Sie sind auf Seite 1von 29

22

Electrolytic Processes
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-1
22.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-2
22.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-2
Electromotive Force and Driving Force for Electrolytic
Reactions Activity, Concentration, and Temperature
Dependence; Pourbaix Diagrams Activation Overpotential
and Tafel Relationship Limiting Current Density, Nernst
Diffusion Layer, Reaction Kinetics Other Factors
Contributing to Overpotential: Hydrogen Overvoltage
Electrocrystallization

22.3 Surface Finishing. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-9


Electrodeposition Electroless Deposition Displacement
Deposition Chemical Conversion Anodizing

22.4 Direct Manufacturing Using Electrolytic Processing . . 22-16


Large-Scale Manufacture Small-Scale Manufacture

22.5 Primary Metal Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-18


Electrowinning and Electrorefining Molten Salt Electrolysis

22.6 Electrolytic Processing of Nanostructures . . . . . . . . . . . . . 22-20


Monolithic Bulk Metal Nanostructures Structurally Graded
Nanomaterials Nanocomposites CMA Nanostructures
Nanocrystalline Particles Nanocrystalline Oxide
Powders Template Nanotechnology by Anodizing

Uwe Erb
University of Toronto

22.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-26


References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22-26

Abstract
This chapter deals with the synthesis of metals, alloys, and composite materials by electrolytic processes using
aqueous solutions. The first part of the chapter covers the fundamentals of electrochemical reactions including
electromotive force and driving force for electrolytic reactions, overpotential, reaction kinetics, and electrocrystallization. This is followed by several examples of processes in the three major application areas: surface
finishing, direct manufacturing, and primary metal production. The final section deals with the application of
electrolytic processes in the synthesis of various nanostructured materials such as bulk metal nanostructures,
nanocrystalline particles, and template materials.
22-1

2007 by Taylor & Francis Group, LLC

22-2

Materials Processing Handbook

Primary
Production

Electrochemical
Synthesis
Surface
Finishing

FIGURE 22.1

Direct
Manufacturing

Main application areas of material processing by electrolytic methods.

22.1 Introduction
On the basis of the underlying fundamental chemical and physical principles, most material processing
routes can be broadly categorized into one of five major groups. These are (1) solid-state processing (e.g.,
rolling, extrusion, severe plastic deformation, surface deformation); (2) liquid-phase processing (e.g.,
casting, rapid solidification, atomization); (3) vapor-phase processing (e.g., chemical or physical vapor
deposition, molecular beam deposition); (4) chemical processing (e.g., precipitation, replacement); and
(5) electrolytic processing (e.g., electrodeposition, calcination, electrowinning). Each processing route has
unique features and is specifically designed to process distinct groups of materials in the most economical
way. This section deals with electrolytic processing of materials which, in many respects, is similar to
chemical processing in terms of the chemical reactions involved. However, the distinguishing feature of
electrolytic processing is that it involves a well-defined interface at which charge transfer takes place.
Most electrolytic reactions have clearly separated cathodic and anodic reactions. In processes such as
electrodeposition or electrowinning an external power supply is required to drive the reaction, while in
cases such as electroless deposition or galvanic conversion the electrons for metal ion reduction come
either from a reducing agent added to the base electrolyte or from electrochemical displacement reactions.
The three major areas in which electrolytic processing is applied in industry are primary production of
materials, surface finishing, and direct manufacturing (Figure 22.1). This chapter will first present some of
the fundamental aspects of electrolytic processing by discussing issues such as electrode potential, charge
transfer, and structure formation (Section 22.2). This will be followed by a description of specific examples
from the three major application areas (Section 22.3 to Section 22.5). Finally, examples of applications of
electrolytic processing in the synthesis of specific nanostructured materials will be presented (Section 22.6).
This chapter is mainly concerned with metallic materials such as pure metals, alloys, and metal
matrix composites. Electrolytic processing of other materials (e.g., semiconductors and ceramics) is
mentioned whenever appropriate. However, a detailed description of electrolytic processing of these
structures is beyond the scope of this review. Furthermore, this chapter deals mainly with electrolytic
processes involving aqueous solutions. Electrolytic processing of materials using organic solutions, molten
salts, or ionic liquids will only be briefly addressed.

22.2 Background
This section summarizes some of the important electrochemical principles of relevance to electrolytic
processing of materials. For a more in-depth treatment of the subject, the reader is referred to the many

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-3

excellent electrochemistry texts that have been published over the past several decades. (e.g., Reference 1
to Reference 10). These give full explanations with figures and derivations of the equations. Important
terms often used in the literature which deal with electrolytic processing are given in italic print.

22.2.1 Electromotive Force and Driving Force for Electrolytic Reactions


When a metal is immersed in an aqueous electrolyte containing ions from the same metal there will be
an exchange of ions between the electrolyte and the solid metal. After a certain period of time a dynamic
equilibrium will be established at which the number of metal ions entering the solution from the solid is
the same as the number of metal ions entering the solid from the solution:
Mz+ + ze M

(22.1)

where M is the solid metal, Mz+ is the charged ion, z is the number of electrons involved and e is the
charge of the electron. Reaction from the left to right is called a reduction reaction. Conversely, the right to
left reaction is referred to as oxidation reaction. At equilibrium the magnitudes of the currents associated
with the anodic and cathodic reactions are the same, but in opposite directions. The current per unit
area is referred to as the exchange current density, i0 . The consequence of this is the establishment of
charge separation through an electrical double layer (Helmholtz layer) where the metal surface has a net
negative charge (inner Helmholtz plane) while the aqueous solution adjacent to it has a positive charge
(outer Helmholtz plane). Outside the Helmholtz layer is the Gouy-Chapman layer, a diffuse region of excess
electric charge that can extend over considerable distances.
As in any other chemical reaction, the driving force for electrochemical reactions can be expressed in
terms of the Gibbs energy of the system. Reactions proceed in the direction that would decrease the Gibbs
energy. For processes that involve electron transfer, the driving force is directly related to the potential of
the reaction of an electrolytic cell according to the equation describing electrochemical work:
G = z F E

(22.2)

where z is the number of electrons taking part in the reaction, F is the Faraday constant
(96,500 coulombs/mol) and E is the potential of the reaction.
The standard potential or electromotive force of a reaction can be measured using an electrochemical
cell consisting of two half cells. In the simplest form the first half cell consists of a solid metal rod, M,
immersed in a solution containing Mz+ ions, usually at a concentration of 1 mol/L. This cell is connected
to a second half cell, the standard hydrogen electrode, via an electrically conductive, but chemically inert
salt bridge. The standard hydrogen electrode (standard half cell) consists of an inert platinum electrode in
a 1 M solution of H+ ions and is saturated with hydrogen gas bubbling through the solution at a pressure
of 1 atmosphere and a temperature of 25 C. The shorthand to describe such an electrochemical cell is
demonstrated for the case of a cell consisting of a rod of zinc immersed in a solution of ZnCl2 (1 mol of
Zn2+ ions) in cell 1, connected to the standard hydrogen electrode in cell 2:
..
Zn(s) | Zn2+ .. .. H+ (aq), H2 (g) | Pt(s).

(22.3)

Here (s), (aq), and (g) describe the solid, aqueous, and gaseous states, respectively, and the two
dotted lines indicate the salt bridge connecting the two half cells. When the two half cells are electrically
connected via an external wiring circuit containing a volt meter, the potential difference between the two
half cells can be measured in units of volt. Defining the standard potential of the hydrogen electrode as
zero on the potential scale, the measured voltage against the Zn half cell is +0.76 V which is a measure
of the electromotive force. Therefore, the half cell reactions in the two cells together with their standard

2007 by Taylor & Francis Group, LLC

22-4

Materials Processing Handbook

potentials can be written as follows:


Cell 1: Zn(s) Zn2+ (aq) + 2e

E10 = 0.76V

(22.4)

Cell 2: 2H+ (aq) + 2e H2 (g)

E20 = 0V

(22.5)

The overall cell reaction is given as:


Zn(s) + 2H+ Zn(aq)2+ + H2 (g)

(22.6)

E = +0.76V + 0V = +0.76V

(22.7)

where E is the standard potential of this electrochemical cell.


The cell reaction with a positive cell potential is consistent with the fact that the zinc metal dissolves
spontaneously when immersed in an acid.
It should be pointed out that the oxidation potential of E 0 = 0.76V for Zn(s) Zn2+ (aq) + 2e is
opposite in sign as for the reduction reaction Zn2+ (aq) + 2e Zn(s) which is E 0 = 0.76V.
A relative ranking in terms of their electromotive force can be obtained by measuring cell reaction
potentials for various elements relative to the standard hydrogen electrode, as described above for the
case of zinc, and summarizing the results in a table of standard potentials at 25 C such as shown in
Table 22.1. The values given in this table are for the reduction reactions. In this table the metals at the top
TABLE 22.1

Element
Gold
Gold
Oxygen
Platinum
Iridium
Palladium
Mercury
Silver
Tellurium
Copper
Ruthenium
Oxygen
Copper
Arsenic
Rhenium
Bismuth
Antimony
Tungsten
Hydrogen
Lead
Tin
Molybdenum
Nickel
Cobalt
Thallium
Indium
Cadmium

Standard Electrode Potentials for Various Elements

Electrode reaction
Au+ + e Au
Au3+ + 3e Au

O2 + 4H+ + 4e 2H2 O
Pt2+ + 2e Pt
Ir3+ + 3e Ir
Pd2+ + 2e Pd
Hg2+ + 2e Hg
Ag+ + e Ag
Te2+ + 2e Te
Cu+ + e Cu
Ru2+ + 2e Ru
O2 + 2H2 O + 4e 4OH
Cu2+ + 2e Cu
As3+ + 2e As
Re3+ + 3e Re
Bi3+ + 3e Bi
Sb3+ + 3e Sb
W3+ + 3e W
2H+ + 2e H2
Pb2+ + 2e Pb
Sn2+ + 2e Sn
Mo3+ + 3e Mo
Ni2+ + 2e Ni
Co2+ + 2e Co
Tl+ + e Tl
In3+ + 3e In
Cd2+ + 2e Cd

Standard
potential, E
(volts), 25 C

Element

Electrode reaction

Standard
potential, E
(volts), 25 C

+1.68
+1.50
+1.23
+1.20
+1.00
+0.98
+0.85
+0.80
+0.57
+0.52
+0.45
+0.40
+0.34
+0.34
+0.30
+0.20
+0.11
+0.10
0.00
0.13
0.14
0.20
0.25
0.28
0.34
0.34
0.40

Iron
Gallium
Tantalum
Chromium
Zinc
Chromium
Niobium
Vanadium
Manganese
Zirconium
Titanium
Aluminum
Hafnium
Uranium
Beryllium
Neodymium
Cerium
Magnesium
Yttrium
Sodium
Calcium
Strontium
Rubidium
Potassium
Arsenic
Lithium
Cesium

Fe2+ + 2e Fe
Ga3+ + 3e Ga
Ta3+ + 3e Ta
Cr3+ + 3e Cr
Zn2+ + 2e Zn
Cr2+ + 2e Cr
Nb3+ + 3e Nb
V2+ + 2e V
Mn2+ + 2e Mn
Zr4+ + 4e Zr
Ti2+ + 2e Ti
Al3+ + 3e Al
Hf4+ + 4e Hf
U3+ + 3e U
Be2+ + 2e Be
Nd2+ + 2e Nd
Ce3+ + 3e Ce
Mg2+ + 2e Mg
Y3+ + 3e Y
Na+ + e Na
Ca2+ + 2e Ca
Sr2+ + 2e Sr
Rb+ + e Rb
K + + e K
As+ + e As
Li+ + e Li
Cs+ + e Cs

0.44
0.53
0.60
0.74
0.76
0.91
1.10
1.13
1.18
1.53
1.63
1.66
1.70
1.80
1.85
2.10
2.34
2.37
2.37
2.71
2.87
2.93
2.93
2.93
2.93
3.05
3.05

Note: Depending on the choice of cell diagram, the sign of the potential may be reversed in other references.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-5

are chemically inert (e.g., gold, platinum, and palladium). Moving down in the table, the metals become
more active (i.e., more susceptible to oxidation). Lithium and cesium show the highest tendencies for
oxidation. Using the values given in Table 22.1, the cell potential of any pair of redox half reactions can be
calculated. For example, for Zn(s) + Cu2+ Zn2+ + Cu(s), the two half reactions are:
reduction half reaction:

Cu2+ + 2e Cu(s);

oxidation half reaction:

Zn(s) Zn2 + + 2e;

which gives:

E0 = +0.34V
E0 = +0.76V

E = (+0.34V) + (+0.76V) = +1.103V.

(22.8)
(22.9)
(22.10)

The table of standard potentials is of considerable technological importance in several areas. It can
be used as a starting point in assessing the corrosion tendency of materials in the absence of corrosion
inhibiting surface effects. It is also widely used as a first guide in the design of electrolytic materials
processing methods to be discussed later in this chapter. However, there are a number of issues to be aware
of when using tables of electromotive force values such as shown in Table 22.1. The first has to do with
the convention regarding the sign of the potentials. Many European texts show the same values as shown
in Table 22.1, but with the opposite sign. This was discussed in great detail by Bockris and Reddy.6 While
this issue is beyond the scope of this chapter, it should be pointed out that the sign of the potentials in
Table 22.1 is consistent with the recommendation given by the International Union of Pure and Applied
Chemistry (IUPAC). Second, the values listed in the table are only valid for standard conditions. There are
a number of factors that can result in considerable changes in the potential values, such as the formation
of complex ions, formation of adsorbed layers, and the like. Third, while the standard potentials give good
information regarding the relative energy states, they provide no information with respect to the kinetics
of processes (i.e., the rates of reactions).

22.2.2 Activity, Concentration, and Temperature Dependence;


Pourbaix Diagrams
The potential of the Mz+ /M electrode as a function of metal ion activity in solution, a (Mz+ ), and
temperature, T , is given by the Nernst equation:
E = E0 +

RT
ln a(Mz+ )
zF

(22.11)

where R, T , F , and z are the gas constant, absolute temperature, Faradays constant, and number of
electrons in the reaction, respectively. For many practical applications using dilute solutions the activity,
a(Mz+ ), can be replaced with the metal ion concentration, c(Mz+ ), without introducing substantial errors.
Thus, Equation 22.11 can be approximated as follows:
E = E0 +

RT
ln c(Mz+ ).
zF

(22.12)

At 298 K (25 C), the term (RT /F ) has the value of 0.0257 V, giving:
E = E0 +

0.0257
ln c(Mz+ )
z

(22.13)

or when converting the natural logarithm to decimal logarithm (factor of 2.303):


E = E0 +

2007 by Taylor & Francis Group, LLC

0.0592
log c(Mz+ ).
z

(22.14)

22-6

Materials Processing Handbook


2
1.6
CuO

Potential, V (SHE)

1.2
Cu2+

0.8
0.4

Cu2O

Cu+

Cu

0.4
0.8
1.2
1.6
0

FIGURE 22.2

8
pH

10

12

14

16

Pourbaix diagram for the watercopper system.

Using this equation it can be shown, for example, that the electrode potential at 25 C of a copper
electrode immersed in aqueous copper sulfate solution (z = 2) changes from 0.34 V in a 1 M solution to
0.25 V in a 0.001 molar solution, that is, by 0.0592/z for each order of magnitude decrease in the ionic
concentration.
When the Nernst equation is considered in conjunction with all possible electrochemical equilibria
when a metal is in an aqueous solution containing its own ions, phase stability regions of various species
can be plotted in potential pH diagrams, also known as Pourbaix diagrams.11,12 An example of a
Pourbaix diagram for the watercopper system is shown in Figure 22.2. These diagrams are useful for
electrolytic processes (as well as in corrosion studies) to identify all thermodynamically possible phases.
However, again it is not possible to draw conclusions from such diagrams when it comes to reaction rates.

22.2.3 Activation Overpotential and Tafel Relationship


When an external potential is applied across the double layer, the electrode becomes polarized and current
will flow either as anodic or cathodic current. The difference of the equilibrium potential, E(i = 0), and
the potential of the same electrode as a result of flowing current, E(i), is usually referred to as the activation
overpotential a :
(22.15)
a = E(i) E(i = 0).
The relationship between overpotential and current is usually expressed in the form of the Tafel equations:
a = a b logi

(22.16)

where the constants a and b for cathodic () and anodic (+) reactions are given as:
anodic:

cathodic:

aa = (2.303RT /zF ) log i0

(22.17)

ba = 2.303RT /zF

(22.18)

ac = (2.303RT /(1 )zF ) log i0

(22.19)

bc = 2.303RT /(1 )zF .

(22.20)

In these equations i0 is the exchange current density and the geometrical transfer coefficient. This
activation overpotential is a direct measure of the interference of the equilibrium conditions at the electrode.
Larger exchange current densities require smaller overpotentials to produce a given net external anodic or
cathodic current.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-7

22.2.4 Limiting Current Density, Nernst Diffusion Layer, Reaction Kinetics


The Tafel equation describes the currentpotential relationship for cases where charge transfer is the ratedetermining step. However, for both anodic dissolution and cathodic deposition of metals the mobility
of the ions in the solution presents a practical limit that results in a deviation from linear Tafel behavior,
in an E-log i plot, at what is usually referred to as the concentration overpotential. Here, we only consider
the case of cathodic deposition. If the deposition is carried out too fast, the ions Mz+ from the bulk of
the electrolyte cannot diffuse fast enough to the solid surface to become incorporated into the growing
deposit. A current density limit will be reached at which the ions are deposited as soon as they arrive at
the electrode. At this point the metal ion concentration at the electrode approaches zero and any further
increase in current density will result in other electrochemical processes, such as the evolution of hydrogen
bubbles. The limiting current density, iL , for a system is given by:
iL =

nFD
c

(22.21)

where is the thickness of the so-called Nernst diffusion layer, D the bulk diffusion coefficient for Mz+
ions in the solution and c the concentration of Mz+ ions in the bulk solution.
In Figure 22.3, the change in the metal ion concentration from c , the bulk concentration in the
electrolyte to zero at the electrode surface is schematically indicated. The thickness of the transition layer
is typically on the order of micrometers. With Equation 22.21 the concentration overpotential, c , can be
given as:


2.303RT
i
log 1
(22.22)
c =
nF
iL
where the current density i is a direct measure of the actual rate of the reduction process.
Both activation and concentration polarization occur at the same electrode. At low reaction rates
(small current densities) activation polarization is predominant while higher reaction rates (high current
densities) are concentration polarization controlled.
The combined effect is summarized in the ButlerVolmer equation:
reduction = log


 

2.303RT
i
i
+
log 1
i0
i0
iL

(22.23)

where is the expression 2.303RT /zF . This equation allows us to determine the kinetics of many
reduction reactions from only three parameters: , i0 and iL .

Mez+
Mez+

Nucleation

Mez+

Growth
CMez+

ze

ze Surface Diffusion

FIGURE 22.3 Schematic diagram showing various steps involved in electrodeposition.

2007 by Taylor & Francis Group, LLC

22-8

Materials Processing Handbook

22.2.5 Other Factors Contributing to Overpotential: Hydrogen Overvoltage


There are many other factors that can affect the overpotential such as adsorped species (adsorption
overpotential), ionically conducting films or oxide films (film or resistance overpotential), formation of
complexed ions (complex ion overpotential), ad-atom diffusion, and surface nucleation (crystallization
overpotential).
One specific type of overpotential, sometimes referred to in short as the overvoltage, is the overpotential
(required polarization) at which gas evolution appears at the electrode as a result of oxidation (e.g.,
oxygen evolution) or reduction (e.g., hydrogen evolution). In particular the hydrogen overvoltage is of
great importance during electrodeposition processes (Section 22.3.1.1).

22.2.6 Electrocrystallization
When a metal is deposited from an aqueous solution onto a cathode, the hydrated ions in the solution
must go through a series of complex steps before they can be incorporated as an atom into the growing
electrodeposit. As pure metals always deposit in crystalline form (amorphous structures are only formed
in certain alloy deposits), electrocrystallization is usually used as the term to describe this series of processes.
Electrocrystallization can be considered to consist of two distinct stages.5,6 The first stage, referred to as
the deposition stage, is the mass transfer through the Nernst diffusion layer. Mass transfer is a function of
the surface and bulk concentrations of ions, diffusion of the ions, and total overpotential. All of these are
strongly dependent upon process parameters such as electrolyte composition, pH, temperature, agitation,
and applied current density or potential. Fundamental mathematical expressions exist to describe these
physical phenomena that give fairly good control during electrolytic processing.
The second stage of electrocrystallization is the actual crystallization process that involves the formation
of adions on the cathode surface, adsorption and desorption reactions, surface diffusion of adatoms, as
well as nucleation and growth of crystals. Some of these are shown in Figure 22.3 in very simplified form.
To date, crystallization phenomena are less well understood than the deposition stage phenomena. They
are usually described by using empirical relationships.
When deposition begins, the surface conditions of the cathode are very important not only in terms
of cleanliness, adsorped layers, or oxide films, but also with respect to atomistic structure. All surfaces
contain numerous defects such as steps, ledges, kinks, vacancies, dislocations, and intersecting grain
boundaries. Processes such as adatom attachment and diffusion, as well as nucleation and growth are
strongly influenced by these defects. Chapter 9 provides more details on surface growth.
Crystalline metal electrodeposits exhibit several growth forms including layers, blocks, pyramids, ridges,
spiral growth forms, dendrites, powders, and whiskers.13 These morphologies have been studied extensively and various models have been proposed to correlate specific growth forms with process parameters
and substrate microstructure. For example, with increasing overvoltage the deposit structure changes
from compact to powdery.
The internal microstructure evolution of the deposit in terms of grain size and shape is strongly dependent upon crystal growth inhibition that can be controlled, for example, by adding certain bath additives that
reduce the surface mobility of adatoms or block certain sites for deposition. With increasing inhibition, the
deposit structure changes from field oriented (FI) to basis oriented and reproduction type (BR), to twin
transition types (TT), to field-oriented texture type (FT) and finally to unoriented dispersion type (UD):14
FI: Field-oriented crystal types such as whiskers, dendrites, or loose crystal powder.
BR: Basis reproduction type consisting of coherent deposits in which grain size and surface roughness
increase with deposit thickness.
TT: Twin transition type showing high twin densities, usually in materials with low stacking fault
energies.
FT: Field-oriented texture type showing coherent deposits with constant small grain size throughout
the deposit.
UD: Unoriented dispersed type with very small grain size.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-9
(b)

Growth Direction

(a)

Substrate

Poly Ni

250 m
Substrate

FIGURE 22.4 Cross-sectional structure of a typical Ni electrodeposit. (a) Schematic diagram and (b) thick deposit
showing transition from equiaxed fine-grained to columnar textured structure.

With increasing deposit thickness considerable structural changes can be observed as a result of the
competition between crystal nucleation and crystal growth. Under certain deposition conditions (e.g.,
BR growth conditions) the structure can change from initially fine grained, equiaxed structure to coarse
grained columnar structure with extensive grain shape anisotropy and crystallographic texture. Figure 22.4
shows a schematic diagram of this transition as well as a cross-section through a nickel electrodeposit with
such a change in the microstructure. It is generally accepted that this transition is a direct result of the
growth competition between differently oriented grains, such that grains with lower surface energy grow
faster than grains with high surface energy.14
Over the past decade, considerable progress has been made in the understanding of the early stages
of electrodeposition, microstructural evolution with increasing deposit thickness and the formation of
dendritic, densely branched, fractal, stringy, and needle-like morphologies.15,16 Many studies in these
areas apply powerful computer simulations (e.g., tight binding approximations, embedded atom models,
Monte Carlo simulations) or experimental methods (e.g., in situ techniques such as scanning tunneling
microscopy, small-angle neutron scattering or microwave reflectivity) to study these phenomena.

22.3 Surface Finishing


The broadest area of applications of electrolytic processes deals with various types of surface-finishing
methods in which a coating is deposited or formed on the surface of finished or semifinished products.
The reasons for applying coatings are to improve/modify surfaces in terms of appearance, corrosion resistance, wear resistance, oxidation resistance, heat and light reflectivity or absorption, electrical conductivity,
magnetic properties, or to build up undersized or worn parts. The various process types include electrodeposition, electroless deposition, metal displacement, oxidation (anodizing) processes, and chemical
treatments.

22.3.1 Electrodeposition
In electrodeposition the workpiece to be plated is made the cathode in an electroplating tank containing
the metal ions in aqueous solution. The workpiece is electrically connected via a power supply with the
anode that is usually made of the metal to be plated either in the form of a solid rod/sheet or, more
commonly, pieces of the metal contained in an inert titanium basket. During electrodeposition, the anode
dissolves slowly and replenishes the plating bath with metal ions as they are reduced on the cathode to
form the coating. In some applications, dimensionally stable anodes are used (e.g., platinum or platinized
titanium) that do not dissolve during the plating process.

2007 by Taylor & Francis Group, LLC

22-10

Materials Processing Handbook


TABLE 22.2

Watts Nickel and Sulfamate Nickel

Nickel sulfate, NiSO4 6H2 O [g/l]


Nickel sulfamate, Ni(SO3 NH2 )2 4H2 O [g/l]
Nickel chloride, NiCl2 6H2 O [g/l]
Boric acid, H3 BO3 [g/l]
Temperature ( C)
Cathode current density [mA/cm2 ]
pH
Anode

Watts Nickel

Sulfamate Nickel

220400

3060
3045
4065
30100
24.5
Nickel

300450
030
3045
3060
5300
3.55
Nickel

22.3.1.1 Pure Metals


Electrodeposition of pure metals from aqueous solutions is the most widely used surface-finishing method.
The basic process at the cathode follows the reduction reaction given in Equation 22.1, for example for
the case of Ni:
Ni2+ + 2e Ni.

(22.24)

A large number of electroplating baths have been developed for nickel plating14 which differ mainly in
terms of metal salt types and concentrations, pH and bath additives used for very specific purposes. Two
commonly used plating solutions for nickel are the Watts and the sulfamate baths, Table 22.2. Nickel
sulfate/nickel sulfamate are the main sources of the nickel ions in the solutions. Nickel chloride also
supplies some nickel ions, but its primary functions are to promote the dissolution of the anode nickel
and to increase the electrical conductivity of the plating solution. Boric acid acts as a buffer in the solution.
There are a number of other bath additives (e.g., saccharin, coumarin) that may be added for specific
purposes such as stress relief or grain size refinement in the deposit. Many of these bath additives are
proprietary chemicals supplied by many electroplating supply companies.
Potentially all metals with standard potentials more positive than the hydrogen reduction reaction
(Table 22.1) could be electrodeposited from aqueous solutions. However, there are other factors that limit
the number of electropositive metals that are routinely electrodeposited. These include the quality of the
resulting deposits in terms of growth form, physical integrity and properties, the plating bath stability,
ease and economics of operation, as well as environmental and health risk concerns when dealing with
toxic chemicals.
On the other hand, there is also a relatively large number of metals with more negative potential
than hydrogen that can still be plated, including nickel, cadmium, zinc, and chromium. The reason for
this is the hydrogen overvoltage relative to the metal deposition overvoltage. The hydrogen overvoltage
is a direct measure of the required activation energy for hydrogen reduction that varies greatly from
metal to metal. The following example will illustrate that the metal zinc can be electroplated from an
aqueous solution onto steel even though this would not be expected from its standard potential relative to
hydrogen. By comparing only the two standard potentials in Table 22.1 for zinc (0.76 V) and hydrogen
(0 V) one would expect hydrogen evolution to occur instead. For zinc deposition, for example, from an
acidic (pH = 1) 1 molar zinc sulfate solution, the reversible hydrogen potential as per Nernst equation is
E = 0.059 pH 0.06V. On the other hand, the activation overvoltage for zinc deposition is very small
and zinc can be plated at a potential very close to 0.76 V (Table 22.1). The hydrogen overvoltage on
steel at pH = 1 increases from 0.56 V at a current density of 10 mA/cm2 to 0.82 V at 100 mA/cm2 .14
Therefore, at a current density of 100 mA/cm2 , the onset of hydrogen evolution is at a more negative
potential than the potential required for zinc deposition, thus allowing for zinc deposition onto steel.
Nevertheless, with increasing negative potential it becomes more and more difficult to deposit metals
from aqueous solutions. Chromium and manganese can still be plated. However, there are no practical
aqueous solutions for electrolytic processing of elements such as Zr, Ti, Al, or other elements with even
more negative standard potentials. For these metals electrolysis requires the use of molten salt or ionic

2007 by Taylor & Francis Group, LLC

Electrolytic Processes
TABLE 22.3

22-11

Important Pure Metal Electrodeposits and Their Applications

Metal
Cadmium
Chromium
Chromium
Chromium
Copper
Cobalt
Gold
Iron
Lead
Nickel
Nickel
Palladium
Rhodium
Silver
Tin
Zinc

Application

Typical thickness(m)

Electrochemical protection of steel and cast iron


Hard chromium for enhanced wear properties on hydraulic shafts, piston
rings, aircraft landing gears; rebuilding of mismachined or worn parts
Decorative coatings over undercoatings such as nickel and coppernickel
Black chromium for solar selective coatings
Printed wiring boards; semiconductor interconnect technology; printing
rolls for paper and textiles; build up of worn parts
Electroplated steel for batteries; magnetic coatings
Electronic components; wire bonding; high reflectivity surfaces; decorative
finish; jewelry
Build up of worn parts; hard coatings on pistons
Corrosion protection of steel
Corrosion resistant coatings; decorative coatings; magnetic coatings
Black coatings for decorative purposes; nonreflective surfaces
Interconnect products; electronic packaging; decorative coatings
Electronic components; decorative coatings
Decorative coatings; electronic coatings; medical-instrument coatings;
jewelry
Corrosion protective coatings; solder applications
Corrosion protection of iron and steel

<25
100250
<5
<5
10200
120
0.510
100500
10100
10200
11.5
0.55
10200
250
5100
520

liquid electrolytes (see Section 22.5.2). Taking all of these factors into consideration, only the elements
given in bold in Table 22.1 are routinely processed using electrodeposition from aqueous solutions.
Table 22.3 lists several pure metal electrodeposits that are of considerable technological importance.
The many different electroplating baths that have been developed for the various metals can be found
in several books dealing with electrodeposited metals.1424 Depending on the application and type of
coating, typical thicknesses range from <1 m to hundreds of micrometers.
The physical, chemical and mechanical properties of electrodeposited metals depend strongly on grain
size and grain shape, impurities codeposited with the metal, crystallographic texture, and internal stress.
To a large extent, these microstructural characteristics can be controlled by the plating parameters such
as bath composition, pH, and temperature as well as agitation of the electrolyte and current density used
during deposition. Extensive tables showing electrodeposit properties are presented in several books on
plating.14,21,23
22.3.1.2 Alloy Deposits
Commercial electrodeposition is not limited to pure metals. A large number of alloy deposits have been
developed that contain two or more elements. Examples of important binary electrodeposits are given
in Table 22.4. In alloy electrodeposition the bath formulations are more complex than for pure metal
deposition. These baths contain several ionic species, and in many cases, complexing agents. For example,
nickelphosphorus alloy deposits can be electroplated from a solution containing nickel sulfate (NiSO4 ),
nickel carbonate (NiCO3 ), phosphoric acid (H3 PO4 ), and phosphorous acid (H3 PO3 ).19 There are several
direct and indirect cathodic reactions during electrodeposition of NiP alloys:
direct reactions:

Ni2+ + 2e Ni

(22.25)

2H+ + 2e H2

(22.26)

indirect reactions: 12H + 12e 12H

2007 by Taylor & Francis Group, LLC

(22.27)

2H3 PO3 + 12H 2PH3 + 6H2 O

(22.28)

2PH3 + 3Ni2+ 3Ni + 2P + 6H+ .

(22.29)

22-12

Materials Processing Handbook


TABLE 22.4

Examples of Nickel, Cobalt, and Zinc Alloy Electrodeposits

Material
NickelChromium
NickelCobalt
NickelCopper
NickelIndium
NickelIron
NickelManganese
NickelPhosphorus
NickelTungsten
CobaltNickel
CobaltMolybdenum
CobaltPhosphorus
CobaltTungsten
ZincCobalt
ZincNickel
ZincTin

Application
Corrosion resistant coatings
Decorative, mirror like deposits
Compositionally modulated alloys
Low friction and microelectronics coatings
Soft magnetic coatings; decorative coatings
High ductility nickel coatings
Abrasion and corrosion resistant coatings
High temperature oxidation/resistant coatings
Magnetic coatings for storage devices
High temperature oxidation resistant coatings
Wear-resistant coatings
High temperature strength coatings
Corrosion resistant coatings
Corrosion resistant coatings
Coatings on fasteners for aluminum panels

Reaction 22.29 shows that phosphorus is reduced together with nickel, which explains the NiP alloy
deposit formation. In this case, the phosphorus in the alloy deposit comes from the phosphorous acid in
the plating bath. The phosphoric acid does not provide P for the alloy formation. Its primary role is to
adjust the pH of the solution.
The composition of the electrodeposit may be quite different from the ratio of the relative concentrations
of bath constituents. For example, for electrodeposited ZnNi electrodeposits, the ratio of the less noble
metal (Zn) to the more noble metal (Ni) in the deposit is larger than in the bath. This effect is known as
anomalous codeposition.19
The electrochemistry of alloy plating is much more complex than for simple metal deposition. The
complexity is due to the presence of more than one ionic species in the solution, their mutual influence
when it comes to double layer formation, electrochemical potentials, electrocrystallization, and deposit
structure formation. Alloy deposition is beyond the scope of this chapter and the reader is referred to the
references that treat alloy formation by describing simultaneous codeposition of several species for very
specific alloy systems.13,19,25
22.3.1.3 Composite Deposits
Metalmatrix composites can be produced by adding second phase particles to the plating bath that are
then codeposited with the metal matrix. An example is shown in Figure 22.5 which represents a nickel
silicon carbide composite consisting of a nickel matrix with about 5 vol.% SiC particles (average particle
size: 0.3 m). The coating was produced from a Watts type plating bath containing about 50 g/l of SiC
powder. The purpose of the SiC particles is to increase the wear and scratch resistance of the nickel coating.
Other hard particles that find application in wear-resistant coatings include alumina, titania, thoria, and
diamond.21 On the other hand, particles such as Teflon or molybdenum disulfide can be incorporated into
the metal to make low friction surfaces. Several models have been proposed26 that describe the particle
incorporation in the deposits in terms of the various steps involved including initial weak adsorption of
particles on the cathode surface, field-assisted adsorption, ionic double layer formation on the particles,
and encapsulation in the metals matrix.
22.3.1.4 Semiconductors
Because of the very large negative reduction potential silicon cannot be electrodeposited from aqueous
solutions. However, several compound semiconductors can be deposited from suitable aqueous
electrolytes.14,27 Examples of the main bath constituents for several IIVI and IIIV compound
semiconductors are given in Table 22.5.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-13

(b)

(a)

10 m

100 nm

FIGURE 22.5 NiSiC composite electrodeposit: (a) scanning electron micrograph and (b) bright-field transmission
electron micrograph.

TABLE 22.5

Main Bath Constituents for Semiconductor Deposition

Semiconductor
InAs
InSb
PbS
CdTe
CdSe
ZnSe
GaAs

Bath Constituents
Indium chloride (InCl3 ); Arsenic chloride (AsCl3 )
Indium chloride (InCl3 ); Antimony chloride (SbCl3 )
Lead nitride (Pb(NO3 )2 ); Sodium sulfite (Na2 SO3 )
Cadmium sulfate (CdSO4 ); Tellurium oxide (TeO2 )
Cadmium chloride (CdCl2 ); Selenium oxide (SeO2 )
Zinc sulfate (ZnSO4 ); Selenic acid (H2 SeO3 )
Arsenic oxide (As2 O3 ); Gallium oxide (Ga2 O3 )

22.3.2 Electroless Deposition


Electroless deposition14,28,29 was originally known as electrode-less deposition, describing the fact that
no external power supply is connected to positive and negative electrodes. Also known in the literature
as autocatalytic reduction or chemical reduction, the process still requires anodic and cathodic reactions.
However, these occur on the same material, and the electrons required for these are supplied by a reducing
agent in the plating bath according to the following general reactions:14
Mz+ + Red M + Ox

(22.30)

where Red is the electron-providing reducing agent and Ox is the oxidation product of the reducing agent.
The partial cathodic and anodic reactions are as follows:
cathodic:
anodic:

Mz+ + ze M

(22.31)

Red Ox + n e

(22.32)

where n is the number of electrons produced during the oxidation of the reducing agent. The two half
reactions occur on the same surface which could be either a metal substrate or a nonconductive substrate
(e.g., glass, polymer) made active by first depositing catalytic nuclei such as tin or palladium. In other
words, electroless deposition is an electrolytic method that can be used to deposit a metal coating onto
nonconductors such as polymers or glass.

2007 by Taylor & Francis Group, LLC

22-14

Materials Processing Handbook


TABLE 22.6

Reducing Agents for Electroless Metal Deposition

Metal
Nickel
Nickel
Nickel
Nickel
Copper
Gold
Gold
Palladium
Palladium

Reducing Agents
Sodium hypophosphite (NaH2 PO2 H2 O)
Sodium Borohydride (NaBH4 )
Dimethylamine borane, DMAB ((CH3 )2 NHBH3 )
Hydrazine (N2 H4 H2 O)
Formaldehyde (HCHO)
Potassium borohydride (KBH4 )
Dimethylamine borone, DMAB ((CH3 )2 NHBH3 )
Hydrazine (N2 H4 H2 O)
Hydroxylamine hydrochloride (NH2 OH HCl)

Some of the commonly used electroless metals that are commercially of considerable importance
are nickel, copper, gold, and palladium. Table 22.6 shows several of the reducing agents used in their
deposition.
The electrochemical reactions during electroless deposition are usually more complex than shown in
Equation 22.30, and involve several intermediate steps. Without presenting all intermediate reactions, one
important step during the deposition of electroless nickel using sodium hypophosphite as reducing agent
is as follows:
H2 PO2 + H P + H2 O + OH .

(22.33)

This reaction shows that phosphorus is codeposited with Ni, similar to the case of electrodeposition
of NiP alloys (Section 22.3.1.2). This explains why electroless nickel can contain substantial amounts
of phosphorus, up to 20 at.%, depending on the bath composition and temperature. A direct consequence
of this is seen in the microstructural evolution of the deposits. For small phosphorus contents, the deposits
are usually crystalline. With increasing phosphorus, the grain size continuously decreases, and for high
phosphorus the deposits become amorphous.14,28,29

22.3.3 Displacement Deposition


Displacement deposition, also known as immersion plating or galvanic plating is similar to electroless
deposition in that it does not require an external power supply. However, in contrast to electroless
plating, no reducing agent is required for this process. In displacement deposition, the electrons for metal
reduction come from the substrate itself. For example, a strip of zinc can be plated with copper from
a copper sulfate solution at room temperature according to the reactions given in Equation 22.8 and
Equation 22.9. In other words, the more electropositive metal copper replaces the more electronegative
metal zinc on the zinc substrate. Of course, the thickness of galvanic Cu deposits is very limited, typically
to a few micrometers, as this reaction needs direct contact between the zinc substrate and the copper
sulfate solution. Once a closed layer of copper has been deposited the reaction will cease. Immersion
plating solutions are commercially available17 for a number of metals (e.g., Cu, Cd, Au, Pd, Pb, Pt, Rh,
Ru, Ag) as well as several alloys (e.g., brass, bronze) in particular to plate substrate materials such as steel,
aluminum, zinc, and copper. Of course, the choice of deposit/substrate combination is largely dictated by
the standard potentials given in Table 22.1.

22.3.4 Chemical Conversion


From a processing point of view, chemical conversion techniques are borderline processes falling between
chemical and electrochemical synthesis. In many cases, no clear electrified interface can be defined at which
charge transfer occurs. Nevertheless, the most important chemical conversion processes are included
here, because of (1) their current importance in surface finishing and (2) their as of yet unexplored

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-15

future application potential in surface nanotechnology. To date, microstructural information on chemical


conversion layers is very limited.
22.3.4.1 Antiquing of Brass, Copper, and Bronze
This process is a popular method in the manufacture of decorative hardware fixtures for many indoor
and outdoor applications.17 The surface color of copper, brass, and bronze articles is modified to various
shades of brown to black or green, including verde green patina (verdigris) similar to natural patina on
aged copper roofs, by an immersion technique. Typical solutions are based on polysulfide salts, selenium
salts, or mild acids, and the color of the conversion layer comes from the compounds that form during
the chemical reaction with the surface. Articles made of steel or zinc die casts are first electroplated with
a thin layer of copper, brass, or bronze and then antiqued using the chemical conversion treatment.
22.3.4.2 Phosphating
Phosphate coatings are chemical conversion coatings applied to the surfaces of steels and aluminum alloys
for articles such as nuts, bolts, and screws, as well as sheet steel for automotive and consumer product
applications.18 These coatings provide good corrosion and wear protection and, in many cases, are applied
as a base for subsequent paint application. Using dilute solutions of phosphoric acid containing zinc, iron,
or manganese salts, coating layers (typical thickness: 1 to 50 m) of strongly adherent phosphate crystals
are formed either during immersion or spray operations.
22.3.4.3 Chromate Conversion Coatings
Chromate conversion coatings are applied on various metals by chemical or electrochemical treatment
with hexavalent chromium (chromic acid) or chromium salts (e.g., potassium chromate). During this
treatment the surface of the metal is converted to a protective film that contains complex chromium
compounds.17,18 The thickness of the film is typically on the order of 1 m and can vary in color from
clear-bright, blue-bright and yellow to brown, olive, and black, depending on the substrate material. The
main purpose of the chromate coating is to provide corrosion protection, hardness, and wear resistance
and to serve as a base layer for subsequent paint application. Chromate coatings are routinely applied on
aluminum, magnesium, and zinc alloys. More recently, trivalent chromium conversion layers have been
developed, in view of the environmental and health risk concerns associated with the use of carcinogenic
hexavalent chromium compounds.17
22.3.4.4 Blackening of Ferrous Metals
This conversion process uses various bath chemistries to produce attractive black finishes on ferrous metals
that also provide moderate corrosion resistance.17 In the hot alkaline nitrate black oxidizing solution, the
black color comes from the black iron oxide (magnetite, Fe3 O4 ) which is formed in boiling alkaline nitrate
solution. A black zinc phosphating process has been developed in which steel, cast iron, or malleable iron
is blackened before coating with a zincphosphate top layer.17

22.3.5 Anodizing
Anodizing of aluminum is a widely used surface-finishing process to improve the surface characteristics
in terms of appearance (color), durability, and corrosion resistance of finished aluminum parts.30,31
In this process, the natural protective aluminum oxide layer that forms in air to a thickness of only a
few nanometers is artificially built up to considerable thickness (up to several tens of even hundreds of
micrometers) by making the aluminum part the anode in a suitable electrolyte, connected via the power
supply to a dimensionally stable (inert) cathode such as graphite or stainless steel. While anodic treatment
in boric acid, borate, or nitrate baths produces only slightly thicker oxide films than air-exposure, sulfuric
acid and chromic acid baths produce initially a thin oxide barrier, followed by the growth of a thicker,

2007 by Taylor & Francis Group, LLC

22-16

Materials Processing Handbook


(a)
Top
View

Alumina

100 nm
(b)

Side
View
Alumina

Aluminum
Pigment
(c)

Aluminum

FIGURE 22.6 Nanoporous aluminum oxide produced by anodizing: (a) top view showing hexagonal pore structure,
(b) cross-sectional view and (c) coloring by pigment addition.

porous layer on top of the initial layer. The reactions are as follows:
cathodic reaction: 6H+ + 6e 3H2 (g)

(22.34)

anodic reactions: 2Al + 3O2 Al2 O3 + 6e

(22.35)

2Al3+ + 3H2 O Al2 O3 + 6H+ .

(22.36)

It should be noted that there are two anodic reactions that produce Al2 O3 . The first Reaction 22.35 is
at the Al/Al2 O3 interface, while the second Reaction 20.36 takes place at the Al2 O3 /electrolyte interface,
which requires the diffusion of Al3+ ions through the Al2 O3 layer already formed on the base material.
Under suitable conditions this porous layer can be produced as irregular or regular arrangements of
hexagonal columns with pores extending from the free surface down to the initial barrier layer. Figure 22.6a
and Figure 22.6b show schematic diagrams of the top and side views of a regular oxide structure without
giving details of the original barrier layer. For requirements such as scratch and corrosion resistance, the
porous aluminum oxide is usually sealed, for example, by boiling in water, which hydrates and expands the
oxide to close the pores. In color anodizing, the pores are impregnated prior to sealing with various dyeing
compounds, mineral pigments, or microparticles formed from the microconstituents of the aluminum
alloy during the anodizing process, to produce a wide variety of colors, in particular for architectural
applications of aluminum (Figure 22.6c).

22.4 Direct Manufacturing Using Electrolytic Processing


Electrodeposition is not restricted to surface-finishing applications, but is also widely used to produce
finished or semifinished products through direct manufacturing. Generally known as electroforming,3235

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-17
VE

VE

Extraction
Collar
Stainless Steel
Mandrel

Nickel Deposit
Forming
During Electroforming

Nickel
Electroform
After Extraction

FIGURE 22.7 Net shape manufacturing of an electroformed product (courtesy of Nickel Development Institute,
Toronto, Canada).

these processes are used to produce various product forms by electrodeposition onto suitably shaped
cathodes, designed as temporary reusable mandrels during the deposition process (Figure 22.7). Following
electrodeposition of the product, the mandrel is separated from the electroform by simple mechanical
means or heating/cooling cycles. To allow for easy separation, mandrel materials such as titanium or
stainless steel are commonly used that have a thin oxide layer on the surface to prevent strong adhesion
of the electroformed product. Many metals have been used in electroforming, including nickel, copper,
iron, gold, platinum, cobaltnickel, and cobalttungsten alloys.

22.4.1 Large-Scale Manufacture


Direct net shape manufacturing by electroforming is applied in areas where other processing techniques
are limited in their capabilities or simply too expensive. The most widely known example of electroformed
nickel are the large printing rolls used as mandrels in textile printing in which the pattern is embedded
in the surface of the electroformed nickel. Other examples in which the same principle was used are the
stampers for records and CDs and the embossing tools for holographic images.
Sheet and foil products with thicknesses <100 m can be produced in a continuous plating process onto
a rotating drum that is partially or fully immersed as the cathode in a plating bath. The foil is mechanically
stripped from the drum once the desired thickness is reached. The foil thickness is controlled by the current
density and the residence time in the plating bath, given by the drum diameter and the rotational speed.
One of the largest commercial applications of this type of foil product is the copper foil used as the metallic
conductor in the manufacture of printed circuit board copper/epoxy laminates. Other foil applications
include stock material for gaskets, pressure control membranes, hydrogen purification membranes, solar
energy absorbers, and microfoils. If the rotating drum surface is modified to contain a surface pattern
produced by using nonconductive paints, adhesives, or lithographic patterning, continuous metal mesh
can be produced for application such as sieves, electric razor foils, printing screens, or centrifuge screens.
A large number of complex free-forms have been produced by electroforming on reusable complexshaped cathodes. Examples include erosion shields for helicopter blades, thrust chambers for rocket
engines, nozzles and precision reflectors and mirrors. Even very intricate art objects have been made by
this process, known as galvanoplasty.32

22.4.2 Small-Scale Manufacture


In recent years, microsystems technology has advanced rapidly and microelectromechanical systems
(MEMS) find increasing use in many automotive, aerospace, medical, defense, and biotechnology

2007 by Taylor & Francis Group, LLC

22-18

Materials Processing Handbook

(a)

Top
View

(b)

X-Ray Irradiation

Side
View
Mask

Original
Mask

Metal Substrate

500 m

Cavity (d)
Etched
Mold

(c)

Metal Substrate
(f)

PMMA
Mold
Metal
Deposit

Metal Substrate
Final
Product

Planarized
Deposit

(e)

MEMS gear

Metal Substrate
FIGURE 22.8 Various steps in the production of a MEMS gear by electrodeposition: (a) mask showing the shape of the
gear, (b) irradiation of PMMA through mask, (c) etched mold with cavity, (d) mold cavity filled with electrodeposited
metal, (e) planarization step and (f) final product removed from mold.

applications.36,37 One specific area of MEMS manufacturing that deals with electrodeposition of small
(<1 mm) metallic components is the process known as LIGA (German acronym for Lithographie
[lithography], Galvanoformung [electroforming], and Abformung [molding]). (More details about such
additive processes may be found in Chapter 26.) Figure 22.8 shows a series of schematic diagrams depicting
the production of a MEMS gear via the LIGA process using polymethylmethacrylate (PMMA) as the mold
material. Basically, LIGA involves two main subprocesses: preparation of a mold into which the metal
is to be deposited and the actual deposition and removal of the final product. Typical metals that find
applications in LIGA products are nickel, copper, gold, and nickeliron alloys.

22.5 Primary Metal Production


While surface finishing is likely the area with the broadest range of electrolytic processes in terms of
demonstrated technological flexibility, primary metal production is certainly the field where the largest
tonnages of materials are handled. Electrolytic processes are of tremendous importance in the extraction
and refining of metals for which pyrometallurgical processes are unfavorable in terms of free energy
of reactions and economic considerations. Electrolytic reduction and refining is applied in numerous
processing steps during the manufacture of many metals including nickel, copper, cobalt, lead, iron,
aluminum, and magnesium. This chapter is limited to two of the most important processing routes:
electrowinning and electrorefining, and molten salt electrolysis.

22.5.1 Electrowinning and Electrorening


Electrowinning and electrorefining are two of the most important electrometallurgy processes in which
huge quantities of metal ions are processed by electrolytic means, usually in very large electrolytic cells.20 In
electrowinning, metals are recovered from solutions derived from primary metal leaching and purification
operations. The main purpose of electrorefining, on the other hand, is to electrolytically dissolve a metal
in a solution and then redeposit it with higher purity than the starting material. The main differences
between the two processes are the anode materials and the electrochemical reactions. In electrowinning the
anode is the material to be refined, while electrowinning operations use dimensionally stable (e.g., inert

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-19

lead) anodes. For this reason the anodic and cathodic reactions for both processes are quite different.
For example, if the electrowinning electrolyte, coming from a sulfuric acid leaching of a metal oxide, is
MSO4(aq) then the reactions are as follows:
cathodic reaction: MSO4(aq) + 2e M + SO2
4(aq)

(22.37)

anodic reaction: H2 O 2H+ + 1/2O2(g) + 2e

(22.38)

In other words, the main reaction at the anode is the decomposition of water. On the other hand, during
electrorefining the reactions are simply as follows:
cathodic reaction: Mz+ + ze M
anodic reaction: M M

z+

+ ze

(22.39)
(22.40)

In both cases, the cathode materials are often starter sheets of the same material that is to be deposited.
Assuming that there are no other cathodic reactions, the weight, W , of the deposit can be calculated using
Faradays law:
W =

I t A
z F

(22.41)

where I is the current through the cell in amps, t the time in seconds and A the atomic weight of the
deposited material.

22.5.2 Molten Salt Electrolysis


As pointed out in Section 22.3.1.1, reactive metals more electronegative than manganese cannot be
processed electrolytically from aqueous solution. For this reason electrolytic methods for such metals
require either water-free organic electrolytes38 (not discussed in this chapter) or molten salt electrolytes.
The latter are sometimes also referred to as fused salt or high temperature inorganic melts. Examples
of materials that have been produced by this method include many refractories such as molybdenum,
niobium, vanadium and tungsten, titanium, zirconium, magnesium, and aluminum, and several rare
earth elements. While many of the fundamental concepts developed for aqueous electrolytic processes
can be applied to molten salt electrolysis, considerable modifications and new considerations are required
because the molten salt solvent is quite different from water as the solvent. In fact, the electrochemistry of
fused salts is an entire research field on its own.39,40
Probably the most widely known molten salt process is the HallHroult process for aluminum production introduced in 1866.41 In this process, molten cryolite (Na2 AlF6 ) is used as the solvent for alumina
(Al2 O3 ) which is usually chemically purified from bauxite in the Bayer process. Cryolite melts at about
940 C, and the HallHroult process operates at temperatures between 960 and 1000 C. The anode
material used is carbon and a typical electrolytic cell operates at 6 V. The cell is lined with carbon that acts
as the cathode onto which Al is deposited. The overall cell reaction is:
2/3Al2 O3 + C 4/3Al + CO2 .

(22.42)

During the process, the carbon from the anode is continuously consumed to produce CO2 . A typical
cell produces about 300 kg of aluminum daily at an energy consumption of about 15 to 26 kWh/kg of
aluminum. Aluminum with a purity of 99% or higher can be produced by this process. It is important to
note that the fused salt electrolyte does not chemically react with the Al deposit.

2007 by Taylor & Francis Group, LLC

22-20

Materials Processing Handbook

22.6 Electrolytic Processing of Nanostructures


As a result of their outstanding physical, chemical, and mechanical properties, nanostructured materials
have received considerable attention over the past 10 years. Since their first introduction in the early 1980s,
several hundred synthesis methods for their production have been described in the literature.4252 Many
of these methods are simple extensions of well-established production routes while others were specifically
designed for various types of highly specialized nanomaterials. Many of these techniques are based on
electrolytic processes.
In the following sections, examples will be presented for the synthesis of a variety of nanostructured
materials by electrolytic processing including monolithic metals and alloys, structurally graded nanometals, nanocomposites, compositionally modulated alloy (CMA) nanostructures, and nanocrystalline
oxide powders.

22.6.1 Monolithic Bulk Metal Nanostructures


Synthesis methods for the production of monolithic bulk electrodeposits with grain sizes <100 nm have
been systematically developed since the early 1980s.5355 Initially, relatively simple electrolyte formulations
and conventional direct current electrodeposition were employed to produce nanostructured deposits
(e.g., NiP56,57 ), with particular attention to structures with relatively narrow grain size distributions
and average grain sizes in the 5 to 50 nm range. Beginning in the late 1980s emphasis shifted toward
the development of more complex alloy systems (e.g., CoFe, NiFe, NiMo, CoMo, ZnNi, NiFeCr,
etc.) and the use of pulsed current plating to promote crystal nucleation at higher achievable current
densities.58,59
In the electrodeposition process there are two competing factors that govern the grain size of the final
deposit. Figure 22.3 showed some of the steps involved in the electrocrystallization process. When a
metallic ion is reduced it can either nucleate a new grain or join an existing grain and contribute to its
growth. Nanostructure formation can be achieved by using operating parameters (e.g., bath composition,
pH, temperature, current density, current on and off times during pulsed plating), which are conducive to
massive nucleation throughout the entire electrodeposition process. It is important to avoid the transition
from initially fine grained to coarse columnar structure often observed in conventional electrodeposition
(Figure 22.4). In other words, nanostructured electrodeposits can be considered an extreme case of UD
type deposits (Section 22.2.6). Their structure consists of nanosized grains throughout the entire thickness
(Figure 22.9 and Figure 22.10).

(a)

(b)

Growth Direction

Nano Ni

Substrate

Substrate

100 nm

FIGURE 22.9 Cross-sectional structure of nanocrystalline nickel electrodeposit on bronze substrate: (a) schematic
diagram and (b) transmission electron micrograph.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-21

200 nm

FIGURE 22.10 Darkfield electron micrograph parallel to the substrate of bulk nanocrystalline nickel (average grain
size: 12 nm) in planar cross-section.

Electrodeposition processes have been developed for the synthesis of nanocrystalline deposits in many
different shapes and forms.54,55 These range from thin and thick corrosion and wear-resistant coatings to
sheet and plate products for structural, magnetic, and electronic applications. Furthermore, the technology can be incorporated with little extra costs in conventional electroplating plants in processes such as
rack plating, reel-to-reel plating, continuous strip plating, barrel plating, and brush plating.
One of the earliest industrial applications of this type of nanocrystalline deposits was the in situ
electrosleeve nuclear reactor steam generator tubing repair technology.60,61 This technology was the first
large-scale structural application of nanomaterials in the world, specifically developed for Canadian and
U.S. nuclear reactors since the early 1990s. In this process, steam generator tubes (e.g., Alloy 600 or 400)
whose structural integrity was compromised by localized degradation phenomena (e.g., intergranular
corrosion, pitting) were repaired by coating the inside of the tubes with a thick (1 mm) nanocrystalline Ni
microalloy to restore a complete pressure boundary. The grain size of the materials was adjusted to be in
the 50 to 100 nm range to give the required combination of strength, ductility, corrosion resistance, and
thermal stability for this application. Figure 22.11 shows some of the tremendous property improvements
that can be achieved when reducing the grain size in nickel electrodeposits from 10 m to 10 nm.59,62

22.6.2 Structurally Graded Nanomaterials


While monolithic nanometals with relatively narrow grain size distributions have excellent properties
in terms of hardness, tensile strength, and wear resistance, their tensile ductility is usually compromised regardless of the processing route.59,63 In the past few years, it has been recognized that considerable
ductility can be restored in these materials through broader or bimodal grain size distributions.63 Recently,
electroplating conditions have been developed for the synthesis of a variety of structurally graded nanometals (Figure 22.12), including bimodal distributions, alternate layers of different grain sizes, and grain
size gradient deposits.59,64 In such structures, a good compromise between strength and ductility can
be achieved, whereby the smaller grains provide the strengthening in the metal while the larger grains
allow for sufficient dislocation activity to result in reasonable ductility values. It should be noted that the
deformation mechanisms in these materials are currently not completely understood.

22.6.3 Nanocomposites
Over the past several years, the concept of composite coatings (Section 22.3.1.3) has been extended to nanocrystalline materials.54,55,59,65 Figure 22.13 shows several examples of submicrocrystalline/nanocrystalline
second phase particles or fibers embedded in a nanocrystalline matrix. Typical examples of second phase

2007 by Taylor & Francis Group, LLC

22-22

Materials Processing Handbook

1000
(a)

(b)
800

Yield Strength (MPa)

Vickers Hardness (GPa)

5
4
3
2

600

400

200
1
0
100

0
100

101

102
103
Grain Size (nm)

104

105

102
103
Grain Size (nm)

104

105

1.0

150
(c)

0.9

125

(d)

0.8
Coefficient of Friction

Wear rate (mm3/mm) 105

101

100
75
50

0.7
0.6
0.5
0.4
0.3
0.2

25

0.1
0.0

0
100

101

102
103
Grain Size (nm)

104

105

100

101

102
103
Grain Size (nm)

104

105

FIGURE 22.11 Effect of grain size in electrodeposited Ni on (a) Vickers hardness, (b) yield strength, (c) sliding wear
rate and (d) coefficient of friction.

materials are Al2 O3 , SiC, or B4 C for improved hardness and wear resistance, or Teflon and MoS2 for
reduced coefficient of friction. Composite coatings can also be produced in situ by first codepositing an
alloy as a supersaturated solid solution, followed by heat treatment to precipitate nanosized second phase
particles in a nanocrystalline matrix.62,66

22.6.4 CMA Nanostructures


The idea of producing CMAs by electrodeposition (Figure 22.14) is not new. Some of the earliest work
was apparently already done by A. Brenner in his 1939 Ph.D. thesis (as cited in Reference 67). Because
of the unusual microstructure (i.e., individual layer thickness <10 nm), unexpected mechanical (e.g.,
strength, elastic constants), tribological and electromagnetic properties (e.g., giant magnetoresistance)
have been reported for such materials, which has led to considerable research efforts since the early 1980s.
The two main approaches to synthesize such materials are rotating substrate plating or potential-stepping
deposition. In the rotating substrate method, the rotating cathode is situated between two physically
separated baths and the thickness of each layer is determined by the plating rate in each bath and the
rotation speed of the cathode. In the potential-stepping method, on the other hand, the cathode is placed

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-23

(a)

(b)

(c)

FIGURE 22.12 Schematic diagrams showing various types of structurally graded nanomaterials: (a) bimodal grain
size distribution, (b) alternate layers with different grain sizes and (c) grain size gradient structure. (Courtesy of
Integran Technologies, Toronto, Canada.)
(a)

(b)

(c)

(d)

FIGURE 22.13 Examples of nanocomposite electrodeposits showing various configurations with different shapes
and sizes of the reinforcing phase in a matrix of electrodeposited nanometal. (Courtesy of Integran Technologies,
Toronto, Canada.)

into an electrolyte containing both species to be plated. By stepping the potential between predetermined
values, the relative deposition rate of one species over that of the other for certain time intervals is
controlled. Using these approaches, numerous compositionally modulated nanostructures have been
produced, including AgPd, CuNi, CuPb, and NiP59,67 with modulation wavelength down to a few
nanometers.
Even nanomodulated ceramic superlattices have been produced by the potential-stepping method. For
example, layered structures consisting of Tla Pbb Oc semiconducting oxides, with varying values for a, b,
c in the alternating layers, were produced with modulation wavelengths between 6 and 13 nm.68 These
materials are of considerable interest for optical and electrical applications.

2007 by Taylor & Francis Group, LLC

22-24

Materials Processing Handbook


Multilayers
Ni
Cu
Ni
Cu
Ni
Cu

Substrate

FIGURE 22.14 Schematic diagram showing a compositionally modulated material with alternating nanometer thick
layers of Ni and Cu.

22.6.5 Nanocrystalline Particles


A number of electrolytic methods have been developed for the synthesis of various nanoparticles either
on substrates or in colloidal solutions. For example, using the galvanic displacement approach (Section
22.3.3), very intricate structures of silver nanoparticles were recently grown on (111) and (100) oriented
single crystals of germanium according to the following displacement reaction:69
Ge Ge4+ + 4e

(22.43)

4Ag+ + 4e 4Ag.

(22.44)

The hexagonal nanoparticles deposited in the form of vertical stacks on the substrate in what was
described as nanoinukshuks, an analogy to the inukshuk structures made by the Inuit people in the Arctic
using flat slabs of rock.
Another example of nanoparticle production using a combination of electrolytic and chemical reactions is the synthesis of -CuI semiconductor nanocrystallites.70 The three-step process involved (1) the
electrodeposition of copper nanocrystals on a basal oriented single crystal of graphite from a Cu2+ containing solution, followed by (2) electrochemical oxidation of Cu to Cu2 O and (3) displacement of oxygen
by iodine in an aqueous potassium iodine solution to obtain semiconducting -CuI nanocrystallites with
particle sizes up to about 20 nm.

22.6.6 Nanocrystalline Oxide Powders


Known as electrochemical deposition under oxidizing conditions (EDOC), one specific technique to
produce nanostructured oxide powders is based on metal dissolution and deposition.59,71,72 As shown in
Figure 22.15, this method consists of three basic steps. At the sacrificial anode, metal is dissolved in an
organic electrolyte (e.g., 2-propanol) containing a salt to increase the electrical conductivity of the bath.
The second step is the cathodic reduction of metal ions to form metal clusters. These clusters are coated
with a stabilizer (e.g., quarternary ammonia salts, betains, or ethoxylated fatty alcohols, not shown in
Figure 22.15). In the final step, the colloidal metal particles are oxidized by introducing oxygen or air into
the bath with simultaneous strong agitation or sonication. Subsequently, the oxide particles are filtered,
washed with ethanol and dried. This method can produce simple oxides (e.g., ZnO, MgO, CuO, Fe2 O3 ,
SnO2 ) or mixed oxides (e.g., CoFe3 O4 , In2 O3 /SnO2 ) with high purity and particle size distributions much
narrower than what can be achieved with other nano-oxide powder synthesis methods such as inert gas
condensation, sol-gel, or flame pyrolysis. EDOC is a relatively inexpensive synthesis method and easy to
scale up for large-scale production.

22.6.7 Template Nanotechnology by Anodizing


Over the past decade, the synthesis of highly ordered structures based on using ordered arrays of nanoporous honeycomb structures of anodic aluminum oxide (Section 22.3.5) has evolved rapidly as part of the

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-25

Power
Supply

Oxygen

MeO
Me x+

Me x+

e
e
Cathode

Sacrificial Anode

Oxygen

Metal Cluster

e
e

FIGURE 22.15 Schematic diagram showing an experimental setup for the production of nanocrystalline powders
via EDOC. (Modified from Dierstein, A. et al., Scripta Mater., 44, 2209, 2001.)

(a)

Deposition of Various Materials

Free-Standing
Alumina
Template

+
Substrate
(b)

100 nm

Nanodots

Substrate
Nanowires

(c)

Substrate
FIGURE 22.16 Use of nanoporous alumina template in nanodot and nanowire synthesis.

general area of template synthesis of nanostructures.59,7375 In these processes, the porous alumina films
are extracted from previously anodized aluminum sheets (usually by dissolving the aluminum base) to
serve as a template (Figure 22.16a) in further synthesis steps for a variety of nanostructures. For example,
Figure 22.16b shows that deposition of other materials (e.g., metals, semiconductors) through such a
template can be used to produce regular arrays of so-called quantum dots, which show very interesting
electrical properties resulting from electronic state confinement in crystals with very small external sizes
(<100 nm). If the pores in thick aluminum oxide layers are filled up completely (Figure 22.16c), this
templating technique can be used to produce nanowires of varying diameters and lengths, which have
been shown to display exceptional mechanical, electrical, optical, or chemical properties.

2007 by Taylor & Francis Group, LLC

22-26

Materials Processing Handbook

Besides nanowires, template synthesis has been used to produce nanocables with a radial
metal/semiconductor junction.76

22.7 Conclusions
Electrochemical reactions provide the basis for an extensive number of electrolytic processes to synthesize
materials in many different shapes and forms. Numerous processing routes have been developed over
the past century for primary metal production, direct manufacturing processes, and surface-finishing
applications. Today, many of the chemical and physical fundamentals of electrolytic processing are fairly
well understood. However, numerous aspects of electrocrystallization require further study to achieve a
better control of microstructure, and therefore properties, of the final material through appropriate choice
of processing parameters such as electrolyte composition, pH, temperature, current density, and electrolyte
agitation. While electrolytic processing is overall a relatively mature technology for conventional material
synthesis, research over the past 20 years has shown that there are tremendous new opportunities for
electrolytic processes in the general area of nanotechnology. Several nanostructured materials produced
by electrolytic processes have already found application in industry, and many more processes are very
close to commercialization. The outlook for electrolytic processing is excellent, and it is expected that
many of the questions regarding electrocrystallization will be answered in the next few years through the
extensive use of powerful computational as well as nanoscale chemical and microstructural analysis tools
currently applied in the development of electrolytic nanomaterials processing technologies.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]

[14]
[15]
[16]
[17]

Vetter, K.J., Elektrochemische Kinetik, Springer Verlag, Berlin, 1960.


Potter, E.C., Electrochemistry: Principles and Applications, Cleaver Hume Press Ltd., London, 1961.
West, J.M., Electrodeposition and Corrosion Processes, Van Nostrand, London, 1965.
Raub, E. and Muller, K., Fundamentals of Metal Deposition, Elsevier, New York, 1967.
Bockris, J.O.M. and Razumney, G.A., Fundamental Aspects of Electrocrystallization, Plenum Press,
New York, 1967.
Bockris, J.O.M. and Reddy, A.K.N., Modern Electrochemistry, Vols. 2, Plenum Press, New York,
1970.
Bloom, H. and Gutmann, F., Electrochemistry The Past Thirty and the Next Thirty Years, Plenum
Press, New York, 1977.
Bard, A.J. and Faulkner, L.R., Electrochemical Methods, Wiley & Sons, New York, 1980.
Reiger, P.H., Electrochemistry, 2nd ed., Chapman and Hall, New York, 1993.
Bockris, J.O.M. and Khan, S.U.M., Surface Electrochemistry A Molecular Level Approach, Plenum
Press, New York, 1993.
Valensi, G., Van Muylder, J., and Pourbaix, M., Atlas of Electrochemical Equilibria in Aqueous
Solutions, Pergamon Press, New York, 1966.
Pourbaix, M., Lectures on Electrochemical Corrosion, Plenum Press, New York, 1973.
Gorbunova, K.M. and Polukarov, Y.M., Electrodeposition of alloys, in Advances in Electrochemistry
and Electrochemical Engineering, Vol. 5, Tobias, C.W., Ed., Interscience Publishers, Wiley & Sons,
New York, 249, 1967.
Schlesinger, M. and Paunovic, M., Eds., Modern Electroplating, John Wiley & Sons Inc., New York,
2000.
Andricacos, P.C. et al., Eds., Electrochemical synthesis and modification of materials, Mat. Res.
Soc. Symp. Proc., 451, 1997.
Merchant, H., Ed., Defect Structure, Morphology and Properties of Deposits, TMS, Warrendale, PA,
1344, 1995.
Stivison, D.S., Ed., Metal Finishing 2004 Guidebook and Directory, Elsevier, New York, 2004.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-27

[18] Wood, W.G., Ed., Metals Handbook 9th ed., Vol. 5: Surface Cleaning, Finishing and Coating,
American Society for Metals, Metals Park, OH, 1982.
[19] Brenner, A., Electrodeposition of Alloys: Principles and Practice, Academic Press, New York, 1963.
[20] Boyer, H.E. and Gall, T.L., Eds., Metals Handbook Desk Edition, American Society for Metals,
Metals Park, OH, 1985.
[21] Safranek, W.H., The Properties of Electrodeposited Metals and Alloys, 2nd ed., American
Electroplaters and Surface Finishers Society, Orlando, FL, 1986.
[22] Dennis, J.K. and Such, T.E., Nickel and Chromium Plating, John Wiley & Sons, New York, 1992.
[23] Dini, J.W., Electrodeposition, Noyes Publ., Park Ridge, NJ, 1993.
[24] Paunovic, M. and Schlesinger, M., Fundamentals of Electrochemical Deposition, John Wiley & Sons,
New York, 1998.
[25] Andricacos, P.C. and Romankiw, L.T., Magnetically soft materials: Their properties and electrochemistry, in Advances in Electrochemical Science and Engineering, Vol., 3, Gerischer, H. and
Tobias, C.W., Eds., VCH, Basel, Switzerland, 227, 1994.
[26] Celis, J.P., Roos, J.R., and Buelens, C., A mathematical model for the electrolytic codeposition of
particles with a metallic matrix, J. Electrochem. Soc., 134, 1402, 1987.
[27] Gomas, W.P. and Goossens, H.H., Electrochemistry of IIIV compound semiconductors, in
Advances in Electrochemical Science and Engineering, Vol., 3, Gerischer, H. and Tobias, C.W., Eds.,
VCH, Basel, Switzerland, 1, 1994.
[28] Okinaka, Y. and Osaka, T., Electroless deposition process: fundamentals and applications, in
Advances in Electrochemical Science and Engineering, Vol., 3, Gerischer, H. and Tobias, C.W., Eds.,
VCH, Basel, Switzerland, 55, 1994.
[29] Mollory, G.O. and Hajdu, J.B., Eds., Electroless Plating: Fundamentals and Applications, American
Electroplaters and Surface Finishers Society, Orlando, FL, 1990.
[30] Evans, U.R., The Corrosion and Oxidation of Metals, Edwards Arnold Publ. Ltd., London, 241,
1960.
[31] Wang, Y.M., Kuo, H.H., and Kia, S., Effect of alloy types on the gloss of anodized aluminum, Plat.
Surf. Fin., 91, 34, 2004.
[32] ASTM Special Technical Publication No. 318, Electroforming Applications, Uses and Properties
of Electroformed Metals, American Society for Testing and Materials, Philadelphia, PA, 1962.
[33] Watson, S.A., Applications of Electroforming, NDI Technical Series No. 10 054, Nickel Development
Institute, Toronto, Ontario, 1989.
[34] Parkinson, R., Electroforming A Unique Metal Fabrication Process, NDI Technical Series No. 10
085, Nickel Development Institute, Toronto, Ontario, 1998.
[35] Parkinson, R., Nickel Plating and Electroforming, NDI Technical Series No. 10 088, Nickel
Development Institute, Toronto, Ontario, 2001.
[36] Lyshevski, S.E., MEMS and NEMS, CRC Press, Boca Raton, FL, 410, 2002.
[37] Lyshevski, S.E., Nano- and Microelectromechanical Systems, CRC Press, Boca Raton, FL, 1, 2001.
[38] Lehmkuhl, H., Mehler, K., and Landau, U., The principles and techniques of electrolytic aluminium deposition and dissolution in organoaluminum electrolytes, in Advances in
Electrochemical Science and Engineering, Vol. 3, Gerischer, H. and Tobias, C.W., Eds., VCH,
Basel, Switzerland, 163, 1994.
[39] Delimarskii, I.U.K. and Markov, B.F., Electrochemistry in Fused Salts, The Sigma Press, Washington,
1961.
[40] Mamantor, G. et al., Eds., Molten Salts, Proceedings Vol. 87-7, The Electrochemical Society Inc.,
Pennington, NJ, 1987.
[41] Sadoway, D.R., Inert anodes for the Hall-Hroult cell, JOM, 53, 34, 2001.
[42] Kear, B.H., Tsakalakos, T., and Siegel, R.W., Eds., Materials with ultrafine microstructures,
Nanostruct. Mater., 1, 1106, 1992.
[43] Komarneni, S., Parker, J.C., and Thomas, G.J., Eds., Nanophase and nanocomposite materials,
Mater. Res. Soc. Symp. Proc., 286, 1459, 1993.

2007 by Taylor & Francis Group, LLC

22-28

Materials Processing Handbook

[44] Yacaman, M.J., Tsakalakos, T., and Kear, B.H., First international conference on nanostructured
materials, Nanostruct. Mater., 3, 1518, 1993.
[45] Schaefer, H.E. et al., Eds., Second international conference on nanostructured materials,
Nanostruct. Mater., 6, 11026, 1995.
[46] Suryanarayana, C., Singh, J., and Froes, F.H., Eds., Processing and Properties of Nanocrystalline
Materials, TMS, Warrendale, PA, 1496, 1996.
[47] Komarneni, S., Parker, J.C., Wollenberger, H.J., Eds., Nanophase and nanocomposite materials II,
Mater. Res. Soc. Symp. Proc., 457, 1588, 1997.
[48] Trudeau, M.L. et al., Eds., Third international conference on nanostructured material, Nanostruct.
Mater., 9, 1771, 1997.
[49] Ma, E. et al., Eds., Chemistry and Physics of Nanostructures, TMS, Warrendale, PA, 1241, 1997.
[50] Muhammed, M. and Rao, K.V., Eds., Fourth international conference on nanostructured materials,
Nanostruct. Mater., 12, 11188, 1999.
[51] Inoue, A. et al., Eds., Fifth international conference on nanostructured materials, Scripta Mater.,
44, 11612372, 2001.
[52] Shaw, L.L., Suryanarayana, C., and Mishra, R.S., Eds., Processing and Properties of Structural
Nanomaterials, TMS, Warrendale, PA, 1222, 2003.
[53] Erb, U., Aust, K.T., and Palumbo, G., Electrodeposited nanocrystalline materials, in Nanostructured Materials, Koch, C.C., Ed., Noyes Publ./William Andrew Publ., Norwich, NY, 179,
2002.
[54] Palumbo, G., Gonzalez, F., Tomantschger, K., Erb, U., and Aust, K.T., Nanotechnology
opportunities of electroplating industries, Plat. Surf. Finish., 90, 36, 2003.
[55] Palumbo, G., McCrea, J., and Erb, U., Applications of electrodeposited nanostructures, in Encyclopedia of Nanoscience and Nanotechnology, Nalwa, H.S., Ed., American Scientific Publishers,
Stevenson Ranch, CA, Vol. 1, 89, 2004.
[56] McMahon, G. and Erb, U., Structural transitions in electroplated NiP alloys, J. Mater. Sci. Lett.,
8, 885, 1989.
[57] McMahon, G. and Erb, U., Bulk amorphous and nanocrystalline NiP alloys by electroplating,
Microstruct. Sci., 17, 447, 1989.
[58] Erb, U. and El-Sherik, A., Nanocrystalline Materials and Process of Producing the Same, U.S. Patent
No. 5,352,266, 1994.
[59] Cheung, C., Erb, U., and Palumbo, G., Nanostructure synthesis by electrodeposition, in Processing
and Fabrication of Advanced Materials, Srivatsan, T.S. and Varin, R.A., Eds., ASM International,
Materials Park, OH, 2004.
[60] Gonzalez, F. et al., Electrodeposited nanostructured nickel for in-situ nuclear steam generator
repair, Mater. Sci. Forum, 225, 831, 1996.
[61] Palumbo, G. et al., In-situ steam generator repair using electrodeposited nanocrystalline nickel,
Nanostruct. Mater., 9, 736, 1997.
[62] Erb, U. et al., Grain size effects in nanocrystalline electrodeposits, in Processing and Properties of
Structural Nanomaterials, Shaw, L.L. et al., Eds., TMS, Warrendale, PA, 109, 2003.
[63] Koch, C.C. and Scattergood, R.O., Grain size distribution and mechanical properties of nanostructure materials, in Processing and Properties of Structural Nanomaterials, Shaw, L.L. et al., Eds.,
TMS, Warrendale, PA, 45, 2003.
[64] Integran Technologies Inc., Toronto, unpublished data.
[65] Zimmerman, A.F. et al., Mechanical properties of nickel silicon carbide nanocomposites, Mater.
Sci. Eng., A328, 137, 2002.
[66] Erb, U., Palumbo, G., and Aust, K.T., Electrodeposited nanostructured films and coatings, in
Nanostructured Films and Coatings, Chow, G.M. et al., Eds., NATO Science Series 3, Vol. 78,
Kluwer Academic Publ., Dordrecht, The Netherlands, 11, 2000.

2007 by Taylor & Francis Group, LLC

Electrolytic Processes

22-29

[67] Paunovic, M., Schlesinger, M., and Weil, R., Fundamental considerations, in Modern Electroplating, Schlesinger, M. and Paunovic, M., Eds., John Wiley and Sons, New York, 37,
2000.
[68] Switzer, J.A., Electrochemical architecture of ceramic nanocomposites, Nanostruct. Mater., 1, 43,
1992.
[69] Aizawa, M. et al., Silver nano-inukshuks on germanium, Nano Lett., 5, 815, 2005.
[70] Hsiao, G.S. et al., Hybrid electrochemical/chemical synthesis of supported luminescent semiconductor nanocrystallites with size selectivity: Copper (I) iodide, J. Am. Chem. Soc., 119, 1439,
1997.
[71] Hempelmann, R. and Natter, H., German Patent, DE 198-40-841, 1998.
[72] Dierstein, A. et al., Electrochemical deposition under oxidizing conditions: EDOC, Scripta Mater.,
44, 2209, 2001.
[73] Al Malawi, D. et al., Electrochemical fabrication of metal and semiconductor nanowire arrays,
in Nanostructured Materials in Electrochemistry, Searson, P.C. and Meyer, G.J., Eds., The
Electrochemical Society, Inc., Pennington, NJ, 262, 1995.
[74] Masuda, H. and Fukuda, K., Ordered metal nanohole arrays made by a 2-step replication of
honeycomb structures of anodic alumina, Science, 268, 1466, 1995.
[75] Sawitowski, T., Beyer, N., and Schulz, F., Bio-inspired antireflective surfaces by imprinting processes, in The Nano-Micro Interface, Fecht, H.J. and Werner, M., Eds., Wiley VCH, Weinheim,
Germany, 263, 2004.
[76] Ku, J.-R. et al. Fabrication of nanocables by electrochemical deposition inside metal nanotubes,
J. Am. Chem. Soc., 126, 15022, 2004.

2007 by Taylor & Francis Group, LLC

Das könnte Ihnen auch gefallen