Vous êtes sur la page 1sur 38

This article was downloaded by: [University of Arizona]

On: 21 July 2012, At: 05:39


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

C R C Critical Reviews in Food Science and Nutrition


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/bfsn19

Internal corrosion and shelf‐life of food cans and


methods of evaluation
a b c
Chaim Mannheim , Nehama Passy & Aaron L. Brody
a
Professor of Food Technology, Department of Food Engineering and Biotechnology, Technion
— Israel Institute of Technology, Haifa, Israel
b
Senior Research Engineer, Packaging Laboratory, Department of Food Engineering and
Biotechnology, Technion, Israel Institute of Technology, Haifa, Israel
c
Manager, Market Development, Container Corporation of America, Oaks, Pennsylvania

Version of record first published: 29 Sep 2009

To cite this article: Chaim Mannheim, Nehama Passy & Aaron L. Brody (1983): Internal corrosion and shelf‐life of food cans
and methods of evaluation, C R C Critical Reviews in Food Science and Nutrition, 17:4, 371-407

To link to this article: http://dx.doi.org/10.1080/10408398209527354

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to
anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should
be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims,
proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in
connection with or arising out of the use of this material.
Volume 17, Issue 4 371

INTERNAL CORROSION AND SHELF-LIFE OF FOOD CANS AND


METHODS OF EVALUATION
Authors: Chaim Mannheim
Nehama Passy
Department of Food Engineering and
Biotechnology
Technion, Israel Institute of Technology
Haifa, Israel

Referee: Aaron L. Brody


Market Development
Container Corporation of America
Oaks, Pennsylvania
Downloaded by [University of Arizona] at 05:39 21 July 2012

I. INTRODUCTION
The term corrosion is conventionally applied to formation of rust on, or oxidation of, a
metal surface. Internal corrosion in food cans is characterized by dissolution of the
container metal in its contents. The ultimate effect of this process is to render the product
unmarketable because of organoleptic changes, loss of vacuum, hydrogen swelling, or
perforation of the container.
While corrosion failures in tinplate containers cannot be expected to disappear,
improved knowledge of the mechanisms involved and development of better materials
and coatings (both interior and exterior) already have alleviated the problem
significantly, and further improvement can be expected with time.
Selection of the most suitable tinplate for a given food product often may present a
problem. While the properties of the tinplate are defined and measurable, the choice may
be affected by the aggressiveness of the product, by technological parameters of the
canning process, by storage conditions, by admissible limits of metal dissolution, and by
considerations of preservation of flavor and color. Following a review of the two main
available types of cans (plain and lacquered) and their methods of manufacture,
corrosion mechanisms and the probable interaction between metal, coating, and medium
will be discussed. Foods are classified, according to their aggressiveness, as tin and iron
dissolvers, and this classification will be presented. Methods of shelf-life measurement,
and selection recommendations for given classes of foods, will be reviewed, as well as
current and future trends in the tinplate industry.

II. THE STRUCTURE OF TINPLATE FOR FOOD CANS

Tinplate has a multilayered structure, as can be seen schematically in Figure 1. The


basic material is steel plate coated on both sides with a tin layer. In addition, there are
layers of tin-iron alloy (FeSn2), formed during the flow-brightening process, and
passivation films containing tin oxides, metallic chromium, and chromium oxides.

A. Steel Base
Three types of steel (designated L, MC, and MR), differing in their chemical
composition, are used for can manufacture. Type L has a low and controlled content of
residual elements such as phosphorus, silicon, copper, nickel, chromium, and
molybdenum, and is suitable for highly corrosive products. Type MC is a rephosphorized
372 CRC Critical Reviews in Food Science and Nutrition

thickness
ooooooooo Oil film o o o o o o o o o o —*• 4lTHi

'#&£%&passivation layer %%liHWz. — l-2mji


300
EE5-3~ alloy layer
HHir
-~ r

"'
80m
H

steel '^Z^ZZZZZ -"- 0.2mm

—_-_—.——I~T alloy layer ——:—.—.-r ~

tin layer

layer
Downloaded by [University of Arizona] at 05:39 21 July 2012

OOOOOOOOOO ji Xjl— OOOOOOOOOO

FIGURE I. Tinplate multilayered structure, a schematic cross section.

steel with high strength but low resistance to corrosion. Type MR — not restricted in
residual elements to the same degree as Type L, except for phosphorus which is kept at a
low level — is the most commonly used plate for moderately corrosive products.
Other types of steel, such as K plate or Special Test Plate (STP), should also be
mentioned. These plates are highly resistant to corrosion and suitable where a long shelf-
life of an acid product is required.
The thickness of the steel sheet used in can manufacture depends on the size of the can
and normally ranges from 0.20 mm for small sizes to 0.29 mm for the institutional sizes.
In some cases, still smaller or still larger gauges are also used.
The industry used to designate both steel thickness and tin coating in terms of "base
box" where one base box used to be 112 sheets of 14 X 20 in., i.e., on an area of 31,360 in.2
(20.232 m2). Each thickness is given as the weight of steel of one base box and tin coating
as the amount of tin in pounds required to coat one base box on both faces (i.e., to cover
an area of 62,720 in.2). In recent years, coatings are usually given in gauge thickness for the
steel plate in grams per square meter for tin coatings.

B. Tin Coatings
Tin coatings are applied to the steel in a chemical or an electrolytic process. The
chemical process in which the steel sheet is immersed in a molten tin bath and "hot
dipped" tinplate is produced, has fallen into disfavor due to its high cost and the
uneven and superfluous thickness of the tin layer (14 to 16 g/m2 on each side) due to
control difficulties.1
In the now prevalent process of electrolytic deposition, basic operations include
pickling, electroplating, flow brightening, passivation, and oiling. The coating obtained
in this fashion is more homogeneous but more porous and less corrosion-resistant
compared to the "hot-dipped" tinplate. The various electroplating techniques differ in
the processes of metal pretreatment, and in type of electrolyte. Electrolytic tinplate with
unequal coating weights on each face (differential tinplate) is very common in the food
canning industry. Its main advantage is the possibility of providing a higher coating
weight on the more aggressive side (usually the inside of the can, except when lacquer is
Volume 17, Issue 4 373

used). This enables a significant saving in tin consumption and decrease of packaging
cost. The range of tin coating weights most common in food cans is 2.8 to 11.2 g/m2.

C. Tin-Iron Alloy
The tin-iron alloy (FeSn2) is produced by diffusion of tin into the steel sheet during the
flow-brightening process of electrolytic tinplate and its thickness is 10 to 30% of the tin
layer. This protective barrier layer plays an important role in corrosion resistance, and
the more compact and continuous the layer, the higher the protection obtained.2"6
Ideally, the compound should exist as a thin, densely packed layer which completely
covers the steel base. If during the flow-brightening process a uniform close-packed alloy
layer is obtained (regardless if the tin is deposited from an acid or alkaline plating bath)
the plate is considered to have K plate qualities having high corrosion resistance.4"6 Its
properties are tested by the Alloy-Tin Couple Test (ATC) which will be discussed later.
When the alloy layer is not continuous, the steel plate is exposed through the pores and
corrosion is accelerated. Electrochemically, the layer is cathodic to tin and iron. Several
authors report on the values of its potential and on the influence of area ratios (tin/iron)
Downloaded by [University of Arizona] at 05:39 21 July 2012

on the couple behavior.7"9 However, it was found that the alloy is easily polarized, thus
contributing little to the galvanic detinning reaction.8 In most cases tin is dissolved down
to the alloy layer rather than to the iron.9 Since the tin and alloy layer are connected
physically, the effective potential is the coupling potential of Sn+FeSn2. The direction of
this potential will determine whether tin alone or both tin and iron will go into solution.9

D. Tin Free Steel (TFS)


The high price of tin, and political problems in the tin-producing countries, forced the
industry to look for alternative materials.10 Chromium and chromium oxide were found
to be suitable coatings, and early developments in tin-free steel were centered in Japan
and have attracted worldwide interest.
Processes and products based on coating of a single film of chromium metal or a
duplex film of chromium and chromium oxide were developed. The protective films are
about 0.05 to 0.1 jum in thickness, equivalent to 500 to 1000 mg/dm2. This type of steel
must always be used with lacquer coatings, and epoxy-phenolic and phenolic-based
enamels have proved satisfactory coatings. The degree of adhesion obtained with TFS is
very high, allowing fabrication and processing operations to be carried out with excellent
lacquer adhesion properties and high corrosion resistance.11'12
Since TFS cannot be soldered, cementing and welding are called for with round can
bodies. In these circumstances, TFS is presently limited to open top ends on tinplate
bodies, drawn two-piece cans for fish products, beer, beverage containers, and aerosols.
TFS is unsuitable for products with high acidity, whether artificial or natural.13 They
cause filiform corrosion under enamel films, blistering of enamel and undermining of
chromium layers.

£. Tin Oxide and Passivation Layers


During production at the tin mill or during warehousing, tin oxide films may form on
the tinplate. These oxides have long been associated with adhesion failures of organic
coatings, yellow surface discoloration and soldering difficulties especially in "hot-
dipped" plate. To prevent uncontrolled oxidation of the tin, and formation of sulfide
stains produced by certain foods, and to afford some protection against corrosion during
storage as well as improved lacquer adhesion, the tinplate is subjected to passivation
treatments.14'15
Passivity is a phenomenon shown by a metal which is able to withstand corrosion or to
corrode at a low rate, in media where it is thermodynamically unstable.16 Three main
types of processes are used for passivation of tinplate: cathodic sodium dichromate
374 CRC Critical Reviews in Food Science and Nutrition

(CDC), which yields a highly effective film against formation of tin oxides; sodium
dichromate dip (SDCD), which yields moderate protection against oxide formation with
limited storage stability; and cathodic sodium carbonate (CSC), the least effective of the
three.14"18
These processes produce a colorless surface film (a few nanometers thick) of
nonstoichiometric oxide. Depending on type and intensity of the treatment, different
amounts of Sn, SnC>2, and different compounds of trivalent chromium oxide and
chromium metal have been identified. A great deal of work has been carried out on a
number of passivation treatments, and various workers have offered different views
concerning the film composition as regards the different stages of chromium
oxidation.14'15'19"31 Passivation produces a surface layer which is not only resistant to
oxidation but also to staining by sulfur containing products and offers improved lacquer
adhesion.25'3'"33

F. Oil Film
The final treatment of the tinplate is lubrication. The lubricant used is either a
Downloaded by [University of Arizona] at 05:39 21 July 2012

vegetable (cottonseed or palm) oil or a synthetic, commonly dioctyl sebacate. The


lubricant film is usually monomolecular in thickness and its weight ranges between 3 to
10 mg/ m2 on each side. It minimizes damage by abrasion and facilitates handling of the
tinplate during can manufacturing.34'35 On prolonged storage the lubricant may undergo
changes through evaporation, polymerization, or soap formation; these changes may
cause difficulties in the lacquer application process due to nonuniform wetting of the
metal surface.

G. Soldering and Welding


Production of a soldered joint comprises the following basic steps: shaping of the metal
parts to fit together; cleaning and degreasing of the joining parts; application of soldering
flux to assist in wetting the surface; application of the solder itself; and removal of excess
solder and cooling of the joint.
The action of the flux consists in removal of the protective tin oxide layer enabling the
molten solder to alloy with the tin coating, spreading rapidly and penetrating the folds of
the can seams. There are many kinds of fluxes, ranging from the highly active zinc
chloride to noncorrosive resin-type compounds.36'37
The most common tin-lead alloys used for soldering tin cans are 2/98 (98% lead) and
40/60 (40% tin). Because of the high cost and relatively high melting point of tin, it is
customary to use solder with minimal tin content.
Due to proposed and actual rulings to reduce lead content in food, it has been
customary, in recent years, to use pure tin for soldering baby food cans. As an alternative
to soldering, welded three-piece cans have taken rapid strides, especially during the past
decade.38'39 This eliminates the lead problem and permits reduction of the sideseam
thickness by 50 to 60%.
Two welding techniques are available, the "rolling electrode" — a high-speed, low-cost
technology, and wire welding. The basic parameters controlling the weld are the current,
pressure (ensuring contact between overlapping edges), and speed. The microstructure of
the weld depends on the heat generated at each part of the interface between the tinplate
surfaces. Welded cans are used for aerosols, beverages, as well as for acid products in
large sizes and more recently also in regular sizes.

H. Organic Can Coatings


The primary function of interior can coatings are prevention of interaction between the
can and its contents. The advent of lacquers has permitted extension of the use of cans to
additional products, including highly aggressive ones. Coating materials are essentially
Volume 17, Issue 4 375

solutions or dispersions of organic resinous matter (natural or synthetic, with or without


pigments) in solvents. The solvent acts as a vehicle which carries the resin and pigments
and ensures a smooth layer during application on the metal surface. Accurate control of
dry film weight, in conjunction with correct stoving (i.e., evaporation of the solvent,
accompanied by a change in chemical structure) is essential for evolution of the intended
properties of the coating as regards chemical and physical resistance.
There are several essential requirements needed from an interior coating. It has to act
as an inert barrier, separating the container from its contents, and not to impart any off-
taste to the food. The coating materials must resist physical deformation through the
bending, formation, or stamping processes and still provide the required chemical
resistance. The lacquer must be flexible, spread evenly, completely cover the substrate,
and adhere to the metal surface. Adhesion failure may occur during, or as a result of,
mechanical deformation during heat processing or undermining by corrosion.9'34'40"44
The interior coating has to withstand sterilization temperatures, and action of acids, as
well as sulfide staining. During the retorting process, protein breakdown occurs with
liberation of sulfides which react with the tin to form black stains, or accumulate in the
Downloaded by [University of Arizona] at 05:39 21 July 2012

headspace and give out an unpleasant smell.45 To overcome this problem, lacquers
pigmented with zinc oxide, which react with the sulfur compounds to form white zinc
sulfide (sulfur-absorbing lacquers), are used. Another type are the sulfur-resisting
lacquers. Enamels are colored with aluminum powder or some other white pigment, and
the black stain is obscurred. These are suitable for solid packs of meat or fish. In the latter
case, waxes (so-called meat release agents) are incorporated in the lacquer to facilitate
removal of the contents.
Soft drinks, and especially beer, are very delicate products and their flavor and clarity
are easily affected if in contact with tin, iron, or an unsuitable varnish. The taste of beer,
for example, is adversely affected by dissolved iron above 0.1 ppm. It is customary to
spray the interior surface of a beverage can with a second coat after the body has been
formed, and also apply an additional side stripe lacquer spray along the side seam of the can
in order to repair any mechanical damage caused during seaming and soldering.
Currently, the substrates used for lacquer application are: tinplate, aluminum, TFS,
and blackplate, with tinplate dominating the market at present. The type of resins most
commonly used for internal tin coating are; oleoresinous, vinylic, acrylic, phenolic, and
epoxy-phenolic.
As mentioned above, control of dry film weight is essential in ensuring anticorrosive
properties. A dry film which is too thin may not cover the surface completely and impair
corrosion resistance. On the other hand, an overthick film, apart from being
uneconomical, makes for brittleness and impaired protection. The common dry film
weights of lacquers are in the range of 2.5 to 8 g/m2 (5 to 12 micron).
Table 1 summarizes some of the common lacquers, their properties and general
applications. Most lacquers contain a yellow pigment and thus are called "golden"
lacquers. With lacquered tinplate, the tin coating on the steel base may be significantly
reduced (50 to 75%), without reducing corrosion resistance.
The rise in energy cost and modern environmental legislation, necessitated
development of new coatings which would not cause pollution during application and
curing and would permit energy saving.46 Water-based interior coatings (which are
equivalent in all respects to the old solvent-based liners) are now being evaluated by can
makers. A typical water-soluble interior coating consists of about 36% solids, 53% water,
and 11 % nonpolluting hydrocarbons. There are still some off-flavor difficulties as well as
FDA approval problems regarding this type of coating. An alternative approach
consisted in development of high solid-content liquid lacquers. The problem here is
proper application due to the high viscosity of the coating. The ultimate aim is powders
which contain no solvent at all, require minimum baking for curing, and permit
Table 1
MAIN TYPES OF INTERNAL CAN LACQUER
Downloaded by [University of Arizona] at 05:39 21 July 2012

General type of resin


and components blended to Sulphide-stain
produce it Flexibility resistance Typical uses Comments

Oleo-resinous (drying oil and Good Poor Acid fruits Good general purpose range
natural or synthetic resins at relatively low cost
Sulphur-resistant Oleo- Good Good Vegetables, soups, (especially Not for use with acid products;
P
resinous (added zinc oxide) can ends or as topcoat over possible intense green color 8
epoxy-phenolic) with vegetables, e.g., spinach
Phenolic (phenol or substi- Moderate — poor Very good Meat, fish, vegetables, soups Good at relatively low cost
tuted phenol with formalde- but film thickness restricted
hyde) by flexibility
Epoxy-phenolic (epoxy resins Good Poor Meat, fish, vegetables, soups, Wide range of properties may
with phenolic resins) beer, and beverages (first be obtained by modifica-
coat) tions o
Epoxy-phenolic with Zinc Good Good Vegetables, soups (especially Not for use with acid products;
Oxide (zinc oxide added) can ends) possible color change with
some green vegetables
Aluminized Epoxy-phenolic Good Very good Meat products Clean but rather dull appear- 3"
(metallic aluminum powder ance
added)
Vinyl, solution, (vinyl Excellent Not applicable Spray on can bodies, roller Free from flavor taints; sensi-
chloride-vinyl acetate co- coating on ends, as topcoat tive to soldering heat and
polymers) for beer and beverages not usually suitable for
direct application to tinplate
Vinyl, organosol orplastisol, Good Not applicable Beer and beverage topcoat on As for vinyl solutions but o
3
(high molecular weight vinyl ends bottle closures, drawn giving a thicker, tougher
resins suspended in a non- cans for sweets, pharma- layer
solvent) ceuticals, tobacco
Acrylic (acrylic resin, usually Very good in some Very good when Vegetables, soups, prepared Attractive clean appearance
pigmented white) ranges pigmented foods containing sulphide of opened cans
stainers
Polybutadiene (hydrocarbon Moderate — poor Very good if zinc- Beer and beverages first coat. Cost and, hence, popularity
resins) oxide is added vegetables and soups if depends on country
with ZnO

From Britton, S. C , in Tin Versus Corrosion, I.T.R.I, Publication No. 510, 1975, 45. (With permission.)
Volume 17, Issue 4 377

recirculation and reuse of the heated air produced by the ovens.- Such coatings have been
developed, received FDA approval, and are in use in several plants.47
Quite often, exterior coatings are used for protection against rusting of the food cans.
The main difference is that exterior coatings do not require FDA approval and are not
subject to critical flavor requirements. The advent of UV-cured inks which contain no
solvents, require no baking, consume less energy, and release minimal emission to the
atmosphere was an important innovation. In addition to UV inks and coatings, water
based, non-varnish white base-coats are also used for this purpose.48

I. Methods of Can Fabrication


The two common methods for manufacturing open-top sanitary cans are the three
piece and the two piece can. The three piece can consists of a cylinder and two stamped
ends. The cylinder is normally produced on a high-speed bodymaker and the most
prevalent side seam is the "lock" type made with 98/2 (98% lead and 2% tin) solder. The
solder is a source of lead dissolution, especially when exposed on the lap portion of the
side seam, where the lead is not in direct contact with the tin surface, and has no cathodic
Downloaded by [University of Arizona] at 05:39 21 July 2012

protection. In order to overcome this problem, high-tin (or even pure-tin) solders are
used in specific cases, for example baby food. These solders, however, are expensive and
weaker in mechanical strength than the lead solder, and therefore unsuitable for general
use. Another solution widely used, is covering the side seam with a lacquer coating after
soldering — side-seam striping.
Welding and cementing are also used for side seams, the welding solution recently
taking significant strides. The ends, with rubber-based sealing compound, are double
seamed to the can body, one by the can maker and the second by the canner after filling.
The two-piece cans, which have made rapid progress, are produced by the drawn and
ironed (D & I) or the drawn-redrawn process (DRD). In the DRD process, a tinplate disc
is drawn into a cylindrical cup which is then redrawn in a sequence of operations to
produce a seamless can body complete with one end. Flow of the tin during the drawing
operations imparts a bright finish.34 The final operation is carried out on a header, which
beads and necks-in the bottom end. This process imposes severe stresses on the metal sur-
face and organic coatings, and the cans have to be resprayed. In this process the bottom and
walls have the same thickness.
D & I cans are usually made by drawing a cup from a coil of lubricated tinplate. The
drawn cups are fed into an ironing press which thins the sidewalls, thereby increasing the
can's height. The sidewall of the modern beer cans is reduced from 0.3 mm to 0.1 mm by
the ironing process. The cans are then decorated on the outside, baked, spray-coated on
the inside, rebaked, and finally necked and flanged. This relatively new development was
discussed to a great extent at the 1st International Tinplate Conference, as well as the
influence of the drawing and ironing operations on the surface properties.49'50
Until recently these two-piece cans were used only for beer and carbonated beverages,
mainly because the thin walls would not prevent paneling due to vacuum. Recently,
beaded two-piece cans were developed and are finding use in the vegetable canning
industry in the U.S. The capital investment for D & I or DRD two-piece cans is very high
and thus only warranted if long production runs can be assured. Another limitation is
that each line can accommodate a single can size only.

III. CORROSION OF TIN PLATE

Corrosion is the destructive attack on a metal through chemical or electrochemical


reaction with the environment.16 It represents the tendency of pure metals and alloys to
378 CRC Critical Reviews in Food Science and Nutrition

change into thermodynamically more stable compounds. In many cases this attack is
accompanied by physical deterioration such as swelling and perforation. The term
"rusting" refers to corrosion of iron or iron-base alloys with formation of corrosion
products consisting largely of hydrous ferric oxide.
Corrosion in food cans is characterized by anodic dissolution of the metal with ion
formation:

M —» M*1 + ze" (I)

The ions may be carried away from the electrode into the solution or may form insoluble
salts which precipitate on its surface. A layer of oxide or hydroxide may be formed on the
surface which may further be oxidized. Such layers may be porous, facilitating further
corrosion (as with tin-plated steel), or compact and nonporous, forming a protective
layer (as with aluminum).
In most cases, spontaneous corrosion is accompanied by hydrogen release or oxygen
reduction. Other cathodic reactions may occur, consuming the electrons released in the
Downloaded by [University of Arizona] at 05:39 21 July 2012

anodic dissolution of the metal. The corrosion reaction in tin cans is essentially
electrochemical and takes place in an aqueous medium through formation of galvanic
cells. Such cells are formed when two metals with different dissolution pressures are
coupled. The more active metal — the anode — dissolves, thus protecting the passive
metal — the cathode. Mechanical strain in the metal surface itself, impurities, and
microscopic faults may lead to formation of local cells generating corrosion currents
which, according to Faraday's law, dissolve an equivalent amount of the metal.
The capacity of a metal to dissolve in the aqueous medium is expressed by a potential
related to the Gibbs free energy, which is a measure of the feasibility of chemical reaction.
On the other hand, the above capacity does not reflect on the reaction rate, which is
measured by current density. Oxidation of a metal and reduction of species in solution take
place at identical rates, i.e., the total anodic and cathodic currents are equal, while the
current densities may differ substantially depending on the relative sizes of the anodic
and cathodic areas. When anodic areas are very small and carry a high current density,
the metal may undergo deep pitting and even perforation even though the average rate of
corrosion may be relatively small.
When a metal or alloy comes into contact with a medium, it assumes a potential
dependent on its nature and on the medium. This corrosion potential (ECOrr) is a mixed
potential resulting from the metal and hydrogen electrode processes.16'51'52 When a
reaction shifts away from equilibrium, i.e., when one direction is favored over the other,
the electrode is polarized. The potential at which the reaction takes place changes by an
amount called overpotential, and defined as t] = EcOrr — Ej, where Ej is the polarized
potential. The anodic overpotential r]a drives the metal dissolution process, and the
cathodic overpotential r]c drives the cathodic process. The current applied to cause
the shift from equilibrium is the net current of reaction and can be measured.
Current/potential curves for anodic and cathodic processes are shown schematically
in Figure 2.
In 1905 Tafel found that both anodic and cathodic polarization can be plotted as linear
functions of current, i.e.:

E = a + b log i (2)

where E is the electrode potential, i is the current density, and a and b are constants. The b
constant is referred to as the Tafel slope which is characteristic of a particular anodic (b a )
or cathodic (b c ) process and is expressed in mvper decade. At the point of intersection the
Volume 17, Issue 4 379

PASSIVEM

MOUC OXIDATION

CATHODIC REDUCTION
( 2 H * + Z*-— H2>

8
a:
u
Ul

ACTIVE (-)
Downloaded by [University of Arizona] at 05:39 21 July 2012

10 100 1000
CURRENT DENSITY (pA/cm 2 )

FIGURE 2. Schematic cathodic and anodic polarization curves.

two processes occur at the same rate, icorr, which is the steady-state corrosion rate. The
potential corresponding to this point is the corrosion potential. If a metal is connected to
a source of current or potential and polarized anodically or cathodically with respect to
its corrosion potential, a current is obtained, which permits electrochemical determina-
tion of the corrosion rate from the E vs. log i plot. This is done by either extrapolation
from high values of Eon on either the anodic or cathodic side to the corrosion potential
(Tafel plots), or by measurements in the vicinity of ECOrr (linear polarization).52"56

IV. CORROSION IN FOOD CANS

A. Plain Tin Cans


1. Reversal of Polarity
The mechanism of corrosion in acidic products in plain tin cans is essentially one of
cathodic protection of the steel where the tin serves as the sacrificial anode. The metallic
surface of the can exposed to the corroding medium consists of a large area of tin and tiny
areas of exposed alloy and steel through pores and scratches in the tin coating.2'8'57"60 For
the tin to be protective, it has to be anodic to steel under anaerobic conditions. However,
tin coupled to steel may be either the anode or the cathode, depending on the corroding
medium. In a dilute aerated acid medium the iron is the anode; it dissolves and liberates
hydrogen. In deaerated acidic food cans this is the initial situation but, later, reversal of
polarity sets in and the tin becomes the anode, thus protecting the steel. This reversal has
been related to the high hydrogen overpotential of tin compared with that of iron,43 and
to the low solubility constant of stannous hydroxide (5 X 1026) which forms and
precipitates on the cathode — reducing its size — compared with that of iron (1.6 X
1015).61 The decrease in stannous ion concentration due to formation of soluble complexes
with some organic ligands (complexing anions) in the product, also accounts in part
for this reversal.5''58'62 Reactions of this kind greatly impair the activity of Sn** ions with
which the tin is in equilibrium, producing a tin potential that is much more active and
actually less noble than iron. Under these conditions, the polarity reverses sign.
380 CRC Critical Reviews in Food Science and Nutrition

a. 3 r a stage
Q.

<

H
Z
UJ

o
u
Downloaded by [University of Arizona] at 05:39 21 July 2012

STORAGE TIME (days)

FIGURE 3. Tin dissolution in acidic foods, schematic rate curve.

The ratio between dissolved tin and iron concentrations, necessary for this polarity
reversal to set in, is obtainable theoretically from the Nernst equation, and was found to
be [ S n ^ / I T e ^ = 5 X 10"" in 0.1 M citric acid at pH = 3.8.16 In other words, the ratio
must drop below this level for tin to become active to iron. Such low levels can be realized
only through formation of tin complexes.
The potential of both tin and steel in fruit acids was found to respond to pH changes.
Also, it was found that the potential of the steel is influenced by the strength of the
complexing environment and the concentration of the stannous ions which form
complexes with sugar, pectins, and organic anions, rather than by ferrous ion
concentration. Other factors such as penetration of the steel lattice by atomic hydrogen,
or adsorbed surface films, play an important role in affecting corrosion rates and the
hydrogen overvoltage on the steel surface, modifying the relative position of the
corrosion potential of tin, steel, and the alloy, and thereby the corrosion mechanism.58'65

2. Rate of Tin Dissolution


Corrosion in acidic food cans comprises three stages67 (Figure 3). In the first stage, the
oil and tin oxide layers are removed from the can surface and the rate of tin dissolution is
high; this initial stage is under cathodic control whereby oxygen and depolarizers are
reduced. The rate and duration of this stage, which ranges from 4 to 15 days, depend on
the type of product and process parameters.59'66 The hydrogen evolved here seems to
diffuse into and pass through the steel.
In the second stage, which is lengthy and slow, the corrosion rate is almost constant.
Dissolution of the tin causes enlargement of existing pores and scratches, exposing the
steel and the alloy. At the same time there is proliferation of galvanic cells, where the
exposed steel at pores provides sites for the cathodic reaction of hydrogen evolution (the
hydrogen being taken up by the depolarizers in the product) and its area dictates the rate;
as the ratio between tin and steel areas decreases in the process, so does polarization.
Exposure of the steel is governed by the integrity of the electrochemically inert alloy
layer; the more homogeneous and continuous the latter, the better the protection
afforded in the final stages of corrosion.34'67'68
The third stage, in which large areas of steel are exposed, is characterized by a high rate
Volume 17, Issue 4 381

of tin and iron dissolution. The hydrogen evolves at a fast rate in gaseous form and
accumulates with loss of vacuum and possible swelling or perforation. This stage,
however, is of little importance, since by this time the can has reached the end of its shelf-
life anyway.
Tin dissolution in a can as function of storage time is schematically shown in Figure 3.
Mathematically, the first stage of corrosion is described by a polynomial of degree n (n =
1. . .3).67

[Sn~] = A, + Ait - A2t2 + A 3 t J (3)

where [Sn++] is dissolved tin concentration, A|, A2, A3 are the coefficients of the
polynomial, and t is storage time. When A2t2 = A3t3, the second stage sets in and is
described by a linear equation:

[Sn~] = a + bt (4)
Downloaded by [University of Arizona] at 05:39 21 July 2012

where a and b are again constants, differing according to product, temperature and
depolarizers.

3. Electrochemical Behavior of Tin-Iron Couple-Possible Situations


The mechanism of tinplate corrosion varies according to the type of food and
depolarizer. The corrosion potential of steel is determined by polarization of local anodes
and cathodes by the corrosion current. To protect the steel from corrosion by an
externally applied cathodic current, the entire surface must be polarized to a potential
equivalent to, or more negative than, that of the local anodes — which is the case with the
Sn/Sn** electrode in many food products.2 According to these relative relationships of
the potential, different types of internal corrosion are recognized as follows:2
Normal detinning — The tin is anodic to steel and affords complete cathodic
protection. Tin dissolution and hydrogen evolution at the exposed steel are slow. The
couple is under cathodic control and the area of exposed steel determines the couple
current. Tin enters into complexes, the hydrogen is attached to the depolarizers, and an
adequate shelf-life is obtained. This situation is characteristic of citrus products, peach
and apricot compotes, and other low-pH products packed in plain tin cans.
Rapid Detinning — The tin is sufficiently anodic to protect the steel but the couple
current is high. Rate of tin dissolution and hydrogen evolution are high and early failures
occur. This may result from exposure of large steel areas resulting in high couple
currents; a faster corroding steel which is more difficult to polarize cathodically; strong
complexing agents in the product, which remove Sn** from solution and render the tin
more anodic. This situation is characteristic of tomato and aggressive citrus products
such as lemon juice.
Partial detinning and pitting — The tin is anodic to steel, but local anodes on the latter
are more anodic than the tin, so that protection is limited. The exposed steel continues to
corrode and early failure takes place due to hydrogen swelling or to perforation. This
behavior is typical of inferior steel quality or of particularly problematic products such as
prunes or pear nectar.
Pitting only — In this case there is reversal of the normal situation, and the steel is
anodic to tin. The effect is a combination of the couple current and local anode currents;
iron dissolves at the exposed surface of the steel and pitting corrosion occurs at a rate
corresponding to the combined current. Early failure takes place due to hydrogen
swelling or to perforation. This mechanism is quite rare and it may appear in highly
corrosive products such as pickles and phosphoric-acid formulated carbonated
beverages; it may be prevented by using enamelled cans.
382 CRC Critical Reviews in Food Science and Nutrition

B. Lacquered Cans
In order to reduce corrosion it is necessary to suppress the corrosion current. This can
be done either by retarding the cathodic or anodic reaction, or by interpolating in the
electrolytic path a very high resistance (such as a polymeric coating). The effectiveness of
a coating is related directly to its ability to act as an impermeable barrier to gases, vapors,
liquids, and ions, thereby preventing corrosive action on the protected surface.69 The
passage mechanism of the first three involves dissolution in the film, diffusion through it
under a concentration gradient, and finally penetration to the other side. Passage of ions
is governed by the electrochemical characteristics of the film.
The corrosion process in lacquered cans is more complex than in plain tin cans and
depends not only on the quality of the base steel plate, the iron alloy layer, and the tin
coating, but also on the passivation layers and the organic coating. The latter is rarely
perfect and may have pores and scratches of varying size, some of which may reach the
alloy or even the steel layer.

/. Ion Transfer Through Lacquers


Downloaded by [University of Arizona] at 05:39 21 July 2012

Transfer of ions into and through an organic coating is complicated by the fact that
they are electrically charged and thus their motion is actually the flow of an electric
current across a potential gradient. The selectivity of organic coatings to ion transfer
depends on the electric charge,70 as well as on the concentration of the electrolyte. The
membrane itself may become charged if the diffusion rates of cations and anions through
it differ. The protection which an organic coating affords to the metal plate depends on its
resistance to this diffusion, which may take place even in the absence of pores, scratches,
or blisters but is certainly accelerated by them.71'72
Resistance-type inhibition by coatings was demonstrated through their protection
ability, evaluated by measurements in sea water.71 Values above 108 ohm/cm2,
maintained over an extended period, were found satisfactory, whereas values below 106
ohms/cm2 resulted in rapid corrosion.
Figure 4 shows, schematically, the effect of sea water on the electrical conductance for
layers of coatings under immersion, and indicates the time at which failure occurs.

2. Coating Thickness
The coating thickness greatly affects the performance of the lacquered food can.
Storage tests showed that there is a minimum thickness needed to assure protection
against corrosion.73 For nonaggressive products such as apricots, beans, etc., a thickness
of 4 to 6 yum suffices while for aggressive products such as tomato concentrate, layers of 8
to 12 jum are needed. The heavier coating reduces the porosity of the organic layer, as can
be seen in Figure 5.74'75 In most cases, the tin coating under the lacquer is between 5 to 12
g/m2, heavier coatings being used in larger cans.
The amount of tin-iron alloy exposure also has been found to influence performance of
lacquered cans.76 Since alloy exposure at the surface of tinplate increases rapidly when
the free tin layer is below 4 g/m2, it is advisable not to use tinplate with less than 5 g/m2
for lacquer coating. This is being reconsidered at present with the development of low
tin coated steel (LTS).

3. Electrochemical Processes
The electrochemical corrosion mechanism in lacquered cans was extensively
investigated9'40'43'77"79 and is described schematically in Figure 6. In most cases, the
product attacks the tin layer which dissolves. The lacquer protects the surface, and tin
dissolution occurs only from both sides of a scratch under the lacquer or through a pore
and cause anodic undermining (Figure 6A). The contact area with the contents is very
Volume 17, Issue 4 383

I layer
2 layers

3 layers

o
<
o
Q

o
o
Downloaded by [University of Arizona] at 05:39 21 July 2012

STORAGE TIME
(DAYS)

FIGURE 4. The effect of the number of coating layers on


lacquer conductance during storage.

small and dissolution is slow. The exposed area increases with time and lacquer
detachment may be observed. This results in the appearance of enlarged pores and
surface breaks of greyish color (that of the alloy layer) in the lacquer film. In terms of
potential, the alloy layer is more passive than tin and iron, so theoretically, dissolution
should proceed up to it. However, due to its extreme thinness and to its coupling effect
with the tin, the potential of the can may be more noble than that of iron, which may also
go into solution.77 It was assumed that in this case the lacquer may serve as cathode due to
the diffusion of protons through it.40 An example is tomato concentrate; while the overall
rate of tin dissolution is reduced in the presence of a lacquer coating, protection of the
iron by the tin is also impaired due to reduction in the ratio of exposed tin to iron areas.
Corrosion is concentrated in small areas and strong local currents may occur. The situa-
tion may be further aggravated by the presence of corrosion accelerators such as nitrates.
Figure 6B shows the case where the alloy and iron are more anodic than the tin. The
alloy layer dissolves rapidly and the iron is attacked. Since no tin is dissolved, the alloy
layer is not laid bare, and so it has no influence on the corrosion process. Eventually,
failure occurs due to pinhole formation (also called pitting corrosion). In this case there is
no undermining of the lacquer, but the corrosion usually starts at a point of
discontinuity, i.e., a pore or scratch in the coating. Cans containing such aggressive
products do not have a corroded appearance and only closer inspection reveals spots of
corrosion, often with deep penetration into the steel. Examples of such products are beets
in acetic acid and berries.77
The ratio of exposed tin to iron is very important in all corrosion processes. If the tin is
anodic, the rate of corrosion increases as more iron becomes exposed, thus iron areas
384 CRC Critical Reviews in Food Science and Nutrition

E
o

v>
O
Q.

O
Downloaded by [University of Arizona] at 05:39 21 July 2012

•z.

COATING THICKNESS
(MICRONS)

FIGURE 5. Numbers of pores as a function of coating


thickness.

=---^--~ pore -=r= pore — —

• jJIIW " ^ i— \ ^ ^ _.«?_.

Fee-

FIGURE 6. Tin and.iron corrosion processes in lacquered tin can.


(From Maercks, O. and Maercks, I., Verpack. Rundsch., 22,4,29,1971.
With permission.)

should be kept as small as possible. Conversely, if the product is an iron dissolver (i.e., the
iron is anodic), it is the tin areas which should be kept to a minimum. In such a case
imperfections near the side seam must be masked by a side seam stripe. Where zero metal
dissolution is required, for example in beer or carbonated beverages, the entire can must
be relacquered after the forming operation.
In light of the above, foodstuffs may be classified into three groups indicating the
direction of aggressivity and characterizing the type of corrosion on prolonged storage:
tin dissolvers, tin and iron dissolvers, and iron dissolvers. It should be borne in mind that
the direction of aggressivity can be altered when the initially present oxygen has become
Volume 17, Issue 4 385

exhausted. Thus, a product which is originally an iron dissolver may become a tin
dissolver in the course of the storage period.

4. Adhesion
Failure of lacquered cans is often due to reduction in the bond between lacquer and
metal surface, eventually causing lifting of the lacquer film. Adhesion is thus required to
prevent anodic reactions, counteract forces developing under the coating due to physical
or chemical factors, and, last but not least, ensure an esthetic appearance.
There are several ways in which loss of adhesion may occur.40'43 In some cases it may set
in even prior to corrosion due to stresses induced during fabrication, but more often
detachment of the coating is due to breakdown processes taking place through or under
the coating, or to underfilm corrosion spreading from exposed metal. The detachment
mechanisms may be classified42 as:
Water displacement — There is always a monomolecular layer of water between the
oxide film on the metal and the organic coating, and as the water coverage increases the
bond between coating and metal is bound to be adversely affected. Theaffinityforwateris
Downloaded by [University of Arizona] at 05:39 21 July 2012

at a minimum at the isoelectric points of the oxide, hence the loss of adhesion due to
water displacement is governed by the pH of the electrolyte.
Cathodic detachment — Cathodic detachment can best be explained by formation of
hydroxyl ions due, for example, to reduction of residual oxygen in the can according to
the reaction O2 + 4e~ + 2H2O — 4OH"; the resulting high pH in turn makes for loss of
adhesion by water displacement.80
Anodic undermining — Anodic undermining is a type of crevice corrosion in which the
surface layer of metal beneath the organic coating is dissolved as mentioned before. The
anode is in the crevice between the metal surface and the coating. This type of corrosion is
promoted by acidic anions like chlorides and sulfates: lower current densities favor
anodic undermining along the surface, and higher current densities corrosion into the
metal.42
One of the major factors affecting adhesion and coating performance is the surface
quality of the tinplate prior to coating. Contaminants such as salt residues were found to
cause failure of the coating due to their ability to attract water (thereby establishing a
conducting film beneath the coating) and to provide ions for carrying the corrosion
current.42 As already noted, the organic coating is not bonded directly to the metal
surface but via the oxides or other active sites formed during passivation of the metal
plate.43 The chemical structure of the passivation films in tinplate and TFS has been
studied by many investigators.22'24'29'31'33'81
Recently an effort was made to relate lacquer adhesion to the structure of the
passivation film which was found to contain three kinds of chromium compounds
(metallic CR, Cr2C>3, Cr (OH)3) and amorphous tin oxide, which crystallizes to SnO,
orthorhombic SnO, and SnO2 during baking.31 It was found that lacquer adhesion
performance of tinplate passivated by various methods and its variation with ageing
could be interpreted by reference to the types of chromium and tin oxides in the layers.31
Only tinplate which yields orthorhombic SnO during baking provides good adhesion;
with the other forms adhesion is poor. Optimal adhesion, as measured by a peel test, was
afforded by a layer containing Cr°—Cr2O3 (4kg/5mm), followed by orthorhomic SnO (3 to
3.5 kg); SnO (2 to 2.5 kg); Sn-Cr2O3 (0.3 to 0.6 kg) and Sn-SnO2 (0.3 to 0.5 kg).31
So far no standards have been formulated for surface treatment of tinplate or TFS
which would ensure adhesion and satisfactory performance of lacquered cans. It is well
known, moreover, that considerable differences exist between passivation treatments
according to the manufacturer, as well as between batches of the same manufacturer,
with the attendant fluctuations in degree of protection. Adhesion, at present, is evaluated
386 CRC Critical Reviews in Food Science and Nutrition

Table 2
SOME CORROSION PROMOTING AGENTS AND
THEIR MODE OF REACTION

Reduction
Corrosion accelerator product Equivalent in weight

Proton (H*) H2 ml!Hi= 5.3 mgSn**


Oxygen (O2) H2O m8 O2 = 10.6 mg Sn"
Sulfur dioxide (SO2) H2S mg SO2 = 5.5 mg Sn"
Sulfur (S) H2S mg S = 3.7 mg Sn"
Nitrate (NO3) NHj mg NO3 = 7.65 mg Sn
Trimethylamine
oxide (TMAO) TMA mg TMAO = 1.57 mg

mostly by measuring the force required to lift a dry lacquer coating from the metal plate
in a peel test, or by assaying a scratched or unscratched surface with scotch tape. While it
Downloaded by [University of Arizona] at 05:39 21 July 2012

is practically certain that films which fail in the scotch tape test afford poor protection,
there is no guarantee that those which pass the test would give satisfactory long-term
performance when in contact with a specific food. A test was developed in which a given
potential is applied to a coated scratched plate, dipped in a certain electrolyte, and the
lacquer adhesion is measured. While this test comes close to stimulating actual
conditions in a can, it still calls for further refinement.82 Several laboratories are now
looking into new physical methods such as transmission electron microscopy, photo
electron spectroscopy (ESCA), and Auger electron microscopy (AES) analysis, in an
attempt to elucidate the factors which affect good bonding and performance.29'83'84 Once
better understanding of the bonding effect is available, it can be assumed that standards
will be drawn up.

C. Effects of Contents, Process, and Storage Temperature


1. Introduction
The properties of the food, the tinplate used for packaging the food, the technological
process and storage conditions are interrelated in affecting the shelf-life of a product; in
fact, a synergistic effect may accelerate corrosion.
The foods themselves are complex biochemical systems covering a wide range of pH
and buffering properties; they contain oxygen, acids, coloring matter, and other
ingredients such as salt, sugar, spices, and phosphates — which may either accelerate or
inhibit corrosion. The most important corrosion accelerators in foods include oxygen,
nitrates, sulfur compounds, trimethylamines, anthocyanins, dihydroascorbic acid, etc.
Some typical corrosion reactions associated with these accelerators and their
stoichiometric equivalents of dissolved tin are presented in Table 2.85
Depolarizers, which have a high redox potential and are readily reduced by hydrogen
formed in cathodic reactions, also make for accelerated corrosion. Some depolarisers are
introduced into the food can in natural form (such as sulfur compounds in meat and
vegetables, anthocyanins in fruit, oxygen), and others as unintentional additives (sulfur-
containing fungicides, nitrates).
Technological factors (vacuum, temperature at fill, and the sterilization regime) also
play a role in determining shelf life. No less important is the influence of storage
conditions.9'86"102

2. Acidity
There is no direct proportionality between acidity and the degree of corrosion of
Volume 17, Issue 4 387

Table 3
AMOUNT OF TIN IN SOLUTIONS OF VARIOUS
ACIDS, O2-FREE, AFTER 35 DAYS AT 30° C

Acid Oxalic Citric Malic Tarta

Dissolved tin (ppm) 20.3 10.5 8.5 5.5


Calculated corrosion 26.8 13.9 11.2 7.2
rate (nA/cm 2 )

From Sherlock, J. C. and Britton, S. C , Br. Corro. J., 7, 180, 1972.


(With permission.)

tinplate. In other words, two products of the same acidity need not necessarily be equally
corrosive.57 Experiments conducted with different organic acids showed that the acids in
pure form were less corrosive than the fruit juices containing them.57'60
Downloaded by [University of Arizona] at 05:39 21 July 2012

Variation in internal corrosion was attributed by many workers to the specific nature
of the organic acids involved, rather than to their concentration.60 Tin was found to be
anodic to iron in 0.75% malic and citric acids, but cathodic to iron in malonic, succinic,
and acetic acids. It was hypothesized (and proved) that the tendency of the acid to form
complexes with the dissolved tin has an important bearing on the relative polarity of tin
and steel, i.e., on the degree of corrosion.8'58'65'103 It also was suggested that the depressed
potential of a tin electrode in fruit juices is due to a decrease in stannous ion
concentration through complex formation.65 The complexing anions in fruits and juices,
and their complexing strength, were found to control the relative position of the
corrosion potentials of tin, steel, and tin-iron alloy, and thereby the corrosion
mechanism.65 Other factors such as natural oxidants and adsorbable materials in fruit
juices and some beverages may also affect the potentials. The stability constants of
complexes formed between stannous tin and some organic acids were found to correlate
with tin pickup.58
It was found that for a given ligand concentration, the more stable the complex
formed, the higher the corrosion rate (Table 3). The descending order of the corrosion
rate in the Table 3 was changed in the presence on nitrates, possibly due to reduction of the
latter and the resulting shift in the potentials.58

3.pH
As in the case of acidity, there is no direct proportionality between the pH level and the
degree of corrosion of tinplate. When a metal is dissolved, the reaction product is not
always an ionic species but often a solid oxide or hydroxide. Using free-energy
considerations, the reversible potentials for the metal in equilibrium with its hydrated
ions or with its. insoluble oxide is obtainable from the Nernst equation, and a plot of the
equilibrium potential E against pH can be drawn.104 These plots, commonly known as
Pourbaix diagrams, are widely used in determining the equilibrium cell potentials as well
as pH regions of stability, corrosion, and passivity, and provide answers as to whether a
particular corrosion process is feasible.54 They indicate graphically which electron-
transfer, proton-transfer, and combined-transfer reactions are thermodynamically
favored for a given metal solution system.54
Diagrams of stability of tinplate in deaerated citrate solutions at 30° C, constructed by
Bombara et al.51 show regions of immunity, resistance, unsuitability, and attack as
function of the potential and pH of the system.51'59'105 The pH of the system also
determines the relative cathodic protection given to steel. In some cases tin is cathodic to
388 CRC Critical Reviews in Food Science and Nutrition

steel over a certain pH range (acetic acid environment of pH 2 to 4.5) while in others it
offers protection up to pH 4; above that level, it may accelerate corrosion.65

4. Sulfur Compounds
Sulfur compounds in foods may derive from agricultural chemicals (in trace amounts),
or as residues in preservation or discoloration processes. Their role in corrosion varies
according to circumstances. Acceleration of corrosion by sulfur compounds was
generally attributed to formation of an iron sulfide layer in Which the iron in the FeS/ Fe
couple is anodic in anaerobic media at pH 3.6 to 4.6." The potential of this couple was
found to be more anodic than that of the Fe/ Sn couple, thus contributing to localized
corrosion.
The frequent cases of accelerated corrosion and failure of plain tin cans with acid foods
such as apricots and peaches were attributed to derivatives of thio- and dithiocarbamic
acid fungicides."5"120 At the same time, pitting corrosion in cans containing products
with fungicidal residues was attributed to inactivation of the tin coating by a protective
film of sulfide. This film has a more cathodic potential and consequently there is a
Downloaded by [University of Arizona] at 05:39 21 July 2012

significant reduction in the tin dissolution rate, and no electrochemical protection of the
steel by the tin.34'115'120
As regards the effect, known as sulfide staining, commonly observed in cans with high-
pH products such as meat, fish, and certain vegetables, it is more an esthetic than
chemical problem, since it does not necessarily cause intensified corrosion.13'34'106"114 A
study of corrosion and preferential sulfide staining of tinplate was believed to follow two
steps:121 first, oxidation of tin, and second, deposition of an insoluble tin-sulfide
precipitate on the surface. Isolation of the stains by special procedures,121'122 and their
chemical analysis confirmed the presence of sulfur, tin, and iron in the stain; X-ray
diffraction indicated that the sulfides consisted of SnS with some FeS, SnS2, and FeSn2.
The potential difference and current flow between crystal grains constitute an
electrochemical mechanism whereby oxidation of the tin takes place on the (220) and
(321) crystal faces and reduction on the (112) crystal face.121 This preferential pattern was
identified by X-ray diffraction techniques.
Residual SO2 accelerates corrosion through its action as a depolarizer, inducing a
negative charge in the double layer (which in turn repels the electrons from the electrode)
and thus shifting the potential in the positive direction.123 Such a shift in potential was
observed in cans containing a model solution of citric acid and trace amounts of SO2; tin
pick-up was also increased. Experiments performed with canned grapefruit juice spiked
with SO2 showed that even 1 ppm accelerated corrosion.67 ATC values, as well as
increased tin and iron pick-up, also were increased by addition of 1 ppm SO2 to an acid
medium.124 Most of the above-mentioned corrosion problems may be avoided by filling
such products in lacquered cans.

5. Nitrates
Nitrates are found in plants, water supplies, and heavily fertilized soils. Extensive
studies demonstrated that green beans, spinach, turnips, lettuce, beets, and radishes
contain several thousand ppm of nitrates*.125"128 During storage, the nitrates are reduced
through ,a number of intermediates to ammonia and act as corrosion accelerators.
Accelerated corrosion of plain tin cans due to nitrates has been investigated and reviewed
thoroughly, due to the seriousness of the economic and toxicological consequence of the

* Nitrates and nitrites are also present as intentional additives in processed meats. However, since these
products are above the critical pH range for detinning to occur via the nitrate-tin reaction system (pH = 5.5),
there is no intensified corrosion when uncoated tinplate is used.126
Volume 17, Issue 4 389

problem, especially in tomato products.14'67'129"141 Corrosion in cans with grapefruit juice


was found to be accelerated by as little as 10 ppm nitrates. Nitrate in combination with an
enlarged headspace accelerated corrosion 2 to 3 times, as was indicated by tin pick up.
Iron pick up was also increased, 5 to 10 times.67 The quality of the tinplate itself (as
indicated by its ATC value) in combination with nitrate was also found to affect
corrosion.96 The dependence on pH was confirmed by several reports and the critical
range was found to be 2.5 to 3.5.130'131'134'142
Admixture of nitrates to a model solution containing citric acid or acetic acid at pH 3.6
was found to increase the corrosion current and tin potential significantly, but hardly
changed the iron potential.143'144 A reduction in corrosion current, and lowering of the tin
potential according to polarization measurements, was attained when pH was increased.
Neither the type of steel (MR or L), nor the type of tin coating (electrolytic or hot-dip)
were found to affect corrosion due to nitrates.
The mechanism of corrosion in the presence of nitrates differs from the regular
detinning process in which electron discharge is confined to small areas of steel where
hydrogen is envolved and the catodic reaction becomes rate-limiting.129'131"132'143'145"149
Downloaded by [University of Arizona] at 05:39 21 July 2012

Here, the nitrates act as electron acceptors, thus opening a new pathway for electrons
produced when tin is dissolved, and shifting the reaction towards increased tin
dissolution. The high hydrogen overvoltage is no longer a barrier to electron discharge at
the tin surface, once the hydrogen evolution reaction has been replaced by the electron-
nitrate reaction. The reduction reactions involved are believed to be as follows:131

4 Sn —• 4 S n " + 8e~ (5)


slow
—> NOI + H 2 O
NOJ + 2e~ + 2H—> (6)
fust
+ 2H2O (7)
Equations (6) and (7) indicate that the rate of detinning depends on the nitrate
concentration and pH. Hydrogen is not produced in these reactions, so that can vacuum
does not change appreciably during nitrate detinning. The first reduction reaction
(nitrate to nitrite) is slow and probably rate determining, but the second (to ammonia) is a
fast one*. In view of the potential level involved (0.86 to 0.94 volt for nitrate reduction vs.
-0.14 volt for tin oxidation), intermediary intervention in the oxidation-reduction system
was postulated to meet conditions for acceleration. 131 The ferrous ion was found to meet
the requirements for such a catalyst in the nitrate-tin system, and possible reactions are
as follows:

NOI + 3Fe*2 + 4H+ >NO + 3Fe<3 + 2H 2 O (8)

(reduction of nitrate to nitric oxide)

2Fe*3 + Sn > Sn+2 + 2Fe<2 (9)

(tin oxidized by ferric ion). Combined, the two reactions give:

2NOJ + 3Sn + 8H* • 2NO + 3Sn*2 + 4H 2 O (10)

* Besides the ammonium ion, nitrous (N2O) and nitric (NO) oxide and hydroxylamine were found among the
reduction products of nitrates.148 In citrate buffered systems, at pH 5, it was found,134 that about 80% con-
version to ammonia occurred, but at high pH the ammonia yield accounted for only a small part of the
nitrate lost. Nitrous oxide was detected by gas-chromatography in headspace of the cans, and its concen-
tration varied during storage.
390 CRC Critical Reviews in Food Science and Nutrition

In principle, there are several ways in which nitrate detinning may be prevented or
reduced:

1. Use of crop varieties and cultivars which do not accumulate nitrates


2. Strict control of the amounts, composition and application timing of fertilizers:
difficult in practice, and probably unacceptable to producers. (It is possible to
modify some cannery operations, such as washing and blanching of the crops, but
these offer only marginal protection against nitrate detinning)
3. Internal lacquering of cans — the best solution to-date
4. Elevated pH: may be effective with some vegetables, but cannot be used with fruit
products because of organoleptic and processing considerations
5. Corrosion inhibitors: a subject of recent interest. This solution involves a long-term
approach and can at best be partial135'137

Certain compounds such as thiourea, ally! thiourea, carbon disulfide, and others were
found to inhibit detinning by blocking nitrate reduction on the tin surface at the nitrite
Downloaded by [University of Arizona] at 05:39 21 July 2012

stage,150 but these are prohibited in foods. Addition of anionic surface-active agents to
solutions containing citric acid and nitrates at pH 3.0 also had an inhibitive effect as
measured by polarization conductance.151 Ascorbic acid (a natural component with
oxidation-reduction properties) added to acid model solutions containing nitrates also
was found to inhibit detinning to some extent.152 Sodium benzoate was tried at various
concentrations but found to be only partially effective.153 Potassium sorbate, sodium
lauryl sulfate and p-amino benzoic acid were tried with mandarin segments (a very
aggressive product) and also found to have a partial inhibition effect.uo Their action was
attributed to formation of a fine film adsorbed on the metal surface, which prevents
interchange reactions and alters the electric field at the metal-solution interface.140
Several authors found natural constituents in cabbage juice, sauerkraut, turnip, and
mustard greens, which inhibit nitrate detinning in acid foods.130'149'150 In the light of the
above, the search for substances with higher protection efficiency against nitrate
corrosion of tinplate is promising and should be continued.

6. Technological Processes and Storage Conditions


The shelf life of canned foods is affected considerably by storage temperature and
processing parameters such as filling and exhausting procedures, processing time and
temperature, proper cooling of the cans etc. Both the quality of the product and the rate
of corrosion are affected.

a. Oxygen
Oxygen has been given much attention in the food industry due to its general
detrimental effect.57'66'67'155"157 It participates in oxidation reaction which results in
deterioration of ascorbic acid, browning, organoleptic changes, and reduction of the
nutritive value. It also acts as a depolarizer, accelerating corrosion by reacting with the
hydrogen formed in the can. (The presence of even small quantities of oxygen prevent
accumulation of hydrogen, through a cathodic reaction as follows:

O2 + 4H* + 4e »2H 2 O)

Oxygen may be present in the headspace of the can, dissolved in the product, and
absorbed or entrained by particulate products. Elimination of this oxygen is a common
practice and is usually done by hot-filling or vacuum filling of the product, exhausting,
steam injection, or vacuum syruping.
The rate of oxygen consumption is high at the beginning of storage and decreases with
Volume 17, Issue 4 391

Table 4
AVERAGE Qio VALUES

Product Average

Test solution 1.4


(1% acetic acid
1.5% salt)
Tomato soup 1.3
Tomato puree 1.15
Oxalic acid (0.5%) 1.1

time. The rate of disappearance depends on the initial concentration, vacuum in the can,
headspace level, filling temperature, and type of product and package. The corrosive
effect depends on the type of product and the type and concentration of acceptors
present. Solubility of oxygen is low in concentrated and viscous solutions, and its
corrosive action is limited due to the lower diffusivity.
Downloaded by [University of Arizona] at 05:39 21 July 2012

Vacuum and headspace level depend largely on the effectiveness of exhausting and on
the closing temperature. Exhausting is especially important in paniculate products such
as melon cubes and apple segments. In this kind of product, removal of oxygen by hot
filling is aided by vacuum syruping. In liquid products hot filling was found to be
satisfactory, whereas in concentrates, preliminary deaeration is called for.158 Increased
headspace and vacuum affect corrosion due to high initial oxygen level.60'66'157

b. Thermal Treatment
As a general rule, the higher the temperature, the higher the rate of the chemical
reactions responsible for degradation of the product, and that of the corrosion reactions
— the shelf-life of the product is reduced accordingly.
Little is known on the effect of sterilization on corrosion phenomena. For aluminum
cans, the corrosion current increased 10 to 15 times during sterilization compared to that
at 20°C.159 With some products this even resulted in perforation. No such extreme
corrosion currents were found for tinplate cans. In experiments on carrot cubes, with
corrosion currents measured under heating and cooling, the product alternated between
iron dissolver and tin dissolver according to temperature. However, since sterilization
time is very short compared to the total storage period, the quantity of metal dissolved
was very small.77
The importance of correct cooling is also emphasized. Fast cooling may cause paneling
of the can and prevent complete drying causing external corrosion. On the other hand, it
was shown that faulty cooling increased corrosion and browning specifically in viscous
products, such as tomato or lemon concentrates.

c. Storage Temperature
The importance of the storage-temperature parameter is due to the fact that at elevated
temperatures, the chemical, electrochemical, and physical reactions in the can-food
product system are accelerated, with possible formation of degradation products. Earlier
data on this point were reviewed by Hartwell,57 and more recently by Mahadaviah.60
The general rule of thumb whereby shelf life is halved with every 10°C increment in
storage temperature (i.e., Qio = 2) is only of limited validity and does not apply to such
phenomena as iron dissolution to the stage of perforation.77 Figure 7 shows the
temperature of the corrosion current measured between tin and iron for some test
solutions and food products (measurements normalized at 20° C to a relative value of 100
units), and it is seen that Qio does not equal 2 for any of the products (see also Table 4).
The 0.5% oxalic acid solution is almost temperature-independent in its aggressivity and
392 CRC Critical Reviews in Food Science and Nutrition

Theoretica! l%Vmegar Tomoto


curve (q=2) 1.5% Salt soup
400

350-
Downloaded by [University of Arizona] at 05:39 21 July 2012

10 20 30 40 50 60

-50 TEMPERATURE , »C

FIGURE 7. The dependence of corrosion current on temperature.


(From Maercks, O. and Maercks, I., Verpack. Rundsch.. 22,4,29,1971.
With permission.)

3% acetic acid even has a negative temperature coefficient. (At 20° C the latter is a tin
dissolver, but between 50 and 60° C becomes an iron dissolver.)
Storage tests on fish preserves at 40° C failed to show perforation corrosion, but at
20° C the cans pitted rapidly.160 This apparent paradox was explained by reversal of the
corrosion effect between these two temperatures.77 Similar phenomena were observed
with other fish products.160 However, it is wrong to conclude from this that increasing
temperature improves storage stability, since corrosion is only one of the processes
affecting shelf-life.77'161'162

V. TOXICITY OF DISSOLVED METALS

In the past few years, authorities have become increasingly concerned about possible
chronic and harmful health effects of certain nonessential metal contaminants in
foods.163'164 Most countries have enacted food laws and regulations, usually in the form
of maximum allowable limits for specified metals in individual foods. With the advent of
new and more accurate analytical techniques,163'164'180 and the accession of new
toxicological and nutritional data, this legislation is necessarily subject to continual
revision.
Toxic elements present in foods as contaminants are the consequence of spreading
industrialization and the attendant pollution of the biosphere. Elements known to be
toxic at very low levels of intake are arsenic, antimony, lead, cadmium, and mercury.163
Volume 17, Issue 4 393

Others, like selenium, are essential diet constituents at low levels but have undesirable
effects at higher levels, or in situations where the balance between intake and excretion is
upset.
Corrosion products in food cans are limited to three metals, tin, iron, and lead, which
are liable to dissolve from the container. Of these, only the lead is toxic and cumulative in
body tissues, hence a hazard.165 Although in all foods lead is cathodic to tin and steel, so
that in coupled form it is cathodically protected against corrosion, lead contamination
cannot be prevented completely. In addition, its low hydrogen overvoltage and its
tendency to form insoluble carbonates and phosphates render this form more cathodic
than the pure metal. In lacquered cans, there is a small exposed solder area in the side
seam, with hardly any tin area for cathodic protection; accidental solder splashes, or dust
sticking to the surface,34'165 also lack cathodic protection and may dissolve in chemical
reactions. The same is the case with plain tin cans. This is the reason for the preference for
tin solder, especially for baby-food cans.
Lead content varies widely within and between food products. Even the raw material
may contribute some of it, although generally the level is well below the regulatory limits,
Downloaded by [University of Arizona] at 05:39 21 July 2012

which is 2 ppm for most foods, 0.5 ppm for baby foods, and 0.2 ppm for soft drinks.166"170
In another survey,170 the lead content of 168 samples of acid foods was found to be in the
range of 0.02 to 8.16 ppm, the average for lacquered cans being 1.45 ppm and for plain
cans 0.46 ppm. In Spanish canned vegetables the average lead content was found to be
less than 1 ppm, with the exception of pickled cucumbers in lacquered cans where it was
8.5 ppm.168 Average lead levels in canned grapefruit juice in Florida, after 1 year of
storage, were 0.2 ppm with extremes of 1.6 ppm in cans with large exposed solder areas or
with splashes, and 0.08 ppm in cans without exposed solder.171 In 1980, the American
canning industry carried out its first "market basket" survey, sampling 24 similar canned
products from retailers' shelves; the average lead level obtained was 0.22 ppm.172
Iron, an essential constituent of our diet, does not constitute a toxicity problem, and a
limit of 50 ppm is under discussion for canned vegetables, fruit, meat, fish and milk.
Tin. Most of the tin present in canned foods is insoluble in the gastric and intestinal
juices and is not absorbed during digestion.60 In solid foods high levels may occur; its
toxicology has been reviewed extensively.174"176 Metallic tin and its salts are considered to
be of low oral toxicity, whereas alkyl derivatives are highly toxic.177 Outbreaks of
poisoning manifested by nausea, vomiting, and other gastrointestinal disturbances were
traced to solid foods and drinks containing high levels (hundreds of ppm) of tin.173'174'178
Experiments on test animals showed not only gastrointestinal disturbances, but also
anemia, tumors, and other pathological changes.174'176'179 However, the tin compounds
administered in the animal diet differed totally from those involved in canned foods and,
besides, the tin dosage far exceeded the quantities available for dissolution in the latter.
The 250 ppm level quoted in the past as the permissible upper limit in canned foods is
based not on toxicological evidence of safety, but rather on the fact that higher levels
produce off-taste and are rarely found under normal conditions of processing and
storage. The tendency today is to reduce it to 150 ppm.167 (In lacquered cans, the tin
content of the food rarely exceeds 100 ppm.)
Rarer toxic metals such as cadmium, mercury, and selenium, derive mostly from
environmental pollution and vary depending on the origin of the crop.163'169

VI. METHODS FOR EVALUATING CORROSION AND SHELF-LIFE

A. Introduction
Quality assurance for any product-package combination during storage requires
preliminary evaluation through laboratory tests. Conventional long-term storage tests
are time consuming and costly, but still indispensable in studying basic problems. At the
394 CRC Critical Reviews in Food Science and Nutrition

same time, reliable rapid tests have been developed to satisfy the need for quick answers
in evaluating new products and markets, in disputed interpretation of acceptance, etc.
The questions that must be answered by these tests include choice of a dominant quality
parameter, such as the time needed to reach a certain deterioration level, its method of
determination, and its critical value.
A proper criterion for shelf-life evaluation has to encompass the entire package-
product-environment system, always bearing in mind that the quality of the product is
the most important consideration. (A criterion such as hydrogen swelling would be
meaningless if the product has already become inedible due to browning reactions or to
textural and other organoleptic changes.)
The procedures enumerated below (for a more detailed review see Britton34) are
suitable as acceptance tests for purchase of the cans; some of them are intended for
routine inspection and quality control, others for research. Some are described more
fully, especially those which have a bearing on can-product interaction and were used for
research purposes by the authors.
Downloaded by [University of Arizona] at 05:39 21 July 2012

B. Tests for Tinplate


/. Tin Coating Weight
Tin coating weight is determined coulometrically with anodic dissolution.34 Suitable
instruments were developed for in-line control.

2. Porosity of Coating
Porosity of the tin coating is an important factor in corrosion resistance. The common
tests include dyeing of a specimen surface, whereby pores and other surface faults in the
tin layer are brought out and examined visually for number and pattern. Another method
is anodic treatment of a specimen in ammonium thiocyanate and recording of the current
between it and a counter electrode; the current correlates with the area of metal exposed.

3. Tin Oxide
Tin oxide is determined by applying a cathodic current in deaerated hydrobromic acid
and reducing it to the metal. The change of potential indicates removal of oxide and the
current measured relates to the amount of oxide through Faraday's law.122'181'183

4. Chromium in Passivation Layers


The underlying principle is anodic oxidation of the passivation film to soluble
hexavalent chromium and the oxidation current is converted into the amount of
chromium through a calibration curve. A chemical method yields separate measure-
ments for the assorted forms of chromium (metal and hydrated oxides).14'15'18"20'34'182
Recently, X-ray spectrometry was found to afford a rapid measurement of tin coating
weight and chromium content in passivation layers.183

5. Special Property Tests


These refer to characteristics which influence the shelf-life of acid products in plain tin
cans (of these tests, only the last-named is still in use for certain fruit packs):
Pickle lag test — A specimen is immersed in hydrochloric acid, and the time needed to
reach a steady rate of hydrogen evolution is recorded.
Iron solution value (ISV) — Specimens are exposed to a standard solution, and the
dissolved iron concentration is measured.
Alloy tin couple test (ATC) — (Used successfully for shelf-life evaluation of canned
citrus products). The couple current flowing between the can metal stripped to its alloy
layer, and a tin electrode immersed in deaerated grapefruit juice, is measured and gives an
indication of uniformity of the alloy layer and tinplate quality.
Volume 17, Issue 4 395

C. Tests for Lacquers


Test methods for lacquered tinplate were detailed by Lempereur and Biston.185
(Physical characteristics of the lacquers properties (specific gravity, solid content,
viscosity, and completeness of cure) need not be dealt with here since they are only of
interest to can manufacturers.

1. Thickness of Film
Nondestructive magnetic or electromagnetic techniques, such as the Pfund gauge or
conductance measurements, may be used. For accurate determination of weight, the
lacquer has to be peeled off or dissolved from a weighed specimen of known area. Some
lacquers are difficult to remove without damage to the tin surface.

2. Continuity of Lacquer
Pores, scratches, and other discontinuities in the lacquer may be examined
qualitatively by dyeing, as in the case of the tin coating. Dipping the plate in a sulfuric-
Downloaded by [University of Arizona] at 05:39 21 July 2012

acid solution of copper sulfate, or in ammonium thiocyanate, or exposure to sulfur


dioxide, brings out such porous and faulty areas.^4 Other methods include anodic
porosity tests in which a fixed voltage is applied to a test specimen serving as anode in a
solution of sodium hydroxide or ammonium thoicyanate (bleeding test), or in a mixture
of sodium benzoate and sodium thiosulfate, and the current between the specimen and a
platinum counter electrode is recorded; the current correlates with the exposed steel
area.34 Similarly, a whole can may serve as cathode in sodium chloride solution, and the
current between it and a carbon anode is recorded.

3. Adhesion
Loss of adhesion has two main causes: mechanical faults during manufacturing and
exposure to a corrosive environment.
The most common testing method is the Scotch Tape technique in which the tape is
applied to a pattern of scratches and then pulled away (the same technique is used after
heavy heat processing in model solutions). In a recent modification,82 the test is carried
out after the specimen has been subjected to a 50 mA current. Another method is to press
by a hot press between two lacquered sheets with an adhesive film and then measure the
peel strength with a tensile testing machine.31

4. Electrochemical Test
A fixed voltage is applied between the lacquered tinplate specimen and a carbon rod in
a sodium chloride and vinegar solution, in a specially designed cell. The current, as well as
the potential drop with time, were found to indicate the degree of protection.77 A simple
test in which the current flowing between a bare tin electrode and the lacquered surface
is measured, was devised by Koehler.40 The result was found to be sensitive to the effect of
depolarizers. Later, it seemed that polarization resistance measurement offers promise as
a means of assessing the corrosion resistance of lacquer-tinplate. It was applied for
evaluation of isolated films and lacquered tinplates and found to be a useful indicator for
performance of the system.186

D. Product Evaluation
1. Chemical Parameters of Product
Undesirable physical and chemical changes undergone by the product during storage
need not necessarily be associated with corrosion. Such effects as browning, loss of
ascorbic acid, textural modifications, and other oxidative processes may be determined
at levels below the threshold of sensory perception.92'187
396 CRC Critical Reviews in Food Science and Nutrition

2. Metal Pick-up
Tin pick-up was found to be a suitable criterion for determining shelf-life for plain tin
cans, iron pick-up — for lacquered cans. Data on tin pick-up as function of storage time,
temperature and quality variables of the tinplate, are reported in literature. Simple
extrapolation of a short term pick-up rate yields a satisfactory prediction of the shelf-
life.187

3. Gas-Chromatographic Headspace Analysis


As metal is dissolved into the product, equivalent quantities of hydrogen are released
and taken up in chemical reactions; some of it diffuses out of the can, and some
accumulates in the headspace. The qualitative and quantitative composition of
headspace gases may contribute information on product deterioration, container
corrosion, and bacterial spoilage. Gas chromatography permits rapid analysis of
mixtures of oxygen, hydrogen, carbon dioxide, and carbon monoxide.155'157'188'189

E. Storage Simulation — Food-Package Interaction


Downloaded by [University of Arizona] at 05:39 21 July 2012

1. Model Solutions
A common method for testing inside coatings as affected by thermal processes involves
use of model solutions (e.g., acetic, lactic, and citric acids) of specific aggressivity,
adjusted by regulating the concentration and adding electrolytes (chlorides, nitrates,
sulfides, etc.). This permits simulation of the resistance of the can to acids, alkalis, fats,
sulfur-containing compounds, etc. The potential of the can is continuously measured and
a sudden drop indicates loss of adhesion.78'79

2. Accelerated Storage Tests


Under an elevated storage temperature, chemical and physical reactions in the can are
accelerated and yield some indication of the shelf-life. These results, however, should be
treated with reservation, since the reaction mechanisms themselves may be affected as
well.187 (In fact, it is known that the electrochemical potentials of the component metals
of tinplate shift at increased temperature.)

3. In-Can Potential Monitoring


Potential measurements are useful in determining galvanic relationships with a
reference electrode (a so-called "in-can" electrode) inserted into the can. Its potential vs.
the former may be continuously monitored during the storage period and the readings
used for evaluating the shelf-life.190

4. Couple-Current Corrosivity Testing


The most common and convenient technique (due to Daly191) involves tin, iron, and
solder electrodes embedded in a resin block and immersed in the tested product.
Anaerobic or aerobic conditions are simulated according to test requirements, and
potentials and currents (magnitude and direction) are recorded under steady-state
conditions. From these data, conclusions as to anodic or cathodic behavior of the system
may be drawn. For plain tin cans, the tin electrode has to be much larger than its iron
counterpart, namely 10 mm2 vs. 0.1 mm2. For lacquered cans, the ratio is 1:1.
In more recent versions of the corrosivity test some laboratories dispensed with the
resin block and used metal-sheet electrodes. A special device for lacquered tinplate,
according to Koehler,40 consists of a T made of glass, with discs of the plate bonded to
the arms and a tin wire inserted into the links. It is also useful for evaluating lacquer
adhesion.
Attempts also were made to simulate actual conditions by using whole cans and
Volume 17, Issue 4 397

immersing additional tin, iron, and solder electrodes into the product. Results, however,
are only representative of the individual systems tested.

5. Electrochemical (Polarization) Testing


Measurements of current-potential relations under controlled conditions, may yield
information on corrosion rates, coatings and films, passivity, pitting tendency, etc. The
basic prerequisites for electrochemical testing are:" 2 the corroding sample must be
submerged in an electrolyte; the primary corrosion mechanism must be electrochemical;
and the rate obtained must be the average over the entire specimen.
When a metal specimen is immersed in a corrosive medium, both reduction and
oxidation reactions occur on its surface (typically, the metal is oxidized and the medium
reduced). Thus, the specimen functions as both anode and cathode, with anodic and
cathodic currents flowing on its surface. Since corrosion processes are usually the
consequence of anodic currents, it is preferable, for study purposes, for the specimen to
act as a single type of electrode.
Potentiodynamic polarization techniques for estimating corrosion rates are based on
Downloaded by [University of Arizona] at 05:39 21 July 2012

the assumption that both cathodic and anodic reactions exhibit Tafel behavior, i.e., E = a
+ b log i where E is the electrode potential of the specimen, i is the current density of the
electrochemical reaction, and a and b are constants. Under steady-state conditions the
specimen assumes a rest potential ECOrr. The predominating current is anodic or cathodic
depending on whether polarization direction is positive or negative. Experimentally,
polarization characteristics are evaluated by plotting the applied potential against log
current response. Based on the above, the corrosion current can be determined by one of
the following procedures:"2

1. The specimen is polarized to levels at least 50 mv more negative than the corrosion
potential; current density data are converted to Ecorr by means of a semilog plot (see
Figure 2) and lcon is obtained.
2. Same as above, with anodic polarization.
3. Polarization is applied in both directions, and lcorT is obtained as the intersection of
the Tafel plots. When applying the Tafel slope technique, it should be borne in mind
that interfering electrochemical processes may distort the log i/E plot before the
linear region has been reached.

Another popular technique, generally known as the linear polarization method, is


based on the Stern-Geary equation:

and consists in scanning over a potential interval (AE) very close to Econ (+25 mv or less)
the resulting current (AI) being assumed linear. For this purpose, the two Tafel slopes (ba,
bc) of the given system must be determined in advance. The method was found to yield
reasonable estimates with satisfactory accuracy. It should be borne in mind, however,
that if AE is taken too far from ECOII it may overshoot the linear portion of the
polarization curve; moreover, AE may vary considerably with the type of corroding
system or else the anodic and cathodic polarization curves may be nonsymmetrical.
The experimental polarization system consists of a closed cell with a set of three
electrodes: the reference electrode (R), the counter-electrode (usually made of Pt), and
the working electrode (W) (Figure 8). The cell is purged with an inert gas for 24 to 48 hr to
equilibrium. The measuring device consists of a potentiostat/galvanostat for converting
398 CRC Critical Reviews in Food Science and Nutrition

DIGITAL
GENERATOR
VOLTMETER

- R
POTENSIOSTAT
W RECORDER
R x-v

LOGARITHMIC
CELL AMPLIFIER
Downloaded by [University of Arizona] at 05:39 21 July 2012

FIGURE 8. Schematic description of a polarization system set-up.

the potential difference between the specimen and the reference electrode, and a
generator for regulating the scanning rate. The current between the specimen and the
counter electrode, processed by a logarithmic current converter, is read on an X-Y
recorder with the potential plotted on one axis and log i on the other.
The above techniques are problematic where lacquered cans are concerned, since full-
depth damage to the film may result in data similar to those for unlacquered tinplate.
Another problem is the very small currents involved (because of the small exposed metal
surface), hence the instrumentation must be extremely sensitive.193
To sum up, potentiodynamic polarization techniques are useful in identifying
interrelations between materials and environments, and in predicting how a material will
behave when exposed to a certain medium. However, they are subject to the same
reservation as use of elevated temperatures in accelerated testing, and are no substitute
for long-term studies which may involve different mechanisms.

VII. CONSIDERATIONS IN CHOICE O F TYPE O F CANS

Choice of the type of can most suitable for a given food product is a function of the
following factors: nature of the food; type and cost of cans and facilities available to the
cannery; and shelf-life and other requirements for can performance, dictated by national
marketing conditions and by legislation.34'85
As food products vary in their corrosive effect according to origin, season, and
precanning processes, the suggestions outlined below can at best serve as guidelines and
offer a range of choices for the major groups of products.34
Distinction should be made between products that withstand contact with the tin
coating and may even benefit from it, and those which need a barrier between the metal
and contents. Thus, the first choice the packer has to make is between a plain tin and a
lacquered can. Several products benefit from dissolved tin, which has a strong reducing
action and helps preserve vitamin C and prevent browning in some products. Provided
the dissolved tin content does not exceed a certain level over an acceptable period, the
plain tin can is the most suitable choice.
However, because of new legislation regarding dissolved metals, as well as
considerations of cost and suitability of the product, most foods today are marketed in
lacquered cans. For less aggressive products, internal coatings prevent corrosion
Volume 17, Issue 4 399

altogether and improve the internal appearance of the can. The lacquer coating permits
reduction in tin coating weight and affords a reasonable shelf-life, provided appropriate
technological and storage requirements are observed. For aggressive products with
strong detinning action, lacquers are essential for acceptable shelf-life. When the product
is an iron dissolver, a second lacquering or side seam stripping is applied to ensure a
reliable continuous coating.
Lacquers have solved difficulties in the canning of fruits and juices pigmented with
anthocyanins, which change color as a result of tin dissolution. Dark fruits pigmented
with anthocyanins, such as red plums and cherries, require cans with high tin coating
weights. Beets and pickles in vinegar, and lemon concentrates fall in the same category.
By contrast, pale colored fruits such as citrus, pineapples, peaches, and apricots benefit
from dissolved tin and are marketed in plain tin cans; or, if a very long shelf-life is
required, type K or STP electrolytic plate is specified. In some cases "hot dip" tinplate is
recommended, but is too expensive.
Products which benefit from dissolved tin but are highly corrosive in the presence of
depolarizers — tomato preserves, melon cubes, green beans, asparagus, etc. — are
Downloaded by [University of Arizona] at 05:39 21 July 2012

marketed in lacquered cans with low tin-coating weight; in some cases stannous ions are
added (legislation permitting) or else the "high tin fillet" seam or tin powder techniques
are applied.194
Sulfur staining products such as meat, fish, and corn are commonly marketed in
lacquered cans with low weight tin coating. Beer (where flavor is impaired by traces of
iron) and soft drinks require complete protection, and therefore low tin-coating weight
cans with double lacquering and side seam stripping are used.

VIII. RECENT DEVELOPMENTS AND FUTURE TRENDS


Dynamic change has been the "order of the day" in the can manufacturing industry
since the early sixties. During that period, cans have become lighter, stronger, more
reliable, easier to open, and more attractive in overall appearance. These trends have
been most evident in the beverage can industry.48'195"196
The principal change is the gradual weight decrease in the steel base and tin coating.
Only 20% of the tin and 65% of the steel used 30 years ago are needed today for a standard
can. Other changes include geometric modifications, such as beads and "necked-in"
features, designed to withstand mechanical stresses.197
Still another important technological innovation is the two-piece can, whose
advantages vs. the three-piece include a seamless and lighter one-piece body, thus
eliminating soldering and one-end seaming and, as a result, fewer points of potential
microleakage; prevention of lead migration into the product; larger surfaces for printing
and decoration; and much higher speed of production. This changeover is justified on
grounds of product integrity or marketing considerations. Seamless (DRD) cans for
fruits and vegetables, with a low bottom lead configuration, were introduced into the
American market recently. Two-piece cans are extensively used in the beer and soft drink
industry. Their acceptance for thermally processed foods, however, is much slower, and
three-piece cans will remain dominant there for some time. This is due both to the fact
that saving in materials is not as great as for beer cans, in view of the danger of buckling
under vacuum, which calls for thicker walls and circumferential beads.172
In comparing the DRD and D & I two-piece can, most can makers today are of the
opinion that the latter would dominate in the future, being more economical to
manufacture. It will be reduced in weight in the future by over 20% and still have the
desirable crush resistance and adequate resistance to bottom buckling and burst
pressure.48
400 CRC Critical Reviews in Food Science and Nutrition

Two further fairly new technologies of can production are high-speed (70 m/ min)
welding and cementing (the latter being mostly used for TFS cans), with the resultant
elimination of lead.
Easy-open ends are made of aluminum or tinplate. To alleviate the litter problem,
nondetachable tabs were developed, while detachable pull-tabs were actually banned in
several states in the U.S.
Tinplate is the only material universally suitable for all products in the food-canning
industry, thanks to its electrochemical behavior. Political and economic constraints,
however, led to other materials (such as TFS and aluminum) being introduced. Several
can companies are conducting active programs on use of black plate in D & I cans, which
may prove suitable for products where the cathodic protection usually supplied by tin is
not needed. Examples are beer, meat, and vegetable products.
Significant changes have occurred recently also in the coating technology (especially
for beverage cans) as a result of environmental and air pollution regulations regarding
solvent emission into the atmosphere, as well as of energy-saving considerations. One
development involved use of water as solvent: water-based interior coatings, epoxy and
Downloaded by [University of Arizona] at 05:39 21 July 2012

acrylic base-coats, and vinyl and acrylic spray topcoats are already in use for beer and
beverage cans.198'199 The ultimate coatings are believed to be powders which contain no
solvents (thus no pollution is involved) and require minimal baking for curing.48'197'199
Acceptable powders have been developed, and the resins used have FDA approval. There
are still problems regarding the average powder grain size, film weight, and the economic
aspects of the process. There are two classes of powders: thermoplastic, which require
just enough heat to be fused but whose adhesion is questionable; and thermosetting, with
good adhesion and process resistance.
Mention should also be made of UV curing of exterior printing inks and coatings.
These inks contain no solvents, require no baking (hence energy consumption is reduced
significantly), and there is no pollution. Basically, the process involves a photopoly-
merizable mixture of a low molecular weight unsaturated polymer dissolved in a blend of
reactive diluents. Reaction is initiated by a blend of photoinitiators which induce
polymerization on exposure to UV light.
Use of UV light in curing lacquers and internal coatings is also possible, but will
probably remain limited to external lacquering since the chemicals involved have to
comply with food regulations and will take very long to gain approval."
Future trends in the tinplate industry are: reduction in amount of tin coatings and pro-
duction of low tin coated steel (LTS), i.e., 0.5 to 2 g/m2;200'201 further reduction in steel
thickness, through use of higher tensile strength steel; and use of blackplate, with an
organic or a phosphate-type inorganic coating.

IX. FINAL REMARKS


Can manufacturing has undergone dynamic changes in the last two decades. These
permit improved strength, reduced can weight and tin coating weights, higher corrosion
resistance, and improved barrier performance of the interior and exterior coatings. The
advent of the two-piece can (as partial substitute for the three-piece can) and of new
welding techniques reduced the incidence of leakage. Increased attention has been given
to safety, to public health considerations (especially regarding dissolved metals), to
energy consumption, to pollution problems, and to availability of raw materials.
At the same time, in spite of these changes, corrosion still poses a problem for both the
container and canning industries. Numerous aspects remain unsolved, and new ones
arise with the advent of new processes, products and formulations.
Predicting the suitability and shelf-life of the system container-food medium by
Volume 17, Issue 4 401

different techniques has helped reduce the danger of failures as a result of poor
compatibility.
The present review summarizes the state of the art of published information as regards
existing container usage and problematics, knowledge of the mechanism of corrosion in
plain and lacquered cans, metal container-food medium interaction, and available
techniques of evaluation. It should be remembered that the industries are in possession of
extensive unpublished information acquired as a result of practical experience and
research.
The closing note here will refer to Britton's remark in his excellent review:34 "It should
be remembered that corrosion resistance is not an absolute property of a metal but has to
be defined in relation to the metal and a particular environment." If all the above will be
taken into consideration while canning foods, it is believed that the number of failures
will decrease, and better products with longer shelf-life will be available.

REFERENCES
Downloaded by [University of Arizona] at 05:39 21 July 2012

1. Anon., Instruction for electrodepositing tin, Tin Research Inst., Middlessex, England, 1971.
2. Kamm, G. G. and Willey, A. R., Corrosion resistance of electrolytic tinplate, Corrosion, 17, 77t, 1961.
3. Kamm, G. G., Willey, A. R., and Beese, R. E., Methods for improving corrosion resistance of
electrolytic tinplate, Mater. Prol., 3, 70, 1964.
4. Kamm, G. G. and Krickl, J. L., K plate for heavily coated electrolytic tinplate applications, 8th
Annual AIME Conf. on Metalworking and Steel Processing, Pittsburgh, U.S.A., 1965, 1.
5. Castell-Evans, J. V., Modified alloy layers in tinplate, Trans. Inst. Met. Finish., 47, 71, 1969.
6. Endlé, J., Properties and production of grade K tinplate and its performance in canning trials,
Br. Corros. J., 7, 216, 1972.
7. Britton, S. C. and Bright, K., Influence of area of the steel component on behavior of a tin-steel
couple, Corrosion, 17, 98t, 1961.
8. Kamm, G. G. and Willey, A. R., The electrochemistry of tinplate corrosion by acid foods, in 1st Int.
Congr. on Corrosion, London, 1961, 493.
9. Maercks, O. and Maercks, I., Einfluss der fullgut aggressivitat die korrosion lackierter weissbleck-
dosen, Verpack. Rundsch., 20, 6, 59, 1969.
10. Mc Farlane, D., Corrosion studies of tin free steels in food canning application, in La Corrosion
des boites metalliques destinees a l'industrie alimentaire, Symp. Int., Liege, 24, 1971.
11. Azzerri, N. and Baudo, G., Comparative corrosion resistance of Tin-Free Steels chromium type to
synthetic and natural food media, Br. Corros. J., 10, 28, 1975.
12. Kamm, G. G., Willey, A. R., and Linde, N. J., Surface and corrosion characteristics of Tin Free
Steel chromium type for beverage containers, J. Electrochem. Soc., 116, 1299, 1969.
13. Nehring, O., Corrosion problems of metallic food cans in, La corrosion des boites metalliques
destinees a l'industrie alimentaire, Symp. Int., Liege, 22, 1971.
14. Britton, S. C., Electrochemical assessment of chromium in passivation films on tinplate, Br. Corros. J.,
1, 19, 1965.
15. Britton, S. C., Examination of the layer produced by chromate passivation treatments of tinplate,
Br. Corros. J., 10, 85, 1975.
16. Uhlig, H. H., Corrosion and Corrosion Control, John Wiley & Sons, Inc., New York, 1971, chap. 5.
17. Tin Mill Products, Commercial Dept., U.S. Steel Co., 1976.
18. Vrijburg, H. G., Spruit, A. C., and Soepenberg, E. N., Investigation on the kinetics and mechanism of
the formation of cathodic dichromate passivation films on electrolytic tinplate, Estel, Ber. Forsch.
Entwicklung Werke, 1, 39, 1975.
19. Rocquet, P. and Aubrun, Ph., Contribution a l'etude des films de passivation du fer blanc,
Corrosion Trait. Prol. Finit., 16, 229, 1968.
20. Salm, D., The composition of passivation films on tinplate, B.H.P Tech. Bull., 16, 1, 2, 1972.
21. Rauch, S. E. and Steinbecker, R. N., Mechanism of cathodic dichromate passivation of tinplate,
Plating (Paris), 62, 246, 1975.
22. Aubrun, Ph. and Rocquet, P., The mechanism of metallic chromium electrode-position in the cathodic
dichromate treatment of tinplate, J. Electrochem. Soc., 122, 861, 1975.
23. Aubrun, J. and Pennera, G. A., Coulometric determination of the metallic chromium content in
tinplate passivation film, in 1st Int. Tinplate Conf., London, 1976, paper No. 25.
402 CRC Critical Reviews in Food Science and Nutrition

24. Coad, J. P., Mott, B. W., Harden, G. D., and Walpole, J. F., Nature of chromium on passivated
tinplate, Br. Corros. J., 11, 219, 1976.
25. Leroy, V., Servais, J. P., Habraken, L., Belgium, C. R. M., Renard, L., Lempereur, J., and
Cockerill, S. A., Secondary ion mass analysis, Auger and photoelectron spectrometry of passivation
layers on tinplate, in 1st Int. Tinplate Conf., London, 1976, paper No. 34.
26. Maeda, S., Asai, T., and Sawairi, T., Structure of passivation films on tinplate in 2nd Int.
Tinplate Conf., London, 1980, paper No. 25.
27. Pappert, W., Die Bestimmung von Chrom-Chromoxid und Zinnoxid auf Weissblech, ESTEL,
Ber. Forsch. Entwicklung Werke, 4, 178, 1976.
28. Gabe, D. R., Mixed oxides films on tin, Surf. Technol., 5, 463, 1977.
29. Albu-Yaron, A., Application of transmission microscopy to the case of structural surface studies of
tinplate, in 2nd Int. Tinplate Conf., London, 1980, paper No. 22.
30. Maeda, S. and Asai, T., Structure of passivation film on tinplate, in 2nd Int. Tinplate Conf.,
London, 1980, paper No. 25.
31. Takano, H. and Watanabe, T., Effect of passivation treatment on lacquer adhesion of tinplate, in
2nd Int. Tmplate Conf., London, 1980, paper No. 36.
32. Carter, P. R., Some factors affecting the surface chromium content of electrochemically treated
tinplate, J. Electrochem. Soc., 108, 782, 1961.
33. Britton, S. C. and Hanox, J. H., Assessment of stain resistance produced by commercial passivation
Downloaded by [University of Arizona] at 05:39 21 July 2012

treatments of tinplate, Sheet Met. Ind., 1, 15, 1965.


34. Britton, S. C., Tin Versus Corrosion, Publ. No. 510, Int. Tin Res. Inst., Middlessex, England, 1975.
35. Marsal, P., Influence des produits phytosanitaires sur la corrosion des boites en fer blanc, Ind.
Aliment. Agric., 369, 1972.
36. Lewis, W. R., Notes on soldering, Tin Res. Inst., Middlessex, England, 1968.
37. Morris, H., Solder for can making, Tin-Printer & Box Maker, March, 4, 1971.
38. Norman, G. F., Welding of tinplate containers — an alternative to soldering, in 1st Int. Tinplate
Conf., London, 1976.
39. Andres, C., Lead-free, welded side seam cans have improved integrity, Food Process., March, 60,
1980.
40. Koehler, E. L., Korriosionsprozess durch und unter organischen uberzugen, Werkst. Korros.,
21, 454, 1970.
41. Collenteur, H. A. P., The phenomenon of corrosion concentration on small defects of internal
lacquered cans, in La Corrosion des Boites Metalliques destinees a l'industry Alimentaire, Symp.
Int., Liege, 1971, 83.
42. Koehler, E. L., The influence of contaminants on the failure of protective organic coatings on
steel, Corrosion, 33, 209, 1977.
43. Koehler, E. L., Corrosion under organic coating, in The U.R. Evans Conf. on localized corrosion,
Williamsburg, U.S.A., 1971, paper No. 51.
44. Sherlock, Y. C., Factors affecting the undermining resistance of lacquers on tinplate, in 1st Int.
Tinplate Conf., London, 1976.
45. Pilley, K. P., Lacquers, varnishes and coatings for food cans, Arthur Holden & Sons. Ltd., 1968.
46. Stroad, R. C., Waterborn coatings in metal packaging, in 1st Int. Tinplate Conf., London, 1976.
47. Anon., Can coatings: waterbased winning compliance battle, Modern Metals, 35, 96, 1979.
48. Wilson, R. T., State of the can making art, in 1st Int. Tinplate Conf., London, 1976, paper No. 4.
49. Carter, P. R., Lewis, L. L., and Murray, M. V., Surface properties of drawn-and-ironed tinplate
containers, in 1st Int. Tinplate Conf., London, 1976, paper No. 12.
50. Fidler, F., Two piece container development — some influences on tin, in 1st Int. Tinplate Conf.,
London, 1976, paper No. 15.
51. Bombara, G., Azzerri, N., and Baudo, G., Electrochemical evaluation of the corrosion behavior of
tinplate, Corros. Sci., 10, 847, 1970.
52. Barnartt, S., Electrochemical nature of corrosion, in Electrochemical Techniques for Corrosion, Natl.
Assoc. Corros. Eng. Meeting, Houston, Texas, 1976, 1.
53. Stern, M. and Geary, A. L., Electrochemical polarization, J. Electrochem. Soc., 104, 56, 1957.
54. Bockris, J. O'M. and Reddy, A. K. N., Modern Electrochemistry, Plenum Press, New York, 1970.
55. Evans, V., The Corrosion and Oxidation of Metals, Edward Arnold Ltd., London, 1976.
56. Mansfeld, F., Polarization resistance measurements — experimental procedure and evaluation of test
data, in Electrochemical Techniques for Corrosion, Natl. Assoc. Corros. Eng. Meeting, Houston,
Texas, 1976, 18.
57. Hartwell, R. R., Certain aspects of internal corrosion in tinplate containers, Ad. Food Res., Vol.
III, Academic Press, New York, 1951, 327.
58. Sherlock, J. C. and Britton, S. C., Complex formation and corrosion rate for tin in fruit acids,
Br. Corros. J., 7, 180, 1972.
Volume 17, Issue 4 403

59. Sherlock, J. C., Hancox, J. H., and Britton, S. C., Rate of dissolution of tin from tinplate in
oxygen-free citrate solutions, II. Effect of coating porosity, Br. Corros. J., 7, 227, 1972.
60. Mahadeviah, M., Internal corrosion of tinplate containers with food products, Ind. Food Pac., 30, 2,
9, 1976.
61. Cheftel, H., Monvoisin, J., and Swirsky, M., Observation on the internal corrosion of tinplate cans
by acid foodstuffs, J. Sci. Food Agric., 11, 652, 1955.
62. Patrick, G. W., Internal corrosion of tinplate food containers, Anti-corrosion, June 9, 1976.
63. Serra, G. and Perfetti, G. A., The diffusion of hydrogen through tinplate containers packed with
grapefruit juice, Food Technol., 17, 114, 1963.
64. Häggman, J., The role of hydrogen in the corrosion of tinplate cans, in 2nd Int. Tinplate Conf.,
London, 1980, paper No. 34.
65. Willey, A. R., Effect of tin ion complexing substances on the relative potentials of tin, steel and
tin-iron alloy in pure acid and food media, Br. Corros. J., 7, 29, 1972.
66. Bakal, A. and Mannheim, C. H., The influence of processing variants of grapefruit juice on the rate
of can corrosion and product quality, Israel J. Technol., 4, 262, 1966.
67. Saguy, I., Mannheim, C. H., and Passy, N., Some factors influencing corrosion of canned acid foods,
J. Food Technol., 8, 147, 1973.
68. Landau, M. and Mannheim, C. H., Evaluation of electrolytic tinplates for juice containers, J. Food
Technol., 5, 417, 1970.
Downloaded by [University of Arizona] at 05:39 21 July 2012

69. Kumins, C. A., Electrochemical properties of protective coatings, Off. Dig., 34, 843, 1961.
70. Kumins, C. A. and London, A., The electrochemistry of vinyl polymer membranes J. Polym. Sci.,
46, 395, 1960.
71. Maitland, C. C. and Mayne, J. E. O., Factors affecting the electrolytic resistance of polymer film,
Off. Dig., 34, 972, 1962.
72. Mayne, J. E. O., The Mechanism of protection by organic coatings, in Trans. Inst. Met. Finish., 41,
121, 1964.
73. Kosek, V., Relation of lacquer coat thickness and of fundamental features of tinned plate to its
corrosion resistance, Obaly, 12, 1, 6, 1966.
74. Maercks, O., Wege zur erhöhung der Korrosionssicherheit Lackierter Weissblech packungen, Mitt
Dtsch. Forsch., Blechverarbeitung, No. 4/5, 1, 1966.
75. Nehring, P., Lebensmittelrechtliche Aspekte von Migration und Korrosion im Hindlbick auf Metall-
verpackung, in Int. Konf. Schutz Verderblicher Guter Verpackung, Munich, Germany, 1976, 138.
76. Harden, G. D. and Walpole, J. F., Influence of the alloy layer on the performance of lacquered
tinplate for food cans, in 1st Int. Tinplate Conf., London, 1976, paper No. 30.
77. Maercks, O. and Maercks, I., Weitre Untersuchungen uber den Enifluss der Fullgut-aggressivitat
auf die Korrosion Lackierter Weissblechdosen, Verpack. Rundsch., 22, 4, 29, 1971.
78. Maercks, O., Weitere Untersuchungen zur Wechselwirkung Zwichen Metall und Lackierung bei der
Korrosion Lackierter Weissblechdosen, Verpack. Rundsch., 28, 6, 45, 1977.
79. Maercks, O., Die Wechselwirkung Swischen Metall und Lackierung bei der Korrosion von Weiss-
blechdosen, Verpack. Rundsch., 29, 4, 25, 1978.
80. Evans, U. R., The Corrosion and Oxidation of Metals, St. Martin Press Inc., New York, 1960.
81. Albu-Yaron, A., Berthelin, N., Christen, J. M., and Pagetti, J., A tentative evaluation of the
potentiokinetic polarization technique in studies of localized corrosion of lacquered tinplate, J. Food
Technol., 14, 9, 1979.
82. Albu-Yaron, A., Semel, A., and Berzin, A., A rapid electrochemical test for the assessment of the
delamination of lacquered tinplate, J. Food Technol., 12, 325, 1977.
83. Johannessen, S. J. and Grande, A. P., An Auger analysis of tinplate samples, in 2nd Tinplate
Conf., London, 1980, paper No. 25.
84. Servais, J. P., Lemperer, J., Renard, L., and Leroy, V., Examination by ESCA of the constitution
and effects on lacquer adhesion by passivation films on tinplate, Br. Corros. J., 14, 126, 1979.
85. Marsal, P., Matching tinplate cans to their contents, in 4th Int. Meet. Canning Ind., Varna,
Bulgaria, 1976.
86. Lambeth, V. N., Field, M. L., Brown, J. R., Regan, E. S., and Blevins, D . G., Detinning by
canned spinach as related to oxalic acid, nitrates and mineral composition, Food Technol.,
22, 132, 1969.
87. Mahadevaiah, M., Gowramma, R. V., Radhakrishnaiah, G., Sastry, M. V., Sastry, L. V. L., and
Bhatnagar, H. C., Studies on variation in tin content in canned mango nectar during storage, J. Food
Sci. Technol., 6, 192, 1969.
88. Horio, T., Iwamdo, Y., and Komura, S., Studies on dissolved tin in food products, Repub. Junior
Coll. Food Technol., 9, 1, 1970.
89. Nehring, P., Interaction between the content and canning material in canned vegetables and fruits,
Ind. Obst. Gemueseverwert., 55, 307, 1970.
404 CRC Critical Reviews in Food Science and Nutrition

90. Wunsche, G., Effect of storage temperature on the stability of preserved foods, Inform. Fischwirt.,
17, 170, 1970.
91. Mannheim, C. H. and Hoenig, R., Storage changes in canned comminuted orange and lemon as
affected by type of can, Confructa, 16, 165, 1971.
92. Seelenberger, P. and Luh, B. S., Effect of post canning on chemical changes and brown discoloration
in canned peaches, Confructa, 16, 145, 1971.
93. Catala, R. and Duran, L., Internal corrosion in cans containing green beans, Influence of tinplate
and of processing and storage conditions, Rev. Agroquim Technol. Aliment., 12, 319, 1972.
94. Royo-Iranzo, J. and Grima, R., Tin and lead contamination of commercial orange juices packed in
cans, Rev. Agroquim Technol. Aliment., 13, 436, 1973.
95. Andrae, W., Effect of can materials on the shelf-life and storage stability of food preserves,
I - I I I , Verpackung, 14, 167, 1973; 14, 185, 1973; 15, 14, 1974.
96. Semel, A. and Saguy, M., Effect of electrolytic tinplate and pack variables on the shelf-life of
canned pure citrus fruit juices, J. Food Technol., 9, 459, 1974.
97. Heikal, H. A., El-Dashlouty, M. S., and El-Sidawi, M. H., Influence of storage time and temperature
on the tin concentration of canned Egyptian fruit and vegetable products, Confructa, 21, 93, 1976.
98. Mahadevaiah, M., Gowramma, R. V., Epieson, W. E., and Sastry, L. V. L., Influence of tinplate
variables on the internal corrosion of tinplate containers with mango and orange products, J. Food
Sci. Technol., 13, 17, 1976.
Downloaded by [University of Arizona] at 05:39 21 July 2012

99. Marsal, P. and Dane, J. M., Detinning of tin cans-influence of can size, Cent. Rech. Fer-Blanc Bull.,
15, 1976.
100. Mergey, C. and Hanusse, H., Plain tinplate corrosion by pear halves in syrup, Br. Corros. J.,
12, 103, 1977.
101. Davis, D. R., Cockrell, C. W., and Weise, K. F., Pitting in canned green beans: effect of cultural
practices, tin coating, vacuum, corrosion accelerators and storage conditions, J. Food Sci., 44, 241,
1979.
102. Davis, D. R., Cockrell, C. W., and Wiese, K. F., Can pitting in green beans: relation to vacuum,
pH, nitrate phosphate, copper, and iron content, J. Food Sci., 45, 1411, 1980.
103. Kamm, G. G., Willey, A. R., Beese, R. E., and Krickl, J. L., The alloy-tin couple test — a new
research tool, Corrosion, 17, 84t, 1961.
104. Pourbaix, M., Atlas d'Equilibres Electrochimiques, Gauthier Villars & Cie, 1963.
105. West, J. M., Electrodeposition and corrosion processes, Van Nostrand-Reinhold Co., London, 1971.
106. Yamada, K. and Tanaka, M., Black discoloration of cans containing fish and shell fish, Seikei-Ku
Kenkyusho Kenkyu Hokoku, 55, 3866, 1959.
107. Pigott, G. M. and Dollar, A. M., Iron sulphide blackening in canned protein foods, Food Technol.,
17, 481, 1963.
108. Thompson, M. H., The mechanism of iron sulphide discoloration in cans of shrimp, Food Technol.,
17, 665, 1963.
109. Grau, R., Theoretical basis of chemical interactions between cans and their contents with special
reference to cured pork products, Fleishwirtschaft, 50, 451, 1970.
110. Gruenwedel, D. W., and Patanik, R. K., Release of hydrogen sulphide and methyl mercaptan from
sulphur-containing amino acids, J. Agric. Food Chem., 19, 775, 1971.
111. Hottenroth, B., Discoloration of tinplate cans interior walls through metal sulphide build-up — a
review, Verpack. Rundsch., Nov. 85, 1972.
112. Gruenwedel, D . W. and Hao, H. C., Model studies regarding the internal corrosion of tinplated
food cans, part 3, J. Agric. Food Chem., 21, 246, 1973.
113. Gruenwedel, D. W. and Patnaik, R. K., Model studies regarding the internal corrosion of tinplated
food cans, part 4, Chem. Mikro. Technol. Lebens., 2, 97, 1973.
114. Kolb, H., Sulphur induced discoloration of tinplate cans containing cystein solution and natural
products fills, Verpack. Rundsch., 25, 91, 1974.
115. Board, P. W., Holland, R. V., and Elbourne, R. G. P., The effect of sulfur containing fungicides
on the corrosion of plain cans of foods, J. Sci. Food Agric., 18, 2322, 1967.
116. Board, P. W., Holland, R. V., and Britz, D., Dithiocarbamates, carbon disulfide and the corrosion of
tin plate, Br. Corros. J., 3, 238, 1968.
117. Elkins, E. R., Farrow, R. P., and Kim, E. S., The effect of heat processing and storage on
pesticide residues in spinach and apricots, J. Agric. Food Chem., 20, 286, 1972.
118. Marsal, P., Effect of pesticide residues in fruit and vegetables on the corrosion of tin cans, Ind.
Aliment. Agric., 89, 369, 1972.
119. Wessel, J. R., Pesticide residues in foods, in Environmental Contaminants in Foods, Western
New York State Inst. of Food Technol., Spec. Rep. 9, 6, 1972.
120. Allouf, R., Jarillon, G., and Calas, J., Contribution to the study of the corrosion of tinplate be
apricots: influence of fungicide residues, Br. Corros. J., 12, 187, 1977.
Volume 17, Issue 4 405

121. Seiler, B. C., The mechanism of sulfide staining in tin foodpacks, Food Technol., 22, 91, 1968.
122. Britton, S. C. and Bright, K., An examination of oxide films on tin and tinplate, Mettalurgia, 56,
163, 1957.
123. Akimov, G. W., Electrode potentials, Corrosion, 11, 477t, 1955.
124. Eden, H., Corrosion problems of tin cans, M.Sc. thesis, Technion, Haifa, Israel, 1978.
125. Ashton, M. R., The occurrence of nitrates and nitrites in foods, The British Food Manuf. Ind.
Res. Assoc., Liter. survey No. 7, 1970.
126. Seale, P. E., Symposium-Nitrates in canned foods, the scope of the problem, Food Technol. Aust.,
25, 10, 1973.
127. Walker, R., Naturally occurring nitrate/nitrite in foods, J. Sci. Food Agric., 26, 1735, 1975.
128. Siciliano, J., Krulick, S., Heisler, E. G., Scwartz, H. J., and White, J. W., Nitrate and nitrite
of some fresh and processed market vegetables, J. Agric. Food Chem., 23, 461, 1975.
129. Cheftel, H., and Monvoisin, J., La corrosion des boites des fer-blanc dans l'industry des conserves.,
Etud. J.J. Carnaud Forges Nasse-Indres. Bull., 12, 90, 1954.
130. Strodtz, N. H. and Henry, R. E., The relation of nitrates in foods to tinplate container corrosion,
Food Technol. 8, 93, 1954.
131. Thomas, G., Mergey, C., Leblanc, C., Hanusse, H., Monovoisin, J., and Cheftel, H., Action des
nitrates et du fer dans le desetamage des boites de fer-blanc par des solutions citriques, Comm. Int.
Perm. Cons., Paris, 1966.
Downloaded by [University of Arizona] at 05:39 21 July 2012

132. Farrow, R. P., Charbonneau, J. E., and Lao, N. T., The tinplate producers — CMI-NCA, Research
program on internal can corrosion, Natl. Canners Assoc., Washington, D.C., 1969.
133. Gerson, L., The influence of fertilizer on can corrosion with green beans, in La corrosion de
boites metalliques destinees a l'industrie alimentaire, Symp. Int., Liege, 1971, 1, 103.
134. Farrow, R. D., Research program on can detinning in canned food storage, Food Technol., 24, 44,
1970.
135. Farrow, R., Lao, N. T., and Kim, E. S., The nitrate detinning reaction in model systems, J. Food
Sci., 35, 819, 1970.
136. Hoff, J. E. and Wilcox, G. E., Accumulation of nitrate in tomato fruit and its effect on detinning,
J. Am. Soc. Hort. Sci., 95, 1, 92, 1970.
137. Farrow, R. P., Johnson, J. H., Gould, W. A., and Charbonnean, J. E., Detinning in canned
tomatoes by accumulations of nitrate in the fruit, J. Food Sci., 36, 341, 1971.
138. Horio, T., Iwamoto, Y., and Miyazaki, M., Mechanism of corrosion and influence of tomato
cultivar and other factors on uptake of nitrate, in 6th Int. Cong. on Canned Foods, CIPC, Paris,
1972, 93.
139. Barbieri, G., Bellucci, G., Massini, R., Milanese, G., Pellizziari, A., Aldini, R., and Rosso, S.,
Behavior of product and containers in tomato concentrate storage, Imballaggio, 24, 209, 3, 1973.
140. Massini, R., Corrosion of tinplate by food preserves, part I and II, Imballagio, 25, 224, 1975; 26,
227, 1975.
141. Epieson, W. E. and Sastry, L. V. L., Corrosion of tinplate cans by vegetables — composition of
Ivy Gourd with special reference to corrosion accelerating and inhibiting compounds, J. Food Sci.
Technol., 15, 113, 1978.
142. Marsal, P., La role des nitrates dans la corrosion du fer-blanc, Bulletin 1977, Cent. Rech. Fer-Blanc,
Thionville, 1977.
143. Kolb, H., Uber dei dorrosions-beschleunigende wirkung von nitrat und nitrit im weissblechdosen,
Verpack. Rundsch., 25, 6, 41, 1974.
144. Kolb, H., Electrochemische Untersuchengen zum Einfluss von Nitrat un Nitrit, Werkst. Korros.,
27, 10, 1976.
145. Board, P. W. and Elbourne, R. G. D., Pitting corrosion in plain cans containing acid foods,
Food Technol., 19, 1571, 1965.
146. Horio, T., Iwamoto, Y., and Komura, S., Internal corrosion of cans, IV. Possible mechanism of the
action of nitrate in canned drinks, Shokuhin Eiseigaku Zasshi, 9, 133, 1968.
147. Board, P. W., The chemistry of nitrate-induced corrosion by tinplate, Food Technol. Aust., 25, 15,
1973.
148. Board, P. W., Corrosion control in the food industry, Food Technol. Aust., 24, 582, 1972.
149. Emilsen, W., Wilson, E., and Dieckmann, R., Natural inhibitory mechanisms in canned food
products, Food Technol. Aust., 26, 28, 1974.
150. Board, P. W. and Holland, R. V., Inhibition of nitrate-induced corrosion of tinplate cans, Br.
Corros. J., 4, 162, 1969.
151. Sherlock, J. C. and Britton, S. C., Promotion by nitrates of the dissolution of tin by acids and
its inhibition, Br. Corros. J., 8, 210, 1973.
152. Albu-Yaron, A. and Semel, A., Nitrate-induced corrosion of tinplate as affected by organic acid
food components, J. Agric. Food Chem., 24, 344, 1976.
406 CRC Critical Reviews in Food Science and Nutrition

153. Albu-Yaron, A. and Semel, A., Inhibitory capacity of sodium benzoate for nitrate-induced corrosion
of tinpalte, J. Food Sci., 41, 703, 1976.
154. Catala, J. M., Cabanes, J. M., and Soler, M., Inhibition of tin dissolution in canned vegetables,
in 2nd Int. Tinplate Conf., London, 1980, paper No. 33.
155. Lopez, A. and Krebs, B. S., Determining headspace gas composition in canned foods, Food Technol.,
21, 366, 1967.
156. Mannheim, C. H., Bakal, A., and Reznik, D., Shelf-life of canned grapefruit juice, Methods for
evaluation and factors affecting it, Food Technol., 3, 115, 1968.
157. Catala, R. L., Duran, E., and Primo, E., Internal corrosion in canned apricots, influence of oxygen
initial content, in La Corrosion des Boites Metalliques, Symp. Int. Liege, 1971, 146.
158. Passy, N. and Mannheim, C. H., The effect of deaeration on quality of concentrated grapefruit juice,
in Tropical Foods Chemistry and Nutrition, Vol. 1, 141, Academic Press, New York, 1979.
159. McHardy, J., Electrochemical testing of aluminium in cans, Mod. Packag., 39, 10, 161, 1966.
160. Hottenroth, B., Uber Lagerversuche mit Fisch Vollkonserven, Verpack. Rundsch., 14, Nr. 2, tech.-
wiss. Beilage, S. 9-14. 4. Nr. 4 tech.-wiss. Beilage, S. 25-33, 1963.
161. Cecil, S. R. and Woodroof, J. G., Long Term Storage of Military Rations, Dept. of the Army,
Quartermaster Research and Eng. Command, 1962.
162. Cecil, S. R. and Woodroof, J. G., The stability of canned food in long-term storage, Food Technol.,
17, 639, 1963.
Downloaded by [University of Arizona] at 05:39 21 July 2012

163. Crosby, N. T., Determination of metals in foods, Analyst, 102, 225, 1977.
164. Noller, B. N. and Bloom, H., Methods of analysis for major and minor elements in foods, Food
Technol. Aust., 30, 11, 1978.
165. Anon., Survey of lead in food, working party on the monitoring of foodstuffs for heavy metals,
second report, Ministry of Agric., Fisheries and Food, London, 1972.
166. Farrow, R. P., Survey of lead in canned foods, Natl. Canners Assoc., Washington, 1974.
167. Epieson, W. E. and Paulus, K., Determination of tin and iron in canned foods — a comparative
study using a complexometric and polarographic method, Lebensm. Wiss. Technol., 7, 1, 47, 1974.
168. Catala, R., Duran, L., and L'lacer, J., Contenido en plomo de conservas vegetales, A.T.A., 17, 2,
197, 1977.
169. Thomas, B., Roughan, J. A., and Watters, D. E., Lead and cadmium content of some vegetable
foodstuffs, J. Sci. Food Agric., 23, 1493, 1972.
170. Thomas, B., Edmunds, J. W., and Curry, J., Lead content of canned fruit, J. Sci. Food Agric.,
26, 1, 1975.
171. Rouseff, R. L. and Ting, R. L., Lead uptake of grapefruit juices stored in cans as determined
by flameless absorption spectroscopy, J. Food Sci., 45, 965, 1980.
172. Anon., Two piece, welded cans, Packag. Eng., Feb., 46, 1981.
173. Benoy, C. J., Hooper, P. A., and Schneider, R., The toxicity of tin in canned fruit juices and
solid foods, Food Cosmet. Toxicol., 9, 645, 1971.
174. Cheftel, H., Tin in foods, Joint FAO/WHO Food Standards Programme, 4th Meeting of the Codex
Committee on Food Additives, 1967.
175. Browning, E., Toxicity of Industrial Metals, 2nd Ed., Butterworths, London, 1969.
176. De Groot, A. P., Subacute toxicity of inorganic tin as influenced by dietary levels of iron and
copper, Food Cosmet. Toxicol., 11, 955, 1973.
177. De Groot, A. P., Feron, V. J., and Til, H. P., Short-term toxicity studies on some salts and
oxides of tin in rats, Food Cosmet. Toxicol., 11, 19, 1973.
178. Nehring, V. P., Zinn in Pfirsichekonserven, Ind. Obst. Gemuseverwert., 8, Sept., 1972.
179. Roe, F. J. C., Boyland, E., and Millikan, K., Effects of oral administration of two tin compounds
to rats over prolonged periods, Food Cosmet. Toxicol., 3, 277, 1965.
180. Williams, E. V., New techniques for the digestion of biological materials — application to the
determination of tin, iron and lead in canned food, J. Food Technol. 13, 367, 1978.
181. Frankenthal, R., Butler, T. J., and Davies, R. D., Coulometric reduction of oxides on tinplate,
Anal. Chem., 30, 441, 1958.
182. Britton, S. C. and Sherlock, J. C., Examination of oxides on tin surfaces by cathodic reduction,
Br. Corros. J., 9, 1974.
183. Feret, F. and Polaczek, T., Estimation of the tin coating thickness and of the chromium content of
passivation films on tinplate by X-ray fluorescence, Br. Corros. J., 16, 46, 1981.
184. Carter, P. R. and Butler, T. J., Accelerated corrosion test for tinplate in grapefruit and other
juices, Corrosion, 17, 72t, 1961.
185. Lempereur, J. and Biston, R., General information on lacquer testing, Bull. INACOL, 23, 285, 1972.
186. Kleniewski, A., Polarization resistance measurements as a guide to the performance of lacquered
tinplate, Br. Corros. J., 10, 91, 1975.
Volume 17, Issue 4 407

187. Habenicht, G., Rapid test methods for characterizing product container reactions with tinplate and
aluminum containers, Seminar on Food Processing and Packaging, Guatemala, UNIDO ID/WG.
152/1, 1973.
188. Vosti, D. C., Hernandey, H. H., and Stran, J. B., Analysis of headspace gases in canned foods by
gas chromatography, Food Technol., 15, 29, 1961.
189. Hoenig, R., Reznik, B., and Mannheim, C. H., Note on a modified device for headspace
evaluation of cans, J. Food Technol., 1, 363, 1966.
190. Reznik, D. and Mannheim, C. H., In-can measuring of electrochemical corrosion, Mod. Packag., 39,
127, 1966.
191. Daly, J. J., Corrosivity tester a-new tool for packaging, Packag. Eng., 10, 9, 96, 1965.
192. Dean, S. W., Electrochemical methods of corrosion testing, in Electrochemical Techniques for
Corrosion, Symp. at the NACE Corrosion, Houston, Texas, 52, 1976.
193. Beese, R. E. and Allman, J. C., Application of electrochemical techniques to the evaluation of the
corrosion barrier properties of modern container coatings, in ASC Symp. Series. No. 78, Modern
Container Coalings, 1978, 91.
194. Warwick, M. E., Role of tin powder incorporated into lacquers for tinplate containers, Br. Corros. J.,
12, 247, 1977.
195. Barry, B. T. K., Tinplate 1978 — an international perspective, Seminar: Innovations in Tinplate,
Australia, 1978.
Downloaded by [University of Arizona] at 05:39 21 July 2012

196. Seale, P. E., The processors perspective of the tinplate container, Seminar: Innovations in Tinplate,
Australia, 1978.
197. Habenicht, G., Tinplate containers in a changing world of technology, 1st Int. Tinplate Conf.,
London, 1976, paper No. 10.
198. Falkenberg, H. R. and McGuiness, R. C., New developments in interior can coating, lacquers and
some possible applications, in 2nd Int. Tinplate Conf., London, 1980, paper No. 38.
199. Guerrier, J., The changing scene of tinplate coatings and lacquers, in 2nd Int. Tinplate Conf.,
London, 1980, paper No. 37.
200. Salm, D., Towers, A., and Kaan, D., The development of a low tin coating mass tinplate: LTS, in 2nd
Int. Tinplate Conf., London, paper number 12, 1980.
201. Kuroda, H., Onada, I., Inui, T., and Kondo, Y., Characteristics of lightly tin-coated steel sheet, in
2nd Int. Tinplate Conf., London, Paper number 13, 1980.

Vous aimerez peut-être aussi