mineral Commodity Summaries (http://minerals.usgs.gov/product.html).
Brininstool, M., 2015. U.S. Geological Survey, Mineral Commodity Summaries (http://minerals.usgs.gov/product.html).
Chen, G.Z., Fray, D.J., Farthing, T.W., 2000. Direct electrochemical reduction of titanium dioxide to titanium in molten calcium chloride. Nature 407, 361–364.
Evans, J.W., DeJonghe, L.C., 2002. The Production and Processing of Inorganic Materials. Warrendale, PA, USA: The Metallurgical Society.
Evans, J.W., Jiricny, V., 2011. Spouted bed electrochemical reactors. In: Epstein, N., Grace, J.R. (Eds.), Spout and Spout-Fluid Beds. Cambridge University Press, pp. 269–282.
Fray, D.J., 2000. Aspects of technology transfer. Metallurgical and Materials Transactions, B 31, 1153–1162.
Habashi, F., 1998. Principles of Extractive Metallurgy Vol. 4: Amalgam and Electrometallurgy. Saint-Foy, Québec, Canada: Métallurgie Extractive Québec.
International Aluminium Institute. Available at: http://www.world-aluminium.org/statistics/primary-aluminium-smelting-energy-intensity/#data
Kamavaram, K., Reddy, R.G., 2002. Electrochemical studies of aluminum deposition in ionic liquids at ambient temperatures. In: Schneider, W. (Ed.), Light Metals 2002.
Warrendale PA, USA: The Minerals, Metals and Materials Society, pp. 253–258.
Keist, J., 2013 In-situ analysis of zinc electrodeposition within an ionic liquid electrolyte. PhD Dissertation, University of CA, Berkeley.
Masterson, I., Evans, J.W., 1982. Fluidized bed electrowinning of copper; experiments using 150 amp and 1,000 amp cells and some mathematical modeling. Metallurgical
Transactions, B 13, 3–13.
Newman, J.S., 1973. Electrochemical Systems. New York, NY: John Wiley & Sons, Inc.
Newton, J., 1959. Extractive Metallurgy. New York, NY: John Wiley & Sons, Inc.
Salas-Morales, J.C., Evans, J.W., Newman, O.M.G., Adcock, 1997a. Spouted bed electrowinning of zinc; part I laboratory-scale electrowinning experiments. Metallurgical and
Materials Transactions, B 28, 59–68.
Salas-Morales, J.C., Evans, J.W., Verma, A., 1997b. Spouted bed electrowinning of zinc; part II investigations of the dynamics of particles in large thin spouted beds.
Metallurgical and Materials Transactions, B 28, 69–79.
Tabereaux, A., et al., 1998. The operational performance of 70 kA prebake cells retrofitted with TiB2-G cathode elements..In: Welch, B. (Ed.), Light Metals 1998. Warrendale PA,
USA: The Minerals, Metals and Materials Society, pp. 257–264.
Wu, B., Reddy, R.G., Rogers, R.D., 2000. Aluminum recycling via near room temperature electrolysis in ionic liquids. In: Stewart, D.L., Stephenes, R., Daley, J.C. (Eds.), Light
Metals 2000. Warrendale PA, USA: The Minerals, Metals and Materials Society, pp. 237–243.
Wu, B., Reddy, R.G., Rogers, R.D., 2001. Aluminum reduction via near room temperature electrolysis in ionic liquids. In: Anjier, J.L. (Ed.), Light Metals 2001. Warrendale PA,
USA: The Minerals, Metals and Materials Society, pp. 237–243.
Xiao, H., Hovland, R., Rolseth, S., Thonstad, J., 1996. Studies of the corrosion and behavior of inert anodes in aluminum electrolysis. Metallurgical and Materials Transactions,
B 27, 185–193.

Further Reading

Biswas, A.K., Davenport, W.G., 1980. Extractive Metallurgy of Copper, second ed. Oxford, UK: Pergamon Press.
Doyle, M., Arora, P., 2001. Report on the electrolytic industries for the year 2000. J. Electrochem. Soc. 148, K1–K41.
Evans, J.W., 1995. Electricity in the production of metals; from aluminum to zinc. Metallurgical and Materials Transactions B 26, 189–208.
Grjotheim, K., Kvande, H., 1986. Understanding the Hall-Héroult Process for Production of Aluminium. Düsseldorf, Germany: Aluminium-Verlag Gmbh.
Grjotheim, K., Welch, B.J., 1980. Aluminium Smelter Technology. Düsseldorf, Germany: Aluminium-Verlag Gmbh.
Hine, F., 1985. Electrode Processes and Electrochemical Engineering. New York, NY: Plenum Press.
Pawlek, R.P., 2015. Primary Aluminum Industry in the Year 2014, Light Metal Age, pp. 9−21.
Pletcher, D., 1984. Industrial Electrochemistry. London: Chapman and Hall.
Roušar, I., Micka, K., Kimla, A., 1986. Electrochemical Engineering. Prague, Czechoslovakia: Academia.

You might also like