Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

VOL II PModern materials

UBLISHED IN THE 13TH TRIENNIAL MEETING RIO DE JANEIRO PREPRINTS 911

Abstract Ageing studies of acrylic emulsion paints


A selection of acrylic emulsion paints
was applied to glass slides, artificially
aged with intense light and examined
with pyrolysis-gas chromatography-
mass spectrometry, Fourier transform Tom Learner*
infrared spectroscopy, size exclusion Tate Millbank
chromatography, thermogravimetric London SW1P 4RG, United Kingdom
analysis and water immersion to Fax: +44 20 7887 8982
monitor any changes with light E-mail: tom.learner@tate.org.uk
exposure. The principal change Oscar Chiantore and Dominique Scalarone
appeared to be the incremental Department of IPM Chemistry
disappearance of surfactant from the University of Torino
surface of the paint films. The via P.Giuria 7
solubility in tetrahydrofuran, a high- 10125 Torino, Italy
swelling organic solvent, was ob- Fax: +39 11 670 7855
served to decrease slightly for several E-mail: chiantore@ch.unito.it, scalarone@ch.unito.it
paints, with an accompanying de-
crease in average molecular weight
Introduction
(MW) in their soluble fractions, both
of which pointed to cross-linking Since their introduction in the late 1950s, acrylic emulsion paints have been widely
reactions. However, in some paints used by artists and conservators. Manufacturers have always claimed these paints
no change was observed and in a few
exhibit high durability, resistance to yellowing and permanent flexibility, because
an increase in solubility (and MW of
the soluble fraction) was noted. they are based on emulsions designed for exterior house paints, in which resistance
Colour measurements of thermally to light, humidity and temperature are clearly required (Hochheiser 1986). On a
aged samples showed acrylic emulsion visual level, it does appear that acrylic emulsion paints are less prone to cracking
paints to be highly resistant to or (natural) discoloration compared to oil paints of a similar age, but very little
yellowing, in contrast to oil and alkyd research has been done to characterize their ageing behaviour.
paints. Most studies have concentrated on unpigmented acrylic emulsions (or disper-
sions). De Witte et al. (1984) and Howells et al. (1984) used different ageing
Keywords methods to examine various commercial products, many of which have been used
in acrylic paint formulations. They reported that films often yellowed under dark
acrylic emulsion paints, light ageing, and warm conditions, but this was not accompanied by any change in removability
polyethylene glycol (PEG) surfactant, or mechanical properties. The most detailed study on the ageing of an artists’
cross-linking, chain scission acrylic emulsion product, by Whitmore and Colaluca (1995), was also undertaken
on an unpigmented medium. Thin films of Liquitex acrylic gloss medium were
aged naturally in the dark as well as artificially by exposure to elevated temperatures
and various sources of ultraviolet (UV) radiation. It was found that in the dark,
films acquired a slightly yellow colour, lost solubility and gained in tensile strength,
all of which were attributed to slight cross-linking of the polymer. Exposure to UV,
however, caused a loss of tensile strength and an increase in solubility, suggesting
a breakdown of the polymer by chain-scission reactions. It was noted that all of
these changes were extremely small – the tensile testing even of UV-aged samples
did not cause them to break – and it was concluded that the product certainly
exhibited high stability when exposed to ambient indoor lighting.
Most acrylic emulsion paints are bound in either a poly (ethyl acrylate-methyl
methacrylate), p(EA/MMA) copolymer or a poly (n-butyl acrylate-methyl meth-
acrylate), p(nBA/MMA) copolymer (Learner 2000). Unfortunately, the chemical
and physical differences between acrylic solutions and emulsions negate much of
the knowledge about ageing of acrylic resins (e.g. Feller et al. 1981, Melo et al.
1999, Chiantore et al. 2000). It is therefore not possible to predict the ageing of
emulsion paints from the relative stability of these copolymers alone, although it
is likely that certain trends about stability may still hold, such as the vulnerability
of tertiary hydrogen atoms to abstraction (present on the backbone of all acrylate
polymers). Unlike solutions, emulsions contain several additional components,
including surfactants, protective colloids, thickeners, anti-foam agents, biocides,
freeze-thaw agents, pH buffers and coalescing solvents (e.g. Learner 2000).
Although some of these ‘additives’ evaporate from the paint film on drying, the
majority are non-volatile and therefore likely to remain in the dried paint film, and
any of these may affect the ageing characteristics of the material. The influence that

*Author to whom correspondence should be addressed


912 ICOM COMMITTEE FOR CONSERVATION, 2002 VOL II

a single minor constituent can have on the ageing of an emulsion system was well
demonstrated by Howells et al. (1984), who showed that a 2% addition of one
thickener had a remarkable yellowing effect on emulsions.
When considering emulsion paints, there are even more variables to consider,
with the presence of pigments, extenders and further additives, such as wetting
agents and dispersing agents. Pigments in particular may have a strong influence
on ageing properties, as they have the potential to act either as catalysts for ageing
reactions, for example by the adsorption of oxygen onto their surfaces, or as
stabilizers, by scattering and absorbing light and/or UV radiation. As part of a
long-term study of the ageing and effects of treatment on acrylic emulsion paints,
thin films of commercially available paints were prepared and exposed to periods
of artificial light ageing, as well as elevated temperatures. For each ageing protocol,
examples of artists’ paints with other binding media, including oil, alkyd and
polyvinyl acetate emulsion, were included for comparison. This paper reports on
preliminary observations from comparative measurements on paints before and
after ageing using a variety of analytical techniques.

Experimental
Samples
Samples of acrylic emulsion paints and other artists’ media for comparison were
kindly provided by a number of artists’ colourmen in 1995. A full list of
manufacturers along with the polymers used as binding media (checked by
pyrolysis-gas chromatography-mass spectrometry; see Learner 2001 for analytical
details) is given in Table 1. A selection of paints was requested, usually including
titanium white, cobalt blue, cadmium red, yellow ochre, mars black, phthalocyanine
blue, ‘azo’ red and ‘azo’ yellow. It should be remembered that formulations alter
constantly, so these copolymers may not be used in current products. Indeed, many
paints containing p(EA/MMA) copolymers have been reformulated with p(nBA/
MMA) emulsions. Thin films were prepared from a range of paints as ‘draw downs’
on cleaned microscope glass slides using a film applicator (Sheen Instruments). The
wet film thickness of emulsion paints was set at 120 µm (and 60 µm for oils and
alkyds), which produced dried films of between 30 and 40 µm. Samples were left
to dry for five weeks prior to any ageing or analysis, for which removal from the
slide was always required (except for colour measurements).

Artificial ageing
In light ageing, samples were placed in front of six daylight colour-rendering
fluorescent tubes (Philips TLD 94), behind a sheet of Perspex VE (ICI) ultraviolet
filter to cut out all radiation below 400 nm. Light intensity was measured as 18,000
lux at an average 25°C and 40% relative humidity (RH). Two slides of each paint
were placed in the light box with a third (control) kept in the dark at ambient

Table 1. List of paint brands used in ageing tests, including the identified binding medium
Manufacturer Range Binding medium Web page
Daler Rowney (U.K.) Cryla p(nBA/MMA) www.daler-rowney.com
Golden (U.S.A.) Artists’ acrylic p(EA/MMA) www.goldenpaints.com
Grumbacher (U.S.A.) Artists’ acrylic p(EA/MMA) www.sanfordcorp.com
Lascaux (Switzerland) Artists’ acrylic p(nBA/MMA) www.lascaux.ch
LeFranc & Bourgeois (France) Flashe pVA/VeoVa www.lefranc-bourgeois.com
Liquitex (U.S.A.) Liquitex p(EA/MMA) www.liquitex.com
Lukas (Germany) Lukascryl p(EA/MMA) www.lukas-online.com
Maimeri (Italy) Brera p(nBMA/styrene/ 2EHA) www.maimeri.it
Spectrum (U.K.) Spectracryl p(nBA/MMA) www.spectrumoil.com
Talens (Netherlands) Rembrandt p(EA/MMA) www.talens.com
Winsor & Newton (U.K.) Artists’ acrylic p(nBA/MMA) www.winsornewton.com
Winsor & Newton (U..K) Griffin alkyd www.winsornewton.com
Key:
p(nBA/MMA) = poly (n-butyl acrylate - methyl methacrylate) copolymer
p(EA/MMA) = poly (ethyl acrylate - methyl methacrylate) copolymer
pVA/VeoVa = poly (vinyl acetate – vinyl versatate) copolymer
p(nBMA/styrene/ 2EHA) = poly (n-butyl methacrylate – styrene – 2-ethylhexyl acrylate)
terpolymer
VOL II Modern materials 913

temperature. A complete set was removed after 2688 hours (16 weeks) and a second
after 5376 hours (32 weeks). Assuming reciprocity, and based on an annual museum
exposure of 3000 hours at 200 lux, this correlates to approximately 90 and 180
years under museum conditions.
In thermal ageing, separate samples (all white paints) were placed in a Fisons 185
HWC environmental oven, one set of samples at 50°C and the other at 60°C (both
at 55% RH). A third set was used as control and kept in the dark at ambient
temperature. Once ageing was complete, the samples were all stored in the dark
at ambient temperature.

Pyrolysis-gas chromatography-mass spectrometry (PyGCMS)


Samples were pyrolysed at 610°C for eight seconds in a Curie point pyrolyser
(FOM-4LX system). The pyrolysis unit was mounted directly onto a HP 5890 gas
chromatograph (Hewlett Packard), held at 40°C for two minutes, then ramped to
350°C at 10°C/min through an SGE BPX-5 column (25 m length; 0.22 mm inside
diameter) and held for a further two minutes. The gas chromatograph was
interfaced to an Incos 50 quadrupole mass spectrometer (Finnigan MAT) using
electron impact at 70eV. Data were processed with Datamaster II software
(Finnigan MAT).

Fourier Transform infrared spectroscopy (FTIR)/attenuated total reflection (ATR)


Transmission spectra were collected (64 scans at 4 cm-1 resolution) on a 2000 series
FTIR (Perkin Elmer) with a diamond cell/4X beam condenser (both Spectra Tech)
accessory. ATR spectra were obtained (200 scans at 4 cm-1 resolution) on a Nic-
Plan FTIR microscope (Nicolet) using a Spectra-Tech ATR objective with zinc
selenide crystal. Both systems were purged with dry air and data were processed
with Omnic 4.1 software (Nicolet).

Size exclusion chromatography (SEC)


Molecular characterization was performed with a modular SEC system, compris-
ing a Waters M-45 pump, a Rheodyne 7010 injection valve, a differential
refractometer ERC 7510 (Erma) and PL Gel type columns (Polymer Labs, 10 µm
particle diameter). Tetrahydrofuran solutions were separated from solid impurities
by filtration through a 0.45 µm PTFE membrane filter. Calibration between the
molecular weights 1600000 and 147 was obtained with PMMA narrow distribu-
tion standards and o-dichlororobenzene and a cubic fit for the calibration curve
was used.

Thermogravimetric analysis
Thermogravimetric measurements were run on a DuPont Thermal Analyst SDT
2960, under nitrogen and air flow of 90 cm3/min and a heating rate of 10°C/min
up to 800°C.

Water immersion
Samples of dry paint (approximately 5 mm2) were weighed on a C-32 Microbal-
ance (Cahn) before and after 24 hours’ immersion in deionized water. Carbon to
hydrogen ratios of these water extracts were measured by dynamic flush combus-
tion at 1600°C with an EA1108 Elemental Analyzer (Carlo Erba).

Colour change
Colour measurements were made with a Minolta CF-221 chromameter on the
CIELAB 1976 colour system using illuminant C and averaging eight readings
from different places over each paint film. All three coordinates were meas-
ured, but the ∆b* value (measuring the yellow-blue scale) gave the best
indication of yellowing.
914 ICOM COMMITTEE FOR CONSERVATION, 2002 VOL II

Figure 1. PyGCMS of Spectacryl yellow ochre acrylic


emulsion paint, from the unaged paint (upper pyrogram)
and after 32 weeks’ light ageing (lower pyrogram)

Results and discussion


Pyrolysis-gas chromatography-mass spectrometry (PyGCMS)
Examination by PyGCMS yielded no measurable differences after ageing. For
example, Figure 1 shows pyrograms from an unaged ‘Spectacryl yellow ochre’
acrylic emulsion paint (Spectrum) and the same paint after 32 weeks’ light ageing.
The binder is a p(nBA/MMA) copolymer and the y-axis is expanded to show the
smaller peaks, such as the group of four dimers/sesquimers (scans 1000–1400) and
four trimers (scans 1650–2000) in detail (for full characterization of peaks, see
Learner 2001). If oxidation played a significant role in ageing, it would be
expected for oxidation products to become visible in the pyrogram. Since no new
peaks were identified and relative areas remained unchanged, these measurements
suggested that oxidation had not occurred in these acrylic copolymers. However,
neither cross-linking nor chain-breaking reactions would have affected the
pyrograms – the polymer is broken into small fragments on pyrolysis – so these
ageing reactions could not be ruled out. It was noted that significant changes were
observed in the pyrograms of oil paints after light ageing, principally the reduction
in area of the octadecenoic (oleic) acid peak.

Fourier Transform infrared spectroscopy (FTIR)


Recording FTIR spectra in transmission mode also yielded no significant
differences between spectra of unaged and aged paints. Figure 2 shows the spectra
of ‘Rembrandt naphthol red acrylic emulsion paint’ (Talens) before ageing and
after both periods of light ageing. The characteristic peaks of the p(EA/MMA)
copolymer medium are C-H stretching at 2985 and 2955 cm-1, C=O stretching
at 1732 cm-1, C-O stretching at 1177 and 1162 cm-1 and a further characteristic
fingerprint region peak at 1030 cm-1. Other peaks from two organic pigments
(1553 cm-1 from PR112 and 1693 cm-1 from PO43) and a chalk extender (1445,
876 and 712 cm-1) are also visible. Although minor changes could be identified
between spectra, these were neither sequential with the amount of light exposure
nor reproducible, and certainly lay within error limits. An unfortunate conse-
quence of this mode of measurement was that sample size and thickness could not
be held constant, so an appreciable variation in relative peak heights and widths
was possible.
VOL II Modern materials 915

Figure 2. FTIR (diamond cell) of Rembrandt Figure 3. FTIR (µATR) spectra of Golden polymer
naphthol red acrylic emulsion paint, from the unaged gloss medium, from the unaged medium (upper
paint (upper spectrum), after 16 weeks’ (middle spectrum), after 16 weeks’ (second spectrum down) and
spectrum) and 32 weeks’ light ageing (lower spectrum) 32 weeks’ light ageing (third spectrum down), and from
the underside of the unaged medium (lower spectrum)

However, when FTIR spectra were collected by attenuated total reflection, a


significant and reproducible change with light exposure was observed. Figure 3
shows the FTIR spectra of unaged ‘Polymer gloss medium’ (Golden) and those after
the two periods of light exposure. The sharp peaks at 2889, 1343, 1110 and 964
cm-1 were seen to decrease in intensity with light exposure. These peaks correspond
to a polyethylene glycol (PEG) component, previously identified as a surfactant in
acrylic emulsion systems (Whitmore 1996) and a class of material whose numerous
ether linkages would make them prone to deterioration with light exposure. The
same incremental reduction of surfactant peaks was also observed in cadmium red
and titanium white (both Golden) emulsion paints, but absorbing bands from
pigments/extenders made the features less obvious, so are not shown here. The
lowest spectrum was taken from the underside of the emulsion film and exhibits no
surfactant peaks, suggesting that it had been drawn to the film’s surface with water
evaporation.

Size Exclusion Chromatography (SEC)


SEC is a standard technique for determining molecular weight (MW) distribution
in polymers (e.g. Chiantore and Lazzari 1996), so it was hoped this technique
would detect increases (i.e. cross-linking) or decreases (i.e. chain scission) in MW.
However, this is only possible in solution and unfortunately none of the acrylic
emulsion paints could be dissolved completely, even in high-swelling organic
solvents such as tetrahydrofuran. SEC measurements were therefore collected from
the soluble fraction of these paints, in an attempt to obtain an indication of the trend
towards chain-breaking or cross-linking. Figure 4 shows chromatograms from
‘Lukascryl cadmium red’ acrylic emulsion paint (Lukas) before and after light
ageing. Since this analysis was run five years after the ageing programme, an
additional run was made with a fresh sample. Two trends were apparent: first was
the incremental disappearance of the low molecular weight component, previously
identified as the PEG surfactant (Chiantore et al. 2001), and second was a reduction
in MW in the soluble fraction. Although this suggests the soluble fraction had
916 ICOM COMMITTEE FOR CONSERVATION, 2002 VOL II

_________
freshly dried sample
––––––
dark aged 5 years
. . . light aged 16 weeks
Refractive index (a.u.)
– – – –
…..…. light aged 32 weeks

_________
freshly dried sample
––––––
dark aged 5 years
. . . light aged 16 weeks
– – – –
…..…. light aged 32 weeks

12 14 16 18 20 22 24
Retention time (min)

Figure 4. SEC chromatograms from Lukascryl cadmium red acrylic emulsion paint

100
insoluble
fraction

90

b
Weight (%)

80
c TGA curves:
d a) paint sample
70 e b) solid residue after THF extraction - unaged sample
c) solid residue after THF extraction - 5 years dark aged
60 d) solid residue after THF extraction - 16 weeks light aged
e) solid residue after THF extraction - 32 weeks light aged
a
50

0 200 400 600 800


Temperature (°C)

Figure 5. TGA curves for Rembrandt titanium white acrylic emulsion paint

Table 2. List of paints examined by SEC (for MW change in soluble fraction)


and TGA (for amount of insoluble material)
Brand Main pigment MW change in Amount of
soluble fraction insoluble material
Lukasryl Cadmium red Decrease Increase
Rembrandt Cadmium red Decrease Increase
Rembrandt Azo red (PR112) Decrease Increase
Rembrandt Titanium white Decrease Increase
Rembrandt Azo yellow (PY3) Decrease Unchanged
Rembrandt Mars black Unchanged Increase
Lukascryl Cobalt blue Unchanged Increase
Rembrandt Phthalo blue Unchanged Unchanged
Lukascryl Titanium white Not tested Unchanged
Rembrandt Cobalt blue Increase Decrease
Rembrandt Yellow ochre Increase Decrease

undergone some chain scission reactions on ageing, the same shift in MW


distribution would be obtained if the higher MW fraction had become insoluble
through cross-linking. Although this MW decrease was seen in six paints tested,
three remained unchanged and two showed an increase (see Table 2).

Thermogravimetric analysis (TGA)


Whereas TGA measurements on the unaged and aged paints appeared similar,
differences were noticed when the technique was applied to the solid residues after
extraction with solvent. Figure 5 shows the curves for ‘Rembrandt titanium white’
acrylic emulsion paint (Talens), with curves obtained from the solid residues after
tetrahydrofuran extraction of paints after varying amounts of light ageing. As with
SEC measurements, a fresh sample was compared to the original ‘unaged’ sample,
VOL II Modern materials 917

which had already naturally aged for five years. For each curve the major loss of
weight corresponded to the volatilization of the acrylic polymer, although this area
% insoluble material after 24 hour

98 would also include organic additives, and the second smaller one around 650°C is
from the loss of carbon dioxide from calcium carbonate extender. For this sample,
97
it was seen that light exposure caused an increase in the insoluble fraction,
water immersion

96 suggesting that a degree of cross-linking had occurred in the polymer. This change
cobalt blue was not always observed with other paints (see Table 2): half of these yielded a
95 yellow ochre
similar increase in insoluble fraction with light ageing, but three were unchanged
phthalo blue
94 and two showed a decrease. It could be seen, however, that an inverse relationship
crimson
unaged
16 exists between the SEC and TGA measurements (i.e. when SEC measurements
32
w eeks
w eeks
yielded a decrease in MW of the soluble fraction, this was accompanied by a
light exposure
li tendency for the amount of insolubility to increase, and vice versa).

crimson Figure 6. Weight loss


Water immersion measurements
phthalo blue in Cryla acrylic
yellow ochre
emulsion paints after All films yielded a decrease in dry weight after 24 hours’ water immersion, ranging
24 hours’ water between 1.8% and 9.3%. This upper limit is a very significant weight loss from a
cobalt blue
immersion
dried film. Larger differences were typically observed for paints containing
organic pigments, which often require larger quantities of water-soluble additives,
such as wetting agents. However, the amount of material extracted was consistently
reduced with light exposure. Figure 6 shows the weight loss for four Cryla
(Rowney) acrylic emulsion paints for unaged and light-aged films, crimson and
phthalo blue being organic pigments.
Elemental analysis was also done on the water in which the paint fragments had
been immersed, to measure quantitatively the amount of organic material that had
been extracted. Figure 7 shows the amount of carbon detected from these water
extracts up to an hour of immersion (on an arbitrary scale). The organic extraction
appeared to occur very rapidly on initial exposure to water and then level off. The
quantity of carbon detected in the extracts was found to be higher from paints
containing organic pigments and also decreased with light exposure.

Colour change
Colour measurements were taken on a series of white paints after thermal ageing,
a commonly used method for discoloration studies. Figure 8 shows the onset of
yellowing at 60°C, viewed as the change in b* CIELAB (yellow-blue) value. A
change of between 0.3 and 0.4 b* units was measured for all acrylic emulsion
paints, but this was well within the lower limit of perceptible change (often
considered ∆b* = 1.0) and negligible compared to all other paint types, some of
which exhibited a ∆b* = 14.0. On this scale it was not possible to differentiate
between the various acrylic brands, hence the thick black band along the bottom
of the graph, although it was noticed that Maimeri white emulsion paint, based on

100000
carbon detected in extract

80000

60000

40000

20000 Crimson
Crimson 16
Crimson 32
0
Yellow ochre
0 10 20 30 40 50 60
Yellow ochre 16
minutes
Yellow ochre 32

Figure 7. Carbon concentration in water extracts (arbitrary scale)


after immersion of Cryla acrylic emulsion paints
918 ICOM COMMITTEE FOR CONSERVATION, 2002 VOL II

14

12

10
W&N Cremnitz w hite oil
Old Holland Ti w hite oil

8 Row ney Ti w hite Georgian oil


W&N Ti w hite oil

delta b*
W&N Griffin Ti w hite alkyd

6 Row ney Ti w hite artists' oil


Grumbacher MAX Ti w hite oil
Plaka w hite casein

4 L&B Flashe Ti w hite PVA


Maimeri Ti w hite acrylic
Row ney w hite gouache
Talens Ti w hite acrylic
2
Spectrum Ti w hite acrylic
Lascaux Ti w hite acrylic
Golden Ti w hite acrylic
0
0 200 400 600 Row ney Ti w hite acrylic

Tim e (hours) Liquitex Ti w hite acrylic

Figure 8. Change in ∆b* values of white paints during heat ageing at 60°C

a styrene-acrylic copolymer, was slightly more prone to yellowing than pure


acrylic emulsions. The three most yellowing paints were all based on linseed oil,
compared to the other oils and alkyd paints that all contained safflower oil. The
additional yellowing effect of lead white (in Cremnitz white) was also observed.
The alkyd paint appeared to discolour slightly more than these less-yellowing oils.
The discoloration of Plaka casein paint was probably due to a small linseed oil
content and the yellowing in the Flashe PVA paint was just perceptible to the eye.

Conclusions
Although this was clearly not a comprehensive study on the ageing of acrylic
emulsion paints, some interesting observations were made and several potentially
useful analytical techniques were identified. The paints certainly displayed a high
resistance to yellowing, especially when compared to other paint types, particularly
oils and alkyds. Although an apparent stability to light ageing was also measured,
the techniques used would only have been sensitive to oxidation reactions and not
cross-linking or chain-breaking reactions, either of which could greatly affect
mechanical properties, such as brittleness and strength.
The use of SEC to determine MW distributions before and after ageing was
limited by the incomplete solubility of these paints in solvents. This partial solubility
before ageing is interesting, as it indicates a significant degree of cross-linking to
be already present. Although this could be the result of rapid cross-linking reactions
on drying (i.e. indicative of an unstable polymer), it is possible that these emulsions
are appreciably cross-linked even in the wet state. One of the advantages of
emulsion formulations over solutions is that polymers of extremely high MW and/
or with cross-linked regions – which provide increased toughness and strength –
can be dispersed in water, without affecting the viscosity of the medium and
therefore its usefulness as a binder.
SEC could still be employed on the soluble polymer fraction and when combined
with TGA measurements on the insoluble component, a general correlation was
observed between an increase in insoluble material and a reduction (or no change)
in the MW of the remaining soluble part. The combination of these two techniques
points to a tendency for acrylic emulsion paints to cross-link with light exposure.
However, there were exceptions to this, and all changes appeared small. The effect
may therefore be pigment-dependent, but no definite relationship appeared in
these preliminary tests. It is clear that monitoring mechanical properties would
benefit this study.
The most evident change was the gradual loss of residual non-ionic surfactant
from the paints’ surfaces with exposure to light, as seen with micro ATR and
associated measurements with its water extraction. The latter phenomenon has
important ramifications for assessing changes that might occur to these paints with
aqueous conservation treatments, although it is not clear how substrate-dependent
is this effect.
VOL II Modern materials 919

Acknowledgements
The authors thank Joy Keeney and Herant Khanjian (both of the Getty Conser-
vation Institute) for doing the elemental analysis and FTIR-ATR experiments,
respectively.

References
Chiantore, O, and Lazzari, M, 1996, ‘Characterization of acrylic resins’, International Journal of
Polymer Analysis and Characterization 2, 395–408.
Chiantore, O, Trossarelli, L, and Lazzari, M, 2000, ‘Photooxidative degradation of acrylic and
methacrylic polymers’, Polymer 41, 6447–6458.
Chiantore, O, Scalarone, D, and Learner, T J S, 2001, ‘Characterization of acrylic paints used
as artists’ materials’, Int. J. Polym. Anal. Characterization, in press.
De Witte, E, Florquin, S, and Goessens-Landrie, M, 1984, ‘Influence of the modification of
dispersions on film properties’ in N S Brommelle, E M Pye, P Smith, and G Thomson (eds.),
Adhesives and Consolidants, preprints of contributions to the Paris Congress. IIC, London
32–35.
Feller, R L, Curran, M, and Bailie, C W, 1981, ‘Photochemical studies of methacrylate coatings
for the conservation of museum objects’ in S P Pappas and F H Winslow (eds.), Photodegradation
and photostabilization of coatings, ACS symposium series 151, American Chemical Society,
Washington D.C., 183–196.
Hochheiser, S, 1986, Rohm and Haas: History of a Chemical Company, University of Pennsylvania
Press.
Howells, R, Burnstock, A, Hedley, G, and Hackney, S, ‘Polymer dispersions artificially aged’
in N S Brommelle et al. (eds), op. cit., 36–43.
Learner, T J S, 2001, ‘The analysis of synthetic paints by pyrolysis-gas chromatography-mass
spectrometry (PyGCMS)’, Studies in Conservation 46, in press.
Learner, T J S, 2000, ‘A review of synthetic binding media in twentieth-century paints’, The
Conservator 24, 96–103.
Melo, M, Bracci, S, Camaiti, M, Chiantore, O, and Piacenti, F, 1999, ‘Photodegradation of
acrylic resins used in the conservation of stone’, Polymer Degradation and Stability 66, 23–30.
Whitmore, P, and Colaluca, V, 1995, ‘The natural and accelerated aging of an acrylic artists’
medium’, Studies in Conservation 40, 51–64.
Whitmore, P, Colaluca, V, and Farrell, E, 1996, ‘A note on the origin of turbidity in films of an
artists’ acrylic paint medium’, Studies in Conservation 41, 250–255.

You might also like