Você está na página 1de 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/286130823

Corrosion of Lead and its Alloys

Article · December 2010


DOI: 10.1016/B978-044452787-5.00098-6

CITATIONS READS

5 1,987

1 author:

Stuart Lyon
The University of Manchester
213 PUBLICATIONS   2,879 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Chromate free corrosion inhibiting coating technology View project

475°C Embrittlement - Comprehensive Characterisation of Microstructure and Corrosion Properties View project

All content following this page was uploaded by Stuart Lyon on 08 June 2018.

The user has requested enhancement of the downloaded file.


Lead and its Alloys
Original article in 3rd Edition of Shreir by:
P.C. Frost, Cookson Group, E. Littauer, Lockheed Missiles and Space Co. and H.C. Wesson,
Lead Development Association

Revised for the 4th Edition by S.B. Lyon (Section Editor)

X.1 Introduction
X.1.1 History and production
Lead is one of the metals known in antiquity; there is evidence of lead production dating back
to before the early Bronze Age with a lead figurine from Upper Egypt dated to approximately
3000BC [1]. However, at that time lead does not appear to have been greatly prized since it is
soft and easily tarnishes. Indeed, early lead ores seem to have been smelted mainly in order to
obtain silver with which it frequently occurs. However, lead production grew very
significantly during the Roman period, where it was useful in its own right for an number of
purposes including plumbing (water transport, water tanks, etc.) and kitchenware and drinking
vessels (despite the knowledge, even in antiquity, that lead was generally harmful to human
health).

Lead ores comprise mainly sulphide mineralisation with smaller amounts of carbonate and
sulphate. Historically lead was mined on its own (in the UK, Derbyshire and Leadhills in
Southern Scotland were historic sources). Modern ores frequently occur along with zinc, and
other species (silver, copper, gold, antimony, bismuth, etc.) in considerably lesser amounts.
Lead sulphide (galena) is easily smelted by roasting in air to produce lead oxide then by
carbothermic reduction, which traditionally used charcoal but now uses coke, in a blast
furnace.

Most primary lead is currently produced in Australia, China, USA, Canada, Mexico and Peru.
Western Europe produces about 8% of the total world ore production of around 3 million
tonnes (2007), mainly from Sweden but also Ireland and Spain. Lead is amongst the most
recyclable of all metals and has by far the highest recycling rate, primarily from batteries, but
also from usage in the building trade. Secondary (i.e. recycled) sources of lead generally
account, in developed economies, for more 70% of lead usage.

X.1.2 Physical properties


Lead is characterised by its relatively high density (in comparison to other engineering
materials), its extremely low hardness and strength and its favourable electrochemical
properties including its good resistance to corrosion [1,2]. However, it should be noted that
lead is by no means the most dense element; for example, tungsten and tantalum are
significantly more dense. Table 1 provides a summary of some physical properties in
comparison with other materials.

The density of lead provides it with useful properties in respect of the attenuation of phonon
vibrations and electromagnetic waves. Thus, the mass attenuation coefficient for lead is
relatively high, and also the neutron absorption cross-section for lead is small, which makes
lead effective for the radiation shielding of X-rays, gamma rays and neutrons. In addition, the
softness and high mass of lead effectively damps the propagation of sound waves.

Pb Sn W Cu Fe
Atomic number 82 50 74 29 26
Atomic weight (g) 207.19 118.69 183.84 63.55 55.85
-3
Density (g cm ) 11.34 7.3 19.3 8.96 7.87
UTS, 99.9% pure (MPa) 12-15 20-30 950-1000 120-170 100-200
Melting point (K) 601 505 3680 1358 1810
Boiling point (K) 2020 2540 5830 2816 3130
Electrical conductivity x 106 (ohm-1 cm-1) 0.048 0.079 0.189 0.596 0.099
-1 -1
Thermal conductivity (W cm K ) 0.35 0.73 1.74 4.01 0.238
-6
Linear expansion coefficient x 10 29.0 23.5 - 16.5 12.2
Crystal structure (room temperature) fcc bct bcc fcc bcc

Table 1: Physical properties of lead compared with some other common metals

X.1.3 Applications
Traditionally, lead as a metal has a number of uses driven by its general resistance to
corrosion, as well as its electrochemical properties in rechargeable lead-acid batteries. The
main uses of metallic lead are [2]:

1. Lead-acid batteries: this currently accounts for over 50% and up to 90% of use in
certain countries; most lead used in this way is recovered and recycled into the
secondary lead market.
2. Construction: the main application in construction is for waterproofing flashings and
roofing of historic buildings, as well as architectural detailing; the traditional use in
water pipes and water storage has now vanished; lead from construction also has a
high recycling rate.
3. Chemical industry: lead lining of vessels was traditionally used in the chemical
industry, particularly in the production of sulphuric acid; however, these uses have
now almost vanished.
4. Metal joining: lead is (with tin) a key component of solder used to join copper (piping,
heat exchangers, etc) and steels, and for electrical contacts; however, lead-free solders
are now widely used in the water and electronics industry.
5. Munitions: lead is traditionally used in the manufacture of ammunition although
depleted uranium and tungsten, which are both significantly more dense than lead, are
used in high performance (e.g. armour piercing) rounds; however for sporting
purposes non-toxic alternatives exist and are increasingly being used.
6. Dense metals: lead is used for radiation shielding, for sound dampening and for
balancing of machinery; weights used in sporting applications (such as fishing weights)
are being phased out in favour of non-toxic alternatives.
7. Cable sheathing: the first electrical cables were sheathed with lead to protect the cable
from moisture ingress. For undersea cables, especially high-voltage power
transmission lines, there is still no adequate alternative; thus, properly applied lead
sheathing has zero water and water vapour transmission.
8. Alloying: lead is used as an addition in a number of alloys, including: solders, pewter,
bearing alloys, etc.; it also may be added to other materials (such as copper alloys and
steels) to improve machinability, although alternatives are now available.
9. Anodes: by virtue of its relative chemical stability and easy passivation, lead is used as
an insoluble anode in many electroplating processes (e.g. for chromium); traditionally
it was also used as a non-consumable anode in cathodic protection systems but this
application area is decreasing as alternatives now exist.
10. Fusible alloys: eutectic alloys containing lead with bismuth, tin, cadmium, indium are
used as low melting point fuses and plugs in certain applications such as sprinkler
systems etc.

X.1.4 Alloying
Apart from tin, indium and bismuth, lead has extremely limited solid solubility for most
elements at room temperature although solubility can increase somewhat at higher
temperatures [4]. Thus, most lead alloys consist of primary lead grains either with second
phase particles (which may be intermetallic compounds), or with a eutectic mixture,
decorating the grain boundaries. Due to the very poor mechanical properties of high-purity
lead, alloying is almost always essential in order to achieve acceptable performance. Most
commercial lead alloys contain a few tenths of a percent of alloying elements with
strengthening caused by precipitation or by the presence of a eutectic; many alloys are age-
hardenable. The main classes of alloy listed below with composition ranges presented in
Table 2:

1. Lead-antimony (+ tin, arsenic): the traditional alloy used for batteries, ammunition,
sheet and chemical plant; still very widely used but the battery application is
increasingly being replaced with lead-calcium.
2. Lead-calcium (+ tin, aluminium): increasingly replacing lead-antimony alloys
particularly in battery applications as they have very low hydrogen evolution
characteristics and can therefore be sealed. Lead-calcium-tin is used for electro-
winning anodes
3. Lead-tin (+ silver, antimony): used for jointing alloys (solders) and bearings.
4. Lead-copper: used generally for sheet products and linings; also for cable sheathing
and in chemical plant.
5. Lead-silver: used for anode materials and higher temperature solders
6. Lead-tellurium (+ copper): because of its low alloy content used for radiation
shielding and, with copper, is a cable sheathing alloy.
Alloy type Specification Application Composition

Pb-Sb Battery grids 1-3% Sb + 0.5-1% S, Se and Cu


Ammunition 0.5-8% Sb + 0.02-3% As
BS EN 12548 Cable sheaths 0.5-1% Sb
Anodes 6-10% Sb + 0.5-1% As
BS EN 12659 Chemical lead < 0.06% impurities
BS ISO 4381 Bearings 9-15% Sb + 1-20% Sn

Pb-Ca Battery grids 0.03-0.07% Ca; + 0.03% Al and 0.3% Sn


Cable sheath/splices 0.04% Ca
Anodes 0.03-0.09% Ca; + up to 0.3% Sn

Pb-Sn BS EN ISO 9453 Solders 2-70% Sn; + up to 2% Sb


BS ISO 4999 Terne plate 15-20% Sn

Pb-Cu BS EN 12588 Construction (sheet) 0.03-0.06% Cu

Pb-Ag Battery grids 0.005-0.05% Ag


Solder 1-6% Ag
Anodes 0.25-2% Ag; + up to 1% Sn or up to 6% Sb

Pb-Te BS 3909 Nuclear/radiation 0.06% Te


BS EN 50307 Cable sheaths 0.035-0.1% Te; + 0.03-0.08% Cu

Table 2: Compositions of typical lead alloys [5].

X.1.5 Mechanical properties

X.1.5.1 Lead
Unalloyed lead has relatively poor mechanical properties [2,3] thus, in addition to a low
ultimate tensile strength (< 12 MPa), it has a very low yield stress (<4 MPa); these are
considerably lower than tin, for example. The low melting point of lead means that, like tin, it
undergoes re-crystallisation and grain growth at room temperature and is susceptible to creep
at stresses above about 2 MPa (about 15% of the UTS), which is very low indeed. Thus, while
lead of purity in excess of 99.99% is commercially available (BS EN 12659) it is not
specified unless its superior corrosion resistance is required. Lead is also susceptible to
fatigue, particularly due to thermal cycling, and this is a key failure mechanism in pure and
alloyed lead. Table 3 provides typical mechanical properties for pure lead as well as a range
of common lead alloys.

Yield Stress, UTS, Elongation, Fatigue Creep failure at


Material MPa MPa % limit, MPa 20 MPa, hours
99.99% Pb 3.5 12.2 55 2.7 -

0.06% Cu 9.0 17.5 55 4.9 -


0.06% Ca 24.3 27.9 30 9.0 10
0.5% Sn 27.9 41.8 15 - 30
0.08% Ca + 1% Sn 46.0 59.7 15 - 850

1% Sb 19.3 37.9 20
2% Sb 24.1 46.9 15
2.75% Sb 55.2 65.5 10
6% Sb 71.0 73.8 8

Table 3: Mechanical properties of selected lead and lead alloys [5].


X.1.5.2 Lead Alloys
Of the elements commonly found in lead alloys, zinc and bismuth aggravate corrosion in most
circumstances, while additions of copper, tellurium, antimony, nickel, silver, tin, arsenic and
calcium affect corrosion resistance only slightly, or may even improve it depending on the
service conditions. Alloying elements that are of increasing importance are calcium,
especially in maintenance-free battery alloys and selenium or sulphur, together with copper as
grain refiners (nucleants) in low antimony battery alloys. Other elements of interest are
indium and silver in anodes [6], aluminium in batteries (to control calcium losses) [7], and
selenium in chemical lead as a grain refiner [8]. In Europe, lead alloy designations are
specified in ISO TR 7003: “Unified format for the designation of metals”, in the format
PBnnnA. The designate “PB” is the chemical symbol for lead, the 3 digits “nnn” define
specific alloy compositions, while the “A” designates the application: “R” for pure lead, “K”
for cable, “A” for battery alloys and “M” for miscellaneous alloys.

Historically, lead for use in chemical plant was specified in BS 334, which defines
compositions for 5 types of lead: A B1, B2, B3 and C. This standard has now been superseded
by BS EN 12659: “Lead and lead alloys” which specifies the composition of effectively pure
lead (formally Type A). This is because satisfactory alternative materials exist such that lead
is now rarely if ever specified in chemical plant. Type A lead should only be used in a
vibration-free environment and where the superior corrosion resistance is of paramount
importance. Historically, for general chemical plant use, type B1 (copper lead) is to be
preferred on account of its much greater structural stability, especially at elevated
temperatures. Type B2 (copper tellurium lead) has extremely good fatigue resistance which is
retained to a greater extent at elevated temperatures than type B1. The main effect of
tellurium is to form a fine-grained uniform grain structure, to enhance work hardening, and to
delay re-crystallisation. The silver content in type B3 also delays re-crystallisation and
promotes a large-grained stable structure which is creep-resistant. Type C (antimony lead) is
used for valves, pump bodies and fatigue-resistant applications, but is not suitable for use at
temperatures above 60°C owing to a rapid increase in creep rate, or in sulphuric acid
concentrations above 60%.

BS EN 12548: “Lead alloy ingots for electric cable sheathing and for sleeves” gives
compositional requirements for a range of lead alloys for this application. The main types are
PB001K (formally type B), containing 0.85% antimony; PB021K (formally type E),
containing 0.2% Sb and 0.4% Sn; and PB012K (formally 1/2C), containing 0.2% Sn and
0.07% Cd. Type B is suitable for use in environments where severe vibration is expected,
while Type E is relatively resistant to vibration compared to Type 1/2C. The performance of
these materials is generally adequate in underground or marine environments.

BS EN 12588: “Rolled sheet lead for building purposes”, lays down requirements for
composition, structure, thickness, freedom from defects, width and length, and marking. The
specified copper content stabilises the structure of the material, conferring resistance to
thermal fatigue cracking caused by grain growth and thermal cycling.

Other UK, European or worldwide specification standards exist for specific applications, for
example: radiation shielding, soft solders and bearing alloys. However, there are no standards
for battery alloys as many of these are proprietary to specific battery manufacturers.
X.2 Electrochemistry
X.2.1 Thermodynamics
Pourbaix et al. [9] studied the Pb-H2O and Pb-H2O-X systems where X is a non-metal, and
have established the domains of thermodynamic stability of lead, lead cations and anions, and
insoluble compounds of lead. Figure 1 shows the Pb-H2O system in the absence of
complexing agents such as acetic acid. Lead can be seen to be a relatively noble metal from
pH > 5, but dissolving to Pb2+ species at lower pH and the plumbite oxyanion at pH > 10.5.
This is due to the amphoteric nature of lead and is a significant factor its in actual
environments. Also, passivation at high potential, due to the formation of PbO2, is evident
across the whole pH domain.

In regions where the dissolution of lead is possible, according to the thermodynamics, the rate
of corrosion may be very slow. This is because the overpotential for hydrogen evolution on
lead is very high in 1M H2SO4: 10-12 A cm-2 with a Tafel slope of 0.125 mV (decade current)-1,
indicating that a 1-electron transfer reaction is rate controlling [10]. Corrosion in near neutral
environments generally results in passivation with the development of insoluble salt film
corrosion products. In strong alkali, the corrosion is more rapid with the formation of
plumbite ions [11]. Consequently, in acidic or moderately alkaline solutions, free from
complexing agents (particularly organic acids), the corrosion of lead is generally negligible.

Figure 1: Pourbaix diagram for lead at a dissolved total metal ion concentration of 10-5 M.

In contrast with the Pb-H2O system, in the presence of significant concentration of sulphate
ion, the zone of thermodynamic stability of PbSO4 expands and lead now passivates by
formation of a salt film of lead sulphate, which is the basis of lead’s low corrosion rate in
sulphuric acid, Figure 2.
Figure 2: Pourbaix diagram for lead-sulphur in 1 M sulphate ion concentration and a
dissolved total metal ion concentration of 10-5 M.

In similar way, lead forms a series of relatively insoluble compounds, many of which are
strongly adherent to the metal surface. Thus, in conditions where a stable continuous film can
form, further reaction is often prevented or greatly reduced. The generally good corrosion
resistance of lead results, therefore, from the formation of relatively thick protective films of
corrosion product.

X.2.2 Dissolution
The standard electrode potential, E°: Pb2+/Pb = − 0.126 V [12], shows that lead is thermo-
dynamically unstable in acid solutions but stable in neutral solutions. In acid conditions, the
exchange current density for the hydrogen evolution reaction on lead is very small (~ 10-12 A
cm-2) but, at higher pH, control of corrosion is usually due to physical blocking of local
anodes (from alloying additions and intermetallic compounds, which generally have higher
exchange current densities) by virtue of insoluble lead salts that frequently form protective
films. At pH > 11, lead will corrode freely at low anodic overpotentials dissolving as the
HPbO2- species. However, at higher potentials, it will passivate, again forming PbO2. Lead
may thus be used as an insoluble anode effectively throughout the entire pH range; an
extremely useful characteristic. In summary, the corrosion rate of lead in an environment is
frequently not controlled by the electrochemical processes but by the chemical dissolution
rate of the corrosion product film.

Under strong cathodic polarisation, disintegration can occur with lead, observable as a grey
cloud of fine metal particles; this is a significant phenomenon in the application of lead to
electrowinning (metal plating) anodes. Hydrogen evolved on the surface of the lead can be
absorbed if the current density is sufficiently high [13] and, above this level, the formation of
lead hydride may occur, which leads to severe disintegration [14].

X.2.3 Oxidation and passivation


Lead is characterised by a series of anodic corrosion products which give a film or coating
that effectively insulates the metal physically from the electrolyte, e.g.: PbSO4, PbCl2, Pb3O4,
PbCrO4, PbO, PbO2, 2PbCO3·Pb(OH)2; of which PbSO4 and PbO2 are the most important,
since they play a part in batteries and anodes. Lead carbonate and lead sulphate are
particularly important also in atmospheric passivation and chemical industry applications.

In aqueous electrolytes, the anodic behaviour of lead varies greatly depending on the
conditions prevailing. Extensive reviews of the aqueous electrochemistry of lead, including its
anodic behaviour have been produced, for example by Kuhn in 1979 [15], and much of this is
still valid. The vast majority of historic and current research is concerned with the behaviour
of lead in sulphuric acid and this is understandable given that the main application is still in
lead-acid batteries. More recently, research has been concerned increasingly with
environmental concerns, including lead removal from contaminated locations and the possible
application of lead as a barrier material in nuclear waste disposal; these specific issues are
dealt with in more detail below.

The dry oxidation of lead was studied in oxygen pressures from 10-2 to 1000 Torr and at
temperatures between -90 to 150°C [16]. Over several weeks, the oxide grew to a limiting
thickness of about 10 nm. Ellipsometry and Auger Electron Spectroscopy confirmed that the
oxide was PbO and that the growth law conformed to the Cabrerra-Mott model (high-field ion
conduction). X-ray diffraction data confirmed that the oxide was epitaxial to the lead but that
it was highly strained and that recrystallisation of the lead substrate promoted oxide growth.

Passivation of lead in sulphuric acid is caused by the potential-dependent, and generally


reversible, formation of PbO, PbSO4 and PbO2. However, the reactions and transformations
are complex and difficult to characterise [17]. At potentials just above the reversible potential
for lead oxidation, a dense layer of PbO can form together with porous PbSO4 [18,19]; the
growth of PbO formed in this way is controlled by the rate of oxygen anion transport through
the oxide. At higher potentials, PbO and PbSO4 transform to PbO2, which occurs in one of
two non-stoichiometric polymorphs, both with a slight oxygen deficiency. Lead anodic
corrosion proceeds in two stages in sulphuric acid. Initially, Pb is oxidized to tetragonal PbO,
as above, then in the second stage this is oxidized to PbO2. If this process is performed in
solid state, α-PbO2 forms; however, β-PbO2 can be formed by a dissolution-precipitation
mechanism [20,21]. The oxide species are O2- anionic conductors; additionally, there is
evidence from studies in batteries, that PbSO4 is a Pb2+ cationic conductor [22].

Passivation of lead also occurs in a range of other media the most important of which is
carbonate/bicarbonate as these equilibria are commonly encountered in natural environments
[23]. The passivation sequence on anodic polarisation is similar to that in sulphuric acid
except for the formation of lead carbonate instead of lead sulphate. Thus, Pb may be directly
oxidised to PbO or Pb(OH)2, which then reacts with CO2 (in the atmosphere) or bicarbonate/
carbonate ions (in solution) to form protective PbCO3,; at higher potentials, PbO2 is formed.
The rate of growth of the passive carbonate film is dependant on the concentration of
carbonate in solution. However, unlike in sulphuric acid, the passivation process in carbonate
media is essentially irreversible. Although pitting in sulphuric acid is possible (e.g. by
perchlorate ions) [24], it is not the common degradation mechanism. However in carbonate
media, pitting is observed with a number of species in solution including perchlorate [25] and
nitrate [26].

In summary, lead is passive at higher potentials at all pH forming lead dioxide, which is a
defect anion conductor. At intermediate potentials and in solutions of near neutral to acid pH,
lead generally passivates by formation of a salt film characteristic of the solution, for example:
PbSO4 in solutions of sulphuric acid. In such cases, the corrosion rate of lead tends to be
controlled by the chemical dissolution of the corrosion product (salt film) and not by the
potential. At intermediate potentials and high pH, lead will tend to dissolve forming the
plumbite ion HPbO2-. In strong alkali (1M NaOH) formation of Pb(OH)2 occurs at less
positive potentials, which transforms to PbO according to a nucleation and growth
mechanism under diffusion control [11].

X.2.4 Corrosion products


The nature (composition and morphology) of the corrosion product on lead will very
frequently enable a good interpretation of the environment in which the material was exposed.
However, note that lead dioxide cannot form under ordinary environmental conditions and
that other metals cannot polarise lead sufficiently for it to form. Lead dioxide corrosion
products can, therefore, only be produced by external polarisation, either deliberately (as in a
battery or anode) or accidentally (e.g. via stray currents).

Compounds of lead Formula Solubility at 25°C


Acetate (CH3COO)2Pb 550 g dm-3
Formate (CHOO)2Pb 16 g dm-3
Nitrate Pb(NO3)2 600 g dm-3

Chloride PbCl2 1 x 10-4 g dm-3


Hydroxide Pb(OH)2 1 x 10-8 g dm-3
Carbonate PbCO3 3.3 x 10-14 g dm-3
Basic carbonate 2PbCO3.Pb(OH)2 as above
Sulphate PbSO4 4 x 10-15 g dm-3
Sulphide PbS 3.4 x 10-28 g dm-3
Phosphate Pb3(PO4)2 10-55 g dm-3

Oxide PbO insoluble


Dioxide PbO2 insoluble

Table 4: Solubility of lead compounds [27]

From an inspection of the more common compounds of lead, it will be seen that, in many
environments, the corrosion product will be relatively insoluble, Table 4. Often, however,
compact protective films are prevented from forming on the surface of the metal. The nature
of the film is influenced by the mode of crystallisation, and in the case of the lower oxides for
example, frequently little protection is afforded. Lead dioxide often forms a good adherent
film, especially when it is produced from a sulphate film or other adherent compounds during
anodising. Concentrated hydrochloric acid gradually dissolves it to form hexachloroplumbic
acid, and with alkalis, plumbites are formed. With sparingly soluble salts of lead, the
compactness of the deposits is likely to be strongly influenced by the concentration of the
relevant anion; with low concentrations tending to result in imperfect coatings.
X.2.5 Galvanic corrosion
Lead is relatively noble compared to most other metals in common use, apart from copper,
and therefore there is normally little concern about its galvanic corrosion since it will usually
be the cathode in a coupled system. Where it is anodic, it will often not corrode significantly
due to the formation of protective corrosion product films. However, increases in temperature
may cause changes in the porosity of the protective film, or a phase change that, consequently,
reduces its protective function. For example at 75°C lead was found to be anodic to steel in
seawater and groundwater due to a polarity reversal [28]. However, the expected order of
potentials was restored if the water also contained a Bentonite clay suspension.

There are continuing concerns regarding galvanic corrosion involving lead in drinking water
systems, where lead-containing solders are traditionally used (but are now being phased out).
Particularly in plumbo-solvent soft waters, lead is leached out from solders in measurable
quantities that may be harmful to health [29]. Clearly the best solution to this problem is not
to use lead however, for the large numbers of historic properties where lead is still present in
pipework, then inhibition of this form of attack, for example using silicates or phosphonates,
is possible [30]. However, the ratio of chloride to sulphate is critical in controlling the leach
rate of lead, and water utilities need to take note of this also in order to minimise the lead
solubility of the water supply [31].

X.3 Corrosion
X.3.1 Atmospheric corrosion
Lead has been (and especially in historic and prestige properties still is) used for roofing,
gutters, flashings, downspouts, etc. and exhibits excellent resistance to air (dry or humid) and
rain water, producing, after an initial period of time, an attractive patina. Initially an oxide
(PbO) is formed, which then converts by reaction with atmospheric carbon dioxide to
plumbonacrite (Pb5O(OH)2(CO3)2) and hydrocerrusite (Pb3(OH)2(CO3)2) [32]. In atmospheres
containing sulphur dioxide from industrial pollution, the general sequence of corrosion
product (patina) formation is: lead oxide → basic lead carbonate → normal lead carbonate →
normal lead sulphite → normal lead sulphate [33]. Lead is found to be reactive to common
atmospheric gases, NOx, SO2, CO2, as well as the vapours of carboxylic acids [34]. As noted
above, outdoor exposures were found to produce anglesite (PbSO4) and/or cerussite (PbCO3)
while indoor exposures often produced lead carboxylates, due to higher levels of organic acid
vapours from wood and wood products. Excluding the carboxylates, these corrosion products
have a low solubility and are protective, however, they can produce a white flocculant ‘run-
off’ in wet weather, which can stain surrounding surfaces in the very early stages of exposure
[35].

In marine environments, the initial oxide film reacts with sodium chloride when wet to
produce basic lead chloride, which may result in corrosion of adjacent materials such as
aluminium [36]. In such environments, the patina stabilises, but takes approximately twice as
long as in other atmospheric environments to do so. A common treatment for new lead is a
resin based patination oil which suppresses the formation of basic carbonates allowing the
slow controlled growth of a strongly adherent normal carbonate patina from the outset [38].

Due to increasing concerns about lead in the environment, corresponding concerns regarding
the release of lead from buildings are also evident. Thus, lead ions can be introduced into the
biosphere via corrosion and subsequent run-off of Pb2+ species primarily from the solubility
of cerrusite (lead carbonate), hydrocerrusite (lead hydroxy carbonate) and anglesite (lead
sulphate) in rainwater. After an initial induction period, the measured release rate for lead ions
was found to be 10-5 moles of Pb per litre, or precipitation run-off per square metre of lead
surface (i.e. 2.1 mg dm-3) in both marine and rural locations [38], which corresponds to a
corrosion rate of about 0.3 μm y-1 in these atmospheres.

It has been known for centuries that certain species of wood are “more aggressive” to lead
than others and this is now understood to be caused by the breakdown of cellulose species in
the timber to volatile organic compounds (VOCs), especially acetic acid [39]. These can be
liberated by new wood, especially oak, and also by varnishes, glues, urea formaldehyde,
plastics, fabrics and drying-oil paints, which can liberate fumes for a considerable time after
application. Acetic acid in the 100-200 ppb range strongly accelerates the corrosion of lead
giving a corrosion product containing lead acetate-oxide-hydrate (Pb3O2(CH3COO)2·H2O) as
well as cerrusite and hydrocerrusite [40]. The corrosion rate was found to be independent of
relative humidity from 45% to 95% RH. This effect of VOCs on lead corrosion is significant
in the context not only of buildings, but also in museums for display of historic artefacts [41].
However the use of fatty acid soaps, such as sodium decanoate, successfully provides 99.9%
inhibition to corrosion, by the formation of lead soaps, which are well known as corrosion
inhibitors [42].

X.3.2 Water
X.3.2.1 Distilled and condensed water
In distilled water free from dissolved gases, corrosion is slight though significant however,
the rate of corrosion is increased by the presence of oxygen. Small amounts of dissolved
oxygen and carbon dioxide cause rapid attack since the formation of a continuous protective
film of lead carbonate does not occur. At moderate CO2 concentrations, passivation of the
lead surface occurs due to PbCO3 formation, but corrosion and in high CO2 contents,
corrosion is increased due to the formation of soluble lead bicarbonate [43].

Condensation corrosion is a common cause of failure in lead-work on buildings. Trapped


water is evaporated from and condensed on the underside of the lead during thermal cycling
in the environment. This repeated condensation causes the production of lead oxide and lead
hydroxide which migrates away from the surface, leaving it unpassivated. Subsequent
reaction with CO2 in the atmosphere produces copious quantities of basic lead carbonate,
resulting in blistering, perforation, and finally disintegration of the lead [37]. Adequate
ventilation and adherence to codes of practice are essential to prevent this [44]. Lead sheet
can also crack through thermal fatigue and, hence, admit water into a building structure. This
is a consequence of over-fixing the lead, using sheets which are too large, or of using lead
containing insufficient copper and can be eliminated by correct installation.

X.3.2.2 Natural waters


The European Union directive on the reduction of hazardous materials [45] stipulates that lead
is no longer permitted for use in water supply, either as pipework or in solders. Where it is
encountered in old buildings, its replacement is strongly recommended. However, because of
the long life of existing installations, water may be conveyed through existing lead pipes, or
pipes secured using lead-containing solders, for years to come. This may not be hazardous if
the waters contain sufficient carbonate, sulphate or silicate, and are alkaline. The current UK
limit for lead in cold water drawn from a water supply tap is 25 ppb (the US limit is lower at
15 ppb); anything above these levels is actionable. Soft waters invariably dissolve lead to
some extent and raised levels of organic acids will render even harder supply waters
plumbosolvent [46]. Thus, water treatments to reduce lead solubility are required.

Lead corrodes slightly in most supply waters dependent primarily on the solution pH. The
presence of oxygen and oxidizers in the aqueous medium affects the corrosion resistance of
the metal more in acidic waters [47]. The most common treatments given to water are to
increase the pH (reduce acidity) and remove organic acids; this is easily carried out by
flocculation (e.g. with aluminium sulphate) and alkalisation (e.g. with calcium hydroxide or
carbonate), which will encourage the formation of a protective carbonate scale [48]. Silicate
and orthophosphate treatments reduce lead solubility however, the common use of zinc
orthophosphate to control corrosion of steel and cast iron water mains, may increase lead
corrosion rates [49,50]. In more saline waters, including seawater, lead usually has a good
corrosion performance owing to the formation of the normal passive film of hydrocerrusite
(Pb3(OH)2(CO3)2) together with Pb(OH)Cl and PbCl2 carbonate-chloride double salts [51],
however, corrosion rates are linked to the activity of Cl- ions particularly in flowing solution
[52].

X.3.3 Buried structures

X.3.3.1 Stray-current corrosion


Stray currents are a source of damage to all buried metal structures and lead pipes and cable
sheaths are particularly susceptible. Although lead can corrode under cathodic (alkaline)
conditions, it is generally the anodic sites on the pipe or cable sheath that are attacked. Lead is
considered to be endangered if the current density is more than 25 mA m-2. This is influenced
by the conductivity of the soil, which is largely determined by the moisture content, but may
be affected by salting of roads in winter. Non-metallic links in pipework may break electrical
continuity. This will produce more numerous corrosion sites, but they are frequently less
intense [53]. The surface will normally be covered with a mostly whitish corrosion deposit
associated with either a smooth pitted surface or a more general rough etched appearance [43].
The corrosion product may comprise oxides, carbonates, hydroxides and PbCl2·Pb(OH)2 and
PbCl2·6PbO·2H2O have additionally been identified. Stray alternating current can also
markedly increase corrosion rates [54], and the use of lead pipes for electrical grounding has
resulted in serious corrosion [53]. Electrolytic corrosion may also occur on the inside of cable
sheaths by the passage of current from the cable sheath to the wire [53]. No protective coating
is fully effective, but many give good protection with modern systems using polymers [55].

X.3.3.2 Underground corrosion


Lead has a long history of use underground in buried structures; historically for water supply,
more recently for lead sheathed electrical cables. Corrosion is promoted by stray currents,
local inhomogeneities in the lead, galvanic corrosion with another metal and the environment
of the soil [43]. Thus, soils may differ in water content, degree of aeration, or the presence of
various chemicals or bacteria and these can cause local or extensive corrosion of lead.
Historically, extensive long-term tests have been conducted on lead in soils and these data are
still valid [56,57]. The worst combination of soils is wet clay and cinders (ashes from coal
combustion). The carbon in the cinder acts as an efficient cathode and severe anodic corrosion
takes place in the clay environment. Moisture held in the clay permits the passage of
relatively high currents. Anodic corrosion can occur when cables are in contact with
dissimilar metals such as steel support racks or copper bonding ribbon. A new (clean) section
of cable may also become anodic to an old (passivated) cable and can then corrode.
Sandy soils and loams of high permeability are less aggressive since water tends to be mobile,
so reducing concentration cells, and frequently drains readily to allow free movement of
oxygen, thus reducing the effect of differential aeration cells. Where air circulates freely, a
stable patina is often formed. Very large grained soils are normally good for the reasons given
above, but under certain conditions severe localised pitting can be caused due to differential
aeration (e.g. caused by differing degrees of soil density and compaction). Soils of low
permeability, such as clays and silts, tend to be most corrosive however, if the soil becomes
dearated the corrosion rate may fall; under such environments microbial effects, due to
sulphate reducing bacteria, often become important and corrosion can then become severe.
Trenches are often lined before cable laying and backfilled after laying, with materials such as
sand and crushed chalk, in order to encourage good drainage and ensure a consistent and
benign local environment.

Sulphates, silicates, carbonates, colloids and certain organic compounds in soils act as
inhibitors if evenly distributed [43]. Nitrates tend to promote corrosion, especially in acid soil
waters, due to cathodic de-polarisation and to the formation of soluble nitrates. Alkaline soils
can cause serious corrosion with the formation of the plumbite anion which decomposes to
give PbO. Organic acids and carbon dioxide from rotting organic matter also have a strong
corrosive action. Pitting corrosion in most groundwater/soil environments does not occur in
the traditional meaning of the term however, areas of non-uniform, general attack did occur,
resulting in pitted surface morphologies [58].

Calcium hydroxide leached from incompletely cured concrete causes serious corrosion of lead.
This is because carbon dioxide reacts with the lime solution to form calcium carbonate, which
is practically insoluble. Carbonate ions are therefore not available to form a passive film on
the surface of the lead [43]. Typically, thick layers of PbO are formed, which may show
seasonal rings of litharge (tetragonal PbO) and massicot (orthorhombic PbO) [59].

To prevent underground corrosion, lead was traditionally protected with coatings of tar,
bitumen, resin, etc.; recently, however, polymeric materials (such as polyvinyl chloride,
polypropylene, etc.) are preferred. No coating is wholly effective unless they completely
insulate the metal from corrosive agents and stray currents. The most successful method of
protection for lead cables is cathodic protection. It is effective at a potential of E° = -0.8 V or
about 0.1 V more negative than its equilibrium potential in the soil in question [60]; both
impressed currents or sacrificial anodes have been used. Excessively negative potentials can
increase the pH of the environment, thus causing corrosion.

Sulphate-reducing bacteria in soils can produce metal sulphides and H2S, which results in the
formation of conductive lead sulphide and increased corrosion rates [61]; deep pits containing
a black mass of lead sulphide have also been observed. Other micro-organisms may also be
involved in the corrosion of lead in soil, including other bacteria and fungi [62].

Interest in lead as a possible barrier material for use in underground repositories for nuclear
waste has prompted a number of studies. Repository environments with static groundwater
would have very low oxygen contents and high ionic strengths and, under these conditions,
lead would corrode at an unsuitably high rate. However, if the repository is flooded and
oxygen levels increase, then lead might be acceptable [63]. Mass loss tests, supported by
electrochemical polarisation experiments, were carried out in natural and simulated
groundwater, and in a range of specific salt solutions [64]. Corrosion rates in aerated
conditions varied between 1-3 µm y-1 in groundwater (which is acceptable) to over 600 µm y-1
in sodium acetate and over 380 µm y-1 in sodium nitrate (which is very high and
unacceptable). Corrosion rates in bentonite clay suspensions likely to be used as repository
backfill, were 10-15 µm y-1 and deemed to be acceptable.

X.3.4 Acids

X.3.4.1 Mineral acids


Traditionally, sulphuric acid was made, stored and conveyed in lead; this is because its
corrosion resistance at moderate temperatures and over almost all the concentration range of
the acid is excellent provided that the protective sulphate film is not disrupted. Rupture of the
sulphate film may be caused by erosion as a result of high velocity liquids and gases
containing acid spray. In such an environment acid-resistant brick is often used with a liner of
lead in between the brick and the (usually steel) vessel. Thermal cycling may also disrupt the
film by thermal fatigue. The corrosion rate of lead is generally less than 125 µm y-1 below
about 55% concentration up to its boiling point (130°C) while acid concentrations to 80%
have similar corrosion rates below 100°C. Above an acid concentration of 85% the corrosion
rate of lead increases to unacceptable values [64]. Lead is no longer used for this service in
industry, its place having been taken by passive alloys such as zirconium, specialist stainless
steels and nickel-based alloys [65].

Nitric acid readily attacks lead if dilute and the metal should not even be used for handling
nitrate or nitrite species, except at extreme dilutions and preferably with a passivating reagent
such as a sulphate, which will confer some protection. Corrosion decreases to a minimum at
65–70% HNO3 and lead has been used for storage of nitric acid in the cold at this
concentration [43]. Hydrochloric acid should generally be regarded as aggressive to lead and
its use cannot be recommended, although a satisfactory life has been obtained with acid of up
to 30% concentration at ambient temperature and 20% concentration at 100°C [43].
Resistance of lead to corrosion by HCl is presumably due to the formation of a protective film
of lead chloride which is only slightly soluble at these concentrations combined with its high
overpotential for hydrogen evolution. In mixed hydrochloric and hydrofluoric acids (e.g. used
for pickling steel) the behaviour of lead is unreliable unless the lead is initially passivated in
hydrofluoric acid first.

The remaining use of lead in industrial processes is as an anode in electrowinning and as an


inert anode in electroplating and for these applications there are few economic alternatives.
Lead has good resistance to phosphoric and chromic acids (e.g. in chromium plating) and, at
high current densities, will tend to passivate anyway with the formation of lead dioxide.

X.3.4.2 Organic acids


Lead is attacked by most organic acids, which produce soluble lead salts, particularly in the
presence of air or other organic oxidants. Aqueous acetic acid, solutions containing acetates,
and acetic acid vapour all rapidly corrode lead and should be avoided if possible. During
corrosion, the protective film is dissolved yielding lead salts of the organic acid. These are
susceptible to rapid carbonation in the presence of CO2 and water, forming basic lead
carbonate, which, in these circumstances, does not form a passive film.

Studies of the electrochemical behaviour of lead at 25°C in acetic, lactic (0.01 M-1.0 M),
oxalic and tartaric (0.01 M-0.15 M) acid solutions demonstrated that lead is readily soluble
both in acetic and lactic acid solutions up to 2000 mV [66]. In these acids, anodic dissolution
appears to be under charge transfer control, with lactic acid more aggressive than acetic acid
with lead passivating at higher potentials. However, in oxalic and tartaric acid solutions a
dependence on the acid concentration is evident. Thus, above a certain specific concentration,
an anodic current peak is followed by a reduction in current associated with the formation of a
passivation salt film that consisting of the oxalate or tartrate species, respectively. As noted
above, corrosion by organic acids is important in the construction sector and also in the
conservation of historic artefacts.

X.3.5 Lubrication oils


White metal (tin-lead) bearings do not normally fail due to corrosion, but where this has
occurred it has been associated with the generation of acidity in the lubricant, the production
of peroxides and the presence of air. Peroxides appear to be the controlling factor, but
corrosion is also reduced in the absence of air. The corrosion product generally consists of
basic lead salts of organic acids [67]. The presence of residual organic acids is thought to be
the main cause of the corrosion of terne (lead) plated steel when used as fuel tanks.

X.3.6 Miscellaneous environments


As can be seen from the Pourbaix diagram, lead has no stable passive species at pH > 10-11
and, hence, lead is not particularly resistant to dilute alkalis, and will dissolve freely as the
plumbite oxyanion. Where free access to carbon dioxide is available, then a passivating salt
film of lead carbonate may form. However, lead is susceptible to lime drips from fresh
concrete and cement mortar, which will tend to disrupt the lead carbonate film formation.
Lead can tolerate concentrated alkalis such as KOH to 50% and up to 60°C and NaOH to 30%
and 25°C although it is explicitly not used for this purpose.

Lead is not generally attacked rapidly by solutions that contain anions where the lead salt is
sparingly soluble and, hence, where lead can passivate by the formation of a salt film. Thus
only nitrates and, to a lesser extent chlorides, are corrosive. The presence of nitrate tends to
pit lead, for example in carbonate solution [23]. In sodium chloride the corrosion rate
increases with concentration to a maximum in 0.05M solution, then decreases due to
formation of a relatively porous film PbCl2. Control of the cyclic voltammetry conditions
allowed the development of a relatively thick and more protective layer [68]. In potassium
bromide adherent deposits are formed, and the corrosion rate increases with concentration.
The attack in potassium iodide is slow in concentrations up to 0.1M but in concentrated
solutions rapid attack occurs, probably owing to the formation of soluble KPbI3. In dilute
potassium nitrate solutions (0.001M and below) the corrosion product is yellow and is
probably a mixture of Pb(OH)2 and PbO, which is poorly adherent. At higher concentrations
the corrosion product is more adherent and corrosion is somewhat reduced [69].

X.4 Specific applications


X.4.1 Lead anodes
Anodes for electroplating and for electrolysis of brine are frequently made of lead and lead
alloys. This is because of the formation of a passive film of lead dioxide at high anodic
potentials where conventional passive alloys, such as chromium containing materials, would
be destroyed by transpassive dissolution. Nevertheless, there is generally a very slow
continued corrosion which leads to thickening of the PbO2 film. The resulting stresses caused
by growth of the oxide layer can cause it to crack and disbond, releasing PbO2 particles into
the electrolyte. Alloying elements are frequently added for strength and to stabilise the film;
rolled or extruded alloys are generally found to be more resistant to this form of degradation
than cast alloys.
Lead-silver (1-2% Ag) anodes for cathodic protection may be used in brine and seawater
applications for cathodic protection of ships and dockside structures [71,72] at current
densities of up to 120 A m-2; although in many applications these have now been superseded
with platinum-doped titanium or niobium alloys, which can operate at still higher current
densities.

Lead-silver, lead-tin, lead-calcium-tin, lead-calcium-silver and lead-tin-silver are all used for
electrowinning, which is generally carried out using sulphuric acid-based electrolytes. These
alloys have replaced the more traditional lead-antimony compositions due to requirements of
higher purity in the deposits. In all electrowinning applications it is essential to keep the
potential of the anode well above the PbO2/PbSO4 equilibrium potential otherwise rapid
corrosion will occur [72]. Electrochemical studies of electrowinning anodes have been
extensively studied and are focussed on: (a) reduction of anode corrosion rate, (b) reduction of
(lead and other metal) contamination in the deposit, and (c) improved current efficiency.
For example, the efficiency and corrosion resistance of calcium and silver containing alloys
appears to depend on the nature of PbO2 passive layer and its electronic and ionic
conductivity [73]. The corrosion rate and current efficiency of anodes can be improved
significantly by the incorporation of minor amounts of foreign species in the electrolyte. Thus,
preconditioning of anodes in a fluoride solution prior to use can result in a compact and more
protective PbF2/PbO2 bilayer that is more corrosion resistant [74].

The corrosion rates for lead-silver anode corrosion rates in zinc electrowinning solutions were
studied at a range of current densities from 2500-10000 A m-2. Increases in acid concentration
and temperature caused increases in corrosion rate whereas, in the absence of bath additions,
the rate was independent of current density [75]. Chloride ions increased the corrosion rate
while fluoride preconditioning reduced it but only in the presence of manganese. The
presence of small concentrations of cobalt and manganese ions can reduce the corrosion rate
of electrowinning anodes considerably. The effect appears to be related to an increase in the
oxygen evolution kinetics that reduces cell overvoltage and causes less disruption of the PbO2
passive layer [76].

Traditional chromium plating uses lead-tin or lead-antimony alloys as the anode material
although coated passive metal anodes (e.g. platinised Ti, Zr, Nb and Ta) have recently been
advocated [77]. Lead anodes offer processing advantages in that they can be formed or cast to
conform to the surface to be plated, hence giving an even current density and uniform
coverage of coating. The electrochemistry of anodes containing tin, antimony and silver have
been studied in electrolytes similar to chrome plating baths. In chromic acid a passive film of
PbCrO4 forms at very low overpotentials and the subsequent formation of PbO2 is masked by
oxygen evolution. The effects of alloying additions were found to be minor although
antimony raises the passive current substantially and silver reduces the overpotential of
oxygen evolution by about 0.2 V [78]. As for electrowinning anodes, the quiescent (non-
polarised) corrosion rate can be very high and it is recommended that anodes are connected at
all times. It has been found that 0.5 g/l of magnesium fluosilicate suppresses corrosion
without affecting the plating process [79].

X.4.2 Lead-acid battery


Lead-acid batteries comprise by far the single most important worldwide application for lead.
The technology of lead acid batteries is still under constant improvement and is technically
still very much relevant [80]; this section can only touch upon very general aspects of the
technology. Lead-acid batteries typically consist of lead alloy supports which carry an
electrochemically active mass, the composition of which differs between positive and
negative plates, and with the state of charge of the battery. Failure normally occurs in the
positive grids of a battery. The main cause of failure is loss of contact between the grid and
the active mass due to ‘grid growth’ which is caused by the change in volume of the active
material during the charge/discharge cycle, and by corrosion of the metal surface, which can
be accelerated by stress. Batteries for automotive, electric vehicle, standby services, etc, all
have different characteristics and requirements, which are met by battery grid design and
choice of alloy [81,82].

Traditionally battery grids have been made from lead with 6-14% antimony with a small
amount of arsenic. High antimony alloys have a significant electrochemical disadvantage,
which is that it reduces the overpotential for hydrogen evolution on the lead leading to
electrolysis of the electrolyte (to hydrogen and oxygen). Thus, batteries using traditional
materials require regular maintenance (topping-up) via additions of distilled water. While
alloys in the region of 5-6% antimony are still used in some industrial, deep discharge and
traction applications, high antimony contents have been largely replaced in automotive
batteries by complex low antimony or antimony-free lead-calcium-(tin) alloys. Modern
automotive batteries now use lead-calcium based alloys that do not require topping up
(provided they are operated within their electrical design parameters). This reduction in
antimony content has been made possible by the introduction of additions of solidification
nucleants, such as selenium and sulphur, which promote fine-grain structure. It is also
increasingly common to find different alloys used in the positive and negative grids and
optimised for these applications.

High antimony alloys exhibit high strength, good castability and give good deep cycling
performance. The latter requires that the active mass has good adhesion to the metal, is
structurally stable during cycling and does not passivate. Antimony reduces shedding of
active material from the cells, produces a surface film of greater porosity which becomes
more porous during cycling, promotes stability of the active mass. It has also been shown that
PbSO4 is more reluctant to nucleate on antimonial lead [83]. Although corrosion rates may
appear quite high, attack is normally of a general nature which allows a satisfactory service
life. This is because the eutectic is preferentially corroded, which reduces intergranular
corrosion. Antimony reduces the oxygen overpotential on the positive grid while Sb5+ ions
migrate from the positive grid to the negative and be reduced to metallic antimony [84]. This
reduces the hydrogen overpotential, leading to excessive gassing, consuming water from the
electrolyte, reducing charge efficiency and liberating stibine (SbH3). During overcharge,
antimony increases the rate of formation of the inner corrosion layer on the positive grid.

Low maintenance batteries, which only require the addition of water infrequently in the
second half of their service life, use low antimony alloys that typically contain less than 3%
antimony, with some alloys containing as little as 0.6%. The most commonly used alloys are
1.3–1.8% Sb. They always contain As to assist hardening, and a nucleating agent such as Se
or S with Cu. These are necessary because the coarse dendritic structure is prone to porosity
and hot cracking during casting. The addition of nucleating agents gives a fine grained
structure with good corrosion resistance. Tin is often added to increase fluidity in casting
alloys.

Lead-calcium-(tin) alloys are used in maintenance-free automotive starting lighting, ignition


(SLI) batteries, in stationary batteries and some traction batteries. It is essential that the
correct calcium content and a suitable calcium-tin ratio is used. In the binary lead-calcium
alloys, a fine grained structure with serrated grain boundaries is produced by a discontinuous
precipitation reaction [85]. The addition of tin changes the nature of the precipitation
reactions to give two areas of stability. One is with high calcium-low tin and the other is in
the region below 1.8% tin and less than 0.07% calcium. Batteries made from these alloys have
a much reduced rate of self-discharge compared with antimonial alloys, thus giving a longer
shelf-life, and maintain a high discharge voltage throughout their life.

X.4.3 Reactor coolants


Lead-based liquid metal coolants (using either liquid lead or liquid lead-bismuth eutectic)
were first used in a nuclear reactor in the 1960’s by the Soviets as an advanced, high power
density, submarine propulsion plant. Reactors designed using such coolants have some
compelling advantages (e.g. operation at atmospheric pressure, boiling point greatly in excess
of the reactor operating temperature, relative non-reactivity to water and air compared with
liquid sodium) and comprise a candidate “Generation IV” fast neutron reactor system as well
as a candidate coolant system for fusion reactors. However, lead alloys are significantly more
corrosive to constructional materials (e.g. steels) then liquid sodium. This is due to dissolved
oxygen in the lead that reacts with the containing material, as well as liquid metal
embrittlement [86,87]. To avoid excessive oxidation of structural alloys, such as martensitic
steels, it is necessary to control the oxygen activity to below 10-24 atm, and this can be
monitored by the use of electrochemical probes [88].

Bibliography
International Lead-Zinc Research Organisation, website at: www.ilzro.org
International Lead Association, website at: www.ila-lead.org:
“Lead and Its Alloys”, D.R. Blasket and D. Boxall pub. Ellis Horwood (1990)
“Engineering Properties and Applications of Lead Alloys”, S. Guruswamy, CRC Press (1999)
“The Electrochemistry of Lead”, ed. A.T. Kuhn, Academic Press, London (1979)

References
[1] “Lead – the Facts”, I. Thornton, R. Rautiu and S. Brush, pub. International Lead
Association (2001)
[2] “Lead and Its Alloys”, D.R. Blasket and D. Boxall pub. Ellis Horwood (1990)
[3] “Engineering Properties and Applications of Lead Alloys”, S. Guruswamy, CRC Press
(1999)
[4] “Smithell’s Metal Handbook”, 7th ed. Chapter XXXX, ???? (???)
[5] “Lead alloys”, D.R. Prengaman, in: “Ullman’s Encyclopaedia of Industrial Chemistry”,
pub. Wiley (2005)
[6] “Consumption of lead-silver alloy anodes in sulphuric acid” F. Hine, Y. Ogata, M.
Yasuda, Bulletin of Electrochemistry, Vol.4, p.61-65 (1988)
[7] “Structure control of non-antimonial lead alloys via alloy additions, heat treatment and
cold working”, R.E. Prengaman, Proceedings 7th International Lead Conference,
Madrid, pub. Lead Development Association (1983)
[8] “Development of improved lead materials for chemical plant”, U. Heubner and M.
Reinert, Proceedings 7th International Lead Conference, Madrid, pub. Lead
Development Association, London (1983)
[9] “Potential-pH diagram of lead and its applications to the study of lead corrosion and to
the lead storage battery”, P. Delahay, M. Pourbaix, and P. Van Rysselbergh, Journal of
the Electrochemical Society, Vol.98, p.57 (1951)
[10] “Modern Electrochemistry – Vol.2 ”, J. O’M Bockris and A.K. Reddy, Table 10.12,
pub. Plenum (1986)
[11] “Lead anodes in alkaline solution”, V.I. Birss and M.T. Sheralier, Journal of the
Electrochemical Society, Vol.134, p.802-808 and p.1594-1600 (1987)
[12] “Standard electromotive force of the lead electrode”, J.J Lingane, Journal of the
American Chemical Society, Vol.60 p.724-5 (1938)
[13] “Hysteresis effects in the hydrogen evolution reaction at lead in aqueous perchloric
acid”, D.J.G. Ives, F.R. Smith, Transactions of the Faraday Society, Vol.63(1), p.217-
33 (1967)
[14] “Cathodic lead disintegration and hydride formation”, H.W. Salzberg, Journal of the
Electrochemical Society, Vol.100 p.146-51 (1953)
[15] “The Electrochemistry of Lead”, ed. A.T. Kuhn, pub. Academic Press, London (1979)
[16] “The growth of thin PbO layers on lead films, I: Experiment and II: Theory”, J.M.
Eldridge and D.W. Dong, Surface Science, Vol.40, p.512-530 and p.531-544 (1973)
[17] “Electrochemical and spectroscopic methods of characterising lead corrosion films”,
K.R. Bullock, Journal of Electroanalytical Chemistry, Vol.222, p.347-366 (1987)
[18] “Ion selectivity and diffusion potentials in corrosion layers of PbSO4 films on Pb in
H2SO4”, P. Ruetschi, Journal of the Electrochemical Society, Vol.120, p.331-336 (1973)
[19] “Processes in solid state at anodic oxidation of a lead electrode in H2SO4 solution and
their dependence on the oxide structure and properties”, D. Pavlov, Electrochimica
Acta, Vol.23, p.845-854 (1978)
[20] “Dependence of the phase composition of the anodic layer on oxygen evolution and
anodic corrosion of lead electrode in lead dioxide potential region”, D. Pavlov, T.
Rogachev, Electrochimica Acta, Vol.23, p.1237-1242 (1978)
[21] “Photoelectrochemistry in the lead-sulphuric acid system”, S. Fletcher, O.B. Matthews,
Journal of Electroanalytical Chemistry, Vol.126, p.131-144 (1981)
[22] “A conductive film model for the lead anode in sulphuric acid “, S.B. Hall and G.A.
Wright, Corrosion Science, Vol.31, p.709-714 (1990)
[23] “Anodic behaviour and passivation of a lead electrode in sodium carbonate solutions”,
E.E. Abd El Aal, S. Abd El Wanees, A. Abd El Aal, Journal of Materials Science,
Vol.28, p.2607-2614 (1993)
[24] “Anodic dissolution of lead in perchloric acid solutions: the effect of sulphuric acid”, E.
Ahlberg and B. Berghult, Electrochimica Acta, Vol.36, p.197-201 (1991)
[25] “Studies on the pitting corrosion of lead in carbonate media”, E.E. Abd El Aal, Anti-
Corrosion Methods and Materials, Vol.48, p.116-125 (2001)
[26] “Pitting corrosion of lead in sodium carbonate solutions containing NO3- ions”, M.A.
Amin, S.S.A. Rehim, Electrochimica Acta, Vol.49, p.2415-2424 (2003)
[27] Solubility products of lead compounds??? CRC Handbook of Chemistry and Physics
[28] “Lead/carbon steel galvanic corrosion evaluation”, C.J. Semino, A.L. Burkart, M.E.
García, R. Cassibba, Journal of Nuclear Materials, Vol.238, p.198-204 (1996)
[29] “Water contamination: Impact of tin-lead solder”, K.S. Subramanian, V.S Sastri, M.
Elboujdaini, J.W. Connor, A.B.C. Davey, Water Research, Vol.29, p.1827-1836 (1995)
[30] “Inhibition of release of lead into water owing to galvanic corrosion of lead solders”,
V.S. Sastri, K.S. Subramanian, M. Elboujdaini, J.R. Perumareddi, Corrosion
Engineering Science and Technology, Vol.41, p.249-254 (2006)
[31] “Chloride-to-sulphate mass ratio and lead leaching in water”, M. Edwards, S.
Triantafyllidou, Journal of the American Water Works Association, Vol.99, p.96-109
(2007)
[32] “The basic lead carbonates”, J.K. Olby, Journal of Inorganic and Nuclear Chemistry,
Vol.28, p.2507 (1966)
[33] "Patination of lead: An infra-red spectroscopic study”, G.C. Tranter, British Corrosion
Journal, Vol.11, p.222 (1976)
[34] “ Chemical mechanisms for the atmospheric corrosion of lead”, T.E. Graedel, Journal of
the Electrochemical Society, Vol.141, p.922-927 (1994)
[35] “Atmospheric corrosion of lead and Its alloys”, A.R. Cook, and R. Smith, in:
“Atmospheric Corrosion”, ed. W.H. Ailor, pub. Wiley, New York (1982)
[36] “Corrosion of aluminium in contact with lead in atmospheric environments”, R.H. Hill,
P.C. Frost, and R. Smith, Proceedings 7th International Lead Conference, pub. Lead
Development Association, London (1983)
[37] “Various aspects of weathering and corrosion of lead in building applications”, R.H.
Hill, P.C. Frost and R. Smith, Proceedings 8th International Lead Conference, Lead
Development Association, London (1985)
[38] “Precipitation runoff from lead”, S.A. Matthes, S.D. Cramer, B.S. Covino Jr., S.J.
Bullard, and G.R. Holcomb in “Outdoor Atmospheric Corrosion”, ed. H.E. Townsend,
ASTM Special Technical Publication Vol.1421, p.265-274 (2002)
[39] “Carboxilic acids in the atmosphere and their effect on the degradation of metals”, A.V.
Echavarría, F.E. Echeverría, C. Arroyave, E. Cano, J.M. Bastidas, Corrosion Reviews
Vol.21, p.395-413 (2003)
[40] “The influence of relative humidity and temperature on the acetic acid vapour-induced
atmospheric corrosion of lead”, A. Niklasson, L.-G. Johansson, J.-E. Svensson,
Corrosion Science, Vol.50, p.3031-3037 (2008)
[41] “Studies of lead corrosion in acetic acid environments”, J. Tétreault, J. Sirois, E.
Stamatopoulou, Studies in Conservation, Vol.43, p.17-32 (1998)
[42] “Inhibition treatment of the corrosion of lead artefacts in atmospheric conditions and by
acetic acid vapour: Use of sodium decanoate”, E. Rocca, C. Rapin, F. Mirambet,
Corrosion Science, Vol.46, p.653-665 (2004)
[43] “Lead and Lead Alloys Properties and Technology”, W. Hofmann, Springer-Verlag,
English Translation by Lead Development Association, London (1970)
[44] “Rolled Lead Sheet - The Complete Manual”, pub. Lead Sheet Association (2007)
[45] “Reduction of Hazardous Substances”, European Commission Directive 2002/95/EC
[46] “Action of natural waters on lead”, G. Miles, Journal of the Society of Chemical
Industry, Vol.67 p.10-13 (1948)
[47] “Corrosion and passivation of lead in aqueous solutions”, W.A. Badawy, F.M. Al-
Kharafi, Corrosion Prevention and Control, Vol.46, p.13-22 (1999)
[48] “Control of lead corrosion by chemical treatment”, B.P. Boffardi, and A.M. Sherbondi,
Paper 445, Proceedings “Corrosion 1991”, NACE, USA (1991)
[49] “Control of lead corrosion”, J.W. Patterson, J.E. O'Brien, Journal of the American
Water Works Association, Vol.71, p.264-271 (1979)
[50] “Effect of phosphate inhibitors on lead release from pipes”, M. Edwards, L.S. McNeill,
Journal of the American Water Works Association, Vol.94, p.79-90 (2002)
[51] “Investigation on lead corrosion products in sea water and in neutral saline solutions”,
A.M. Beccaria, E.D. Mor, G. Bruno, G. Poggi, Werkstoffe und Korrosion, Vol.33,
p.416–420 (1982)
[52] “Corrosion of lead in seawater”, A.M. Beccaria, E.D. Mor, G. Bruno, G. Poggi, British
Corrosion Journal, Vol.17(2), p. 87-91 (1982)
[53] “Protection of cables and conduits by isolating layers against stray currents”, F. Glander,
W. Glander, Zeitschrift fuer Metallkunde, Vol.44 p.97-101 (1953)
[54] “The influence of superimposed a.c. on the anodic corrosion of lead in aqueous
sulphuric acid”, J.M. Costa and T.P. Hoar, Corrosion Science, Vol.2, p.269-274 (1962)
[55] “New developments in lead-sheathed cables”, J. Dyba and F. Goodwin, Proceedings
IEEE International Symposium on Electrical Insulation, USA, (1998)
[56] “Underground Corrosion”, M. Romanoff, National Bureau of Standards, publication
#579, (1957)
[57] “Some experiments in the mechanism of corrosion of lead pipes in soils”, W.W.
Robson, and A.R. Taylor, Report MM/19/54, Associated Lead Manufacturers Ltd.
(1954)
[58] “Electrochemistry of lead in simulated groundwater environments”, E.A. Joerg and O.F.
Devereux, Corrosion, Vol.52, p.953-957 (1996)
[59] “Formation of lead monoxide as a cable-sheath corrosion product”, E.F. Wolf, C.F.
Bonilla, Transactions of the Electrochemical Society, Vol.79, p.307 (1941)
[60] “Potential criteria for the cathodic protection of lead cable sheath”, K.G. Compton,
Corrosion, Vol.12 p.553-60 (1956)
[61] “Pit corrosion and surface-film formation on lead”, E.L. Schmeling, B. Roschenbleck,
Werkstoffe und Korrosion, Vol.12 p.215-23 (1961)
[62] “Underground biodeterioration of medium tension electric cables”, J.L. Pintado and F.
Montero, International Biodeterioration & Biodegradation, Vol.29, p.357-365 (1992)
[63] “Corrosion resistance of lead alloys under nuclear waste repository conditions”, F.E.
Goodwin, Corrosion Prevention and Control, Vol.32(2), p.21-24 (1985)
[64] “Lead corrosion behaviour in sited media of an underground repository”, R.O. Cassibba
and S. Fernandez, Journal of Nuclear Materials, Vol.161, p.93-101 (1989)
[63] “Corrosion resistance of lead alloys under nuclear waste repository conditions”, F.E.
Goodwin, Corrosion Prevention and Control, Vol.32(2), p.21-24 (1985)
[64] “Corrosion Engineering”, M.G. Fontana, …. (XXXX)
[65] “Corrosion in the Petroleum Industry”, ed. L. Gaverick, pub. ASM International (1994)
[66] “Electrochemical behaviour and corrosion of lead in some carboxylic acid solutions”,
S.S. Abd-El Rehim, N.H. Amin, L.I. Ali, N.F. Mohamed, Journal of Chemical
Technology and Biotechnology, Vol.72 p.197-201 (1998)
[67] “The role of peroxides in the corrosion of lead by lubricating oils”, B.S. Wilson, F.H.
Garner, Journal of the Institute of Petroleum, Vol.37, p.225-38 (1951)
[68] “Cyclic voltammetric behaviour and some surface characteristics of the lead electrode
in aqueous NaCl solutions”, A.M.A. El-Halim, M.H. Fawzy, A. Saty, Journal of
Electroanalytical Chemistry, Vol.316, p.275-292 (1991)
[69] “Colloid-chemical phenomena on surfaces of metals and retardation of corrosion in salt
solutions, IX: Corrosion of lead in neutral solutions of potassium salts”, A. Vaivads, L.
Liepina, Latvijas PSR Zinatnu Akademijas Vestis, Vol.8, p.119-29 (1954)
[70] “Cathodic Protection, Theory and Practice”, L.L. Shreir and P.C.S. Hayfield, in V.
Ashworth and C.J.L. Booker, pub. Ellis Horwood (1986)
[71] “Service experience with lead-silver alloy anodes in cathodic protection of ships”, K.N.
Barnard, G.L. Christie, and D.G. Gage, Corrosion, Vol.15, p.581-586 (1959)
[72] “The anodic corrosion of lead in sulphuric acid solutions”, J.J. Lander, Journal of the
Electrochemical Society, Vol.103 p.1-8 (1956)
[73] “Corrosion properties of lead anodes: 1. Calcium- and silver-bearing lead alloys” and
“… 2. Lead alloys doped with silver, titanium, tin, cobalt, manganese, and silicon”, L.F.
Kozin, V.F. Kozin, Protection of Metals, Vol.33, p.131-136 and 549-555 (1997)
[74] “Preconditioning of lead-silver alloy anodes for use in electrowinning of metals”, P.
Ramachandran, K. Balakrishnan, Bulletin of Electrochemistry, Vol.12, p.352-354 (1996)
[75] “Corrosion rates of lead based anodes for zinc electrowinning at high current densities”,
R.H. Newnham, Journal of Applied Electrochemistry, Vol.22, p.116-124 (1992)
[76] “Influence of Co2+ and Mn2+ ions on the kinetics of lead anodes for zinc
electrowinning”, C. Cachet, C. Le Pape-Rérolle, R. Wiart, Journal of Applied
Electrochemistry, Vol.29, p.813-820 (1999)
[77] “On the use of platinized and activated titanium anodes in some electrodeposition
processes”, M.G. Pavlović, A. Dekanski, Journal of Solid State Electrochemistry, Vol.1,
p.208-214 (1997)
[78] “Filming behaviour of lead anodes for chromium electroplating”, M.J.P. McBurney,
D.R. Gabe, Surface Technology, Vol.9, p.253-266 (1979)
[79] “The resistance of lead to corrosion in chromium-plating solutions”, V.E. Carter, H.S.
Campbell, Metal Finishing Journal, Vol.8, p.103-7 (1962)
[80] “Battery Technology Handbook”, ed. H.A. Kiehne, pub. CRC Press (2003)
[81] “A look back at forty years of lead-acid-battery development: A survey especially
regarding stationary applications”, D. Berndt, INTELEC, International
Telecommunications Energy Conference (Proceedings), pp.269-275 (2005)
[82] “Advanced battery systems - The end of the lead-acid battery?”, J. Garche, Physical
Chemistry Chemical Physics, Vol.3, pp.356-367 (2001)
[83] “The cycle life of various lead alloys in 5M H2SO4” and “Electro-reduction processes
of lead and lead alloys in 5M sulphuric acid”, S. Webster, P.J. Mitchell, N.A. Hampson,
J.I. Dyson, Journal of the Electrochemical Society, Vol.133 p.133-137 and p.137-139
(1986)
[84] J.L. Dawson, J. Wilkinson and M.I. Gillibrand, in “Power Sources 3” Proceedings of
the 7th International Symposium in Non-Mechanical Electrical Power Sources,
Brighton, UK, ed. D.H. Collins, Oriel Press, Newcastle-upon-Tyne, p.1-9 (1970)
[85] “Lead-calcium alloy development: Quality improvement”, J.L. Caillerie, L. Albert,
Journal of Power Sources, Vol.67, p.279-281 (1997)
[86] “Corrosion and compatibility considerations of liquid metals for fusion reactor
applications”, P.F. Tortorelli, O.K. Chopra, Journal of Nuclear Materials, Vol.103 p.
621-632 (1981)
[87] “Compatibility tests on steels in molten lead and lead-bismuth”, C. Fazio, G. Benamati,
C. Martini, G. Palombarini, Journal of Nuclear Materials, Vol.296, p.243-248 (2001)
[88] “Active control of oxygen in molten lead-bismuth eutectic systems to prevent steel
corrosion and coolant contamination”, N. Li, Journal of Nuclear Materials, Vol.300, p.
73-81 (2002)

View publication stats

Você também pode gostar