You are on page 1of 16



CHAPTER FIVE

MURALES IN SAN SPERATE (SARDINIA):


IDENTIFICATION AND CHARACTERISATION
OF MODERN PAINT MATERIALS EXPOSED
TO OUTDOOR CONDITIONS.

V. PINTUS, M. ANGHELONE, E. DRAGANITS


AND M. SCHREINER

Introduction
Since the late 1960´s “murales”–-which owe their name in Italian to
the plural form of the Spanish word “mural” (wall)–-of great cultural and
historical value have been painted in San Sperate, a village approximately
25 km from Cagliari, Sardinia, Italy. Today, the village and its murales
represent a unique museum where the many highly visible murales offer
inhabitants and visitors the opportunity to appreciate their beauty and
messages in the context of the whole village.
The famous painter-sculptor Pinuccio Sciola is considered to be the
pioneer of introducing the influence of the Mexican muralismo movement
to Sardinia (Concu 2012, Olita and Pes 2007). Sciola together with other
artists started to paint the old ládiri (mud-brick) walls of the historical
centre, turning the initial idea of a village-museum into a reality (Concu
2012)--one that has lasted so far. His project was mostly focused on the
realisation of an outdoor museum which expresses the locals’ state of
mind and the cultural changes they were experiencing, thus involving all
people of the community. Based on the aim and message that the artist
wanted to impart, the selection of the wall was the first step for a murale.
Sometimes defects or imperfections of the wall, as well as the presence of
windows and balconies, stimulated new ideas for the artist, which were
then integrated into the theme of the murales. Once the idea of a painting
Murales in San Sperate (Sardinia) 87

was formed, the next step was the application of a smooth intonaco
(plaster layer) over the wall, which was especially useful with walls built
of ládiri or when the surface was rough or had imperfections. Commonly a
further layer of lime-whitewash paint was then applied over the intonaco,
which once dry was ready to be painted. For the paintings mostly
washable exterior paints were used (Concu 2012) because of their low
cost, fast drying quality, easy application, and adaptability to different
supports. Industrial varnishes might also have been employed (Concu
2012).
The subjects of the murales in San Sperate, which are mostly based on
daily life or social or environmental motifs, differ from later ones found in
other villages in Sardinia such as Orgosolo and Villamar, which show
essentially political or ideological scenes relevant to Sardinians. In the last
few decades, artists such as Luciano Lixi and Angelo Pilloni have carried
out many murales in San Sperate using synthetic paints in different styles
and techniques. These modern and contemporary artworks are painted on
diverse surfaces and are commonly exposed to harsh outdoor conditions
including strong sunlight, wide temperature variations, wind, precipitation,
aerosols, microbiological colonisation, and vandalism. Therefore rigorous
state-of-the-art documentation and study is essential in order to appreciate
their importance as cultural heritage, as they are under constant threat.
Although interest in the preservation and conservation of this
important cultural heritage is increasing, little information about applied
paint materials or about techniques is available. There are several books
and Internet sites about murales in general, especially those from Sardinia
(Rubanu and Fistrale 1998, Barnoux 2001, Olita and Pes 2007, Piras 2010,
Concu 2012). Most of them are more or less designed as picture books
containing abundant, colourful images of mural art. They commonly
mention historical background, the village where the mural can be found,
the name of the artist and possibly the year of creation, but usually little
further information. Moreover, the lack of the exact location within a
village is frequently also missing. In San Sperate, a complex street pattern
containing more than 300 murales makes the search for a specific mural a
long-lasting undertaking within the labyrinth of narrow streets.
In this work a scientific study of three different murales in San Sperate
was carried out. It aimed to characterize and identify for the first time the
materials used for these valuable modern and contemporary wall paintings
as well as to understand their painting techniques, thus providing
preliminary documentation useful for conservation and preservation
purposes. Therefore the geotagging method was used for a complete
88 Chapter Five

photographic documentation of each of the murales while optical


microscopy (OM) was employed for studying the stratigraphy of all
samples, giving important information about the painting technique.
Additionally, different analytical techniques such as pyrolysis-gas
chromatography/mass spectrometry (Py-GC/MS), Fourier transform
infrared spectroscopy-attenuated total reflectance (FTIR-ATR), and
Raman spectroscopy were used for the investigation of the organic and
inorganic compounds and to determine possible degradation products.

Experimental
Samples
Three different murales in San Sperate were selected for this study.
Because the three selected murales are untitled, and for the sake of
simplicity, in this manuscript the name Murales 1 (Fig. 2-la), Murales 2
(Fig. 2-lb), and Murales 3 (Fig. 2-lc) have been assigned to the murale at
via Risorgimento 92, the murale at via Risorgimento 14, and the murale at
via Sassari 1, respectively.
Sampling was carried out as non-destructively as possible; small paint
fragments were taken only from areas where the paint had already started
to flake off. The name and the type of the collected samples for each
murale are listed in Table 2-1.

Fig. 2-1: Images of the three investigated murales in San Sperate and sampling
area with the relative cross-section. a) Murales 1 in via Risorgimento 92, b)
Murales 2 in via Risorgimento 14, c) Murales 3 in via Sassari 1, and d) typical wall
in San Sperate made with lddiri (mud bricks).
Murales in San Sperate (Sardinia) 89

Photographs geotagging
The process called “geotagging”–commonly used for adding location
information to photos (Luo et al. 2011)–was employed for the
documentation of the mural paintings. The determination of a murale’s
location was based on the Global Navigation Satellite Systems (GNSS), of
which presently only the American Global Positioning System (GPS) and
the Russian Global Navigation Satellite System (GLONASS) are fully
globally operational (Hofmann-Wellenhof et al. 2008).
GNSS enables position determination (longitude, latitude and altitude)
with the help of small electronic devices that receive and calculate signal
travel time based on location and timing of signals from a number of
satellites in orbit around the Earth. As in line-of-sight connection with at
least 4 satellites is required for location calculation, at least 24 satellites
have to be in orbit.
The photo-geotagging of murales was done using the Solmeta
Geotagger Pro (Solmeta Technologies, China), which was mounted on a
Nikon D90. The geotagger recorded the GPS data (accuracy <10 m) and
direction of view directly into the exif file of each photo. Geosetter 3.4.16
free software was used for handling the position data of the pictures.
Geosetter cannot only read and modify photo metadata, but it also shows
their location and even the direction of view on a variety of maps
including Google Maps and OpenStreetMap (Ramm and Topf 2010). In
the map representation in GeoSetter, based on the direction and focal
length recorded in the metadata, even the part of the landscape covered by
the specific photos is indicated.
The automatic recording of location data into the metadata of every
photo takes place automatically while shooting a photo, and therefore no
additional time is required. The photos are unequivocally located by their
geographical coordinates and since all data are recorded automatically, any
confusion, mistakes or reading errors are unlikely. Since geotagging
automatically saves location and directional data into the metadata of
every single photo, it greatly enhances the ability to document murales or
any other outdoor work of cultural importance.
90 Chapter Five

Optical microscopy (OM)


For the investigation of the morphology and stratigraphy of the
samples, a LEITZ Metalloplan optical microscope (6.3x, 10x, 20x, 30x,
60x objectives and 8x oculars) was employed using visible light (Vis). The
microscope was equipped with a UV lamp Nikon HB-10101AF in order to
observe the samples under ultraviolet (UV) light, although no fluorescent
compound was detected. The images of the samples were acquired by
Camera Nikon Digital Sight DS-Fi1 and then collected and evaluated with
the NIS-Elements F2.20 Software.
To prepare cross-sections, tiny representative fragments taken from the
selected samples were embedded into a mixture of a polyester resin type
GT (Kurt & Wolf CO KG, Austria) with an amount of 2% of MÄK
(methyl-ethyl-ketone peroxide) hardener resin (Kurt & Wolf CO KG,
Austria) (see Fig. 2-1). Afterwards they were allowed to dry for 24 hours
at room temperature. Once dried all cross-sections were polished by using
abrasive grinding papers based on silicon carbide (SiC) with different grit
sizes (from 600 to 2400 grit) in order to observe their stratigraphy under
the optical microscope without any interference of a rough surface.

Pyrolysis gas chromatography mass spectrometry (Py-GC/MS)


To identify and characterize the sample material, each paint layer was
carefully separated and scraped under the optical microscope with the help
of a scalpel in order to avoid the interference of different types of binder or
organic pigment from other layers in the resulting pyrograms.
Consequently, Py-GC/MS analysis was carried out on between 0.1 and 0.3
mg of each layer, which were put in a sample-cup (ECO-CUP Frontier
Lab, Japan) for analysis. Py-GC/MS analyses were performed with a PY
2020iD (Frontier Lab, Japan) pyrolyser unit combined with a GCMS-
QP2010 Plus (Shimadzu, Japan) equipped with a capillary column Ultra
Alloy-5 (5% diphenyl 95% dimethylpolysiloxane) with 0.25 mm internal
diameter, 0.25 ȝm film thickness, and 30 m length (Frontier Lab, Japan).
NIST 05 and NIST 05 s Library of Mass Spectra were used for identifying
the compounds.
The GC column temperature conditions were as follows: initial
temperature 40 °C, held for 5 min, followed by a temperature ramp of 6
°C/min to 280 °C, held for 3 min. The helium gas flow was set at 1 mL
min-1 and mass spectra were recorded under electron impact ionization at
70 eV. For the analyses the pyrolysis temperature was set at 600 °C and
Murales in San Sperate (Sardinia) 91

held for 12 seconds. The temperature of the pyrolysis interface and


injector was set at 280 °C.

Fourier transform infrared spectroscopy-attenuated total


reflection (FTIR-ATR)
FTIR–ATR analyses were performed with an Alpha FT-IR Platinum
ATR instrument (Bruker Optics, Germany) equipped with a deuterated
triglycine sulfate detector (DTGS) and with a diamond crystal. Spectra
were acquired in the spectral range between 4000 and 370 cm-1
performing 64 scans at 4 cm-1 resolution. The resulting spectra were
collected and evaluated with the spectrum software Opus from Bruker
Optics.

Raman spectroscopy
The LabRAM Aramis confocal micro-Raman spectrometer (HORIBA
Scientific, Japan) equipped with Nd-YAG 532 nm (green), HeNe 632.8
nm (red), AlGaAs diode 785 (NIR) lasers and with a confocal microscope
Olympus L-BXFM (Olympus Corporation, Japan) having 10x, 50x LWD,
50x, and 100x objectives, was used. Additionally, the confocal microscope
was coupled to a 460 mm focal length spectrograph with 300, 600, 1200
and 1800 g/mm interchangeable gratings. The acquisition of the spectra
was performed with LabSpec Software, while the evaluation of the
obtained spectra was done with ACD/SpecManager.
The identification of inorganic pigments was based on the comparison
between the acquired spectrum of the sample and the reference spectrum
of the IRUG (Infrared Raman Users Group) database. Synthetic organic
pigments were characterized and identified according to the literature
(Scherrer et al. 2009).
92 Chapter Five

Table 2-1: List of the analytical techniques used for the investigation
of the samples from each murales and the main results detected by
every method.

Analytical technique
Optical miscroscopy (OM) Raman spectroscopy Py-GC/MS FTIR-ATR
Intonaco/plaster -
Layer nr. Thickness (Njm) Colour Pigment Binder
Whitewash paint

Murales 1 - via Risorgimento 92


SBl (Sample_Blue)
0 220 White Calcite (CaCO3) + Titanium dioxide rutile (TiO2) Vinyl-Acrylic
1 40 Blue Copper phthalocyanine (CuPC) Vinyl-Acrylic
SY (Sample_Yellow)
0 320 White Calcite (CaCO3) + Titanium dioxide rutile (TiO2) Vinyl-Acrylic
1 30 Yellow Monoazo pigment PY74 Vinyl-Acrylic

Murales 2 - via Risorgimento 14


SY (Sample_Yellow)
0 240 White Calcite (CaCO3) + Titanium dioxide rutile (TiO2) Vinyl-Acrylic Calcite (CaCO3) - silica (SiO2) -
1 80 Yellow Monoazo pigment PY74 Vinyl-Acrylic iron oxide (Fe2O3)
(Carbonated Portland Cement
SBr (Sample_Brown)
based plaster)
0 120 White Calcite (CaCO3) + Titanium dioxide rutile (TiO2) Vinyl-Acrylic -
1 90 Yellow Monoazo pigment PY74 Vinyl-Acrylic Calcite (CaCO3) + Titanium
Dark red- dioxide rutile (TiO2) + vinyl-acrylic
2 40 Iron hydroxyde (Fe(OH)2) Acrylic
brownish (Whitewash paint)
SGBl (Sample_Grayish-Blue)
0 500 White Calcite (CaCO3) + Titanium dioxide rutile (TiO2) Vinyl-Acrylic
1 40 Blue Copper phthalocyanine (CuPC) Vinyl-Acrylic
2 2 Gray Calcite (CaCO3)

Murales 3 - via Sassari 1


SRW (Sample_Red-White)
1 85 Red Iron oxide (Fe2O3) Vinyl-Acrylic
2 35 Red Synthetic organic red Vinyl-Acrylic
3 110 White Calcite (CaCO3) + Titanium dioxide rutile (TiO2) Vinyl-Acrylic

Results and discussion


Photographic geotagging and documentation
The exact location of the three investigated murales was carried out by
geotagging the acquired photos. Figure 3-1 displays the location of the
investigated murales in the satellite and street pattern view from Google
Earth and in the OpenStreetMap. Additionally, it indicates the direction of
their exposure.
The data obtained from geotagging and photography of the
geographical position (longitude, latitude and altitude), street name and
number, exposition, year, artist’s name (where known) is listed in Table 3-
1.
Additionally, a preliminary evaluation of the state of conservation of
each murales was carried out. Murale 1 was mainly characterized by the
Murales in San Sperate (Sardinia) 93

disaggregation in several areas of the paint films applied onto the lower
part of the wall near where it meets the sidewalk (Fig. 2-1 a). The paint
layers as well as the intonaco support were flaking off particularly badly
from Murale 2, especially along all central part of the wall (Fig. 2-lb). The
lower right side of this murale was signed by "San Sperate Paese Museo-
23-2-89", essential information for the documentation of murales.

Fig. 3-1 : Geotagging of the three investigated murales (1, 2, and 3) in San Sperate,
Sardinia (Italy) using the satellite view of Google Earth (left) and the map of
OpenStreetMap (right).

Table 3-1: List of the main data acquired by geotagging the photos of the
three investigated morales and integrated by additional information.
Geotagging
Longtude Latitude Attude i
Foto dU-ecton Muraes exposton Street name and number Artist Year

E9°0'14.64" N39°21'20.85" 40m so0 290° Murales 1 - via Risorgimento 92 x x

E9°0'20.71" N39°21'38.16" 43m 285° 105° Murales 2 - via Risorgimento 14 x 1989

E9°0'20.74" N39°21'20.96" 40m 125° 305° Murales 3 - via Sassari 1 Luciano Lixi 1996
94 Chapter Five

Murale 3, which was inscribed by the signature of the artist and the
year “Luciano Lixi-96”, was also characterised by a loss of the paint
layers, especially from the middle to the lower area of the wall (Fig. 2-1c).
Furthermore, several long cracks run through the murales in different
places.

Analytical techniques
Different, complementary analytical techniques were used for the study
of selected murales in San Sperate. These were used for the identification
of the type of materials contained in the paint, such as synthetic organic
binder, organic and inorganic pigments, the plaster support for the paint, as
well as to understand the painting technique. The results obtained by every
analytical technique on each sample taken from three different murales are
reported in Table 2-1.
0.45

b)
1013
0.40

1222
0.35
0.30

1728
0.25
ATR Units

a)
0.20

1420
1010
0.15

875
0.10

469
710
535
0.05

1796

4000 3500 3000 2500 2000 1500 1000 500


-1
Wavenumber cm

Fig. 3-2: FTIR-ATR spectra of the a) intonaco support having calcite (CaCO3),
silica (SiO2), and iron oxides (Fe2O3), and b) whitewash paint–a characteristic of
the first layer of all samples–which additionally contains a vinyl-acrylic resin.
Murales in San Sperate (Sardinia) 95

Prior to the analytical analysis a documentation of the state of


conservation of each sample was carried out under the optical microscope.
A peculiarity that was observed in most of the samples is the simple
stratigraphy, mainly based on only two paint layers above the intonaco
support (plaster). FTIR-ATR measurements yielded that the main
constituents of the plaster were mostly constituted of calcite (CaCO3)
(bands at 1796, 1420, 875, and at 710 cm-1), silica (SiO2) (Si-O-Si strong
and broad absorption in the IR region between 1100-1000 cm-1 with a
maximum at 1010 cm-1 and at 798 cm-1) and iron oxides (Fe2O3) (bands
at 469 and 535 cm-1) (Fig. 3-2a), which are the main elements of
carbonated Portland cement based plaster (Kootker 2007).
The first layer – meant as the first paint level applied on the carbonated
Portland cement based plaster – had a white colour and a homogeneous
quality, which in some cases contains several quartz crystal inclusions of
small dimensions (Fig. 2-1a and 2-1b). Raman measurements showed that
calcite (CaCO3) together with titanium white pigment (TiO2) composed of
rutile characterized the first layer of all samples analysed.
The Figure 3-3a shows the CO32- stretching band at 1087 cm-1, the
CO32- bending band at 713 cm-1 and peaks at 154 and 281 cm-1 relative
to the lattice vibration of calcite, while the peaks at 143, 448, 611 cm-1 are
due to the different Ti-O vibrational modes of rutile. The presence of
calcite was also confirmed by FTIR-ATR measurements and
complemented by some peaks at 1728, 1222, and at 1013 cm-1 of a vinyl-
acrylic resin (Fig. 3-2b), especially detected by Py-GC/MS (Fig. 3-4a).
These compounds are a clear indication of the use of lime-whitewash paint
as a preparation layer above the intonaco for the subsequent coat of
colour. Figure 2-1d shows a traditional wall in San Sperate made by ládiri
(mud clay bricks) covered first by a layer of intonaco and then whitewash,
which clearly corresponds to the stratigraphy found in the investigated
samples.
The second layer was essentially based on a unique, thin and
homogeneous paint colour (Fig. 2-1a). The presence of further paint layers
in some samples was mainly due to the application of a shading colour
effect in certain areas of the murales (Fig. 2-1b and 2-1c), which was
confirmed by a macroscopic investigation of the wall painting.
96 Chapter Five

d)

293
226

412

613
498

c)

1530
749

1452
683

1343
485
Raman Intensity

b)
1329
1264

1594
13531403
1509
1160
1091
803

a)
448

611
281

1087
143

713

100 300 500 700 900 1100 1300 1500 1700


Raman Shift (cm-1)

Fig. 3-3: Raman spectra of: a) white layer of sample SBl–Murale 1 (532 nm,
1mW, 3x20 sec), b) yellow layer of sample SY–Murale 1 (532nm, 1mW, 5x10
sec), c) blue layer of sample SBl–Murale 1 (632.8 nm, 1mW, 40 sec), and d) red
layer of sample SRW–Murale 3 (785 nm, 1 mW, 40 sec).
Murales in San Sperate (Sardinia) 97

The type of pigment of the second layer was identified by Raman


spectroscopy. Synthetic organic pigments such as monoazo yellow in
specimen SY of Murale 1 (Fig. 3-3b) and the sample SY and SBr of
Murale 2, copper phthalocyanine in sample SBl of Murale 1 (Fig. 3-3c)
and in SGBl of Murale 2, and an unidentified synthetic organic red in
SRW of Murale 3 (Fig. 3-3d), were mainly used. According to the Raman
spectrum (Fig. 3-3b), the monoazo yellow pigment belongs to the
acetoacetic arylide group.
Table 3-2: Main pyrolysis products of the paints/whitewash and
sample red-white of Murale 3 at their corresponding retention time
(RT min) and their respective molecular weight (m/z).

Paints/whitewash Sample red-white (Murales 3)


Peak RT min m/z
(vinyl-acrylic based binder) (p(2-EHA/MMA) based binder)

1 2.1 Carbon dioxide 28, 44


2 2.5 Acetone 43, 58
3 3.3 Acetid acid 43, 60
4 3.9 Benzene 52, 78
5 4.9 MMA 69, 100
6 6.7 Toluene 65, 91
7 7.4 2-ethylhexene 55, 70, 112
8 11.0 Styrene 51, 78, 104
9 11.2 n BA 55, 73, 83, 128
10 13.1 2-Ethylhexanal 57, 72
11 13.4 Benzaldehyde 51, 77, 105
12 14.2 Benzonitrile 50, 76, 103
13 15.5 2-Ethylhexanol 57, 70, 83, 112
14 15.9 Benzene, 1-propynyl 63, 89, 115
15 16.6 Acetophenone 51, 77, 105, 120
16 19.9 Naphthalene 51, 75, 102, 128
17 20.8 2-EHA 55, 70, 83, 112, 127
18 21.9 VeoVa 55, 73, 87, 101, 116, 130
19 23.0 Phthalic acid 50, 76, 104, 148
20 31.7 2-EHA/MMA sesquimer 55, 70, 83, 115, 143
21 32.1 2-EHA/MMA dimer 57, 70, 83, 95, 112, 127, 141, 155, 173
Phthalic acid, diisobutyl ester
22 33.8 57, 104, 121, 149
(DIBP)

23 35.4 Phthalic acid, dibutyl ester (DBP) 57, 104, 149

24 40.3 2-EHA sesquimer 57, 71, 83, 115, 133, 245, 284
25 40.9 2-EHA dimer 57, 71, 83, 112, 127, 145
26 45.2 2-EHA-2-EHA-MMA trimer 57, 71, 93, 139, 195, 227, 257, 324, 368
27 45.5 2-EHA-2-EHA-MMA trimer 57, 71, 93, 139, 195, 227, 257, 324, 368
28 47.4 2-EHA trimer 57, 71, 93, 121, 149, 167, 195, 227, 284

The absence of the band at 1140 cm-1 corresponding to C-N


symmetric stretching vibration, and the presence of bands at 1264 cm-1
(amide III vibration) and at 1329, 1353 and 1594 cm-1 (aromatic
nitrogroup vibration) allowed the identification of the yellow pigment as a
PY74 (Vandenabeele et al. 2000, Fremout and Saverwyns 2012). The
spectrum acquired from the blue pigment presents typical features of a
98 Chapter Five

copper phthalocyanine blue (Fig. 3-3c): bands due to pyrrole C=C and azo
C-C stretches between 1450-1530 cm-1, a peak at 1343 cm-1 related to the
pyrrole C-C stretching, and macrocyclic ring breathing bands at 683-749
cm-1 (Lutzenberger 2009). In Figure 3-3d the red pigment can be easily
identified as iron oxide, all the peaks at 226, 293, 412, 498, and 613 cm-1
are related to the different Fe-O vibrational modes. Moreover, observing
the cross-section (Fig. 2-1c), and in particular the compact texture and the
fine granulometry of the second layer containing the iron oxide, it can be
assumed that it is a synthetic version of the red pigment.
For the identification and characterization of the type of binder of each
paint layer Py-GC/MS analyses were carried out. The main pyrolysis
compounds achieved at their corresponding retention times (RT min) are
listed in Table 3-2. The obtained results indicated that the main chemical
composition of the most analysed samples was based on a vinyl-acrylic
copolymer as for the binder detected in the lime/whitewash layer (Fig. 3-
4a).
b)
5 17

28
Relative intensity

21 25
10 13 26
27
20 24

1 a)
2

18
3
11 16
4 14
6 8 9 12 15 19 22 23

10 20 30 40
Retention time (min)

Fig. 3-4: Pyrograms of the a) white layer of sample SBl of Murale 1 characterized
by a vinyl-acrylic binder and b) brown layer of sample SBr of Murale 2 mainly
based on an acrylic binder–p(2-EHA/MMA). The identified peaks are listed in
Table 3-2.
Murales in San Sperate (Sardinia) 99

The pyrogram in Figure 3-4a clearly indicates that the binder is a vinyl
acetate/styrene-nbutyl acrylate copolymer. Carbon dioxide (RT 2.1 min–
m/z = 28, 44), acetone (RT 2.5 min–m/z = 43, 58), acetic acid (RT 3.2
min–m/z = 43, 60), benzene (RT 3.9 min–m/z = 52, 78), toluene (RT 6.7
min–m/z = 65, 91), vinyl versatate product (VeoVa) (RT 21.9 min–m/z =
55, 73, 87, 101, 116, 130) as well as a phthalic acid (RT 23.0 min–m/z =
50, 76, 104, 148), phthalic acid, diisobutyl ester (DIBP) (RT 33.8 min–m/z
= 57, 104, 121, 149), phthalic acid, dibutyl ester (DBP) (RT 35.4 min–m/z
= 57, 104, 149) based plasticizers are the main pyrolysis products of the
vinyl acetate in the copolymer. On the other hand the two close peaks of
styrene (RT 11.0 min–m/z = 51, 78, 104) and n-butyl acrylate (RT 11.2
min–m/z = 55, 73, 83, 128) are related to the styrene-n-butyl acrylate.
Vinyl versatate and the phthalic-based compounds are two different types
of plasticizer normally included in the polyvinyl acetate formulation. They
are added due to the hardness and brittleness of the PVAc homopolymer
that prohibits the formation of a continuous film (Learner 2004). The
vinyl-acrylic type of copolymer is commonly the main constituent of the
formulation of household paints for exterior use.
Another type of binder, the 2-EHA/MMA (2-ethylhexyl
acrylate/methyl methacrylate) copolymer was detected in the sample SRW
of Murale 3 by Py-GC/MS, where the peaks of MMA (RT 4.9 min–m/z =
69, 100), 2-ethylhexene (RT 7.4 min–m/z = 55, 70, 112), 2, ethylhexanal
(RT 13.1 min–m/z = 57, 72), 2-ethylhexanol (RT 15.5 min–m/z = 57, 70,
83, 112), and 2-EHA (RT 20.8 min–m/z = 55, 70, 83, 112, 127) are the
most intense (Fig. 3-4b). Several other peaks, present in the latter parts of
the pyrograms, were identified as sesquimers, dimers, and trimers (see
Table 3-2). Due to the very high adhesion of the brown layer with the
lower yellow layer, it was not possible in this case to properly separate the
two layers. Therefore, the vinyl-acrylic resin of the yellow layer was
detected in the pyrogram of the brown layer.
Besides the type of support, pigments, and binders that constituted the
materials used for the murales, an effect of the degradation processes on
the Murale 2 was detected by Raman spectroscopy. A thin layer of calcite
(CaCO3) covering the blue layer of sample SGBl was identified by Raman
measurements. The formation of calcite above the surface of the murales
indicates a process of solution and crystallization of calcium carbonate
contained in the layers below (intonaco and whitewash paint), which is
mostly caused by the constant variation in relative humidity (Cather 1991).
An evidence of this degradation reaction was given by the alveolar
formations found on the surface caused by crystallization of calcite.
100 Chapter Five

Conclusions
The combination of geotagging photographs and optical microscopy
(OM) with different analytical techniques such as Py-GC/MS, FTIR-ATR,
and Raman spectroscopy allowed for the collection of comprehensive
information about different murales in San Sperate such as the painting
technique and paint materials used, as well as information on their state of
preservation.
The geotagging of the selected murales in San Sperate added precise
spatial information to photos such as exact location of murales (street,
street number, house and even specific house walls), increasing the ease of
access for their protection as cultural heritage and scientific investigations
as well as for the benefit of visitors just enjoying their beauty.
The observation of the collected samples under the optical microscope
in visible light revealed the existence of two simple layers, mainly based
on the application of a paint colour–characterized by a mixture of organic
pigments detected by Raman measurements–with a vinyl-acrylic binder
identified by Py-GC/MS analysis–above a whitewash paint layer made by
calcite, titanium white, and a vinyl-acrylic copolymer obtained by Raman
spectroscopy, FTIR-ATR, and Py-GC/MS analysis. These two layers were
found above the intonaco support composed mostly of a carbonated
Portland cement based plaster, as confirmed by FTIR-ATR analysis.
Additionally a type of deterioration reaction, the formation of calcite
crystals on the surface of the Murale 2 due to large variation in relative
humidity within the wall, was detected analysed by Raman spectroscopy.
All of this information provides the first scientific data of murales in
San Sperate and contributes to the documentation, study and conservation
of murales as cultural heritage.

References
Barnoux, Y. Murales della Sardegna. Cagliari: Ettore Gasperini, 2001.
Campos, R. “On urban graffiti: Bairro Alto as a liminal space”. In The wall
and the city, edited by A.M.Brighenti. Trento: Professionaldreamers,
2009.
Cather, S. “The Conservation of wall paintings”. In Proceedings of a
symposium organized by the Courtauld Institute of Art and the Getty
Conservation Institute, Los Angeles: J. Paul Getty Trust, 1993.
Concu, G. Murales. L'arte del muralismo in Sardegna. Nuoro: Imago
Multimedia, 2012.
Murales in San Sperate (Sardinia) 101

Fremout, W. and S. Saverwyns, “Identification of synthetic organic


pigments: the role of a comprehensive digital Raman spectral library”.
Journal of Raman Spectroscopy, 43, (2012): 1536-1544.
Hofmann-Wellenhof, B., H. Lichtenegger and E. Wasle, GNSS - Global
Navigation Satellite Systems: GPS, GLONASS, Galileo & more. Wien:
Springer, 2008.
Kootker, L. M., “On the binder of mortars from Nikopolis, Greece. A
physico-chemical characterisation in view of their conservation”. In
IGBA-Rapport , 2007.
Learner, T. Analysis of Modern Paints. Los Angeles: The Getty
Conservation Institute, 2004.
Luo, J., D. Joshi, J. Yu and A. Gallagher, “Geotagging in multimedia and
computer vision – a survey” Multimedia Tools and Applications 51 (1),
(2011):187-211
Lutzenberger K., Synthetische organische Pigmente des 20. Jahrhunderts
und Moeglichkeiten ihrer zerstoerungsarmen analytischen
Identifizierung, Munich: Herbert Utz Verlag, 2009.
Olita, O. and N. Pes, San Sperate: all´origine dei murales. Cagliari:
AM&D edizioni, 2007.
Perra, M. ARTSTREETS|STREETSART. San Sperate: Il progetto dello
spazio pubblico di un paese-museo. Cagliari:Università degli Studi di
Cagliari, MA thesis.
Piras, D. Murales. Pontedera: Bandecchi & Vivaldi, 2010.
Ramm, F. and J. Topf, OpenStreetMap: Using and enhancing the free map
of the World. Cambridge: UIT Cambridge, 2010.
Rubanu, P. and G. Fistrale, Murales politici della Sardegna: Guida –
storia – percorsi. Bolsena: Massari, 1998.
Scherrer, N. C., S. Zumbuehl, F. Delavy, A. Fritsch and Kuehnen R.,
“Synthetic organic pigments of the 20th and 21st century relevant to
artist’s paints: Raman spectra reference collection” Spectrochimica
Acta Part A-Molecular and Biomolecular Spectroscopy, 73 (3) (2009):
505-524.
Vandenabeele P., L. Moens, H. G. M. Edwards, Dams R., “Raman
spectroscopic database of azo pigments and application to modern art
studies” Journal of Raman Spectroscopy 31 (6) (2000): 509–517.

You might also like