Tioz Pigment Technology: A Review: Juergen H. Braun, Andrejs Baidins and Robert E. Marganski Du

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Progress in Organic Coatings, 20 (1992) 105-138 105

TiOz pigment technology: a review

Juergen H. Braun, Andrejs Baidins and Robert E. Marganski


Du Pant Chemicals, Research and Development Division, Ja&son Laboratory,
Wilmington, DE 19898 (US!)

Contents

1 Introduction .............................................. 106


2 Background. .............................................. 106
2.1 Industrial milestones. ..................................... 106
2.2 Reviews .............................................. 107
3 Manufacture .............................................. 107
3.1 The sulfate process. ...................................... 107
3.2 The chloride process. ..................................... 109
3.3 Surface treatments ....................................... 110
3.4 Grinding .............................................. 111
3.5 shlrry ............................................... 112
4 Product characteristics ....................................... 113
4.1 Pigment grades ......................................... 113
4.2 Crystal phase .......................................... 114
4.3 Composition ........................................... 115
4.3.1 Uncoated pigments .................................. 115
4.3.2 Coated pigments .................................... 115
4.3.3 Slurries .......................................... 116
4.4 Particle size ........................................... 116
4.5 Surface .............................................. j17
4.6 Color ................................................ 119
4.7 Hazards .............................................. 119
5 Performance .............................................. 120
5.1 Optics ............................................... 120
5.1.1 Light scattering. .................................... 120
5.1.2 Measurement ...................................... 121
5.1.3 Gloss ........................................... 122
5.2 Dispersibility. .......................................... 123
5.3 Durability ............................................. 124
5.3.1 Catalytic activity .................................... 125
5.3.2 Control .......................................... 126
5.3.3 Measurement ...................................... 126
5.4 TiOz substitutes. ........................................ 127
6 Applications .............................................. 129
6.1 Coatings. ............................................. 129
6.2 Plastics .............................................. 130
6.3 Paper ................................................ 131
6.4 Ink ................................................. 132
6.5 Other ............................................... 133
6.5.1 Pigmentary uses .................................... 133
6.5.2 Non-pigmentary uses of pigment. ......................... 133
References ................................................. 133

0033-0655/92/$5.00 0 1992 - Elsevier Sequoia. AU rights reserved


106

1 Introduction

In recent decades, TiOa has replaced all other materials as a source of


light scattering in coatings, plastics and inks because appropriately sized
TiOz particles provide more light scattering and at lower cost than other
compositions of matter. The light scattering performance of TiOa is unique
because its r-utile phase has the highest refractive index of colorless, stable
substances. Furthermore, pigmentary TiOz is inert, non-toxic, thermodynam-
ically stable and less costly than most pigmentary sized particulates. Thus
TiOa pigment provides the lowest cost bright hiding and bright opacity for
most applications.
This paper considers peer-reviewed literature published since individual
subjects were summarized in the review articles cited below. Because pertinent
literature is sparse, reviews are supplemented by personal insights, necessarily
subjective and incomplete. The non-peer-reviewed literature is more extensive
but does not usually communicate new findings.

2 Background

2.1 Industrial mil.eston.es


1920s Beginnings of the TiOa pigment industry
Extended TiOa pigments produced.

1930s The TiOa industry begins to grow rapidly


Anatase pigments, made by the sulfate process, begin to compete with
lithopone and white lead.

1940s Manufacture of r-utile pigment


Rutile pigments, made by seeding the sulfate process, begin to replace
anatase pigments in coatings and plastics applications because of their
higher hiding and better chalk/fade resistance.

1950s Chloride process


The chloride process introduces pure r-utile pigments that are much
brighter and offer bluer undertone.

1960s Durable grades


Silica encapsulation of rutile pigments qualifies white and pastel paint
6lms for demanding outdoor applications.

19 70s Aqueous dispersions


Aqueous pigment slurries provide pigment of stir-m ease of dispersion
for waterborne coatings.
107

1980s Commodity phase of the industry


Most manufacturers offer specific products for awide range of performance
requirements and begin to compete on basis of product quality and
service.
2.2 Reviews
The progress of the technology is summarized in the books and articles
listed in refs. l-l 1, respectively.

3 Manufacture

The TiOa industry is secretive and competitive. Progress in the manufacture


of TiOz pigment is documented in a large number of patents. But even for
insiders, the signif?icance of any individual patent to the progress of the
industry is all but impossible to gauge. Accordingly, the manufacturing section
of this review is limited to a brief overview.
TiOa pigment is manufactured in four stages:

1 Titanium ore is converted to a purified intermediate, either aqueous


titanyl sulfate solution or anhydrous titanium tetrachloride.
2 The intermediate is converted into crystalline, size-optimized pigment
particles.
3 The surface of the pigment particles is coated by aqueous precipitation
techniques.
4 The pigment is Cnished by grinding with or without dry after-treatments.

TiOa pigments are produced from a variety of ores by either of two


processes, the older sulfate and the newer chloride process. Sulfate is a
relatively low-technology, batch manufacture; chloride a high-technology,
continuous process. Size-optimized pigment particles from either process are
surface-treated from an aqueous environment with inorganic oxides and
hydroxides, generally in a batch mode. The product is filtered, washed and
dried. The dried pigment can receive a mushing treatment with organics and
must then be ground and packed or slurried with appropriate dispersants.
The discharges of the TiOa processes - soluble iron compounds, dilute
acids and miscellaneous inorganic contaminants of the ore - have become
international environmental issues, decisive to the viability of existing plants.
The costs of waste disposal have been responsible for large increases in the
manufacturing costs of the pigment.

3.1 The sulfate process (see Fig. 1)


The sulfate process employs aqueous chemistry. Titanium-bearing ore
is dissolved in sulfuric acid, and the resulting titanyl sulfate is hydrolyzed
to hydrous titania, which is calcined to grow titanium dioxide crystals to
pigmentary dimensions. The agglomerates of pigment crystals are preground,
surface-treated and fluid-energy milled to develop optical properties, durability
and dispersibility.
108

Alummum Sails

Fig. 1. Flow charts of titanium dioxide manufacturing processes.

The conversion from ore into pigment begins with the digestion of
ilmenite or titania-rich slag in hot, concentrated sulfuric acid fortified with
sulfur trioxide. Reaction conditions are kept reducing, usually by the addition
of scrap steel to keep iron and other impurities in their more soluble, lower
oxidation states. Rutile and anatase ores dissolve too slowly for this process.
The resulting titanyl sulfate solution is hydrolyzed thermally to precipitate
out hydrous titanium dioxide, containing 7% adsorbed sulfuric acid. Straight
hydrolysis yields only anatase on calcination. To obtain the rutile pigment,
rutile seed crystals (generated by alkaline hydrolysis of titanyl sulfate or
titanium tetrachloride) are added during the hydrolysis step.
The high-surface-area hydrous titanium dioxide adsorbs colored impurities
from the ore. For a whiter product, they are washed out with dilute acid.
Having Ti 3+ dissolved in the acid helps washing by keeping impurities in
their lowest, more soluble oxidation states, and by exchanging the adsorbed
impurity for innocuous Ti3+. All impurities except niobium and phosphorus
can be removed in this way.
109

The 0.05-pm hydrous titania crystals are grown to 0.2 pm r-utile or


anatase crystals of light-scattering dimensions in a kiln. The product enters
at about 350 “C, where water vapor and adsorbed sulfuric acid are driven
off, and exits at about 900 “C, where most of the crystal growth and phase
conversion takes place. The niobium impurity (about 0.2% as Nb205) in-
corporates itself on calcination into the titanium dioxide crystal. Since titanium
has four valence electrons and niobium five, the extra electron enhances the
semiconductor properties of the crystal, resulting in light absorption with a
consequent degradation of color. This effect is very prominent in r-utile
pigments. To restore the white color, the kiln feed is doped with trivalent
aluminum to compensate for the fifth electron of niobium.
Phosphorus plays an important and desirable role in crystal growth
regulation [ 121. It is not incorporated into the crystal structure, but remains
on the particle surface. As the individual titania particles grow and coalesce,
the surface area decreases and the phosphate concentration on the surface
increases. At a critical particle size the phosphate layer blocks further growth,
thus all crystals reach the same diameter. Adjustment of the hot-end tem-
perature of the kiln, and the addition of a mixture of sodium and potassium
sulfates as secondary growth regulators, enables the manufacturer to optimize
particle size for maximum light scattering.
The pigment leaves the kiln as large lumps of strongly aggregated crystals.
The discharge is cooled slowly to relax the crystal lattice, otherwise the
pigment may become off-color or even photochromic. For further processing
to improve durability, dispersion and optical properties, the lumps are ball-
milled into constituent crystals and aggregates to form a line suspension
suitable for wet treatment.
Besides pure pigment, a composite titanium dioxidwalcium sulfate
mixture was once a major product. Its production utilized waste sulfuric
acid, cutting costs and decreasing pollution. The advent of emulsion paints
eliminated composite pigments from the market: the growth of gypsum
crystals in water and the coagulation of the vehicle emulsion by calcium
ions degraded their shelf stability [ 131.
The sulfate process was thoroughly reviewed in the last edition of
Barksdale’s Titanium in 1966 [2]. Two short reviews have been published
meanwhile [ 14, 151. Recent advances in sulfate process technology have
been limited to the recycling and disposal of its major waste products: dilute
sulfuric acid and ferrous sulfate. To avoid the emission of sulfur oxides from
the kiln, a hydrothermal process for growing hydrous titania to crystals of
pigmentary dimensions has been extensively investigated [16-l 81.

3.2 The chloride process (see Fig. 1)


The chloride process, commercialized by Du Pont in the early 195Os,
offers waste disposal, energy and quality advantages over the sulfate process
[ 19, 201. The process generates i-utile crystals in a flame of titanium
tetrachloride and oxygen. The product has a narrower particle size distribution
than sulfate r-utile [21]_ The process is more energy- and materials-
110

efficient since chlorine is recycled. The waste itself is less environmentally


invasive.
The flow chart shows the five process steps to rutile titanium dioxide
pigment. They are (step 1 - optional depending on ore purity) beneficiation
of the ore to a level which will allow it to be used in chlorination (step 2).
Next, the solid blow-over and iron chlorides are separated from the gaseous
exit stream of the chlorinator (step 3). Distillation (step 4) of the TiC14
permits the removal of practically all impurities. Finally, the purified TiCl,
is oxidized to rutile via a large, controlled flame reactor (step 5).
The following chemical reactions are involved:
TiOa (ore) + 2Cla + C - TiCl, (impure gas) + CO,

TiC14 (impure gas) - TiC1, (pure liquid)

TiC14 (pure gas) + Oa (gas) - TiOs (pure solid) + 2C1,


Natural r-utile, leucoxene (a weathered ilmenite) and slag can be chlorinated
without beneficiation. Ilmenite can be processed by simultaneous beneficiation
and chlorination of the ore. Synthetic r-utile can be obtained from ilmenite
after pyrometallurgical treatment or by selective chlorination in a fluidized
bed [22].
Pulverized and dried coke and ore are fed into the chlorinator and
fluidized with chlorine at a temperature from 900-1700 “C. The chlorination
is exothermic. Thus, the reactor must be cooled internally and externally.
The exit stream of the chlorinator contains TiCl,, miscellaneous metal chlorides
and blow-over solids (coke, ore and gangue). Blow-over solids and most of
the metal chlorides are removed from the gas stream by various separation
techniques [ 23 1.
The gaseous TiC14 is condensed and further purified by distillation and
chemical treatments. Impurities can be reduced to typically about lo-20
ppm. In contrast, the sulfate process has no easy way to remove the
contaminants [ 15 1. TiCL, is then oxidized in a reactor with a geometry designed
to minimize pluggage and yield the desired particle size. The patent literature
lists numerous reactor designs. Since the oxidizing reaction is only slightly
exothermic, the reactants must be preheated to sustain the reaction.
Co-oxidants such as Al& are added to the TiC14 stream to promote
the formation of r-utile and to moderate pigment aggregation. The chloride
process can also be made to produce anatase; however, the method is not
commonly used.
Fina.lly, the TiOZ is collected by a suitable bag house or cyclone. Most
of the chlorine released in oxidation is recycled to the chlorination reactor.
Residual chlorine associated with the solid TiOZ is removed by aqueous
hydrolysis. In some cases, the pigment is calcined in a kiln to hydrolyze
chlorides and drive off residual chlorine.
3.3 Suflace treatments
Titanium dioxide pigments are coated in water by either precipitation
or adsorption of various mate&& onto the uncoated base. The most common
111

additives are silica and alumina. Other hydrous inorganic oxides used include
zirconium, tin, zinc, cerium and boron. More elements have been reported.
Uncoated pigment from the chloride or sulfate process is slurried in water.
Residual chlorine is reduced and residual acid neutralized. Sometimes, the
base pigment is treated with a dispersing agent as part of a premilling step,
preceding wet treatment.
Uncoated chloride pigment generally contains c. 1% alumina. Most of
this is at the surface of the rutile crystal. The aluminum oxide is generated
from aluminum trichloride injected into the oxidation reactor to promote
the formation of r-utile and to moderate particle agglomeration.
Inorganic oxides are precipitated from solution by the hydrolysis of
acidic or basic salt. Generally, hydrochloric or sulfuric acid is used to
neutralize sodium silicate and sodium aluminate. TiOa surface coatings can
be applied in distinct layers, coprecipitated as mixed metal oxides or a
combination of sequential precipitations and partial redissolutions between
surface treatments. The surface precipitation on sulfate TiOa may be different
from precipitation on chloride pigment.
Hydrous silica and alumina exist in various phases. If sodium aluminate
or aluminum sulfate are added to a neutral slurry, an amorphous gel will
form initially. When the coordinately bound water in the gel is eliminated
on standing, a pseudo-boehmite structure results. Pseudo-boehmite crystallizes
in a ribbon-like morphology [ 241. These ribbons can enhance dispersibility
and gloss.
The ratio of the various forms of silica and alumina will depend on the
starting reagents, the ionic strength, pH and temperature. Mixing, sequence
and rate of reactant addition, and slurry solid concentration will also be
factors. The optimum conditions for good end-use properties are closely
guarded secrets in the TiOa industry.
After the inorganic coating is applied, additional organic or silicone
treatments may be added. Organics can be added in the aqueous phase or
after the pigment is dried. The patent literature also covers less practical
surface treatments. Titanium dioxide pigment can be coated in the vapor
phase, during the oxidation step or in a fluidized bed reactor. However, the
physical nature of these types of coatings is usually quite different than
those applied in aqueous solution. TiOa can also be calcined after wet
treatment and drying to enhance photochemical stability. Hydrothermal surface
modifications have also been reported.

3.4 G&ldin$J
Conventional grinding devices cannot crush large r-utile or anatase
crystals to the uniform submicron size required for efficient light
scattering. Consequently, crystals are grown to the proper size from
precursors in the manufacturing process. Because some degree of
aggregation in the growth process is unavoidable, grinding is necessary
to free the crystals so they can be dispersed efficiently in end-use
applications.
112

The mill that can efficiently break pigment agglomerates into single
crystals of pigmentary dimensions is the fluid-energy mill. Its most popular
variety, the micronizer, consists of a disk-shaped chamber, where steam and
pigment are injected tangentially at supersonic speeds. The incoming pigment
agglomerates hit each other and the outer wall of the micronizer and fracture.
The fragments lose speed and are hit in turn by other particles, resulting
in further attrition.
A fluid-energy mill not only grinds pigment but also classifies particles.
Steam flows from the periphery, where it is injected, to the center, where
it is discharged. The flowing steam drags pigment along, but centrifugal
force keeps the coarser particles to the periphery, so only the finer particles
are removed with the steam. The aerosol then goes to a gas-solid separator
such as a cyclone or a bag filter, where the pigment is recovered.
To achieve a uniform grind and good classiilcation, pigment must flow
at a uniform rate into the micronizer so as not to disturb the flow pattern.
If the pigment is sticky it adheres to the conveying equipment and forms
chunks. Micronizing aids are thus needed to improve the dry flow. Stable
polyhydroxy compounds, such as pentaerythritol, trimethylol propane and
triethanolamine, are suitable.
The fluid-energy mill, though very effective in breaking pigment ag-
glomerates into individual crystals, uses a large amount of high-pressure
steam to grind the pigment. Consequently, pregrinding is widely practiced
to minimize the work needed in the fluid-energy mill. Large agglomerates,
such as sulfate process kiln discharge, are preground in ball mills. A ball
mill is a drum filled with balls which are lifted by the rotation of the drum
to cascade down under gravity and fracture the agglomerates. Once large
agglomerates are broken, grinding becomes inefficient. The lower limit for
ball-milling TiOa pigments is about 7 pm.
Media and sand mills accelerate balls by means other than gravity. In
these devices millimeter-sized grinding media cascade over pigment agglom-
erates, grinding them into individual crystals by attrition. These mills can
reduce pigment agglomerate size to less than a micrometer. The preground
pigment is then surface-treated to improve its pigmentary properties, filtered,
washed, dried and ground in a fluid-energy mill to deagglomerate it to single
crystals and sintered aggregates.
The essentials of fluid-energy milling have been reviewed by Temperley
and Blythe [25 ], and in technical detail by Muschelnautz et al. [26]. Scale-
up theory has been devised by Tanaka [27, 281. F’urther studies have been
published by Menyhart and Miskiewicz [29], Zahradnicek and Loffler [30]
and by Nakayama and Inui [31]. Ball and media mills have been described
by Patton [lo]. Detailed unit operation calculations have been published by
Tanaka [32].

3.5 Slurry
Aqueous, stabilized and pumpable slurries were introduced by Du Pont
in the early 1970s to eliminate the need for dispersion operations by the
113

customer. The use of slurry also reduces pigment loss and eliminates dusting
and bag handling. Moreover, a slurry facility is readily automated.
The initial shu-ries, based on anatase, were used primarily in paper. A
few years later, rutile slurry was made for use both in paper and in water-
based paint. Slurries contain about 60-80 wt.% pigment depending on the
grade. Dispersing aids are used to stabilize the system during transportation
and storage.
Because slurries must have a shelf-life of weeks, the dispersant formulation
must retard settling by inducing some minimal level of flocculation. This
can be achieved by formulating a secondary minimum in the particle interaction
potential Alternatively, polymeric dispersants can be used for steric stabi-
lization [46].
The slung recipe must also include biocides to inhibit microbe growth
[33]. Microbe colonies tend to adapt to a given biocide. Hence, it is common
practice to alternate biocides.

4 Product characteristics

The function of any pigment is to modify the incident light. Ti02 pigment
serves this function by scattering light within the end-use environment. It
is essential that the pigment is stable, insoluble, safe in its application and
ultimately safe as a waste. Color and particle size are decisive for optical
performance. Composition, crystal phase and surface characteristics are of
lesser importance.

4.1 Pigment grades


TiOz pigments are used in several industries, in a wide variety of
applications and in many binder, plastic and fiber environments. A single
TiOa grade could serve all these diverse needs but would perform only
moderately well. Optimal performance in any application demands specialized
pigment grades that satisfy the specific requirements of optics, surface
chemistry and dispersion technology.
Grades are designed to serve specific performance requirements, usually
associated with individual industries. They are differentiated by durability,
intended end-use, pigment concentration, undertone and as dry powders or
aqueous slurries. All TiOz pigments contain minor amounts of inorganic
oxides and oxyhydrates, while some have organic additives, dispersants or
conditioners. TiOa/extender mixtures have disappeared from most markets.
Calcium sulfate-extended TiOZ pigments were very popular. They performed
well in solvent-based paints but not in aqueous systems because calcium
sulfate crystals grow in water to unacceptable sizes and traces of calcium
and sulfate ions flocculate vehicle emulsions.
Different pigment grades tend to differ in composition in ways readily
apparent to appropriate analysis. Ti02, either rutile or anatase, is the optically
active ingredient. Other components adapt the pigment to the intended end-
114

use. Almost all commercial pigments are composed of 80% or more Ti02,
most contain ahunina, some more hydrous alumina than others. Slurry grades
are aqueous dispersions, sold by pigment content but dispersed in 20-40
wt.% water with small concentrations of pigment dispersants. Durable grades
contain either silica or other inorganic additives. Flat grades contain more
hydrous oxides of high surface area than enamel grades. Blue undertone
pigments consist of smaller particles than their neutral and red undertone
equivalents.
US Federal and ASTM specifications [ 34 ] distinguish four types of TiOa
pigments by composition, performance, application and TiOa content. This
differentiation into only four types fails to capture the choices that are now
demanded by the consumer. The wrong grade of TiOa pigment for any spectic
application will still be totally inert, be white and hide well. But the inappropriate
pigment will fall short of expectations on performance characteristics of
secondary importance*, such as durability, gloss or rheology.
Minor variations of pigment design are identined by special codes in
addition to grade designations by some manufacturers, others include them
within their grade structure. Involved are accommodations to specific re-
quirements of certain market segments including testing, packaging and
selection criteria. Corresponding performance differences are generally quite
subtle, sometimes detectable, never obvious.

4.2 Crystal phase


Two crystal modifications of titanium dioxide serve as pigments: r-utile
and anatase. Rutile pigments scatter light more effectively than anatase and
are less prone to chalk. Because of the light-scattering advantage, r-utile has
replaced anatase in almost all coatings and plastics applications. In paper,
the advantage of r-utile over anatase is less decisive because the light-scattering
advantage is partially off-set by a density disadvantage. Also, anatase absorbs
less UV light than r-utile and is more compatible with fluorescent brighteners.
Thus rutile is expected to replace most but not all anatase in paper applications.
Pigment, like most materials, sells by weight but scatters light by volume.
The TiOz phases differ in their densities and refractive indices as shown by
the data listed in Table 1. Thus, a pound of r-utile contains fewer but more
effective particles than a pound of anatase although the two phases have
nearly the same ratio of density/refractive index. As a result, r-utile is more
effective than anatase but less so in dry hiding applications where refractive
index is less important.
Rutile absorbs slightly more violet light than anatase and is slightly more
yellow in bulk. However, little, if any, of this yellowness extends into pigment
applications. Titanium dioxide cam-rot undergo phase transitions in the am-

*In marked contrast to most color pigments where an inappropriate grade could fail
grossly in an application. Much or almost all of the color strength can be lost by recrystallization,
flocculation, poor dispersibility or incompatibility of pigment surfactants and modifiers with
application systems.
115

TABLE 1
Density and refractive index data for TiOz crystal phases

Crystal phase Density, p Average refractive Ratio


(g n-J-9 index (n) p/n

rutile 4.3 2.7 1.6


anatase 3.8 2.5 1.5

bience of paint, paper and plastics manufacture or usage. This is in stark


contrast to most organic pigments, many of which phase convert readily,
usually with dramatic loss of optical performance. The TiOa phase composition
can be measured readily and reliably by a conventional X-ray diffractometer.
One per cent of anatase in i-utile and vice versa is detectable. Phase mixtures
of lower concentrations can be analyzed by Raman spectroscopy [35].

4.3 Cornposition
4.3.1 Uncoated pigments
Uncoated pigments contain 93% or more titanium dioxide. Minor con-
stituents, either carried over from the ore or added in the manufacturing
process, can be important in determining pigmentary properties. Trace
constituents are unimportant except for transition metals such as iron,
chromium and vanadium, which degrade color via a semiconductor mechanism

Sulfate pigments retain up to 0.3% niobium pentoxide and 0.3% phos-


phorus pentoxide from the ore. They also contain up to 0.2% alumina, added
to compensate for pentavalent impurities to minimize discoloration. Chloride
process pigments contain practically no unwanted impurities because the
titanium tetrachloride intermediate can be efficiently purified. Chloride process
pigments contain about 1% pyrogenic alumina [ 36 ] added for better process
control and for photochemical stability improvement.
Besides inorganic constituents, most uncoated and many coated pigments
contain up to 0.5% of an organic grinding aid to improve flow in the micronizer
and thus achieve a more uniform grind [37]. The grinding aids are usually
polyhydroxy compounds (for example, trimethylol propane, triethanolamine)
that can undergo some pyrolytic degradation in the fluid-energy mill.

4.3.2 Coated pigments


Nondurable grades contain up to 5% alumina in the form of precipitated
hydrous ahnnina for ease of dispersion. Durable grades have an additional
coating. It usually consists of about 2% silica, but lately other components
have appeared, such as anion (for example, sulfate, phosphate) stabilized
amorphous hydrous alumina [ 38,39 1, coprecipitated hydrous ahrmina-zirconia
[40] or zirconium-tin [41] oxides. These additives are usually present in up
to 1% quantities by weight of the titanium dioxide, calculated as the oxide.
116

Superdurable grades of r-utile pigments are obtained by encapsulating


individual particles in glassy silica sheaths, 4% silica or more by weight of
titanium dioxide, using a method developed by Werner [42]. The integrity
of such a coating is determined by how well it protects pigment from attack
by hot concentrated sulfuric acid.
Paper laminate pigments require outstanding photochemical stability to
resist darkening in light. Such stabilization can be obtained in two ways.
One is to add ceric hydroxide in addition to the usual hydrous aluminum,
silicon and phosphorus oxide coatings to enhance photostability [ 431. The
other way is to add equal amounts of hydrous silica and alumina, with light
stability developed by heating (600-700 “C) after wet treatment [44].
Pigments intended for dry flat paints contain 5-10% hydrous silica and
3-6% hydrous alumina in the form of a porous coating for better spacing
and improved optical efficiency in vehicle-starved formulations [45].

4.3.3 Slurries
Predispersed pigments sold as aqueous slurries contain from 60% to
almost 80% titanium dioxide by weight. They are stabilized with a carboxylic
acid dispersant [46]. Odorless amines are added for pH control [33, 471.
Together with preservatives, about 1% of organic material is present. Pigment
loading in shu-ry is limited by the concentration at which the slurry becomes
too thick to pump. Slurries of lightly treated grades are available at higher
concentration than slurries of heavily treated, dry flat grades.

4.4 Particle size


TiOz pigment particles are submicroscopic with size distributions narrower
than many so-called monodisperse particulates. Primary particles are individual
crystallites, averaging from 0.1 pm to 0.3 pm with a geometric standard
deviation of about 1.3. Aggregates are associations of crystallites sharing
grain boundaries. Agglomerates are associations of crystahites and aggregates,
bonded by weak forces. Agglomerates are usually broken by paint grinding;
aggregates can be broken in some high-intensity equipment. Crystallites do
not break in conventional mills. TiOa pigments are not ordinarily degraded
in paint and plastics manufacturing.
Appropriately ground, pigment dispersions contain < 5 wt.% of particles
smaller than 0.10 and larger than 1.0 pm. The mean particle size of pigment
grades is tailored to the required optical performance. Small particle size
pigment grades are made for applications at low pigment volume concentration
(PVC) and blue undertone; large particle size pigment for higher PVCs and
redder undertones.
Claridge and Craker have discussed size-controlling steps in pigment
manufacture [48]. Effective pigment particle size in an acrylic vehicle was
measured by scattering of visible light in a transmission mode and comparison
with data based on electron microscopy [49]. The particle size of pigment,
vibration milled in alkyd resin, has been examined by Barcucci et al. [50]
117

using tint [under] tones, i.e. the red, blue and green reflectance data of gray
tints.
The size measurement of submicroscopic particles is a science. Each
method yields results unique to itself. Electron microscopy does not show
actual particle dimensions but their projections. Settling methods measure
settling cross-sections, not geometric dimensions. Optical methods measure
features of light scattering that are particularly difficult to translate into
geometric particle sizes of r-utile where the refractive index is extreme, the
birefringence strong and the optical dispersion high [51]. Sieving and elec-
trozone measurements are not applicable to particles as small as those in
TiOa pigments. Cross comparisons between methods require careful allowance
for what it is that the method ‘sees’. At that, some disagreement is likely.
For lack of suitable size standards, many methods can operate only in a
fingerprint mode, i.e. provide precise comparisons of similar particulates of
identical composition but not size data of absolute accuracy.
Paint grind gauges do not respond to pigmentary particle sizes but reveal
the presence of minute quantities of agglomerates composed of millions of
primary particles. The relative, optically effective particle size of TiOs can
be measured with great precision by the carbon black undertone test [ 521
that uses light-scattering sensitivity to particle size by ratioing red and blue
light reflectances of a pigment dispersion in black oil or grease. A similar
method uses tint tone values calculated from red, blue and green reflectance
data of gray paints [ 501.
Particle size measurements with sedimentation analyzers have been
reported by Sennet et al. [53], by Mukhopadhyay and Ramrajkar [54] and
by Kalita et al. [ 551. Murley [56 ] used gravitational and centrifugal settling,
and analyzed TiOz dispersions attained by different milling equipment. Barnett
and Simon [ 5 11 considered the impact of the optical characteristics of rutile
on size analysis by light scattering. The effective particle size of pigment
has been studied by automated, electron microscopic image analysis by
Winker and Dulong [57] and by Hess and McDonald [ 581 for pigment
recovered from an end-use system. Pigment crystallites are too large for size
analysis by line broadening of X-ray diffraction lines.

4.5 Suflace
The surface of titanium dioxide pigment is not just titanium and oxygen.
While pigment particles grow, insoluble components accumulate on the surface.
Those components are either impurities present in the ore and not removed
in the purification process, or additives designed to control crystal structure
and growth, and to regulate agglomeration. This fortuitous surface is not
necessarily suitable for a given end-use application, so it may have to be
modified [59].
Sulfate process pigments emerge from the kiln with a phosphate layer
on their surface, imparting excellent water dispersibility to them. Such
unmodified pigments are used in aqueous applications where photochemical
reactivity is of little consequence, such as paper.
118

Unmodified chloride process pigment surface consists of pyrogenic


aluminum and titanium oxides [36]. The nearly anhydrous nature of this
surface is advantageous in the high-temperature processing of plastics: no
water is present to desorb, form bubbles and blow holes in the extruded
fihS.
Since unmodified pigments tend to cake and flow poorly, they are treated
with up to 0.5% of a micronizing aid, usually a polyhydroxy compound, such
as trimethylol propane (TMP), to improve dry flow [37]. Silicones improve
dry flow even more effectively, but they make the pigment hydrophobic and
unsuitable for some applications, such as in heat-sealing or printable films
[601.
To facilitate dispersion, the pigment surface is modified by precipitating
fluffy hydrous aluminum oxide particles onto it. This lowers the van der
Waals’ forces by several orders of magnitude, resulting in decreased par-
ticle-particle attraction [ 6 11. The effect is analogous to covering wax balls
with line down to keep them from sticking; thus the morphology of the
covering particles is important [62, 631. Even though other oxides (e.g.
silica) can also form fine particles, none aid dispersion as well as hydrous
alumina [64].
If durability is required, pigment particles are encapsulated in an im-
pervious oxide coating to prevent contact between the degradable organic
binder of the paint and the photochemically active titanium dioxide surface:
the better the integrity of the layer, the greater the durability. The encapsulating
agent is usually silica [42], although combinations of amorphous alumina,
zirconia, tin and phosphorus oxides [40, 411 have been found to function
well. To preserve dispersibility, the impervious layer must be under the fluffy
alumina layer.
In very high PVC paints where optical efficiency is decreased by crowding,
a fluffy silica layer is precipitated on titanium dioxide surface as a spacer
to prevent the close approach of pigment particles [45, 651. Here silica has
been found to be the most effective spacer; the alumina precipitated on top
of the silica serves mainly to maintain dispersibility.
The hydrous oxide structure affects the surface area of the pigment.
The calculated specific surface area of 0.2-pm rutile spheres is 7 m2 g-l,
close to the published values for untreated pigments. Precipitated hydrous
alumina has a specific surface area of c. 200 m2 g-i, so each per cent of
alumina adds about 2 m2 g-’ of surface area. The contribution of hydrous
silica to the surface area depends on the method of preparation: the glass
capsule around superdurable pigments contributes little to the surface area,
while each per cent of the fluffy silica precipitated on dry flat grades contributes
several square meters of surface area.
The bulk density (packing in air) of pigment is also affected by the
surface structure, as hydrous oxide coatings prevent the close approach of
pigment particles in proportion to the thickness of the coating. The bulk
density varies from 0.5 g cmp3 for fluffy, heavily coated pigments to almost
1 g cmp3 for uncoated pigments. This indicates a very loose packing: a
119

cubic centimeter of randomly packed r-utile spheres is calculated to weigh


about 2.7 g.
Oil absorption (packing in linseed oil) also depends on pigment surface
structure, except that here the dispersing action of free acid in the oil breaks
down floes, resulting in rather tight packing. Untreated pigment oil absorption
values of c. 12 lb. linseed oil per 100 lb. pigment are what would be calculated
for randomly packed r-utilespheres. The oil absorption of high gloss pigments
ranges from 15 to 20, while that of pigments for dry flat hiding paints range
as high as 35 lb. linseed oil per 100 lb. pigment.
A detailed and well-referenced review of surface chemistry of pigments
and fillers prior to 1983, including titania, alumina, silica and other oxides,
is found in a book by Solomon and Hawthorne [66].

4.6 Color
Titanium dioxide is an almost colorless, intrinsic dielectric which has
some semiconductor properties due to small amounts of contaminant [67].
Rutile has an absorption edge at the violet end of the visible spectrum. This
imparts a slight yellow hue to the solid. The anatase absorption edge is at
a shorter wavelength than r-utile. Thus, anatase is less yellow. For well-
purified pigment, dry powder color does not correlate with end-use color.
TiOs particle size has an effect on the color of tints. Smaller particles
scatter blue light more efficiently [ 681, and as a result, pigment with a smaller
average particle size will give a bluer tint to coatings and plastics (or coatings
at incomplete hiding). Pigment with larger average particle size will appear
redder in the same systems.
The chloride process removes most ore components that cause color.
The most common impurities, iron and vanadium, are typically reduced to
less than lo-20 ppm.
Dry TiOa color is usually measured on pressed powder pellets and is
reported in a variety of color systems [69]. These describe color via three
dimensions: hue, value (brightness) and chroma (purity). The most commonly
used are the Munsell and CIE (Commision International de 1’Eclairage) systems
[70, 711.

4.7 Hazards
The hazards of TiOs pigment and the relative lack thereof are detailed
in the Material Safety Data Sheets [72] which must accompany any US
shipment.
TiOz presents no pronounced health hazards; it is neither corrosive nor
acutely toxic and does not appear to be a significant carcinogen nor embryo
toxin in the work place. As a dry powder, TiOz can become a nuisance dust
that may require control [72].
The potential toxicity of r-utile and anatase dust has been discussed by
Zitting et al. [73]. Upon prolonged exposure to high dust loadings, TiOz
can become a mild irritant to the lung [74]. In vitro studies of TiOa and
pigment toxicity indicate generally mild effects [73, 751. Industrial hygienists
120

have found adverse effects of silicon compounds [76] but mild [ 76, 771 or
no effects [78] for titanium dioxide.
TiOa pigment cannot burn nor explode, neither as a dry powder nor as
aqueous shrrry. Neither the dry pigment nor the slurry is corrosive nor
reactive. The hazards of pigment dispersions in organic liquids reflect the
characteristics of the liquids.

5 Performance

The TiOz industry ‘sells’ light scattering. Thus, optical effectiveness is


the primary performance characteristic of the pigment. However, TiOB pigment
is well optimized. Optical performance differences are small between products
developed for a similar end-use. Thus, secondary characteristics may assume
major commercial signihcance.
5.1 Optics
5.1.1 Light scat&ring
The primary value of TiO, pigment is its ability to scatter visible light
from a polymeric matrix cost effectively. All other performance characteristics
of the pigment - dispersibility, chemical characteristics, effects on durability
and gloss of paint films, application characteristics of paint - are less
important.
The particle size of TiOz pigment has been optimized experimentally
and confirmed by Mie theory. Optimal size depends on the wavelength of
light. For white pigment applications, size is optimized for green light where
the human eye is most sensitive. A rutile sphere of 0.2 pm is optimal. Mie
calculations can specify an optimal diameter for the scattering of an elec-
tromagnetic wave by a dielectric particle with great precision but the result
is limited to single, true spheres. The quantitative translation of data derived
from theory, applicable only to one sphere, into (a) real particles of elongated,
angular shapes, (b) from a single, isolated particle to assemblies of massive
numbers and (c) from mono size particles to particle size distributions remains
unreliable.
Particle crowding diminishes light scattering substantially. Crowding
affects smaller particles to a greater extent than larger ones. A simple physical
model that explains the effects of pigment crowding on light scattering has
been proposed by Fitzwater and Hook [79]. They consider the loss of
scattering caused by the overlap of optical cross-sections, which, for pigment-
size particles, are significantly larger than their geometric cross-sections.
The optical effects of combined light scattering and absorption by
homogeneous thin films are quantified by the Kubelka-Munk model. This
serves exceedingly well in spite of some theoretical limitations. More complex
models have been proposed but have not established themselves. Ku-
belka-Munk equations are the basis of most performance measurements:
hiding, opacity and tinting strength. The model has been well explained by
Judd and Wyszecki [ 711.
121

Several theoretical efforts have been reported in recent years. Mie


calculations for a variety of pigments have been discussed by Kunitomo et
al. [SO]; Bhavsar et al. have examined the theoretical relation between
Kubelka-Munk and Mie calculations [ 81, 821; a multiple scattering technique
for the computation of reflectances of pigmented surfaces has been evaluated
by Shah and Mehta [83]. Ross has calculated the theoretical maximum light
scattering power of r-utile spheres and spherical bubbles in resin [84].
In a more practical vein, Simpson has summarized factors associated
with the opacity of pigments [85, 861, and Steig has discussed the effect
of porosity, dilution and TiOa concentration, recommended a laboratory
technique and solved formulating problems [87].
The light scattering of air voids in resin has been calculated and compared
with the scattering of r-utile in resin and in air and with foams [88, 891.
Practical implementations of void hiding have been proposed and the trade
names Pittment@ - air voids generated within paint films [90] - and
Ropaque@ - resin spheres with a single central air void [ 91, 921 - have
been coined.
5.1.2 Measurement
Until recently, pigments and paint films were evaluated by visual com-
parison. The eye was more sensitive than available instruments and the
mathematics of Kubelka-Munk are too complex for routine calculations.
Old practices continue through much of the coatings industry, but two
developments make quantitative evaluation of pigments now appropriate:
optical instruments have become more sensitive and more reliable than the
eye and the least powerful of computers can calculate results on the spot.
The light scattering and light absorption of paint films can be quantified
independently from reflectance measurements of thin* films drawn over black
and white substrates. Additionally, film thickness or film weight and com-
position must be measured. F’ilms on transparent substrate can be measured
in the transmission mode.
With spectrophotometers, reflectances can be measured for any specific
and narrow wave band. By measuring wave band by wave band across the
spectrum of visible light, the applicability of hiding power measurements
may be extended from black and white to color. Paints of different hue can
be compared and the results expressed separately for white and color pigment
performance.
Tinting strength tests measure the relative light scattering of a white
pigment by overwhelming the intrinsic absorptions of binder and pigment
by the addition of colorant. Thick films of infinite hiding power+ may be
measured. The same test principle can be used to measure the color strength
of a color pigment. Vial has evaluated the DIN method for the determination
of the relative scattering power of white pigments [93]; Jansen has presented

*Thin enough to show obvious contrast between regions drawn over black and white
backgrounds but thick enough to look uniform rather than mottled.
+Thick enough so that further increases in thickness do not affect reflectance.
122

a tinting strength method [94]; while Shafey et al. have developed and
evaluated a diffuse spectral reflectometer [951. An infrared back-scattering
method for the measurement of pigment flocculation in paint films has
been described [96] and developed into an instrumental technique [97,
981. Billmeyer et al. have concerned themselves with the color&tic and
spectral aspects of measurements based on Kubelka-Muuk equations
[99, 1001.

5.1.3 Gloss
Gloss is a quality of light that is specularly (mirror-like) reflected by a
surface. All levels of gloss have their coating markets. Low gloss - flat
surfaces - prevails in wall paints; intermediate levels of gloss in trim paints
and objects of daily use. The highest glossiness is demanded for automotive
and some appliance llnishes, higher, in fact, than paint films can deliver.
Incentives to improve the gloss of automotive finishes are high. Glossiness,
i.e. the wet look, sells cars.
Gloss is highest for surfaces of clear liquid and amorphous solids.
Particulates in the paint film - pigments, extenders and some additives -
reduce its gloss because they roughen its surface. Low and intermediate
levels of gloss are deliberately formulated into paints through coarse flattening
agents and large extenders to achieve pleasing satin tones. Where gloss is
intended to be at its highest, the presence of pigments degrades it.
Since the volume concentration of white pigments in paint lilms must
be much higher for hiding than the concentration of color pigments and
black, the burden of improvement falls on TiOz producers. In response to
customer pressure, they have increased the gloss performance of grades that
serve the automotive and industrial markets.
Matters are complicated by fundamental differences between measured
gloss that guides pigment development and perceived gloss that sells cars.
Several theoretical and numerous empirical studies are beginning to clarify
the issues. Insights into the nature of gloss lead the way. Thus Braun [ 1011
has described the manifestations of gloss in terms of three distinct dimensions:
(i) the physics involved in the optics of specular reflection and its measurement;
(ii) the physiology encompassing the capabilities and limitations of the human
eye; and (iii) the psychology that translates perception into consumer values.
The two operational definitions of gloss, one based on a human perception
of a distinctness-of-image, the other based on instrumental measurement
of an intensity-of-image, differ profoundly in substance but correlate well
enough for most purposes of paint evaluation and pigment development.
Simpson has provided an in-depth and authoritative review of the subject
[102];-Germak and Verma [103] have studied objective predictors of paint
appeal; while Uchida and Maruo [ 1041 have measured gloss by an aperture-
contrast method and observed disagreements with visual ratings.
Simpson [ 1051 has reviewed two optical equations that quantify the
intensity of specularly reflected light in terms of (a) refractive index and
123

optical angle, and (b) surface roughness and wavelength. The same subject
has also been studied in detail by Kawabata [ 1061.
Braun [ 1011 has proposed a mechanism by which pigment causes surface
roughness. This mechanism is seen as an interplay between the surface
tension of the wet lilm with the compressive strength of the gel structure
that develops within the film while it dries. The effect of pigment and extender
particle size and concentration has been reviewed and studied by Simpson
[ 105 1, pigment grades by Davies [ 1071 while Biglieri and Di Paolo have
analyzed the effects of pigment dispersion [lOBI.

5.2 Dispersibility
Dispersion is the incorporation of dry pigment in a liquid matrix. It
consists of the replacement of the air-solid and solid-solid interfaces with
a liquid-solid interface (including supercooled liquids, such as plastics).
During dispersion, three distinct processes occur: (1) penetration of the
liquid phase into inter-particle pores; (2) breaking of particle-particle bonds;
and (3) distribution of particles throughout the dispersion medium.
If the medium is of such a low viscosity that Brownian motion can bring
particles into contact with each other, a stabilizer (usually called a dispersant)
is added. It adsorbs onto pigment particles and prevents them from sticking
to each other and reagglomerating into random structures. These structures,
or floes, are easily disrupted by shear, but re-form when the shear is removed.
Floes degrade optical properties.
Stabilizers can prevent flocculation by electrostatic repulsion and by
steric hindrance. Electrostatic repulsion takes place whenever dispersed
particles acquire a charge. Steric hindrance occurs when one segment of a
highly solvated molecule is adsorbed onto the pigment surface while another
segment extends into the dispersion medium. For another particle to approach
close enough to stick it must push aside the adsorbed polymer, and for that
to happen the associated solvent must be displaced, an improbable event.
In non-aqueous systems steric repulsion is the chief stabilizing mechanism
because a low dielectric constant limits ionization. In aqueous systems both
mechanisms are operative, and modern dispersants are designed to utilize
both. Their long, hydrophilic polymer chains, with easily ionizable groups
attached to the end that extends into the dispersion medium, provide both
the charge for electrostatic repulsion and the high degree of hydration for
steric stabilization.
To be effective as a steric stabilizer, one segment of the dispersant
molecule must be securely anchored on the pigment surface, while the other
end must extend into the dispersion medium. Since carboxy end-groups,
each with the potential of forming two hydrogen bonds, are strong anchors,
modern dispersants are derivatives of carboxylic acids. To attract such
dispersants, pigment surface constituents must be insoluble with a slightly
alkaline reaction.
Hydrous aluminum oxide, with its chemical stability, point-of-zero-charge
at pH 9 and solubility of 5X 10e7 mol l-‘, satisfies all the requirements for
124

a surface modifier for a pigment. Anisotropic hydrous alumina particles create


sufficient surface roughness for a significant decrease in the Londonhran der
Waals’ attraction among pigment particles [ 611. The high surface area of
hydrous alumina provides anchoring points for the dispersant molecules.
In paint manufacturing today, the most popular dispersers are high-
speed mixers with serrated disk blades as agitators. They have a small high-
shear region around the perimeter of the blade, through which the rest of
the batch circulates and where the agglomerates are broken. Dispersion is
usually evaluated by the Hegman gage, a shallow, tapered channel into which
the paint is poured, and the excess scraped off. The thin, elongated wedge
of paint is then examined to see at what depth the diameter of the largest
agglomerates exceeds the depth of the channel. This point is visible as a
transition from a smooth to a streaked paint surface.
For the dispersion of pigment in high-viscosity polymers and plastics,
counter-rotating blades and twin-screw mixers are generally used. Because
of the high viscosity, reflocculation cannot occur, so only a thin hydrous
alumina coating is needed on pigment to keep the particles apart and insure
a good dispersion.
In the powerful plastic dispersing machinery required to overcome high
viscosity, pigment clumps are frequently compressed and form grit. To reduce
mechanical pressures, high-boiling compounds with low surface energy, such
as stearates and siloxanes [ 371, lubricate pigment and minimize interparticle
attraction. This kind of surface treatment retains the naturally high bulk
density of titanium dioxide pigment, enabling manufacturers to make highly
loaded pigmented concentrates of manageable viscosity.
The addition of organic compounds has its drawbacks: the oily additives
migrate to the surface of plastic films and hurt printability and heat-sealing
performance [60]. The problem can be overcome by co-extrusion: trapping
the pigmented film between two clear films.
All aspects of dispersion, including machinery and methods of assessment,
have been thoroughly reviewed in the book by Patton, covering literature
prior to 1978 [lo]. Since then, an engineering view of dispersion with selected
bibliography has been published in a book by Nelson [ 1091. Dispersion in
non-aqueous systems has been reviewed by McKay [ 1 lo]. Papers dealing
with various aspects of dispersion during paint preparation include millbase
preparation [ 111,112 I, dispersant evaluation [ 113,114], ultrasonic dispersion
[ 1151 and theoretical studies of agglomerate disruption in dispersion machinery
11161. The dispersion of pigments in plastics has also been covered [60,
117, 118 I. Novel methods of assessing the quality of dispersion include the
free volume microprobe [119], the electron microscopy of paint films [120,
121 I and the measurement of flocculation gradient by the scattering of
infrared light [97, 1221.

5.3 Durability
Durability is the continuance of decorative and protective performance
of paint films and their components under the influence of weathering. Lack
125

of outdoor durability manifests itself as discoloration and chalking of the


TiO, with concomitant erosion and gloss loss. The instability of substrates
introduces additional degradation effects.
The durability of the paint 6lm differs decisively from the durability of
its components. By itself, any TiOz is absolutely stable. Exterior binders are
practically stable. But a composite of unimproved TiOz and binder degrades
rapidly. Degradation occurs because paint films are slowly oxidized by air.
Sunlight triggers the degradation reactions. In the dark, paint films can last
for centuries, even millennia. In light, durability depends on binders, pigments
and the conditions of exposure.
Only the W portion of sunlight causes degradation because it has an
energy content sufficient to break chemical bonds. Paint vehicles, all organic
and some inorganic pigments degrade irreversibly; only a few inorganic
pigments are thermodynamically stable. Titanium dioxide is stable but its
catalytic characteristics are activated by W radiation.
Literature reviews abound. In 19 72, Sullivan published a definitive review
[3]; in 1974, Voeltz et al. [123] addressed the subject of Ti02 surface
reactions; in 1983, Simpson [105, 1241 evaluated the wavelength effects; in
1981 and 1986, Colling and Dunderdale [125] considered film surface and
subsurface; and in 1987, Braun [ 1261 summarized previous fhrdings.

5.3.1 Catalytic activity


Titanium dioxide affects the durability of paint films in two distinct and
opposing ways:

1 As a strong W absorber, TiOs protects the paint 6lm.


2 As a W-activated oxidation catalyst, TiOz degrades binders.

The protective function of TiOz has not received much attention, possibly
because it is not important for the binders that are used in durable paints.
The destructive effect of Ti02, commonly called chalking, has been the subject
of extensive study.
A model of chalking mechanisms has been developed and elegantly
confirmed through scanning microscopy by Kaempf et al. [ 127, 1281. They
have described the characteristics of weathered paint films in terms of binder/
pigment combinations. During degradation, paint films lose mass [ 1291 while
pigment concentration near the surface increases substantially [130].
The chemistry of chalking has been outlined by Voeltz et al. [ 131-1331
who demonstrated that W light, water and oxygen are essential to the TiOs-
catalyzed degradation of binder. They proposed the following reactions,
supported by an abundance of data by numerous other investigators:
Ti*‘...OH- +hv - Ti3’ + OH’ (1)
Ti3’ + OZ - Ti4+...02- (2)
Ti*+...Oa- +H,O - Ti4+...0H+HOz’ (3)
126

The sequence continues (1)-(2)-(3)-(1)-(2)-(3)-(l). . . , with an overall


reaction:

HzO+02+hv= OH’+HO; (4)


Hydroxy and peroxy radicals activate, oxidize and degrade the binder:

3HO’+ 3HO;+ 2(-CHz-) - ZC02 + 5HaO (5)


The chain of chalking events is cyclical with respect to TiOz and can be
disrupted at the TiOa surface by exclusion of either UV light, water or oxygen.
Lacoste et al. [ 1341 have considered the photocatalytic activity of TiOz
in the context of zinc oxide and cadmium selenide. Keifer [ 1351 has related
photochemical activity to the drying time of alkyd paints. Pappas and Fischer
[ 1361 have studied the involvement of singlet oxygen in chalking.

5.3.2 Control
The TiOa industry has learned to control chalking in several ways. The
first substantial improvements were made by a switch of crystal phases. The
industry had started with anatase pigments because anatase is the product
that crystallizes naturally from hydrolysates of titanyl sulfate. Thus, anatase
is easier to make by the sulfate process but i-utile is far less photoreactive.
In the 194Os, the industry introduced r-utile grades for paints and plastics.
A second quality leap came in the 1960s with the encapsulation of rutile
in amorphous silica [ 421. The glass prevents contact between the catalytic
surface of rutile and the organic vehicle and provides a surface for the
recombination of free radicals. Intermediate levels of ‘durability’ are attained
by partial encapsulation in silica and/or alumina, and by bulky coatings of
hydrous aluminum and silicon oxides.
Alternative approaches to chalking control involve attempts to (a) re-
combine holes and electrons at the TiOa surface via semiconductor mechanisms
and (b) prevent hydroxylation of the TiOa surface, i.e. interfere with reaction
(3) (see p. 125) of the chalking sequence.
Day and Egerton [ 1371 have used surface analytical methods to study
the effects of TiOa surface coatings on photocatalytic activity.

5.3.3 Measurement
Progress in the control of chalking in the 1960s outpaced advances in
testing and created problems for product development and promotion. Instead
of quantitative durability measures, the coatings industry accepts ratings
based on simulations of natural exposures. Ideally, durability would be
measured in terms of times and activities:

‘Durability’ = f(time of exposure, reactivity of components, etc.)

with such a function providing time factors of durability. With quantitative


rather than comparative durability data, results from different sources could
be compared and degradation processes could be modeled. Simms’ ‘accel-
eration shift factors’ [ 1381 are a practical approximation. But unfortunately,
127

‘durability’ cannot be measured as a direct function of time because degradation


of a paint 6lm itself involves multiple chains of events with links that are
not time-dependent and others that are specific to certain systems. The
literature of outdoor and accelerated exposure testing [ 139-1561 is too
extensive to be reviewed in this broader context.
The ‘durability’ of a pigment can be predicted from a measurement of
its potential contribution to the degradation of paint films, using a ‘catalytic
activity coefficient’ as a quantitative measure. A test described by Braun
[157] follows the UV-energized, cyclic reduction/oxidation of titanium ions
on the surface of a pigment which drives the chalking mechanism. The
reduction of lead salt ‘counts’ redox cycles by the accumulation of metallic
lead. Black, metallic lead darkens an initially white paste. Lead accumulation
is readily measured by reflectance. Only a single measurement of reflectance
is required to characterize the catalytic activity of a pigment via a mathematical
model based on the first principles of physics.
Other tests [ 1571 have been based on the oxidation of isopropanol
[158-l 631 and mandelic acid, the decomposition of peroxides [ 1641, the
formation of defect electrons [ 165 1, the generation rate of radicals as observed
by ESR spectroscopy [ 1661, the oxidation of iodide [167] and the reduction
of silver salts [ 1681.

5.4 Ti02 substitutes


Recent TiOa shortages have prompted a search for TiOa substitutes.
Numerous products have appeared. ‘Classic’ white pigments, porous beads,
void hiding plastics and a variety of extenders are being promoted as partial
replacements for TiOa. The total replacement of TiOa pigment has not been
advocated.
White pigments are particulates of some of the very few substances that
have refractive indices higher than 2.0 (Table 2). They must also be inert,
stable, non-toxic and colorless. Three classic white pigments belong to this
category: lithopone, zinc sulfide and zinc oxide. These pigments became
obsolete because they were and are far less hiding and cost-effective than
TiOa. Lead white, the widely used of the classic white pigments, is now
considered too toxic for pigment use.
Extenders are inexpensive, inert, colorless ground minerals and precip-
itated particulates that (i) do not hide in the polymer matrix because their
refractive indices are similar to that of the matrix, (ii) can hide indirectly
by creating porosity in polymer-starved matrices, and (iii) do hide in very
porous films by scattering light at the extender/air interface. Frequently,
nowadays, extenders are mentioned in the context of spacing TiOa particles
[ 1691. Through crowding, white pigments lose as much as two-thirds of their
scattering power [ 1701. Uniformity of spacing can minimiie detriments.
The effectiveness of extenders and coatings on TiOa particles has become
the subject of controversy [171-l 73 1.As yet, matters have not been resolved,
in part undoubtedly because of the complexity of paint formulations. Because
of the large number of variables involved in formulating paints, experimental
128

TABLE 2
Representative refractive indices of colorless substances

Substance Refractive index*

media vacuum 1 .oooo


air 1.0003
water 1.3
silicate glasses 1.5-1.9
organic polymers 1.4-1.6

extenders almost all 1.5-1.6

pigments white lead 1.9-2.1


zinc oxide 2.0
zinc sulfide 2.4
anatase titanium dioxide 2.5
rutile titanium dioxide 2.7

miscellaneous substances diamond 2.4


most lead compounds 1.6-2.2
most oxides 1.4-2.0
most ionic salts 1.4-1.7
most organic compounds 1.4-1.6

*Rounded off for ease of comparison.

designs are exercises in many-dimensional spaces that make each author’s


data unique and meaningful comparisons between publications all but im-
possible.
Extended TiOZ pigments were the first products of the TiOZ industry.
Calcium sulfate was used almost exclusively. Extended pigments became
obsolete, largely because they are incompatible with water-borne paints. In
water, calcium sulfate recrystallizes to large-sized particles. Steig has described
the demise of these products as the “death of a giant” [ 131.
The recent non-peer-reviewed literature on the partial substitution of
extenders for TiOa pigment is extensive but short on reliable evidence.
Unfortunately, statistical validations of the conclusions are the exception
rather than the rule, although it is granted that this is not an easy task when
dozens of many interdependent variables are involved.
Voids of appropriate size scatter light as if they were particles. Their
scattering effectiveness is low because the difference between the refractive
indices of air and the polymer matrices is only modest [ 174, 1751 (Table
2). A theoretical study of pigmented microvoid coatings of hypothetical
concentric structures leads to a counter-intuitive result: concentric layers of
high and low refractive indices cause a pronounced reduction in the scattering
coefficients [ 176 1.
Pigmentation by microvoids developed within paint binder was advocated
widely as ‘pittmentation’ [ 177, 1781. Pittment? voids were not added as such
to the paint but created within paint flms from the evaporation of cheap
129

hydrocarbon droplets, emulsitled in the paint. The technology did not catch
on.
Beads composed of voids, polymer and pigment do scatter light effectively
[ 1791. The beads themselves have to be large enough to accommodate
numerous pigment particles and voids. Because of their size, they are flattening
agents not suited to gloss paints. A process to make pigmented, veciculated
beads called Spindrift? [ 1801 received some commercial attention.
Irregular voids, and particularly interconnected pores, increase the light
scattering of paint films but severely degrade the mechanical and chemical
performance of the ms. It is through the creation of pores that extenders
contribute to hiding. Color increases the hiding of paint films. Off-color
extenders and slightly colored polymers can boost hiding incidentally or
deliberately. Hiding is improved at the expense of brightness.

6 Applications

Half of all TiOa pigment produced is consumed by the coatings industry


and a quarter by the paper industry. Fifteen per cent goes into plastics; less
than 3% into inks. All other end-uses - floor tiles, ceramics, fibers, fabrics,
roofing granules, welding rods, foods, etc. - consume c. 7%. Underdeveloped
countries consume modest quantities, mostly in coatings.

6.1Coatings
According to Lambourne [ 1811 the term ‘paint’ and ‘surface coating’
are often used interchangeably. Surface coating is the more general description
of any material that may be applied as a thin continuous layer to a surface.
Paint was traditionally used to describe pigmented materials as distinct from
clear films which are more properly called lacquers or varnishes. Because
coatings are applied in thin layers, a relatively high amount of pigment must
be used to achieve hiding. The amount of pigment in a coating is commonly
specified as the pigment volume concentration (PVC). Different kinds of
paints (high gloss, flat) will have different levels of PVC.
The coatings market is divided into several categories:

Trade sales - Trade sales coatings are paints sold through retail stores
for application by the consumer or professional painter. They are usually
air-dry paints applied by brush or roller. Spray application is becoming
increasingly popular. Trade sales coatings are the coatings equivalents of
thermoplastics resins.

Industrial coatings - Industrial coatings are formulated for application


as part of a manufacturing process. Application of these paints tends to be
less tolerant of operating error than the typical trade sales paint. These
coatings are also more sensitive to substrate contamination and surface
heterogeneity [ 1821. Most industrial coatings are coatings equivalents of
thermoset plastics.
130

The application of industrial coatings includes brush, roller, dip, elec-


trodeposition and spray [ 1831. Some industrial paints also come in two-part
packages that are mixed prior to use. Film drying and/or cure techniques
are similarly diverse. Thermoset or baked finishes are an important part of
the industrial market. Infrared or ultraviolet light can be used to crosslink
suitable polymers. Electron beams or radiofrequency radiation are also used
[ 184-1861.
Industrial coatings cover a wide range of services. This includes wood
products, metal furniture, coil coatings, appliances and transportation equip-
ment. Automotive and automotive aftermarket coatings are usually considerd
as a separate business.

Automotive jinishes - Automotive finishes have the most critical per-


formance requirements of the industry. These include the need for photo-
durability, corrosion protection and chip resistance. Pigments suitable for
automotive market demands serve as a benchmark for other less critical
applications [ 187 1. Pigment characteristics important to automotive finishes
are gloss, gloss retention and hiding power.
Titanium dioxide pigment is used in some automotive topcoat finishes.
Exceptions are metallic and deepcolor topcoats.

Primers - Primer-sealers come under four basic categories: anticor-


rosive, sacrificial, barrier and film reinforcers. Titanium dioxide is used to
impart white hiding and in some cases film integrity. There has been a recent
move to brighter primers which contain TiOa.

The adhesion, gloss and general appearance of a paint film are largely
determined by the physical condition of the surface to which the paint is
applied. Few previously unpainted materials satisfy these requirements. For
the best results, the final coat should be applied to a surface which is
uniformly smooth but not glossy, is easily wetted by the topcoat and may
have a level of porosity that permits some penetration of the topcoat for
good adhesion. The use of cathodic electrodeposition of primer on car bodies
has recently become widespread. Electrocoating provides vastly superior
corrosion protection [ 188 1.

6.2 Plustics
Titaniurndioxide pigment is used to opacify plastic materials. Unpigmented
plastic is translucent to transparent and therefore not aesthetically appealing.
In some applications, TiOz is used to improve photodurability.
The term plastic derives from the Greek term plmtikos which means
to mold or form. The application of the term plastic to polymeric substances
was introduced around the turn of this century during the time these materials
were introduced into the marketplace. The use of plastic has become ubiquitous
[ 1891.
131

Plastics have distinct advantages over more traditional materials such


as metal or wood. These include lower raw material cost, lower fabrication
cost, lighter weight, a greater strength to weight ratio, and better durability
and corrosion resistance.
Plastics are dehned as either thermoset or thermoplastic. Thermoset
plastics, once hardened, cannot be reworked by the application of heat.
Thermoplastic materials can be remelted and formed again. Thermoplastics
account for about 90% of the plastic market.
The requirements for TiOa in plastics are good dispersibility in a polymer
system, blue undertone and good heat stability. Blue undertone is required
because humans perceive bluish whites as cleaner than yellow ones. A few
applications, light shades among them, call for yellow undertone pigment
because the transmitted light is bluish.
Plastic objects usually have a much lower pigment concentration than
coatings because they are usually much thicker than films. Hence, opacity
can be achieved with less TiOa. The vast amount of plastic produced, however,
makes the plastics industry a major TiOz user.
In many cases the TiO, for plastic does not have an inorganic coating
such as hydrous alumina, because its water vaporizes at the high temperatures
of plastic processing creating bubbles in the hnished part. In thin-film
extrusions, this effect is known as lacing. Uncoated pigments for plastics
are usually treated with an organic polyol, silane or siloxane to enhance
dispersibility.
In stabilized plastics such as polyolefins, TiOB accelerates yellowing of
the plastic. The phenomenon is caused by the reaction of titanium with
phenolic groups attached to the antioxidant additives in the plastic. The
chemistry involves the base-catalyzed formation of a yellow titanium phenate
complex. Suitable organic surface treatments have been shown to reduce or
eliminate this effect. Other mechanisms of yellowing not related to TiOa can
be a factor. This includes the effects of smog, ozone and other photoreactions.

6.3 Paper
Titanium dioxide is added to high-quality or thin papers to impart
brightness and opacity. Other papers are pigmented exclusively with clays
and calcium carbonates, which cost less than paper pulp but decrease strength
and increase weight. The manufacture of paper has been described by Casey
[ 1901 and the use of pigments in paper has been reviewed by McGinnis [8]
in 1984 and Alince et al. [191] in 1985.
Because of the rough surface of paper, gloss and photochemical stability
are not as important in paper as they are in paint. Consequently, untreated,
and less vigorously ground, anatase and rutile pigments can be used. Thus
paper pigments tend to be cheaper than paint pigments. Since paper is made
in aqueous systems, most of the titanium dioxide is sold to the industry as
a slurry.
Because of its higher optical efficiency, rutile is the pigment of choice
especially in coated paperboard. The higher abrasiveness of r-utile crystals
132

can be overcome by coating them with a mixture of hydrous silica and


alumina. An advantage of anatase is its lower absorption of light in the far-
violet and near-UV. This results in a blue-white color perceived as pleasing
to the eye. Anatase is compatible with the fluorescent brighteners that are
quenched by r-utile.
The paper industry uses pigment in two applications: wet-end addition
and paper coatings. In the wet-end addition, an aqueous pigment dispersion
is added to the paper pulp. To make pigment particles stick to the fibers,
flocculating agents and retention aids are added. Up to 6% by weight of
titanium dioxide can be added this way without decreasing the strength of
the paper. Addition conditions can affect retention [192] and opacity [193].
In the coating operation, a shu-ry of 5-10% titanium dioxide, SO-85% clay
and 10% binder is spread on printing papers for better printability, or on
paperboard or speciality papers for improved appearance.
In the paper industry, titanium dioxide is used in conjunction with clay,
calcium carbonate and other extender-type pigments. Its levels are adjusted,
depending on the quality of the paper pulp, to meet performance and cost
goals. Magazine, book and business paper contains O-5% titanium dioxide.
Up to 9% titanium dioxide is used in envelope, bible and opaque offset
paper, and up 30% in the papers that decorate plastic laminates. Papers for
laminates require a special, light-stable pigment resistant to photodarkening.
The thick, dense plastic layer on top of the pigmented paper prevents the
access of oxygen, so any photoreduced titanium stays in that state for a
long period of time, darkening the pigment and discoloring the laminate.
The required photostability can be attained either by annealing finished
silica-alumina-treated pigment in a kiln [ 1941 or by binding ceric ion to the
pigment [195]. Ce4+ ion re-oxidizes the Ti3+ formed by photoelectrons
reacting with Ti4+ ions. The resulting Ce3+ ion is subsequently re-oxidized
by a photogenerated hole, regenerating Ce4’ and restoring the starting
conditions.

6.4 Ink
The term ink comes from the Greek root enkaien which means to burn
in. Today, the concept of an ink spans a wider range of application than
the traditional arts of letter-press, gravure and lithiographic printing [196-l 981.
Inks have performance requirements different from coatings. They are
usually applied in a much thinner film than a normal coating. Ink formulation
is adjusted by the printer to meet the specific requirements of drying speed,
type of application equipment as well as the nature of the substrate being
covered [ 1991.
For white hiding and high brightness, titanium dioxide is the pigment
of choice. The type of TiOs can affect ink rheology, abrasiveness, gloss and
redispersibility. Additional ink applications with TiOz include inks for wood
molding, marking pens and decorative sheets. Ink correction fluid for paper
depends on titanium dioxide to hide errors. Inks for concealed writing
133

(scratch-off lottery tickets) likewise use TiOa because of the superior hiding
power. Even the marking on gelatin capsules can contain titanium dioxide.

6.5 Others [2, 2001


Other applications of titanium dioxide can be classified as pigmentary
or non-pigmentary. In some non-pigmentary applications, pigment is used
simply because it is the cheapest pure material of proper particle size. In
others, pigment is a source of titanium.

6.5.1 Pigmmatary uses


TiOa is used as the pigment in linoleum, asphalt tiles and other floor
coverings, granule coatings for asphalt roofing shingles, fabric and imitation
leather coatings (e.g. oil cloth), rubber and other elastomers (e.g. whitewall
tires), sealants (e.g. caulk), delustrant in textile fibers to get a soft white
look, food coloring (can be used up to 1% by weight), cosmetics, photographic
paper and film coatings, and dental materials.

6.5.2 Non-pigmentary uses of pigment


In ceramics and glazes, the titanium dioxide is dissolved in the glass,
then precipitated as a pigmentary particle size by heat treatment to whiten
the ceramic objects. In dental waxes and other high value-in-use plastics,
titanium dioxide is used as an inert filler and rheology modifier.
Pigmentary titanium dioxide powders are reacted with alkali-metal com-
pounds to make t&mates for high-performance dielectrics. They can also be
fused into large r-utile and titanate crystals for gem and optical use.
Other uses are as a photoconductor in xerography and as a mild abrasive
in polishing powders. Occasionally, sintered titanium dioxide pigment ag-
glomerates are used as catalyst supports.

References

1 F. B. Steig Jr., in R. W. Tess and G. Poehlein (eds.), Opaque White Pigments in Coatings,
ACS Symp. Ser. No. 285, 2nd edn., Applied Polymer Science, New York, 1985.
2 J. Barksdale, Titanium, 2nd edn., The Ronald Press, New York, 1966.
3 W. F. Sullivan, Prog. Org. Coat., I (1972) 157.
4 T. C. Patton, Pigment Handbook, Vol. 1, Wiley, New York, 1973, p. 1.
5 H. B. Clark, in R. R. Myers and J. S. Long (eds.), Treatise on Coatings, Vol. 3, Dekker,
New York, 1975, Part 1, p. 479.
6 G. D. Parfitt and K. S. W. Sing, Characterization of Powder Surfaces, Academic Press,
London, 1976.
7 T. Entwistle, in A. D. Wilson and H. J. Prosser (eds.), Suflace Coatings 2, Vol. 2, Elsevier
Applied Science, New York, 1988, pp. 183-223.
8 W. J. McGinnis, in R. W. Hagemeyer (ed.), Pigments for Paper, TAPPI Press, New York,
1984, p. 241.
9 Kronos Guide, European Titanium Pigment Division, National Lead Co., Kronos Titania
Companies Press, Leverkusen, Germany, 1968.
10 T. C. Patton, Paint Flow and Pigment Dispersion, 2nd edn., Wiley, New York, 1979.
11 R. E. Day, Prog. Org. Coat., 2 (1974) 269.
134

12 R. M. McKinney, US Put. 2150235 (1939).


13 F. B. Stieg Jr., J. Paint Technol., 44 (1972) 63.
14 H. Reichmann, Pigm. Resin Techrwl., 3 (1974) 12.
15 R. S. Darby and J. Leighton, in The Modern Inorganic Chemicals Industry, Chem. Sot.
Spec. Publ. No. 31, Chem. Sot., London, 1977, p. 354.
16 F. Izumi, BULL Chem. Sot. Jpn., 51 (1978) 1771.
17 F. Izumi and Y. Fqjiki, Bull. Chem. Sot. Jpn., 49 (1976) 709.
18 A. Matthews, Am. Mineral., 61 (1976) 419.
19 D. S. BuII, Paint Resin, 52 (1982) 15.
20 K. Nakata, Chem. Econ. Eng. Rev., 2 (1970) 25.
21 B. H. R. Mayes, Polym. Paint CoIour J., 178 (1988) 117.
22 V. G. Neurgoankar, J. Chem. Technol. Biotechnol., 36 (1986) 27.
23 R. H. Perry and D. W. Green, Perry’s Chemical Engineer’s Handbook, 6th ed., McGraw-
Hill, New York, 1984.
24 R. Tettenhorst and D. A. Hofmann, Clays CZuy Miner., 28 (1980) 373-380.
25 H. N. V. Temperley and G. E. K. Blythe, Nature (&ndon., 219 (1968) 1218.
26 E. Muschelnautz, G. Giersiepen and N. Rink, Chem.-Ing.-Tech., 41 (1970) 6.
27 T. Tanaka, Ind. Eng. Chem., Process Des. Dev., 11 (1970) 238.
28 T. Tanaka, Ind. Eng. Chem., Process Des. Dev., 12 (1973) 213.
29 M. Menyhart and L. Miskiewicz, Powder Technol., 15 (1976) 261.
30 A. Zahradnicek and F. Loffler, Int. Chem. Eng., 19 (1979) 40.
31 N. Nakayama and K. Inui, Sprechsaal, 120 (1987) 89.
32 T. Tanaka, J. Chem. Eng. Jpn., 5 (1972) 425.
33 M. R. BaIoga, Polym. Paint Colour J., 170 ‘(1989) 364.
34 ASTM Ti02 Pigment SpecZfLcaticms, ASTM Standards, Section 6, 06.02, D476-73, ASTM,
Philadelphia, PA, 1979.
35 R. J. CapweII, F. Spagnolo and M. A. DeSesa, Appt. Spectrosc., 26 (1972) 537-539.
36 H. H. Schaumann, US Pat. 2 798819 (1957).
37 J. WinkIer, Farbe+Lack, 94 (1988) 263.
38 H. W. Jacobson, US Pat. 4416699 (1983).
39 M. A. Claridge and P. L. Cowe, US Pat. 3 767455 (1973).
40 P. B. Howard, US Pat. 4052223 (1977).
41 M. Matsunaga, T. Usami, H. Okuda and H. Futumato, US Pat. 4 405 376 (1983).
42 A. J. Werner, US Put. 3437502 (1969).
43 B. Barnard and W. T. Laverick, US Pat. 4239548 (1980).
44 W. T. Siuta, US Pat. 3 035 966 (1962).
45 F. B. Stieg Jr., J. Paint Technol., 39 (1967) 701.
46 R. L. De Colibus, US Pat. 4177081 (1979).
47 W. H. Morrison, Can. Pat. 1206392 (1986).
48 M. A. Claridge and W. E. Craker, Inst. Chem. Eng. Symp. Ser. No. 63, (Technol.), D411/
l-D4/II/13, 1981, pp. 42-45.
49 R. A. Slepetys, J. Paint TechnoL, 44 (1972) 91.
50 U. Barcucci, F. Bigheri, F. Gambino and R. Borsso, XIIth FATlPEC Congress, Congress
book, Vol. 12, (1974), pp. 249-254.
51 C. E. Bamett and G. G. Simon, J. Colloid Interface Sci., 34 (1970) 580.
52 H. H. Schaumann, US Pat. 2488440 (1949).
53 P. Sennet, J. P. OIivier and G. K. Hickin, Tappi, 57 (1974) 92.
54 6. Mukhopadhyay and R. J. Ramrsjkar, Paintindia, 35 (1985) 25.
55 C. C. KaIita, L. D. Brown and D. M. Kirkpatrick, Am. Inkmaker, 63 (1985) 2730, 32,
34-35 and 37-38.
56 R. D. Murley, XIIth FATIPEC Congress, Congress book, Vol. 12, (1974), p. 377.
57 J. WinkIer and L. DuIong, Farbe+L.ack, 89 (1983) 332.
58 W. M. Hess and G. C. McDonald, Rubber Chem. TechnoL, 56 (1983) 892.
59 F. B. Stieg Jr., J. Paint Technol., 43 (1971) 36.
60 D. Williams, Sot. Plost. Eng., Tech. Pap., 25 (1979) 307.
135

61 J. Czamecki and T. Dabros, J. Co&id Inter$ace Sci., 78 (1980) 25.


62 T. F. Swank, US Pat. 3523810 (1970).
63 T. F. Swank, US Pat. 3595822 (1971).
64 A. S. Hare and J. C. Vickerman, J. Chum. Sot., Faraday Trans. I, 79 (1983) 185.
65 F. B. Stieg Jr., J. Coat. Technol., 61 (1989) 67.
66 D. H. Solomon and D. G. Hawthorne, Chemistry of Pigments and Fillers, Wiley, New
York, 1983.
67 K. Vos, J. Phys. C, 10 (1977) 3893-3915 and 3917-3938.
68 M. Kerker, The Scattering of Light and Other Elect romagnetic Radiation, Academic
Press, New York, 1969.
69 F. W. Billmeyer and M. Saltzman, Principles of Color Technology, 2nd edn., Wiley, New
York, 1981.
70 R. S. Hunter and R. W. Harold, The Measurement of Appearance, 2nd edn., Wiley, New
York, 1987.
71 D. B. Judd and G. Wyszecki, Color in Business, Science and Industry, 3rd edn., Wiley,
New York, 1975.
72 Material Safety Data Sheet, Ti-Pure Titanium Dioxide Pigment, Jan. 1992.
73 A. Zitting and E. Skytta, Int. Arch. Occup. Environ. Health, 43 (1979) 93-97.
74 R. Elo, K. Maatta, E. UksiIa and A. U. Arstila, Arch. Pathol., 94 (1972) 417-424.
75 I. M. Nuqja and A. U. ArstiIa, Environ. Res., 29 (1982) 174.
76 K. Maatta and A. U. Arstila, Lab. Invest., 33 (1975) 342.
77 E. M. Ophus, L. Rode, B. Gylseth, D. G. Nicholson and K. Saeed, Stand. J. Work Environ.
Health, 5 (1979) 290-296.
78 J. G. Hoogerbeets, VeQlcroniek, 56 (1983) 296.
79 S. Fitzwater and J. W. Hook, J. Coat. Technol., 57 (1985) 39.
80 T. Kunitomo, H. M. Shafey and T. Teramoto, B&l. JSME, 22 (1979) 1587-1594.
81 M. C. Bhavsar, P. M. Vora, H. S. Shah and K. T. Mehta, Paint Resin, 56 (1986) 32-36;
ibid., 56 (1986) 19-20 and 22.
82 M. C. Bhavsar, P. M. Vora, H. S. Shah and K. T. Mehta, Paintindia, 35 (1985) 23-28.
83 H. S. Shah and K. T. Mehta, Dyes Pigm., 9 (1988) 283.
84 W. D. Ross, J. Paint Techrwl., 43 (1971) 49.
85 L. A. Simpson, Polym. Paint Colour J., 179 (1989) 158.
86 L. A. Simpson, Polym. Paint Colour J., 179 (1989) 246.
87 F. B. Stieg Jr., J. Coat. Technol., 49 (1977) 54; ibid., 60 (1988) 95.
88 J. A. Seiner and H. L. Gerhart, X7th FATIPEC Congress, Congress book, Vol. 11, (1972),
p. 127.
89 W. D. Ross, Ind. Eng. Chem., Prod. Res. Develop., 13 (1974) 45.
90 R. W. Andrew and B. Lestarquit, Polym. Paint Colour J., I74 (1984) 442.
91 J. W. Hook III and R. E. Warren, in G. D. ParfItt and A. V. Patsis (eds.), Organic Coatings:
Science and Technology, Vol. VII, Dekker, New York, 1984, p. 299.
92 D. M. Fasano, J. Coat. Technol., 59 (1987) 109.
93 F. Vial, Dtsch. Farben-Z., 28 (1974) 365.
94 H. D. Jansen, Polym. Paint Colour J., 172 (1982) 770.
95 H. M. Shafey, Y. Tsuboi, M. Fujita, T. Makino and T. Kunimoto, ALAA J., 20 (1982)
1747-1753.
96 J. G. BaIfour and M. J. Hird, J. Oil Colour Ch.em. Assoc., 58 (1975) 331.
97 D. J. Rutherford and L. A. Simpson, J. Coat. Technol., 57 (1985) 75.
98 L. A. Simpson, Adv. Org. Coat. Sci. Technol. Ser., 10 (1986) 121.
99 D. G. Phillips and F. W. Billmeyer, J. Coat. Technol., 48 (1976) 30.
100 F. W. BiIImayer and R. L. Abrams, J. Paint Technol., 45 (1973) 31.
101 J. H. Braun, J. Coat. Technol., 63 (1991) 43.
102 L. A. Simpson, Prog. Org. Coat., 6 (1978) 1.
103 G. W. Cermak and M. Verma, J. Coat. Tech??&, 57 (1985) 39.
104 T. Uchida and M. Maruo, Shikizai Kyokaishi, 47 (1974) 164.
105 L. A. Simpson, Polym. Paint Colour J., 176 (1986) 408.
136

106 H. Kawabata, J. Jpn. Sot. Color Mater., 41 (1968) 1; ibid., 41 (1968) 14; ibid., 42
(1969) 2.
107 H. V. Davies, Au.%. Oil Colour Chem. Assoc., Proc. News, 17 (1980) 13.
108 F. Biglieri and V. Di Paolo, XIVth FATIPEC Congress, Congress book, Vol. 14, (1978),
p. 165.
109 R. D. Nelson Jr., Dispersing Powders in Liquids, Elsevier, Amsterdam, 1988.
110 R. B. McKay, Surfactant Sci. Ser., 21 (1987) 361.
111 L. Cutrone, J. Coat. Technol., 56 (1984) 105.
112 C. C. Tatman, J. Coat. Technol., 53 (1981) 57.
113 J. W. Joudrey and M. Schnall, Proc. 16th Water-Borne Higher-Solids Coat. Symp.,
(1989), p. 521.
114 A. C. D. Cowley, J. Oil Colour Chem. Assoc., 71 (1988) 310.
115 J. Stoffer, J. Gordon, M. Fahim and L. Wallace, Proc. 13th Water-Borne Higher-Solids
Coat. Symp., (1986), pp. 410-431.
116 J. Winkler and L. Dulog, .I Coat. TechnoZ., 59 (1987); J. Wirdder, E. Klinke, M. N.
Sathyanarayana and L. Dulog, J. Coat. Technol., 59 (1987) 45-53; J. Winkler, E. Klinke
and L. Dulog, J. Coat. Technot., 59 (1987) 35-41.
117 H. P. Schreiber, M. Y. Boluk and C. Dufour, J. Appl. Polym. Sci., 39 (1990) 465-
470.
118 D. Gupta, Sot. PZast. Eng., Tech. Pap., 25 (1979) 311.
119 B. Mayo and J. P. Pfau, J. Coat. Technol., 59 (1988) 49.
120 M. A. Claridge et al., Adv. Org. Coat. Sci. Technol. Ser., 10 (1988) 145.
121 J. Prosser, Polym. Paint Colour J., 175 (1985) 390.
122 J. E. Hall, R. Benoit, R. Bordeleau and R. Rowland, J. Coat. Technot., 60 (1988) 49-61.
123 H. G. Voelz, G. Kaempf and H. G. Fit&y, Prog. Org. Coat., 2 (1974) 223-235.
124 L. A. Simpson, Aust. Oil Colour Chem. Assoc., Proc. News, 20 (1983) 6.
125 J. H. Coiling and J. Dunderdale, Prog. Org. Coat., 9 (1981) 47.
126 J. H. Braun, Prog. Org. Coat., 15 (1987) 249.
127 G. Kaempf, W. Papenroth and R. Helm, J. Paint Technol., 46 (1974) 56-63.
128 G. Kaempf and W. Papenroth, XBIth FATIPEC Congress, Congress book, Vol. 13, (1976),
pp. 298-304.
129 W. Hughes, Au&. Oil Colour Chem. Assoc., Proc. News, 11 (1974) 7.
130 J. Dunderdale, W. E. Craker and J. H. Colling, XIIth FATIPEC Congress, Congress book,
(1974), p. 69.
131 H. G. Voelz, G. Koempf and H. G. Fitzky, Farbe+Lack, 78 (1972) 1037-1048.
132 H. G. Voelz, G. Koempf and A. Klaeren, Farbe+Lack, 82 (1976) 805-810.
133 H. G. Voeb, G. Kaempf, H. G. Fit&y and A. Klaeren, ACS Symp. Ser. No. 151, Am.
Chem. Sot., Washington, DC, 1981, pp. 163-182.
134 J. Lacoste, R. P. Singh, J. Boussand and R. Arnaud, J. Polym. Sci., Part A: Polym.
Chem., 25 (1987) 2799-2812.
135 S. Keifer, XUth FATIPEC Congress, Congress book, Vol. 12, (1974), p. 517.
136 S. P. Pappas and R. M. Fischer, .I Paint Techrwl., 46 (1974) 65.
137 R. E. Day and T. A. Egerton, CoUoids Su& 23 (1987) 137.
138 J. A. Simms, J. Coat. Technol., 59 (1987) 45.
139 L. Cutrone, D. V. Moulton and L. A. Simpson, XIXth FATlpEC Congress, Congress book,
Vol. 19, (1988), pp. 23-46.
140 D. Corless, Sue Coat. Aust., 26 (1989) 18.
141 L. Cutrone, D. V. Moulton and L. A. Simpson, Pigm. Resin Technol., 18 (1989) 16-19.
142 New York Society of Paint Technology, J. Paint Technol., 45 (1973) 55.
143 H. G. Voelz, G. Kaempf and A. Klaeren, Farbe+Lack, 86 (1980) 1047-1055.
144 W. Papenroth and P. Koxholt, Wochenbl. PapiqfXrr., 107 (1979) 373-378.
145 R. Schwindt, XIVth FATIPEC Congress, Congress book, Vol. 14, (1978), p. 617.
146 R. Epple and A. Englisch, XIIth FATIPEC Congress, Congress book, Vol. 12, (1974),
pp. 241-248.
147 E. Hoffman and A. Saracz, J. Oil Colour Chem. Assoc., 55 (1972) 101-113.
137

148 P. B. Mitton and D. P. Richards, Tech. Pap., Reg. Tech. Corlf:, Sot. Plast. Eng., Philadelphia
Sect., (1972) 42-60.
149 E. Hoffmamr and A. Saracz, J. Oil Colour Chem. Assoc., 55 (1972) 1079-1085.
150 D. P. Richards and G. W. Bovenizer, J. Paint Technol., 44 (1972) 90-96.
151 R. Schwindt, XVth FATIPEC Congress, Congress book, Vol. 15, (1980), p. 111-24.
152 K. Wolny, Kunststofle, 75 (1985) 407.
153 R. Schwindt, Dtsch. Farben-Z., 33 (1979) 13.
154 K. Rehacek and M. Bradac, KVth FATIPEC, Congress, Congress book, Vol. 15, (1980),
pp. 111-5-23.
155 K. M. Oesterle, Farbe+Lack, 86 (1980) 879.
156 W. H. Daiger and W. H. Madson, J. Paint Technol., 39 (1967) 399.
157 J. H. Braun, J. Coat. Technol., 62 (1990) 37.
158 J. R. Brand, Plast. Compd., 10 (1987) 27.
159 G. hick Jr., T. H. Strickland and J. S. Zannucci, ASTM Spec. Tech. PubL 781, (1982)
35-42.
160 G. Irick Jr., J. Appl. Polym. Sci., 16 (1972) 2387.
16 1 P. R. Harvey, R. Rudham and S. Ward, J. Chem. Sot., Faraday Trans. 1, 79 (1983)
1381-1390.
162 T. A. Egerton and C. J. King, J. Oil Colour Chem. Assoc., 62 (1979) 386-391.
163 A. Mackor, Th. P. Koster and K. Liu, Sixth FATlPEC Congress, Congress book, Vol. 19,
(1988), pp. 197-215.
164 J. Lacoste, R. Arnaud and J. Lemaire, J. Polym. Sci., Polym. Chem. Ed., 22 (1984)
3885-3893.
165 K. Hauffe and R. P. Viswanath, Chem.-Ztg., 104 (1980) 295-298.
166 L. Balducci, F. Gallia, F. Gambino, M. Vista, L. Burlamacchi and G. Pedulli, KVUth
FATTPEC Congress, Congress book, Vol. 17, (1984), pp. 175-191.
167 P. R. Harvey and R. Rudham, J. Chum. Sot., Faraday Trans. I, 84 (1988) 4181-4190.
168 V. F. Nosach et al., USSR Pat. 1070458 (1984).
169 J. H. Braun, J. Coat. Technol., 60 (1988) 67.
170 R. J. Bruehlman et al., Ofi Dig. Fed. Sot. Paint Technol., 33 (1961) 252.
171 F. B. Stieg Jr., J. Coat. TechnoL, 53 (1981) 75; ibid., 53 (1981) 65; ibid., 59 (1987)
96.
172 L. Cutrone, J. Coat. Technol., 58 (1986) 83.
173 J. Hook, J. Coat. Technol., 58 (1986) 81.
174 W. D. Ross, Ind. Eng. Chem., Prod. Res. Dev., 13 (1974) 45.
175 W. D. Ross, J. Paint. TechnoL, 43 (1971) 50.
176 M. Kerker, D. D. Cooke and W. D. Ross, J. Paint Technok, 47 (1975) 33-42.
177 J. A. Seiner, US Pat. 3 654 193 (1972).
178 J. A. Seiner and H. L. Gerhart, Ind. Eng. Chem., Prod. Res. Dev., 12 (1973) 98.
179 A. Baidins and E. W. Gillow, US Pat. 3 979342 (1976).
180 R. W. Kershaw, Aust. Oil Co&a- Chem. Assoc., (Aug.) (1971) 4.
181 R. Lamboume (ed.), Paint and Surface Coatings - Theory and Practice, Wiley, New
York, 1987.
182 G. P. A. Turner, in R. Lamboume (ed.), Paint and Su@x.ce Coatings - Theory and
Practice, Wiley, New York, 1987.
183 S. B. Levinson, Application of Paints and Coatings, Fed. Ser. Coat. Technol., Philadelphia,
PA, 1988.
184 N. S. AIlen and M. Edge, J. Oil Colour Chem. Assoc., 73 (1990) 438-445 and 453.
185 S. J. Bett, P. A. Dworjanyn and J. L. Gamett, J. Oil Colour Chcm. Assoc., 73 (1990)
446-453.
186 J. R. Costanza et aL, Radiation Cured Coatings, Fed. Ser. Coat. Technol., Philadelphia,
PA, 1986.
187 B. N. McBane, Automotive Coatings, Fed. Ser. Coat. Technol., Philadelphia, PA, 1987.
188 H. J. Sireitberger, J. Oil Colour CA-cm. Assoc., 73 (1990) 454.
189 V. Wigotsky, Plast. Eng., 46 (1990) 20.
190 J. P. Casey, in Pulp and Paper Chemistry and Chemical Technology, Chapt. XII, XIX,
Interscience, New York, 1960.
191 B. Alince, W. Bichard and P. Lepoutre, TAPPI J., 68 (1985) 122-123.
192 B. Alince, Colloids Su& 39 (1989) 39.
193 B. Alince, TAPPI J., 70 (1987) 114.
194 W. T. Siuta, US Pat. 3 035 966 (1962).
195 B. Barnard and W. T. Laverick, US Pat. 4239548 (1980).
196 J. W. Vanderhoff, in J. K. Carver and R. W.-Tess (eds.), Applied Polymer Science, Am.
Chem. Sot., Washington, DC, 1975.
197 Printing Ink Handbook, 4th edn., National Association of Printing Ink Manufacturers,
Harrison, NY, 1980.
198 A. M. Wells, Noyes Data Corporation, Park Ridge, NJ, 1976.
199 D. E. Bissett, The pritiirq Ink Manual, 3rd edn., Northwood, London, 1979.
200 Kirk-Othmm Encyclopedia of Chemical Technology, 3rd edn., Wiley, New York, 1983.

You might also like