Sie sind auf Seite 1von 75

Prog. Polym. Sci.

25 (2000) 12611335

Photostabilization of coatings. Mechanisms and performance


J. Pospsil*, S. Nespurek
Institute of Macromolecular Chemistry, Academy of Sciences of the Czech Republic, Heyrovsky Sq. 2,
162 06 Prague 6, Czech Republic

Abstract
The lifetime of organic coatings is reduced in outdoor applications by attacks of solar radiation, oxygen and
atmospheric pollutants. Undesirable mechanical, physical and chemical consequences of the resulting degradation
can be substantially restricted by properly selected photostabilizers. Structures of absorbers of UV radiation
(UVA) and photoantioxidants (hindered amine stabilizers, HAS), mechanisms of their activity, processes respon-
sible for loss of their durability, interferences with other coating components (catalysts, pigments, extenders) and
application in various coatings are outlined. 2000 Elsevier Science Ltd. All rights reserved.

Keywords: Coatings; Photostabilization; UV absorber; Photoantioxidant

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1262
2. Durability of coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1263
2.1. Physical, chemical and biological deteriogens affecting long-term durability of coatings . . . . . .1264
2.1.1. Solar tropospheric radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1264
2.1.2. Atmospheric pollutants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1265
2.1.3. Microbial deteriogens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1268
2.2. Degradation phenomena in coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1269
2.2.1. Coating photochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1270
2.2.2. Coating hydrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1277
2.2.3. Assessment of weathering of coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1277
3. Photostabilization of coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1278

* Corresponding author. Tel.: 420-2-20403-292; fax: 420-2-367981.


E-mail address: posp@imc.cas.cz (J. PospIsil).
Abbreviations: A: absorbance; BP: benzophenone; BT: benzotriazole; Ch: chromophore, e : molar absorption; EPDM: ethy-
lenepropylenediene terpolymer; ESIPT: excited state intramolecular proton transfer; EPR: electron paramagnetic resonance;
HAS: hindered amine stabilizer; HMW: high-molecular weight; HS: high-solid; IC: internal conversion; IMBH: intramolecular
hydrogen bond; IR: infrared; ISC: intersystem crossing; LMW: low-molecular-weight; NMR: nuclear magnetic resonance; OA:
oxanilide; PI: photoinitiator; Q: quencher; TA: 1,3,5-triazine; TSA: thermosetting acrylics; UVA: UV absorber; YI: yellowness
index

0079-6700/00/$ - see front matter 2000 Elsevier Science Ltd. All rights reserved.
PII: S0 0 7 9 - 6 7 0 0 ( 0 0 ) 0 0 02 9 - 0
1262 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

3.1. Light screening pigments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1279


3.2. UV absorbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1279
3.2.1. Activity mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1281
3.2.2. UVA durability loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1290
3.3. Photoantioxidants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1301
3.3.1. Activity mechanism of HAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1303
3.3.2. Adverse effects and depletion of HAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1307
3.4. Bifunctional photostabilizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1309
3.5. Physically persistent photostabilizers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1311
4. Performance of photostabilizers in coatings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1316
4.1. Physical factors in coating stabilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1327
4.2. Influence of pigments and fillers (extenders) on performance of photostabilizers . . . . . . . . . . . .1329
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1331
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1331
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1331

1. Introduction

Organic polymeric materials such as plastics, fibers, elastomers, coatings and polyblends are vulner-
able to harmful physical and chemical effects of the environment. The relevant processes are classified as
long-term heat aging and weathering. They are initiated by temperature, mechanical forces, chemical
catalysis and solar radiation triggered processes.
Safe long-term application of all polymers that have become commercial products is dependent on the
inherent resistance of the particular polymer system, including stabilizing additives. This permits the
polymers to withstand chemical and physical stresses.
This paper deals exclusively with organic coatings, a rather specific organic material from the point of
view of outdoor durability. The term organic coating used throughout the text encompasses conventional
paints, varnishes, enamels, lacquers, water emulsions, organosols and powder coatings [1,2]. The market
usually classifies coatings as trade sales coatings, industrial coatings, marine coatings and automotive
finishes, the last category having the most critical performance requirements [2].
Organic coatings are used in thin layers to protect various substrates (metal, wood, plastics) against
corrosion during weathering and mechanical damages and for decorative purposes [13]. Coatings are
considered to last for a long time in service in hostile environment in most different applications. A good
adhesion to the substrate is mandatory.
Coatings are rather complicated multicomponent systems formed by a combination of organic cross-
linkable film forming substances (binders) with relatively high concentrations of pigments [1,4]. Most
paints used today still contain organic solvents. Water-borne and solvent-free paints (such as powder
coatings and UV-radiation curable paints) are already on the market [1]. Radiation curable coatings
profit from a rapid cure of solvent-free systems [5]. Their commercial application involves, however,
some specific problems encountering selection of proper photoinitiators (PI) compatible with photo-
stabilizers.
A variety of the chemical composition of the binder (e.g. oleoresinous binders, alkyds, polyesters,
acrylates, urethane systems, silicones, epoxies, amino resins) prepared by different technologies [1],
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1263

Fig. 1. Scheme of automotive paint structure. A monocoat, B clearcoatbasecoat system.

various regular additives and processing aids such as pigments or photostabilizers [5], or rather specific
additives, such as biocides [6], corrosion inhibitors or antistatic additives [7] and flame retardants [8]
increase diversity of composition of the coating systems together with various catalysts or initiators used
for curing of various monomers or prepolymers and crosslinkers, and are characteristic of very different
responses of particular coating systems to environmental attacks. Differences in the inherent resistance
of each of the coating components to degradation contribute to the integral effect. Consequently, coat-
ings are by no means simple systems from the point of view of aging and stabilization.
Various modified acrylate-based resins used today in automotive coatings are of specific interest. A
cured pigmented binder layered over the pigmented primer and electrocoat layer forms the traditional
thermoset monocoat paint systems. Present clearcoatbasecoat paint systems consist [9] of a pigment-
free, clear polymer layer, a pigmented basecoat, primer and (over a metal substrate) an electrocoat layer
(Fig. 1). The latter is not used over plastic substrates. The layer thickness shown in Fig. 1 is not intended
to be representative of actual dimensions. Typical dimensions of individual layers (in mm) are: clearcoat
4050, basecoat 3050, primer 2030, electrocoat 2030, phosphate layer 2.
The development of high-performance coatings requires a complex of analytical data and testing
methods allowing a better understanding of relations between chemical properties of binder components,
coating network structure, additives behavior and distribution, product performance and long-term
durability.

2. Durability of coatings

The primary and economically important purpose of application of paints is to protect substrates from
depletion in harsh environment. The decorative value of the coating cannot be underrated. To fulfil the
two functions, the coatings themselves must remain intact for long periods of time under attacks of
hostile weathering. This implicates high inherent resistance of coatings against weathering.
Polymers are very different from each other as far as their inherent sensitivity to oxidation and
1264 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

photodegradation is concerned. The stability differences are due to variety of chemical structures of the
binder and crosslinker, manufacture processes, polymer morphology and additives or impurities
embedded in the polymer matrix [9]. Polymers containing structure-born chromophores (Ch), photo-
active pigments or adventitious impurities are particularly sensitive to photoprocesses [10]. The coating
degradation is a result of a concerted attack of various deteriogens and accounts for the chemical,
physical and mechanical modification of the coating. Some changes in physical properties of UV
exposed polymers, such as discoloration, have been directly related to chemical changes [11].

2.1. Physical, chemical and biological deteriogens affecting long-term durability of coatings

During outdoor service, coatings are exposed to environmental deteriogens affecting negatively their
physical or mechanical properties and chemical composition by the formation of new functional
moieties or fragmentation of the crosslinked macromolecule. Physical and mechanical changes are
consequences of minor chemical changes. Solar radiation and atmospheric pollutants are principal
deteriogens responsible for the reduced durability of coatings. In some coatings, microorganisms
have an adverse effect. Various deteriogens act in concerted processes and have mostly a complementary
or even a synergistic effect.

2.1.1. Solar tropospheric radiation


Solar radiation reaching Earths surface is characterized by wavelengths from approximately 295 up
to 2500 nm. The terrestrial radiation is classified as UV-B (285315 nm, having energy of 426
380 kJ mol 1; fortunately, the higher-energetic part of UV-B having wavelengths 280295 nm is
filtered off by the stratosphere and does not reach the Earths surface, UV-A (315400 nm, having
energy between 380 and 300 kJ mol 1 and being less dangerous for organic materials than UV-B, visible
(400760 nm) and infrared (7602500 nm) [12].
The actinic part of the tropospheric radiation (the energetic UV radiation) constitutes only about 1
5% of the total radiation. It, however, represents the most dangerous radiation component responsible for
initiation of most effects in materials under outdoor weathering by direct photolysis of covalent bonds
Visible light (3953% of solar radiation) photosensitize chromophores and accelerates polymer degra-
dation in this way. Infrared light (4260% of the total radiation) triggers thermodegradation of poly-
mers, dark pigmentation in particular [13].
The intensity of the UV radiation is dependent on the season, day time and location of the sun exposed
Earths surface. Currently, a partial depletion of the stratospheric ozone layer filtering the shorter
wavelengths of the extra-terrestrial UV-B radiation is monitored. Satellite and ground-based data indi-
cate a long-term decrease in global ozone levels by 2:7 ^ 1:4% per decade. Recently, the rate may have
increased to about 4% per decade, outside the tropics. This is expected to account for a higher level of the
UV-B radiation reaching Earths surface and, logically, for increased danger of photodegradation [14].
An increase in UV-B radiation of 1.9% per year since 1989 was reported for the summer time in Toronto,
in a direct correlation with the ozone loss. The increased UV radiation is manifested by a broad band
centered about 310 nm (changes in the ozone level affect mostly the terrestrial UV-B because ozone
itself does not essentially absorb radiation having wavelengths higher than 320 nm. The tail of the
absorbance is responsible for the effective 295 nm cutoff of the terrestrial UV). Therefore, only a
negligible absolute radiation increase in wavelengths shorter than 295 nm is expected providing a proper
control of all anthropogenic emissions catalyzing ozone depletion [14]. Anyway, even the present
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1265

Fig. 2. Relative spectral energy distribution in: (1) sunlight (Miami Average Optimum, as measure at 45S on 3-20-84); (2)
Sunshine Carbon Arc, Corex-D filtered, 3 6500W Xenon Arc lamp with borosilicate inner and outer filters, 4 FS-40 Fluorescent
Sun lam used in QUV tester as per ASTM G53. (Reproduced, with permission from Ciba Specialty Chemicals, Inc., Basel.)

intensity of the actinic solar UV radiation reaching the troposphere is sufficient to trigger serious
photochemical degradation of coatings during long-term outdoor exposure.
The intensity and the nature of the UV light are critical variables, which must be considered in
checking durability of modern long-lasting coatings. Even small local changes in UV levels due to
the ozone depletion may affect the lifetime prediction of coatings calculated from contemporary accel-
erated experiments. Moreover, we should consider, that besides chromophores present in polymers, also
some constituents of the polluted atmosphere are species absorbing UV light and involved in light
triggered processes (see Section 2.2.2) These include ozone, nitrogen dioxide, sulfur dioxide, nitric
acid, alkyl nitrates, acyl nitrites, low-molecular-weight (LMW) hydrocarbons, aldehydes and ketones
[15]. A complex of photoactive air pollutants is involved in complicated weathering mechanisms in a
manner limiting reproducibility of weathering trials due to unpredictable changes in atmospheric factors
and their interferences with solar radiation. This also explains difficulties in correlation of results
obtained in different localities and accelerated trials.
This paper does not deal with weathering tests of coatings. However, it should be noted that accel-
eration of degradation processes using artificial radiation sources with increased intensity in UV-B
deviates from outdoor tests by changing relative importance of individual steps of the degradation
mechanism and relations to mechanistic pathways of photostabilizers. Moreover, activation spectra
indicate considerable differences in spectral sensitivity between various polymers [16]. Only a reason-
able acceleration provides a base for acceptable correlation between outdoor and laboratory weathering
tests. Emission spectra of light sources used in various accelerated weathering devices are shown in Fig.
2 and compared with the Miami Average Option sunlight.

2.1.2. Atmospheric pollutants


Pollutants play an increasing role in environmental degradation of organic polymers. Oxidizing and
1266 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

acid pollutants arising from both the anthropogenic activities and natural sources contaminate the
Earths troposphere. From the point of view of adverse effects on synthetic and biological polymers,
photooxidants and precursors of acids, such as carbon dioxide, sulfur oxides and nitrogen oxides, should
be mentioned [12]. Most of them, together with various volatile aliphatic organic compounds and their
oxidation products are involved in reactions characteristic of photochemical smog. Some pollutants
absorb actinic solar radiation and react in photochemically excited states. This is characteristic of some
polynuclear aromatic hydrocarbons emitted into atmosphere by fuel combustion. They are potential
sources of singlet oxygen and sensitizers of oxidation of hydrocarbon polymers. The actual material
problems such as color, transparency, surface properties occur as a result of a combined attack of several
atmospheric deteriogens [17].

2.1.2.1. Atmospheric photo-oxidants. Besides oxygen, the regular atmospheric oxidant, strong oxidizing
species are formed in troposphere and participate in polymer degradation. They consist of relatively
stable (ozone, peroxyacetyl nitrate, nitrogen oxides) and unstable oxidants (atomic and singlet oxygen,
hydroxy radicals).
Tropospheric photoreactions involve nitrogen dioxide and nitric oxide arising from natural sources
(tropic forests, agriculture, lightning) or various anthropogenic industrial and domestic activities (auto-
motive exhausts, power plants, combustion of biomass or fossil fuels). Nitrogen dioxide photolyses to
ozone and atomic oxygen, Eqs. (1) and (2) [18]. Photoreactions Eqs. (3)(6) are characteristic of ozone
photolysis and HO and HOO radical formation.
hn430 nm
NO2 ! NO O 1

O O2 ! O3 2

O O3 ! 2O2 3

O H2 O ! 2HO 4

hn340 nm
O3 ! O O2 5

O3 HO ! O2 HOO 6
Excited nitrogen dioxide NO2 is formed according to Eq. (7) and accounts for nitrogen trioxide and
singlet oxygen 1O2 Eqs. (8)(10). The latter is, together with ozone, a specific oxidant for unsaturated
organic polymers [12].

NO O3 ! NO2 O2
hn
7

NO2 O3 ! NO3 O2 8

NO2 O2 ! NO2 1 O2 9

NO2 O ! NO 1 O2 10
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1267

A typical normal global ground-level concentration of ozone is very low (13 ppm) but sufficient to
initiate oxidation of saturated polymers. However, ozone concentration in polluted atmosphere reaches
up to the 2050 ppm range [19,20]. Therefore, potentially very harmful levels of ozone and other related
atmospheric oxidants can now be seen in urban and rural areas.
Air photopollutants trigger degradation of synthetic polymers and biopolymers (PH) during long-term
heat aging and weathering [20]. Eqs. (11)(15) show schematically processes due to ozone-triggered
initiation of degradation. Consequently, ozone is a strong photo-oxidant in organic polymers Eq. (16)
and is a precursor of hydroxy radicals HO via Eqs. (4), (6) and (13).
PH O3 ! POOOH ! POO HO or PO HOO 11

POO O3 ! PO 2O2 12

PH O ! P HO 13

P O2 ! POO 14

P O3 ! PO O2 15

PH HO ! P H2 O 16

PH NO2 ! P HNO2 17

P NO2 ! PO NO 18
Processes in Eqs. (17) and (18) are characteristic of nitrogen dioxide-initiated polymer oxidation [20].
Peroxyacetyl nitrate arises [21] in reaction of alkylperoxyls generated in the troposphere by oxidation of
LMW hydrocarbons with nitrogen dioxide Eq. (19)
2ROO NO2 ! R 0 COOONO2 RO 19
It is evident that all oxidizing photopollutants forming components of the troposphere are harmful for
polymers in long-term contact. Moreover, they participate in a specific way in oxidative deterioration of
stabilizers. For example, nitrogen oxides oxidize phenolic moieties in antioxidants and phenolic UVA
[22] and ozone depletes HAS [23].

2.1.2.2. Acid atmospheric pollutants. Mineral acids arising from oxides of sulfur and nitrogen are
responsible for heavy damages of organic matter in environment, including organic polymers. Sulfur
dioxide originating from combustion of fossil fuels or volcanic eruptions is photo-oxidized in atmo-
sphere to sulfur trioxide [12] and hydrated with water vapor to sulfuric acid (a part of the acid rain, dew
or fog) forming acid deposits [17,24] etching coating surfaces, and participating under solar radiation in
accelerated degradation of coatings [25] and surface extraction of basic stabilizers (HAS) after forma-
tion of water-soluble salts. Moreover, excited sulfur dioxide photo-sulfonates and oxidizes carbon-chain
polymers and photosensitizes formation of singlet oxygen in the troposphere [15] Eqs. (20) and (21)

SO2 ! SO2
hn
20
1268 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

SO2 O2 ! SO2 1 O2 21
A combination of sulfur dioxide, humidity and solar radiation deteriorates pigments and coatings and
accounts for their rapid fading [15,26]. Atmospheric SO2 is oxidized with alkylperoxyls Eq. (22) and is a
potential decomposer of polymeric hydroperoxides Eq. (23) resulting in free-radical species. The latter
may participate in accelerated photo-oxidation [12].
POO SO2 ! POO SO2 ! PO SO3 22

POOH 2SO2 ! POSO2 HOSO2 23


As far as organic coatings are concerned, SO2 is one of the most aggressive deteriorating factors of the
environment. In the form of acid deposits, it stimulates atmospheric corrosion of metallic substrates,
deteriorates binders by enhancing their permeability and forming durable spots, and pigments by fading
mostly in photoassisted processes [2426]. Permeation of the acrylatemelamine basecoatclearcoat
film by acid rain water results in depletion of the melamine crosslinks by transformation of melamine
segments into water-extractable cyanuric acid. Swelling of the coating, formation of bubbles, blisters,
crazes, cracking or micropeeling are the consequences [24] together with extraction of the basic HAS.
The binder degradation is dependent on the concentration of sulfuric acid. The levels of 5070% of
H2SO4 are reached when water from the acid rain drops is evaporated on heated surfaces [25]. Synergism
was observed in photodegradation of coatings exposed to UV radiation and acid precipitation. Mela-
mine-free coatings, such as epoxycrosslinked coatings, have much higher resistance to acid rain etching
[26]. As a consequence, HAS embedded in these coatings is more resistant to acid depletion.
Nitric acid is formed in the atmosphere by oxidation and hydration of NOx, returns to the Earth in the
form of acid rain and forms oxidizing and etching deposits [24]. The acid rain may contain small
amounts of hydrochloric acid, arising from burning of coal.

2.1.3. Microbial deteriogens


Synthetic polymers, crosslinked in particular, are relatively inert to microbial degradation [27].
Microorganisms are not able to assimilate polymers directly. Crosslinking increases the packing density
of the polymer while inhibiting penetration of enzymes having molecular weights in the range 10 000
1 000 000. Moreover, in most cases, microorganisms do not degrade synthetic polymers having mole-
cular weights higher than 20 000. The susceptibility of polymers to bacteria, mildew and fungi is
effectively promoted by photo-oxidative degradation of the polymer matrix resulting in LMW fragments
with oxygenated structures (ketones, carboxylic acids, lactones, alcohols) increasing hydrophilicity of
the polymer substrate and, consequently, physical binding of water needed for biodegradation reactions.
The oxygenated fragments are consumed preferentially on the polymer surface [27]. Moreover, sunlight-
induced physical alterations of coating surfaces, such as erosion, roughness, fissures and microfractures
(microcracks) are formed. Microorganisms can be fixed mechanically in the surface irregularities
together with dirt, exogenous carbon pollutants which are serving as foothold and nutrient sources
[7]. It was reported that reduction of polymer surface alteration using a proper photostabilizer, HAS
in particular, reduces dramatically the accumulated biological growth. UVA can enhance the light-
fastness of biostabilizers added to coatings [7].
The self-supporting character of photo- and biodegradation was confirmed experimentally. The
biodeterioration is manifested by discoloration and formation of blotches by indelible dyes (pink,
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1269

yellow, gray, black) which are metabolic products of microorganisms, odor development due to meta-
bolites of mold and mildew, enhanced surface dirt uptake due to surface stickiness, permeability to
solvents, material brittleness and poor appearance [28].
A large number of microorganisms, their metabolic versatility and almost unlimited variety underline
the danger arising from microorganisms [27]. Paints based on oil and oleoresinous materials are the most
vulnerable. Polymer biodeterioration depends on environmental conditions (humidity, temperature,
types of microorganisms, pH and chemistry and morphology of the material [28].
Surface growth of microorganisms should be considered as the principal biodegradation mode of
coatings and paints. Biological defacement of coatings by mildew is a common problem especially in
trade sales coatings [7]. Binders with hydrolysing moieties, such as aliphatic polyesters, polyester-
urethanes, polyether-urethanes and even methacrylates are more susceptible. The biodeterioration is
limited by crosslinking of the substrate [27] and application of proper biocides [6].

2.2. Degradation phenomena in coatings

Consumer pressure includes requirements for prolonged life span of coatings ensuring long-term
serviceability in protecting substrates. Modern automotive clearcoatbasecoat paints are expected to
retain their protective power and attractive appearance over the lifetime of the vehicle, i.e. up to five
years in harsh outdoor exposure (Arizona or Florida) are required [9].
Some general phenomena are characteristic of coating failure [29,30]. Reduction of gloss due to
erosion, yellowing and surface roughness or checking (formation of slight breaks in the film not
penetrating to the underlying substrate) and cracks are considered the first indication of the failure.
This is a result of free-radical photochemically triggered processes, including photo-oxidation or photo-
lysis and surface hydrolysis. In clearcoatbasecoat systems, the integral resistance is based on durability
of the clearcoat.
Erosion of the resin from the surface of the coating is the next step in most cases, leaving loose
pigment particles on the surface. This phenomenon is called chalking. It reduces gloss, can lead to a
complete film erosion and results mostly in drastic color changes by a shift to lighter shades due to
increased surface reflectances. The mechanism of chalking was well described [31,32].
The color is also changed as a result of photochemical changes taking place either in the pigment itself
or pigment-triggered in the system binder-coating [12]. Color fading takes place in pigmented basecoat
as a result of radiation penetration through the clearcoat. Photodegradation of the resin in the basecoat
can be accompanied by blistering due to moisture penetrating via degraded clearcoat and delamination
[33] (loss of adhesion of layers), flaking or scaling (loss of adhesion, detachment of pieces of the film
from the substrate owing to stressstrain factors) or peeling (lifting of large sheets of the coating from
the substrate).
Acid rain and acid deposits induce hydrolytic degradation and pigment discoloration [30] (see Section
2.1.2.2). As a consequence, acid attacks affect the clearcoat appearance (spotting) and durability.
Another deleterious effect is due to road salts (chemical spotting including corrosion of metal substrates)
[9].
Thermal and mechanical processes contribute to cracking of coating surfaces as a result of great
differences in thermal expansion between the coating and the substrate, and by mechanical vibrations.
Microcracks of any origin enable retention of dirt from the atmosphere resulting in blotchy appearance,
irregular color changes and enhanced biosusceptibility of mildew growth [30].
1270 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

A proper understanding of relationships between exposure conditions (radiation, temperature, atmo-


spheric pollutants, humidity), chemical composition and morphology of the resin, history of coating
formulation and curing, photochemistry of pigments or fillers (extenders), chemical changes occurring
in individual coating layers is mandatory to ensure topcoat durability and a proper selection of stabilizers
[9].

2.2.1. Coating photochemistry


All organic polymers commonly used in coatings are sensitive to a greater or lesser extent to UV light-
triggered free-radical photo-oxidation under outdoor exposure conditions [9,34]. Weathering of coatings
varies with the composition of the resinous system and interactions between the resin and involved
photoactive components. According to the first law of photochemistry, the photochemically induced
changes are effective after absorption of a photon resulting in electronically excited states [12]. Each
absorbed photon activates only one macromolecule or chromophore (Ch) in their ground state S0; thus,
the excited singlet states (Si) are generated. The energies of the higher singlet states deactivates ther-
mally very fast by vibrational relaxation and the lowest excited singlet (S1) state or, after the intersystem
crossing (ISC), the first triplet state (T1) are formed. Both S1 and T1 are the starting species for all the
following photophysical or photochemical processes encountering photodegradation and photostabili-
zation of coatings. This differentiates photodegradation from thermal oxidation arising from the ground
state S0.
Theoretically, polymers without inherent chromophores, such as most aliphatic polymers, including
acrylates, do not absorb light in the terrestrial UV range of 295400 nm. Aromatic polymers, such as
aromatic polyamides, polyesters, polycarbonates, polyurethanes or epoxy resins include in their mole-
cules chromophoric moieties making them inherently sensitive to direct photolysis and photo-oxidation.
However, not even commercially available chromophore-free polymers are any ideal systems and
absorb actinic radiation due to structural inhomogeneities and light-absorbing impurities present mostly
in trace amounts [10]. The level of structural inhomogeneities (unsaturation, oxygenated species),
arising during fabrication and curing and sensitizing degradation, gradually increases during polymer
weathering together with polymer branching and chain scission. The progressive degradation is
enhanced by non-polymeric impurities such as metallic contaminants and photoactive pigments [10].
The photochemical reaction of a polymer taking place on a high-energy excited state potential level
may result in generation of some high-energy products, free radicals. The electronic excitation is
localized and restricted within a small number of atoms involved in the chromophore itself. The
excitation energy can be, however, transferred to another part of the same macromolecule or can migrate
intermolecularly [12] by an exciton (quantum of excitation energy) hopping mechanism. Some physi-
cally defined species can arise transiently, such as excimers (excited dimers, aggregates between
aromatic moiety containing molecules in S1 and S0 states [S1 S0] arising from the exciton migration)
or exciplexes (i.e. excited charge-transfer complexes existing in electronically excited states, for exam-
ple between excited Ch and an acceptor of the energy (A), [Ch A] . This enhances the number of sites
of polymer degradation by electronic energy transfer from the light-absorbing impurity to originally
low- or non-absorbing groups in the polymer (tertiary CH bonds, some peroxides). Electron deloca-
lization also participates in energy deactivation contributing to enhanced polymer stability [35] and for
physical interferences with photostabilizers.
The absorption of a quantum of UV radiation by a chromophoric impurity does not always result in
photodegradation, although enough energy is available to dissociate chemical bonds in a polymer PH
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1271

Fig. 3. Simplified energy level (Yablonski) diagram characterizing physical pathways and sites of chemical transformation of
an excited organic molecule. S0, S1, S2 singlet electronic states, T1, T2 triplet electronic states. Processes:
0
(1) electronic
excitation (hn); (2) relaxation of higher singlet states into the first singlet state; (3) fluorescence (hnF); (4), (8), (9) and
(14) internal conversion (IC); (5) vibrational relaxation (VR); (6) autoinitiation of the excited state; (7) and (13)
intersystem crossing (ISC); (10) phosphorescence (hn 0 P); (11) singlettriplet absorption (at high light intesity only);
(12) triplettriplet absorption.

and generate macroalkyls P . This is due to photophysical processes encountering excited states [12]:
conversion to thermal energy, conversion between states (ISC), energy transfer or radiative dissipation
(luminescence), schematically described in the Yablonski diagram (Fig. 3).
The excited first-singlet state of the chromophore (S1) can release (dissipate) a part of its energy by a
radiative process, fluorescence, or by non-radiative internal conversion (IC) to heat. The first triplet state
(T1) of the chromophore is generated from S1 by non-radiative intersystem crossing. Triplet state T1 loses
a part of its energy by a radiative process, phosphorescence, and radiationless intersystem crossing to the
singlet ground state S0. The photophysical mechanism involving singlet states is not depleting for
polymer [12].
The photochemical reaction occurs, as mentioned above, from both photoexcited S1 or T1 states. The
probability of a chemical reaction in the excited state increases with their lifetime. The long-living triplet
state provides usually 10 5 times more time for photodegradation than S1 state and is more liable to
primary photochemical conversions. If thermodynamically or kinetically possible, chemical reactions
via S1 occur as well.
Formation of carbon-centered free radicals P is characteristic of initiation of photodegradation.
Macroalkyls are formed from a polymeric substrate PH by direct or sensitized photolysis, thermolysis
or metal (M) catalysis Eq. (24).

Ch; hn; D; M
PH ! P 24
1272 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Scheme 1.

The initiating process resulting in primary radicals is mostly vaguely defined, because most coatings are
rather complex systems and contain various inherent or adventitious initiating or sensitizing species. The
resistance of resins to photodegradation is related to the ease of formation of the primary radicals by
abstraction of hydrogen from the activated CH bonds (formation of intrachain radicals 1) or by break-
ing weakened CC bonds in the backbone (formation of terminal radicals 2).

Polymer-bound moieties such as CH2NH , CH2O , CH2CHyCH or COCHyCH are


oxidation-sensitive and enhance radical formation [12]. On the contrary, fluorinated or silicone-based resin
segments are resistant in the initiation phase. Free radical formation is, consequently, one of the decisive
factors in the intrinsic paint photochemistry. In automotive coatings, the clearcoat is directly attacked by
solar radiation. Its intrinsic resistance is, therefore, decisive for the entire finish durability [9].
Photochemically induced free radicals eventually react to form final products through dark (second-
ary) reactions. In oxygen-deficient atmosphere, carbon-centered radicals undergo self-reactions: dispro-
portiation (resulting in two molecular fragments in the case of radicals 2 and enhanced in-chain
unsaturation in 1) and recombination (responsible for branching or crosslinking). However, in the
presence of oxygen, oxidation chain reaction is triggered preferentially [10,12]. It involves propagation,
Eqs. (25) and (26), chain branching, Eqs. (27) and (28), and termination steps, Eq. (29).

P O2 ! POO 25
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1273

POO PH ! POOH P 26

D; hn; M
POOH ! PO HO 27

PO HO PH ! P POH HOH 28

P ; PO ; POO ; HO ! non-radical products 29


The role of free migrating radicals HO is underrated. They are active reactants yielding hydrogen
peroxide and are involved in initiation of photodegradation [12]. In Eq. (30), the alkoxy radical PO
undergoes b-scission resulting in carbonyl species.

PO ! CO
hn
30
The chain oxidation of polymers is schematically described as a cyclic process showing principal
pathways, free-radical primary products and sites of effective application of various stabilizers enhan-
cing photostability of coatings [36]. Photochemical and oxidative processes in an oxidizing polymer
(PH), and sites of effective application of stabilizers in the photo-oxidative mechanism are shown in
Scheme 1 where the abbreviations used are: CB AO, chain-breaking antioxidant; HD AO, hydroper-
oxide-decomposing antioxidant; UVA, UV absorber; LSP, light-screening pigment; Q, quencher; MD,
metal deactivator; Ch, chromophore; X, (photo)initiator.
Hydroperoxides and carbonyls are dangerous chromophores [10,12]. Hydroperoxides, much more
depleting species than carbonyls due to an easy photolysis, Eq. (27), were intensively studied as
initiating species in polyolefins and polystyrene. Excited carbonyls (Ch ) are involved in their triplet
states in sensitized formation of macroalkyls, Eq. (24), in formation of singlet oxygen by energy transfer
to ground state oxygen, Eqs. (31) and (32), and in sensitized homolysis of alkylhydroperoxides via an
exciplex, Eq. (33), (the exciplex increases the absorption of alkylhydroperoxide in the near UV).

Ch ! 1 Ch ! 3 Ch
hn
31

3
Ch 3 O2 ! Ch 1 O2 32

3
Ch ROOH ! ChROOH ! Ch RO HO 33
The energy transfer from excited chromophore may take place intermolecularly or intramolecularly,
attacking in the latter case the non-excited segments of the molecule and enhancing locally the level of
oxidation products. Polymer-bound carbonyl species are photolyzed according to Norrish I reaction to
terminal C-centered radicals and acyl radicals [12], Eq. (34), both participating in the chain branching of
the photo-oxidation. The acyl radical is a precursor of an acylperoxyl CH2C(O)OO .

CH2 COCH2 CH2 ! CH2 C O CH2 CH2


hn
34
The dangerous role of singlet oxygen, capable of yielding hydroperoxides by direct addition to unsa-
turated bonds CHyCH should be considered in pigmented coatings [12].
The briefly mentioned photochemically triggered processes are irreversible and account for changes
in the physical properties of coatings, such as embrittlement, changes in color and tensile strength,
1274 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 4. Hydroperoxide concentration versus xenon arc exposure time in an acrylic-melamine coating. Dry (B, A) and humid
(W, X, dew point of 25) exposure. Open symbols: unstabilized coating; filled symbols: coating stabilized with 1%HAS 61
(R H). (Reprinted from Polym Degrad Stab 1993;40:349, 1993, with permission from Elsevier Science [40].)

cracking, loss of gloss [7,37]. Oxygenated moieties (carbonyls, hydroxyls) increase hydrophicity of the
degraded matrix.
Each of diverse polymer types used as components in coating systems has its unique susceptibility to
photooxidation managed by spectral (wavelength) sensitivity (activation spectra) characteristic of a
particular polymer [16]. The photooxidation sensitivity is complicated by manufacturing and curing
histories of the polymer, including various additives and their transformation products [10]. The detailed
information about photochemistry of various polymers used in coatings is beyond the scope of this
paper. The reader shall find information in references dealing with oxidative processes in polyacrylates,
polystyrene, aromatic polyesters, aromatic or aromatic epoxides [7,12,38,39].
Some results obtained in mechanistic photooxidation studies are mentioned because of their relevance
to activity mechanisms and performance of photostabilizers. It was found in weathered acrylic-mela-
mine and acrylic-urethane coatings that the rate of formation and accumulation of the primary molecular
oxygenated product triggering propagation, alkylhydroperoxide, is a function of the initial free-radical
formation rate [9,34,4042]. The concentration of hydroperoxides rises more rapidly in coatings with a
high rate of radical formation, reaches a maximum and then decreases (Fig. 4). The hydroperoxide level
was found always higher in urethane-crosslinked systems (it is probably due to photolysis of urethane
crosslinks) in comparison with melamine-crosslinked acrylates. It may be extrapolated that in the former
system, the high level of hydroperoxides contributes significantly to the propagation step during long-
term photo-oxidation. Co-monomers used in synthesis of acrylic copolymers account for a rather
different behavior of construction units during photodegradation. Poly(methyl methacrylate co-butyl
acrylate) is an example [43]. Irradiation of poly(methyl methacrylate) results in chain scission and
formation of LMW fragments. Methyl methacrylate segments in the copolymer yield relatively stable
alkylhydroperoxides on tertiary CH bonds, followed by lactone and CyC unsaturated moieties. Poly(butyl
methacrylate) degrades by crosslinking and formation of short chain fragments. Relatively unstable
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1275

secondary alkylhydroperoxides and peroxyacids are formed in deeper stages of oxidation of butyl
acrylate segments of the copolymer.
The evidence on differences in the rate of hydroperoxide formation is important for explanation of
differences in activity of photoantioxidants (HAS) deactivating hydroperoxides and scavenging alkyl-
peroxyls [44].
Differences in weathering resistance were reported even for acrylic-melamine clearcoats having
different processing history and polymer variables [4547]. The more photo-oxidation-resistant clear-
coat undergoes an initial increase in crosslink density, while the more photo-oxidation sensitive one
undergoes primarily chain scission. After a prolonged weathering, the crosslinking exceeds chain scis-
sion also in this acrylate.
The photoinitiation rate was found to be proportional to the amount of initiator used for the coating
synthesis. The initiator residues influence the photo-oxidation: peroxide-based initiators are more dele-
terious in this respect than 2,2 0 -azobis(isobutyronitrile). Residues of photo-oxidation-sensitive solvents
used in solvent-borne coatings are certainly another photosensitive impurity, as indicated by comparison
of effects of xylene and heptane-2-one in acrylates [47].

2.2.1.1. Involvement of pigments, dyes and fillers (extenders) in photodegradation. In this section, a
general description of pigments and extenders (fillers) and their negative influences on polymer durabil-
ity are included. Application of pigments as light screens and interactions of pigments or fillers with
photostabilizers are included in Sections 3.1 and 4.2, respectively.
Pigments form an integral part of numerous protective and decorating coating systems [48]. They give
the desired color, hiding, hardness and protective effects. Pigments are essentially polymer-insoluble
finely divided particles dispersed in the coating vehicle and are classified as white, colored and metallic
flake pigments [30]. Titanium dioxide and zinc oxide are the principal white pigments.
Color pigments selectively absorb some wavelengths of the visible light. The selection is based on
color, cost, transparency, lightfastness and resistance to heat [30]. Carbon black, azo pigments, metal
complex pigments (phthalocyanine, azomethine, oxime-based pigments), carbonyl pigments (quinacri-
dones, anthraquinones, thioindigo, perinones, perylenes, quinophthalones, isoindolinones, diketopyrro-
lopyrroles), azomethine pigments (dioxazines), triarylmethane and related basic dye pigments or
pigments for special effects (metal flakes, pearlescent) are common colored pigments for coatings [48].
Dyes are solubilized in the polymer matrix and can develop brighter and cleaner colors than pigments.
They belong to various rather photosensitive chemical classes (derivatives of triarylmethane, anthra-
quinone, quinophthalone, perinone, azomethine, azine, furanone) [49].
Extenders (barium or calcium sulfates, calcium carbonate, clay, kaolin or silica) are inert inorganic
fillers (rather than pigments) with low refractive indices and negligible hiding [1,50]. They are generally
white or weakly colored (e.g. gray, the color being influenced by minor impurities). The tinting strengths
of extenders are low. They are used to control rheological properties of paints, improve mechanical
properties and resistance of coatings against water and aggressive gases [30].
Pigment selection can be critical in coating formulations expected to have high outdoordurability.
Some pigments are able to deteriorate the longevity of coatings by photocatalysis or sensitization. In this
respect, titanium dioxide, a versatile pigment in coatings has to be mentioned because of its strong
photocatalytic effect [2]. In 1996, almost 60% of the TiO2 production was consumed in coatings. The
two grades, rutile and anatas, differing in crystal structure are available. The size of particles of pigmen-
tary TiO2 is about 200 mm. Rutile has a low light absorption and high refractive index [30]. The light
1276 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

scattering at 560 nm increases with TiO2 concentration in the polymer matrix up to 10%. Anatas has a
lower refractive index and the light scattering is less effective. As a consequence, both grades of TiO2
produce different effects in the polymer matrix.
Owing to the outstanding properties and ability to act either as a light screen or thermo- and photo-
catalyst, TiO2 pigments were studied in detail in sunlight-exposed polymers. The photocatalytic effect is
influenced by particle size and surface treatment of the pigment and results in formation of free radicals
in the binder. The mechanism of the initiation process was studied [31,32]. After irradiation of TiO2 with
light of wavelengths less than 405 nm, the absorbed energy promotes electrons from the valence band to
the conduction band, leaving positively charged holes in the valence band. Holes and electrons may
move within the crystal lattice. Some free holes and electrons are eliminated by recombination.
However, those remaining on the crystal surface trigger chemical reactions. Positive holes combine
with hydroxyl groups on TiO2 surface and account for HO radical formation, Eqs. (35) and (36).
Electrons reduce Ti 4 to Ti 3. The latter is reoxidized by adsorbed oxygen, Eq. (37). In the following
steps, oxygen anion-radical, Eq. (38), O-centered radicals HO and HOO , Eqs. (36) and (37), and
hydrogen peroxide, Eq. (39), are formed.

Ti4 HO ! Ti4 HO e
hn
35

Ti4 HO ! Ti4 HO 36

O2 H
Ti4 e ! Ti3 ! Ti4 O2 ! Ti4 HOO 37

e O2 ! O
2 38

2HO ! H2 O2 39
The net photocatalytic process yielding H and HOO radicals, triggering the formation of macroalkyls P
taking part in photo-oxidative chains, involves humidity as a source of hydrogen, Eq. (40).
TiO2
H2 O O2 ! HOO HO 40
The heterogeneous photocatalysis of polymer degradation by TiO2 can be limited either by admixing
transition metals (zinc, aluminum) acting as recombination centers for electron and holes or by coating
TiO2 particles with alumina or silica promoting destruction of HO radicals [31].
The photocatalyzed oxidation of a pigmented paint system is localized in the vicinity of TiO2 crystals
and results in erosion (mass loss) of the polymer material around TiO2 particles, the pigment particles
laying bare in the upper paint layers and sticking on being touched. This chalking effect was well
described [31,32] (see Section 2.2.). Anatas pigments show a far higher degree of chalking than rutile
pigments and are less suitable for outdoor exposed paints [3032].
The photocatalytic effect of white pigments (titanium dioxide, zinc oxide, cadmium sulfide) was
studied in hydrocarbon-based polymers, such as polyolefins, ethylenepropylene thermoplastic elasto-
mers or ethylenepropylenediene terpolymer (EPDM) and was shown to be concentration-dependent
[51]. The original UV screening effect observed at lower concentrations was converted in photocatalysis
at increased concentrations. Surface-passivated TiO2 lowers the amount of formed carbonyl groups.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1277

Effect of extenders on polymer photodegradation depends strongly on the surface active sites and
content of photoactive metal impurities.
Photoexcited chromophores of organic colored pigments and dyes are able to initiate formation of
macroalkyls in the polymer matrix, Eq. (24), and sensitize photooxidation of polymers [12] or stabilizers
containing the phenolic moieties [22]. The final effect of pigments on polymers, either protecting or
promoting photo-oxidation changes, depends on the light-absorbing characteristics of pigments, content
of metallic impurities, surface properties, humidity, inherent polymer sensitivity and the mode of
the energy transfer from the excited pigment to the resin. Chemical or physical interactions on the
interfaces between the filler and the matrix polymer or stabilizer cannot be underrated. Permanent
(covalent) or dissociating H-bonds can be formed. It was shown that some yellow, orange and red
pigments had a detrimental effect on polypropylene [52]. No explanation based on the structural effect
was presented.
The lightfastness of a dye or a pigment is determined largely by the photochemical characteristic,
triplet state energies and physical state of the colorant itself. Ionization, free radical formation or
oxidation of the dye are involved, some of the processes being irreversible. The dye photodegradation
is manifested by changes in shade or depth of the shade (fading accounting for both reactions asso-
ciated with the dye itself or dyepolymer interactions) and by dye-induced (sensitized) degradation of
the polymer material (phototendering) [53,54]. Dyes with higher activation energies of photoconduc-
tivity show a high lightfastness.
The phototendering is mostly due to the formation of excited triplet states of the dye inducing
formation of singlet oxygen, hydrogen peroxide or macroalkyls by H-abstraction from the polymer.
Humidity and oxygen have a marked accelerating effect on the tendering rate [54]. There is obviously no
exact correlation between chemical structure of the dye and its tendering activity.

2.2.2. Coating hydrolysis


Resins having in the backbone or crosslinks ester or melamine moieties undergo hydrolysis. During
hydrolysis of melamine-crosslinked acrylate coating, the bonds between the acrylic copolymer and the
melamine crosslinker are broken and new melaminemelamine crosslinks are formed [55,56]. Residues
of acid catalysts used for curing and environmental acidity (see Section 2.1.2.2) account for severe
hydrolytic damages. Urethane- and epoxy-crosslinked acrylates are inherently more resistant than
melamine-crosslinked systems [9,56].

2.2.3. Assessment of weathering of coatings


Natural weathering trials and various accelerated weathering tests expected to simulate more or less
the outdoor weathering chemistry of coatings provide data mandatory to understand the inherent resis-
tance of coatings to photoinduced processes and to determine the performance of photostabilizers used
to prolong lifetime of coatings [9]. Various physical and mechanical properties affected by weathering
are monitored, such as loss of gloss, yellowing, pigment color change, chalking, checking, haze, hard-
ness, embrittlement, flexibility retention, mass loss, surface cracking, adhesion of layers, flaking, peeling
or acid spot resistance.
Formation of carbonyl groups, hydroperoxides, residual UVA or free-radical products of HAS after
natural or accelerated weathering are mostly monitored to describe chemical changes in degradation of
1278 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

coatings. Correlation of changes of selected properties under natural and accelerated tests serves for
elaboration of schemes for predicting coating service durability [34].

3. Photostabilization of coatings

The general scheme of the photo-oxidation of organic polymers (binders, Scheme 1) indicates sites of
the effective application of various classes of stabilizers inhibiting or at least reducing photodegradation.
Additives with different chemical structures, interfering with the initiation and/or propagation steps of
the chain oxidation have been used in coatings. Their structures have to be adjusted to the chemical
nature of the polymer to be stabilized and application conditions.
To prevent the polymer matrix from attacking by UV radiation, light screening pigments (see Section 3.1)
and UV absorbers (UVA, see Section 3.2) are used. Photoantioxidants (see Section 3.3) deactivate peroxidic
species arising in photo-oxidation. Other photostabilizers not included to Sections 3.13.5 are not common
at present in coatings and are mentioned only briefly. Quenchers (Q) deactivate excited species in polymers,
such as carbonyls [CO] , Eq. (41), by collision energy transfer within a distance 1015 A or by resonance
transfer over extensive regions of space (50100 A) as a result of Coulombic interactions [57].

CO Q ! CO Q 41

Quenchers dispose the energy as heat or fluorescent and phosphorescent radiations. Their application in
coatings is not common. Most Q are salts or chelates of nickel [58] thus imparting green color to the polymer.
Compatibility problems and ecological objections against contamination of the environment by nickel lower
the application possibilities of Q.
Metal deactivators and photostable hydroperoxide decomposing antioxidants prevent photolysis of
hydroperoxides [58]. These groups of stabilizers improving photostability are not used in coatings.
UVA and HAS contribute to polymer protection by more than one mechanism. Mostly, the inherent
chemical efficiency of the parent stabilizer and of its transformation products is decisive for the final
effect. HAS are the best example: the true stabilizing species are formed during sacrificial transforma-
tions of the parent amine [59].
Because of the complexity of species that have to be deactivated in photodegrading coatings, a
complex protection requires a proper combination of stabilizers (UVA with HAS) differing in activity
mechanisms [5,58]. Some bifunctional stabilizers (see Section 3.4) provide beneficial protection by
exploiting intramolecular cooperation between two active moieties bound in one molecule [58].
Chemical structures of various photostabilizers (Sections 3.23.5), activity mechanisms based on
their inherent chemical efficiency controlled by functional groups and proper molecular architecture (see
Sections 3.2.1 and 3.3.1), involved photophysical processes, physical persistence managed by molecular
weight and compatibility with the polymer matrix (see Section 3.5), processes depleting activity of
photostabilizers (see Sections 3.2.2 and 3.3.2), effect on durability of coatings and potential interferences
with other additives, pigments and fillers (see Section 4) are outlined.
To satisfy expectations for effective commercial application, photostabilizers for coatings have to
fulfil a set of requirements [5,58].

High efficiency at appropriate concentration and economically acceptable price.


J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1279

Optimum relations between inherent chemical efficiency, solubility, and partitioning in multilayer
systems.
High inherent thermal stability at curing temperatures, chemical and photochemical stability to
environmental pollutants and oxidation products arising in the binder.
Absence of photoinitiation and sensitization effects.
Absence of discoloring and staining in the original form or in transformation products.
Chemical inertia related to the protected polymer matrix and involved contaminants.
Physical resistance to volatility and leaching.
Optimum cooperation with other components of the stabilizing system (other additives, photoinitia-
tors, pigments).
Commercialization in processing forms in accord to rules of industrial hygiene (dust-free systems,
one-pack concentrates of stabilizer blends).
Toxicity corresponding to legislative requirements and environmental rules.

3.1. Light screening pigments

The photostabilizing effect of pigments can be attributed to screening (minimizing penetration) of


harmful radiation, selective absorption of the incident light and, to some extent, deactivation of photo-
excited species [54].
Potentials of light screening by pigments depend on construction of the coating. Photodegradation
takes place preferentially on the coating surface. In mono-coat systems, the penetration of light to the
paint bulk is restricted by the pigment. In clearcoatbasecoat systems, the light passes through the
unpigmented clearcoat to the pigmented basecoat and primer [1,5]. Hence, the pigment cannot protect
the surface layer (clearcoat).
The protective effect of common inorganic pigments, ultrafine TiO2 (diameter 20 mm) and carbon
black is fully exploited in polyolefins. Application in automotive coatings has some drawbacks [5,7].
The pigments protect only domains of the binder, which are beneath the screener resulting in poor
surface protection, and cannot be used in clearcoats. The potential photocatalytic effect of TiO2 on
binder degradation was described in Section 2.2.1.1.
Carbon black acts as a UV absorber through energy level transitions in its polynuclear polyconjugated
structure arranged in parallel ring-forming layers [60]. Contribution of free radical scavenging to the
final effect is considered. Excellent photostabilization can be obtained with 2.55% carbon black having
particles of 1525 mm [53]. Adsorption of stabilizers on carbon black surface as well as binding of
phenolic or amine moieties of stabilizers via H-bonds to oxo groups located on edges of layers forming
carbon black are potentially able to reduce activity of UVA and HAS.

3.2. UV absorbers

Absorbers of UV radiation (UVA) are colorless or nearly colorless compounds having high absorption
coefficients in the UV part of the solar spectrum. They protect coatings against photoinduced damages
by absorbing the harmful actinic solar radiation preferentially to binders [5,7,61]. Formation of excited
chromophoric impurities is thus prevented. UVA must absorb, above all, radiation in the region between
290 and 350 nm (having the highest energy in the solar spectrum). However, the range of the absorption
should be extended up to 400 nm taking into account protection of binders in the presence of pigments
1280 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

and dyes. UVA transform the absorbed radiation energy into less harmful thermal energy via a photo-
physical process involving ground state and excited state molecules [61] (see Section 3.2.1). Light
screening, quenching of excited states and scavenging of alkylperoxyls is considered to be involved
to some extent in the UVA mechanism [62]. The two first activities are beneficial. The third one,
characteristic of phenolic UVA, can be considered as a process depleting the excited state intramolecular
proton transfer [22,58] (ESIPT, see Section 3.2.1).
An ideal UVA would be expected to absorb all UV terrestrial solar radiation but no visible light and to
have high inherent photostability [58]. To achieve maximum performance, some UV transparency or a slight
yellow discoloration of UVA may be tolerated. Compounds with different phenolic and non-phenolic
structures fulfil commercial expectations for long-term performance. The phenolic-type UVA involve
compounds forming OHO bridges, such as salicylates 3 (Ar phenyl, 4-tert-butylphenyl), 4-substi-
tuted-2-hydroxybenzophenones (BP) 4 (R CH3, C8H17, C12H25), 2,2 0 -dihydroxybenzophenones 5, bridged
BP, e.g. 6, carbamate 7, 3-hydroxyflavone 8 or xanthone 9. Various other phenolic compounds capable of
participating in the ESIPT mechanism by OHO tunnelling, such as 7-hydroxyindan-1-one, 3,4-benzo-
tropolone or 3-hydroxypyridinone are potential photostabilizers. It was reported; however, that not all
aromatic hydroxy ketones are lightfast [63]. For example, 1-hydroxyfluoren-9-one is light sensitive.
Compounds involving O H N tunnelling are the second group of phenolic UVA. More structural
types are involved, such as 2-(2-hydroxyphenyl)benzotriazoles (BT), 1014, 2-(2-hydroxyphenyl)-
pyrimidine 15 and 2-(2-hydroxyphenyl)-1,3,5-triazines (TA) 16, 17. Various other compounds with
potential OHN tunnelling were elucidated as UVA, e.g. salicylidene anilide 18, 2-(2-hydroxyphenyl)-
benzimidazole 19 (X NH), benzoxazole 19 (X O) or benzothiazole 19 (X S), 3-(2-hydroxyphenyl)-
pyrazole 20 or 5-(2-hydroxyphenyl)pyrazole 21.
Non-phenolic UVA include oxanilides (OA) 2225, formamidine 26, benzylidenemalonate 27 or 2-
cyanoacrylates 28, 29. Derivatives of BT, TA and OA are the most widely used UVA in coatings
[5,7,61].
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1281

3.2.1. Activity mechanism


The effectiveness of UVA in coatings is determined by their absorption characteristics, inherent
photostability and efficiency to dissipate in harmless way the absorbed energy, by concentration in
the coating, the film thickness, chemical interaction with the binder and other additives under photo-
degradation conditions.
An absorption maximum between 330 and 350 nm in the UV absorption spectrum of UVA is consid-
ered a necessary condition for an adequate efficiency in coatings [5]. UVA having l max too far in the
short-wavelength region are unable to cover sufficiently the harmful long-wavelength region up to ca
380 nm. On the contrary, UVA having l max over 350 nm may account for inherent yellow tinge. This can
adversely affect color in the coating. Derivatives of BT and TA have, according to their spectral
characteristics, the most beneficial properties in coating protection. Absorption spectra of representative
UVA are shown in Fig. 5.
Phenolic UVA with structures of substituted BP, BT and TA having intact intramolecular H-tunneling
show two absorption maxima, one in the UV-B at about 300 nm and another in UV-A above 320 nm (as
measured in chloroform) [64]. Triazines show the strongest absorption in UV-B within the phenolic
UVA, and are followed by BT and BP. The second (UV-A) maximum for BT is in the region 340
350 nm, for TA at 335340 nm and for BP at 320330 nm. The exact position of the long-wavelength
1282 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 5. Absorption spectra of UV absorbers in chloroform solution at concentration 1 10 4 M, optical pathlength 1 cm: (1) 4
R C8H17); (2) 10d; (3) 17a (5 10 5 M); (4) 29 (R i-C8H17); (5) 22 (R C2H5).

absorption maximum is influenced by substituents on the UVA molecule and solvent polarity used for
spectral measurements (an interruption of the H-tunneling may take place in hydroxylic solvents, see
Section 3.2.2.1). Some BP may cause undue yellow initial color due to their spectral properties [63]. UV
absorption spectra of BT show a bathochromic shift caused by the chlorine atom in position 5 (such as in
11, R Cl) and by substitution of positions 3 0 and 5 0 . Benzotriazoles have, in comparison with BP,
higher molar absorption coefficients and steeper absorption edges towards 400 nm.
Oxanilides 2225 have only one absorption maximum [64] above 300 nm. Their photoactivity is
essentially attributed [65] to the inner filter effect in the wavelength region 280340 nm. In this wave-
length region, the absorption coefficients of OA are higher than 10 4 dm 3 mol 1 cm 1, i.e. much larger
than that of polymer-bound carbonyl chromophores or peroxidic linkages. The position of the absorption
maxima is independent of medium polarity.
2-Cyanoacrylate 29 absorbs mostly in the near UV region [66] between 290 and 320 nm. Its activity is
based on the inner filter effect combined with some quenching power.
UVA with steep absorption edge reaching into the long-wavelength UV region absorb more
effectively the harmful radiation. Surprising stabilization effects can be achieved using blends of
UVA from various groups due to superposition of their absorptions.
Absorbance (A) may be regarded as a measure of the screening effect of a particular UVA [5,7,66,67].
An effective UVA should have high absorption coefficients in the wavelength range in which the
polymeric material is most susceptible to photodegradation [67]. Absorbance (A) is defined by the
LambertBeer law, Eq. (42):
A log I0 =I ecd 42
where A is the absorbance, I0, I the intensity of incident and emergent light, respectively, e the molar
absorption coefficient of UVA in l mol 1 cm 1, c the concentration of UVA in mol l 1, and d the film
thickness in cm.) It is assumed that the higher is the absorption coefficient at a wavelength maximum
(usually, the l max in UV-A is considered), more radiation will be screened and the greater will be the
stabilizing effect. The inherent photostability of UVA under measured conditions is suggested.
According to Eq. (42), the absorbance depends on e and concentration of UVA and film thickness of
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1283

the unpigmented coating. This reveals that the final effect of a particular UVA having a fixed e changes
by varying either its concentration or film thickness. Because the total light absorbance by UVA is a
function of the pathlength, UVA are less effective in protecting thin coating films and coating surface
[7,61]. To enhance the stabilizing effect, either the concentration of UVA or the film thickness must be
increased. The former approach is mostly used: relatively high levels of UVA (0.253%) have to be
1284 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

added to commercial coatings [5,7,61,66]. Because of solubility and compatibility problems, this may
limit the effective application of UVA in very thin coatings.
In the description of activity mechanisms, it is necessary to distinguish between phenol-type UVA 3
21 and the non-phenolic UVA 2229. Much more attention has been paid to the former type.
In principle, UVA must absorb terrestrial UV light more effectively and faster than the UVA-doped
coating, and must transform and dissipate the absorbed energy in its singlet state (S), i.e. before UVA-
depleting photoprocesses involving ISC to the triplet state (T) can be initiated [68,69].
The activity mode of phenolic UVA, compounds with intramolecular hydrogen bonds (IMHB),
has been the subject of numerous investigations. The mechanism was described for salicylates [70],
benzophenones [64], 3-hydroxyflavone [71] and related molecules, hydroxychromanone, xanthone
[63,68], salicylidene anilide [72], benzotriazoles [67,68,73,74], triazines [68], 3- and 5-(2-hydroxy-
phenyl)pyrazoles [78], 2-(2-hydroxyphenyl)benzimidazole [79], 2-(2-hydroxyphenyl)benzothiazole
[71] or 2-(2-hydroxyphenyl)benzoxazole [63].
Generally, these phenolics absorb damaging UV radiation and transform it into vibrational (thermal)
energy via highly efficient radiationless deactivation pathways. The activity was explained in terms of
the ESIPT mechanism. Theoretical models that have been proposed for explanation of proton-trans-
ferred reactions were reviewed [71,7983]. The understanding of ESIPT is an indispensible part of
photochemistry and photophysics, fully exploited in explanation of polymer photostabilization.
ESIPT is responsible for the exceptional photostability of phenolic UVA and its preservation from
depletion by structural elements is a necessary condition [58]. The molecular architecture enabling
formation of IMHB and involving the ESIPT process does not impart a priory an effective UV absorbing
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1285

stabilization potential, as shown for derivatives of 2-(2-hydroxyphenyl)benzoxazole 19 (X O), benz-


imidazole 19 (X N) or thiazole 19 (X S). These compounds are light sensitizers [63].
Steady-state and ultrafast laser-based picosecond fluorescence spectroscopy, time-resolved emission
and transient-absorption spectroscopy, femtosecond resolution techniques, laser induced opto-acoustic
spectroscopy or phase fluorometry were used to elucidate very rapid photophysical processes in phenolic
UVA [7377,84].
The very rapid ESIPT mechanism is the pivot of the photophysics of UVA. It is facilitated in
planar five- and six-membered rings having IMHB between the phenolic hydrogen and hetero
atom, either a nitrogen atom (in BT, TA and other N heterocycles) or oxygen atom (in salicylate,
BP, flavone or xanthone) in the ground state S0 of the phenolic form [71]. First excited singlet S1
is formed after absorption of light. The molecule remains in the phenolic form, showing,
however, a strongly enhanced acidity of the phenolic group (for about 7 pKa units) in comparison
with S0 [71,72,7981]. The effect of the electronic excitation may be dramatic and is reflected in
phototriggered changes in pK values of nitrogen and oxygen bases and acids, respectively [72,80]. The
rate of protonation of basic species accounting for a tautomer may be increased by 1114 orders of
magnitude in accordance with the Forster theory in molecules containing both acid and basic functions
in one molecule in a reasonably close proximity [71] promoting an intramolecular proton transfer which
is the basis of the large Stokes shift in the fluorescence spectrum. The acid group in the molecule of UVA
is invariably a phenolic hydroxyl. The basic acceptor is usually either a heterocyclic nitrogen atom or
oxygen of a carbonyl function [79,81], both functions being in a position favoring IMHB in the ground
state.
The electronic distribution in the first excited singlet state S1 favors a fast proton transfer to the
heteroatom (the rate is about l0 11 s 1) and accounts for proton-transferred (zwitterionic) excited
singlet state of the tautomer form S1 0 with strong IMHB [85]. The excited tautomer S1 0 is believed to
dissipate the excitation energy by a rapid internal conversion (IC), a radiationless non-degradative
process, accounting for the ground state of the tautomeric form S0 0 . Fluorescence emission with an
unusually large Stokes shift from the proton transferred S1 0 was reported [70,72,74,76,85,86]. In most
cases, the fluorescence of the primary excited form does not appear at all. The lifetime of the proton-
transferred state S1 0 is very short. The dynamics of the tautomer formation encompasses a large range of
time domains, ranging from femtoseconds to microseconds. A majority of the relevant processes occurs
in the picosecond domain [71].
In the final phase of the ESIPT, proton jumps back (a reverse proton transfer) and the original ground
state S0 of the phenolic form is regenerated. The transformation S0 ! S1 ! S1 0 ! S0 0 ! S0 occurs on an
ultrafast timescale (40 ps) for highly photostable UVA molecules [67]. The fast physical process
S1 0 ! S0 accounts for the long service time of UVA. In spite of the present knowledge of the ESIPT
mechanism, some so far unknown deactivation channels of the initially populated electronic states are
not excluded [71].
Heat energy is released to the medium [87]. The mechanism can be repeated as long as the IMHB and
planarity in the UVA molecule remain intact [71,84]. ESIPT can be, therefore, considered the key rapid
step in deactivation of excited UVA molecules [63,64,67,68,71,72,75,77,80,85,88,89]. Fig. 6 illustrates
schematically photophysical processes taking place in phenolic UVA.
The main result of the cycle in ESIPT is the conversion of the electronic excitation energy to thermal
vibrational energy by the radiationless processes. All experimental results indicate that both the structure
of the UVA and the nature of the surrounding environment determine the efficiency of the energy
1286 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 6. Scheme of excited state intramolecular proton transfer (ESIPT) mechanism of phenolic UVA. S0, S0 0 and S1, S1 0
ground states and first excited singlet states of the phenolic and tautomer form, respectively, T1, T1 0 triples states, IC
internal conversion, ISC intersystem crossing, hnF 0 fluorescence, hnP 0 phosphoresence.

dissipation by phenolic UVA. The best lightfastness and photostabilizing performance is exhibited
by compounds with the highest inherent chemical resistance to breaking IMHB in photo-oxidizing
binder.
Geometric arrangement of the whole UVA molecule is of fundamental importance for the intramo-
lecular H-tunnelling [67,75,76,85,87,88]. Various theoretical findings in elucidation of photophysics of
chromophoric molecules are of top importance for development of the architecture of UVA. It is a
general rule in aromatic molecules that electron-donating substituents become stronger donors in the
excited state, while acceptors will attract the electrons more strongly [80] in a good correlation with s
and s Hammett substituent constants. This is because ionic resonance structures contribute much more
to the excited states than to the ground state [90]. The effective ESIPT mechanism proceeds only in
planar structures and is strongly influenced by the polarity of the environment. Loss of coplanarity (e.g.
in BT, twisting of the two rings against each other is about 56 due to the solutesolvent effects results in
disruption of ESIPT [67,69,73,75,88].
The transfer of proton between two functions in UVA accounts for electronic and structural rearran-
gements associated with significant changes in dipole moments, molecular geometry and a rather large
(10 000 cm 1) fluorescence shift [71]. As a consequence, the dynamics of the proton transfer process
can be strongly influenced by solvents namely with regard to the formation of H-bonds. Intersystem
crossing to a long-living triplet state takes place in non-planar structure and accounts [70] for depletion
of the UVA efficiency and sensitization of the polymer matrix oxidation, Eq. (43) (see Section 3.2.2).
The photoproducts are formed from the excited non-planar structure of UVA [67].

UVA ! S UVA ! T UVA ! Products


hn ISC binder
43

It can be assumed from photophysical elucidations, that the lightfastness of UVA is an inverse of the
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1287

quantum yield of a photochemical reaction [75,89] and is proportional to the rate of IC, the latter being
inversely proportional to the energy difference between the first excited state S1 and the ground state S0.
The difference is smaller in the keto form (S1 0 ) than in the enol form (S1) (Fig. 6). This was confirmed in
a number of experiments with UVA forming IMHB O H O or O H N. It was evidenced very early
[69] that effective phenolic UVA have much higher rates of the IC than that of ISC and of fluorescence
decay.
Some specific features encounter individual groups of phenolic UVA. Esters of salicylic acid 3 have
low initial absorption and can be used only in systems with low photosensitivity [63]. Aromatic sali-
cylates 3 (R 4-methylphenyl, 4-tert-octylphenyl, 2,4-di-tert-butylphenyl) act by the ESIPT mechan-
ism [72] involved in the original structure and via 2,2 0 -dihydroxybenzophenone 5 (R H) resulting
from photo-Fries rearrangement of 3. Yellowing under UV radiation is a serious disadvantage of
common aromatic salicylates [64]. Moreover, the photo-Fries reaction is too slow and cannot be the
substantial mechanism of 3. A proper substitution, such as in 2,6-dimethyl-4-octadecylphenyl salicylate
provides a non-yellowing and non-rearranging salicylate ester acting presumably by ESIPT.
Derivatives of benzophenone are lightfast UVA with an effective ESIPT [66]. A variety of mechan-
istic pathways following excitation to S1 and even formation of T1 and T1 0 states was reported [64,66,91].
It is rather difficult, on the basis of available experimental data, to decide which of proposed cycles
operates predominantly in the ESIPT. The ISC is possible among the pathways.
Transient absorption revealed that the population of the ground state S0 of the excited species is in a
non-polar (hexane) solution in the range 35 ^ 5 ps. In ethanol, the kinetics is more complex, with the
1.5 ns lifetime due to the contribution of the intermolecularly H-bonded S0 and involves an excited
triplet state. The performance of BP is optimized by a proper substitution on 4-alkoxy group [66].
Besides absorbing UV light, BP were reported as weak quenchers of excited carbonyls with a negligible
power to quench the singlet oxygen [64].
Some other less common compounds forming OH O IMHB were proposed for specific applica-
tion as UVA. 5-Hydroxy-7-methoxy-2,2-dimethyl-6-octylchroman-4-one was found to be an active but
rather photosensitive UVA [64]. The importance of the presence of phenyl group in position 2 and
coplanarity of the molecule preserving formation of IMHB in polar medium was reported [81] for 3-
hydroxyflavone 8. Xanthone 9 is an effective lightfast UVA. It absorbs, however, close to the visible
light region and is, consequently, yellowish [63].
Much attention has been paid to UVA forming OH N IMHB. Photophysical elucidation of
salicylidene anilide 18 reveals that proton transfer from the excited tautomer occurs within 5 ps at
temperatures above 4 K in both protic and aprotic solvents [72]. Fluorescence kinetics observed at
low temperatures appears to manifest excited state vibrational relaxation taking place on the 10 ps
time scale. Fluorescence has two components arising probably from cis and trans keto tautomer
forms. 2-(2-Hydroxyphenyl)benzoxazole 19 (X O), benzothiazole 19 (X S) and benzimidazole
19 (X N) involve IMHB and the ESIPT mechanism. In spite of this, benzothiazoles and benzimid-
azoles are light-sensitive and not suitable as UVA [63]. In electron ground state, the enol form of 2-(2-
hydroxyphenyl)benzoxazole 19 (X O) is the stable tautomer [92]. A very rapid proton-transfer takes
place from S1 to S1 0 with high fluorescence yield of the proton transferred species. This is a result of the
absence of an efficient radiationless deactivation process. S1 0 is depopulated by thermally activated IC
and fluorescence to S0 0 and by ISC to the ketonic triplet state T1 0 (the latter process proceeds within a few
nanoseconds) [92]. Intersystem crossing to the triplet state from the proton transferred singlet form
S1 0 ! T1 0 is followed by reverse proton transfer within the triplet system T1 0 ! T1 and by radiationless
1288 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

deactivation of the energetically low-laying T1 to the singlet ground state S0. An equilibrium is estab-
lished between triplet keto and enol states [79]. H-tunneling in the metastable triplet state is populated
nearly exclusively from the excited enol form [93]. In spite of a sufficient lightfastness, the benzoxazole
provides no UV protection.
3- or 5-(2-hydroxyphenyl)pyrazoles 20, 21 undergo ESIPT in the first excited singlet state
and yield a singlet state tautomer which either emits very weakly or is completely non-emitting
[78]. Application of 20, 21 as UVA is potentially possible similarly to that of planar derivatives of
1-(2-hydroxyphenyl)pyrazole.
UVA of the 2-(2-hydroxyphenyl)benzotriazole class provide excellent photoprotection to coatings.
Photophysics and photochemistry of BT was studied in detail. Excitation of BT is accompanied by
fluorescence emission from the proton-transferred form at about l max 630 nm and Stokes shift
10 000 cm 1 attributed to the S1 0 state in non-polar glasses [67] and in the crystalline form [76].
The ESIPT mechanism was generally accepted for BT [67,71,73,74,76,77,80,84,85,8789,94]. The
lifetime of the proton-transferred state S1 0 of 10a and other BT is short, ranging from 90140 ps and
reducing the probability of participation of S1 0 in any photoinitiation step and depletion of photoche-
mical transformation of BT [84]. Owing to electronic rearrangement following excitation, the hetero-
cyclic nitrogen becomes more basic while the phenolic hydroxyl more acidic [67]. The ESIPT in BT is
followed by internal vibrations of the planar molecule (stretching vibrations and out of plane bending
vibrations of hydrogen of the IMHB, torsional vibrations around the central bond CN in BT
[63,68,69,72,76,80,84]. The IMHB in BT provides the energy for the planar arrangement [76]. Its
stability as well as probability of ESIPT increases with the proton-donor and proton-acceptor ability
of the phenolic hydroxyl and nitrogen atom, respectively.
X-ray and UV spectroscopic investigation in the range of 35012 K confirmed [84] the planar
arrangement of solid 10a and 10d [84]. The strength of the IMHB and lightfastness are enhanced by
methyl and tertiary alkyl groups in positions 3 0 or 5 0 [63,67,73,85,88]. Very strong IMHB and high yield
of proton-transferred fluorescence was reported for 10b. Methyl group located in position 6 0 (in ortho
position to the CN bond connecting phenolic and BT rings) results in loss of coplanarity; the derivative
30 has only a very weak IMHB and a noticeable dihedral angle (56.354.9, almost a perpendicular
position) between the two rings [76,85]. This is a difference from 10ac (having one bulky substituent in
position 5 0 ) or from 10d, f, g having a bulky substituent in position 3 0 favorizing a stronger IMHB. All
these BT have virtually coplanar structure with a dihedral angle having zero value [67,76,85,95].
A potential existence of an undefined excited intermediate X between excited singlet states S1 and S1 0
was postulated in the BT mechanism [84]. It was speculated that the intermediate X can have a config-
uration twisted around the CN central bond (angle 90) and/or an out-of-plane bent H-atom in IMHB.
Absorption spectra of BT in various solvents and polymer matrices differing in polarity reveal the
existence of planar and non-planar ground state forms [88]. The equilibrium between them is affected by
both the polarity and the hydrogen bonding strength of the medium. The excited-state non-radiative
processes are thus affected by solvation. Influence of the microenvironment acting as an acceptor (A:) of
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1289

Scheme 2.

a proton of the phenolic hydroxyl is shown in Scheme 2. Following excitation, it is possible to observe
emission from both S1 and T1 states in BT either intermolecularly bonded with a solvent matrix or in
derivatives like 30 having the intramolecular bond broken by steric effects of the substituent in position
6 0.
The specific role of the bulky ortho substituents to phenolic hydroxyl restricts opening of the
intramolecularly H-bonded six-membered ring hindering the rotation of the hydroxy group and blocking
ISC to excited triplet form. This substitution mode is fully exploited in most commercialized BT-based
UVA. Moreover, the importance of the substitution of the position 5 0 by a propionate group, such as in
10i to 10l regenerating ESIPT after the depleting consumption of the phenolic part [22,58,96] should be
noted (for mechanistic details, see Section 3.2.2.2).
1,2,4-Triazoles have a remarkable lightfastness. However, they discolor polymers [63]. Present inter-
est is concentrated on asymmetrically substituted 2-(2-hydroxyphenyl)-1,3,5-triazines having aryls in
positions 4 and 6. They are lightfast [63], have a low basicity and high resistance to protonic polar
environment [74]. This makes them excellent UVA for coatings. Experimental data reveal larger long-
term performance of TA in comparison with effective BT [97].
Influence of substituent effects on photophysics of TA was exploited to optimize properties of TA as
UVA [74]. Derivatives having only one aryl, such as 2-(2-hydroxyphenyl)-4,6-dimethoxy-1,3,5-triazine
exhibit a proton-transferred fluorescence at l max 510 nm with both a large Stokes shift (9 890 cm 1) and
high quantum yield [74,77,98]. Introduction of the second aryl into positions 4 or 6 effectively weakens
or completely quenches the proton-transferred fluorescence [74,75]. The latter is therefore strongly
affected by the electron-donating capacity of substituents on the TA ring [77]. It has been assumed
that these substituents accelerate the radiationless deactivation by offering additional vibration modes
[77,84,98]. Both kinetic aspects and electron density factors are involved in the drastic decrease in the
quantum yield of the proton-transferred emissions in 2,4-diaryl-6-(2-hydroxyphenyl)-1,3,5-triazines
[77].
Substitution of aryls in positions 2,4 for phenoxy groups results in both short- and long-wavelength
fluorescence. The existence of two ground state conformers was thus suggested [75].
The ESIPT mechanism of TA was studied in detail [74,75,77,98]. It was proposed that for the excited
2,4-diaryl-6-(2-hydroxy-5-substitutedphenyl)-1,3,5-triazines the IC of the S1 state is associated with an
out-of-plane bending vibration of the hydroxy group. The temperature-dependent radiationless deacti-
vation is related to the strength of the IMHB. Triaryltriazines have entirely closed-ring structures
because of strong IMHB [98] enabling effective ESIPT after excitation. ESIPT is characteristic of green
fluorescence with a large Stokes shift (10 000 cm 1). The low emission yields are consistent with an
efficient radiationless relaxation process [75].
1290 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Scheme 3.

The proton transfer rate in the excited singlet state S1 ! S1 0 in poly(methyl methacrylate) (PMMA)
matrix was reported [74,98] to be 2 10 11 s 1 at 77 and 300 K, respectively, having no potential barrier.
The ground state proton transfer S0 0 ! S0 occurs relatively slowly (10 3 s 1). No proton-transferred
phosphorescence was observed in TA with three aryls [77]. X-ray and spectral (IR, NMR) data and
effects of solvent effects measured by UV spectroscopy indicate that IMHB in triazines is stronger than
that in benzotriazoles [74,76,77,85]. This results [77] from a longer O H distance and shortened N H
distance and definitely a much more linear H-bond (the angle O H N in TA is 159.6). Rotation
around the C2C2 0 bond allows the switching of H-tunneling either to N1 or N5 within the 13C NMR
time scale.
Not much information is available on the photophysical mechanism of non-phenolic UVA [65]. The
class of oxanilides 2225 is notably important for coatings. Single bonds in the oxanilide bridge can
exist in various geometrical structures. IR and Raman spectroscopic investigation indicates the trans-
planar geometry and presence of intramolecular H-tunneling between carbonyl and imine groups. This
implies a planar structure and a small solvent sensitivity of the UV absorption spectra. Experimental data
indicate probability of involvement of the ESIPT mechanism [65] (Scheme 3).
The photochemical stability of oxanilides is due to a very short lifetime of the first excited singlet state
S1, similar to that of the phenolic UVA. It is suggested from the photoelectron spectra that the lowest p
p transition is characteristic of a net electronic charge transfer from the aniline moieties towards the
carbonyl groups [65]. The ESIPT yields an excited tautomer (S1 0 ) capable of undergoing intramolecular
conversion to S0 0 and reverting rapidly to the original oxamide structure. The excitation energy is
dissipated. No luminescence was detected in dilute solutions of 22. Fluorescence with a lifetime less
than 5 ns was measured for 22 at 77 K. Experiments in polyethylene show that oxanilides retard the
Norrish II reaction. Neither triplet quenching nor free radical scavenging have been observed with OA in
model experiments [65].
Cyanoacrylate 29 is able to transform the absorbed radiation energy into heat because a potentially
large number of vibration modes. Formation of charge-separated species after photoexcitation was
speculated [99] (Scheme 4).

3.2.2. UVA durability loss


High durability of photostabilizers is required to guarantee long-term outdoor performance of organic
coatings. During years of research and industrial development, the structures of UVA have been

Scheme 4.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1291

optimized to reach high inherent chemical and physical durability during the long-term outdoor protec-
tion of coatings [58]. However, photostabilizers are very reactive compounds in degradation of polymers
and the long-term protection of the coating proceeds at the expense of stabilizer chemistry. This has been
categorized as sacrificial consumption meaning a sum of chemical transformations resulting from
effective protection of polymers during processing or curing and final outdoor application. Besides
the sacrificial consumption, the stabilizers are depleted by interaction with reactive species in the
polymer matrix, by direct or sensitized photolysis or photo-oxidation, oxidizing and acid atmospheric
pollutants or adsorption on fillers and pigments. These processes have been categorized as depleting
consumption [58].
Both sacrificial and depleting consumption reduces the concentration of the active form of the
stabilizer below a level no more capable of protecting the degrading polymer in prolonged weathering.
It is evident that the inherent photochemical resistance of UVA and oxidative sensitivity of the coating
are principal factors affecting the rate of UVA degradation underconditions of the exposure [58,99101]
(radiation source and humidity in particular).
The activity mechanism of UVA has been based on reversible physical processes (see Section
3.2.1). However, irreversible chemical and physical transformations reducing UVA durability and
performance cannot be fully avoided in oxidizing coatings [100,102] affecting the ultimate lifetime
of coatings. Attention paid to UVA durability is understandable especially in durable coatings used
as protective films (clearcoats) in weatherable application of photosensitive plastics such as poly-
carbonate or poly(vinyl chloride) or in two-layer automotive coatings. The screening effect of
stabilized coatings lasts for a limited period of time only due to a progressive photoaging of UVA
[99,102]. Several mechanisms of photostabilizer efficiency loss in coatings have been considered
[9,101,103].

Interruption of IMHB by formation of intermolecular H-bonds in matrices with H-acceptor moieties


[67,73,88].
Photochemical transformations via radiation-assisted free-radical attack or direct photolysis of
excited states of UVA having weakened IMHB [88,100]. Photostability of UVA drops in the presence
of oxygen and with increased photo-oxidative sensitivity of the polymer matrix.
Physical losses during cure or high-temperature weathering and reduction of effective concentration
due to migration of UVA from the clearcoat to the basecoat or plastic substrate as checked by UV
spectroscopy [104].
Interruption of IMHB by formation of intermolecular H-bonds in matrices with H-acceptor moieties
[67,73,88].

If follows that high levels of UVA are required to stabilize coatings not only because of the Lambert
Beer law (see Section 3.2.1) but also to compensate high rates of UVA loss. In commercial application of
coatings, all depleting pathways may operate, depending on the environment. Matrix-assisted photo-
oxidation of UVA predominates in polymers with very low resistance to degradation. In relatively
unreactive matrices, such as PMMA, the polarity promoting interruption of IMHB is probably respon-
sible for the depletion. The principal modes of loss of UVA durability are outlined in detail in the
following sections.

3.2.2.1. Formation of intermolecular hydrogen bridges. The intramolecular H-tunneling mechanism of


1292 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

phenolic UVA is vulnerable in matrices of solvents, coatings and engineering polymers having construc-
tion units with basic moieties capable of intermolecular H-binding with acid phenolic hydoxy groups
[67,73,84,8789,94,99,101,103]. Intermolecular H-bonds also arise in deeply oxidized polymers with
high levels of hydroxy or carbonyl groups [10]. Intermolecular H-complexes with 2-hydroxybenzophe-
none 31 and 3-hydroxyflavone 32 are examples [71].

This depletion mode of UVA activity is more common for phenolic UVA with sterically unhindered
hydroxy groups. This was confirmed by formation of intermolecular H-bonds in ground state of BP in
ethanolic solution. A low efficiency of BP in easily oxidizable polymers (such as rubber-modified
polystyrene or acrylonitrilebutadienestyrene polymers) or in polypropylene processed under very
severe conditions is a consequence [105].
Addition of the second hydroxy group into position 2 0 in BP (such as in 5) improves UVA durability in
PMMA film. Probably, the second group HO enhances the probability of IMHB in BP [66]. The light-
fastness of 2-hydroxybenzophenones can be also slightly improved by substitution of position 3 by
methyl group, shielding the IMHB. A practical exploitation of this structural modification in BP is no
longer attractive due to commercial availability of BT and TA-based UVA. However, the unfavorable
effect of polar matrices with acceptor behavior also affects the activity of 2-(2-hydoxy-phenyl)benzo-
triazoles (see Scheme 2) creating competitive intermolecular H-tunneling [67]. The long-wavelength
absorption band of 10a at 350 nm can be observed only in molecules with intact IMHB, i.e. in non-polar
(inert) environment [84,88]. The intensity of this band decreases in polar hydroxylic solvents, in
mixtures of polar with non-polar solvents, and even in the presence of traces of moisture in hydrocarbon
solvents [67,73,76,88,106]. This is because of the partial transformation of the planar BT molecule with
IMHB to a distorted structure characteristic of intermoleculer H-bonding. BT can act as UVA only if the
IMHB remains intact [85,106]. Phosphorescence and moderately red shifted fluorescence at l max
400 nm are observed [76,89] in BT 10a indicates destruction of IMHB. This is attributed to inter-
molecularly bonded species, which cannot be involved in ESIPT [67,73,76,94]. A solvent-assisted
excited state intermolecular proton transfer was considered for this system, characterizing an exchange
of the proton with the acceptor medium [73,88].
It is now well established that the efficiency of the photophysical processes in phenolic UVA depends
strongly on the stabilizer structure and polarity of the microenvironment [88]. A comparison of the
performance in non-polar plastics, such as polystyrene, with that in PMMA reveals a lower effect in the
latter polymer, due to the intermolecular H-bonding to carbonyl groups [99].
Analysis of absorption spectra, fluorescence yields and UVA lifetimes has allowed to assess the
proportion between the planar and non-planar forms of BT 10a in various polymers [94]. It was found
that 11 mol% of the non-planar structure is present in poly(vinyl acetate), 86% in poly(N-vinylpyrroli-
done). This indicates a reduced performance of BT in these polymers in comparison with non-polar
polymers.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1293

The disruption of the IMHB by the polar environment, loss of planarity due to rotation around the N2
C1 bond [67,88] and excitation of the intermolecularly bonded UVA with sites of the environment
presumably more basic than N1 or N3 of the triazine ring generates a low-lying (n,p ) reactive triplet
[85,88] involved in triggering the photochemical depletion of the UVA itself and sensitizing polymer
degradation via an excited solutesolvent complex 33 taking place in excited state intermolecular proton
transfer [67,88,103].

Comparison of the lifetimes of intramolecularly bonded and planar BT and that of the intermolecu-
larly bonded BT is indicative [107]. The first excited singlet state of the planar form is very short living
(less than 200 fs). That of the non-planar BT form lives over 1.4 ns. A difference in the lightfastness was
also observed between the planar and non-planar form of polymer-bound BT moiety in poly[2-(2-
hydroxyphenyl-4-methacryloyloxyphenyl)benzotriazole-co-methyl methacrylate]. The non-planar
form is slowly photolyzed. The process was monitored by the decrease in intensity of the 390 nm
band after long-term irradiation of the copolymer. The planar form characteristic of absorption at
550 nm was not affected by irradiation [67,108].
A proper molecular architecture can favor the IMHB and planar structure of the BT-type UVA
[67,88]. Steric hindrance of the phenolic hydroxyl by a bulky substituent, such as tertiary alkyl or
a,a-dimethylbenzyl in 10d, 10f10h, shields the IMHB effectively, diminishes the danger of its break-
ing and prevents the formation of photodegrading intermolecularly bonded species. Properly substituted
BT with sterically shielded IMHB and relevant planarity have exceptional photophysical properties even
in polar matrices [63,106].

3.2.2.2. Photoassisted free-radical depletion. Phenolic UVA are subjected to a slow radical-induced
photoassisted degradation in the course of years of outdoor exposure of the stabilizer-doped polymer
matrix [100,101,109]. The changes were observed in various thermosetting (TS) acrylic melamine
clearcoats or UV-cured acrylic coatings from the onset of weathering [99,110,111]. UVA lose in clear-
coats their protective strength after a prolonged exposure [5,100,104]. As a result, the penetrating light
causes both surface and bulk oxidation; the latter may trigger a catastrophic failure, including loss of
mechanical properties, and adverse adhesion of clearcoats to underlying basecoat [37,45].
The lightfastness of UVA in oxidizing matrices depends on their photosensitivity under experimental
conditions [99,109,112]. Chemical (dark) and photochemically triggered processes participate mostly in
concerted pathways in UVA losses in acrylate clearcoats [99,100,104]. Photolysis and photo-oxidation
are preferentially effective in radiation exposed surface layers.
All the major classes of commercially available UVA photodegrade in coatings [99,100]. The kinetics
of phototriggered losses of UVA in solid matrices depends on the photochemistry of the coating itself.
Any generalization is difficult because of lack of data. Zero-order kinetics was reported for solid films
and coatings absorbing at short wavelengths (350 nm), first-order in matrices absorbing the less
harmful long-wavelength light [99,100].
A progressive loss of the stabilizing effect of BT 10h in photodegraded UV-cured polyurethaneacrylate
1294 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

coating was shown to follow the first-order exponential kinetic law [109,111] (Fig. 7). The time intervals
(t 50) required to cut by half the original concentration of UVA were in the range of 140 h for the rather
unstable BP and 4180 h for the lightfast BT 10h. The values of the lightfastness increased with the
UVA concentration and coating thickness [109].
The rates of the phototriggered losses of various UVA were compared in various polymeric films:
PMMA, silicone hardcoat and UV-cured acrylate [100,101,111]. In PMMA, BP 4 (R i-C8H17) was the
least stable while TA 17a degraded at about one half the rate. The rating of UVA losses in PMMA show
a series BP cyanoacrylate, OA BT TA. Almost an analogous rating was found for UV-cured
acrylate; however, the decay rates were considerably faster than in PMMA. Different rating of losses was
found in the silicone hardcoat matrix: TA cyanoacrylate, OA, BT BP. The data indicate the neces-
sity of testing UVA in each particular coating.
Scavenging of alkylperoxyls or alkoxyls by phenolic UVA was mentioned as a process potentially
contributing to the integral performance of UVA in polymer stabilization by the chain-breaking anti-
oxidant effect [113]. This scavenging accounts, however, for sacrificial oxidation of the phenolic moiety
of UVA resulting in phenoxyls, quinone methides and substituted cyclohexadienones [22,96,114]. A
mechanism characteristic of phenolic antioxidants is involved accounting for chemical disruption of
ESIPT. Phenolic moieties are strongly vulnerable under photo-oxidative conditions, including photo-
sensitization [22,115].
Chemical reaction of phenolic UVA with POO arising in the photooxidized matrix was confirmed by
isolation of some transformation products. It was reported [105] that BP 4 (R C8H17) oxidizes via the
relevant phenoxyls and unstable alkylperoxycyclohexadienone 34 into a derivative of benzophenone 35
and xanthone 36. Oxidation of BP 4 is faster in easily oxidizing polymers, in polymers with enhanced
levels of photosensitizing chromophores (carbonyls) or in polymers that have been subjected to severe
processing.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1295

Fig. 7. Loss profile of UVA upon QUV-B aging in a polyurethaneacrylate coating: (1) 1 (R C8H17); (2) 23; (3) 11 (R Cl);
(4) 10h; (5) 10h 61 (R CH3). (Reproduced, with permission from Decker C, Zahouilly K, Polym Mater Sci Eng 1993;68:70
[109].)

Photoassisted oxidation of BP with POO proceeds faster than the dark oxidation. Abstraction of the
phenolic hydrogen from the excited BP is considered the primary step [116]. This is especially effective in
polar matrices where the intermolecularly bonded excited singlet state has a much longer lifetime [67,88].
Photolysis of BP 4 resulting in benzoic acid 37 as the principal identifiable product was reported [99,100].
Various trace products arise together with 37 by the oxidative coupling of the phenolic fragment of BP.
Model studies were performed with BT [117]. Alkylperoxycyclohexadienone 38 and related quinone
39 were isolated in the dark chemical reaction taking place at high levels of tert-butylperoxyls generated
in benzene solution from tert-butylhydroperoxide in the presence of cobalt(II) acetylacetonate.
Initiation of the oxidative tranformation-depletion of phenolic UVA, including the lightfast BT is based on
principles of chemistry of phenolics [22,96,114]. The oxidative transformation is preferred at sites with
localized high concentrations of alkylperoxyls and hydroperoxides in the polymer matrix [22,88,96].
Hindered BT-type UVA are not readily attacked by low levels of alkylperoxyls in aliphatic hydro-
carbons [110]. The importance of concentration of oxidizing species accounting for participation of
oxygen is evident from model measurements with 13 (R t-C8H17) in heptane solution [97] and in
polycarbonate. Photodegradation in heptane was approximately 2.5 times faster than in polycarbonate. It
was suggested that the result account for photo-oxidation, because the permeability of oxygen in heptane
is about 1000 times higher in comparison with polycarbonate.
A proper substitution of the phenolic moiety is beneficial to regeneration of the IMHB in phenolic BT
after disruption of IMHB by oxidation with alkylperoxyls [58,96]. The regeneration mechanism is
exploited in systems substituted by propionate esters in the para position to the phenolic hydroxy
group, such as in UVA 10i10l, generally 41 (BT benzotriazole moiety). The regeneration pathway
of IMHB from quinone methide 42 formed from 41 involves [96] an intramolecular rearrangement
resulting in phenolic derivative of cinnamic acid 43.
1296 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Scheme 5.

A more recent study states [37,110,118] that abstraction of hydrogens from the phenolic hydroxyl in
BT takes place most probably in the excited state and in the presence of free radicals, rather than in the
ground state. Experiments were performed with BT 10a, 10g and 10h. Their photolysis was studied in
solvents differing in polarity under radiation with fluorescent UV-A lamp at a air temperature of 40C
and a dew point of 25C, in the presence of free radical initiators, 2,2 0 -azobis(isobutyronitrile) and
cumylhydroperoxide [118]. No free-radical attack was observed in non-polar solvent (cyclohexane)
in thermally initiated homolysis of initiators. This is consistent with the strong influence of the
IMHB and planar structure, protecting the BT from depletion (see Section 3.2.1). Photoreaction in
ethanolic solution in the absence of oxygen results in adducts formed by the solvent or initiator and
the phenolic part of BT. In oxygen atmosphere, free triazole 40 arises by photolysis of the N2C1 bond
of the non-planar BT molecule [110,118]. Formation of oxidation products from the phenolic fragment
can be anticipated.
To explain the photolysis-photo-oxidation of the open form of BT, we consider oxidation of the
excited solutesolvent complex 33 into mesomeric radical 45 and its reaction with alkylperoxy or
hydroperoxy radicals XOO accounting for unstable peroxycyclohexadienone 46 and its photolysis to
benzotriazole 40 and products of oxidative coupling of phenolic fragments (Scheme 5).
Suggestions from photophysical measurements on the decisive role of polarity of the environment on
photostability of phenolic UVA were confirmed by photochemical experiments. It is assumed that the
open form of BT (with destroyed IMHB) behaves as a hindered phenolic antioxidant, including the
related tranformations [22,114].
Results with 10h in two acrylicmelamine clearcoats differing in photo-oxidation rate, development
of carbonyl levels and sustained alkylhydroperoxides reveal [46,110] that the UVA is destroyed by
photo-oxidation and not by pure photolysis. This suggests a photoassisted participation of oxygen-
centered radicals. The UVA loss was lower in the more oxidation-resistant clearcoat [99,119]. Degrada-
tion of the excited non-planar form of UVA is suggested [110,119,120].
Analysis of results obtained in various clearcoats reveals [5,99,110,120] that binder type, method of
crosslinking and irradiation conditions during weathering influence durability of UVA. Matrices with
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1297

Fig. 8. Effect of binder type on the stability of UV absorber 10h (1.5%). Dry film thickness 30 mm, exposure: UVA-340 lamp,
40C dry conditions. (Reproduced with permission from Light stabilization of paints, 1997. 1997 C.R. Vincentz Verlag,
Hannover [5].)

high degradation rates promote UVA losses more efficiently than durable acrylics [99]. Durability of 10h
in various matrices is an example [5] (Fig. 8). The UVA stability monitored as the time period to 50%
loss and increases from HS acrylatemelamine clarcoat (3000 h) to one-pack polyurethane coating
(5300 h) and fluoropolymermelamine binder (6800 h). The loss of UVA evidently increases
with the increasing rate of formation of free radicals. This is also shown in another series of experiments
with stabilized coatings at two different exposures, using UV-B and UV-A radiation sources (Fig. 9). It is
evident at the same time [109,120,121] that addition of radical deactivating HAS 61 (R CH3) extends,
under both the radiation conditions, durability of BT 10i.
The experiments with BT in solutions and acrylate melamine clearcoats reveal that a combination of
radiation, oxygen-centered free radicals and moderate polarity of the coating matrix are depleting factors
affecting UVA durability in outdoor applications [5,67,88,99,100,109,110,118,119]. Photoassisted loss
of UVA in clearcoats of automotive coatings begins at the surface and extends more deeply with

Fig. 9. Effect of radiation source and addition of a free radical scavenger 61 (R CH3) on the stability of UV absorber 10i.
Clearcoat: a two-pack polyurethane (acrylic polyol/isocyanate), cure 30 min at 130C, dry film thickness 35 mm, irradiation at
40C, dry conditions. (Reproduced with permission from Light stabilization of paints, 1997. 1997 C.R. Vincentz Verlag,
Hannover [5].)
1298 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

continuos irradiation [99]. This increases the danger of phototriggered damages in underlying basecoat
accounting for delamination. Loss of UVA durability due to coating polarity proceeds in the bulk of the
matrix. Combination of both processes is detrimental for UVA-doped coatings after 35 years of
outdoor exposure [99,100].

3.2.2.3. Non-phenolic UVA. Losses of non-phenolic UVA were observed in weathered coatings
[100,101]. There is, however, a lack of information on their depletion modes. Oxanilides are resistant
against attack of free radicals and hydroperoxides. Their sacrificial consumption was postulated to be a
consequence of quenching of excited carbonyls [65]. No changes in UV/visible spectra were detected in
22 (R H, R 1 C2H5) after prolonged (100 h) irradiation at 303 and 313 nm in nitrogen and air atmo-
sphere, even in the presence of a benzophenone sensitizer.
Mechanism of the loss of durability of 2-cyanoacrylates is not known at all. Interference of alkylhy-
droperoxy radicals with charge-separated species in cyanoacrylates was noted recently, without mechan-
istic details [100,101].

3.2.2.4. Interferences with coating components and curing systems. Coating formulations consist of
various components enabling the technological process and mandatory to achieve expected protective
and decorative properties and durability [1,4]. Serious consequences such as inadequate curing, coating
discoloration or loss of UVA durability account for interaction of UVA with some coating components.
Interaction of UVA with catalysts. Phenolic part of BP, BT or TA-based UVA tends to react with
strongly basic catalysts (such as tertiary amines), basic components of the binder (e.g. amino-functio-
nalized acrylates) and/or metal-based catalysts (such as zinc 2-ethylhexanoate or aluminium tris(acety-
lacetonate) [5,10,22,63]. Transformation of the phenolic hydroxy group participating in IMHB into an
anion interrupts ESIPT, results in intensive yellowing [7] and promotes oxidaton of the phenolic moiety
into cyclohexadiene structures [10,22,58,96]. Discoloring complexes can be formed also from phenolic
UVA and metallic driers, such as cobalt naphthenate or 2-ethylhexanoate [7] [6162]. Oxanilides show a
slightly lower tendency to discoloration than phenolic UVA. The discoloration is undesirable especially
in clearcoats over a white basecoat.
Strong yellowing was evidenced after thermal aging at room temperature in two-pack epoxy varnish
based on an amino- and epoxy-functionalized acrylate [5] doped with 10h, a blend 10j with 10k, and 4
(R C12H25). Triazine 17d was found to be non-interacting with the components of HS thermosetting
acrylate (TSA) clearcoat (Fig. 10). Triazine 17d caused only negligible color changes in 40 mm acry-
latemelamine clearcoat prepared by catalysis with 2% aluminium tris(acetylacetonate) and baking at
130C over a white basecoat. On the contrary, a high discoloration (measured by yellowness index, YI)
resulted from BT 10h and BP 4 (R C12H25, all UVA were used at 2% level).
Interactions of UVA with photoinitiators. This problem arises in UV-cured coatings due to an undesirable
overlapping of photochemical properties of UVA and photoinitiators (PI) during curing [123]. Conventional
UVA interfere with the curing process by a preferential absorption of light in comparison with PI (this results
in inadequate curing [111]). However, UVA loose at the same time their durability due to photolysis
during photocuring. This forced a search of specific improvement of high lightfastness of UVA in UV-
curable coatings with long-term resistance to sunlight and selection of suitable PI [5,109,124].
The UVA lightfastness can be substantially increased by addition of HAS (e.g. by a combination of
BT 10h with HAS 61 (R CH3)). A proper selection of PI, UVA and their combination with HAS was
beneficial to durable UV-curable polyurethaneacrylate-based protective coatings [100,102,109,112].
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1299

Fig. 10. Effect of phenolic UV absorber on the yellowness index (YI) of a 30 mm two-pack epoxy clearcoat based on
aminofunctional/epoxyfunctional acrylate over a white basecoat. Curing: 30 min at 60C; W, unstabilized, X 2% 17d; O,
2% (blend) 10j 10k, K, 10h. (Reproduced with permission from farbe + lack 1994;100:919. 1994 C.R. Vincentz Verlag,
Hannover [169].)

This suggests that the UVA loss during curing is partly due to radicals arising in the cured matrix. The
radicals can be scavenged by HAS (see Section 3.3).
Structures and activity mechanisms of PI for UV-cured coatings were well described [5,103,124,125].
Some polymer-bound PI moieties were introduced [126]. Problems, however, arise, in a simultaneous
application of PI and UVA.
In spite of the competition between UVA and PI for light absorption, suitable combinations for UV-
curing of coatings can be selected [124]. It was found that the highest polymerization rate during UV
curing must not necessarily be at the absorption maximum of PI. A properly through-cure is achieved in
systems where the light necessary for photolysis of PI is not completely suppressed by UVA. Very good
results were obtained using aryl-a-hydroxyketones 47 and 48 or their blends with bis(acylphosphine) 49.
Hydroxyketones affect beneficially surface hardness of the coating, the phosphine improves through-
cure [5,124]. Fig. 11 shows pendulum hardness values of UV-cured aliphatic epoxyacrylate/aliphatic
urethaneacrylate/tripropyleneglycol diacrylate clearcoat, inititated with 3% 48 or with a 3:1 blend of
47 and 49. Excellent stability of the coating was obtained with TA 17d. The selection of a proper system
UVA-PI is influenced by the color of the basecoat.

The durability problems and interference of UVA with UV-curing process can be solved by applica-
tion of a delayed-action UVA-precursor. Aryl benzoates 52 belong to this category [58]. Unfortunately,
most of them turn yellow after photo-Fries rearrangement to hydroxybenzophenone 53.
Derivatives of BP, BT or TA having hydroxy functions blocked with a photolysable group are
effective [127]. These compounds have a reduced absorption in the region 350400 nm, do not interfere
seriously with curing and lose stepwise the protective group by photolysis. UVA 50 (X H,
R 1 SO2 C6H5, R 2 i-C8H17, e 20 000 dm 3 mol 1 cm 1 at 345 nm) is formed by photolysis
1300 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 11. Effect of UV absorbers 10i and 17d and photoinitiators 4749 on the curing of a UV-curable 40 mm clearcoat based on
aliphatic epoxy acrylate/aliphatic urethane acrylate/tripropylene glycol diacrylate, curing: 2 80 W cm 1. Pre-baked silver
metallic basecoat. (Reproduced with permission from Light stabilization of paints, 1997. 1997 C.R. Vincentz Verlag,
Hannover [5].)

of 50 (X SO2 C6H5, R 1 H, R 2 i-C8H17) at 300 nm. Photolysis is exploited also in UVA-progen-


istors having hydroxy groups blocked by acetylation. Compounds photolysable during the cure, such as
benzotriazoles 50 (X COCH3, R 1, R 2 H) [128] or 50 (X COCH3, R 1 H, R 2 CH3) [102,129]
and diacetate 51 [130] are effective latent UVA for acrylate coatings.
Photolysis of the acetyl group in 50 (X COCH3) during the UV cure of acrylates with a UV-B 313 or
UV-A 340 lamp yields a free hydroxy group 50 (X H).The process can be monitored [7,112] by
appearance of the absorption band at 340 nm.

Other photolysable BT-based compounds, e.g. 50 (X CONHC6H5, R 1 H, R 2 CH3) or 50


(X P(O)(aryl)2, R 1 H, R 2 CH3) [131] were considered UVA progenistors as well. Compounds
with free ortho-position to the blocked hydroxyl function undergo photo-Fries rearrangement during
curing [127]. The photolysis of the progenistor continues in the cured coating under solar radiation. As a
consequence, the stabilizing power in the coating increases with time. A concentration gradient of the in
situ formed UVA has been formed from the coating surface to the lower layers. The distribution profile
of the formed UVA provides the strongest protection just in the surface layer, most severely attacked by
light [102,112]. Moreover, the delayed-action acetylated stabilizer has a much less detrimental effect on
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1301

the curing speed than conventional UVA [102]. This reveals that in situ generation of UVA is a
technique particularly well suited for stabilization of UV-cured coatings.

3.3. Photoantioxidants

Scavenging of chain propagating alkylperoxyls and deactivation of hydroperoxides is a very effective


approach to photostabilization of organic materials. Selection of suitable additives is limited to
compounds having high inherent photostability. Conventional phenolic antioxidants (strong alkylper-
oxyl scavengers) generally suffer from a low photostability [22,115]. 4-Hydroxybenzoates 52 are
1302 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

considered photostable phenols [58]. The aromatic ester (52, R 2,4-di-tert-butylphenyl) is a free-
radical scavenger with contributing UV absorbing effect due to hydroxybenzophenone 53 arising
from 52 by photo-Fries rearrangement [22]. The aliphatic ester 52 (R C16H33) performs as a photo-
stable antioxidant only (no photo-Fries rearrangement is possible).
Aromatic amines, another class of strong scavengers of alkylperoxyls, are not suitable photoantioxidants
for coatings due to their photolability and strong discoloration. Fortunately, invention and development of
HAS solved the problem of photoantioxidants and was really a breakthrough in polymer stabilization [23].
Recent optimization of the inherent chemical efficiency and physical persistence substantially broadened
application of HAS in coatings. The common application levels of HAS are generally 11.5%.
Examples indicate the diversity in structures of HAS. Most commercialized additives contain the 4-
substituted 2,2,6,6-tetramethylpiperidine moiety differing in molecular weight, substitution on the imino
group and the number of piperidine moieties in the molecule. Mononuclear 5458, dinuclear 6067 or
polynuclear piperidines 6872, and piperazinones 59 and 73, the only commercialized HAS with
heterocyclic moiety other than that of piperidine, are examples of commercially available or developing
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1303

LMW and HMW HAS. Potentially, HAS structures can be derived from various other heterocyclic
amines, such as derivatives of hindered oxazolidine 74. Imidazolidin-4-one 75, hexahydropyrimidine 76
or decahydroquinoxalin-2-one 77.
Because of importance of HAS for polymer photostabilization, effective additives combining HAS
and UVA moieties in one molecule (see Section 3.4) and oligomeric and polymer-bound HAS (see
Section 3.5) were synthesized.
The success of HAS as photostabilizers in a large number of commodity or engineering
polymers and coatings is impressive. Secondary (NH) and tertiary (NR) HAS can be used univer-
sally [23]. O-alkylhydroxylamines (NOR) and acylated HAS (NCOCH3) having lower basicity were
introduced for application in coatings containing acidic components or contaminants [23]. The optimum
effectiveness in coating has been obtained using combination of UVA with HAS [5,58,121,132135].

3.3.1. Activity mechanism of HAS


Hundreds of research papers have tried to explain the intricate activity mechanism of HAS [23]
1304 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

outperforming other classes of stabilizers by their complex activity in processes induced by solar and
high-energy radiations and thermal oxidation at temperatures below 120C. Due to their commercial
importance, individual pathways of the HAS mechanism have been continuously re-evaluated. Most
mechanistic studies were performed in model (liquid) substrates and polyolefins and can be exploited
also for coatings [23,132,136].
Generally accepted mechanistic pathways are based on chemical transformations of HAS with oxidiz-
ing species arising in photo-oxidized polymer matrix and identification of nitroxide (NO ), the key
free-radical intermediate accounting for the photoantioxidant performance [23]. Molecular architecture
of HAS excludes involvement of absorption of UV radiation or quenching of excited carbonyls in the
activity mechanism. Some rather inefficient quenching of singlet oxygen was reported. Activity of HAS
is independent of the thickness of the oxidizing polymer film.
Mechanistic analyses in non-polar polymers revealed a possibility of formation of charge-transfer
complexes [HASO2], [polymer O2], and/or [HAS ROO ] at the very early stages of the HAS
stabilization mechanism [62]. Eqs. (44)(46) describe the primary sacrificial process accounting for
antioxidatively active NO from secondary HAS (NH) by oxidation with alkylhydroperoxide, alkyl-
peroxyl and peroxy acid [23]. HAS react slowly with simple alkylhydroperoxides (ROOH) and more
efficiently in sites with accumulated ROOH. Complexing of relatively basic NH with acid ROOH
[NH HOOR] effectively increases probability of reaction (44) in non-polar polyolefins [137]. A
similar acidobasic complex cannot be expected in more polar coating matrices.
2 NH ROOH ! NOHNOR ROHHOH 44

ROO
NOH; NOR ! NO products 45

2 NH 3RCOOOH ! 2 NO 3RCOOH H2 O 46

Nitroxides arise in the primary step also from tertiary HAS (NR, e.g. 5558, 6164, 69, 70, R CH3),
N-acyl HAS (e.g. 56, 57, R COCH3) and O-alkylhydroxylamine 61 (R i-C8H17) [23]. Tertiary HAS
are oxidized with alklylperoxyls, alkylhydroperoxides or singlet oxygen to salts of secondary HAS with
formic acid (78), Eq. (47), N-acyl HAS with alkylhydroperoxides to salts of secondary HAS with acetic
acid (79), Eq. (48), and O-alkylhydroxylamine-type HAS directly into nitroxide according to general Eq.
(45) [138]. Salts of NH with weak organic acids have not a restrictive effect on oxidative transforma-
tion to nitroxides [139], i.e. 78 or 79 generate NO .
ROO ; ROOH;1 O2
NCH3 ! N H2 OCOH 47
78

NCOCH3 ROOH ! NOR N H2 OCOCH3 Products 48


79

The mechanism yielding nitroxides from NOR is consistent with regenerative mechanism of HAS
[23,140,141] (Scheme 6). According to this mechanism, alkylperoxyls or acylperoxyls oxidize NOR
to NO and are themselves reduced into alcohols or carboxyl acids, and carbonyl species via an
intermediary complex 80. The earlier mechanistic proposals included formation of peroxidic species
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1305

Scheme 6.

which were never experimentally verified.

NH; NCH3 ; NOR; NCOCH3 ! NO


oxidation
49
Generation of nitroxides is indicative for the performance of various classes of HAS Eq. (49).
Nitroxides are generally considered the key intermediates responsible for photoantioxidant activity of
HAS in deactivation of most species arising in a photo-oxidizing polymer matrix, i.e. of P , PO , POO ,
POOH, PC(O)OO or PC(O)OOH. Trapping of macroalkyls P at rates approximately 10 9 M 1 s 1
accounting for O-alkylhydroxylamines NOP, Eq. (50) is considered a specific and the most effective
contribution of HAS to the photoantioxidant mechanism [142]. HAS are unique in this respect because
of general shortage of alkyl scavenger applicable in polymers [58]. Deactivation of alkyls in the
oxidizing environment is a complicated task. It must be effective to compete with alkyl self-termination
reactions (disproportionation and recombination accounting for branching, crosslinking or chain scis-
sion) and oxidation of alkyls [23].
O-Alkylhydroxylamines formed according to Eqs. (45) and (50) thermolyse in an inert atmosphere to
hydroxylamine and an olefinic species (Eq. (51)). Nitroxide is formed [23] in air atmosphere, Eq. (52).
NO P ! NOP 50

D; N2
NOP ! NOH Olefin 51

D; O2
NOP ! NO POO 52
Formation of O-alkylhydroxylamines was unambiguously identified in the aged polyolefin matrix doped
with HAS and has to be considered as a reservoir of nitroxides [141] due to an easy regeneration
according to Scheme 6. It is understandable that trapping of alkylperoxyls by O-alkylhydroxylamines
representing both the antioxidant chain-breaking process and principal regenerative process of nitrox-
ides, Eq. (53), has attracted so much attention [23]. Eq. (53) shows at the same time the unique
mechanistic feature combining cyclic scavenging of alkyl and alkylperoxyl radicals with regeneration
of nitroxides, the latter originally formed in sacrificial transformation of HAS. The regeneration of
nitroxides is not dependent on the structure of the parent HAS (Eq. (53), X H, R, OR, COR).
There is no analogy to such complementary mechanistic effects in any other class of polymer stabilizers.
P POO P
NX ! NO ! NOP ! NO ! etc:
oxidation
53
1306 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 12. Photo-deacetylation of HAS 57 (R COCH3) (W,B) to HAS 57 (R H) (X, A) in toluene solution; nitrogen atmo-
sphere (W,X), air atmosphere (B,A). (Reproduced with permission from farbe + lack 1993;99:25 1993 C.R. Vincentz Verlag,
Hannover [143].)

The crucial role of nitroxides as photoantioxidant species is generally accepted [23,136,140,143]. They
are formed by the sacrificial processes from all types of commercially available HAS, Eq. (49). The rate
of their generation from various N-substituted HAS is of practical importance. Experimental results
show [143] that the conversion of NH or NCH3 proceeds faster than that of NCOCH3. Conse-
quently, acyl-HAS have a lower performance in comparison with NH or NH3. They can be, however,
used in formulations containing acid components. The rate of the sacrificial transformation of O-alkyl-
hydroxylamines into nitroxides is comparable with that of secondary and tertiary amines. Consequently,
they have very beneficial properties because of their inertness to the acid environment.
A series of experiments was performed to follow the photodeacylation (photolysis) of 56 and
57 (R COCH3) and 93 into the relevant secondary HAS, according to Eq. (54). The loss of the
acyl group was used as a step determining the rate of NO formation. Toluene solutions and HS TSA
doped with the three acyl-HAS were irradiated [143] in air or under nitrogen with a lamp Haereus TQ
150.
hn
NCOCH3 ! NH products 54

A low rate of photolysis was observed in both substrates for 56 (R COCH3). The low conversion to 56
(R H) was explained using simulated model calculation by the MNDO method. This revealed that the
spiro-HAS 56 exists in a H-bridged boat form resistant to photolysis for steric reasons. Secondary
amines were effectively formed from the onset of photolysis of 57 (R COCH3) and 93. Fig. 12
exemplifies deacylation of 57 in toluene solution in air or nitrogen [143]. Photolysis of acyl-HAS in
HS TSA after a short-term and long-term irradiation is evident from Fig. 13. The NH arising in Eq.
(54) is a source of NO . It follows from the deacylation experiments that fast 57 (R COCH3) and 93
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1307

Fig. 13. Photo-deacetylation of acyl-HAS in HS TSA: A 93, W 57 (R COCH3), K 56 (R COCH3). (Reproduced with
permission from farbe + lack 1993;99:25 1993 C.R. Vincentz Verlag, Hannover [143].)

are able to deactivate alkylperoxyls and alkylhydroperoxides faster from the onset of coating weathering
than 56 (R COCH3). On the other hand, the slow HAS 56 (R COCH3) is considered as a reservoir
for the long-term stability, providing nitroxides in the later phases of photodegradation. A combination
of slow and fast HAS is a beneficial approach to stabilization of coatings by non-basic acyl-HAS
[143].
A potentially negative role of nitroxides in polymeric matrices with weak RH bonds should be
mentioned. Nitroxides are prone to abstract H atoms, Eq. (55). The reaction results in hydroxylamine
NOH (an antioxidant species) and macroalkyls participating in chain initiation and propagation.
Degradation of the polymer matrix is thus promoted.
NO RH ! NOH R 55
Induction of degradation by H-abstraction according to Eq. (55) was mentioned in acrylate coatings
[9,34,144] and considered as a specific chain-initiating step and one of reasons why HAS do not
completely suppress formation of alkylhydroperoxides [9].

3.3.2. Adverse effects and depletion of HAS


The regenerative mechanism (Scheme 6) is characteristic of excellent long-term HAS performance. In
spite of the effective regeneration, some depleting processes limit the lifetime of HAS either in their
original or transformed form [23]. Understanding of the adverse processes was used for structural
improvements. Some depleting factors resulting in irreversible transformation cannot be eliminated.
Processes limiting the effectiveness of HAS are outlined.
Polymer-bound ketones (products of polymer photo-oxidation or components of construction units)
photolyse by the Norrish reaction to acyl radicals C (O). The latter recombine with nitroxides to N-
acyloxy derivative 81 which cannot be converted into nitroxides and account for irreversible loss of HAS
activity. Photolysis of the ester group in position 4 of the piperidine molecule of 61 accounts for a
1308 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

volatile 4-oxo derivative 82 which can be further degraded to open-chain products. A low decrease in the
ester absorption of HAS 61 was also observed [139] in HAS-doped polyolefins exposed to mineral acids
(HCl, HNO3).
HAS themselves do not absorb solar UV radiation. However, the derived nitroxides absorb light in the
range from 300 to 320 nm and form excited (p,p ) states. As a consequence, the piperidine ring
fragments and undergoes further transformations. Open-chain nitroso compound 83 n 1 and nitro
compound 83 n 2 were identified [145] together with nitrogen-free fragments 84 and 85.
Ozone, the tropospheric oxidizing pollutant involved in photo-oxidation of polymers (see Section
2.1.2.1) reacts with secondary and tertiary amines. An intermediary adduct [NHO3] was proposed
[146] in ozonation of 2,2,6,6-tetramethylpiperidine accounting for 2,6-dimethyl-6-nitroheptan-2-ol 86.
A complex participation of ozone in processes encountering outdoor degradation of HAS-doped poly-
olefins was reported recently [20]. Ozonation of crosslinked binders containing HAS was not studied,
can be, however, anticipated.
Relatively basic secondary HAS (NH) and tertiary HAS (NCH3) are applicable in most coatings
free of acid compounds [7,147]. Difficulties arise in contact with acid species, such as acid polymeriza-
tion catalysts, hydrochloric acid arising from dehydrochlorination of plastic substrate, poly(vinyl chlor-
ide) in particular, halogenated flame retardants or pesticides, various acid products from sulfur-
containing additives [7,23] and acid atmospheric pollutants (see Section 2.1.2.2).
Strong mineral acids (hydrochloric and nitric acids in particular; sulfurous acids is somewhat less
detrimental) and organic acids, e.g. p-toluenesulfonic acid deactivate basic HAS by protonation [139].
Salts such as 87 (Subst H or CH3, X mineral acid rest) are no more prone to form nitroxides, the
principal intermediate in the cyclic HAS mechanism. Weak carboxylic acids forming salts 78 and 79 do
not deplete HAS.
A dual problem arises with acid catalyzed resins such as HS TSA coatings, with acidic pigments and
paints based on acid monomers. Secondary and tertiary HAS as well as derived nitroxides bind acid cure
catalysts (p-toluenesulfonic acid, dodecylbenzenesulfonic acid) and prevent, consequently, hardening of
coatings [7,61,143,147] besides depletion of the HAS activity [145].
The fact that nitroxides are basic species as well is mostly neglected in discussions of the mechanism
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1309

of non-basic (non-interacting)HAS, i.e. acyl-HAS and O-alkylhydroxylamines. These HAS are resis-
tant to acidic species until their sacrificial transformation into nitroxides. The latter as well hydroxyla-
mine NOH, another suggested product of sacrificial transformation of HAS, form salts 88 and 89 with
strong acids [23,139]. Due to a very low concentration of nitroxides in the coating matrix, the danger of
acid deactivation is substantially lower than that with NH or NCH3.
Salts 8789 are partly water-soluble. This implies that acid exposure of HAS-doped films decreases
HAS longevity by a repeated (rain) water extraction [139]. Moreover, the salts 8789 precipitate in the
coating and form seeds.
Basic HAS accelerate crosslinking of resins catalyzed by basic catalysts and shorten irregularly their
pot life [5]. Addition of a basic HAS to water-borne paints may exhibit stability problems because of an
interference with the neutralization step performed with amines. Owing to their basicity, secondary and
tertiary HAS catalyze hydrolysis of polycarbonate [148]. This fact should be considered in polycarbo-
nate articles coated by HAS-doped protective films where long-term migration from the crosslinked
resin to polycarbonate is possible.
To improve the resistance of HAS to acid environment and to increase application possibilities in
coatings, structural modifications on the amino group lowering basicity were performed. O-Alkylhy-
droxylamine 61 (R OC8H17) [61,147] and acylamines 56 or 57 (R COCH3) [143] were developed
and introduced commercially. The non-interacting HAS are suitable for stabilization of acid-catalyzed
clearcoats and for polymers doped with halogenated flame retardants. The influence of structural modi-
fication of the amino group on pKa values is evident [5,23,149].
Substitution on the amino group of the piperidine cycle influences also the speed of nitroxide forma-
tion (see Section 3.3.1). The improved resistance of acyl-HAS, the least basic HAS, to acid environment
is compensated by a lower rate of formation of NO in comparison with O-alkylhydroxylamines [150],
another non-interacting HAS. The latter may be considered as particularly beneficial photoantioxidant in
acid environment.

HAS pKa HAS pKa


Secondary (NH) 8.09.7 Hydroxylamine (NOH) 4.36.1
Tertiary (NCH3) 7.59.2 O-alkylhydroxyl-amine, (NOR) 4.24.4
Tertiary oligomeric
(NCH2 CH2O ) 6.5 N-acyl (NCOCH3) 2.0
Nitroxide (NO ) 7.49.6

3.4. Bifunctional photostabilizers

The necessity of application of combinations of two stabilizers acting by different mechanisms (UVA
and HAS) to achieve an integral photoprotection of coatings was the driving forces for development of
bifunctional stabilizers. An autosynergism has been envisaged [58]. Optimized architecture of a mole-
cule having two different stabilizing centers suggests a balanced contribution of individual functional
moieties to the integral effect. Both active moieties remain in the polymer matrix in the closest proximity
and assure an intramolecular cooperation whenever possible. By comparison of various systems, it was
1310 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

observed that mostly only one of the functions provides a principal contribution to the final effect,
depending on the attacking environment. The second function has a complementary or supporting effect.
Various bifunctional stabilizers have been reported in the patent literature. Some were selected for the
commercial development. The activity mechanisms have been mostly extrapolated from those of the
relevant monofunctional additives.
HAS do not possess any UV absorbing potency. This drawback can be eliminated by built-in UVA
moieties. Compounds such as 2-hydroxy-4-{4-[(2,2,6,6-tetramethylpiperidin-4-yl)amino]butoxy}ben-
zophenone 90 [151], a HAS containing 2-(2-hydroxyphenyl)benzotriazole moiety 91 [152], this struc-
ture exploits the optimum substitution of the phenolic moiety assuring planar structure of the BT part
and potential regeneration of the IMHB after consumption of the phenolic part), triazine moiety 92 [153]
or oxamide moiety 9395 [154] are examples of bifunctional photostabilizers designed for coatings.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1311

Deactivation of the actinic solar radiation by the UVA part is intramolecularly complemented by the
free-radical scavenging photoantioxidant effect of the HAS moiety. Bifunctional stabilizer containg
HAS and hindered phenolic ring 96 has also found some application in coatings [5].

3.5. Physically persistent photostabilizers

Besides chemical and photochemical depletion, physical losses by volatility and leaching into envir-
onment limit exploitation of the inherent chemical efficiency of photostabilizers applied in coatings
exposed to harsh conditions [155]. Some of these losses are due to a low compatibility with the polymer
1312 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

matrix [7]. The problem of the physical persistence can be solved by application of stabilizers with
increased molecular weights [155,156]. Additives having molecular weights higher than 500 (arbitrarily
classified as HMW), oligomers with a proper molecular architecture and molecular weight between
3500 and 5000 and polymer-bound stabilizers generally fulfil requirements for physically persistent
high-performing stabilizers. They are less likely to evaporate at elevated temperatures of curing.
HAS such as 5658 or 61 and UVA 6, 10h, 10j, 10k, 10k, 13, 14, 17b, 17d or 17f are examples
of sufficiently persistent photostabilizers in crosslinked systems. Compounds 97105 are examples of
oligomeric HAS. They can be used very effectively in combination with migrating LMW and HMW
HAS [23].
Polymeric stabilizers prepared by polyreactions and reactions on polymers [155,156] are promising
for applications in coatings. Most common structures of LMW UVA (BT in particular) and HAS with
optimum inherent chemical efficiency can be functionalized with polymerizable groups. Polymeric
UVA and HAS are beneficial whenever thin films consisting of photodegrading polymers have to retain
durability over an outdoor long-time service period. Various monomers functionalized with UVA
moieties 106114 and HAS moiety 115119 were synthesized (only typical structures were selected
as examples) and can be copolymerized with conventional acrylic monomers commonly used in binders.
For example, monomer 110 was copolymerized with alkyl vinyl ether, reactive fluoropolymers and a
mixture of acrylic monomers to form a durable coating [157]. Benzotriazole 109 was used in a mixture
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1313

of acrylates and urethaneacrylate for UV-curable weather-resistant coatings for protection of a poly-
carbonate substrate [158]. UVA, such as 107, 109, 111 or 113, copolymerizable with hydrogel-forming
acrylates, are used in very specific applications, such as non-extractable UV blocks in contact lenses,
considered as protective clearcoats over the eye cornea [159].
Homopolymerization accompanies copolymerization of functionalized and conventional vinyl
1314 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

monomers. Consequently, photostabilizers are present in the coating matrix in the form of copolymers,
homopolymers and residual functionalized monomers [155]. Homopolymers with accumulated photo-
stabilizing moieties have mostly the lowest performance (as compared at an equimolar level) due to
heterogeneous accumulation of UVA moieties in the polymer matrix [155]. Compatibility problems may
arise during blending of copolymerized UVA with PMMA. Tailor-made functionalized copolymers
must be prepared with respect to the conventional polymer system expected to be stabilized.
Incorporation of functionalized monomers by free-radical (thermo)initiated reactive grafting, a
method successfully exploited in polyolefins [105] is not applicable to crosslinked coatings. However,
a successful incorporation of photostabilizers into the coating network was performed by condensation
or addition reaction with reactive functionalized monomers. For example, 4-[bis(2-hydroxyethyl)a-
mino]-2,2,6,6-tetramethylpiperidine 120 was recommended as stabilizer for a permanent chemical
incorporation into pigmented thermosetting acrylic coatings, alkyd-melamine coatings or two-compo-
nent polyurethane coatings [160]. HAS 121 n 0 3; R C4-14H9-29] containing two or more reactive
aliphatic isocyanate groups forms a built-in stabilizing system for lacquers and participates (at n
1 3 in reactive crosslinking [161].

Benzotriazole 10j used in a blend with the diester 10k can react with isocyanate-based components of
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1315

Scheme 7.

the binder and form [134] a polymer-bound UVA 122 (Scheme 7). Benzotriazole 112, having in the ester
group a rest of polyunsaturated acid reacts with Urushi lacquers and forms a polymer-bound stabilizer [162].
Application of a photograftable HAS 118 in acrylates is worth considering. This HAS can be added in
its monomeric form to a binder, a photografting proceeding during UV-cure or outdoor exposure [163].
A UV barrier coating containing an in-chain BP moiety 124 is formed [155] by a stepwise photo-Fries
rearrangement of polyester 123. Yellowing accompanies the rearrangement.

Photophysical mechanistic features characteristic of monomeric photostabilizers are also considered


for copolymeric and polymer-bound systems. Measurements with monomeric UVA 111 and 113 and
their copolymers with styrene, methyl methacrylate or methacrylic acid confirmed that IMHB, the key
requirement for the ESIPT mechanism, is also characteristic of copolymerized phenolic UVA
[74,76,95,129,164]. A strongly Stokes-shifted, temperature-dependent proton transfer fluorescence
evidenced the presence of IMHB. The radiationless deactivation process is concluded [74,76] to origin
from the proton-transferred state S1 0 .
The results of decay-time measurements of proton-transferred emissions indicate that the slowest
rates are displayed in the pure monomeric BT 111 (70 ps) [84], followed by its copolymerized species
[74] (3040 ps). The latter values are comparable with those measured for physical mixtures of
dispersed BT 10a in polystyrene [107].
The photochemical mechanism in polymer-bound UVA is more complex than in pure monomeric
UVA studied under comparable conditions. It was reported [129] that the lifetime of the excited
tautomeric form (S1 0 ) of poly[2-(2-hydroxy-5-vinylphenyl)benzotriazole-co-methyl methacrylate] rank-
ing about 1 10 9 scan ensures the outdoor photostability of the BT moiety up to 20 years. The rapid
1316 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

physical decay pathway of the excited BT chromophore in poly[2-(2-hydroxy-5-vinylphenyl)benzotria-


zole-co-styrene] effectively competes with the deleterious photochemical pathway in the polymer back-
bone and photostabilizes the copolymer [73,95,164].
In poly[2-(2-hydroxy-5-vinylphenyl)benzotriazole-co-methyl methacrylate], the electronic energy
transfer from the excited UVA species to the polymer chain resulting in initiation of oxidative degrada-
tion may potentially compete with ESIPT. The UVA chromophore thus promotes chain scission or
crosslinking of the matrix. Moreover, intermolecular H-bonds can be formed between pendant UVA
species and H-acceptors in the same or neighboring macromolecular chain, both prolonging the reactive
excited state and resulting in shortened UVA durability [95]. A low level of polymer-bound BT was
considered to be efficient for quenching of the mobile exciton with an additional quenching resulting
from long-range energy transfer from the excited traps to the BT moiety [165].
Copolymers of functionalized BP with methyl methacrylate or butyl acrylate were synthesized
[155,156] and used for photophysical studies. The decay mechanism of excited poly(2-hydroxy-3-
allyl-4,4 0 -dimethoxybenzophenone-co-methyl methacrylate) 125 (irradiated with 297400 nm light)
was reported to involve more pathways, besides ESIPT [129]. A small non-zero triplet yield was
postulated and accounting for photodegradation of 125. The triplet yield was considerably high in
ethanolic solution of 125 promoting intermolecular H-bonding (see Section 3.2.2.1). IR measurements
at 3580 cm 1 indicated photo-oxidation of PMMA segments in the copolymer 125 due to photocatalysis
by the excited BP moiety. This resulted in crosslinking via free radicals formed on tertiary CH bonds of
the methyl methacrylate unit of 125. On the contrary, non-functionalized PMMA degraded much more
slowly by chain scission. This indicates that the mechanism of PMMA photodegradation is strongly
influenced by the bound-in UVA chromophore.
Studies of temperature dependence of fluorescence and photostability of poly[2-(2-hydroxy-4-meth-
acryloyloxyphenyl)benzotriazole-co-methyl methacrylate] 126 have extended the insight into the
excited relaxation mechanism occurring in the polymer-bound BT [67,108]. The temperature-insensitive
short-wavelength band having l max at 390 nm was assigned to an emission arising from a low amount of
the non-planar form of bonded BT with disrupted IMHB. The red-shifted and temperature-sensitive
emission at 550 nm arises from the fluorescence of the proton-transferred BT-tautomer. The room
temperature fluorescence lifetimes of the 390 and 550 nm bands have been determined to be 895 and
45 ps, respectively [108]. The rapid radiationless depletion of the excited proton-transferred BT-tauto-
mer accounts for the high photostability of the polymer-bound planar form of BT. Some photodegrada-
tion of the non-planar form was observed after a long-term irradiation of the copolymer 126. The
depletion process was manifested by a decrease in the intensity of the 390 nm band. The planar form
important for ESIPT was not changed during irradiation.

4. Performance of photostabilizers in coatings

Coatings are expected to increase outdoor durability of many objects. However, they are themselves
attacked by weathering. Their inherent resistance is not sufficient to withstand the harsh conditions. A
proper selection of stabilizers and their combinations fitting the type of binder, catalyst or pigments
guarantees the highest possible long-term performance under expected application conditions. General
requirements of end-use customers include increased stability at lower stabilizer concentration levels.
To protect coatings against cracking, gloss loss, color change, chalking, blistering or delamination,
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1317

two classes of photostabilizers are used: UVA (see Section 3.2) and HAS (see Section 3.3.). UVA having
structures of BP, BT, TA and OA protect coatings against delamination by filtering the harmful radiation
initiating formation of free radicals. HAS reduce loss of gloss, yellowing and cracking by radical
scavenging accounting for a substantially reduced levels of hydroperoxides and carbonyl groups
[5,132,136,143]. Basicity is the principal chemical property affecting application of HAS. The effect
of HAS can be fully exploited particularly in fast degrading coatings yielding high levels of alkyl-
peroxyls [100,109]. HAS also has a positive influence on the lightfastness of UVA. Based on the
knowledge of the mechanism of UVA depletion (see Section 3.2.2), the protective action of HAS is
considered to result from suppression of the alkylperoxyl level in the coating accounting for light-
assisted radical depletion of IMHB in phenolic UVA. On the contrary, UVA prevents the photolysis
of the HAS-developed nitroxide [100,102,166]. The effect of both stabilizer types is complementary, but
it is difficult to speak about synergism. For example [166], the addition of 1% of BT prolongs the time
period of nitroxide permanence in acrylic coating containing 25% styrene and doped with 1% 61
(R H) (Fig. 14).
Concentration levels of photostabilizers in coatings are rather high. Application ranges/preferred
concentrations (in %) are for UVA 1-3/1-2, HAS 0.5-2/1-1.5, HAS-UVA blends 0.5-2/1 of each [5,7]
depending on the film thickness, photosensitivity of the binder and pigment.
Application of photostabilizers in coatings is accompanied by various peculiarities arising from the
complicated structure (one-coat, two-coat systems) and composition (different chemistry, solvents) of
coatings. Owing to the validity of the LambertBeer law (see Section 3.2.1), UVA functioning by
absorption of UV light have only a low and strongly concentration dependent efficiency in thin coating
films. Typically, 1.52% of effective BT must be added to prevent light transmission in a 40 mm
clearcoat. To maintain the same level of protection, 2.53% UVA must be used in 20 mm clearcoats
[7]. This at the same time shows insufficient protection of surface layers of coatings. Data obtained with
two-layer coatings show complicated relations involved in the system [5,7,9]. UVA are generally added
1318 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 14. Formation of nitroxide radicals from HAS 61 (R H, 1%). In (K) coating A (an acrylic copolymer with 25% styrene),
(W) coating B (styrene-free acrylic copolymer) and (X) coating A doped with 1% benzotriazole-based UVA, during artificial
weathering. (Reprinted from Eur Polym J 1983;19:11, 1983, with permission from Elsevier Science [166].)

only to the clearcoat, and are not able fully protect its surface. On the contrary, the lower layers (base-
coat, primer, substrate) are mostly well protected against penetrating radiation due to the filtering effect
of the full depth of the stabilized clearcoat. The weathering behavior of the whole two-coat system is
influenced by the pigment [167]. Depending on the reflecting power of the pigmented basecoat, the
active UV radiation passes several times the clearcoat. This increases the intensity of the incident light
[5]. Moreover, pigments in the basecoat absorb heat energy and can develop temperatures up to 100C.
Thermodegradation is thus promoted and contributes to the integral effect of weathering.
Benzotriazoles and triazines outperform other classes of UVA in various clearcoats over pigmented
basecoats [9,69] and prevent fading or color changes of pigmented basecoats [7]. HS TSA clearcoats
doped with 12% 10a and 12% HAS 61 (R H) protected effectively acrylic melamine basecoat
pigmented with Chromophthal Orange 2G. Particularly high levels of UVA must be added to maintain
the desired shade and color intensity with photosensitive pigments [7]. Depending on the pigment, a
combination of two UVA complementing each other in spectral characteristics (e.g. 1.7% of 10i with
0.8% of 17d) blended with 1% of 61 is beneficial [5].
Physical factors influence the final effect of UVA as well. Some UVA migrate from the clearcoat to
the basecoat [168] and lower the light screening power of the former. Losses of UVA during exposure of
clearcoat in Weatherometer are dependent on the initial concentration of UVA. To limit migration of
UVA to the primer, it is desirable to cure the primer prior application of basecoatclearcoat system [9].
The migration problem can be diminished using UVA with increased molecular weights and/or forming
covalent bonds with components of the binder, such as 122 (see Section 3.5).
UVA have to be non-interacting with other coating components [121] (see Section 3.2.2.4). Phenolic
UVA differ in their resistance to reactions with aminofunctional acrylates, basic and metal-containing
catalysts accounting for discoloration of the coating [5]. Phenolic triazine 17d is very resistant being a
solution to the yellowing problem as shown by changes of YI in amino-epoxy-functionalized acrylate
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1319

Fig. 15. Gloss retention in 40 mm two-pack aminofunctional/epoxyfunctional acrylate clearcoat over silver metallic basecoat after
three years Florida exposure, curing 30 min at 60C: W, unstabilized, A, 2% 17d 1% 61 (R OC8H17), K, 2% 23 1% 61
(R OC8H17). (Reproduced with permission from farbe + lack 1994;100:919. 1994 C.R. Vincentz Verlag, Hannover [169].)

clearcoat doped with various UVA [169]. Excellent performance of 17d (2%) combined with 1% 61
(R OC8H17) after three years of weathering in Florida [169] is shown in Fig. 15 and compared with the
performance of a blend of OA 23 and HAS 61 (R OC8H17).
Various other less common phenolic UVA, such as xanthone 9, substituted 2-(2-hydroxyphenyl)pyr-
imidine 15 or silylated triazine 17g used at concentration of 2% stabilize effectively acrylate resins
against weathering.
Lighfastness, thermostability and chemical inertness of oxanilide 23 were exploited in stabilization of
two-layer TSA coatings [133]. Excellent gloss retention was obtained with 1% 23 and 1% 58 (R H)
after three years of Florida exposure. The residual content of 23 after natural weathering was 96%.
Triazinyloxanilide 25 was recommended for acrylic melamine clearcoat [170].

Fig. 16. Effect of additional stabilization with HAS 61 R CH3 on the gloss retention of a UV absorber-doped HS two-pack
polyurethane clearcoat over waterborne silver metallic basecoat, cure 30 min at 90C, 30 months Florida exposure. (Repro-
duced with permission from Light stabilization of paints, 1997. 1997 C.R. Vincentz Verlag, Hannover [5].)
1320 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 17. Accelerated weathering (UV-NOC, UVB-313, 50C) up to cracking of an electron beam-cured clearcoat over silver
metallich basecoat: 1 unstabilized, 2 1.5% 23 1% 61 (R CH3), 3 1.5% blend (10j 10k) 1% 61 (R CH3). (Repro-
duced with permission from Polym Paint Colour J 1992;182(4311):406. 1992 C.R. Vincentz Verlag, Hannover [171].)

Combinations of UVA with HAS account for excellent protection of solvent or water-based,, liquid
solvent-free, UV-cured or powdered clearcoats in two-coat automotive coatings [5,7,111] exposed to
high temperatures during stoving and long-term weathering in tropical regions especially with dark
pigmentation. Modern systems are based on BT or TA-type UVA and HAS with properly selected
basicity. Triazines 17a and 17d and benzotriazoles 10h to 10k are suitable for two pack polyurethane
clearcoats [5]. In most tested systems combinations with HAS impaired a better protection than UVA or
HAS used as single additives. This is shown on combinations 1% 61 (R CH3) with BT 10i, TA 17d or
OA 23 (Fig. 16).
One percent of the blend 10j/10k with 1% 61 (R CH3) was excellent in the protection of two-layer
metallic polyurethane clearcoat over the pigmented basecoat during the three years of Florida weath-
ering. Benzotriazole blend 10j/10k was more efficient than OA 23 [171] (Fig. 17).
Excellent protection of clearcoat was obtained by a 1:1 combination of 17a with photograftable HAS
118 (total concentration 23%). No cracks and loss of gloss were observed after 1000 h exposure at
313 nm or in a long-term Florida test.
A blend of UVA and HAS was also recommended for water-borne coatings [5]. Application of BT 10i
and TA 17d or their combinations with 61 (R CH3) is an effective stabilization of acrylate melamine or
two-pack urethane water-borne clearcoats.
Photostabilizers having melting points between 70 and 110C are suitable for powder coatings
[5]. Stabilizers with higher melting points can be used if they are well soluble in the formed
resin. BT 10h, TA 17a and 17c used at 2% level with 1% 69, 96, or 97 were recommended for
an acrylic polyol-blocked or carboxy-epoxy-functionalized acrylate clearcoats over the water-borne
silver metallic basecoat.
Photostabilization of mono-coat finishes and coatings for non-automotive applications with high
concern to protect various substrates has its particular problems. Pigments present in the surface layer
of the coating are covered by a very thin polymer film and are attacked by radiation more intensively
than pigments in basecoats. Degradation is manifested by surface dullness and chalking. UVA can
protect effectively mono-layer coatings only at rather low pigment-binder ratios, e.g. 1:8. Blends
of 17a or 17c with HAS 61 (R CH3), 61 (R OC8H17) or 96 are suitable for epoxy coatings,
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1321

Fig. 18. Effect of UV absorber and HAS on the weathering resistance (Florida exposure) of an one-coat finish. Binder: two-pack
polyurethane (acrylic polyol/isocyanate) pigmented with Pigment Blue 28, cure 45 min at 80C, substrate: epoxy primer; W,
unstabilized; A, 2% 10h, 2% 61 (R CH3), delamination after 27 months; X, 1% 10h 61 (R CH3). (Reproduced with
permission from Light stabilization of paints, 1997. 1997 C.R. Vincentz Verlag, Hannover [5].)

thermoplastic acrylates, polyester-triglycidyl isocyanurate or long-oil alkyd coatings [5,167]. Polyur-


ethaneacrylate coating was effectively stabilized with 1% BT 10h and 1% HAS 61 (R CH3) (Fig.
18). Both photostabilizers used as single additives, 2% each, did not sufficiently protect the coating.
It was reported [172] that BT 10a encapsulated in PMMA microbeads increases stability of linseed
oil-based coating. At the same time, the lightfastness of 10a was prolonged. PMMA is oil-insoluble,
resistant to oxidation products of the oil and protects UVA from direct attack by alkylperoxyls generated
in the natural oil. The dispersed bead-dosed material (1030 wt% of particles doped with 10% 10a)
outperforms 10a dissolved in the oil and assures long-term durability of the coating.
The rapid development of UV-curable coatings expects that the requirements on their outdoor durabil-
ity will be fulfilled. This includes a proper selection of photostabilizers and photoinitiators. In this
respect, application of HAS is without problems. They do not absorb light in the UV region and do
not interfere with PI. A major concern is the competition for UV absorption between UVA and PI. A
strong absorption by UVA effective up to 380 nm results in lower residual doses available for PI and in

Fig. 19. Effect of photostabilization on the loss of the carbamate group upon QUV aging at 40C of an aliphatic polyurethane
network: W, no additive; K, 0.5% 10h, A, 0.5% 61 (R CH3); X 0.5% 10h 0.5% 61 (R CH3). (Reprinted with permission
from J Polym Sci, Polym Chem Ed 1991;29:739. 1991 [Wiley-Liss, a subsidiary of] John Wiley & Sons, Inc. [111].)
1322 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

incomplete cure on coating-substrate interfaces. An increased consumption of UVA during the UV-cure
is another phenomenon [111,119]. The aim has been to select the most efficient and non-inter-
acting system for the through-cure and durability performance [29]. The present understanding of
problems of photostabilization of UV-cured coatings is based on extensive research
[29,100,102,103,109,111,119,122,123,128,169,171,173]. One of the fundamental problems is the over-
lapping of the absorption spectra of UVA and PI [173]. Application of acylphosphine-based PI 49
absorbing light up to 430 nm in combination with hydroxyketones 47 and 48 solves the problem of
UV-curing of various paint systems doped with UVA [29,174,175]. Benzotriazoles 10h or 10i and
triazine 17a are suitable UVA [5,173175]. For example as in Ref. [111], UV-cured aliphatic polyur-
ethaneacrylate coating was effectively stabilized with BT 10h and HAS 61 (R CH3) (Fig. 19).
Another example is the effective outdoor protection of a 40 mm layer with 22.5% of 17d and 1% 61
(R CH3 or OC8H17) or 97. The real-time IR spectroscopy revealed a good through-cure of the resin in
the presence of photostabilizers and improved resistance to weathering in outdoor application [29].
An excellent protection against yellowing during weathering was attained in a very thin (13 mm)
complex aliphatic polyester/triurethane triacrylatehexanediol diacrylate coating system cured with 3%
of a blend 75% 47/25 % 49 in the presence of 3% TA 17d and 1% non-interacting HAS 61 (R OC8H17)
on polycarbonate substrate [5]. The same combination assured long-term resistance (at 2000 h QUV-A
weathering) of UV-cured polyurethaneacrylate coating [176]. A good through-cure and outdoor
permanence was obtained [173,174] in UV-curable powder clearcoats using 3% of a blend of PI 47/
49 and a blend of stabilizers 2% 17a with 1% bifunctional HAS 96.
Some specific mechanistic features and their relevance to performance were observed in HAS-doped
coatings. It is rather difficult to check the durability and transformations of HAS in coatings. Only the
originally added HAS and developed nitroxides can be determined after extraction by spectroscopic
methods. The activity was studied in details with 61 (R H) in acrylatemelamine and acrylate
urethane coatings [34,36,44,136,144,177179]: it reveals some differences from stabilization of poly-
olefins where association with alkylhydroperoxides promotes effective deactivation of hydroperoxides
[180]. An analogous association is not expected in polar coatings. As a consequence, higher concentra-
tion of HAS (12%) have to be used in coatings [5,135] in comparison with polyolefins (0.11%).
Moreover, HAS have in polyolefins a pronounced heat stabilization effect chracterized by the absence of
induction period [23]. An analogous thermostabilization effect was not checked in coatings. The differ-
ences in HAS mechanism in various polymeric substrates are understandable as the HAS chemistry is
strongly dependent on the degradation chemistry of the substrate under exposure conditions. The
differences observed in HAS effectiveness in various coatings can be explained by differences in
formation rates and accumulation of nitroxides [136]. Therefore, data on the net rate of generation of
nitroxides at very early stages of photodegradation of particular coatings exposed to well-defined
radiation conditions, the time to reach the maximum concentration of nitroxides and their decay at
long exposure time after the concentration turn-out point are of top importance [181]. It was reported
[9,136] that HAS ensure activity in coatings at ca 1 10 6 1 10 7 mol g 1 concentration level of
nitroxides. On extended exposure, this concentration slowly changes due to participation of nitroxides in
the cyclic mechanism, Eq. (53) and H-abstraction from the substrate, Eq. (55). The presence of nitroxide
derived from 61 (R H) was detected from the start of formation of peroxidic species in the acrylic-
melamine clearcoat [136]. Changes in the nitroxide level were monitored [144,177,181]. The concen-
tration reached a maximum and slowly decreased (Fig. 20).
The active consumption of HAS accounting for nitroxides is dependent on the rate of formation of
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1323

Fig. 20. Nitroxide level versus exposure time for the acrylatemelamine coating exposed at a dew point of 25C. The photo-
initiation rates of free radicals were: W, 12 10 8 mol g 1 min 1, A, 7 10 8 mol g 1 min 1 and K, 1.4
10 8 mol g 1 min 1. The level of HAS 61 (R H) was 2%. (Reprinted from Polym Degrad Stab 1986;14:53, 1986, with
permission from Elsevier Science [181].)

oxidizing species in the coating, the energy of the radiation source and environmental humidity [178]. A
faster consumption of HAS is consequently observed in more oxidation-sensitive coatings and with light
sources having a higher portion of UV-B radiation.
In accordance with the inherent sensitivity of the polymer matrix to photodegradation, the lifetime of
HAS was found shorter in acrylatemelamine resins than in acrylateisocyanate-based two-pack poly-
urethane coating [182]. Fig. 21 shows that enough HAS 61 (R CH3) (determined as nitroxide) was
available in an acrylateisocyanate based clearcoat doped with a blend of HAS with 2% of a mixture of
BT 10j and 10k and exposed over silver metallic basecoat for five years in Florida exposure [5]. For
comparison, the same coating was exposed to a UVB-313 lamp and a less energetic xenon weatherom-
eter. Under the latter conditions, less alkyl radicals are formed in the coating. As a consequence, less
nitroxides are consumed for their scavenging and the level of nitroxides remains high. Alkylradical
scavenging by nitroxides accounts for polymer-bound species NO-coating. The latter is converted to
nitroxide according to general Scheme 6.
Humidity of the environment influences the rate of generation rate nitroxides from 61 (R H) in
acrylatemelamine coatings. This phenomenon was not observed in urethaneacrylate coatings. The
authors suggest [144,181] formation of formaldehyde during hydrolysis of melamine crosslinks and its
involvement in the process forming strong oxidizing agents, Eq. (56).

HCHO ! HC O ! HCOOO ! HCOOOH


oxidation RH
56

Oxidation of HAS with peroxy acids and participation of peroxy acids and peroxy acyls in the regen-
eration cycle (Scheme 6) was reported [141].
Impact on generation of alkylhydroperoxides and carbonyl species in coatings [44,136,181] and
1324 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 21. Long-term effect of 1% HAS 61 (R CH3) in a 40 mm acrylate/isocyanate-based two-pack polyurethane clearcoat on
glass substrate; exposure: 1 QUV fluorescent lamp UVB-313, 2 Xenon weatherometer, 3 Florida natural weathering (on top of a
silver metallic basecoat, the clearcoat also contains 2% 10j 10k. (Reproduced with permission from Light stabilization of
paints, 1997. 1997 C.R. Vincentz Verlag, Hannover [5].)

checking nitroxide level by EPR spectroscopy were elucidated in accordance with the photoantioxidant
effect of HAS. Addition of HAS to urethaneacrylate coating reduced levels of alkylhydroperoxides and
prolonged the lifetime of the coating [9,44,182]. The influence of the presence of 2% 61 (R H) on the
level of alkylhydroperoxides arising in two various acrylic-melamine coatings (differing in the photo-
sensitivity of the acrylate copolymers end-groups; binder A contains photoreactive groups [136]) during
weathering in ATLAS Ci 35 Xenotest [44,136] is shown in Fig. 22. HAS is rather ineffective in

Fig. 22. Hydroperoxide concentration versus near-ambient exposure time in melamine crosslinked acrylate A (with photo-
sensitive end groups, W, X) and acrylate N (A, B). Open symbols: unstabilized coatings, filled symbols: stabilized with 2%
HAS 61 (R H). (Reprinted from Polym Degrad Stab 1993;41:323, 1993, with permission from Elsevier Science [44].)
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1325

Fig. 23. Nitroxide level for an acrylatemelamine coating with photosensitive end groups doped with equimolar amounts of
HAS W, 61 (R H), A, 61 (R CH3); K 56 (R COCH3), and X 98; [standard: 2% 61 (R H)], under near-ambient exposure
conditions. (Reproduced, with permission from Gerlock J.L., Bauer D.R., Mielewski D.F., 11th International Conference on
Advances in Stabilization and Controlled Degradation of Polymers, Luzern, 2426 May 1989, p. 25 [136].)

oxidation-resistant acrylatemelamine coating with low rate of radical formation [182]. The oxidation
rates are, however, effectively reduced in photo-oxidation-sensitive melamine-croslinked binder A. The
nitroxide formation and decay curves in the latter monitored in the presence of various HAS (at
concentrations equimolar to 2% of 61, R H) and using a rather severe exposure (UV-B radiation)
are shown in Fig. 23. Differences between HAS are evident. Secondary HAS 61 (R H) and tertiary
HAS 61 (R CH3) build-in rapidly nitroxides to comparable peak levels from which decay starts at
comparable rates. Acyl-HAS 56 (R COCH3) and oligomeric tertiary HAS 98 are much less
effective in reaching the peak concentration of nitroxides. The order of rating of HAS according
to the formation-decay curves of nitroxides was comparable with the HAS rating based on build-
up curves of carbonyls [136]. A comparison of various HAS in photosensitive acrylate binder
exposed to borosilicate glass-filtered xenon arc light (near-ambient conditions) revealed a simi-
lar rating according to the nitroxide formation rate, peak concentration and turn-on-time curves to that in
Fig. 23. Concentration of nitroxide formed from 61 (R H or CH3) exceeds the level 1 10 6 mol g 1
after a few hundreds of hours. Concentration of nitroxides from 56 (R COCH3) and 98 did not reach
this level even after 8000 h exposure, i.e. in a time period when it is too late to prevent the major
oxidation damages of coatings.
Substitution of the piperidine nitrogen affects the application potential of HAS in coatings. Secondary
and tertiary HAS are applicable in coatings without any acid contaminants. O-Alkylhydroxylamines and
acyl-HAS were developed for coatings cured with acid catalysts [61,147,183]. Model experiments
proved differences in rates of nitroxide generation from various N-substituted HAS according to Eq.
(53). If the conversion rate is low, the HAS is not sufficiently effective, particularly of the early stages of
exposure of the coating [143]. High rates of conversion to nitroxides were reported [9,136,147] for
NH, NR and NOR, moderate rates for tertiary oligomeric species NCH2CH2OC(O) , and low
rates for NCOCH3. Examples show differences in performance between various categories of HAS.
Common tertiary HAS 61 (R CH3) generates nitroxides in UV-B exposed acrylatemelamine clear-
coat more quickly than the non-basic acyl-HAS 56 (R COCH3) [9,132,136,147]. The practice
1326 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 24. Effect of N-substitution of the piperidine ring on the stabilizing effect of HAS (measured up to crack formation) in HS
TSA clearcoat over silver metallic basecoat, cure 30 min at 90C. Stabilization: W, unstabilized, A, 1% 56 (R COCH3) 2%
(blend) 10j 10k; X, 2% 56 (R COCH3) 2% (blend) 10j 10k; B, 1% 61 (R OC8H17) 2% (blend) 10j 10k (no
cracking observed after 24 months). (Reproduced with permission from Light stabilization of paints, 1997. 1997 C.R.
Vincentz Verlag, Hannover [5].)

confirms a lower performance of NCOCH3 in coating protection in comparison with NH or NCH3.


The disadvantage of acyl-HAS was eliminated by introduction of another non-interacting (non-basic)
HAS, O-alkylhydroxylamine 61 (R OC8H17) [5,147]. Its conversion to nitroxide is fast as that of 61
(R CH3). Both HAS assure excellent protection of HS TSA clearcoats under Florida exposure. Besides
formation rates of nitroxides, interactions of HAS with the components of the coating formulation must
be considered. Fast secondary and tertiary HAS fail in stabilization of acid-cured coatings, in spite of
high rates of nitroxide formation. Application of non-basic HAS such as 56 (R COCH3), 57
(R COCH3) or 61 (R OC8H17) eliminates curing problems in acid-catalyzed HS TSA [5,147,184]
in spite of a lower rate of generation of nitroxides from acyl-HAS. For example, mechanical properties
of baked HS TSA catalyzed with 0.5% p-toluenesulfonic acid checked by pendulum hardness fully
failed with 1% of amines 61 (R H, or CH3). On the other hand, a satisfactory through-cure was
obtained with 1% 61 (R OC8H17), 56 (R COCH3) and 57 (R COCH3). Non-interacting HAS
protected the coating against photocracking and loss of gloss as well [5,143,183].
There is a difference in performance between non-interacting HAS due to differences in their rate of
nitroxide formation. Tests in 40 mm HS TSA acid-catalyzed clearcoat revealed [183] that 61
(R OC8H17) outperformed acyl-HAS 56 (R COCH3) (both HAS were used in combination with
BT 10j and 10k) [5,143,183,185] (Fig. 24).
O-Alkylhydroxylamines provide a photoantioxidant optimum for acid-catalyzed coatings [5,147,
183]. The acid cure is not retarded and the durability of the coatings is excellent. Moreover, there are
no interactions with acid pigments and transition metals. O-Alkylhydroxylamine 61 (R OC8H17) used
in combination with 2% BT 10i also reduces the rate of formation of melamine-melamine crosslinks in
HS acrylatemelamine clearcoats catalyzed with dodecylbenzenesulfonic acid. Testing the durability of
HS acrylate melamine clearcoat applied over a silver metallic basecoat during QUV exposure reveals
that the non-stabilized clearcoat cracks after 2000 h after a rapid loss of gloss. Addition of 2% BT 10i
extends the service life to about 3000 h. A combination of 1% of HAS 56 (R COCH3) with 2% of 10i
postponed the time-to-cracking but the gloss continued to decrease. A blend of 2% of 10i with 1% of 61
(R OC8H17) resulted in a very slow loss of gloss and the time-to-cracking was postponed to 5000 h
[147]. These data indicate a better performance of NOR in comparison with NCOCH3.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1327

Fig. 25. Concentration of HAS 61 (R H, open symbols) and its mono-N-methyl derivative (closed symbols) versus exposure
time in the melamineacrylate coating exposed at 60C to fluorescent tube FS-20 at a dew point of 25C. (Reprinted from
Polym Degrad Stab 1986;14:17, 1986, with permission from Elsevier Science [178].)

Some specific features in HAS chemistry were revealed in acrylatemelamine coatings


[144,178,179,181]. A drop in concentration of 61 (R H) and formation of its monomethylated analog
was observed during weathering at a 25% dew point. About 15% of the added secondary amine was
methylated with formaldehyde released by hydrolysis of the melamine crosslink [178,181]. Formation of
the monomethylated HAS is exemplified in Fig. 25. Methylation of secondary HAS with formaldehyde
released during weathering of cured melamine-containing automotive coating was also confirmed by
transformation of HAS 56 (R H) into 56 (R CH3) [179].
Excellent durability of coatings based on application of a combination of HAS with UVA is fully
exploited in practice. This promoted development of stabilizers containing both moieties in one mole-
cule. Bifunctional stabilizers (see Section 3.4) were effectively used for stabilization of two-coat auto-
motive lacquers [154,186]. Aplication of UV-absorbing HAS allows to determine at the same time the
permanence of photoantioxidants in the matrix by UV spectroscopy [133]. Stabilizers such as 93, 94
(R H, COCH3) or 95 (having high absorption coefficients at l max 295296 nm) are photostable and
provide at 3% a strong UV filtration effect on TSA clearcoats.

4.1. Physical factors in coating stabilization

Physical losses of photostabilizers from the surface layers resulting from volatility or extraction
reduce the durability of coatings [155,186]. However, the losses of stabilizers from the topcoats due
to migration into lower layers are considered as more serious. Solubility of stabilizers in particular layers
of the polymer matrix, adsorption on fillers or pigments and molecular architecture of stabilizers
(including molecular weight) are reflected in compatibility with the matrix, affect migration during
cure and weathering and distribution of stabilizers in the whole coating system. Migration of stabilizers
in crosslinked organic coatings is an undesirable process [187]. Rather complicated relations manage
partitioning of stabilizers in coatings over plastic substrates and in multilayer coatings.
Microtome technique yielding 25 mm cuttings of coatings was adapted to check distribution of
1328 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

Fig. 26. Distribution of HAS (W) and UV absorber (X) (a proprietary blend HAS/UVA was used) within an automotive coating
system (CC clearcoat, BC basecoat, UC undercoat with high solubility for stabilizers). (Reproduced with permission
from Eur Coat J 1991;10:622. 1991 C.R. Vincentz Verlag, Hannover [190].)

photostabilizers through the clearcoat-basecoat system of automotive coatings [187191]. Extracts of


the cuttings were analyzed by chromatographic and spectroscopic methods. This technique provides
information on effects of the chemical composition of the binder, curing or weathering conditions on
migration and concentration profiles of stabilizers within the whole coating system.
The migration of photostabilizers is strongly affected by their solubility in individual layers of the
coating. Experiments were performed with UVA and HAS added either to the basecoat or clearcoat. In
the first case, stabilizers (HAS in particular) migrate from the pigmented basecoat to both the clearcoat
and undercoat (primer). At high solubility of stabilizers in the undercoat, the clearcoat remains rather
unprotected (Fig. 26). On the contrary, using the undercoat with very low solubility of stabilizers, the
latter migrate preferentially to the clearcoat and enhance its protective power [190]. Similarly, high
solubility of stabilizers in the plastic substrate in comparison with the coating promotes migration of
stabilizers into the substrate [187].
Photostabilizers incorporated in clearcoat migrate slowly towards the primer through the basecoat
[134,168]. This reduces protection of the clearcoat. Only trace amounts of the stabilizers reach the
primer [190] (Fig. 27). Faster migrating stabilizers result in less durable coatings. Low-migrating HMW
and oligomeric stabilizers and non-migrating polymer-bound stabilizers have beneficial behavior.
Oxamide 23 was used to check the influence of the UVA concentration in the basecoat on the
stabilizer distribution in the whole two-coat system [154]. Compound 23 was incorporated into both
the clearcoat and basecoat. After accelerated weathering (fluorescence tube UV-B 313 nm), concentra-
tion of 23 was reduced in the clearcoat, especially in its upper layers. Formation of the UVA concentra-
tion gradient in the clearcoat in the two-coat coatings is due to migration into basecoat [154]. Under
comparable testing conditions, no losses of 23 were observed in mono-coat coating after 1000 h weath-
ering at 40C.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1329

Fig. 27. Distribution of HAS (W) and UV absorber (X) (a proprietary blend HAS/UVA was used) within an automotive coating
system (CC clearcoat, BC basecoat, UC undercoat). (Reproduced with permission from Eur Coat J 1991;10:622.
1991 C.R. Vincentz Verlag, Hannover Ref. [190].)

The rate of photostabilizer migration in the clearcoat during cure is to a large degree determined by the
viscosity of the thermosetting resin. Experiments were performed [187] with a styreneacrylate copo-
lymer crosslinked with melamine and doped with UVA 17a and HAS 35 (R COCH3). A fast migration
occurs in the early stages of the cure, when the resin viscosity is low. The stabilizer migration slows
down by several orders of magnitude as the crosslinking proceeds.
The microtome technique also provided information on HAS concentration profiles. After a wet-on-
wet application in a conventional solvent system, HAS 54 (R H) is homogeneously distributed
through the entire coating, even though the stabilizer was added only to the clearcoat. During baking,
HAS migrates into the basecoat. This reduces substantially the HAS protective level in the clearcoat.
After 1600 h weathering in Atlas weatherometer, the concentration gradient of HAS increases from the
upper part of the clearcoat up to the clearcot-basecoat boundary layer [154]. Decay of HAS 57 (R H)
in the clearcoat of the two-coat metallic finish under various radiation conditionswas checked by
microtome technique too. The decay rate monitored after Atlas xenotest exposure was closer to Florida
natural exposure than in tests performed with fluorescence tube UVCON 313 nm.

4.2. Influence of pigments and fillers (extenders) on performance of photostabilizers

Understanding of potential interactions between photostabilizers and pigments or fillers (extenders) is


necessary to achieve optimum durability of coatings. The efficiency of UVA and HAS can be signifi-
cantly retarded by immobilization or chemical interference on pigment/filler surfaces. Useful informa-
tion arises from model experiments in coatings [187,192] and observations in polyolefins [193].
Calcium carbonate and talc reduce performance of phenolic UVA and HAS in polypropylene (as
monitored by influence on the growth of carbonyl groups). A high adsorption was reported for basic
HAS (HMW in particular) on silica gels. It was proposed that nitroxides derived from HAS can be bound
1330 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

to HO groups on silica surface by hydrogen bonds. The performance of HAS 61 (R H) was reduced in
polyethylene pigmented with uncoated titanium dioxide. No problems were observed [194] using coated
TiO2. UV-grade carbon black influenced negatively performance of secondary HAS 61 (R H) in
polypropylene. Tertiary HAS 61 (R CH3) and secondary oligomeric HAS 100a were not adversely
affected, as confirmed by heat and light stability tests [150].
Important data on interactions in coatings or model systems were published recently [187,182].
Elucidation of adsorption of HAS 61 (R H or CH3) on some pigments and fillers in xylene suspension
revealed that almost 100% of the tested HAS was adsorbed on carbon black Monarch 1300. Secondary
HAS was adsorbed (in %) on talc (25), TiO2 (17) and calcium carbonate (75). Tertiary HAS was more
resistant and the measured adsorption was in the range 67%, with the exception of 40% adsorption on
calcium carbonate [150]. Spectroscopic investigation revealed that chemisorption (covalent bonding) is
the dominant interaction mechanism in inorganic pigments (carbon black and red ferric oxide) contrary
to physical adsorption by van der Waals forces dominating on titanium dioxide or silica filler particles
[192,195]. In the latter case, the physically adsorbed photostabilizers desorb on cure and migrate within
the coating. The chemisorbed stabilizers remain bound to the pigment surface. Moreover, a part of
chemisorbed HAS-containing ester moiety hydrolyses on curing [192]. A quantitative analysis of
extracts of microtomed cuttings of coatings indicated a lower level of stabilizers than originally
added to the formulation [187,189]. The physical loss by volatilization was approximately 3%. It was
concluded that a part of stabilizers is present in a non-extractable form [192,195]. Model experiments
disclosed high adsorption of phenolic UVA 17a, 10i, blend 10j with 10k and HAS 61 (R OC8H17) on
inorganic pigments carbon black and red ferric oxide. Only negligible amounts were adsorbed on
organic pigments Irgacolor Yellow, Sicotan Yellow, Heliogen Green and Blue or Irgazin Red.
Analysis of HAS 61 (R OC8H17) chemisorbed on ferric oxide revealed that the stabilizer was almost
quantitatively hydrolyzed to 127 and 128. A part of the dicarboxylic acid 128 was in the form of
carboxylate anion. Adsorption on ferric oxide was also observed with HAS 57 (R COCH3), 61
(R H, CH3) and UVA 10i. The chemisorbed acyl-HAS 57 (R COCH3) was transformed to the
secondary amine analog 57 (R H). Chemisorbed binuclear HAS 61 (R CH3) converts in salts of
tertiary amine 129 (R CH3) and 130 (R CH3). The secondary amine 61 (R H) was converted into
the relevant salts 129, 130 (R H). Transformation products 129, 130 cannot form nitroxides crucial for
the HAS mechanism and are lost as photoantioxidants.

Benzotriazole 10i added to acrylatemelamine coating doped with 10% ferric oxide was hydrolyzed after
cure almost quantitatively into the related carboxylate and octyl alcohol [192]. This is considered a conse-
quence of chemisorption of 10i on the pigment. According to [192], triazine 17a is resistant to chemisorption
on ferric oxide.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1331

The strength of the physical adsorption of UVA 10i and 17a on ultrafine particles of titanium dioxide
and silica was tested in two-layer clearcoats. Only one layer was UVA-doped [192]. Analyses of extracts
of microtomed cuttings disclosed that physically adsorbed UVA were desorbed during the cure and
migrated into the second layer. The results indicate that a higher concentration of photostabilizers is
necessary to protect coatings with chemisorbing pigments [192].

5. Conclusions

Two classes of photostabilizers, UVA and HAS, effectively protect coatings against atmospheric
degradation. Mechanism of phenolic UVA based on ESIPT and the role of transformation products
of HAS are well understood. This enabled explanation of physical and chemical factors limiting perma-
nence of both the stabilizer classes and a substantial enhancement of the inherent chemical efficiency by
structural modifications. All mechanistic data are exploited for optimum protection of coatings in hostile
environment.

Acknowledgements

The financial support by grant KONTAKT ME-184/99 from the Ministry of Education, Youth and
Sport of the Czech Republic, grant No. A 1050901 from the Grant Agency of the Academy of Sciences
of the Czech Republic and grant No. 12/96/K from the Academy of Sciences of the Czech Republic is
gratefully appreciated. The authors thank Mrs D. Dundrova for technical cooperation in preparation of
the manuscript.

References

[1] Paul S, editor. Surface coatings: science and technology, 2nd ed. Chichester: Wiley, 1996.
[2] Braun JH, Baidius A, Marganski RE. Prog Org Coat 1992;20:105.
[3] Weiss KD. Prog Polym Sci 1997;22:203.
[4] Stoye D, Freytag W, editors. Paints, coatings and solvents, 2nd ed. Weinheim: WileyVCH, 1998.
[5] Valet A. Light stabilizers for paints. Hannover: C.R. Vincentz Verlag, 1997.
[6] Paulus W. Microbiocides for the protection of materials. A handbook. London: Chapman and Hall, 1993.
[7] Schirman PJ, Dexter M. In: Calbo LJ, editor. Handbook of coating additives, vol. 1. New York: Dekker (Marcel), 1987
(225pp.).
[8] Cullis CF, Hirschler MM. The combustion of organic polymers. Oxford: Clarendon Press, 1981.
[9] Bauer DR. J Coat Technol 1994;66(835):57.
[10] Pospsil J, Horak Z, Krulis Z, Nespurek S. Macromol Symp 1998;135:247.
[11] Hill HMK, Ojunga-Andrew M, Wilson RC. Prog Org Coat 1994;24:147.
[12] Rabek JF. Photodegradation of polymers. Berlin: Springer, 1996.
[13] Olayan HB, Hamid HS, Owen ED. J Macromol Sci: Rev Macromol Chem Phys C 1996;36:671.
[14] Pickett JE. Polym Degrad Stab 1994;43:353.
[15] Leighton PA. Photochemistry of air pollutants. New York: Academic Press, 1961.
[16] Andrady AL. Adv Polym Sci 1997;128:47.
[17] Davis A, Sims D. Weathering of polymers. London: Applied Science Publishers, 1983.
[18] Johnston HS. Ann Rev Phys Chem 1992;43:1.
[19] Finlayson-Pitts BJ, Pitts JN. Chem Ind 1993:796.
[20] Gugumus F. Polym Degrad Stab 1998;62:403.
1332 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

[21] Wheeler EL. Gummi, Fasern, Kunstst 1990;43:612.


[22] Pospsil J. Polym Degrad Stab 1993;40:217.
[23] Pospsil J. Adv Polym Sci 1995;124:87.
[24] Rodgers WR, Garner DP, Cheever GD. J Coat Technol 1998;70(877):83.
[25] Trubiroha P, Schulz U. Polym & Polym Compos 1997;5:359.
[26] Haagen H. farbe lack 1983;89:410.
[27] Huang SJ. Comprehensive polymer science. In: Eastmond GC, Ledwith A, Russo S, Sigwalt P, editors. Polymer
reactions, vol. 6. Oxford: Pergamon Press, 1989 (597pp.).
[28] Eilender AL, Oppermann RA. In: Calbo LJ, editor. Handbook of coating additives, vol. 1. New York: Dekker (Marcel),
1987 (177pp.).
[29] Decker C, Biry S. Prog Org Coat 1996;29:81.
[30] Wichs ZW. In: Kroschwitz JK, editor. Encyclopaedia of polymer science and engineering, Supplement volume. 2nd ed.
New York: Wiley, 1989 p. 53.
[31] Braun JH. Prog Org Coat 1987;15:249.
[32] Voltz HG, Kaempf G, Fitzky HG, Klaerer A. ACS Symp Ser 1981;151:163.
[33] Lindberg B, Hauser CM. In: Allen NS, Rabek JF, editors. New trends in the photochemistry of polymers. London:
Elsevier, 1985 (187pp.).
[34] Bauer DR. Polym Degrad Stab 1995;48:259.
[35] Nespurek S. Mater Sci Engng C 1999;89:319.
[36] Pospsil J, Klemchuk PP. In: Pospsil J, Klemchuk PP, editors. Oxidation inhibition in organic materials, vol. I. Boca
Raton, FL. CRC Press, 1990 p. 1.
[37] Gerlock JL, Smith CA, Cooper VA, Dustiber TG, Weber WH. Polym Degrad Stab 1998;62:225.
[38] Allen NS, Regan CJ, McIntyrer R, Johnson BW, Dunk WAE. Prog Org Coat 1997;32:9.
[39] Gardette JL, Mailhot B, Lemaire J. Polym Degrad Stab 1995;48:457.
[40] Bauer DR, Mielewski DF. Polym Degrad Stab 1993;40:349.
[41] Bauer DR, Mielewski DF, Gerlock JL. Polym Degrad Stab 1992;38:57.
[42] Mielewski DF, Bauer DR, Gerlock JL. Polym Degrad Stab 1991;33:93.
[43] Allen NS, Regan CJ, McIntyrer R, Johnson BW, Dunk WAE. Polym Degrad Stab 1997;58:149.
[44] Mielewski DF, Bauer DR, Gerlock JL. Polym Degrad Stab 1993;41:323.
[45] Nichols ME, Gerlock JL, Smith CA. Polym Degrad Stab 1997;56:81.
[46] Bauer DR, Gerlock JL, Mielewski DF, Peck MCP, Carter RO. Polym Degrad Stab 1990;27:272.
[47] Gerlock JL, Bauer DR, Briggs LM, Hudgens JK. Prog Org Coat 1987;15:197.
[48] Herbst W, Hunger K. Industrial organic pigments. 2nd ed. Weinheim: VCH, 1997.
[49] Damm W, Hermann E. In: Gachter R, Muller H, editors. Plastics additives handbook. Munich: Hanser, 1990 (637pp.).
[50] Wypych G. Handbook of fillers. 2nd ed. Toronto: ChemTech, 1999.
[51] Lacoste J, Singh RP, Boussard J, Arnaud R. J Polym Sci: Polym Chem Ed 1987;25:2799.
[52] Klemchuk PP. Polym Photochem 1983;3:1.
[53] Allen NS. Polym Degrad Stab 1994;44:357.
[54] Allen NS, McKellar JF. In: Allen NS, McKellar JF, editors. Photochemistry of dyed and pigmented polymers. London:
Applied Science Publishers, 1980 (247pp.).
[55] Bauer DR. J Appl Polym Sci 1982;27:3651.
[56] Mori K, Tachi K, Muramatsu M, Torita K. 24th Congress FATIPEC, vol. A, Interlaken, 1998. p. A-101.
[57] Suppan P. Chemistry and light. London: Royal Chemical Society, 1994.
[58] Pospsil J, Nespurek S. In: Hamid HS, editor. Handbook of polymer degradation, New York: Dekker (Marcel), 2000
(191pp.).
[59] Pospsil J. Adv Polym Sci 1995;124:87.
[60] Donnett JB, Bansal RC, Wang M-J, editors. Carbon black, science and technology, 2nd ed. New York: Dekker (Marcel),
1993.
[61] Galbo JP. In: Salamone JC, editor. The polymeric materials encyclopaedia, vol. 5. Boca Raton, FL: CRC Press, 1996
616pp.
[62] Gugumus F. Polym Degrad Stab 1993;39:117.
[63] Heller HJ. Eur Polym J 1969;Supplement:105.
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1333

[64] Rabek JF. Photostabilization of polymers, principles and application. London: Elsevier, 1990.
[65] Allan M, Bally T, Haselbach E, Suppan P, Avar L. Polym Degrad Stab 1986;15:311.
[66] Gugumus F. In: Pospsil J, Klemchuk PP, editors. Oxidation inhibition in organic materials, vol. II. Boca Raton, FL: CRC
Press, 1990. p. 29.
[67] Ghiggino KP. J Macromol Sci A: Pure Appl Chem 1996;33:1541.
[68] Heller H-J, Blattmann HR. Pure Appl Chem 1972;30:145.
[69] Heller H-J, Blattmann HR. Pure Appl Chem 1973;36:141.
[70] Weller A. Z Elektrochem 1956;60:1144.
[71] Formoshino SJ, Arnaut LJ. J Photochem Photobiol A 1993;75:21.
[72] Barbara PF, Renzeipis PN, Prus LE. J Am Chem Soc 1980;102:5651.
[73] Ghiggino KP, Scully AD, Leaver IH. J Phys Chem 1986;90:5089.
[74] Keck J, Stuber GJ, Kramer HEA. Angew Makromol Chem 1997;252:119.
[75] Bigger SW, Ghiggino KP, Leaver IH, Scully AD. J Photochem Photobiol A 1987;40:391.
[76] Woessner G, Goeller G, Kollat P, Stezowski JJ, Hauser M, Klein UKA, Kramer HEA. J Phys Chem 1984;88:5544.
[77] Stueber GJ, Kieninger M, Schettler H, Bush W, Goeller B, Franke J, Kramer HEA, Hoier H, Henkel S, Fischer P, Port H,
Hirsch T, Rytz G, Birbaum J-L. J Phys Chem 1995;99:10 097.
[78] Catalan J, Fabero F, Guijaro MS, Charamnut RM, Maria MDS, Foces-Foces Mde la C, Cano FH, Elguero J, Sastre JR. J
Am Chem Soc 1990;112:747.
[79] Ormson SM, Brown RG. Prog React Kinet 1994;19:45.
[80] Arnaut LJ, Formoshino SJ. J Photochem Photobiol A 1993;75:1.
[81] Le Gourrierec D, Ormson SM, Brown RG. Prog React Kinet 1994;19:211.
[82] Hilbert F. Adv Phys Org Chem 1986;22:113.
[83] Kosower EM, Huppert D. Ann Rev Phys Chem 1986;37:127.
[84] Goeller G, Rieker J, Maier A, Stezowski JJ, Daltrozz E, Neureiter M, Port H, Wiechmann M, Kramer HEA. J Phys Chem
1988;92:1452.
[85] Rieker J, Lemmert-Schmitt E, Goeller G, Roessler M, Stueber GJ, Schetter H, Kramer AEA, Stezowski JJ, Hoier H,
Henkel S, Schmidt A, Port H, Wiechmann M, Rody J, Rytz G, Slongo M, Birnbaum J-L. J Phys Chem 1992;96:10 225.
[86] Foster T. Pure Appl Chem 1970;24:443.
[87] Kramer HEA. farbe lack 1986;92:919.
[88] Ghiggino KP, Scully AD, Bigger SW. ACS Symp Ser 1988;381:57.
[89] Otterstedt JEA. J Phys Chem 1973;58:5716.
[90] Jaffe HH, Lloyd Jones H. J Org Chem 1965;30:964.
[91] Klopfer W. J Polym Sci: Symp 1976;57:205.
[92] Prieto MFR, Nichel B, Grellmann KH, Mordzinski A. Chem Phys Lett 1988;146:387.
[93] Eisenberg H, Nichel B, Ruth AA, Al-Soufi W, Grellmann KH, Novo M. J Phys Chem 1991;95:10509.
[94] Ghiggino KP, Scully AD, Bigger SW, Leaver IM. J Polym Sci: Polym Chem Ed 1987;25:1619.
[95] Gupta A, Scott GW, Klinger D, Vogl O. ACS Symp Ser 1983;220:191.
[96] Pospsil J, Nespurek S, Zweifel H. In: Al-Malaika S, Golovoy A, Wilkie C, editors. Additives and modifiers for
polymers. Oxford: Blackwell, 1999 (36pp.).
[97] Rytz G, Hilfiker R, Schmidt E, Schmitzer A. Angew Makromol Chem 1997;247:213.
[98] Schizuka H, Machii M, Higaki Y, Tanaka M, Tanaka I. J Phys Chem 1985;89:320.
[99] Pickett JE, Moore JE. Polym Degrad Stab 1993;42:231.
[100] Pickett JE. Macromol Symp 1997;115:127.
[101] Pickett JE, Moore JE. Adv Chem Ser 1996;249:287.
[102] Decker C, Biry S, Zahouilly K. Polym Degrad Stab 1995;49:111.
[103] Decker C. Adv Chem Ser 1996;249:319.
[104] Bonenkamp JE, Maecker NL. J Appl Polym Sci 1994;54:1593.
[105] Al-Malaika S. In: Allen G, Bengton JC, editors. Comprehensive polymer science, vol. 6. New York: Pergamon Press,
1989 (539pp.).
[106] Woessner G, Goller G, Rieker J, Hoier H, Stezowski JJ, Daltrozzo E, Neureiter M, Kramer HEA. J Phys Chem
1985;81:3629.
[107] Wiechmann M, Port H, Frey W, Laermer F, Elsaesser T. J Phys Chem 1991;95:1918.
1334 J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335

[108] Dux R, Ghiggino KP, Vogl O. Aust J Chem 1994;47:1461.


[109] Decker C, Zahouilly K. Polym Mater Sci Engng 1993;68:70.
[110] Gerlock JL, Tang W, Dearth MA, Korniski TJ. Polym Degrad Stab 1995;48:121.
[111] Decker C, Moussa K, Bendaikha T. J Polym Sci: Polym Chem Ed 1991;29:739.
[112] Decker C, Zahouilly K, Biry S. Polym Degrad Stab 1995;47:109.
[113] Hutson GV, Scott G. Eur Polym J 1974;10:45.
[114] Pospsil J. Adv Polym Sci 1980;36:69.
[115] Pospsil J. In: Allen NS, editor. Developments in polymer photochemistry, vol. 2. London: Applied Science Publishers,
1981.
[116] Chakraborty KB, Scott G. Eur Polym J 1979;15:35.
[117] Hodgeman DK, Gellert EP. J Polym Sci: Polym Chem Ed 1980;18:1105.
[118] Dearth MA, Korniski TJ, Gerlock JL. Polym Degrad Stab 1995;48:111.
[119] Decker C, Zahouilly K. J Polym Sci: Polym Chem Ed 1998;36:2571.
[120] Gerlock JL, Prater T, Kaberline S, de Vries J. Polym Degrad Stab 1995;47:405.
[121] Valet A, Sitek F, Berner G. farbe lack 1988;94:725.
[122] Decker C, Bendaikha T. In: Patsis AV, editor. Advances in the stabilization and controlled degradation of polymers.
Lancaster, PA: Technomic, 1989 (143pp.).
[123] Decker C. Chimia 1993;47:378.
[124] Valet A, Jung T, Kohler M, Chang, CH. International UV/EB Processing Conference and Exhibition RadTech 98,
Chicago, 1998. p. 396.
[125] Valet A, Jung T, Kohler M. farbe lack 1998;2:42.
[126] Carlini C, Angiolini L. Adv Polym Sci 1995;123:127.
[127] Olson DR. J Appl Polym Sci 1983;28:1159.
[128] Valet A, Koehler M. Eur Pat Appl EP 458741, 1991 (Chem. Abstr 1992;116:108 390.)
[129] Gupta A, Scott QW, Klinger D. ACS Symp Ser 1981;151:27.
[130] Birbaum J-L, Rytz G, Vien Van Toan, Valet A, Wuerms A. Eur Pat Appl EP 711804, 1996 (Chem Abstr
1996;125:88154.
[131] Hida K, Jpn Kokai Tokkyo Koho JP 10060306, 1998 (Chem Abstr 1998;128:206 024.
[132] Bauer DR, Gerlock JL, Mielewski DF. Polym Degrad Stab 1990;28:39.
[133] Avar L, Bill R, Hess E. farbe lack 1986;92:915.
[134] Jurgetz A, Rothbacher H, Bliefert C. farbe lack 1985;91:921.
[135] Berner C, Remboldt M. farbe lack 1981;87:930.
[136] Gerlock JL, Bauer DR, Mielewski DF. 11th International Conference on Advances in Stabillization and Controlled
Degradation of Polymers, Luzern, 1989. p. 25.
[137] Sedlar J, Petruj J, Pac J, Zahradnckova A. Eur Polym J 1980;16:663.
[138] Kurumada T, Ohsawa H, Oda O, Fujita T, Toda T, Yoshioka T. J Polym Sci: Polym Chem Ed 1985;23:1477.
[139] Carlsson DJ, Zhang C, Wiles DM. J Appl Polym Sci 1987;33:875.
[140] Step EN, Turro NJ, Gande ME, Klemchuk PP. Macromolecules 1994;27:2529.
[141] Klemchuk PP, Gande ME. Makromol Chem, Macromol Symp 1989;28:117.
[142] Bowry VW, Ingold KU. J Am Chem Soc 1992;114:4992.
[143] Bechtold K, Hess E, Ligner G. farbe lack 1993;99:25.
[144] Bauer DR, Gerlock JL. Polym Degrad Stab 1986;14:53.
[145] Wiles DM, Jensen JPT, Carlsson DJ. Pure Appl Chem 1983;55:1651.
[146] Lucki J, Rabek JF, Ranby B. J Appl Polym Sci 1988;36:1067.
[147] Bramer D, Holt M. 13th International Conference on Advances in Stabillization and Controlled Degradation of Poly-
mers, Luzern. 1991. p. 23.
[148] Gaines GL. Polym Degrad Stab 1990;27:13.
[149] Avar L, Bechtold K, Hess E. 22nd Fatipec Congress, vol. 3, Budapest, 1994. p. 45.
[150] Kikkawa K. Polym Degrad Stab 1995;49:135.
[151] Allen NS, Edge M, He J, Chen W, Kikkawa K, Minagawa M. Polym Degrad Stab 1994;44:99.
[152] Ravichandran R, Galbo JP. Eur Pat Appl EP 389427, 1990 (Chem Abstr, 1991;114:63 325).
[153] Slongo M, Birbaum J-L, Rody J, Valet A. Eur Pat Appl EP 453405, 1991 (Chem Abstr 1992;116:41 487).
J. Pospsil, S. Nespurek / Prog. Polym. Sci. 25 (2000) 12611335 1335

[154] Bohnke H, Avar L, Hess E. J Coat Technol 1991;63(799):53.


[155] Pospsil J. Adv Polym Sci 1991;101:65.
[156] Bailey O, Vogl O. J Macromol Sci: Rev Macromol Chem Phys 1976;14:267.
[157] Valet A, Meuwly R, Slongo M. Eur Pat Appl EP 526399, 1993 (Chem Abstr 1993;119:182 935).
[158] Lilly KL. Eur Pat Appl EP 736577, 1996 (Chem Abstr 1996;125:331 679).
[159] Lastuvkova H, Pospsil J, Nespurek S, Bandlitz S, Habicher WD. Polym & Polym Compos 1999;7:165.
[160] Boehnke H, Hess E. Ger Offen DE 4010444, 1990 (Chem Abstr 1991;114:83 963).
[161] Gras R, Wolff E. GerOffen DE 19650045, 1998 (Chem Abstr 1998;129:68 886).
[162] Bartus J, Simosick WJ, Vogl O. Polym J 1995;27:703.
[163] Malk J, Ligner G, Avar L. Polym Degrad Stab 1998;60:205.
[164] Ghiggino KP, Scully AD, Bigger SW, Vogl O. J Polym Sci: Polym Lett 1988;26:505.
[165] OConnor DB, Scott GW, Coulter DR, Yavronian A. J Phys Chem 1991;95:10 252.
[166] Gerlock JL, Van Oene H, Bauer DR. Eur Polym J 1983;19:11.
[167] Berner G, Rembold M. Org Coat Sci Technol 1983;6:55.
[168] Jurgetz A, Nomayr H, Rotbacher H. 18th Congress FATIPEC, vol. 1/A, Venezia, 1986. 263pp.
[169] Valet A. farbe lack 1994;100:919.
[170] Slongo M, Birbaum J-L, Valet A. Ger Offen DE 4008125, 1990 (Chem Abstr 1991;114: 165 660).
[171] Valet A. Polym Paint Colour J 1992;182(4311):406.
[172] Atley RD, Heywood NW, Harlan H. Eur Coat J 1998(1-2):46.
[173] Valet A, Roger D. Surf Coat Aust 1997;34(9):22.
[174] Valet A. farbe lack 1996;102:40.
[175] Valet A. Polym Paint Colour J 1995;185:31.
[176] Decker C, Zahouilly K. Polym Degrad Stab 1999;64:293.
[177] Gerlock JL, Bauer DR, Briggs LM. ACS Symp Ser 1985;280:119.
[178] Gerlock JL, Riley T, Bauer DR. Polym Degrad Stab 1986;14:73.
[179] Ligner G, Hess E. Surf Coat Int 1993;76:233.
[180] Carlsson DJ, Chan KA, Durmis J, Wiles DM. J Polym Sci: Polym Chem Ed 1982;20:575.
[181] Gerlock JL, Bauer DR, Briggs LM. Polym Degrad Stab 1986;14:53.
[182] Bauer DR, Dean MJ, Gerlock JL. Ind Engng Chem Res 1988;27:65.
[183] Valet A. Paintindia 1992;42(7):35.
[184] Berner G, Rembold M. Org Coat 1984;6:55.
[185] Valet A. farbe lack 1990;96:689.
[186] Avar L, Boehnke H, Hess E. 19th Congress, FATIPEC, vol. I, Aachen, 1988. p. 317.
[187] Haacke G, Andraves FF, Campbell BH. J Coat Technol 1996;68(855):57.
[188] Bohnke H, Hess E. Eur Coat J 1990;5:222.
[189] Haacke G, Brinen JS, Larkin PJ. J Coat Technol 1995;67(843):29.
[190] Ligner G, Hess E. Eur Coat J 1991;10:622.
[191] Ligner G, Hess E. farbe lack 1991;97:676.
[192] Haacke G, Longordo E, Brinen JS, Andrawes FF, Campbell BH. J Coat Technol 1999;68(888):87.
[193] Allen NS, Gardette J-L, Lemaire J. Dyes & Pigments 1982;3:295.
[194] Allen NS, Bullen DJ, McKellar JF. J Mater Sci 1978;13:2692.
[195] Haacke G, Longordo E, Andrawes FF, Campbell BH. Prog Org Coat 1998;34:75.

Das könnte Ihnen auch gefallen