Sie sind auf Seite 1von 13

Microchemical Journal 124 (2016) 949–961

Contents lists available at ScienceDirect

Microchemical Journal

journal homepage: www.elsevier.com/locate/microc

Accelerated UV ageing studies of acrylic, alkyd, and polyvinyl acetate


paints: Influence of inorganic pigments
Valentina Pintus a,1, Shuya Wei b, Manfred Schreiner a
a
Institute of Science and Technology in Art, Academy of Fine Arts, Schillerplatz 3, A-1010 Vienna, Austria
b
Institute of Historical Metallurgy and Materials, University of Science and Technology Beijing, 30 Xueyuan Road, 100083 Beijing, China

a r t i c l e i n f o a b s t r a c t

Article history: The stability of two types of acrylic binding media, alkyd and polyvinyl acetate (PVAc), four widely used synthetic
Received 26 March 2015 binders in modern and contemporary art, to UV light also including the UV-B range (315–280 nm, middle UV) for
Received in revised form 25 June 2015 simulating sunlight outdoor conditions was studied and compared by double-shot and single-shot Py-GC/MS,
Accepted 9 July 2015
FTIR-ATR, and colour measurements. Thermally assisted hydrolysis and methylation (THM-GC/MS) analyses
Available online 21 July 2015
were used for the alkyd. Additionally, the influence of inorganic pigments on the photo-oxidative stability of
Keywords:
the binding media was also considered. For this purpose, the binders in their pure form as well as mixed with
Py-GC/MS eight different inorganic pigments (titanium white—anatase and rutile, cadmium yellow, cadmium red, hydrated
FTIR-ATR chromium oxide green, ultramarine blue, raw umber Cyprus, and ivory black) were exposed to the accelerating
Colour measurements artificial UV ageing for different periods of time and analysed before and after UV exposure. After UV ageing, the
Acrylic double-shot Py-GC/MS detected the effect of photo-oxidative processes of the binders in a much more detailed
Alkyd way than by the single shot. For instance, photo-oxidation of the acrylics resulted in a production of oligomers
Polyvinyl acetate even during the thermal desorption step of double-shot Py-GC/MS and in a decrease of the EA and nBA main
monomers in the pyrolysis second step. Contrarily, the aged alkyd samples were mostly characterised by a de-
crease of unsaturated fatty acids and especially by the increase of free ortho-phtalic acid detected by double-
shot Py-GC/MS and THM-GC/MS, which is also reflected in the FTIR-ATR results. On the other hand, an increase
of free acetic acid was observed for the aged polyvinyl acetate by double-shot Py-GC/MS. Colour measurements
recorded a greater sensitivity of the alkyd paints, shown by to the bigger shift of L*, a*, and b* coordinates and
change of E* values. Of additional interest was the higher sensitivity to the UV light for synthetic binding
media in combination with pigments.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction cross-linking reactions when the alkyl side groups are short [4].
On the other hand, the photo-oxidation of alkyds is similar to that
Currently, the binding media of most paints consist of synthetic ma- of oil paints because of their similar chemical composition based
terials such as acrylic, alkyd, and polyvinyl acetates, resulting in a wide on the addition of monobasic fatty acids into the polyester
range of different types with properties, and as a consequence, the ma- conformation—formed by a polyhydric alcohol and a polybasic carbox-
jority of modern and contemporary art is composed of synthetic mate- ylic acid. Auto-oxidation of the alkyd can continue with excessive de-
rials. Therefore, the investigation of the factors influencing their gree of cross-linking producing a very stiff and brittle material, chain
ageing behaviour is becoming more and more important in heritage scission with prevalent β-scissions, loss of volatile products such as al-
science. The stability of these binders to UV light in outdoor conditions dehydes, alcohols and carboxylic acids and fading (yellowing) [5]. The
is particularly important. UV ageing studies of synthetic paint materials fragments formed by the β-scission reactions can be converted into
have been carried out in detail in the last 20 years [1–3], mainly consid- free, low molecular weight compounds that can stay in the paint film
ering indoor conditions that simulate parameters relevant for art ob- or can remain in the network, through other cross-links along the chains
jects housed in museums, such as wavelengths between 400 and or can be lost by evaporation or by solvent treatment [6]. The photo-
315 nm (UV-A, near UV). Acrylic binders based on an emulsion and oxidative degradation of polyvinyl acetates (PVAc), which are formed
formed by the four basic components, water, monomer, initiator, and during the polymerisation of vinyl acetate, follows similar mechanism
surfactant, are prone to photo-oxidation mostly by chain scission reac- steps as those of acrylic polymers, which through chain scission and
tions and cross-linking. The chain scissions tended to prevail over the cross-linking reactions produces a series of reactive intermediates and
radicals [7].
E-mail address: v.pintus@akbild.ac.at (V. Pintus). In contrast to studies based on the photo-oxidation of these synthet-
1
Tel.: +43 1 58816 8680; fax: +43 1 58816 8699. ic materials induced by the UV light in indoor conditions, the influence

http://dx.doi.org/10.1016/j.microc.2015.07.009
0026-265X/© 2015 Elsevier B.V. All rights reserved.
950 V. Pintus et al. / Microchemical Journal 124 (2016) 949–961

of UV-B light (315–280 nm, middle UV) on the stability/degradation of 10–20 μm. Additionally, different mock-ups of pure Plextol® D498 and
materials in artworks exposed to outdoor conditions has been inves- Primal® AC33 mixed with inorganic pigments (Table 2) such as
tigated only rarely. UV-B radiation may cause discolouration, cracks, titanium white (anatase and rutile), cadmium yellow, cadmium red,
and other damage to synthetic materials while the UV-A (400–315 nm, hydrated chromium oxide green, ultramarine blue, raw umber Cyprus,
near UV) is usually considered less harmful to polymers. UV-C (280– and ivory black were prepared. Each mock-up was made by mixing
100 nm, far UV) which could cause chain scissions and/or cross- with a paint brush the binder and the pigment in a mixing ratio of
linking of the chemical structure of the polymers is absorbed by the about 3:1. Once a proper paste consistency was obtained, the paint
atmosphere. Furthermore, volatile compounds, which can be formed was cast on glass slides. The dried film thickness of these paint samples
during photo-oxidative reactions and/or any changes, are likewise rare- was approximately in the range of 30–40 μm. In total, 12 identical spec-
ly studied by Py-GC/MS, especially in the double-shot mode [8]. This imens for each pure acrylic binding medium and three for each pigment
technique represents a useful and powerful method for the identifica- mixed with an acrylic emulsion were made in order to obtain sufficient
tion and characterisation of volatile compounds, polymeric materials, samples for the tests in the UV chamber. After drying the samples for
and additives [8]. 24 hours at room temperature, the glass plates were arranged in the
Another issue is that the photo-oxidative stability of synthetic UV chamber, except for unaged reference samples.
binders can be influenced by the addition of pigments into the system. The alkyd Medium 4 (Lukas®, Dr. Fr. Schoenfeld GmbH & Co.,
They may either have a protective effect by absorbing and/or screening Germany) (Table 1), was cast on three glass plates, producing a dried
the UV light or they may be photo-active and therefore catalyse or accel- film thickness in a range of 10–20 μm. Alike for the acrylics, the alkyd
erate the photo-degradation of the polymer [9]. The mechanism by oil binding medium was mixed with the selected inorganic pigments
which pigments may act as photo-sensitisers is not well understood, and consequently cast onto glass slides (Table 2). In order to obtain
and detailed data on the influence of inorganic pigments on the stability enough samples for the consequent exposure in the UV chamber,
of synthetic paint films are still lacking. Additionally, synthetic paints three identical samples for each pigment mixed with the alkyd binding
are characterised by other components that can also contribute signifi- medium were made. Similar to the acrylic samples, pure alkyd binding
cantly to their degradation under UV. Thus UV ageing studies of pure medium and their mixtures with pigments were left dry for 24 hours
synthetic binding media does not accurately reflect the UV ageing of at room temperature.
synthetic paints and the exact context of the degradation still remains Polyvinyl acetate Mowilith® 50 (Kremer Pigmente GmbH & Co. KG,
unclear in many cases. Therefore, more studies concerning the influence Germany) was chosen (Table 1) for making specimens, which was
of inorganic pigments on synthetic binding media are desirable. dissolved in acetone, cast on glass slides and then allowed to dry at
Partially based on previous studies [10–12], the aim of this work is to ambient conditions (average thickness of the paint film was about
investigate the stability when exposed to UV light including the UV-B 10–40 μm). The selected inorganic pigments used for the acrylic
range for simulating sunlight outdoor conditions of four of the most mock-ups were mixed with the polyvinyl acetate Mowilith® 50 to pre-
widely used binding media in modern and contemporary art, two differ- pare mock-up samples on glass slides (Table 2) and left to dry at room
ent types acrylic, alkyd, and polyvinyl acetate (PVAc), as well as the ef- temperature.
fect of different inorganic pigments on their chemical behaviour. For The inorganic pigments used for all mock-ups are products of
this purpose, two types of acrylic binding media, alkyd and polyvinyl Kremer except the white pigments, which are part of the material col-
acetate (PVAc) both pure and also mixed in the laboratory with eight lection of the ISTA (Institute for Science and Technology in Art) at the
different inorganic pigments (titanium white—anatase and rutile, Academy of Fine Arts Vienna.
cadmium yellow, cadmium red, hydrated chromium oxide green, ultra-
marine blue, raw umber Cyprus, and ivory black) were analysed before 2.2. UV exposure
and after UV exposure by the double-shot of Py-GC/MS, single-shot of
Py-GC/MS while thermally assisted hydrolysis and methylation (THM- UV exposure of the samples was carried out in a UVACUBE SOL 2/
GC/MS) analyses were used for the alkyd, FTIR-ATR, and colour mea- 400F UV chamber, produced by Dr. Hönle GmbH UV-Technology,
surements. A comparison between the double-shot and single-shot Germany. The UV light radiation source was supplied by a 910 W/m2
techniques of Py-GC/MS for their application for UV ageing studies of Xenon arc solar simulator with an incorporated H2 filter, which pro-
modern paint materials was also done. vides radiation with wavelengths between 295 and 3000 nm thus sim-
ulating sunlight outdoor conditions. The chamber temperature was
2. Experimental 48.8 °C. No control of the relative humidity (RH) was possible in the
UV exposure chamber used, therefore the RH varied between 30% and
2.1. Sample preparation 35% depending on the RH of the ambient atmosphere. The accelerated
UV exposure of the acrylic and alkyd samples was carried out for 31
The investigated binders and mock-up samples are listed in Tables 1 and 83 days, respectively, while the polyvinyl acetates were UV aged
and 2, respectively. Approximately 60 mg of two pure acrylic binding for 60 days. According to the ASTM 2565 – 99 standard [13], the
media (Plextol® D498 and Primal® AC33 purchased from Kremer exposure time of the specimens under UV light was established from a
Pigmente GmbH & Co. KG, Germany) (Table 1) were cast separately long-term evaluation and control of the material. This evaluation
on glass plates, which produced a dried film thickness in a range of was periodically carried out under an optical microscope to observe

Table 1
List of the binders investigated and their chemical and technical properties.

Binders

Commercial name Company Type of binder Composition Product number

Plextol® D498 Kremer® Pigmente GmbH & Co. KG, Germany Acryl Acqueous dispersion of a thermoplastic acrylic polymer based on: 76000
poly(n-butyl acrylate / methyl methacrylate), p(nBA/MMA)
Primal® AC33 Kremer® Pigmente GmbH & Co. KG, Germany Acryl Acqueous dispersion of a thermoplastic acrylic polymer based on: 75200
poly(ethyl acrylate / methyl methacrylate), p(EA/MMA)
Medium 4 Lukas®, Dr. Fr. Schoenfeld GmbH & Co., Germany Alkyd Solution of alkyd resin in mineral spirit 2224
Mowilith® 50 Kremer® Pigmente GmbH & Co. KG, Germany PVAc Polyvinyl acetate in grainy state 67040
V. Pintus et al. / Microchemical Journal 124 (2016) 949–961 951

Table 2
List of the mock-up samples investigated and their chemical and technical properties.

Mock–up samples

Binder Chemical Constitution


Pigment Color index:
Color composition of number of
classical name generic name
Acryl: Acryl: pigment pigment
Alkyd: PVAc:
Plextol® Primal®
Medium 4 Mowilith®50
D498 AC33

Titanium white
White anatase TiO2 PW 6 69

Titanium white
White rutile TiO2 PW 6 54

Yellow Cadmium yellow CdS PY 37 21120

Red Cadmium red CdS, xCdSe PR 108 21060

Hydrated
Green chromium oxide Cr2O3•H2O PG 18 44250
green

Blue Ultramarine blue Na8Al6Si6O24•Sx PB 29 45010

Raw umber
Brown Cyprus Fe2O3•H2O, MnO2 PBr 8 40610

Black Bone (ivory) black C, Ca3(PO4)2, CaCO3 PBk 9 47150

any change in the morphology of the surface, by Fourier transform at- MS (5% diphenyl / 95% dimethyl siloxane) capillary column with a
tenuated total reflection (FTIR-ATR) analysis (Section 3.3) and colour 0.25 mm internal diameter, 0.25 μm film thickness, and 30 m length
measurements (Section 3.4). (Frontier Laboratories, Japan) was used for the analyses of the alkyds
By using the Hönle UV-Meter (Dr. Hönle GmbH UV-Technology, and polyvinyl acetates afterwards. NIST 05 and NIST 05s Library of
Germany), it was observed that during the UV ageing of the alkyd sam- Mass Spectra were used for the identification of the compounds. The
ples, the UV radiation intensity of the xenon lamp reached a value of ap- GC column temperature conditions used for the acrylics and polyvinyl
proximately 177 W/m2, thus exposing the samples to a more gentle acetates were as follows: initial temperature 40 °C, held for 5 min
ageing than the acrylic and polyvinyl acetate specimens. Unfortunately, followed by a temperature increase of 10 °C/min to 292 °C. For the
the extended use of UV lamps results in a gradual decrease of the inten- alkyds, the oven initial temperature was set at 40 °C for 5 min, and
sity over time, which has to be taken into account for the UV ageing then followed with a gradient of 6 °C/min up to 280 °C for 10 min.
studies. The helium gas flow was set at 1 mL/min and mass spectra were record-
ed under electron impact ionisation at 70 eV. For the single-shot analy-
2.3. Pyrolysis gas chromatography mass spectrometry (Py-GC/MS) ses, the pyrolysis temperature was set at 600 °C and held for 12 s.
The double-shot parameters were as follows: For the thermal de-
For the UV-B ageing studies, all samples were analysed before and sorption of the pure acrylic binding media and their mixtures with pig-
after UV exposure with double-shot and single-shot Py-GC/MS. ments, the temperature was set at 50 °C, held for 2 min and increased by
Double-shot Py-GC/MS is based on a two-step analysis: 1) Thermal de- 20 °C/min to 300 °C and held there for 2 min while for the alkyds the
sorption of the samples at lower temperature to detect volatile com- temperature was set at 50 °C and increased by 20 °C/min to 250 °C. Ad-
pounds and 2) pyrolysis of the same sample as second step. ditionally, for the polyvinyl acetates, the temperature was set at 100 °C,
Between 80 and 300 μg of the samples were scraped from the glass held for 2 min and increased by 20 °C/min to 250 °C and held there for
plates with a scalpel and put in a sample cup (ECO-CUP Frontier Lab, 2 min. The second-step pyrolysis was carried out for all samples at
Japan) for analysis. Unlike acrylic and the polyvinyl acetate paints, about 600 °C and held for 12 s.
100 μg of alkyd samples were treated with 2 μL tetramethylammonium
hydroxide (TMAH) reagent (25 wt% aqueous solution of TMAH, Sigma-
Aldrich, USA) in the sample cups in order to perform thermally assisted 2.4. Fourier transform infrared spectroscopy-attenuated total reflectance
hydrolysis and methylation (THM-GC/MS) analysis. Py-GC/MS and (FTIR-ATR)
THM-GC/MS analyses were performed with a PY-2020iD (Frontier Lab,
Japan) pyrolyzer unit combined with a GCMS-QP2010 Plus (Shimadzu, FTIR-ATR analyses were performed with an Alpha FT-IR Platinum
Japan). For the analyses of the pure acrylic binding media as well ATR instrument (Bruker Optics, Germany) equipped with a deuterated
as those mixed with pigments, the GC/MS unit was equipped with a triglicine sulphate detector (DTGS) and with a diamond crystal. Spectra
capillary column SLB-5ms SUPELCO, USA (30 m length × 0.25 mm were acquired in a spectral range between 4000 and 370 cm− 1
internal diameter × 0.25 μm film thickness) using bonded and performing 64 scans at 4 cm−1 resolution. The resulting spectra were
highly cross-linked 5% diphenyl / 95% dimethyl siloxane. Due to the collected and evaluated with the spectrum software OPUS® of Bruker
change of the column after the initial measurements, the Ultra Alloy-5 Optics, Germany.
952 V. Pintus et al. / Microchemical Journal 124 (2016) 949–961

The infrared absorptions bands obtained for the investigated pig- generally formed under pyrolysis conditions in a lesser amount. On
ments were identified by comparing the acquired spectrum of the pig- the other hand, the pyrolysis products of the polyvinyl acetate recorded
ment sample with the reference spectrum of the IRUG (Infrared in the single-shot mode of the Py-GC/MS and shown in Fig. 1c are main-
Raman Users Group) database. ly based on acetone (m/z = 43, 58), 1-3 cyclopentadiene (m/z = 39, 66),
acetic acid (m/z = 43, 60), benzene (m/z = 52, 78), and toluene (m/z =
2.5. Colour measurements 65, 91) [12]. The thermal desorption step of the double-shot mode
allowed the detection of the diethyl phthalate (DEP) type of plasticiser
A SPM50 (Gretag-Macbeth AG, Switzerland) instrument was used to usually used in PVAc (Fig. 2c). The detection of the Di-tert-butyl
obtain spectra in the visible range. Reflectance measurements were car- dicarbonate (DTBD) type of reagent is probably due to the impurity of
ried out using a D65 lamp in the range of 380–730 nm. Reflection was the PVAc product, as it is also the case for octanone (Fig. 2c).
measured relative to the white standard of the instrument and a 10° The main components of the alkyd binder—detected by thermally
Standard Observer was used. For the UV ageing studies, the total colour assisted hydrolysis and methylation (THM-GC/MS) analysis—are
values (ΔE*) are obtained according to the Commission Internationale benzoic acid, methyl ester at RT 17.1 min (m/z = 51, 77, 105, 136),
de l’Eclairage (CIE) 2000. trimethyl ether of pentaerythritol at RT 19.7 min (m/z = 55, 69, 71,
75, 85, 101, 114, 128), phthalic acid, dimethyl ester at RT 26.1 min
3. Results and discussion (m/z = 77, 92, 133, 163, 194), and the fatty acids from oil such as
methylated fatty acid peaks related to azelaic acid (2C9:0) at RT
The comparison between the results obtained from unaged and aged 27.7 min, palmitic acid (C16:0) at RT 34.6 min, oleic acid (C18:1) at
samples reveals that the effects of the photo-oxidative reactions were RT 37.4 min, stearic acid (C18:0) at RT 37.8 min, and linoleic acid
most determinable after 83 days of UV exposure. Therefore, the compar- (C18:2) at RT 38 min. Furthermore, the intermediate oxidation products
ison between the results acquired was made between the unaged and of 2-octenal, nonanal, 2-decenal, 2,4 decadienal, 2-undecenal and the
83-day aged samples. dryer compound 2-ethylhexanoic acid were identified in the thermal
desorption step of the double-shot mode of Py-GC/MS (Fig. 2b [11]).
3.1. Impact of UV ageing to the paints After 83 days of UV ageing, the main evidence of photo-oxidative
processes on both acrylic binders observed in the thermal desorption
The visual appearance of paint colours is a fundamental characteris- pyrogram of the double-shot techniques are the peaks of the
tic for the evaluation of their state of preservation. Any variation of the sesquimers, dimers, and trimers, which were not registered in the
material properties to the influence of UV light can correspond to a unaged material (Fig. 2a). The thermal stability of the samples has evi-
change of colour, opacity, and brittleness. Thus, a systematic investiga- dently been changed by the photo-ageing processes leading to the pro-
tion of these chemical changes is an important point of study for the duction of oligomers already detectable during thermal desorption.
conservation and restoration of paints. Additionally, the most intense peak of n-butyl acrylate (BA) of Plextol®
After the ageing periods, some of the samples showed a slight colour D498 and of ethyl acrylate (EA) of Primal® AC33 in comparison to the
change and different opacity. The colour and opacity of blue mock-ups methyl methacrylate (MMA) peak decreased gradually after UV ageing,
were the most affected by UV light. By scraping samples from the whereas the MMA peak became the most intense one (Fig. 1a). This re-
aged films for analysis, it was observed that they had become more frag- duction in intensity of the EA and BA peaks shows a higher sensitivity to
ile and brittle compared to unaged samples. These changes in their mac- photo-oxidation reactions of these monomers compared to a monomer
roscopic properties can be considered characteristic for UV ageing and with lower structural units such as MMA. A further indication for a
can be related to their chemical changes described in the next three photo-oxidation process concerns the octyl phenol polyethoxy ethanol
Sections 3.2, 3.3, and 3.4. surfactant. No traces of octyl phenol previously detected in the thermal
desorption step in the double-shot mode were found anymore in the
3.2. Py-GC/MS analysis of unaged and aged samples pyrogram of the aged Plextol® D498 and Primal® AC33 by plotting
the corresponding single ion mass profile at m/z = 135 [10].
In order to investigate the photo-oxidative deterioration of the sam- After 83 days of UV ageing, one of the main signs of photo-oxidative
ples when exposed to UV light, each of the unaged and aged acrylic and processes of the alkyd binder is shown by a decrease of unsaturated
polyvinyl acetate (PVAc) binder sample types were analysed by either fatty acids such as oleic (C18:1) and linoleic acid (C18:2) followed by a
the single-shot or double-shot mode of the Py-GC/MS while the ther- decrease of oxidation products (cis-9-hexadecenal, cis-9-octadecenal,
mally assisted hydrolysis and methylation (THM-GC/MS) and the and octadecanal) from unsaturated fatty acids and increasing of short
double-shot mode of Py-GC/MS were used for the analysis of the unaged chain products (1-octene and hexanal). Additionally, an increase of dicar-
and aged alkyd samples. boxylic acid such as azelaic acid (2C9:0) and suberic acid (2C8:0) was also
The pyrograms of the binders that had been mixed with pigments detected. These results obtained by THM-GC/MS and second-step pyroly-
did not display any difference in the Py-GC/MS results in comparison sis of double-shot of Py-GC/MS (Fig. 1b [11]) suggest oxidation, cross-
to the pyrograms obtained for the pure binders. The Py-GC/MS results linking, and chain scission of unsaturated fatty acids, meanwhile dicar-
obtained from both aged acrylic binding media mixed with pigment, boxylic acid are produced [6,14]. Other indications of the photo-
as well as those of the aged alkyd and PVAc mock-ups in comparison oxidation of the alkyd binder are given by the total decrease of
with the respective aged binders also did not show any change or for- the group of aldehydes (2-octenal, nonanal, 2-decenal, 2,4-decadienal,
mation of new peaks, apart a more pronounced alteration of the binders 2-undecenal) and the dryer compound (2-ethylhexanoic acid) priory
themselves described in the following section. detected in the unaged binder by the thermal desorption step of the
The results achieved with the second-step pyrolysis of double-shot double-shot of Py-GC/MS (Fig. 2b), probably due to further oxidation
Py-GC/MS and single-shot Py-GC/MS show that the Plextol® D498 is or cross-linking of such products.
based on a co-polymer of n-butyl acrylate (nBA) and methyl methacry- After UV ageing, it was possible to observe an increase of the phthalic
late (MMA) while the Primal® AC33 is characterised by a co-polymer of anhydride (PhA) peak in the thermal desorption step of the double-shot
ethyl acrylate (EA) and methyl methacrylate (MMA) [10]. In addition to of Py-GC/MS (Fig. 2b). By calculating the peak area ratio of the phthalic
the peaks of the main monomers, the pyrograms of each acrylic binder anhydride in the thermal desorption step (PhA1) and in the pyrolysis
are also characterised by the formation of several and different oligo- step (PhA2), it was possible to see an increase of the PhA1/PhA2 in
mers (sesquimers, dimers, and trimers) at higher retention time (RT) the aged alkyd, which indicates that more free ortho-phthalic acid was
[10]. Oligomers are molecules consisting of few monomer units and formed during UV ageing [11]. The main pyrolysis product phthalic
V. Pintus et al. / Microchemical Journal 124 (2016) 949–961 953

Fig. 1. Pyrograms of unaged and aged acrylic Primal® AC33 mixed with ultramarine blue (PB29) (a), alkyd (b) [11], and polyvinyl acetate (PVAc) mixed cadmium yellow (PY37) (c), ob-
tained with the pyrolysis step of double shot Py-GC/MS. (Peaks: 1) Benzene, 2) 1-Octene, 3) Hexanal, 4) 2-Methyl-1-pentanol, 5) Heptanal, 6) Benzaldehyde, 7) 1-Decene, 8) Octanal,
9) Octanoic acid, 10) Benzoic acid, 11) Phthalic anhydride, 12) 1 (3H)-Isobenzofuranone, 13) Pentadecane, 14) Benzeneacetic acid ethyl ester, 15) trans-9-Octadecenal, 16) Pentadecane,
17) Tetradecanal, 18) n-Hexadecanoic acid, 19) cis-9-Hexadecenal, 20) cis-9-Octadecenal, 21) Octadecanal, 22) 2-oxo-octadecanoic acid methyl ester, 23) 1,3-Benzenediol, monobenzoate,
24) trans-2-Hexenyl benzoate [11]).

anhydride of the alkyd is generally formed by the cleavage of a C\\O oxidised by Norrish I photo-cleavage as main initiation step of photo-
bond during thermal decomposition of the alkyds based on ortho- degradation [15–19]. Scheme 1 shows 2 different possibilities of cleav-
phthalic acid and the rise of phthalic anhydride indicates that free age of the ester group creating 4 different primary radicals. The mecha-
ortho-phthalic acid was formed during UV exposure. Therefore, the in- nism of photo-cleavage indicated in path I illustrates the formation of
crease of PhA1/PhA2 detected by the double-shot Py-GS/MS in the carboxyl and alkyl radicals (Ia). The next step (Ib) is based on the ab-
UV-aged alkyd may suggest that the carbonyl phthalic ester undergoes straction of a hydrogen atom by the polymer carboxy radical from the
photo-oxidation reactions through Norrish type I reaction. It has been polymer chain to yield a phthalic acid end group, which under the
reported that aromatic polyesters are prone to photolysis and photo- same mechanism of photolysis and photo-oxidation pathways of Ia)
954 V. Pintus et al. / Microchemical Journal 124 (2016) 949–961

Fig. 2. Pyrograms of unaged and aged acrylic Primal® AC33 mixed with ultramarine blue (PB29) (a), alkyd (b), [11], and polyvinyl acetate (PVAc) mixed with cadmium yellow (PY37) (c),
obtained with the thermal desorption step of double shot Py-GC/MS.

and Ib) may form free ortho-phthalic acid (1c and 1d). Likely the and abstraction of hydrogen (IIb), cleavage (IIc), and again abstraction
mechanisms of photo-cleavage of path I, the path II displays the final of hydrogen (IId).
formation of free ortho-phtahlic acid (IId) through the initial The photo-oxidation process also affected the chemical stability of
production of acyl and alkoxy radicals (IIa) followed by oxidation the polyvinyl acetate (PVAc) binder. The amount of acetic acid—which
V. Pintus et al. / Microchemical Journal 124 (2016) 949–961 955

was detected in the aged PVAc with both methods of Py-GC/MS—was The FTIR-ATR results of the unaged and aged samples are summarised
higher than in the unaged ones indicating the degradation of the film below and arranged according to the corresponding sample types,
by the UV ageing procedure (Figs. 1c and 2c). The increase in intensity i.e. a) binders and b) mock-ups.
of the peak related to the acetic acid shows a growth of free acetic
acid, which is in agreement with the Norrish type II mechanism based
on the formation of an excited carbonyl group following absorption of 3.3.1. Binders
light [20,21]. Furthermore, a decrease of the peak related to the DEP By considering the following infrared regions, it is possible to de-
type of plasticiser in the binder was observed after UV ageing (Fig. 2c), scribe the UV degradation of all investigated binders:
according to the slow evaporation of such plasticisers, resulting in em- O\\H stretching (3,650–3,100 cm−1): A gradual broadening in this
brittlement and degradation of the paint film [22,23]. infrared region, probably due to the formation of alcohol groups and/
or hydroperoxidic structures [3], was shown by the UV-aged acrylic
Plextol® D498 and Primal® AC33 binders (Fig. 3a). The acrylic samples
3.3. FTIR-ATR analysis of unaged and aged samples showed a continuing increase in absorption as well as a band broaden-
ing of the OH groups with a maximum peak around 3,495 cm−1 as was
The FTIR-ATR absorption bands of both acrylics—Plextol® D498 and likewise observed in the alkyd resin after the UV exposure (Fig. 3c). The
Primal® AC33—alkyd and polyvinyl acetates binders are summarised in changing of this infrared region is mostly due to the formation of new
Table 3. Some of the selected pigments used for preparing the mock-ups alcohol groups along the fatty acid portion through either ß-scissions
show characteristic IR absorption bands, which in some cases overlap or Norrish type I reactions during the photo-oxidation degradation in
the absorption bands of the binders. For instance, the C\\O/C\\C skeletal a similar way to the natural and photo ageing of the binder simulating
vibrations of the binders (1,250–900 cm−1) are masked by the ultrama- indoor conditions [24–26].
rine blue pigment by the overlapping Al, Si\\O4 asymmetric stretching The unaged and UV-aged polyvinyl acetate (PVAc) binder and mock-
bands at 1,069, 986 with a weak shoulder at 1,095 cm−1. Similarly to ul- ups did not show any type of absorption in the O\\H stretching region
tramarine blue pigment, bone black pigment (C, Ca3(PO4)2, CaCO3) masks between 3,650 and 3,100 (Fig. 4c).
the C\\O/C\\C infrared region of the binders with a strong peak at C\\H stretching (3,100–2,800 cm−1): After 83 days of UV ageing, the
1,020 cm−1 and two less intense peaks at 1,086 and 960 cm−1 of the sharp peak at 2,890 cm−1 of polyethoxylated surfactant of both acrylics
phosphate groups (PO3− 4 ). Moreover, the raw umber Cyprus pigment was no longer present according to the Py-GC/MS results (Fig. 3a).
based on iron oxide Fe2O3 with water manganese oxide MnO2 has a During UV exposure, the methylene groups of the alkyd binder were
band around 1,000 cm−1 of the iron oxides. On the other hand, the cad- prone to diminish possibly through Norrish type I and II reactions [27]
mium red and cadmium yellow pigments do not absorb in the considered or due to the oxidation of double bonds [27,28], which is shown by a
mid-infrared region (4000–370 cm−1). gradual decrease in absorption in the asymmetric and symmetric
Generally, the most prominent differences in the absorption bands of (C\\H)CH2 stretching at 2,926 and 2,855 cm−1, respectively (Fig. 3c).
the unaged and aged binders were obtained for the acrylic and alkyd, The weak band of the aromatic _C\\H stretching at 3,070 cm−1 detect-
while the polyvinyl acetate (PVAc) seemed to be the most stable. Normal- ed in the alkyd binder did not change after 83 days of UV ageing.
ly, the differences between the unaged and aged samples were more By comparing the C\\H stretching region between 2,800 and
prominent when the binders were mixed with the pigments. On the 3,100 cm− 1 of the IR spectra of the unaged and aged PVAc binding
other hand, after 83 days of UV ageing, the acquired infrared spectra of medium, no significant changes in the weak absorption of the doublet
the unaged and aged pigments did not show any remarkable difference. of the –CH3 and –CH2 asymmetric stretching vibration at 2,973 and
2,926 cm−1 were noticed (Fig. 4c).

Table 3
ATR infrared absorptions of the acrylics Plextol® D498 and Primal® AC33, alkyd Medium 4, and polyvinyl acetate Mowilith® 50 binders investigated.

Bond type Acrylic Plextol® D498 Acrylic Primal® AC33 Alkyd Polyvinyl acetate (PVAc)

p(nBA/MMA) p(EA/MMA) Medium 4 Mowilith® 50


−1 −1 −1
(cm ) (cm ) (cm ) (cm−1)

O\\H stretching – – 3,495 –


C\
\H stretching – – 3,070 –
2,955, 2,931, 2,874 2,949, 2,888, 2,860 2,926–2,855 2,973, 2,926
2,852 shoulder – – –
C_O stretching 1,726 1,723 1,720 1,729
C_C stretching – – 1,600, 1,580, 1,507, 1,489 –
C\
\H bending – 1,513 – –
– 1,466–1,451 1,465–1,451 1,433
1,449, 1,436 shoulder – – 1,449, 1,435
1,386 1,381 1,385 1,370
1,361, 1,343 1,360, 1,343 – –
C\
\O and C\
\C stretching – 1,279 – –
1,236, 1,159, 1,143 1,240, 1,149 1,254 1,225, 1,120
– – 1,174, 1,166 –
1,115 1,109 1,114
1,063 1,061 1,068, 1,040 –
1,022 shoulder 1,021 1,026 1,018
990 – – –
963 964 974 –
944 – – –
C\
\H rock 842 843 – 844
808 – – 795
– – 774 –
754 759 – –
– – 740–709 –
956 V. Pintus et al. / Microchemical Journal 124 (2016) 949–961

Fig. 4. FTIR-ATR spectra of the acrylic Plextol® D498 [10], alkyd, and polyvinyl acetate
(PVAc) mixed with ultramarine blue (PB29), (a) (b) and (c), respectively, before (solid
line) and after UV ageing (pointed line).

carboxyl acids, ketones, and aldehydes [24–27]. Particularly, the pro-


Fig. 3. FTIR-ATR spectra of the acrylic Primal® AC33 and alkyd mixed with raw umber gressive absorption at 1,773 cm−1 in the acrylics and between 1,600
Cyprus (PBr8), (a) and (b) respectively, and of the alkyd binder (c) before (solid line) and 1,710 cm− 1 in the acrylics (Figs. 3a and 4a) and alkyd (Figs. 3c
and after UV ageing (pointed line). and 4b) indicates respectively the formation of a γ-lactone structure [3,
4,29] or open chain anhydrides [30] and the growing of unsaturated
C_O stretching (1,750–1,700 cm−1): The difference of the carbonyl molecules with a C_C bond due to the polymer chain scissions [3] or
stretching absorption and broadening band in all four different UV- ketones products as a consequence of the reaction of oxygen molecules
aged binders are very similar. with radicals and followed by ß-scission [31]. In case of the PVAc binder,
The IR spectra of the UV-aged acrylic binders in comparison to the IR the weak broadening of the carbonyl peak in the range between 1,700
spectra of the unaged samples are characterised by a gradual reduction and 1,600 cm−1 (Fig. 4c) suggests the increase of free acetic acid [32]
of the carbonyl group absorption at 1,726 cm−1, which may be related after UV ageing according to the Norrish type II mechanism [33],
to the loss of the butyl group of the Plextol® D498 (Fig. 4a) and ethyl corresponding with the results obtained by the Py-GC/MS analysis
group of the Primal® AC33. Although a loss of an ester group took with the double shot.
place in the alkyd and PVAc binders during UV ageing with a conse- C_C of aromatic ring (1,650–1,450 cm−1): The two characteristic
quent formation of the free ortho-phthalic acid and free acetic acid, sharp and intense peaks at 1,600 and 1,580 cm−1 followed by two less
respectively, the absorption of the carbonyl peak at 1,720 cm−1 in the intense and weak peaks at 1,507 and 1,489 cm−1 of the aromatic C_C
alkyd binder as well as at 1,729 cm−1 in the PVAc binder remained un- stretching of the polybasic acid part of the alkyd did not show any signif-
changed (Figs. 3c and 4c). This is probably due to the compensation of icant change after UV ageing (Fig. 3c).
the loss of the ester groups with some new carboxyl products absorbing C\\H bending, C\\O and C\\C stretching, and C\\H rocking (1,450–
in the same region and formed by reaction of secondary macro radical 700 cm−1): A gradual reduction of the absorption of the typical peaks
with oxygen [4]. of the PEG type surfactant at 1,343, 1,110, and 963 cm − 1 was ob-
The broadening of the carbonyl stretching band of the acrylics and served in both acrylic binders after UV exposure, which agrees with
alkyd shows the formation of newly photo-degraded products such as
V. Pintus et al. / Microchemical Journal 124 (2016) 949–961 957

the Py-GC/MS results in which no detection of any traces of the surfac- at 1,068 and 1,040 cm−1 as well as the aromatic C\\H out of plane bend-
tant found in the aged samples. ing at 774, 740, and 709 cm−1 of the alkyd binder after UV exposure.
Where the doublet of the most intense symmetric CH2 bending Similar to the other infrared regions of the PVAc binder/mock-ups,
at 1,465 cm− 1 and the less intense asymmetric CH3 bending at no noticeable difference could be determined in the wavelength region
1,451 cm−1 was shown in the IR spectrum of the unaged alkyd binder, between 1,450 and 700 cm−1 by comparing the unaged and UV-aged
after UV ageing, the asymmetric CH3 bending of the doublet became the specimens.
most intense one, complemented by an increase in absorption of the
symmetric CH3 bending at 1,385 cm−1 (Figs. 3b,c and 4b). The absorp- 3.3.2. Mock-ups
tion variation of the CH3 to CH2 may be explained by the decrease of the All of the differences in the IR spectra recorded between the unaged
fatty acid chain length and increase of short chain products in the aged and UV-aged binders were more pronounced in the mock-ups, especial-
alkyd binder due to the chain scission degradation of the unsaturated ly when both acrylic binders were mixed with the ultramarine blue pig-
portion of the fatty acid in the alkyd structure. This result agrees with ment and when the alkyd binder was mixed with the ultramarine blue
the higher amount of short chain products detected in the aged alkyd and raw umber Cyprus pigments. Particularly, the absorption bands of
samples by the thermal desorption step of double-shot of Py-GC/MS. A the ultramarine blue at 1,095, 1,067, 988, 688, 653, 580, and 441 cm−1
slight decrease of the C\\O stretching phthalic group at 1,254 and of the raw umber Cyprus at 3,175, 1,016, and 890 cm−1 became
and 1,114 cm− 1 was observed after UV ageing of the alkyd binder more evident in comparison to the absorption bands of the binders
(Figs. 3b.c and 4b). As it has already been reported in the Py-GC/MS (Figs. 3a,b and 4a–c), mostly due to the photo-degradation of the organ-
Section 3.2, the formation of the free ortho-phthalic acids from the car- ic part in the mock-ups. Furthermore, noticeable changes of the absorp-
bonyl phthalic ester (Schema 1) may explain the gradual reduction in tion spectrum of Plextol® D498 mixed with ultramarine blue are shown
absorption of the phthalic C\\O stretching band due to the C\\O bond by the progressively decrease of the C\\O stretching peaks at 1,144 and
scission from the alkyd structure. Furthermore, after UV exposure of 1,063 cm−1 and another peak at 846 cm−1 related to the C\\H rocking
the alkyd binder, it was noticed that the peak at 974 cm−1 of the out as well as the disappearing of the C\\O stretching peaks at 1,159 and
of plane _CH bending of trans-structures gradually became a shoulder 1,114 cm− 1 (Fig. 4a [10]). This effect related to the C\\O stretching
(Fig. 3c), mostly due to the oxidation of the double bonds [27,28] or bands and to the C\\H rocking band may suggest the loss of an ester
possibly through Norrish type I and II reactions [27]. No significant group in agreement with the reduction of C_O absorption and to the
change was noticed for the aromatic C\\H in plane deformation peaks fast oxidative chain scission of the polyethoxylated surfactants when ex-
posed to UV light [34,35].

Schema 1. Suggested Norrish type I photolysis and photo-oxidation pathways of carbonyl phthalic ester of alkyd.
958 V. Pintus et al. / Microchemical Journal 124 (2016) 949–961

In contrast to all of the other pigments, most of the principal absorp- bond in the main chain, no variations were detected in the infrared
tions gradually increased when the cadmium red and particularly the spectra of the UV-aged PVAc mock-ups.
cadmium yellow pigment was mixed with both acrylic binders as well
with the alkyd corresponding to an increase of the ester side groups. 3.4. Colour measurements of unaged and aged samples
Similar to the FTIR-ATR results obtained on the aged polyvinyl ace-
tate (PVAc) binder, no obvious differences could be determined by com- The colour measurement results of all investigated unaged and 83-day
paring the spectra of the unaged and UV-aged PVAc mock-ups. Possibly UV-aged acrylics (Plextol® D498 and Primal® AC33), alkyd, and PVAc
due to the low amount of the acetic acid formed during UV ageing mock-ups and pigments are summarised in Table 4, which includes the
through the Norrish type II mechanism (as shown by the results of the shift in the values of the lightness/darkness (L*), redness/greenness (a*),
Py-GC/MS double-shot method) and complemented by the C_C double yellowness/blueness (b*), and total colour (E*). Additionally, the CIELAB
diagrams shown in Figs. 5 and 6 highlight the pattern change of a* and
b* values after UV ageing, indicating colour changes, while in the left
side of the figure it is possible to see the differences in L* values corre-
Table 4 sponding to the brightening and darkening of the paint films.
Shifts in the L*, a*, b*, and E* coordinates of the acrylic Plextol® D498 (a) and Primal® The ΔL*, Δa*, Δb*, and ΔE* values recorded for the investigated pig-
AC33 (b), alkyd (c), and polyvinyl acetates (d) mock-ups as well as of the pigments
ments show a slight difference in the shift of the L*, a*, and b* coordi-
(e) before and after UV ageing.
nates as well change colour (Table 4). The general behaviour of most
a) of the analysed pigments was given by a slight increase in L* values
Acrylic mock-up (Plextol® D498) ΔL* Δa* Δb* ΔE* 2000 corresponding to a brightening of the pigment during the UV exposure.
Titanium white anatase 1.54 0.41 −0.41 1.20
Additionally, the cadmium yellow pigment had the greatest ΔE* value
Titanium white rutile −0.49 0.12 −0.472 0.56 (ΔE* = 5.58) based on the shift of L* of 6.43 and of a* and b* of −6.75
Cadmium yellow −2.37 −3.56 −6.8 2.54 and −1.71, respectively.
Cadmium red 1.83 −0.68 −4.9 2.63 Generally, the differences in the shift of the L*, a*, and b* coordinates
Hydrated chromium oxide green 3.01 6.11 0.55 3.63
between the unaged and UV-aged mock-ups are more prominent in the
Ultramarine blue −5.71 1.76 0.98 4.23
Raw umber Cyprus −0.54 −2.43 −3.87 3.68 case of the pigments mixed with the alkyd binder, which shows this
Bone (ivory) black) −0.76 0.07 0.22 0.58 binder to be least stable when exposed to UV light. In particular, the ul-
tramarine blue mock-up, and to a slightly lesser extent, the raw umber
b) Cyprus mock-up, had the greatest shift of a* (Δa* = − 28.0) and b*
Acrylic mock-up (Primal® AC33) ΔL* Δa* Δb* ΔE* 2000 (Δb* = 37.36) between unaged and aged samples showing a strong re-
Titanium white anatase 2.31 0.44 0.32 1.59 duction in red and blue, respectively. The ultramarine blue alkyd mock-
Titanium white rutile 0.83 0.11 −0.2 0.57 up had the highest shift of L* (ΔL* = 13.12) towards the brightening.
Cadmium yellow −1.28 −1.59 −9.03 2.00 The brightening of the surface could be explained by smoothing of the
Cadmium red 0.64 −1.46 −2.96 1.22
film surface during UV exposure [4].
Hydrated chromium oxide green −6.16 0.00 1.54 4.78
Ultramarine blue 3.07 0.18 −3.09 2.65
The acrylic Plextol® D489 and Primal® AC33 mock-ups showed sim-
Raw umber cyprus 0.8 −1.64 −2.71 2.56 ilar colour change results, where the ultramarine blue, raw umber
Bone (ivory) black) −1.67 0.05 0.34 1.20 Cyprus, and the hydrated chromium oxide green had the greatest shift
in the values of a*, b*, and L*. In contrast, the colour results for the
c)
PVAc mock-ups changed substantially less than for the alkyd and both
Alkyd mock-up ΔL* Δa* Δb* ΔE* 2000 acrylic mock-ups after UV exposure. This latter result agrees with the
Titanium white anatase −4.02 −0.43 −2.53 3.33 FTIR-ATR results.
Titanium white rutile 2.08 0.45 2.26 2.57 Additionally, all four type of binders mixed with the titanium white
Cadmium yellow −0.86 −1.68 2.70 1.44 anatase were shown to be slightly more sensitive to the UV light than
Cadmium red 3.67 −6.81 −7.23 4.50
Hydrated chromium oxide green −4.08 4.37 2.48 4.70
the titanium white rutile according to the more pronounced difference
Ultramarine blue 13.12 −28.00 37.36 16.01 of the a*, b*, and L* values after UV ageing (Table 4), which is completely
Raw umber Cyprus 11.66 −6.74 −10.84 12.68 in accordance to the literature [36].
Bone (ivory) black) 1.25 0.08 −0.20 0.92
4. Conclusions
d)

PVAc mock-up ΔL* Δa* Δb* ΔE* 2000 The UV photo-oxidative stability—including the UV-B range for
Titanium white anatase 1.29 0.12 −3.74 3.41 simulating sunlight outdoor conditions—of four different binding
Titanium white rutile 0.33 0.16 0.64 0.65 media such as acrylics in two different co-polymer forms (Plextol®
Cadmium yellow −1.43 0.43 −0.17 1.08
D498 and Primal® AC33), alkyd, and polyvinyl acetate (PVAc)
Cadmium red −1.01 −0.01 0.77 0.96
Hydrated chromium oxide green 1.78 −2.56 1.34 2.19 when mixed with different types of inorganic pigments, particularly
Ultramarine blue 1.91 −2.94 0.64 1.98 used in modern and contemporary art, was investigated through
Raw umber Cyprus 1.83 −0.44 −0.88 1.59 the single-shot and double-shot Py-GC/MS, FTIR-ATR, and colour
Bone (ivory) black) −1.28 −0.02 0.05 0.90
measurements.
e)
After UV exposure, these techniques recorded different ageing be-
haviours of the acrylics, alkyd, as well as polyvinyl acetate (PVAc). The
Pigments ΔL* Δa* Δb* ΔE* 2000
double-shot technique of Py-GC/MS was especially effective at detect-
Titanium white anatase 3.00 −0.10 −3.30 3.38 ing the effect of the photo-oxidative reactions in a more precise and de-
Titanium white rutile 4.65 0.43 −1.63 3.14
tailed manner relative to the single-shot method. Depending on the
Cadmium yellow 6.43 −6.75 −1.71 5.58
Cadmium red 3.18 0.50 −0.22 2.94 analysed binder, the comparison between the thermal desorption of
Hydrated chromium oxide green 0.86 3.02 −0.65 1.34 the double-shot technique results obtained from the unaged and aged
Ultramarine blue 1.64 −0.24 −0.71 1.44 samples shows main evidence for photo-oxidative deterioration.
Raw umber Cyprus 4.46 −1.82 −1.33 3.98 Concerning both acrylic binding media, the detection of dimers, trimers,
Bone (ivory) black) −1.07 0.39 1.63 1.84
and sesquimers can be detected at a lower temperature than the normal
V. Pintus et al. / Microchemical Journal 124 (2016) 949–961 959

Fig. 5. CIELAB diagram of the acrylic Plextol® D498 and Primal® AC33, alkyd, and polyvinyl acetate mock-ups investigated before and after UV ageing showing the shifts of their L*, a* and
b* coordinates.

temperature used for the pyrolysis. This can be considered an effect of hand, newly formed products including unspecific aldehyde, lactones,
the UV light on the investigated materials, showing the impact of UV and acidic oxidation products were detected by FTIR-ATR analysis
on their thermal stability. Additionally, the total decrease of the non- from both acrylic binders.
ionic surfactant after UV exposure, which was possible to detect in the The results obtained from the thermal desorption of the double-shot
unaged binders only by plotting its mass profile in the thermal desorp- technique of the unaged and aged alkyd binder showed a higher
tion pyrogram, was sign of the sensitivity of the acrylic binders to the UV amount of phthalic anhydride indicating that more free ortho-phthalic
light. The pyrolysis step of the double-shot Py-GC/MS, as well as the acid was produced during the UV ageing, which is shown by FTIR-ATR
single-shot Py-GC/MS, showed the higher sensitivity of the BA and EA by the gradual reduction in absorption of the phthalic C\\O stretching
monomers of the Plextol® D498 and Primal® AC33, respectively, in band due to the C\\O bond scission from the alkyd structure. This result
comparison to the MMA monomers by the gradual decrease of the BA is particularly important because it is the first time this phenomenon
and EA peaks in the pyrograms of the aged binders. On the other has been detected during photo-oxidative studies of alkyd binders
960 V. Pintus et al. / Microchemical Journal 124 (2016) 949–961

Fig. 6. CIELAB diagram of the pigments investigated before and after 83 days of UV ageing showing the shifts of their L*, a*, and b* coordinates.

and especially because the formation of free ortho-phthalic acid was measurements demonstrated the higher sensitivity to the UV light of
induced by the effect of UV light including the UV-B range. Furthermore, the alkyd in comparison to the acrylics and PVAc mock-ups showed
the decrease of unsaturated fatty acid and increase of saturated by the bigger shift of the main coordinates L*, a*, b*, and E* coordinates.
dicarboxylic acid was detected by THM-GC/MS and the pyrolysis Thus, the UV ageing of different types of binders either alone or when
second step of the double-shot method. In the thermal desorption mixed with different inorganic pigments shows that their different
step of double-shot Py-GC/MS of the aged polyvinyl acetate (PVAc) chemical composition has a first order effect on their resistance to UV-
binder, the detectable amount of acetic acid increased, while the radiation.
amount of the DEP type of plasticiser highly decreased in the PVAc In this work, the Py-GC/MS technique, especially the double-shot
aged samples. This latter result was also evinced by FTIR-ATR by the mode, proved to be suitable for the photo-oxidative ageing studies of
broadening of the carbonyl peak. acrylics, alkyd, and PVAc binders, whereas the FTIR-ATR method and
Apart from more accentuated differences already noticed between colour measurements provided useful information about the influence
the unaged and aged binders, the results obtained with the double- of inorganic pigments on the UV stability of the analysed binding media.
shot and single-shot Py-GC/MS did not show any difference between
the unaged and aged mock-ups, whereas the FTIR-ATR results demon-
strated that each pigment induces to each binder a different ageing be- Acknowledgements
haviour. For instance, the cadmium yellow pigment mixed with both
acrylics and alkyd seems to promote the formation of ester side groups This work has been funded by Regione Sardegna (Italy), “Programma
during UV exposure, which corresponds to an increase of all principal Master and Back anno 2009” Alta Formazione and the Austrian Science
infrared absorptions in the IR spectra. Additionally, the colour Fund, Project no. L699-N17. We thank Anthony J. Baragona (Institute of
V. Pintus et al. / Microchemical Journal 124 (2016) 949–961 961

Science and Technology, Department of Conservation Science, University [19] P. Malanowski, S. Huijser, F. Scaltro, R.A.T.M. van Benthem, L.G.J. van der Ven, J.
Laven, G. de With, Molecular mechanism of photolysis and photooxidation of
of Applied Arts, Vienna, Austria) for helping with the English corrections. poly(neopentyl isophthalate), Polymer 50 (2009) 1358–1368.
[20] J. Lemaire, J. Gardette, J. Lacoste, P. Delprat, D. Vaillant, Mechanism of photo-
oxidation of polyolefins: prediction of lifetime in weathering conditions, in: R.L.
References Clough, N.V. Billingham, K.T. Gillen (Eds.), Polymer durability: Degradation stabili-
zation and Lifetime predictions, Boston American Chemical Society 1996,
[1] T. Learner, O. Chiantore, D. Scalarone, Ageing studies on acrylic emulsion paints, Pre- pp. 577–598.
prints ICOM Committee for Conservation 13th Triennial Meeting, James & James [21] N.J. Turro, Modern Molecular Photochemistry, University Science Books, Mill Valley,
(Science Publishers), Rio de Janeiro, 2002. 911–919. 1991.
[2] P.M. Whitmore, V.G. Colaluca, The natural and accelerated aging of an acrylic artist’s [22] M.T. Doménech-Carbó, G. Bitossi, L. Osete-Cortina, J. de la Cruz-Ca˜ nizares, D.J. Yusá-
medium, Stud. Conserv. 40 (1995) 51–64. Marco, Characterization of polyvinyl resins used as binding media in paintings by
[3] M.J. Melo, S. Bracci, M. Camaiti, O. Chiantore, F. Piacenti, Photodegradation of acrylic pyrolysis–silylation–gas chromatography–mass spectrometry, Anal. Bioanal. Chem.
resins used in the conservation of stone, Polym. Degrad. Stab. 66 (1999) 23–30. 391 (2008) 1371–1379.
[4] O. Chiantore, L. Trossarelli, M. Lazzari, Photooxidative degradation of acrylic and [23] M.F. Silva, M.T. Doménech-Carbó, L. Fuster-Lopeˇız, S. Martıˇın-Rey, M.F.
methacrylic polymers, Polymer 41 (2000) 1657–1668. Mecklenburg, Determination of the plasticizer content in poly(vinyl acetate) paint
[5] K.J. Saunders, Organic Polymer Chemistry, Chapman and Hall, London, 1973. medium by pyrolysis–silylation–gas chromatography–mass spectrometry, J. Anal.
[6] M. Lazzari, O. Chiantore, Drying and oxidative degradation of linseed oil, Polym. Appl. Pyrolysis 85 (2009) 487–491.
Degrad. Stab. 65 (1999) 303–310. [24] R. Ploeger, D. Scalarone, O. Chiantore, The characterization of commercial artist’
[7] J. Ferreira, M.J. Melo, A.M. Ramos, M.J. Avila, Eternity is in love with the productions alkyd paints, J. Cult. Herit. 9 (2008) 412–419.
of time: Joaquim Rodrigo’s classical palette in a vinyl synthetic medium, in: T. [25] R. Ploeger, S. Musso, O. Chiantore, Contact angle measurements to determine the
Learner, P. Smithen, J.W. Krueger, M.R. Schilling (Eds.), Modern paints uncovered, rate of surface oxidation of artists’ alkyd paints during accelerated photo-ageing,
The Getty Conservation Institute, Los Angeles, 2007 (pp.). Prog. Org. Coat. 65 (2009) 77–83.
[8] M. Herrera, G. Matuschek, A. Kettrup, Fast identification of polymer additives by [26] C. Duce, V. Della Porta, M.R. Tine, A. Spepi, L. Ghezzi, M.P. Colombini, E. Bramanti,
pyrolysis-gas chromatography/mass spectrometry, J. Anal. Appl. Pyrolysis 70 FTIR study of ageing of fast drying oil colour (FDOC) alkyd paint replicas,
(2003) 35–42. Spectrochim. Acta A 130 (2014) 214–221.
[9] N.S. Allen, Photofading and light stability of dyed and pigmented polymers, Polym. [27] F.X. Perrin, M. Irigoyen, E. Aragon, J.L. Vernet, Artificial ageing of acrylurethane and
Degrad. Stab. 44 (1994) 357–374. alkyd paints: a micro-ATR spectroscopic study, Polym. Degrad. Stab. 70 (2000)
[10] V. Pintus, M. Schreiner, Characterization and identification of acrylic binding media: 469–475.
influence on UV light on the ageing process, Anal. Bioanal. Chem. 399 (2011) [28] W.J. Muizebelt, J.C. Hubert, R.A.M. Venderbosch, A.J.H. Lansbergen, R.P. Klaasen, K.H.
2961–2976. Zabel, Aluminum compounds ad additional crosslinkers for air-drying high-solids
[11] S. Wei, V. Pintus, M. Schreiner, A comparison study of alkyd resin used in art works alkyd paints, J. Coatings Technol. 70 (1998) 53–59.
by Py-GC/MS and GC/MS: the influence of ageing, J. Anal. Appl. Pyrolysis 104 (2013) [29] N.S. Allen, M.J. Parker, C.J. Regan, R.B. McIntyre, W.A.E. Dunk, The durability of
441–447. water-borne acrylic coatings, Polym. Degrad. Stab. 47 (1995) 117–127.
[12] S. Wei, V. Pintus, M. Schreiner, Photochemical degradation study of polyvinyl ace- [30] L.J. Bellamy, The infrared spectra of complex molecules, Chapman and Hall, London,
tate paints used in artworks by Py-GC/MS, J. Anal. Appl. Pyrolysis 97 (2012) 1975.
158–163. [31] D. Lien-Vien, N.B. Colthup, W.G. Fateley, J.G. Grasselli, The handbook of infrared and
[13] ASTM International D2565–99 (Reapproved 2008) Standard Practice for Xenon-Arc raman characteristic frequencies of organic molecules, Academic, New York, 1991.
Exposure of Plastics Intended for Outdoor Applications. [32] M.T. Doménech-Carbó, M.F. Silva, E. Aura-Castro, L. Fuster-López, S. Kröner, M.L.
[14] J. Mallegol, J. Gardette, J. Lemaire, Long-term behaviour of oil-based varnishes and Martínez-Bazán, X. Más-Barberá, M.F. Mecklenburg, L. Osete-Cortina, A.
paints. Photo- and thermooxidation of cured linseed oil, J. Am. Oil. Chem. Soc. 77 Doménech, J.V. Gimeno-Adelantado, D.J. Yusá-Marco, Study of behaviour on simu-
(2000) 257–273. lated daylight ageing of artists’ acrylic and poly(vinyl acetate) paint films, Anal.
[15] M. Day, D.M. Wiles, Photochemical degradation of poly(ethylene terephthalate). I. Bioanal. Chem. 399 (2011) 3155–3304.
Irradiation experiments with the xenon and carbon arc, J. Appl. Polym. Sci. 16 [33] J.L. Ferreira, M.J. Melo, A.M. Ramos, Poly(vinyl acetate) paints in work of art: a pho-
(1972) 175–189. tochemical approach. Part 1, Polym. Degrad. Stab. 95 (2010) 453–461.
[16] M. Day, D.M. Wiles, Photochemical degradation of poly(ethylene terephthalate). II. [34] D. Scalarone, O. Chiantore, T. Learner, Ageing studies of acrylic emulsion paints. Part
Effect of wavelength and environment on the decomposition process, J. Appl. II. Comparing formulation with poly(EA-co-MMA) and poly(nBA-co-MMA) binders,
Polym. Scie. 16 (1972) 191–202. Preprints ICOM Committee for Conservation 14th Triennal Meeting, James & James/
[17] M. Day, D.M. Wiles, Photochemical degradation of poly(ethylene terephthalate). III. Earthscan, The Hague, 2005. 350–358.
Determination of decomposition products and reaction mechanism, J. Appl. Polym. [35] P. Whitmore, V. Colaluca, E. Farrell, A note on the origin of turbidity in films of an
Scie. 16 (1972) 203–215. artists' acrylic paint medium, Stud. Conserv. 41 (1996) 250–255.
[18] T. Grossetete, A. Rivaton, J.L. Gardette, C.E. Hoyle, M. Ziemer, D.R. Fagerburg, H. [36] M. Schwartz, R. Baumstark, Waterbased acrylates for decorative coatings, Vincent
Clauberg, Photochemical degradation of poly(ethylene terephthalate)-modified co- Verlag, Hannover, 2001.
polymer, Polymer 41 (2000) 3541–3554.

Das könnte Ihnen auch gefallen