You are on page 1of 38

Progress in Reaction Kinetics and Mechanism. Vol. 26, pp337–374.

2000
1468-6783 © 2001 Science Reviews

MECHANISM OF ACTION OF TITANIUM DIOXIDE


PIGMENT IN THE PHOTODEGREDATION OF
POLY(VINYL CHLORIDE) AND OTHER
POLYMERS

Terence J. Kemp and Robin A. McIntyre

Department of Chemistry, University of Warwick, Coventry CV4 7AL, UK

Contents
ABSTRACT 338

1. INTRODUCTION 338
1·1 Properties of TiO2 338
1·2 Role of TiO2 as a pigment in polymers 339
1·2·1 Pigmenting of poly(vinyl chloride) 341
1·2·2 Pigmenting of polyolefins 341
1·3 Photodegradation of unpigmented poly(vinyl chloride) 341
1·4 Photodegradation of unpigmented polyolefins 345
1·5 Photodegradation of unpigmented polystyrene 348
1·6 Photoactivation of TiO2 348
1·7 Modifications of TiO2 352
1·7·1 Coating of TiO2 352
1·7·2 Doping of TiO2 353

2. POLY(VINYL CHLORIDE) – TiO2 SYSTEMS 353


2·1 Gross action of TiO2 in PVC photolysis 353
2·2 Mechanisms operative in PVC–TiO2 systems 354
2·3 Mechanism of chalking 361
2·4 Mechanism of greying 361
2·5 Mechanism of pinking 362
2·6 Effect of modification of TiO2 on photodegradation 363
2·6·1 Effect of coating 364
2·6·2 Effect of doping 366

337
338 Terence J. Kemp and Robin A. McIntyre

3. POLYOLEFIN – TiO2 SYSTEMS 368


3·1 Polyethylene 368
3·2 Polypropylene 369

4. OTHER POLYMER – TiO2 SYSTEMS 369


4·1 Polycarbonate 369
4·2 Acrylates and paints 370
4·3 Polystyrene 371

5. CONCLUSIONS 371

6. ACKNOWLEDGEMENTS 372

7. REFERENCES 373

ABSTRACT
The extensive use of titanium dioxide as a pigment and optical agent in polymers,
particularly poly(vinyl chloride), polyethylene and alkyd resins, has prompted
many investigations of the various deleterious effects (loss of gloss and mechanical
properties, chalking, pinking) of long-term exposure to sunlight of the pigmented
polymers. This short review summarises the fundamental semiconductor processes
undergone by TiO2 on UV irradiation and the consequent chemical action of the
reactive intermediates so generated. It also describes the strategies developed to
counteract photodegradation induced by TiO2, in particular its encapsulation in a
coating by near-UV transparent metal oxides, such as Al2O3 and SiO2 , with wide
band gaps. The effects of doping TiO2 with transition metal ions are also noted.

Prog React Kinet Mech 26:337–374 © 2001 Science Reviews

KEYWORDS: anatase, chalking, dehydrochlorination, free radical, greying,


hydroxyl radical, photodegradation, pinking, polyene, polyethylene, polystyrene,
poly(vinyl chloride), rutile, semiconductor, superoxide ion, titanium dioxide

1. INTRODUCTION
1·1 Properties of TiO2
Titanium dioxide has found extensive use as a white pigment in a variety of
formulations, being best known for its incorporation into paints and polymers,
Mechanism of action of titanium dioxide pigment in the photodegradation 339

especially poly(vinyl chloride) (PVC) as used in window profile, facia boards,


etc. and in polyethylene (PE) to give the familiar white plastic bag (Section 1·2).
It also finds uses as a colorant in coatings, paper and inks. The superlative light
scattering performance of one of its phases, rutile (see below), is associated with
its having the highest refractive index of colourless, stable compounds. It has no
toxic properties, which prompted its replacement of white lead as the principal
colorant in white paint. Its semiconductor properties, summarised in Section 1·6,
have led to its incorporation into a variety of devices aimed at converting light
energy into electrical energy (in photovoltaic cells) or chemical energy (in water-
splitting systems). Its highly reactive nature on photoactivation has led to its
application as a photoactivated catalyst in photomineralisation systems [1,2],
particularly in the form of a powder coating on a glass support, and in gas-phase
systems, such as the photocatalytic reactions of water vapour with alkynes and
alkenes reviewed by Anpo [3]. Accordingly, a huge literature on this material has
built up since the mid-1970s and this review will confine itself to the role of TiO2 as
a pigment in polymers to give systems which can experience excitation by incident
near-UV light to instigate a variety of processes within its polymeric host leading to
discoloration, loss of gloss, surface deterioration and other undesirable features.
Titanium is abundant in nature, occurring as ilmenite (FeTiO3), rutile
(TiO2), anatase (TiO2) and leucoxene (TiO2. x FeO.yH2O). Titanium dioxide pig-
ments exist in two crystal phases, rutile and anatase (Figure 1), which differ in
terms of lattice structures, refractive indices (Table 1) and densities, with rutile
having a density 4·2 g/cm3 and anatase one of 3·9 g/cm3.
In the 1930s, anatase was made by the sulphate process and it competed as
a pigment for paints with compounds such as lithopone and white lead (PbCO3)2
Pb(OH)2. More recently, pigments based on rutile have begun to replace anatase
because of its light-scattering advantage over anatase of about 20% [4].
Two processes are currently used to produce TiO2, namely the sulfate
process and the chloride process, the latter involving flame technology. The chloride
process, pioneered by DuPont in the 1950s, yields a product with a narrower par-
ticle size distribution, an important factor in predicting optical and electronic
behaviour.

1·2 Role of TiO2 as a pigment in polymers


From the 1920s TiO2 pigment production had increased by 1988 to an estimated
2·9 million tonnes per year [5]. The main reasons for this are (i) that TiO2
340 Terence J. Kemp and Robin A. McIntyre

Figure 1 Structures of anatase TiO2 (a) and rutile TiO2 (b).

possesses a very high light scattering efficiency and (ii) it is an inert, non-toxic
material that is safe to use, especially for packaging applications in the food
industry.
The plastics industry accounts for a major fraction of the world’s con-
sumption of TiO2, it being the most widely used pigment in plastics. TiO2 finds
extensive use in products such as white carrier bags (PE), window profiles
(PVC), etc. Loadings of TiO2 in polymers vary from 10% or more, where strong
white coloration is required, to as low as 0·01% when it is used to tint a deep

Table 1. Refractive indices for selected white pigments and relevant polymers [11].
White pigment RI Plastic RI
Rutile TiO2 2·73 Polystyrene 1·60
Anatase 2·55 Polycarbonate 1·59
Zinc Oxide 2·02 Polyethylene 1·50–1·54
Basic carbonate, white lead 1·94–2·09 Acrylic 1·49
Lithopone 1·84 Poly(vinyl chloride) 1·48
Calcium carbonate 1·63
Silica 1·41–1·49
Mechanism of action of titanium dioxide pigment in the photodegradation 341

shade. Large volumes of plastics are produced using titanium dioxide as the pig-
ment and so the influence of the pigment on polymer degradation or stability is
of major economic importance [5].
The topic of degradation of polymers containing various solids, either as
particles (metal oxides, metal powders, etc.) or fibres (glass, metal), as fillers,
reinforcers or colorants, has been covered in books [6,7] and a review article [8].
The photochemistry of pigmented polymers in general is a regular feature in the
invaluable series of reviews on polymer photochemistry by Norman Allen [9].
There are a number of review articles on the role of TiO2 as a pigment in polymers
[10–15]. This review is confined to the action of TiO2 particles used as colorants
and protectants against UV radiation.

1·2·1 Pigmenting of poly(vinyl chloride)


Rigid PVC formulations are a major outlet for TiO2 pigments throughout the
developed world. In the USA, PVC siding formulations contain 8–10 phr
(although 4–5 phr is required for colour). The extra pigment is beneficial to the
polymer matrix as it ensures good UV stability [16]. The durable (coated) grades
of rutile help to protect PVC from UV attack, the main concern being to ensure
that its natural tendency to yellow in sunlight is suppressed and masked [16].

1·2·2 Pigmenting of polyolefins


Polyolefins are the other major outlet for TiO2 pigments in plastics. Applications
range from PE masterbatches (colour concentrates) through to packaging and
internal pipes where there is no requirement for good exterior weatherability.
Recently, environmental issues have dictated the need for making packaging bio-
and photodegradable, in order to minimise litter and waste disposal problems. A
growing market has also emerged for white or white/black coextruded weather-
able PE films for agricultural and outdoor packaging applications. In this appli-
cation, highly durable, heavily coated grades of TiO2 are usually employed, with
the retention of mechanical properties being of importance.

1·3 Photodegradation of unpigmented poly(vinyl chloride)


A considerable literature exists concerning this process and we merely note a few
key or recent papers. PVC as prepared in bulk is not purely –(CH2CHCl)n– but has
a number of intrinsic defects developed as a result of the processing system. Thus
extrusion or injection moulding procedures involve subjecting polymers to
342 Terence J. Kemp and Robin A. McIntyre

mechanical stresses termed mechano-degradation with consequent scission of the


polymer chain to give macroalkyl radicals, which are effective initiators for
thermal oxidation. Scott [17] has outlined a mechanism regarding the mechano-
oxidation of PVC during processing to give centres capable of absorbing UV
light to induce further degradation (Scheme 1):

Scheme 1. Mechano-oxidation of poly(vinyl chloride) during processing [17].

Gardette and Lemaire [18] have devised a simplified scheme for the photo-
discoloration of PVC (Scheme 2)
The photo-oxidation of PVC is initiated by Cl· atoms (formed during gen-
eration of polyenes of increasing length by successive photochemical excita-
tions). The first step involves absorption of light by a chromophoric defect iden-
tified as a -chlorinated diene. The polymer is subsequently photo-oxidised and
the discoloration is then caused by residual polyenes. In more detail [19,20],
when the PVC matrix is exposed to UV light of wavelengths longer than 300 nm,
these competing reactions are activated, as follows:

(i) A process of colour development which can proceed to complete darkening


via a mechanism of photodehydrochlorination:
Mechanism of action of titanium dioxide pigment in the photodegradation 343

Scheme 2. Schematic for photo-discoloration of PVC [18].

→ [(CH=CH)n – CH• – CH2] + Cl•


hv
(i) –(CH=CH)n – CHCl – CH2-
→ –(CH=CH)n+1 – CHCl – CH2- + HCl

(i) Formation of -chlorinated polyenes results from multiple consecutive photo-


chemical processes activated by photons of wavelength  from 300–800 nm.
The HCl evolution has been studied using mass spectrometry [21].
(ii) A bleaching process (photo-oxidation of chlorinated polyenes by O2 in the
PVC matrix).
(iii) An oxidation process induced by free Cl atoms that have not reacted with
neighbouring H atoms in the dehydrochlorination process.

This leads to functional groups that are readily observed by FT-IR spectroscopy,
e.g. , ′-chlorinated ketones 1745 cm–1; acid chlorides 1785 cm–1; -chlorinated
acids 1718 cm–1.
Schemes 3 and 4 show the summary and mechanism, [18,19,22] respec-
tively, for the photodegradation of PVC.
344 Terence J. Kemp and Robin A. McIntyre

Scheme 3. Summary of the photodegradation of PVC.

Scheme 4. Mechanism of photodegradation of PVC [19].


Mechanism of action of titanium dioxide pigment in the photodegradation 345

The accumulation of these oxidised groups, but not of the chromophoric polyene
groups, is directly correlated with the changes of the mechanical properties of the
PVC matrix. A PVC system can therefore suffer a severe loss of functional
properties without significant change in colour, or darken without modification of
mechanical properties. The balance will be dependent upon the oxygenation of
the exposed system.
Following earlier approaches [23], Audouin and colleagues [24] have
carried out depth profiling of thermally stabilised, unpigmented, rigid PVC sam-
ples following photoageing at 40°, 55° and 70°C. Microtome sections (20 m)
revealed carbonyls and chain scissions predominated in the surface layers (espe-
cially in the first 100 m but extending in attenuated form to 200 + m, see
Figure 2). By contrast, the development of the conjugated polyene systems and
crosslinking predominate in a subcutaneous layer (~300–400 m). Similar depth
profiles of these products are found in TiO2 – doped samples (Section 2·2).

1·4 Photodegradation of unpigmented polyolefins


This process has been researched and reviewed exhaustively, and again we refer
to only a few sources [25–28]. The exemplar polymers are polyethylene and
polypropylene. Essentially a polyethylene molecule denoted PH is considered to
form a radical P· following mechano-chemical damage during processing, which
then reacts with oxygen:

Initiation:
PH → P• (1)
Propagation:
P• + O2 → PO2• (2)
peroxyl radical

PO2• + PH → POOH + P• (3)


hydroperoxide

Chain branching:
POOH →
uv
PO• + •OH (4)
alkoxyl radical

2 POOH → PO2• + PO• + H2O (5)


346 Terence J. Kemp and Robin A. McIntyre

Figure 2 Thickness distribution of (a) carbonyls and (b) polyenes in unpigmented and pig-
mented bulk PVC samples after 1400 h of photoageing in photoreactor at 70 C [41].
Mechanism of action of titanium dioxide pigment in the photodegradation 347

PO• + PH → POH + P• (6)

•OH + PH → P• + H2O (7)

Chain scission:
• •
R1R2CH—CR3—CH2R4 → R1 + R2CH=CR3CH2R4 8)


R1R2CH—O• → R1 + R2CH=O (9)

Chain termination merely involves radical – radical reactions of the type

2PO2• → inert products (10)

P• + PO2• → PO2P (11)

2P• → PP (12)

Once carbonyl compounds begin to be formed, they act as agents for further
degradation either by undergoing further scission reactions:


R1CCH2R2 → R1C+ •CH2R2
hv
(13)
 
O O

R1C CH2CH2CH2 R2→ R1C CH3 + CH2=CHR2


hv
(14)
 
O O

or by attacking C–H bonds in an adjacent polymer molecule

R1R2C=O → 1(R1R2C=O)*→ 3(R1R2C=O)


hv
(15)


(R1R2C=O) + PH → R1R2COH + •P
3 hv
(16)

A further pathway is the generation of the aggressive singlet oxygen:


348 Terence J. Kemp and Robin A. McIntyre

3(R R C=O)
1 2
+ 3O2 → 1(R1R2C=O) + 1O2 (17)

(18)

The net effects are chain scission, resulting in the loss of mechanical properties,
and the build-up of oxidation products, notably aldehydes, ketones and car-
boxylic acids, usually characterised in situ by IR spectroscopy.
Allen has reviewed early work on the photodegradation of polypropylene
[29].
Depth profiling of UV-degraded polypropylene samples shows pro-
nounced depth effects for unstabilised, unpigmented samples (Figure 3) [30]. The
reduction in Mw is ca. two orders of magnitude at up to 100 m depth, and is
significant even at 1 mm depth. Unpigmented samples stabilised with a
Tinuvin/Chimassorb /Irganox B215 mixture gave less degradation which was
much less depth-dependent.

1·5 Photodegradation of unpigmented polystyrene


A useful account of the mechanisms of weathering of styrenic polymers is due to
Lemaire et al. [31]. The same products are formed at long and short wavelength
irradiation, although their distributions are different. Short wavelength light
(254 nm) is absorbed in the surface layers to give high local concentrations of
oxidised products, whereas long wavelength light ( > 300 nm) is more penetrat-
ing. Reliance on IR spectroscopy to chart the progress of reactions can give a dis-
torted picture as the considerable number of volatile products can escape detec-
tion. A single mechanistic scheme was formulated by the authors to account for
the considerable variety of products (Scheme 5). The authors also survey photo-
oxidation of various styrene copolymers such as acrylonitrile–butadiene–styrenes
[31]

1·6 Photoactivation of TiO2


Titanium dioxide is a semiconductor with a band gap of 3·1 eV for rutile; this
corresponds to the near-UV region which renders the material colorless, although
its highly refractive, normally powder form, imparts an intense white coloration.
Mechanism of action of titanium dioxide pigment in the photodegradation 349

Figure 3 Molecular size distributions for (unstabilised) polypropylene at different depths after
34 weeks exposure: (a) near the surface; (b) in the interior [30].
350 Terence J. Kemp and Robin A. McIntyre

Scheme 5. Photo-oxidation mechanism of polystyrene at short and long wavelengths [31].


Mechanism of action of titanium dioxide pigment in the photodegradation 351

The diagram below is a simplified model of the band model of TiO2.

Figure 4 Generation of hole-electron pair in TiO2 on absorption of one UV photon

When UV light is absorbed by the surface, an electron is promoted to the


conduction band, leaving behind an electron hole (h+) in the valence band. Both
of these species are very reactive; an electron can travel to the surface of the TiO2
particle and transfer to an adsorbed molecule such as O2, (eqn 19) losing energy
in the process, while the electron hole can travel to the surface and react with a
surface hydroxide ion to yield the extremely reactive •OH radical.

e¯ + O2 → O2¯ • (19)

hole+ + OH– → OH• (20)

The high-energy electron can also recombine with the positive hole (eqn 21) to
release heat

e¯ + h+ → heat (21)

The O2¯• radical, known as superoxide radical ion, quickly reacts with H2O to
form a hydroperoxyl radical, HO2•, (eqn 22).

O2¯• + H2O → HO2• + OH– (22)

The overall reaction can thus be summarised by eqn (23)

UV light + O2 + H2O → HO2• + OH• (23)

Note here that no TiO2 is used in the overall reaction, i.e. it functions as a catalyst.
352 Terence J. Kemp and Robin A. McIntyre

•OH radicals will react with organic molecules impinging on the TiO2
surface, causing them to become oxidised in a sequence of reactions termed
mineralisation [1,2]. Essentially, the oxidation is induced by UV light, and degra-
dation that occurs as a result of UV absorption by TiO2 is referred to as catalytic
or photocatalytic degradation.
A proposed additional pathway in the process is ion-annihilation (eqn 24)
to form singlet O2 which then attacks any sites of unsaturation in the polymer to
yield a reactive unsaturated hydroperoxide [32]:

TiO2+• + O2–• → TiO2 + 1O2 (ion-annihilation) (24)


(hole)

RCH=CHCH2R′ + 1O2 → RCH=CHCH(OOH)R′ (25)

The vast majority of photodegradation studies involving TiO2 refer to aqueous


media, with the TiO2 either maintained as an agitated suspension or as a thin-film
coating on a support, usually glass. The situation in a TiO2-pigmented polymer is
altogether different in that the diffusion of reactive species is, by comparison,
extremely slow. However, the fundamental process of electron–hole pair genera-
tion, with subsequent reactions of these species, is common to solution and solid
polymer systems, as detailed below.

1·7 Modifications of TiO2


Given the many uses of TiO2, much effort has been made to modify TiO2 to
extend its range of applicability. These include: (i) doping the TiO2 with transi-
tion metal ions to moderate its semiconductor properties and extend its spectral
range; (ii) attaching redox-active coordination compounds to the surface to
extend its light-harvesting capabilities into the visible region; (iii) coating the
exterior of the TiO2 particles with a relatively thin coating of silica, alumina, zirco-
nia, or combinations of these, in order to reduce its photoactivity towards its poly-
mer host. In this review we shall restrict coverage to doping and coating of TiO2.

1·7·1 Coating of TiO2


Unusually, the surface of TiO2 is not just made up of titanium and oxygen. As the
particles of the pigment begin to grow, insoluble components can accumulate on
the surface. These components are either impurities that are present in the ore
Mechanism of action of titanium dioxide pigment in the photodegradation 353

(which are not removed during the purification process), or additives, which are
designed to control the crystal structure and growth.
Such a surface may need to be modified as it may not be optimal for its
chosen application [33]. For example, in the sulfate process, the pigment particles
become surrounded by a phosphate layer which imparts a water dispersibility to
them, resulting in an application in the paper industry. To increase dispersion, the
pigment surface is modified using hydrous aluminium oxide; this will lower the
van der Waals’ forces and decrease the particle-particle attraction [34].
To increase their durability and diminish their potential to act as centres
for photodegradation, the pigment particles are encapsulated in an “impervious
oxide” coating, which prevents direct contact between the degradable organic
polymer and the photochemically active TiO2 surface.
The encapsulating medium is usually silica [35], but it can be combina-
tions of amorphous alumina, zirconia, tin and phosphorus oxides [36,37]. In order
to maintain dispersibility, the impervious layer must be located underneath the
alumina layer. The durability is dependent upon the thickness and density of the
coating. For general usage, high durability grades have coating levels of 5–7%
whereas superdurable grades have up to 10% [5]. These coating levels exceed
considerably that required (1%) to give the best dispersibility; accordingly, heav-
ily coated grades are more difficult to disperse in melt plastics and offer less opac-
ity with their lower TiO2 content. The bulk of the pigment used by the plastics
industry is usually light-to-medium coated rutile. Conversely, there is little use of
anatase or uncoated rutile. For weatherable applications, the manufacturers will
usually choose between general purpose, highly durable and superdurable grades.

1·7·2 Doping of TiO2


To be most effective, the dopant should have a different formal charge from the
Ti4+ [38], thus TiO2 has been shown to be activated towards catalytic hydrogen
generation systems by dopants such as Mg2+[39].

2. POLY(VINYL CHLORIDE) – TIO2 SYSTEMS


2·1 Gross action of TiO2 in PVC photolysis
The presence of TiO2 pigment introduces two main effects, namely

i) screening of the interior layers of the PVC from penetrating UV light;


ii) activation of the TiO2
354 Terence J. Kemp and Robin A. McIntyre

Clearly these effects act in opposite directions and their relative impact depends
on a variety of factors (loading, grade, coating of particle, irradiation wave-
length). The net results of these competing effects are assessed in the following
sections.

2·2 Mechanisms operative in PVC – TiO2 systems


In their 1991 paper, Gardette and Lemaire [18] record that ‘no noticeable photo-
catalytic effect of the pigments was observed’, this referring to zinc oxide and
two forms of rutile. Thus the IR changes in the carbonyl region upon irradiation
with polychromatic light at > 300 nm were the same for rutile-doped PVC, as
for unpigmented PVC, with peaks emerging at 1718, 1745 and 1785 cm–1,
although a fourth weak maximum appeared at 1805 cm–1. The rather broad UV-
visible changes occurring with rutile-doped PVC were essentially the same as for
unpigmented PVC. As regards the rates of photo-oxidation, as revealed by IR
changes, these decreased monotonically as the level of TiO2 (both grades) was
increased through the loadings 2, 4, 8 and 16% TiO2. The changes induced in the
UV-visible spectrum (at 450 nm) depended in a more complex way upon the
loading of rutile (Figure 5).
Irradiating at shorter wavelengths ( = 254 nm) followed much the same
pattern except that the contribution of the band at 1745 cm–1 was relatively
smaller owing to the ready decarbonylation of the , ′-dichloroketone at these
short wavelengths. The pigments exerted no protective effect on the ketones. The
UV-visible spectra developed on irradiating unpigmented PVC at 254 nm showed
a band at 284 nm attributed to an -chlorinated conjugated triene, but this could
not be detected in the pigmented samples because of their huge background
absorption at this wavelength.
The authors interpreted the absorption changes at 450 nm at pigment con-
centrations above 2% as follows:

(i) In the first 50 h of irradiation, polyene sequences are formed by direct exci-
tation of the PVC in unpigmented zones. The polyenes formed are protected
by screening by TiO2 and the retrodiffusion effect operative at irr >400 nm.
The authors note that any photo-oxidation of a polyene will produce com-
pounds with reduced lengths of conjugation, i.e. they will absorb at shorter
wavelengths than 450 nm. It is the inhibition of photo-oxidation that leads to
the discoloration.
Mechanism of action of titanium dioxide pigment in the photodegradation 355

(ii) After 50 h of irradiation, the polyenes have reached a concentration able to


compete effectively with the pigment in absorbing the incident light, and
hence a competition develops between photo-discoloration and photo-
bleaching, leading to a reduction and then levelling-off of the rate of dis-
coloration observed in Figure 5. This effect is not observed in weakly or
unpigmented samples because in these the polyenes are formed in the bulk
material, beyond the reach of diffusing oxygen, whereas in the pigmented
samples, all discoloration is confined to the surface layers, reaching a depth
of 40 m.

The protective effect of the pigment increases with pigment concentration


in the first 50 h of irradiation; thereafter the rate of discoloration becomes virtu-
ally independent of the pigment concentration. The authors conclude that no cata-
lytic effect of the pigment is operative at short wavelengths (where the pigment
absorbs) as the pattern of product distribution is largely unaffected.
The same authors [40] demonstrated that TiO2-pigmented PVC exhibits a
latent discoloration that manifests itself only after a period of storage in the dark.
This effect was shown to be reversible, and the discoloration removed by further

Figure 5 Changes in D450 nm versus irradiation time (SEPAP 12-24//60C) for PVC/TiO2 coated
systems: ●: unpigmented sample; ◆: 2 % TiO2; ●: 4 % TiO2; ▲: 8 % TiO2; ▼: 16 % TiO2 [17].
356 Terence J. Kemp and Robin A. McIntyre

irradiation. The mechanism of this intriguing phenomenon can be summarised


[40] in Scheme 6.

Decrease of
conjugation

 >400 nm

Scheme 6. Mechanism of reversible discoloration of PVC-TiO2 system.

The main features are envisaged as follows:

(i) conjugated polyenes are produced by the well-established process of dehy-


drochlorination;
(ii) shorter polyenes, with abs <400 nm, accumulate in the PVC matrix, being
protected by the TiO2 particles;
(iii) longer-chain polyenes, with abs >400 nm, are continuously photo-oxidised
provided the supply of O2 is maintained;
(iv) thermolysis of an irradiated sample induces the isomerisation of those poly-
enes absorbing just below 400 nm, shifting their electronic absorption to
above 400 nm, i.e. developing the ‘latent’ discoloration;
(v) fresh irradiation of such a thermally – developed sample will induce photo-
isomerisation to the original isomer absorbing just below 400 nm.
(vi) A second thermolysis will restore the yellow colour and the reversibility can
be demonstrated through many further cycles.

The dichotomy between the roles of TiO2 as a screening agent and as a


photoactivated source of aggressive free radicals has been explored by examin-
ing the depth profiles of photoaged industrial PVC – (5%) TiO2 samples using IR
and UV-vis spectra of microtomed layers [41] of 20 m thickness to determine
carbonyls (at 1720 cm–1) and polyenes with 10-12 double bonds (at 500 nm). The
thickness distributions shown in Figure 2 confirm that the thickness of the
degraded layer (TDL) for pigmented samples (~ 200 m) is less than that for the
unpigmented samples (~ 400 m). Moreover, the TDL is lower for the carbonyls
Mechanism of action of titanium dioxide pigment in the photodegradation 357

than for the polyenes, as shown previously be Gardette et al. [23]. The distribu-
tion curves for the polyenes show clear maxima, Figure 2, [41, 42] at 50 m
depth for pigmented material and 200 m depth for unpigmented, whereas the
data for the carbonyls reveal maximum oxidative damage at the surface for both
types of material. Also revealing was these authors’ determination of the light
penetration of pigmented and unpigmented samples (Figure 6), which showed the
ratio of the depths to be ca 5 (for unpigmented/pigmented samples) whereas the
ratio of the thicknesses of the degraded layers was less than 3 for carbonyls and
polyenes (Figure2). By examining the kinetics of carbonyl development using
irradiation sources of varying UV outputs, the authors hypothesized that the

Figure 6 Thickness distribution of attenuation T (%) at 365 nm for virgin pigmented (●) and
unpigmented (◆) PVC samples [41].
358 Terence J. Kemp and Robin A. McIntyre

observed penetration of the photoaging was due to radiation of wavelength


~390 nm, close to the absorption cut-off of TiO2. They note [41, 42] that pigments
displaying a strong screen effect can be very efficient at preventing discoloration
if they are able to absorb most of the incident light in the layer where oxygen
is available and can scavenge most of the macroalkyl radicals. The coloured
polyene sequences are formed below this ‘oxidation layer’ and if light attenuation
becomes sufficiently severe then little ‘subcutaneous’ polyene will be formed.
Depth profiles of the kinetics of formation of OOH/OH groups were also
featured in a study of the photodegradation of a PVC-TiO2 formulation (100 parts
PVC, 4 or 10 parts TiO2, 2 parts Irgastab 17M (dibutyltin-S, S′– di(iso-octylthio-
glycolate), 0·3 parts stearic acid ) [43].
The concentration of OOH and CO groups decreased in the presence of
TiO2 and CaCO3; in the presence of CaCO3, OOH groups were formed to a depth
of 60–80 m while in the presence of TiO2, only to 20–40 m. The uptake of O2
decreased in the presence of both pigments. The kinetics of deydrochlorination
were greatly affected by addition of either TiO2 or CaCO3, with TiO2 exerting a
particularly strong suppressant effect (Figure 7), although this may result from
reaction of the HCl released with the pigments. Irradiation-induced deterioration
of mechanical properties of PVC, such as elongation-at-break and tensile
strength, were heavily suppressed by TiO2 at 4 or 10 parts loading while CaCO3,
by comparison, exerted very little protective effect. The picture emerging from
these various measurements was that TiO2 was a much more effective pigment on
almost every count. These authors propose that •OH and HO2• radicals generated
at the TiO2 surface play the main role in instigating oxidative attack.
A feature of more recent studies of the PVC-TiO2 photosystem has been the
incorporation of additional experimental techniques with the aim of building up a
holistic picture of the process. Cho and Choi provide a particularly comprehensive
approach [44] covering UV-visible spectroscopy, weight loss kinetics, gel perme-
ation chromatography (GPC), scanning electron microscopy (SEM), FTIR and
X-ray photoelectron spectroscopy (XPS) of PVC films containing 0–2% TiO2
(Degussa P25). Analysis of plots of apparent absorbance at 350 nm versus film
thickness indicated, for example, that the active UV light that can excite TiO2 is
able to penetrate completely a film of 30 m thickness at a 1·5% loading of TiO2.
The weight loss kinetics were particularly illustrative as seen from Figure
8. Weight loss rates were much higher for the pigmented than the unpigmented
system, and under air compared to under nitrogen. The maximal figure was 27%
Mechanism of action of titanium dioxide pigment in the photodegradation 359

Figure 7 Kinetics of formation of HCl expressed as conductivity units (S) during UV irra-
diation of: (x) pure PVC and PVC with :(▲) 4 parts CaCO3; (▲) 10 parts CaCO3; (●) 4 parts
and (●) 10 parts TiO2 [43].

for PVC-TiO2 under air. This severe loss has to be associated with oxidative
pathways arising from the presence of TiO2; the corresponding experiment with
‘pure’ PVC gave a much lower weight loss. The volatile products gave FTIR
spectra showing CO2 and H2O to be the main components. In parallel with this
weight loss, the GPC chromatograms showed chain scission to be very significant
(Figure 9), with the value for Mw falling to 31·5% of its original value in a 300 h
photolysis; ‘pure’ PVC film showed a drop of ca 50% in the same time. SEM sur-
face images taken after various irradiation times reveal that degradation of the
PVC matrix starts from the PVC–TiO2 interface, leading to the production of
cavities around the TiO2 particle aggregate; the cavities ultimately become inter-
connected, to a size of ~10 m. ‘Pure’ PVC films, however, gave cavities of size
only 10–20 nm; and increasing the irradiation time led only to a greater density
of holes, but not a change in their diameter. Irradiation of PVC–TiO2 films under
360 Terence J. Kemp and Robin A. McIntyre

Figure 8 Weight loss of pure PVC and PVC–TiO2 (1·5 wt.%) composite films during irradia-
tion under air or nitrogen atmosphere [44].

nitrogen gave little sign of surface degradation. The main results from XPS study
of the irradiated film surface was that the ratio C(ls) to O(ls) decreased from 17·1
to 1·76 as irradiation continued from zero to 300 h. The authors [44] are con-
vinced that the presence of TiO2 induces a variety of additional radical processes
in the photodegradation, centred around the particle itself and culminating in
severe additional weight loss, additional bond scission and much increased develop-
ment of cavities.
Several studies have focused on the weatherability of TiO2-loaded PVC.
Weathering of specimens in the semi-arid climate of Broken Hill, Australia [45]
was followed by monitoring colour, gloss, and tensile and impact strength, thus 2
years of exposure gave a dark, yellowish-red coloration while 1 year robbed all
formulations containing TiO2 of >95% of their gloss. By contrast, 2 years expo-
sure gave virtually no reduction in yield stress. As regards ultimate tensile
strength, loading with TiO2 at least at a 2% level preserved the figure at above
90% of the original after 2 years exposure. In summary, a loading of 2% is opti-
mal in preserving materials properties.
Mechanism of action of titanium dioxide pigment in the photodegradation 361

Figure 9 Gel permeation chromatograms of PVC–TiO2 (1·5 wt.%) film photodegraded under
air. Numbers in parentheses are the average MW of PVC samples (The chromatograms are
sequentially shifted upwards for clarity) [44].

2·3 Mechanism of chalking


PVC–TiO2 formulations as used in external building components (window pro-
file, fascia boards, etc.) show an initial high gloss. On weathering, the gloss turns
to a bright white matt appearance, a phenomenon known as chalking. Essentially,
oxidative degradation of the PVC resin causes the TiO2 particles to become pro-
gressively exposed, leading to the matt appearance [12]. The best remedial action
is to incorporate light stabilisers into the formulation and to use less-photochemi-
cally active grades of TiO2.

2·4 Mechanism of greying


PVC–TiO2 samples stabilised with lead(II) compounds, and placed under water
at 60°C, develop, on UV irradiation, a grey colour. The grey colour is attributed
[46,47] to a combination of Ti3+ formed on photoreduction of Ti4+, and metallic
lead, formed on photoreduction of Pb(II). The intensity of the grey colour is
related to the photoactivity of the TiO2, and indeed provides a quantitative test of
the latter [46,47]. Lemaire et al. [22] comment that greying can also be due in part
362 Terence J. Kemp and Robin A. McIntyre

to the presence of very long polyene chains produced by dehydrochlorination of


the PVC, noting that greying can be induced by artificial weathering of samples
not immersed in water; thermolysis of these at 65°C causes the colour to change
from grey to pink, via equation (26).

2·5 Mechanism of pinking


The propensity of particular formulations of PVC containing TiO2 to develop a
pink colour on weathering, known as pinking, has provoked a vigorous debate
concerning not only the mechanism but even the chromophore responsible. This
level of interest is unsurprising in view of the economic significance of these
grades of PVC–TiO2 in their utilisation in window profile, fascia boards, etc.
What is generally accepted is that pinking becomes a significant problem for pro-
file facing northwards in cool and humid climates, as in Northern Europe [22];
south-facing elevations, or exposure in hot, sunny climates, do not cause pinking.
It is also generally accepted that this dependence on elevation is due to the rela-
tively stronger UV component in northern light.
Lemaire et al. [22] attribute the pink colour to the presence of extended
polyene chains produced by dehydrochlorination of the PVC, according to the
same mechanism they advocated for yellowing both of PVC and PVC–TiO2.
They note that the majority of recorded cases of pinking deal with PVC–TiO2
samples containing, as a stabiliser, lead(II) compounds. However, they repudiate
[22] the hypothesis that lead contributes to the pink coloration, possibly as lead
orthoplumbate Pb3O4 which is brownish-pink and which contains Pb(IV), noting
that they have observed pinking with formulations based on Ca/Zn heat stabilis-
ers, but under more severe artificial weathering conditions. These authors also
describe artificial weathering test regimes to enable prediction of pinking behav-
iour. These all depend on UV irradiation at 60°C followed by extended thermal
treatment, i.e. maintenance at 65°C for 150 h. The thermolysis is considered to
convert very long polyene sequences, which import a grey colour to the sample,
to shorter sequences to be associated with the pink colour:

(26)
Mechanism of action of titanium dioxide pigment in the photodegradation 363

The Bandol conference paper of Walraevens and Georges [47], available


as a press release on the Benvic website [46] describes two key experimental
results which, while not altogether excluding the polyenic sequence explanation,
point to the presence of a second pathway. These results are as follows:

(i) removal of the pink surface layer, followed by extraction from it of organic-
soluble material by tetrahydrofuran solvent gave an insoluble pink residue.
The X-ray photoelectron spectrum of the pink residue gave broad peaks in
the region typical for lead. These could be deconvoluted to reveal compo-
nents for Pb(IV) and Pb(II). Residue from an ‘unpinked’ control sample gave
the peaks only for Pb(II).
(ii) Formulations made from what is described as ‘low-active TiO2’ gave no
pinking, while progressively more active grades gave progressively greater
levels of pinking. The categorisation of ‘activity’ of the TiO2 was based on
its propensity to experience greying in the authors’ ‘greying test.’

Thus these authors propose that the pink colour is due, at least in part, to optical
transitions within a Pb(IV) compound, possibly Pb3O4. Their correlation between
‘greying activity’ and ‘pinking’ activity does not fit well with the dehydrochlori-
nation mechanism of Lemaire et al. [22]. However, it must be noted that the latter
group have achieved pinking with lead-free samples, albeit under more vigorous
artificial weathering conditions than the lead-stabilised samples. The present
authors find the photoelectron spectral results particularly convincing, but clearly
this debate is far from over.

2·6 Effect of modification of TiO2 on photodegradation


While there is a considerable literature on attempts to modify the behaviour of
TiO2 by applying external coatings of various metal oxides or photosensitisers
[48], by platinising (inserting nanoparticles of platinum metal at the surface)[49],
or by doping the TiO2 lattice with ions of charge different from that of Ti [38, 39],
these have generally sought to increase the photoactivity, spectral range or photo-
conductivity of the TiO2. In the field of protection of polymer hosts from aggres-
sive species generated at the surface of TiO2 on photoactivation, attention has
been focused on the application of coatings of Al2O3, SiO2, ZrO2, etc. as outlined
below.
364 Terence J. Kemp and Robin A. McIntyre

2·6·1 Effect of coating


Much effort has been expended on the development of grades of TiO2 coated with
external layers of alumina, silica or zirconia or combinations of these, and these
are termed and marketed as “durable” or “superdurable”, the implication being
that such materials offer superior retention of optical, materials and surface prop-
erties when utilised as pigments in PVC. The pronounced effect of modifying
TiO2 can be seen from carbonyl development kinetics of several grades at a 4 %
loading (Figure 10) [50]. The field has been reviewed [4, 10], there is a consid-
erable commercial literature, and there exist a number of company websites
[14,15] describing the pigments available for purchase.
The aim of applying a coating of a metal oxide such as SiO2 or Al2O3 to
encapsulate a TiO2 particle is to impede movement of h+/e- species generated
within the TiO2 core to the exterior of the composite particle. There will be ad-
ditional protective effects owing to scattering of incident UV photons at the coat-
ing surface. The energetics of migration of h+ and e- generated within the TiO2 to
the outer surface of the exterior coating can be depicted, following Diebold [12]
as in Figure 11, where the situations for coatings with identical, large band gaps
but different frontier energies are presented. In Figure 11A, electrons and holes

Figure 10 Plot of extents of carbonyl development in PVC versus time of irradiation for a
loading of 4 % of different grades of TiO2 [50].
Mechanism of action of titanium dioxide pigment in the photodegradation 365

Figure 11 Location of bond energies for TiO2 coated by insulators with identical band gap but
different relative frontier energies [12].
366 Terence J. Kemp and Robin A. McIntyre

are repelled equally, although some electrons and holes will have sufficient
energy to penetrate the coatings, orbitals to reach the outer surface. In Figure 11B,
the electrons face a prohibitively high barrier, but holes can penetrate easily, to
confer comparable oxidative stress on the polymer matrix. In Figure 11C it is the
electrons which have the easy path, the holes being trapped at the TiO2-coating
interface. Since the holes are the major source of oxidative degradation, the con-
figuration in Figure 11C represents the best strategy for achieving protection of
the polymer. Diebold [12] observes that while most, if not all, of the commercial
durable grades of TiO2 rely on encapsulation, there is a fine balance to be struck
between achieving completeness of encapsulation (by utilising a thick coating)
and preservation of the optical qualities of TiO2, (by utilising the thinnest possi-
ble layer). Calculation shows [12] that at least three monolayer equivalents of
coating oxide must be deposited to achieve 95 % coverage.
Very detailed studies have been carried out of the effects of a more exten-
sive range of TiO2 pigments on retention of brightness and gloss on weathering
[51]. The overall conclusions were (i) an inorganic surface modification using
low amounts gives only minor improvements; (ii) by using various sophisticated
inorganic multicomponent surface treatments, outstanding photostabilities are
achieved, whether the TiO2 core is derived from the chloride or sulfate manu-
facturing processes; (iii) not only the gross formula of surface modification but
also the details of the modification procedure are decisive for pigment
quality.(This paper also deals with the effects of stabilisers). In a development of
this work [52], loss of gloss was correlated with surface roughness measurements
and SEM images of weathered PVC-TiO2 plates.

2·6·2 Effect of doping


The effects of doping TiO2 with various transition metal ions have been studied
by a number of groups with reference to its photoactivity, but there has been
much less work with regard to polymer degradation. One obvious reason for this
is the wish to avoid coloration of the TiO2 by the dopants, which will be unavoid-
able except at very low dopant concentrations. The siting of such ions can induce
several effects: interstitial M2+ or M3+ions can act as hole traps:

h+ + Mz+ → M(z+1)+ (27)

while interstitial M5+, M6+ ions can act as electron traps:


Mechanism of action of titanium dioxide pigment in the photodegradation 367

e– + Mz+ → M(z–1)+ (28)

On balance, eqn (27) is more likely to lead to reduced photoactivity, and hence
protection of the polymer than eqn (28), although the latter will remain signifi-
cant in that e-/h+ recombination will be reduced.
A second effect is on the band-gap; transition metal ions absorb in the
visible region and, for example, Co2+-doped TiO2 is activated towards the photo-
degradation by visible light of gaseous acetaldehyde [53]. A particularly detailed
and wide-ranging study is due to Ikeda et al. [54] who doped TiO2 with V, Cr, Fe,
Co, Co, Cu, Mo and W; in general the photoactivities of these materials were less
than those of the undoped TiO2, and were reduced further as the doping level
increased. Femtosecond pump-probe transient absorption spectroscopy [54]
showed a transient peak at 620 nm, in agreement with Colombo and Bowman
[55], attributed to trapped electrons generated by photo-excitation. The decay
constant of the transient (kR) increased on doping with all of the transition-metal
ion investigated to an extent dependant on dopant concentration. In general, the
photochemical activity decreased with increase in kR. A similar conclusion was
reached for Mo- and V-doped TiO2 from surface photovoltage spectra (SPV) [56];
the SPV values fall from 6400 V for TiO2 to only 1·3 and 3·9 for Mo- and
V-doped samples respectively.
Partial reduction of TiO2 to give a number of Ti3+ centres has an enhanc-
ing effect on the photoactivity of the semiconductor attributed to the difference
between the Fermi level and the redox potential of an O2/H2O2 containing environ-
ment, leading to electron depletion from the surface and reduced surface h+/e–
recombination [57]. Conversely, oxidation of the semiconductor lowers the Fermi
level of the particles, increasing the surface recombination rate and hence reduc-
ing the quantum yields. Milling or grinding the TiO2 introduces bulk defects
where electron and holes recombine [57].
One of the few papers relating to the effect on polymer degradation of
metal-ion doping of TiO2 refers to the non-transition metal ion Al3+; here the
effects on artificial weathering of alkyd paint films were explored, based on loss
of gloss and development of chalking [58]. The Al2O3 dopant partly dissolved in
the TiO2 host while part was segregated to the surface; the latter could be
removed selectively by etching. It was found that Al3+ in the bulk was more
important than at the surface, operating by providing recombination sites. Of the
two modes of incorporating Al2O3 into rutile, only the one providing oxygen
368 Terence J. Kemp and Robin A. McIntyre

vacancies is effective, presumably by trapping an electron which then acts as a


centre for h+.

3. POLYOLEFIN – TIO2 SYSTEMS


An overview of the ageing and stabilisation of filled polyolefins has been given by
Allen and colleagues [59] following an earlier review by Allen and McKellar [60].

3·1 Polyethylene
Initial studies [61] concentrated on changes in mechanical and optical properties
of PE-TiO2 films on UV irradiation. Irradiation-induced decreases in weight and
tensile strength and the development of an enhanced white coloration were
observed, which TEM revealed to be due to the development of voids around the
TiO2 particles.
Allen et al. [62,63] reported a significant irradiation wavelength effect in
the photoageing of TiO2-pigmented low density polyethylene (LDPE) films as
determined by development of the carbonyl index; with long-wavelength
(365 nm) excitation all types of TiO2 pigment, except the superdurable, heavily
coated rutile grade, behaved as sensitisers whereas with a narrow band 254 nm
source, the pigments behaved as UV screeners and protected the LDPE. Under
high intensity polychromatic irradiation all the pigments behaved as photosensi-
tisers. Increasing the temperature of the photoirradiation from 50ºC to 90ºC
increased the rate of oxidation by an order of magnitude. The principal degradation
products were carboxylic acids (1710 cm-1), esters (1740 cm-1) and peroxyesters
(1770 cm–1). The effectiveness of the various grades of TiO2 followed the
expected sequence, i.e.

Anatase ~ Rutile > Rutile > Rutile > Rutile


Lightly coated Medium coat Heavy coat

Typical differential absorbance versus time plots are shown in Figure 12.
In an earlier paper [64], the effect was studied of adding rutile to com-
mercial PE containing the anti-oxidant 4,4′- thio-bis(tert-butyl-meta-cresol),
known as Santanox R. The TiO2-free system shows pronounced photo-yellowing
while addition of coated rutile markedly inhibited this discoloration. As regards
photo-oxidation as measured by the carbonyl index, with uncoated rutile the
photo-stabilisers Irgastab 2002 and UV 531 were more effective than Tinuvin
Mechanism of action of titanium dioxide pigment in the photodegradation 369

Figure 12 Differential absorbance at 1710 cm-1 versus irradiation time (3:1) (h) for LLDPE
films (100 m thick) at 50C containing (X) no TiO2 and 0·5 % w/w of pigment (▼) A, (▲) B,
(■) C, (■) D, (●) E [62].

770, while in the presence of coated rutile, the reverse was the case. The authors
suggest that the stabiliser can be adsorbed onto the pigment particle, particularly
so with the uncoated pigment. The significance of stabiliser-pigment interaction
is also noted in an earlier paper [65].

3·2 Polypropylene
Allen et al. [66] have extended their infrared kinetic spectroscopic studies of
synergistic/antagonistic effects of TiO2 pigments and a range of antioxidants and
light stabilisers from polyethylene (Section 3·1) to polypropylene.

4. OTHER POLYMER – TiO2 SYSTEMS


4·1 Polycarbonate
While no photodegradation studies on polycarbonate (PC) – TiO2 systems have
been carried out, a detailed study of PC–rutile showed that incorporation of only
370 Terence J. Kemp and Robin A. McIntyre

2% of pigment into undried material gave a reduction in ultimate tensile strength


from ca. 64 kN mm-2 to ca 50 kN mm-2. Together with thermally stimulated dis-
charge current measurement and SEM data, this was attributed to a densification
of PC around the TiO2 particles, reducing polymer dynamics [67].

4·2 Acrylates and paints


TiO2 finds one of its major applications as a pigment for paint and the light-
stability of such bulk products has attracted much attention [68, 69]. The signifi-
cance of the level of dispersion of the pigment has been noted [69], as has the need
to achieve a balance between direct and photocatalytic degradation of the binder
[69]. The photodegradation of a commercial acrylic emulsion based on methyl
acrylate/methyl methacrylate/ butyl acrylate, doped with either rutile or anatase,
was monitored [70] by FTIR measurements of the CO2 generated [71]. The rate of
CO2 evolution depends critically on the nature of the pigment, thus anatase has a
very large photocatalytic effect while rutile offers some protection, Figure 13.
A particular interest was the relationship established between UV light
intensity (I) and the level of CO2 evolution after a fixed time (Figure 14), i.e. a
half-power dependence.

Figure 13 Carbon dioxide evolution from unpigmented (U, +), anatase-pigmented (A, ▲), and
rutile-pigmented (R, ●) acrylic films [70].
Mechanism of action of titanium dioxide pigment in the photodegradation 371

Figure 14 Carbon dioxide evolution after 210 minutes as a function of variation in incident UV
intensity. The UV intensity was varied by using neutral density filters, either singly or in com-
bination [70].

This kinetic feature is well-known from photomineralisation studies


involving TiO2 [1].

4·3 Polystyrene
Polystyrene has been manufactured on an industrial scale for about 100 years and
various fillers has been utilised including glass fibres, glass beads and Al2O3 [72],
but to our knowledge, no systematic study has been published of the degradative
effects of TiO2 as a filler or colorant of polystyrene. Our preliminary study of
TiO2-pigmented polystyrene films (Figure 15) indicates that all grades [50] of
TiO2 at 4 % loading reduce carbonyl formation, with Oxonica grade BFB exer-
cising the strongest protective effect.

5. CONCLUSIONS
The interactions of photo-excited TiO2 particles with a polymer host are complex.
372 Terence J. Kemp and Robin A. McIntyre

Figure 15 Plot of extents of carbonyl development in polystyrene versus time for a loading of
4 % of different grades of TiO2 [50].

The intrinsic photochemistry of the polymer, principally dehydrochlorination and


surface oxidation, is partly screened by the TiO2 to become diminished. However,
the TiO2 undergoes UV excitation to develop hole-electron pairs (h+/e-) which
either recombine, become trapped at a defect site, or migrate to the surface to
form reactive species such as •OH, O2•– and polymer–derived radicals, all of
which enhance degradation. The principal means of modifying TiO2 to obviate its
photodegradative action is to coat it with an even but thin coating of SiO2, Al2O3
or similar oxide to reduce migration of the aggressive h+ and e- to the exterior
surface of the coated particle. In TiO2–PVC systems containing stabilisers based
on metal salts such as lead(II), additional pathways emerge giving unacceptable
levels of coloration. The mechanism here is controversial but there is evidence
that with lead-based stabilisers, a Pb(IV) species is at least one of the colorants.

6. ACKNOWLEDGEMENTS
We thank Oxonica Ltd. for financial support of R.A.M. through a Faraday
Partnership.
Mechanism of action of titanium dioxide pigment in the photodegradation 373

7. REFERENCES
[1] A. Mills and S. Le Hunte, J. Photochem. Photobiol. A: Chem., 1997, 108, 1.
[2] A. L. Linsebigler, G. Lu and J. T. Yates, Jr. Chem. Rev., 1995, 95, 735.
[3] M. Anpo, Res. Chem. Intermediates, 1989, 11, 67.
[4] H. J. Braun, J. Coatings Technol., 1997, 69, 59.
[5] R. E. Day, Polym. Degrad. Stab., 1990, 29, 73.
[6] M. Bryk, Degradation of Filled Polymers, Ellis Horwood, Chichester, 1991.
[7] Yu. A. Gorbatkina, “Adhesive Strength of Fibre-Polymer Systems“, Ellis Horwood,
Chichester, 1992.
[8] N.S. Allen, Polym. Degrad. Stab., 1994, 44, 357.
[9] N.S. Allen, Polymer Photochemistry in Photochemistry, Vol. 31, Specialist Periodical
Report, Royal Society of Chemistry, London, 2000, pp. 335-391 (847 references) and
previous volumes in this series.
[10] H. J. Braun, A. Baidins and R. E. Marganski, Prog. Org. Coatings., 1992, 20, 105.
[11] http://www. dupont.com/tipure/profiles/H-88382.pdf
[12] M. Diebold, http://tipure-europe.asp.dupont.com/engnical/photodegradation/introduction.
html
[13] A. Piscopo, D. Robert and J. V. Weber, J. Photochem. Photobiol. A: Chem., 2001, 139, 253.
[14] http://www.nl-ind.com/kronos/
[15] http://www.huntsman.com/tioxide/
[16] A. L. Andrady and N. D. Searle, J. Appl. Polym. Sci., 1989, 37, 2789.
[17] G. Scott, Polym. Degrad. Stab., 1985, 10, 97.
[18] J-L. Gardette and J. Lemaire, Polym. Degrad. Stab., 1991, 33, 77.
[19] J-L. Gardette, S. Gaumet and J. Lemaire, Macromolecules, 1989, 22, 2576.
[20] J-L. Gardette and J. Lemaire, Polym. Degrad. Stab., 1991, 34, 135.
[21] R. S. Davidson and R. R. Meek, Polym. Photochem., 1982, 2, 1.
[22] J. Lemaire, N. Siampiringue, R. Chaigneau, P. Delprat, G. Parmeland, P. Dabin and
C. Spriet, J. Vinyl. Additive Technol., 2000, 6, 69.
[23] J-L. Gardette, S. Gaumet and J. Phillipart, J. Appl. Polym. Sci., 1993, 48, 1885.
[24] C. Anton-Prinet, G. Muir, M. Gay, L. Audouin and J. Verdu, J. Materials. Sci., 1999,
34, 379.
[25] G. Scott, Polym. Degrad. Stab., 1985, 10, 97
[26] G. Scott, Brit. Polym. J., 1984, 16, 271.
[27] J. R. White and A. Turnbull, J. Mater. Sci., 1994, 29, 584.
[28] J. F. Rabek, Polymer Photodegradation, Chapman and Hall, London, 1995.
[29] N. S. Allen, Engineering Plastics, 1995, 8, 247.
[30] T. J. Turton and J. R. White, Polym. Degrad. Stab., 2001, 74, 559.
[31] J. Lemaire , J-L. Gardette, B. Mailhot and X. Jouan in Current Trends in Polymer
Photochemistry ed. N. S. Allen, M. Edge, I. R. Bellobon and E. Selli, Ellis Horwood,
New York, 1995, chap. 12.
[32] G. Kaempf, W. Papenroth and R. Holm,J. Paint Technol., 1974, 46, 56.
[33] F. B. Stieg, Jr., J. Paint. Technol., 1971, 43, 36.
[34] J. Czarnecki and T. Dabros, J. Colloid Interfac. Sci., 1980, 78, 25.
[35] A. J. Werner, US Patent, 3 437 502, 1969.
[36] P. B. Howard, US Patent, 4 652 223, 1977.
374 Terence J. Kemp and Robin A. McIntyre

[37] M. Matsunga, T. Ubani, H. Okuda and H. Futamoto, US Patent, 4 405 376, 1983.
[38] M. A. Fox and M. T. Dulay, Chem. Rev., 1993, 93, 341.
[39] J. Kiwi and M. Grätzel, J. Phys Chem., 1986, 90, 637.
[40] J-L Gardette and J. Lemaire, J. Vinyl Additive Technol., 1997, 3, 107.
[41] C. Anton-Prinet, G. Mur, M. Gay, L. Audouin and J. Verdu, Polym. Degrad. Stab.,
1998, 61, 211.
[42] C. Anton-Prinet, G. Mur, M. Gay, L. Audouin and J. Verdu, Polym. Degrad. Stab.,
1998, 60, 265.
[43] T. A. Skowronski, J. F. Rabek and B. Ranby, Polym. Degrad. Stab., 1984, 8, 37.
[44] S. Cho and W. Choi, J. Photochem. Photobiol. A: Chem., 2001, 143, 221.
[45] L. S. Burn, Polym. Degrad. Stab., 1992, 36, 155.
[46] http://www.benvic.com/pages/pr2001_1.htm
[47] W. Walraevens and I. Georges, Bandol Conference, France, Sept. 2000, 12pp.
[48] Md. K. Nazeeruddin, R. Humphry-Baker, M. Grätzel and B. A. Murrer, J. Chem. Soc.,
Chem. Commun., 1998, 719.
[49] A. N. Shipway and I. Willner, J. Chem. Soc., Chem. Commun., 2001, 2035.
[50] T. J. Kemp and R. A. McIntyre, unpublished work.
[51] U. Gesenhues and J. Hocken, J. Vinyl Additive Technol., 2000, 6, 80.
[52] U. Gesenhues, Polym. Degrad. Stab., 2000, 68, 185.
[53] M. Iwasaki, M. Hara, H. Kawada, H. Tada and S. Ito, J. Colloid Interface Sci., 2000,
224, 202.
[54] S. Ikeda, N. Sugiyama, B. Pal, G. Marci, L. Palmisano, H. Noguchi, K. Uosaki and
B. Ohtani, PCCP, 2001, 3, 267.
[55] D. P. Colombo, Jr. and R. M. Bowman, J. Phys Chem., 1996, 100, 18445.
[56] Z. Luo and Q-H. Gao, J. Photochem. Photobiol A: Chem., 1992, 63, 367.
[57] A. Heller, Y. Degani, D.W. Johnson, Jr. and P.K. Gallacher, J. Phys Chem., 1987, 91, 5987.
[58] U. Gesenhues, J. Photochem. Photobiol. A:Chem., 2001, 139, 243.
[59] N. S. Allen, M. Edge, T. Corralis, A. Childs, C. M. Liauw, F. Catalina, C. Peinado,
A. Minihan and D. Aldcroft, Polym. Degrad. Stab., 1998, 61, 183.
[60] N. S. Allen and J. F. McKellar, Chem. Soc. Rev., 1975, 4, 533.
[61] B. Ohtani, S. Adzuma, S-I. Nishimoto and T. Kagiya, Polym. Degrad. Stab., 1992, 35, 53.
[62] N. S. Allen and H. Katami, Polym. Degrad. Stab., 1996, 52, 311.
[63] N. S. Allen, H. Katami and F. Thompson, Eur. Polym. J., 1992, 28, 817.
[64] N. S. Allen, D. J. Bullen and J. F. McKellar, J. Materials Sci., 1978, 13, 2692.
[65] N. S. Allen, D. J. Bullen and J. F. McKellar, J. Materials Sci., 1977, 12, 1320.
[66] N. S. Allen, M. Edge, T. Corrales and F. Catalina, Polym. Degrad. Stab., 1998, 61, 139.
[67] K. M. Blackwood, R. A. Pethrick, F. I. Simpson, R. E. Day and C. L. Watson, J. Mater.
Sci., 1995, 30, 4435.
[68] T. A. Egerton, Titanium Compounds, in Vol. 24 of Kirk-Othmer Encyclopaedia of
Chemical Technology, Wiley, New York, 1997, p.225.
[69] M. P. Diebold, www2.coatings.de/particles/ecspapers/diebold/diebold.htm
[70] P.A. Christensen, A. Dilks, T. A. Egerton and J. Temperley, J. Mater. Sci., 2000, 35, 5353.
[71] P.A. Christensen, A. Dilks, T. A. Egerton and J. Temperley, J. Mater. Sci., 1999, 34, 5689.
[72] P. Švec, L. Rosík, Z. Horák and F. Vecerka, Styrene-based plastics and their modification,
Ellis Horwood, New York, 1990.

You might also like