Вы находитесь на странице: 1из 19

DK4036_book.

fm Page 1 Monday, April 25, 2005 12:18 PM

76
Pigment Dispersion

76.1 Introduction ......................................................................76-1


76.2 A Brief Introduction to Pigments....................................76-2
Pigment Definition • Pigment Particles
76.3 The Dispersion Process.....................................................76-4
Pigment Wetting • Particle Deaggregation and
Deagglomeration • Dispersion Stabilization
76.4 The Role of Surface Energy..............................................76-6
Surface Energy and Surface Area • Surface Energy and Pigment
Wetting • Surface Energy and Destabilization of the
Dispersion • Surface Energy and the Acid–Base Concept
76.5 Mechanisms for the Stabilization of Dispersion.............76-8
Charge Stabilization • Steric or Entropic Stabilization
76.6 Surface Treatment .............................................................76-9
Surfactants • Polymeric Dispersants • Surface Modifying Agents
76.7 Surface Treatment during Pigment Manufacture.........76-10
76.8 Surface Treatment of Pigments: Application ................76-11
Organic Pigments • Inorganic Pigments
76.9 The Characterization and Assessment of
Dispersion ........................................................................76-17
Theodore G. Vernardakis 76.10 Conclusion.......................................................................76-17
BCM Inks USA, Inc. References ...................................................................................76-18

76.1 Introduction
The dispersion of pigments in fluid media is of great technological importance to the coatings manu-
facturers who deal with pigmented systems. The basic aim is to change the physical state of pigments to
achieve desired effects in specific application systems. The dispersion process involves the breaking down
and separation of the aggregated and agglomerated particles that are present in all pigments in their
normal form after their manufacture. Dispersion is not considered to be a process of pulverization but
rather a process of particle separation, homogeneous distribution of the particles in a medium, and
stabilization of the resultant system to prevent reaggregation, reagglomeration, flocculation, and settling.
The process of dispersion must be done efficiently and in the shortest time possible to draw out of the
pigment its maximum color properties at the least cost.
The topic of pigment dispersion in fluid media has been covered extensively in the literature.1−6
Theoretical aspects of pigment dispersion apply equally well to inorganic and organic pigments. In this
chapter, the practical examples of surface treatments apply primarily to organic pigments, but similar
treatments can be carried out on inorganic pigments as well.

76-1

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 2 Monday, April 25, 2005 12:18 PM

76-2 Coatings Technology Handbook, Third Edition

76.2 A Brief Introduction to Pigments


76.2.1 Pigment Definition
Materials are colored by the use of pigments or dyes. Pigments are colored, black, white, or fluorescent
particulate organic or inorganic solids; usually they are insoluble in, and essentially physically and
chemically unaffected by, the vehicle or substrate in which they are incorporated. They alter appearance
by selective adsorption and/or by scattering of light.7
Pigments usually are dispersed in vehicles or substrates for application (e.g., in inks, paints, plastics,
or other polymeric material). Pigments retain a crystal or particulate structure throughout the color-
ation process.
As a result of the physical and chemical characteristics of pigments, pigments and dyes differ in their
application: when a dye is applied, it penetrates the substrate in soluble form, after which it may or may
not become insoluble. When a pigment is used to color or opacify a substrate, the finely divided, insoluble
solid remains throughout the coloration process.
Organic pigments are highly colored, inert synthetic compounds that are usually brighter, purer, and
richer in color than inorganic pigments. Generally, however, they are less resistant to sunlight (some fade
badly on exposure to light), to chemicals (greater tendency to bleed in solvents), and to high processing
temperatures (lower heat stability); quite often too, they are more expensive than inorganic pigments.
Pigments are classified by the Colour Index according to specific pigment name and constitution number.
For example, phthalocyanine blue is known by the C.I. name Pigment Blue 15, and its C.I. number is
74160, while titanium dioxide is C.I. Pigment White 6, C.I. 77891. The great number and variety of
organic and inorganic pigments make it impossible to treat them all in this chapter. References should
be consulted for information on pigment types, chemical and physical properties, methods of preparation,
grades, specifications, and applications. See, for example, References 8−11.

76.2.2 Pigment Particles


Pigments are normally produced in a wet presscake form, which upon drying and grinding or spray
drying assumes the form of a fine dry powder. Presscakes, either at their normal pigment content (20 to
40%) or as “high solids” (50 to 60%), are used by the manufacturers of aqueous pigment dispersions for
paint, textile, and ink applications, as well as by those who produce flushed colors for oil ink or coatings
applications. Dry pigment powders are used in a host of other systems such as solvent inks, coatings,
and plastics. Pigments in the presscake or dry powder form are composed of fine particles, normally in
the submicrometer size range. Their color properties are generally influenced by particle size and particle
size distribution; therefore, an assessment on the degree of dispersion must, above all, be considered in
terms of these critical measurements.12 In general, color properties, such as strength, transparency, gloss,
rheology, and lightfastness of all pigmented systems, are affected to a greater or lesser extent by the size
and distribution of the pigment particles in the dispersion. For example, phthalocyanine blue is first
prepared commercially in a “crude” pigment form having a large particle size, up to 25 µm. As such, it
has little color value and must therefore be reduced to smaller, finer particles to enhance its coloristic
properties. After particle size reduction (down to 0.03 to 0.15 µm), an excellent pigment is obtained,
which exhibits a high degree of tinctorial strength, transparency, and gloss. Typical electron micrographs
of these two materials, showing particle size, are reproduced in Figure 76.1.
Pigment particles normally exist in the form of primary particles, aggregates, agglomerates, and
flocculates. Primary particles are individual crystals and associated crystals as they are formed during
the manufacturing process (Figure 76.2). They may vary in size depending on the conditions of precip-
itation and growth, which are controlled by the pigment manufacturer. The scanning electron photomi-
crograph of Figure 76.2 for micronized sodium chloride (although this is not a pigment) is used only to
illustrate the individual and associated crystals that make up the primary particles of a compound.
Aggregates are collections of primary particles that are attached to each other at their surfaces or crystal
faces and show a tightly packed structure. Agglomerates consist of primary particles and aggregates joined

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 3 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-3

FIGURE 76.1 Scanning electron photomicrograph of copper phthalocyanine blue crude (top) and transmission
electron photomicrograph of copper phthalocyanine blue pigment (bottom) showing particle size differences; Pig-
ment Blue 15.

at the corners and edges in a looser type of arrangement. Aggregates are formed during the manufacturing
process in the course of the ripening period of the precipitates. Agglomerates, most often, are formed
during the drying of the presscakes and the subsequent dry milling of the pigment lumps. Figure 76.3
shows typical arrangements of aggregated and agglomerated pigment particles.
Flocculates consist of primary particles, aggregates, and agglomerates, generally arranged in a fairly
open structure, as shown in Figure 76.4. Flocculates may be broken down easily under shear, but they
will form again when such shear forces are removed and the dispersion is allowed to stand undisturbed.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 4 Monday, April 25, 2005 12:18 PM

76-4 Coatings Technology Handbook, Third Edition

b
a

100 µ 100 µm

FIGURE 76.2 Scanning electron photomicrograph showing primary particles: (a) individual crystals and (b) asso-
ciated crystals of micronized NaCl.

a a

0.5 µm
b

FIGURE 76.3 Transmission electron photomicrograph showing (a) aggregated and (b) agglomerated pigment par-
ticles. D&C Red No. 30, Vat Red 1.

76.3 The Dispersion Process


The primary purpose of dispersion is to break down pigment aggregates and agglomerates to their
optimum pigmentary particulate size (down to individual single particles, if possible) and to distribute
these pigment particles evenly throughout a medium (i.e., the carrier). Usually the carrier is a liquid or
a solid polymeric material that is deformable at high temperatures during processing. To achieve the
optimum benefits of a pigment, both visual and economic, it is necessary to obtain as full a reduction
as possible to the primary particle size. After all, the color strength of a pigment depends on its exposed

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 5 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-5

FIGURE 76.4 Transmission electron photomicrograph showing flocculated pigment particles. Dimethylquinacri-
done magenta, Pigment Red 122.

surface area: the smaller the particle size, the higher the surface area, and thus the stronger the color.
Furthermore, the pigment is generally the most expensive constituent of any pigmented system; therefore,
the user normally wants to obtain optimum performance with the smallest possible amount of pigment.
Ideally, a good pigment dispersion consists chiefly of primary particles, with only a minimum of loose
aggregates and agglomerates. In practice, reduction to the primary particle size is largely determined by
the nature of the pigment (i.e., its dispersibility), by the dispersion system and processing equipment,
and by the end-use requirements of the product.
Dispersion should not be confused with pulverization. The latter is simply a comminution process
whereby large pigment lumps are broken down to smaller units, which constitute the powder form.
Pulverization does not break down the aggregated, agglomerated, and flocculated particles into primary
particles. Dispersion, however, accomplishes this effectively.

76.3.1 Pigment Wetting


It is generally recognized that the dispersion process consists of three distinct stages: wetting, deaggre-
gation−deagglomeration, and stabilization. The wetting stage involves the removal from the surface of
the pigment particles of adsorbed molecules of gas, liquid, and other materials and their replacement
with molecules of the vehicle. In other words, the pigment−air interface in dry pigment powders or the
pigment−water interface in presscakes is replaced by the pigment−vehicle interface. This is accomplished
through preferential adsorption. The efficiency of wetting depends primarily on the comparative surface
tension properties of the pigment and the vehicle, as well as the viscosity of the resultant mix.

76.3.2 Particle Deaggregation and Deagglomeration


After the initial wetting stage, it is necessary to deaggregate and deagglomerate the pigment particles.
This is usually accomplished by mechanical action with devices such as ball mills, bead mills, and two-
roll mills. As the pigment powder is broken down to the individual particles, higher surface areas become
exposed to the vehicle and larger amounts of it are required to wet out newly formed surfaces. During
this stage of deaggregation, the amount of free vehicle in the bulk diminishes; therefore, the viscosity of
the dispersion increases. At higher viscosities, shear forces are greater and the breaking down and
separation of particles become more efficient. It is this process of mechanical breakdown of the aggregates
and agglomerates that demands a high energy input and can become quite costly. Some easily dispersible

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 6 Monday, April 25, 2005 12:18 PM

76-6 Coatings Technology Handbook, Third Edition

pigments have been developed to aid in the reduction of energy requirements. Such pigments are
produced by surface treatment of the pigment during manufacture, with the purpose of reducing or
inhibiting agglomeration−aggregation formation. In many cases, such treatments are highly specific to
a single ink, paint, coating, or plastic medium.

76.3.3 Dispersion Stabilization


The third stage of great importance in the dispersion process is the stabilization of the pigment dispersion.
This ensures that complete wetting and separation of the particles has been reached, and also that the
pigment particles are homogeneously distributed in the medium. If the dispersion has not been stabilized,
flocculation may occur as a result of clumping together of the pigment particles. Flocculation is generally
a reversible process. Flocculates typically break down when shear is applied and will form again when
the shear is removed. Where a pigment dispersion is not stabilized by the action of resin molecules in
the vehicle, the use of surfactants or polymeric dispersants can be considered. Such additives may be
used directly during pigment manufacture, or they may be incorporated in the vehicle.

76.4 The Role of Surface Energy


It is well known that molecular forces at the surface of a liquid are in a state of imbalance. The same is true
of the surface of a solid, where the molecules or ions on the surface are subject to unbalanced forces of
attraction normal to the surface plane. Such atoms do not have all their forces satisfied by union with other
atoms. As a result, there is a net force, which tends to pull the surface molecules into the bulk. The opposing
force, which resists this inwardly pulling force, is known as the surface tension or surface energy. All liquids
and solids have surface energies to a greater or lesser degree. To satisfy these surface forces, liquids and
solids tend to attract and retain on their surfaces dissolved substances in the solution or gasses from the
surrounding atmosphere. These forces are short-ranged attractive forces, known as van der Waals or London
forces, and they play a very important role in particle aggregation, wetting, and dispersion stabilization.

76.4.1 Surface Energy and Surface Area


Pigments having a very small particle size exhibit high surface area and, consequently, high surface
energies. As large pigment particles are broken down into several smaller particles, new surfaces are
constantly created, contributing to a higher surface area and thus a higher surface energy.
Let us assume that a pigment powder has a surface area S of 60 m2/g and a density ρ of 1.0 g/cm3. Its
basic particle diameter D from

6
D=
ρS

will be 0.1 µm. If these particles are cubic in structure, and if, for the sake of simplicity, we assume that
a 1 cm3 of pigment is broken down into particles 0.1 µm in size, then 1 ´ 1015 particles will be produced.
We assume also that the particles are in perfect cubic packing. To get an idea of the area created by the
new surface, we need only compare the surface area of 6 cm2 for the 1 cm cubic particle to the surface
area of 600,000 cm2 (60 m2) for the 1 ´ 1015 cubic particles that are 0.1 µm in size. The increase in surface
area is 100,000-fold. The new surfaces produced are tremendously large. The surface energies associated
with these new surfaces are also quite large. These van der Waals surface energies create the attraction
between the submicrometer particles that come together to form the aggregates and agglomerates.

76.4.2 Surface Energy and Pigment Wetting


Surface energies play an important role in the wetting and stabilization of pigment dispersions. For
wetting to be effective, the wetting energies of the pigment−vehicle interface must be greater than the

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 7 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-7

γLV

Vapor

Liquid
θ

γSV Solid γSL

FIGURE 76.5 Partial wetting of a solid surface by a liquid in accordance with the Young-Dupré equation.

sum of the adsorption energy (this is because of substances absorbed on the pigment surface) and the
attractive energy that holds the pigment particles together. Generally, lower energy (low surface tension)
liquids, such as aliphatic and aromatic hydrocarbons, will spread over, or wet, higher energy surfaces.
Quite often, it happens that a liquid does not spread over a pigment surface completely. This occurs
when a high-energy liquid (high surface tension), such as water, will not entirely wet out a high-energy
surface. In this case, the wetting energy is equal to or less than the sum of the adsorption and interparticle
attraction energies, and wetting may be either partial or nonexistent. The liquid will not spread entirely
over the surface, as shown in Figure 76.5. The relationship that describes such a system is given by the
Young-Dupré equation as

γSV = γSL + γLV⋅cos θ

where γSV, γSL, and γLV are the interfacial energies at the solid−vapor, solid−liquid, and liquid−vapor
interfaces, respectively, and θ is the contact angle. For complete wetting, the contact angle is zero (cos θ
becomes unity), and the liquid spreads entirely over the solid surface. For θ > 0, wetting either is
incomplete or does not occur.

76.4.3 Surface Energy and Destabilization of the Dispersion


Surface energies play an important role in the destabilization of the dispersion. Particles dispersed in
liquid media are in constant motion (thermal or Brownian movement). As they move through the
medium, they collide with other pigment particles. The frequency of these collisions depends on the size
of the particles and on the viscosity of the medium. During such collisions, the particles will attract and
may join with other particles because of the powerful short-range London−van der Waals attractive forces,
which, in effect, are surface energies. These forces are electrical and are due to the interaction of the
dipoles that are present in the particles, as permanent dipoles (polar particles) or induced dipoles
(nonpolar but polarized particles).
Once the particles have come together, they may reaggregate or form flocculates if their surface is not
protected, and they will settle to the bottom of the container. This is an undesirable effect for the ink,
paint, or coatings manufacturer. Therefore, to prevent reaggregation or flocculation, such dispersions
must be stabilized.

76.4.4 Surface Energy and the Acid–Base Concept


The idea of surface energy in pigments has been closely related to the acid–base concept, advanced by
Sorensen,13 who has used it to describe the interaction between pigments, binders, and solvents, to
obtain optimally stable pigment dispersions having the best application properties in fluid ink systems.
Such interrelation between the surface energies of these three components in a dispersion can be
characterized by their acid−base properties. Pigments can be classified as acidic (electron acceptors),
basic (electron donors), amphoteric (electron acceptors and donors), or neutral. Binders and solvents
can be similarly characterized.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 8 Monday, April 25, 2005 12:18 PM

76-8 Coatings Technology Handbook, Third Edition

Electrical
Double
Layer

Particle

Charge Stabilization

Steric
Barrier
Anchor
Groups
Particle
Associated Solvated
Solvent Chains

Steric or Entropic Stabilization

FIGURE 76.6 Charge and steric or entropic stabilizations.

Acidic pigments should be used with basic resins (polyamide, melamine, alkyd), while basic pigments
should be used with acidic resins (vinyl, acrylic, maleic). Amphoterics can be used with both resins.
Neutral pigments should be surface treated to improve their dispersion characteristics. The solvent must
have the same acid–base character as the pigment, whereby the interaction between the solvent and the
pigment surface is minimized and at the same time the interaction between the resin binder and the
pigment surface is maximized. In other words, there should be no competition between solvent and binder
for the pigment surface; to obtain maximum dispersion and stability, only the binder should adsorb.

76.5 Mechanisms for the Stabilization of Dispersion


76.5.1 Charge Stabilization
Dispersions may become stable through two generally accepted mechanisms: charge stabilization and
steric or entropic stabilization. Charge stabilization is due to electrical repulsion forces, which are the
result of a charged electrical double layer surrounding the particles as shown in Figure 76.6. The charged
electrical double layer developed around the particles extends well into the liquid medium, and since all
the particles are surrounded by the same charge (positive or negative), they repel each other when they
come into close proximity.

76.5.2 Steric or Entropic Stabilization


Steric stabilization is due to steric hindrance resulting from the adsorbed dispersing agent, the chains of
which become solvated in the liquid medium, thus creating an effective steric barrier that prevents the

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 9 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-9

other particles from approaching too close. This phenomenon is also called entropic stabilization, because
as the coated particles approach each other, the solvated chains of the adsorbed dispersant lose some of
their degrees of freedom, resulting in a decrease in entropy. Such lowering in entropy gives rise to repulsive
forces, which keep the particles away from each other. This type of steric or entropic stabilization is also
represented in Figure 76.6.

76.6 Surface Treatment


76.6.1 Surfactants
Surface active agents or, simply, surfactants are substances that are used to lower the interfacial tension
between a liquid and a solid. Such is the case for pigments in fluid media, with the expressed purpose
of improving pigment dispersibility by improving pigment wetting characteristics, preventing reaggre-
gation, and increasing the stability of the dispersion. A surfactant molecule typically contains two groups
of opposite polarity and solubility. The hydrophilic group is the polar, water-loving part, while the
lipophilic group is the nonpolar, oil loving part of the molecule.
Surfactants are characterized by their HLB value (hydrophile–lipophile balance), which is a ratio of
the hydrophilic to lipophilic groups on the molecule and gives an indication of their solubility in water
or oil-solvent systems. High HLB values mean that the surfactant is soluble in water (an abundance of
hydrophilic groups). Low HLB values, on the other hand, mean that the surfactant is soluble in oil or
solvents (an abundance of lipophilic groups).
Surfactants attach themselves to the pigment particles via preferential adsorption, as shown in Figure
76.7 for aqueous and nonaqueous systems. In aqueous systems, the lipophilic (or hydrophobic) groups
are adsorbed on the particle surface, and the hydrophilic (or lipophobic) groups extend into the bulk of
the aqueous phase to form an effective, protective barrier around the particle. In the case of nonaqueous
solvent systems, the hydrophilic groups of the surfactant are attached to the particle surface, and the
lipophilic groups (tails) extend into and are solubilized by the solvent.
Surface treatments are effective for pigments because their surfaces contain polar or polarized func-
tional groups, which can serve as adsorption sites for the hydrophilic or lipophilic groups of the surfac-
tants. For instance, organic pigments typically contain groups such as nitro (—NO2), hydroxyl (—OH),
carbonyl (—C**=O), amide (—NH—C**=O), methoxy (—O—CH3), chlorine (—Cl), bromine (—Br),
sulfonate (— SO −3 ), carboxylate (—COO–), and metal ions such as Ba+2, Ca+2, Mn+2, and Cu+2, which
can function as the anchoring sites for the hydrophilic or lipophilic groups of the surfactants.
It is well known, however, that classical surfactants do not always improve the dispersion characteristics
of pigments especially when pigment surfaces are low in polarity or nonpolar and are dispersed in
nonpolar vehicles. With such pigments and vehicles, dispersants and surface modifying agents of other
types must be used to improve wetting and dispersibility and to prevent flocculation of pigment particles.

76.6.2 Polymeric Dispersants


Polymeric dispersants or “hyperdispersants”14 are claimed to be more effective dispersion stabilizers for
nonaqueous systems. These substances have a two-part structure, one consisting of an anchoring func-
tioning group (or groups) and the other consisting of a polymeric solvatable chain to which the functional
group is attached. They are, in effect, polymeric surfactants or dispersants but were developed for use
in specific nonaqueous systems, where classical surfactants have limitations. When they are used as
dispersants for organic pigments, it is preferable that they have multiple anchoring groups on one
polymeric chain, because organic particles are not as strongly polar as inorganic particles. Such dispers-
ants may be of a fatty polyester type, containing a carboxy group at the end [e.g., poly(12-hydroxystearic
acid)] with the carboxy group functioning as the anchor and the polyester group as the solvated chain.
Others with multiple anchor groups are fatty polyureas and polyurethanes, which may even contain
polymeric solvatable groups instead of the long fatty chains.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 10 Monday, April 25, 2005 12:18 PM

76-10 Coatings Technology Handbook, Third Edition

Hydrophilic Groups Lipophilic Groups

Particle

Aqueous System

Hydrophilic Lipophilic
Groups Groups
Particle

Non-aqueous System

FIGURE 76.7 Surfactant attachment on pigment particles in aqueous and nonaqueous dispersions.

76.6.3 Surface Modifying Agents


Surface modifying agents are another group of additives that can be used to aid the dispersion of organic
pigments in organic media. These agents are often pigment derivatives (e.g., large flat dye molecules),
which provide improved resistance to flocculation and greater stability to the dispersion. The pigment
derivative is adsorbed onto the pigment surface via the van der Waals attractive forces, which act over a
large area, because such large planar dye molecules lie flat on the pigment surface. They may be used
either alone or in conjunction with a polymeric dispersant. When used alone, they introduce or increase
on the surface of nonpolar or low polarity pigments, the number of polar sites, which are necessary to
interact with the resin in the vehicles, to stabilize the dispersion. When used together with the polymeric
dispersant, they provide anchoring sites on which the anchor groups of the dispersant will become
attached. In this context, they can be used synergistically with dispersing agents, at which time they are
called colored synergists.

76.7 Surface Treatment during Pigment Manufacture


Generally, surface-treated pigments are more easily dispersible, produce more stable dispersions in fluid
media with improved flow, and impart higher strength and gloss to the printed films, when compared
with untreated pigments. Surface treatments can be carried out at different stages of pigment manufac-
ture. Some pigments are prepared directly as finished products, while others are in the form of a crude
pigment that must be conditioned into the pigmentary state.
Use of surfactants is typically made at the initial stage of pigment manufacture. During the precipitation
of the intermediate (e.g., diazo in the preparation of azo pigments), surfactants are used to wet out and
control the fineness of the precipitate, and they may also act as promoters to accelerate the azo coupling
reaction. At the second stage, during the precipitation of the pigment (e.g., in the azo coupling reaction),

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 11 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-11

Barium Salt of Abietic Acid Fatty Acid

CH3


(CH2)n


C— O— O—C C — O—
O — O O
Ba H

Pigment

FIGURE 76.8 Treatment of pigment surfaces with rosins and fatty acids.

surfactants may be used in the dispersion of the pigment particles as they are being formed — for
example, in azo yellows, which are precipitated in the pigmentary state, or in the dispersion of the
precursor (dyestuff), as in the case of metallized azo reds (which are first formed as sodium salts), to
control the salt formation (barium, calcium, etc.), and thus produce the final pigment. At the third stage,
during the conditioning of the pigment, surface treatments are used for pigment particle dispersion, for
coating the pigment surface to prevent aggregation, and for controlling the growth of crystal particles.
If particles are too difficult to filter, use of a specific additive (flocculant) sometimes induces controlled
flocculation and facilitates filtration. Complex formation with additives may also be carried out during
this conditioning stage to stabilize the particles and increase dispersibility, as is the case with diarylide
yellows, which may be surface treated with fatty amines to produce Schiff base stabilization and result
in easily dispersible pigments.

76.8 Surface Treatment of Pigments: Application


76.8.1 Organic Pigments
The published literature dealing with surface treatments of organic pigments, patented or otherwise, is
so extensive that no attempt is made to review it, although it may be referred to occasionally. Readers
are urged however, to consult the review by Hayes,15 which covers the role of classical surfactants,
polymeric dispersants, and pigment derivatives in surface treatments. Other reviews of interest are those
by Topham,16 Merkle and Schafer,17 and Hampton and McMillan,14 the latter dealing specifically with
polymeric dispersants. Further examples will be presented here.
It is well known to the pigment manufacturer that rosination is perhaps the oldest surface treatment
known, especially for azo pigments, where rosin (abietic acid) is precipitated onto the pigment surface
as the barium or calcium salt. It can also be used to treat other pigments, such as copper phthalocyanine
blue,18 and for a host of similar applications, in a polymerized form. Along the same lines, long-chain
carboxylic acids (fatty acids) have also been used to treat pigment surfaces.19 A likely arrangement of
these molecules adsorbed on the surface is shown in Figure 76.8. The hydrophilic anchor groups are
attached to the surface, with the lipophilic groups projecting outward. The use of rosin has been men-
tioned because of its historical significance and because it is still widely used today, as it is one of the
least expensive surface treating agents.
In the course of a study by the author for the development of a diarylide yellow AAOT (Colour Index
Pigment Yellow 14) for flexographic ink applications, it was found that a desirable product was one
prepared in the presence of an amine type ethoxylated guanidine weakly cationic surfactant, in combi-
nation with a polar tetramethyl decynediol solvent.20 It appears that these two surface active agents worked
synergistically to produce a strong, transparent and nonflocculating pigment, as opposed to products in
which the surfactant or the solvent or both were absent from the preparation. Transmission electron
micrographs and particle size distributions for such pigments are shown in Figure 76.9 and Figure 76.10.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 12 Monday, April 25, 2005 12:18 PM

76-12 Coatings Technology Handbook, Third Edition

Diamaver = 0.104 mm

No. of Particles
16

0.05 0.10 0.15 0.20


Particle Diameter (µm)

FIGURE 76.9 Transmission electron photomicrograph and particle size distribution of an untreated diarylide yellow
AAOT, Pigment Yellow 14.

It is apparent that the particle size of the treated sample is smaller (average particle diameter = 0.073
µm) and much more uniform (narrow distribution) than that of the untreated sample (average particle
diameter = 0.104 µm and wider distribution). To show pigment flocculation in flexographic inks, optical
photomicrographs were obtained. Figure 76.11, a micrograph of the liquid ink on a glass slide with cover
for the untreated pigment, exhibits flocculation. Figure 76.12, on the other hand, represents the liquid
ink prepared with the surface-treated pigment and shows that this is a nonflocculating pigment.
Polymeric dispersants21 such as poly(12-hydroxystearic acid) are reportedly used both as free acid
and as a salt with a variety of organic toners; these agents show more effectiveness when reacted with
a primary amine (3-dimethylaminopropylamine, 3-octade-cylaminopropylamine, etc.). The latter types
can be used with pigment derivatives to produce a synergistic effect on the pigment surface for improved
dispersion. An example is copper phthalocyanine sulfonic acid.22 The mechanism of synergism is
illustrated in Figure 76.13 for the surface treatment of copper phthalocyanine blue. A great number of
other phthalocyanine derivatives have also been prepared and used as pigment stabilizers for phthalo-
cyanine blue.23
Phthalocyanine pigments may be conditioned from the crude state to the pigmentary form, for
example, by milling the “crude” with a phthalocyanine derivative24 such as a sulfonated phthalimidom-

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 13 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-13

24
Diamaver = 0.073 µm

No. of Particles
16

0
0.05 0.10 0.15 0.20
Particle Diameter (µm)

FIGURE 76.10 Transmission electron photomicrograph and particle size distribution of a surface-treated diarylide
yellow AAOT, Pigment Yellow 14.

ethyl phthalocyanine25 in the absence of any milling of grinding aid.26 These large planar molecules appear
to lie flat on the copper phthalocyanine surface, as shown in Figure 76.14, and they impart stability to
the dispersions when used in printing inks, paints, and coatings, without any additional conditioning of
the milled product.
Pigment derivatives are by no means limited to phthalocyanines. Quinacridone pigments have been
surface treated with sulfonated quinacridone derivatives27 either as the sulfonic acid form or as the metal
sulfonate salt, with a wide range of metals possible. As in the preceding cases, the planar sulfonated
quinacridone molecules appear to lie flat on the quinacridone pigment surface and thus improve con-
siderably the dispersion properties of the pigment, especially when used in coating applications. Figure
76.15 represents the arrangement of sulfonated quinacridone derivative on the pigment surface.
Pigment derivatives of azo red,28 oranges, and yellows29 have also been used for surface treating the
corresponding pigments. With azo yellows, treatments can be carried out in situ with fatty amines to
produce easily dispersible products through a Schiff base reaction between the —C**=O (carbonyl)
groups of the pigment and the —NH2 groups of primary amines, to form —C**=N— Schiff bases.29−32
Derivatives of monoarylide and diarylide yellow pigments can also be prepared by reacting the pigment
with a primary diamine and a glycidyl ether33 to produce a Schiff base. The structure of one of these

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 14 Monday, April 25, 2005 12:18 PM

76-14 Coatings Technology Handbook, Third Edition

FIGURE 76.11 Optical photomicrograph of liquid ink prepared with untreated diarylide yellow AAOT. Shows
flocculation of pigment particles. Same pigment as that of Figure 76.9.

FIGURE 76.12 Optical photomicrograph of liquid prepared with surface-treated diarylide yellow AAOT. Does not
show flocculation of pigment particles. Same pigment as that of Figure 76.10.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 15 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-15

(CH2)5 — CH3
O


HO — CH — (CH2)10 — C — O — H :
n
Poly (12-Hydroxystearic Acid)

CuPc — SO3H : Copper Phthalocyanine Sulfonic Acid

NH2 — (CH2)3 — NH — (CH2)17 — CH3 :


3-Octadecylaminopropylamine

Polyester
CH3


Chain
(CH2)17


− + O C — NH — (CH2)3 — NH
CuPc — SO3 H

Copper Phthalocyanine Blue

FIGURE 76.13 Surface treatment of copper phthalocyanine blue, Pigment Blue 15, showing the synergistic effect
between sulfonated copper phthalocyanine and a polymeric dispersant on the pigment surface.

O
C—

CuPc — CH2 — N — (SO3H)y


C—
O x

Sulfonated Phthalimidomethyl
Copper Phthalocyanine

CuPc — CH2 — N — C — (SO3H)y


C
O
x

Copper Phthalocyanine Blue

FIGURE 76.14 Surface treatment of copper phthalocyanine blue, Pigment Blue 15, with a sulfonated copper phth-
alocyanine additive, surface modifying agent.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 16 Monday, April 25, 2005 12:18 PM

76-16 Coatings Technology Handbook, Third Edition

H O


N C
(SO3M)x — — (SO3M)y
C N


O H
Sulfonated Quinacridone Red
M = H, Al, Mg, Zn, Cu, Ni, Cd, Cr, Co, Mn

H O


N C
(SO3M)x —
— (SO3M)y
C N


O H

Quinacridone Red

FIGURE 76.15 Surface treatment of quinacridone red, Pigment Violet 19, with a sulfonated quinacridone additive,
surface modifying agent.

H2N — R — NH2 : Primary Diamine


O

R′OCH2 — CH — CH2 : Glycidyl Ether


H—N N—H

O C Cl Cl C O

H—C— N N— — —N N — C— H

(R′OCH2CHCH2)2N — R — N C C N —R —N(CH2CHCH2OR′)2

OH CH3 CH3 OH

FIGURE 76.16 Diarylide yellow AAA, Pigment Yellow 12, derivative; Schiff base.

derivatives in shown in Figure 76.16 for Pigment Yellow 12, AAA yellow. Again, the planar pigment
molecule appears to lie on the pigment surface, and the long chains project outward into the vehicle to
produce stabilization of the dispersion.

76.8.2 Inorganic Pigments


Titanium dioxide, in the two naturally occurring crystal forms, anatase and rutile, is the most important
white pigment, which provides maximum opacifying power. Normally, TiO2 pigments are not used in
their pure form because of their poor dispersibility in a variety of resins and solvents. Generally, they
are surface coated with small amounts of alumina, silica, or both (up to 3% total, on TiO2) to increase
the functionality of the surface (active adsorption sites for the resin molecules) and to improve dispers-
ibility and impart stability to the dispersion, especially in alkyd resin paint systems. For alumina-coated
titanium dioxide,34 the highly basic sites on the alumina surface, which are much more basic than the
sites on the TiO2 surface, cause specific adsorption of the acidic functional groups of the alkyd resin

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 17 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-17

molecules. The remaining parts of the resin molecules (long chains) extend away from the surface,
creating a considerable amount of steric hindrance around each pigment particle, thus resulting in steric
stabilization of the dispersion.
Alumina-coated titanium dioxide, iron oxide red, and other inorganic pigments and fillers can be
surface treated with alkanolamines (aminoalkanols), having the general formulas

R 1 − CH − CH 2 − NH 2 , R 2 − CH − CH − R 3 , etc.
| | |
OH OH NH 2

where R1, R2, and R3 are alkyl groups containing from 1 to 22 carbon atoms in the chain.35 The dispers-
ibility of these pigments is increased considerably when used in paint formulations containing air drying
resin vehicles. The stability of the dispersion is similarly improved because of the steric stabilization
imparted to the pigment particles by the R1, R2, and R3 long chain alkyl groups.
Organic isocyanate adducts36 are used as effective dispersing agents for several classes of inorganic
pigments, including zinc oxide, iron oxides, Prussian Blue, cadmium sulfide, ultramarine, vermilion, and
chrome pigments (zinc, barium, and calcium chromates). These agents improve the dispersion charac-
teristics and the flocculation resistance of the above-listed pigments when incorporated into conventional
alkyd paint vehicles with organic solvents, where these systems also contain a substantial amount of
titanium dioxide.

76.9 The Characterization and Assessment of Dispersion


The extent to which a pigment is dispersed in the medium or the degree of dispersion is normally assessed
in terms of color strength, gloss, brightness, and transparency, and it also has an effect on the rheological
properties of the system.37−39 Since all these properties are governed by the size and distribution of the
pigment particles in the dispersion, one can, today, measure these properties using any of the latest
particle size analyzers based on the light scattering principle of the dispersed particles.12 With these
instruments, a very dilute suspension is required, and it is necessary to know the refractive index and
viscosity of the suspending medium. The average particle diameters and the particle size distributions
obtained are those of individual particles, aggregates, agglomerates, and flocculates in the dispersion.
The advantages of these instruments are that they are quite easy to operate, they give results rapidly, and
they allow the dispersion process to be followed at different times and at different stages.
One such instrument is the Coulter model N4 Submicron Particle Analyzer. Figure 76.17 represents
the particle size results for a green-shade phthalocyanine blue, C.I. Pigment Blue 15:3, in an aqueous
dispersion. The distribution is quite narrow, and the mean particle diameter is 0.117 µm. These results
are very similar to those obtained from inspection of the transmission electron micrographs of Figure
76.1 for the same phthalocyanine blue pigment in the dry powder form, showing that very little aggre-
gation exists in the dispersion.
Such particle size analyzers, based on light scattering, can be used very effectively to study particle size
changes that occur during the dispersion of pigments in fluid systems. Furthermore, time studies may
be carried out on the flocculation of pigments by determining particle size immediately after dispersion
and then later, after the dispersions have been allowed to stand for certain periods. This gives a measure
of the stability of the dispersion.

76.10 Conclusion
There is no question as to the desirability and effectiveness of a fully dispersed and stabilized pigmented
system. Such a dispersion brings out the optimum color properties of the pigment in terms of color
strength, gloss, transparency, and rheology. When a pigment is completely dispersed, it contains a larger

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 18 Monday, April 25, 2005 12:18 PM

76-18 Coatings Technology Handbook, Third Edition

SDP Intensity Results


Cumulant Results
Sample ID : CuPc Aqueous Disp.
Sample ID : CuPc Aqueous Disp.
Mean Diameter = 117 nm
Mean Diameter = 118 nm
S.D. = 30 nm C.V. = 26%
95% Limits = 118 to 118 nm
Size S. D. Amount
Standard Deviation = 37 nm
1: 117 nm 30 nm 100 %

20

15 SDP Differential Intensity


Amount in (%)

Size (nm) Amount (%)


10
31.6 0
46.4 0
5 68.1 4
100 53
147 41
215 2
0 316 0
10 100 1000 464 0
Particle Diameter (nm)
SDP Differential Intensity

FIGURE 76.17 Particle size results of a copper phthalocyanine blue, Pigment Blue 15:3, aqueous dispersion by the
Coulter Model N4 Submicron Particle Size Analyzer, used to assess the degree of dispersion (SDP = Size Distribution
Program).

number of primary particles; therefore, a smaller amount is required to produce the necessary coverage
and color strength than would be necessary for a pigment that was not as well dispersed and contained
a larger number of aggregates, agglomerates, and flocculates.
The trend today is toward production of more easily dispersible pigments, as counterparts to the easily
dispersible azo yellows, which are already used widely in certain printing ink systems. Pigment manu-
facturers are always improving pigment dispersibility, through the use of surface treatments, in terms of
surfactants, polymeric dispersants, and pigment derivatives. The end result is the achievement of complete
dispersion easily and quickly. Since this is an energy-intensive process, in terms of the dispersion equip-
ment utilized, less energy is required, which results in greater economic benefits for the pigment user.

References
1.G. D. Parfitt, Ed., Dispersion of Powders in Liquids, 2d ed. New York: John Wiley, 1973.
2.T. C. Patton, Paint Flow and Pigment Dispersion, 2d ed. New York: John Wiley, 1979.
3.V. T. Crowl, J. Oil Colour Chem. Assoc., 55, 388 (1972).
4.O. Hafner, J. Oil Colour Chem. Assoc., 57, 268 (1974).
5.W. Carr, J. Oil Colour Chem. Assoc., 61, 397 (1978).
6.D. M. Varley and H. H. Bower, J. Oil Colour Chem. Assoc., 62, 401 (1979).
7.H. M. Smith, Polym. Paint Color J., 175, 660 (1985).
8.P. A. Lewis, Ed., Pigment Handbook, Vol. 1, 2d ed. New York: John Wiley, 1988.
9.G. D. Parfitt and K. S. W. Sing, Eds., Characterization of Powder Surfaces. London: Academic Press,
1976.
10. H. P. Preuss, Pigments in Paint. Park Ridge, NJ: Noyes, 1974.
11. W. M. Morgans, Outlines of Paint Technology, Vols. 1 and 2, 2d ed. London: Charles Griffin, 1982.

© 2006 by Taylor & Francis Group, LLC


DK4036_book.fm Page 19 Monday, April 25, 2005 12:18 PM

Pigment Dispersion 76-19

12. T. G. Vernardakis, Am. Ink Maker, 62(2), 24 (1984).


13. P. Sorensen, J. Paint Technol., 47, 31 (1975).
14. J. S. Hamptom and J. F. MacMillan, Am. Ink Maker, 63(1), 16 (1985).
15. B. G. Hays, Am. Ink Maker, 62(6), 28 (1984).
16. A Topham, Prog. Org. Coat., 5, 237 (1977).
17. K. Merkle and H. Schafer, in Pigment Handbook, Vol. III. T.C. Patton, Ed. New York: John Wiley,
1973, pp. 157−167.
18. A. E. Ambler and R. W. Tomlinson, U.S. Patent 3,296,001 (January 3, 1967), ICI.
19. T. C. Rees and R. J. Flores, U.S. Patent 4,032,357 (June 28, 1977), Sherwin-Williams.
20. T. G. Vernardakis, Dyes Pigments, 2, 175 (1981).
21. J. F. Stansfield and A. Topham, U.S. Patent 3,996,059 (December 7, 1976), ICI.
22. P. K. Davies, L. R. Rogers, J. F. Stansfield, and A. Topham, U.S. Patent 4,057,436 (November 8,
1977, ICI.
23. Anon., British Patent 1,544,839 (April 25, 1979), BASF.
24. W. H. McKellin, H. T. Lacey, and V. A. Giambalvo, U.S. Patent 2,855,403 (October 7, 1958),
American Cyanamid.
25. V. A. Giambalvo and W. Berry, U.S. Patent 3,589,924 (June 29, 1971), American Cyanamid.
26. S. L. Johnson, G. McLaren, and G. H. Robertson, U.S. Patent 4,448,607 (May 15, 1984), Sun
Chemical.
27. E. E. Jaffe and W. J. Marshall, U.S. Patent 3,386,843 (June 4, 1968), DuPont.
28. J. Mitchell and A. Topham, U.S. Patent 3,446,641 (May 27, 1969), ICI.
29. J. Mitchell and A. Topham, British Patent 1,139,294 (January 8, 1969), ICI.
30. Anon., British Patent 1,080,115 (August 23, 1967), KVK.
31. F. Dawson, J. Mitchell, L. R. Rogers, W. Todd, and A. Topham, British Patent 1,096,362 (December
29, 1967), ICI.
32. G. H. Robertson, U.S. Patent 4,220,473 (September 2, 1980), Sun Chemical.
33. R. J. Schwartz and T. Sulzberg, U.S. Patent 4,468,255 (August 28, 1984), Sun Chemical.
34. M. J. B. Franklin, K. Goldsbrough, G. D. Parfitt, and J. Peacock, J. Paint Technol., 42, 740 (1970).
35. H. Linden, H. Rutzen, and B. Wegemund, U.S. Patent 4,167,421 (September 11, 1979), Henkel.
36. F. Hauxwell, J. F. Stansfield, and A. Topham, U.S. Patent 4,042,413 (August 16, 1977), ICI.
37. W. Carr, J. Oil Colour Chem. Assoc., 65, 373 (1982).
38. K. Tsutsui and S. Ikeda, Prog. Org Coat., 10, 235 (1982).
39. R. Polke, Am. Ink Maker, 61(6), 15 (1983).

© 2006 by Taylor & Francis Group, LLC

Вам также может понравиться