You are on page 1of 15

Progress in Organic Coatings 154 (2021) 106180

Contents lists available at ScienceDirect

Progress in Organic Coatings


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

Resistance to artificial daylight of paints used in urban artworks. Influence


of paint composition and substrate
E.M. Alonso-Villar a, b, *, T. Rivas a, b, J.S. Pozo-Antonio a, b
a
Depto. de Enxeñaría dos Recursos Naturais e Medio Ambiente, Escola de Enxeñaría de Minas e Enerxía, Universidade de Vigo, 36310, Spain
b
CINTECX, Universidade de Vigo, 36310, Spain

A R T I C L E I N F O A B S T R A C T

Keywords: Solar radiation is one of the deterioration factors par excellence of modern urban artworks, as among other
Paint manifestations, it causes loss of colour. This deterioration can compromise the durability of the works, although
Urban artwork it is not known to what extent this durability can vary depending on the substrate supporting the paint. We
Contemporaneous murals
evaluated the susceptibility to sunlight of 10 paints of differing compositions as typically used by urban artists,
Conservation
Cultural heritage
by exposing them, applied to concrete and brick, to artificial sunlight ageing test in laboratory conditions. Colour
Acrylic paints changes over time were studied, along with mineralogical (x-ray diffraction), chemical (scanning electron
Alkyd paints microscopy-energy dispersive x-ray spectroscopy) and physical (contact angle and peeling tests) changes. The
durability of the paints on these substrates depends mainly on the paint composition (base, pigment and filler)
but also on the type of substrate, as this influences paint adhesion capacity. The greatest colour modification was
detected in fluorescent paints, which lose colour due to degradation of the organic phase, and with aggravation
of defects during drying, eventually detach from the substrate.

1. Introduction abundant, even including a glossary of terms related to alteration forms


and agents [2], the deterioration processes that affect contemporary
Contemporaneous murals are artistic creations that are increasingly urban artworks is little documented and no conservation guidelines are
present in urban spaces. Whether independent or commissioned art­ as yet available.
works, they are created by applying modern materials in different ways A main conservation issue for contemporary urban artworks is the
(paint, dabber, spray graffiti, etc) to a wide variety of substrates (brick, fact that they are composed using modern paints, whose formulation
metal, concrete, wood, etc) and structures (stationary urban furniture, makes them very susceptible to solar radiation and colour loss [3]. Like
building walls and civil structures, etc). As with ancient mural artworks, ancient paint, modern paint is composed by a binder and pigments, but
interactions between the paints and deterioration agents (rainwater, sun it also contains extenders, solvents and additives [3–6]. The variety of
irradiation, pollutants, etc) lead to different alteration patterns that binders, which was limited in ancient paints (natural oils and fats,
affect both aesthetics (e.g., colour changes such as fading and yellowing) casein, collagen and egg), is much greater in modern paints, and in­
and integrity (loss of cohesion and of material), compromising their cludes low molecular resins (alkyd, polyurethane, phenolic, epoxy and
durability. Although the ephemeral nature of many independently polyester resins) and/or high molecular resins (acrylic, vinyl and
created artworks in modern urban spaces is recognized [1], there is also nitrocellulose type). Pigments in modern paints are mainly of organic
a certain interest in the creation of more durable artworks, as evidenced nature, the most frequent being phthalocyanines and anthraquinone
in commissions by local authorities and other administrative bodies. derivatives and especially azo-type chromophores (R-N=N-R) such as
Ideally, both the artists and the commissioners should have the best monoazo, diazo and azo-polycyclic compounds. Extenders, e.g., tita­
possible guidance in selecting the best materials and optimal physical nium oxide (TiO2, used mainly as a white pigment), calcium carbonate
conditions that ensure durability (substrate type, pre-treatment of the (CaCO3), kaolinite (Al2Si2O5(OH)4), talc (Mg3Si4O10(OH)2) and barite
surfaces, minimization of weathering and anthropic agent impacts, etc). (BaSO4) influence on the coating properties such as density, flow,
While the literature on the deterioration of ancient mural paintings is hardness and permeability [6]. Solvents, which are organic, facilitate

* Corresponding author.
E-mail address: enalonso@uvigo.es (E.M. Alonso-Villar).

https://doi.org/10.1016/j.porgcoat.2021.106180
Received 21 October 2020; Received in revised form 28 January 2021; Accepted 30 January 2021
Available online 3 March 2021
0300-9440/© 2021 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

the mixing and viscosity of paint components and are crucial for drying exposure. This resistance, and consequently the durability, will depend
and polymerizing organic components. Finally, additives may include how well the paint and substrate are integrated and how each reacts to
anti-settling agents, viscosity modifiers, surfactants or emulsifying deterioration factors. For instance, the alkalinity of concrete can influ­
agents, anti-foaming agents, driers, plasticizers, ultraviolet stabilizers, ence the colloidal stability of the paint and lead to surface defects, while
anti-skinning agents, biocides, flow-modifying agents, etc. the anisotropic deformation of wood due to humidity interferes with
A consensus in the scientific community is that enhancing the adhesion. Paint adhesion will likely also be very different for, e.g.,
resistance of modern paints and coatings to the harmful effects of solar sandstone versus glass, while the absorption capacity of the substrate
and ultraviolet radiation is a key challenge. Organic polymers are pho­ will also influence the thickness of paint layers. The type and quality of
todegraded through interactions between photons in sunlight and the substrate is not usually taken into account in selecting suitable paints
organic bonds, with/without the participation of oxygen [3,7] by means for urban artworks, possibly due to the scant research on substrate in­
of photolysis, photooxidation and/or photoreduction. These reactions fluence on paint durability. In the research into paints subjected to
lead to the formation of very reactive free radicals in the polymer, accelerated ageing cited above, substrate was generally not taken into
which, in turn (and depending on polymer type, molecule length, tem­ account. The only exception [18] was a study of the resistance to pho­
perature and film thickness) cause chain cleavage, cross-linking and the todegradation of paints applied to three different rock types as
formation of low molecular weight intermediates. These processes can substrates.
affect any organic polymer in the pigment or in the binder. In the In this work, we evaluated colour changes, after exposure to artificial
pigment, resistance to photodegradation will depend on the chromo­ daylight for 7 months, in 10 types of paint of different compositions
phore type and concentration; as for the binder, degradation can lead to widely used in urban artworks in Galicia (NW Spain). The paints were
colour changes (because the degradation (loss) of binder leaves the applied to concrete and brick, substrates typically used in contemporary
pigments more exposed to radiation) or to modifications of other optical mural art in this geographical region. After the exposure test, mineral­
properties [3]. ogical, chemical and physical changes in the specimens were studied in
There is an extensive literature on the mechanisms that lead to order to evaluate paint durability in relation to the paint itself, the
photodegradation processes in polymers, coatings and paints. In most substrate and the paint-substrate interaction. Our findings are expected
experimental work on paint photodegradation, polymer samples are to inform artists and commissioners of urban artworks by guiding them
exposed to different levels of light irradiation (ultraviolet (UV) radiation in their choice of paints and substrates in the interest of ensuring
or artificial daylight) in accelerated ageing tests [8–18]. Most of those artwork durability.
experimental studies, as well as basic polymer studies [3–7], describe
the chemical reactions that define different photodegradation mecha­ 2. Materials and methods
nisms: bond cleavage, free-radical formation, chain scissions,
cross-linking and loss of volatile products (alcohols, aldehydes and 2.1. Paints and substrates
carboxylic acids). According to those studies, the mechanisms through
which an organic paint degrades when exposed to sunlight or UV radi­ Visited in a fieldwork phase were around 100 contemporary murals
ation depend on the type of paint (alkyd, acrylic, vinyl, polyurethane, in 4 cities/towns of NW Spain (Vigo, A Guarda, Ordes and Carballo),
phenolic, epoxy, etc) and the structural complexity of the polymer commissioned by local entities with the aim of promoting urban art.
(chain length, unsaturation and aromatic or aliphatic compounds). Poly Selected were 10 widely used differently applied paints from different
(methyl methacrylate) (PMMA) resins are the most photostable, while commercial houses. Table 1 lists the selected paints along with repre­
polyolefins, polystyrene, polyamides and polyurethanes are the least sentative works; 4 were applied by brush (B-type) and 6 were applied by
photostable, with polyethylene and polyvinyl resins falling somewhere spray (S-type), 2 of which were fluorescent (SF-type).
in between [7]. Susceptibility to photodegradation is also influenced by The 10 paints were applied to aluminium foil, left to dry for 1 month
temperature, film thickness [10], the presence of extenders such as TiO2, under laboratory conditions (15 ± 5 ◦ C; RH 60 ± 10 %) and then scraped
whose photocatalytic properties may accelerate degradation of organic and analysed as follows:
paint components [12] and binder-pigment interactions, with the
pigment itself, depending on the range of UV radiation absorbed, (1) X-ray diffraction (XRD), using a Siemens D5000 (Siemens,
potentially acting as a protector against binder photodegradation Munich, Germany), applying the random powder method and
[19–21]. using Cu-Kα radiation with Ni filter. Each mineral phase was
The chemical and physical changes experienced by paints exposed to identified using the X’Pert HighScore (Malvern Panalytical B.V.,
ageing tests also depend on the type of light source used: UV radiation Almelo, Netherlands).
(A, B and C), natural daylight, or artificial daylight. Photodegradation of (2) Diamond crystal Attenuated total reflection (ATR)-Fourier
alkyd-based paints is slower in daylight than in UV-A radiation condi­ transform infrared spectroscopy (FTIR), using a Thermo Nicolet
tions [22], degradation of acrylic paints (ethyl and methyl methacry­ 6700 (Thermo Fisher Scientific, Waltham, MA, USA). Each
late) is slower in UV-A than under UV-B radiation conditions [8], while infrared (IR) spectrum was recorded at 2 cm− 1 resolution over 32
red and blue alkyd-based spray paints have recently been demonstrated scans from 400 cm− 1 to 4000 cm− 1.
to deteriorate faster in response to UV-C compared to UV-B radiation
[18]. After initial characterization, the paints were applied to concrete and
Urban artworks are potentially exposed to a risk of deterioration, brick (the most commonly used substrates in urban artworks in NW
given their ongoing exposure to sunlight – although at variable in­ Spain). The concrete was prepared using a granite aggregate of gran­
tensities depending on the geographical orientation of the surfaces. ulometry 20− 50 μm and Cosmos CEM-II/BM (VL) 32.5 N cement. Slabs
From the point of view of the materials that make up urban artworks, it measuring 100 × 100 × 2 cm were prepared, left to set for 1 month (20
would seem sensible not only to use sunlight-resistant paints, but also to ◦
C; RH 70 %), then cut into specimens measuring 7 × 7 × 2 cm for
take into account the substrate type and possible interactions between painting and testing. Brick specimens measuring 7 × 7 × 2 cm were cut
this substrate and the paint. Typical substrates for painted urban art­ from single 24 × 11 × 4 cm hollow bricks, washed and brushed to
works – concrete, wood, ornamental stone, metal, brick, glass, etc – are remove particles adhering to the surface, dried in an oven at 40 ± 0.1 ◦ C
compositionally very different, with their different properties and re­ to constant weight and stored for 1 month under laboratory conditions
sponses to deterioration factors influencing adhesion between paint and (15 ± 5 ◦ C; RH 60 ± 10 %). To identify the constituent mineral phases of
substrate (which depends on substrate absorption capacity and surface the 2 substrate types, these were analysed by XRD following the pro­
properties) and resistance of both paint and substrate to environmental cedure described above.

2
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Table 1
Trade names and manufacturers of the 10 selected paints, urban artworks where applied and composition according to manufacturer (n.s, not supplied).
ID TRADENAME (MANUFACTURER) COLOUR ARTWORK COMPOSITION

BY P7-Reveproa Matte Silk Outdoor Primary Yellow Xestando – Lidia Cao (Ordes, 2018) PY138 (quinophthalone yellow)
Yellow (PROA)
BR Teppisol Matte Fire Red (EGA) Red Génesis y Gnosis – Peri Helio (Vigo, 2015) n.s.
BG Green RAL 6024 (KROMO) Green Artemisa vulgaris – Doa Oa (Carballo, 2018) n.s.
BP Pink RAL 4010 (KROMO) Pink Artemisa vulgaris – Doa Oa (Carballo, 2018) n.s.
SY Light Yellow RV 1021 (MONTANA Yellow A Guarda escrita nas estrelas – Nuvi & Titanium (IV) oxide-rutile, monoazo yellow, disazo pyrazolone
COLORS) Éxfico (A Guarda, 2018) orange
SP Tutti Frutti RV 151 (MONTANA COLORS) Pink A Guarda escrita nas estrelas – Nuvi & Rutile, transparent yellow iron oxide, monoazo BONA
Éxfico (A Guarda, 2018) manganese blue, quinacridone red 122
SO Orange RV 2004 (MONTANA COLORS) Orange A Guarda escrita nas estrelas – Nuvi & Rutile, transparent yellow iron oxide, disazo pyrazolone orange
Éxfico (A Guarda, 2018)
SG Valley Green RV 6018 (MONTANA Green A Guarda escrita nas estrelas – Nuvi & Rutile, monoazo yellow, transparent yellow iron oxide,
COLORS) Éxfico (A Guarda, 2018) chlorinated copper phthalocyanine
SFO FLUOR Orange (MONTANA COLORS) Fluor Minero asturiano de padre gallego – Spok n.s.
Orange (Ordes, 2015)
SFG FLUOR Green (MONTANA COLORS) Fluor Minero asturiano de padre gallego – Spok n.s.
Green (Ordes, 2015)

2.2. Mockup preparation To investigate mineralogical, chemical and physical changes that
could explain colour changes, mockups for which colour changes were
Substrates were painted with the different paints following the visible to researchers were analysed. It was unanimously found that
manufacturer’s recommendations regarding application. The 4 B-type perceived colour changes corresponded to a threshold of ΔE*ab>5 CIE­
paints were applied by brush in as many layers as necessary to ensure Lab units, corroborating previous researchers [24], according to whom a
complete coverage: 1 application for red, pink and green (BR, BP and colour change is visually perceptible to an inexperienced observer when
BG, respectively) and 3 applications for yellow (BY) at intervals of 48 h. ΔE*ab>3.5 CIELab units, while two different colour changes can be
The 4 S-type green, orange, pink and yellow paints (SY, SP, SO and SG, definitely distinguished when ΔE*ab>5 CIELab units.
respectively) and 2 SF-type green and orange paints (SFG and SFO, Consequently, for the mockups for which ΔE*ab>5 CIELab units, the
respectively) were applied by spray – at a 45◦ angle at a distance of 30 following analyses were performed before and after the exposure test:
cm from nozzle to substrate surface – for the time necessary to fully and
evenly cover the surface. Mockups were left to dry for 1 month in a dark (1) Micromorphological and compositional analysis, via scanning
laboratory environment. electron microscopy (SEM) with energy-dispersive x-ray (EDX)
using a QUANTA 200 in both secondary electron (SE) and back­
scattered electron (BSE) detection modes. Observation conditions
2.3. Artificial solar ageing test
were a working distance of 9− 11 mm, accelerating potential of
20 kV and specimen current of ~60 mA. The specimens were
The mockups were irradiated using 4 300 W OSRAM Ultra Vitalux
studied by observing the surface directly, and by studying the
artificial daylight bulbs placed at a vertical distance of 50 cm from the
paint layer in cross-sections to the mockup surface; for this,
surfaces of the mockups and at a distance from each other of 30 cm to
specimens embedded in epoxy resin (EpoThin 2 Epoxy Resin and
ensure even and equal irradiation of all the specimens. Testing consisted
EpoThin 2 Epoxy Hardener) then cut and polished to a mirror
of 210 repeated cycles of 22 h of exposure to radiation followed by 2 h of
shine.
darkness (representing 4620 h of exposure in total). The maximum
(2) Analysis of water-solid contact angle of the paint surface, using a
temperature reached on the mockup surfaces was 40 ± 1 ◦ C.
Phoenix-300 Touch SEO contact angle analyser, with an outer
diameter of 0.012′′ (0.3048 mm) and an inner diameter needle of
2.4. Analyses 0.005′′ (0.127 mm). Measured were 3 drops of 8 μL per mockup
following Spanish standards [25].
Changes in mockup appearance and colour were monitored as (3) Adherence to the substrate, evaluated by a peeling test –
follows: following [26] – based on using double-sided Tesa Powerbond
adhesive tape, applied and removed 3 times consecutively to each
(1) Stereoscopic microscope visualization of mockups before and mockup. Peeled-off material was determined as the tape weight
after the exposure test using a Nikon SMZ800. differences for averages per mockup before and after application.
(2) Colour characterization of mockups in the CIELab and CIELCH (4) Mineralogical analysis by XRD, using the equipment and method
spaces [23] before the exposure test, during the exposure test described above. Analysed for each mockup was the powder
(every 15 days, representing 330 h of irradiation), and after obtained by scraping the paint surface, taking care not to remove
testing (4620 h of irradiation). Measurements (6 per mockup) material from the substrate.
were made using a Minolta CM-700d spectrophotometer with a (5) ATR-FTIR analysis, using the equipment and method described
pulsed xenon lamp (with UV cut filter), in specular component above, to analyse the powder obtained from scraping the paint
excluded (SCE) mode, spot diameter 3 mm, illuminant D65 and surface.
observer angle 10◦ . Variables were measured as follows: lumi­
nosity L*, which takes values from 0 (black) to 100 (white); a* 3. Results
(positive values for red and negative values for green); b* (posi­
tive values for yellow and negative values for blue); C* (chroma); 3.1. Pre-test characterization of paints, substrates and mockups
and h (hue). To evaluate colour changes, differences were
measured as ΔL*, Δa*, Δb*, ΔC*ab, ΔH*, while ΔE*ab reflected Table 2 shows the mineralogical composition of the concrete and
overall change, with the original mockup colour taken as the brick substrates. The concrete was composed of quartz, potassium
reference value. feldspar, sodium plagioclase and micas (the aggregate) and calcium

3
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Table 2
Mineralogical composition (presence) of paints and substrates (by XRD) and organic composition (by FTIR) (main group). Q: quartz, M: micas; KF: potassium feldspar;
P: plagioclase (albite); T: talc; C: calcite; B: barite; K: kaolinite; Cl: chlorite; TiO2: titanium dioxide; G: goethite (n.a.: not applicable; PVA: polyvinyl acetate).
ID Q M KF P T C B K Cl TiO2 G OTHER ORGANIC PHASE

Concrete x x x x x n.a.
Brick x x x x n.a.
BY x x x x PVA
BR x x x PVA
BG x x PVA
BP x x PVA
SY x x x K2Cr2O7 Alkyd
SP x Alkyd
SO x x Alkyd
SG x x Alkyd
SFO x x Acrylic
SFG x x Acrylic

carbonate. No C-S-H phases or non-hydrated cement phases were The position of the band corresponding to C– – O (carbonyl) stretch­
detected. The brick was composed of quartz, microcline, biotite and ing was used to differentiate PVA from alkyd and acrylic paints: in PVA
albite. paints, this band occurred at a slightly higher wavenumber (1728− 1735
Table 2 also shows the mineral phases detected in the 10 paints as cm− 1) [27], as was also observed for the BY, BR, BG and BP paints.
applied to aluminium foil and confirms a remarkably different compo­ Likewise, the band corresponding to C(=O)-O ester at around 1231
sition. B-type paints all contained calcite and SY, SO and SG paints had cm− 1 also made it possible to discriminate between the different paint
traces of goethite, an inorganic pigment typical of reds and yellows types: this band appears at a slightly higher wavenumber for alkyd
colours. TiO2 and talc were also identified in 8 and 5 of the 10 paints, paints [27]. Thus, for the SY, SP, SO and SG paints this band lays be­
respectively. Barite was occasionally detected in BY, while phyllosili­ tween 1250 cm-1 and 1255 cm-1, and between 1227 cm-1 and 1231 cm-1
cates (chlorite and kaolinite), also commonly used as fillers, were for the remaining paints. Another typical fingerprint for alkyd paints is
evident in BR and BY paints. Potassium chromate was detected in SY the existence of characteristic doublets at 1602− 1584 cm− 1 – assigned
paint (calcium, potassium and zinc chromates are frequently used as an to C–– C aromatic stretching [6,28] – and at 1114− 1066 cm− 1 – assigned
organic pigment in some yellowish paints, e.g., PY36). to C-H in plane bending [27]. Doublets were clearly identified only in
Table 2 also shows the generic polymer of the organic phase for each the SY, SO, SP, SG paints (Table 3).
paint, while Table 3 shows the FTIR fingerprint bands used to identify The SO, SP, SG and SY paints differed from the B-type paints in the
the different paint components (polymer base, chromophores, fillers shape of the bands assigned to the stretching vibration of the C–H group
and, occasionally, inorganic pigments). in the region between 2850 cm− 1 and 2960 cm− 1. In the S-type paints,
B-type paints typically have a polyvinyl acetate (PVA) base, S-type this vibration manifested in 2 intense bands, differentiated at 2854 cm− 1
paints are alkyd-based and SF-type paints are acrylic-based. Identifica­ and 2925 cm− 1, typical of aliphatic C–H stretching due to alkyd poly­
tion was carried out taking into account the typical fingerprint bands of mer oil content. In the B-type paints, 2 bands were observed at higher
each large group of organic polymers (Table 3). The bands are also frequencies (2878 cm− 1 and 2927 cm− 1), with the small shoulder at
depicted in Fig. 1, which shows spectra for a representative paint for 2954 cm− 1 typical of vinyl paints [6,27].
each group (PVA-based BR, alkyd-based SP and acrylic-based SFG). Another band that discriminated between the different paint types

Table 3
FTIR fingerprint bands used to identify organic phase compounds in the paints. FTIR bands assigned to inorganic fillers and consulted references are also indicated.
BOND TYPE WAVENUMBER (cm− 1) ASSIGNED TO REFERENCE PAINT/ID

O-H stretching – intramolecular bonded 3615− 3670 OH of phyllosilicates and/or FeOOH [30,31] BY, BR, SY, SO, SG, SFO, SFG
O-H stretching vibration 3430− 3300 Acryl, PVA, alkyd [15,16,30] SFO, SFG, SN,
N-H stretching Doublet 3295− 3380 Amide [30] Chromophore
2925
C-H stretching Acryl, PVA, alkyd [27,30] BY, BR, BG, BP, SY, SP, SO, SG, SFO, SFG
2885
1721− 1723 Alkyd and acrylic resins SY, SP, SO, SG, SFG, SFO
C=O carbonyl stretching [27]
1728− 1735 PVA BY, BR, BG, BP
N-H def. primary amide 1645− 1556 Azo group [30] SY, SG, SFO, SFG
C=C aromatic group doublet 1602− 1584 Alkyd [27,28] SY, SP, SO, SG
N=N 1505− 1555 Azo compounds [29] SY, SG, SFO, SFG
C-H bending; C=C aromatic stretching 1485 – [30] SFO, SFG
C-O asymmetric stretching 1420 Calcite BY BR, BG, BP
1434
C-H bending PVA [27] BY, BR, BG, BP
1374
1227− 1235 Acryl and PVA BY, BR, BG, BP, SFO, SFG
C(=O)-O ester [27]
1250− 1252 Alkyd SY, SP, SO, SG,
C-H/C-C/C-O/C-C(=O)-O stretching 1150 Acryl [27] SFO, SFG
C-H/C-C/C-O/C-C(=O)-O stretching doublet 1114− 1066 Alkyd [27] SY, SP, SO, SG
C-H/C-C/C-O/C-C(=O)-O stretching 1026− 1020 PVA [27] BP, BV
Si-O 1018 Talc [30] BY, BR, SY, SFO, SFG
OH out-of-plane deformational mode 905 FeOOH [34,35] SY, SP, SG, SO
C=O carbonate 875 Calcite [30] BY, BR, BG, BP
C-H bending of aromatic units 750− 728 Acryl, alkyd [15,30] BY, BR, SP, SO, SP, SG, SFO, SFG
C-O out-of-plane 714 Calcite [30] BY, BR, BG, BP
Ti-O 690− 614 TiO2 [33] BG, BP, SY, SP, SO, SG
Si-O 670 Talc [30] BY, BR, SY, SFO, SFG
Fe-O stretching 460 FeOOH [30,34] SY, SO, SP, SG

4
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Fig. 1. FTIR spectra (absorbance) of SP, SFG and BR paints scraped from the aluminum foil, BR applied to brick (BR-B) and to concrete (BR-C).

was located in the fingerprint band of C–O stretching vibration between manufacturer about the nature of the SFG and SFO paints, but the FTIR
1200 cm− 1 and 900 cm-1. Distinctive was absorption at around 1026 cm- information confirmed them as having azo-type chromophores.
1
and around 1150− 1170 cm-1 for the PVA-based and acrylic-based Regarding paint fillers and inorganic pigments, talc was identified by
paints, respectively [27,28]. Observed in the BP and BG paints was ab­ FTIR (confirming XRD results) in the 5 paints (BY, BR, SY, SFO and SFG)
sorption at around 1020 cm-1 that was not evident for the remaining from the bands at 3670 cm− 1 (O-H stretching), 1018 cm− 1 (asymmetric
paints (in BY and BR this effect overlapped with that for talc), while in Si-O-Si stretching) and 670 cm− 1 (O-H stretching) [30–32]. The pres­
the SFO and SFG paints, a peak was observed at 1150 cm-1 that was not ence of TiO2 was corroborated by FTIR only for 6 paints (BG, BP, SY, SP,
evident for the remaining paints. SO and SG), from the wide band lying between 614 cm− 1 and 690 cm− 1
For the SY, SG, SFO and SFG paints, a band was clearly detected [33]. The presence of calcite was confirmed by FTIR in the 4 paints for
between 1505 cm− 1 and 1556 cm-1 that could be assigned to the which it was also detected by XRD (BG, BP, BY and BR), from the bands
stretching vibration of the N–– N azo group (Table 3) [29]; these same at 1420 cm− 1 but especially at 875 cm− 1, which does not overlap with
paints had a band between 1644 cm− 1 and 1556 cm− 1 that was assigned any organic component bands. Oxyhydroxylated forms of iron
to the N–H bending of primary amide [30]. The SFO and SFG paints also (FeOOH-goethite type) added as an inorganic pigment were verified by
showed a doublet at 3295 cm− 1 and 3380 cm− 1 that corresponded to the FTIR for 3 paints (SY, SG and SO). In the SP mockups, where this
N–H stretching vibration of the amides. All these bands could be pigment was not identified by XRD, there appeared to be FTIR evidence
attributed to azo-type chromophore compounds, as, according to the of its presence in that the characteristic 905 cm-1 band was detected [34,
manufacturers, SY, SP, SG and SO paints have mono- and diazo-type 35].
chromophores. We were unable to obtain information from the Fig. 2 shows images of mockup surfaces, where the appearance after

5
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Fig. 2. A-F: Stereoscopic micrographs of the surfaces of BR on concrete (A) and on brick (B), SY on concrete (C), SP on brick (D), SFO on concrete (E) and SFG on
brick (F). G-J: SEM micrographs of the surfaces of SO on concrete (G) and on brick (H) and SFO on concrete (I) and SFG on brick (J).

6
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

applying the paints were found to differ depending on the paint and the mockups there were numerous bubbles measuring 70 μm in diameter
substrate. In the BR concrete and brick mockups, small bubbles were (Fig. 2-G), while in the SO brick mockups, few bubbles there were
observed (Fig. 2-A and 2-B). Bubbles were also observed in SY and SP (Fig. 2-H) measured around 15 μm in diameter.
mockups, but differing depending on the substrate: in the concrete SEM also confirmed the formation of craquelure in the SF paints
mockups (Fig. 2-C), the bubbles were large (up to 2 mm in diameter) and (Fig. 2-I and 2-J), less dense (but with deeper and more open cracks) in
a few had burst, whereas in the brick mockups (Fig. 2-D) the bubbles concrete (Fig. 2-E and 2-I), and denser (but with finer cracks) in brick
were much smaller and more numerous and most had burst, leaving (Fig. 2-F and 2-J).
many holes on the surface. In the SFO and SFG mockups, an apparently Visualization of mockup cross-sections also confirmed differences
denser craquelure developed in the concrete mockup (more fissures per between the 3 paint types (Fig. 3):
unit area) compared to the brick mockup (Fig. 2-E and 2-F). The same
figures show the influence of surface roughness on the final finish: (1) Thickness. The 3 types of paints were thicker when applied to
concrete surfaces were rougher than brick surfaces (e.g., Fig. 2-A versus concrete, especially B-type paints (150 μm on concrete and 30 μm
2-B). on brick; Fig. 3-A) and SF-type paints (125 μm on concrete and 40
SEM observations of mockup surfaces corroborate the above obser­ μm on brick; Fig. 3-B) and less so for S-type paints (very similar
vations, confirming the differences in roughness (Fig. 2-G and 2-H). for both substrates at 50− 70 μm).
Likewise, the formation of bubbles was also evident, with larger bub­ (2) Texture. BR (PVA-based) was made up of an organic phase (low
bles on concrete (Fig. 2-G) than on brick (Fig. 2-H). For example, in the contrast according to the BSE detector, indicating great richness
BR mockups, bubbles left holes measuring 40 μm in diameter on con­ in C) that agglutinated a large quantity of Ca-rich and Si, Mg and
crete, but measuring less than 15 μm on brick. For the SO concrete O-rich particles (corresponding to calcium carbonate and talc

Fig. 3. A-B: Stereoscopic micrographs of cross-sections of BR on concrete (A) and SO on brick (B). C-F: SEM micrographs of BR on concrete (C) and on brick (D), SO
on concrete (E) and SFO on brick (F).

7
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Table 4
L*, C*ab and H* variations after artificial solar ageing testing and global colour change (ΔE*ab) in paints applied to concrete and brick. In brackets, change behaviour: l,
linear; q, quick initial change; -, undefined. In grey colour, samples with ΔE*ab>5 CIELab units.

fillers, respectively) measuring 1− 30 μm (Fig. 3-C and 3-D). In and were assigned to talc, appeared wider in the paints on both sub­
contrast, the alkyd-based SO paint (Fig. 3-E) looked very strates (Fig. 1). This could be attributed to 2 kinds of contributions:
different, with an organic phase (very low contrast under BSE
detector, so very rich in C) of granular texture; these granules (1) Silicates in both substrates (quartz, feldspars and micas in con­
appeared surrounded by very small titanium particles (high crete, and feldspars and micas in brick). The Si-O bond vibration
contrast under BSE, sizes less than 1 μm). Similarly, the SFO and mode in the different structural configurations of those minerals
SFG paints (Fig. 3-F) also had a different texture than the other ranges between 1100 cm− 1 and 1066 cm− 1 [31,36].
paints, as they were formed of very low contrast C- and S-rich (2) Hydrated C-S-H gel, with absorption at around 976 cm− 1 [37].
grains with well-defined edges, embedded in a matrix composed Note, however, that, on comparing the intensity ratios of the
of small titanium fragments (less than 1 μm) and, occasionally, Si- bands of the organic phase and of the peaks corresponding to the
, O- and Mg-rich grains (corresponding to talc). Another impor­ silicates, the intensity of the bands assigned to silicates was
tant difference between the 3 types of paints was the amount of observed to be higher in paints applied to concrete than to brick;
inorganic fillers. B-type (PVA-based) paints (Fig. 3-C and 3-D) this could be due to the dragging of substrate particles during the
had more filler than S-type (alkyd-based) paints (Fig. 3-E), scraping of the paint. This possibility is remote in the case of
while the SF-type (acrylic-based) paints (Fig. 3-F) had the least brick, much smoother and more resistant to scraping than
amount of inorganic fillers. concrete.

The FTIR spectra of the paints applied to brick and concrete differed
3.2. Paint durability during and after artificial daylight ageing
from the spectra of the paints themselves. The band at around 1018
cm− 1, which corresponded to the asymmetric Si-O-Si stretching mode
Table 4 shows colour parameter variations, after 210 days of

Fig. 4. Variations in ΔE*ab and ΔC*ab for SO and SFG on concrete during exposure.

8
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

artificial daylight ageing. ΔL*, ΔC*ab, ΔH* and ΔE*ab all underwent Cross-section visualization under SEM confirmed a physical separa­
modifications reflecting colour changes. The mockups were divided into tion between the paint layer and the substrate for brick mockups of all
2 groups according to the ΔE*ab reached at the end of the test: the group the paints in this group; this separation measured as much as 25 μm for
that showed ΔE*ab<5 CIELab units and the group that showed ΔE*ab>5 SO (Fig. 6-D). In cross-sections of the SFO and SFG mockups, SEM evi­
CIELab units. As mentioned, ΔE*ab = 5 CIELab units was established as denced no textural or compositional differences in paint layers in depth
the threshold value at which colour changes become visible to the from the exposed surface that would explain the loss of colour observed
human eye. under the stereoscopic microscope (Fig. 5-G and 5-H).
The mockups that showed ΔE*ab<5 CIELab units were BP, BG, BY, The physical separation of the paint layer may partly account for the
SG, SY and SP applied to concrete and brick. The colour change in the results of the peeling test before and after exposure to artificial daylight
green and pink paints (BG, BP, SG and SP) was greater for the concrete exposition (Fig. 7-A). Several observations can be made in relation to
mockups than for the brick mockups, while the reverse occurred for the Fig. 7-A:
yellow paints (BY and SY). For all the paints with the exception of SY,
the greatest contribution to ΔE*ab was chroma C*ab (which suffered a (1) The amount material extracted for SFG, SFO and, most especially,
reduction, causing a colour shift towards grey that was independent of for BR was greater for concrete than for brick, both before and
the substrate), while for SY, the greatest contribution to ΔE*ab was hue after exposure; no differences between substrates were detected
H* (which increased). for SO. The greater amount of material extracted may be related
The mockups that suffered ΔE*ab>5 CIELab units were BR, SO, SFO to the fact that the adhesive tape dragged away grains of aggre­
and SFG applied to concrete and brick. In these paints, colour changes gate from the concrete; for brick specimens (smoother and more
were greater for brick than for concrete mockups, except for SO, where compact), only paint, which weighs less, would has been
the reverse was the case. The changes in SFO and SFG were notably removed.
greater and occurred much faster than in BR and SO. For SFG and SFO, (2) The amount of material extracted for SFG and SFO on either
ΔE*ab reached 5 CIELab units a mere 15 days after starting the test, substrate and for BR only on concrete was greater after exposure
compared to 90 days for the SO concrete mockup, 105 days for the SO to artificial daylight. This is striking because, observed in the
brick mockup, 150 days for the BR concrete mockup and 135 days for brick mockups but not in the concrete mockups, was a clear post-
the BR brick mockup. The parameter that most contributed to change test separation between paint and substrate that would facilitate
was again C*ab, which suffered a reduction, indicating colour fading. material removal by the adhesive tape.
Within the group of mockups in which ΔE*ab>5 CIELab, paint colour (3) Greater amounts of material were removed for SFO and SFG,
parameters evolved differently over time, as follows: whose pre-test craquelure intensified after exposure to daylight.
(4) Although material losses were greater after the exposure test, the
(1) Linear behaviour (Fig. 4-A and 4-C; (l) in Table 4). The increase or pre- and post-test differences were only statistically significant
decrease was progressive and constant for almost all BR and SO for the BR concrete mockups, while, for the remaining cases,
colour parameters for concrete and brick. standard deviation values for the 3 removals of material were
(2) A quick initial change (increase or decrease), during the first 45 very high both before and after the exposure test.
days, which then slowed down ([q] in Table 4). This occurred in
the ΔE*ab evolution of SF paints and in the H* evolution of SO Paint deterioration was also evident from data on variations in the
paint. In the case of SF paints, the evolution of ΔE*ab over time surface contact angle (Fig. 7-B). Before artificial daylight exposure,
was determined by the behaviour of chroma (C*ab) (Fig. 4-B and contact angles were greater than 90◦ (corresponding to hydrophobic
4-D). surfaces) for the SFO and SFG mockups (and greater for brick than for
(3) Undefined behaviour ((-) in Table 4). No clear trend was detected concrete), but barely reached 80◦ for BR and SO. After artificial daylight
in colour parameter change evolution, as exemplified by the exposure, contact angles were reduced for all paints, with only SFO on
evolution in the L* coordinate for BR. This behaviour occurred in concrete having a contact angle close to the theoretical hydrophobicity
those samples that suffer the lowest ΔE*ab. value (89.20◦ ±1.11◦ ). In the concrete mockups, the contact angle was
most reduced in SFG and least reduction in SFO; in the brick mockups,
Fig. 5 shows visible colour changes in the surfaces of some of the the greatest reduction in contact angle occurred in SFO and the least
painted mockups before and after the exposure test. Colour changes reduction occurred in BR.
were intense in SF-type paints (SFO and SFG, Fig. 5-A to 5-D) applied to Colour, texture and surface property changes were observed not to be
both concrete and brick. Furthermore, in the mockups initially affected due to mineralogical changes. On comparing pre- and post-test miner­
by craquelure, some loss of the paint layer – greater for brick mockups alogical composition of the paints, there was no evidence of the disap­
than for concrete mockups – left a partially exposed surface along with pearance of mineral phases present in the non-exposed paints or the
apparently unaltered paint (compare Fig. 5-A, 5-B, 5-C and 5-D). Colour appearance of new mineral phases.
changes in BR and SO (Fig. 5-E and 5-F) were less intense, despite these From the molecular point of view, FTIR confirmed changes, although
paints having ΔE*ab>5 CIELab units. Outside of colour changes, the only in the organic phase. All specimens experienced a reduction in
mockups showed no physical deterioration. bands at 3400− 3300 cm− 1 (O-H stretching), 2925 cm− 1 (C-H stretch­
Cross-sections viewed under a stereoscopic microscope confirmed ing), 1721− 1735 cm− 1 (C=O carbonyl stretching) and 1227-1252 cm− 1
that, in the B-type (PVA) and S-type (alkyd) paints, colour changes (C(=O)-O ester). To rule out substrate influence and to quantify and
occurred evenly throughout the paint layer, whereas in the SF-type relate band reductions to paint deterioration, the ratios between the
(acrylic) paints, colour changes occurred in the microns closer to the intensity of FTIR bands of the organic phase of the paints and a FTIR
paint surface (compare Fig. 5-G and 5-H). band of the filler present in the paints (which remained unaltered after
BR, SO, SFG and SFO, as mockups for which ΔE*ab>5 CIELab units, test) were calculated from spectra obtained before and after the test,
were further analysed to explore possible reasons for the deterioration in obtaining the percentage variations of these ratio after the test. If this
colour. SEM study of the surfaces of these mockups confirmed the sig­ percentage is positive, it would mean that the organic polymer FTIR
nificant colour loss in SFO and SFG. In concrete mockups (Fig. 6-A), the band suffered an increase (relative to the unalterable reference band,
craquelure widened after exposure. while in the brick mockups (Fig. 6- that of the filler). If it is negative, it would mean that the polymer band
B), circular segregations rich in titanium (Ti) were clearly observed suffered a decrease, always relative to the filler. Bands corresponding to
(Fig. 6-C). SEM study of the mockups of B- and S-type paints did not the stretching vibration modes of O-H (3394 cm-1), C-H (2925 cm-1) and
evidence any noticeable differences. C(=O)-O (1732 cm-1) were selected as organic phase FTIR bands; the

9
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Fig. 5. Stereoscopic micrographs of mockups before (A, C, E and G) and after (B, D, F and H) testing. A-B: SFO surface on brick; C-D: SFG surface on concrete; E-F:
cross-section of SO on brick; G-H: cross-section of SFO on brick.

10
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Fig. 6. A and B: SEM micrographs (BSE detector) of SFO on concrete (A) and on brick (B); C: spectrum obtained by EDS of SFO on brick before (left) and after (right)
testing. D: Cross-section of SO on brick.

FTIR band used as reference was the band at 1018 cm-1 (asymmetric Si- loss of carbonyl groups in the polymer. However, no widening of the
O-Si stretching mode) assigned to talc, whose intensity should remain carbonyl peak towards lower wavenumbers (at around 1700− 1680
the same after the test, given that talc is inert to solar radiation. Since cm− 1) was detected, and, according to the literature, this loss leads to
talc was also absent in the 2 substrates, any influence of brick or con­ the formation of ketone groups [10,11,21].
crete composition could be ruled out in calculating FTIR intensities Notable was the reduced intensity in the band at 3400− 3300 cm− 1
ratio. In the case of BR mockups, the band at 670 cm-1 (O–H stretching) (O-H stretching). In most published studies, organic polymer deterio­
of talc was selected because the band assigned to C(=O)-O stretching ration has usually been associated with increased intensity of that band
vibration for PVA (1020 cm-1) overlaps the band at 1018 cm-1 of the reflecting oxidation processes (formation of alcohols, carbonyls, perox­
Si–O group of talc. For SO paint, changes could not be quantified ides, etc). The disappearance of this band may indicate a rapid reaction
because titanium (614 cm-1 and 690 cm-1) – the only inert inorganic of radicals with free OH groups - formed from scission reactions - to form
compound that could be used as reference – overlaps with the C–H ester and ketones. The non-increase or widening of the carbonyl group
stretching and bending vibrations of the aromatic ring of the alkyd base band, indicating a subsequent ketone-formation reaction, would sug­
of the paint. gest, as a more likely hypothesis, the reduction of the O–H stretching
Table 5 shows percentage variations of the calculated FTIR in­ band at 3400− 3300 cm− 1, and a loss of low-molecular weight molecules
tensities ratio. In all the cases, the calculated ratios were reduced. The through volatilization.
percentage variation varied depending on the substrate: for BR on
concrete, the band that changed most – it almost disappeared – was 4. Discussion
O–H stretching (3400− 3300 cm− 1), while for BR on brick, the band
that changed most was the C– – O carbonyl stretching (1721− 1735 Characterization of the 10 different paints applied to brick and
− 1
cm ). SFO and SFG also behaved differently depending on the sub­ concrete revealed compositional differences: brush-applied paints were
strate: in concrete mockups, the most changed band was that of C– –O PVA-based, spray paints were alkyd-based and fluorescent paints were
carbonyl stretching (1721− 1735 cm− 1), while in brick mockups this acrylic-based. FTIR confirmed that the paints for which the factory
band was reduced by 81.1 % for SFO but only 13.4 % for SFG. The composition was known had azo-type chromophores, while FTIR points
reduction on the intensity of organic phase FTIR bands after the test may to possibly azo for several paints for which no manufacturer information
indicate volatilization of small molecules resulting from radiation- was available (SFG and SFO). Mineralogically, the usual minerals were
induced polymer degradation [10]. More specifically, the reduction of used as fillers and extenders. All the B-type paints had calcite as a filler,
the intensity of C–H stretching band at 2925 cm− 1 suggests hydrogen while some also included talc (BY and BR) and others included TiO2 (BP
abstraction processes. Some authors associate the change in this band and BG). S-types paints all contained TiO2 which can act as both a white
with destruction of the backbone structure of the polymer [10,11]. pigment and an opacifier, and SG, SO and SY included goethite as a
Regarding the band at 1721 cm− 1, the reduction would suggest partial possible inorganic pigment. Fluorescent paints (SFO and SFG)

11
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

Fig. 7. Results of peeling test (A) and contact angle measurements (B) for BR, SO, SFG and SFO on concrete and on brick before and after testing.

paint and the substrate, as concrete and brick are both porous and also
Table 5
have an absorption capacity that could favours bubble formation during
Percentage variations after the test of the ratio between the intensity of the
application. The size of the bubbles may also be related to the size of the
organic phase bands (3394 cm− 1, 2925 cm− 1 and 1732 cm− 1) and the intensity
pores in the substrate: SEM observations of the cross-sections showed
of the reference band considered (talc; 670 cm− 1 for BR and 1081 cm-1 for SO,
SFG and SFO). Values for concrete and brick mockups are shown. that porosity in concrete was more heterometric than in brick, which,
compared to concrete, has pores that are much smaller in diameter.
CONCRETE BRICK
RATIO There could also be a relationship with the paint formula, specifically
BR SFG SFO BR SFG SFO with surface active agents with emulsifying properties [4], as these
OH str. (3394) /Si-O − 94.5 − 81.1 − 25.9 − 0.5 − 10.1 − 13.4 compounds favour the formation of foam bubbles that remain intact
(talc) during drying, so their absence should facilitate foam bubble breakage
CH str. (2925) / Si-O − 76.9 − 38.9 − 27.8 − 35.2 − 35.5 − 42.0 during drying. However, this did not seem to happen in our study, as the
(talc)
C=O carbonyl str. − 43.9 − 77.3 − 60.6 − 68.9 − 13.2 − 80.3
same paint formed broken bubbles in brick and whole bubbles in con­
(1732) / Si-O (talc) crete, representing, therefore, a significant influence on the substrate.
Craquelure is related to shrinkage during polymer drying processes
[6] – in this study, during oxidation – or to an excess of paint. Inter­
additionally included talc. estingly, however, in our case, this only occurred with the 2 fluorescent
The paint finishes were different once applied to the concrete or paints (SFO and SFG). Since craquelure did not occur in S-type paints,
brick. PVA-based B-type paints (the only paints that had calcite as a we can rule out any link to the application method, because both (SF and
filler) obtained a smooth even finish, with no defects. The finish for S type) are applied as a spray. Furthermore, since craquelure occurs a
alkyd-based S-type and acrylic-based SF-type paints (with TiO2 as a few seconds after application (with the paint still fresh), so retraction
filler) had defects: bubbles in the yellow and pink alkyd paints (SY and and contraction phenomena associated with oxidation due to drying can
SP) and craquelure in the fluorescent paints (SFO and SFG). The be ruled out. It seems more likely, considering the speed of crack for­
different failure morphologies could be related to the different nature of mation just after application, that the craquelure formation is related to
the paint organic phases (craquelure- acrylic paints; bubbles-alkyd rapid solvent evaporation.
paints) although the paint formulation seems to be more decisive The presence of bubbles in the SY and SP paints did not seem to
since bubbles form in only two (SY and SP) of the 4 alkyd S-type paints. compromise their resistance to deterioration, since none of the corre­
The influence of the substrate on defect intensity was clear: 1) craque­ sponding mockups underwent colour change of ΔE*ab>5 CIELab units.
lure affecting SF-type was denser in brick than in concrete, although Colour resistance seems to have been due to the alteration of the organic
cracks were wider and deeper in concrete samples; 2) bubbles were phase of the paints. There were no mineralogical changes and, from the
larger on concrete and smaller on brick. chemical point of view, FTIR confirmed a reduction in the typical bands
Bubble formation could be related to the interaction between the

12
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

of C–H, O-H and C=O functional groups, suggesting the destruction of except for SO (in this case, greater for concrete). While there was
the backbone structure of the polymer and the loss of compounds of low no evidence that the paints photodegraded more when applied to
molecular weight through volatilization. The loss of an organic phase brick, 2 arguments would support that brick can favour the colour
was very evident in one of the most degraded paints, SFO (Fig. 6-B), with change to be more intense:
SEM detecting on the corresponding surface a relative enrichment of Ti. (a) The paint layer on brick was less thick than on concrete. For
The reduction of the contact angle in all the paints after the exposition, the same irradiation and exposure time conditions, paint
specifically in those of water-repellent character (SF- type), was clear applied less thickly deteriorates more intensely and to a
evidence of the deterioration of the organic (non-polar) phase. greater degree.
Colour fastness during testing appeared to depend on 3 factors: (b) SEM confirms a clear separation between the paint layer and
the brick in all the brick mockups. Polymers adhere less well
(1) Type of organic base. The 2 acrylic-based paints (SFG and SFO) to smooth surfaces than to rough surfaces [6] and brick is
experienced the greatest colour change (ΔE*ab>40 CIELab units), much smoother than concrete. Mechanical forces (a third
very different from change in the PVA (B-type) and alkyd (S-type) mechanism to add to primary chemical bonding and sec­
paints (no more than ΔE*ab>10 CIELab units). However, ondary -polar- bonding) that contribute to polymer adhesion
considering that acrylic resins are highly resistant to solar radi­ to substrates are stronger on surfaces with numerically more
ation [7,10], the greater colour deterioration of acrylic-based chemically active locations (a consequence of surface
paints (SFO, SFG) may be related to the stability of the fluores­ roughness and a greater specific surface). In this sense, the
cent pigments. Initial colour variations in these paints were least favourable substrate is brick, which is much smoother
substantial (already at 15 days, ΔE*ab>5 CIELab units, and after than concrete, also having a lower roughness and specific
testing, ΔE*ab>41 CIELab units), so it is reasonable to assume surface. Therefore, it is not unreasonable to think that paint
exponential behaviour and colour stabilization at some point. adhesion on brick was already initially weak; also, brick has a
The great colour change in the SF-type paints may also be related to lower absorption capacity than concrete. During irradiation
the surface failure (craquelure) evident before the exposure test and testing, polymer degradation together with thermal oscilla­
its intensification after exposure to artificial daylight. tions (due to slight increases in temperature) may have led to
Also, another possible reason for the greater degradation of SF-type paint separation from the substrate. What is unquestionable
paints may be the presence of TiO2. It is known that, in polymers is that colour was influenced by the existence of a gap be­
with low photooxidative stability, the scattering effect of UV radia­ tween the paint layer and the substrate.
tion caused by TiO2 particles improves the ultimate response of the
polymer; but, in contrast, for polymer systems with good photooxi­ The peeling test results did not reflect the SEM finding that BR, SFG
dative stability (like acrylic resins), the addition of TiO2 often leads and SFO adhesion to brick decreased with artificial daylight exposure.
to a reduction in resistance to photooxidation through its photo­ This reduced adhesion would imply that the amount of material
catalytic effects [38]. For the SF-type paints, SEM confirmed a rela­ removed after the test were greater for brick than for concrete, yet the
tive Ti enrichment of the surface (Fig. 6-B) that could point to a loss opposite happened, both before and after the exposure test. This may be
of the organic phase. However, there are several reasons that would partially due to the tape removing small grains of aggregate from the
indicate the low probability that TiO2, in SF-type paints, is the factor concrete: the cohesion of the substrate (lower for concrete than for
that contributes the most to colour degradation. First of all, the TiO2 brick) consequently affected test performance, revealing itself as a
polymorph with the highest photocatalytic effect is anatase. The limitation of the method. The fact that, after the test, the amount of paint
other polymorphs rutile and brookite, due to their lighter effective removed by the tape increased in concrete mockups (and not only in
mass, smaller particle size and longer lifetime of photoexcited elec­ brick mockups) indicates that exposure to artificial daylight likely also
trons and holes [39] have a lesser catalytic effect. The TiO2 identified affected paint adhesion to the concrete – although this could not be
by XRD in these samples is rutile so, without ruling out a possible verified by stereomicroscopy or SEM. Peeling test demonstrated the
contribution, it is unlikely that the intense degradation of SF paints is lesser durability of SF paints, as the amount of removed material was
due exclusively to a photocatalytic effect of TiO2. On the other hand, especially high for SF-painted specimens, especially after artificial
S-type paints, of an alkyd composition (and therefore, according to daylight exposure; this lesser durability may be due to the craquelure
the literature, less resistant to UV radiation than acrylic resins [7, observed in the SF mockups that increased after testing, especially in
10]) and also containing TiO2-rutile, suffer a much less intense concrete (in which cracks widened and deepened).
degradation than SF paints. For all these reasons, the most likely
causes of colour degradation of SF paints would be related to 1) a 5. Conclusions
possible greater susceptibility to photodegradation of their specific
fluorescent pigments and 2) the contribution to the colour change of We tested the durability to artificial daylight of 10 different colour
the increase, after the exposure, of craquelure that initially appeared (including fluorescent) paints, as used in urban art, applied by brush or
after the drying of the paints. spray to concrete and brick. Prior characterization of the paints and
(2) Colour (chromophore type). In the PVA-based (B-type) and alkyd- mockups (paint plus substrate) confirmed differing compositions
based (S-type) paints, pink, yellow and green colours resisted regarding the organic base: spray paints (S-type) were alkyd-based,
best, as indicated by their small ΔE*ab values, while orange and brush-applied paints (B-type) were PVA-based and fluorescent paints
red paints resisted worst (ΔE*ab = 8 to 11 CIELab units). It cannot (SF-type) were acrylic-based. There were also differences in the nature
be ruled out that, in addition to the deterioration of the organic and quantity of fillers: SF-type paints were much poorer in fillers than S-
base, the durability of the paints is influenced by the resistance to type and B-type paints, and S-type and SF-type paints were richer in TiO2
sunlight of the chromophore itself. Unfortunately, we did not than B-type paints.
have information on the type of pigment in each paint, nor was Coating failures observed after the application of the paints were
FTIR able to shed any light on the issue related to paint composition (bubbles in S-type paints and craquelure in
(3) Substrate type. In the specimens that experienced less colour SF-type paints) and to the substrate type: craquelure in SF-type paints
change (ΔE*ab<5 CIELab units), no relationship was observed was denser in brick mockups than that in concrete mockups; the cracks
between the magnitude of the colour change and the substrate. In were finer and more superficial in brick samples whereas they were
the specimens that experienced more colour change (ΔE*ab>5 wider and deeper in concrete. No coating failures were evident for B-
CIELAB units), change was greater in paints applied to brick, type paints.

13
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

The artificial daylight caused a colour degradation of the paints that [4] J. Bieleman, Additives for Coatings, WILEY-VCH Verlag GmbH, Weinheim (Federal
Republic of Germany), 2000, 389p.
was related to organic phase degradation: FTIR confirmed a decrease in
[5] P. Bamfield, Chromic Phenomenal: Technological Applications of Colour
the C-H, O-H, C(=O)-O stretching vibrations bands, suggesting chain Chremistry, The Royal Society of Chemistry, 2001, 389p.
scissions and the formation of low molecular weight volatilizing com­ [6] D.G. Weldon, Failure Analysis of Paints and Coatings, John Wiley & Sons, 2009.
pounds. The relative enrichment of Ti on the SF-type paint surface ISBN:9780470697535.
[7] D. Feldman, Polymer weathering: photo-oxidation, J. Polym. Environ. 10 (2002)
confirmed a loss of the organic phase. 163–173.
Colour degradation varied depending paint composition (type of [8] P.M. Whitmore, V.G. Colaluca, The natural and accelerated aging of an acrylic
base and filler), paint colour and substrate type. The loss of colour was artist’s medium, Stud. Conserv. 40 (1) (1995) 51–64.
[9] M.J. Melo, S. Bracci, M. Camaiti, O. Chiantore, F. Pianceti, Photodegradation of
much more intense in SF-type (acrylic) paints (ΔE*ab>41 CIELab units) acrylic resins used in the conservation of stone, Polym. Degrad. Stab. 66 (1999)
than in either S-type (alkyd) or B-type (PVA) paints (ΔE*ab = 7 to 11 23–30.
CIELab units). The formulation regarding pigment seems to influence [10] O. Chiantore, L. Trossarelli, M. Lazzari, Photooxidative degradation of acrylic and
methacrylic polymers, Polymer 41 (2000) 1657–1668.
photostability: of the B-type colours, only red (BR) degraded, and of the [11] O. Chiantore, M. Lazzari, Photo-oxidative stability of Paraloid acrylic protective
S-type colours, only orange (SO) degraded. The intense colour degra­ polymers, Polymer 42 (2001) 17–27.
dation of the SF paints seems to be related to lower pigment photo­ [12] N.S. Allen, M. Edge, G. Sandoval, A. Ortega, C.M. Liauw, J. Stratton, R.B. McIntyre,
Interrelationship of spectroscopic properties with the thermal and photochemical
stability, while coating failure in the form of craquelure undoubtedly behaviour of titanium dioxide pigments in metallocene polyethylene and alkyd
contributed to their lesser durability. Brick substrate adhesion of the based paint films: micron versus nanoparticles, Polym. Degrad. Stab. 76 (2002)
paints experiencing the greatest colour change was null after the expo­ 305.
[13] T. Learner, O. Chiantore, D. Scalarone, Ageing studies of acrylic emulsion paints,
sure test, and this loss of adhesion may contribute to intensifying the
in: R. Vontobel (Ed.), 18th Triennial Metting, Río de Janeiro, 22-27. September
colour change. 2002, Preprints, ICOM Committee for Conservation, 911-19. London: James &
In selecting materials for urban artworks that are intended to last, James, 2021.
administrations and artists need to consider the impact of exposure to [14] D. Scalarone, O. Chiantore, T. Learner, T. Ageing studies of acrylic emulsion paint,
part II: comparing formulations with poly (EA-co-MMA) and poly (n-BA-co-MMA)
daylight. Since SF paints show zero photoresistance, their use is strongly binders, in: James & James, in: I. Verger (Ed.), 14th Triennial Meeting The Hague,
discouraged. Of the spray (alkyd-based S-type) and brush-applied (PVA- 12-16 September 2005, Preprints, ICOM Committee for Conservation, vol 1, 2021,
based B-type) paints, colour (and therefore pigment type) seems to be a pp. 850–857.
[15] M.T. Doménech Carbó, M.F. Silva, E. Aura-Castro, L. Fuster-López, S. Kroner, M.
determining factor in durability/resistance, with reddish-orange colours L. Martínez- Bazán, X. Más-Barberá, M.F. Meeklenburg, L. Osete-Cortina,
seeming to be the least durable/resistant. It is also important to take into A. Doménech, J.V. Gime-no-Adelantado, D.J. Yusá-Marco, Study behavior on
account the substrate and to properly condition artwork substrates simulated daylight ageing of artists’ acrylic and poly (vinyl acetate) paint films,
Anal. Bioanal. Chem. 399 (9) (2011) 2921–2937, https://doi.org/10.1007/
before applying paint, to ensure good adhesion (if brick) and reduce loss s00216-010-4294-3.
of surface material (if concrete). [16] S.M. Cakić, I.S. Ristić, J.M. Vladislav, J.V. Stamenković, D.T. Stojiljković, IR-
change and colour changes of long-oil air drying alkyd paints as a result of UV
irradiation, Prog. Org. Coat. 73 (2012) 401–408.
CRediT authorship contribution statement [17] V. Pintus, S. Wei, M. Schreiner, Accelerate UV ageing studies of acrylic, alkyd and
polynivyl acetate paints: influence of inorganics pigments, Microchem. J. 124
E.M. Alonso-Villar: Conceptualization, Methodology, Data cura­ (2016) 949–961.
[18] P. Sanmartin, J.S. Pozo-Antonio, Weathering of graffiti spray paint on building
tion, Writing - original draft, Visualization, Investigation, Writing - re­
stones exposed to different types of UV radiation, Constr. Build. Mater. 236 (2020),
view & editing. T. Rivas: Conceptualization, Methodology, Data 117736.
curation, Writing - original draft, Visualization, Investigation, Supervi­ [19] A. Ciccola, M. Guiso, F. Domenici, F. Sciubba, A. Bianco, Azo-Pigments effect on
UV degradation of contemporary art pictoral film: a FTIR-NMR combination study,
sion, Writing - review & editing. J.S. Pozo-Antonio: Conceptualization,
Polym. Degrad. Stab. 140 (2017) 74–83.
Methodology, Data curation, Writing - original draft, Visualization, [20] A. Ciccola, I. Serafini, M. Guiso, F. Ripanti, F. Domenici, F. Sciubba, P. Postorino,
Investigation, Supervision, Writing - review & editing. A. Bianco, Spectroscopy for contemporary art: discovering the effect of synthetic
organic pigments on UVB degradation of acrylic binder, Polym. Degrad. Stab. 159
(2019) 224–228.
Declaration of Competing Interest [21] C. Duce, C.V. Della Porta, M.R. Tiné, A. Spepi, L. Ghezzi, M.P. Colombini,
E. Bramanti, FTIR study of ageing of fast drying oil colour (FDOC) alkyd paint
replicas, Spectrochim. Acta A. Mol. Biomol. Spectrosc. 130 (2014) 214–221.
The authors report no declarations of interest.
[22] F. Rasti, G. Scott, The effects of some common pigments on the photo-oxidation of
linseed oil-based paint media, Stud. Conserv. 25 (4) (1980) 145–156.
Acknowledgements [23] CIE S014-4/E, Colorimetry Part 4: CIE 1976 L* a* b* Colour Space, Commission
Internationale d’eclairage, CIE Central Bureau, Vienna, 2007.
[24] W. Mokrzycki, M. Tatol, M, Color difference DeltaE-A Survey, Match Graph Vis. 20
This work was supported by CONSERVATION OF ART IN PUBLIC (2012) 383–411.
SPACES - CAPuS. Erasmus + Knowledge Alliance KA2: Cooperation for [25] UNE EN 828:2013, Adhesives. Wettability. Determination by Measurement of
innovation and the exchange of good practices, 2017-2021. Grant Contact Angle and Surface Free Energy of Solid Surface, Asociación Española de
Normalización (AENOR), 2013.
Agreement Number 2017-3674/001-001. E. M. Alonso-Villar’s research [26] M. Drdácky, J. Lesák, S. Rescie, Z. Slízková, P. Tiano, J. Valach, Standardization of
was partially financed by the same project. J.S. Pozo-Antonio’s research peeling test for assessing the cohesion and consolidation characteristics of historic
was supported by the Spanish Ministry of the Economy and Competi­ stone surfaces, Mater. Struct. 45 (2012) 505–520.
[27] T. Learner, Analysis of Modern Paints, GCI Publications, 2005, 210pp.
tiveness through grant IJCI-2017-32771. FTIR, XRD and SEM analyses [28] H. Lobo, J.V. Bonilla, Handbook of Plastic Analysis, CRC Press, 2003, 656 pp.
were performed at the University of Vigo’s Research Support Centre for [29] F. Ahmed, R. Dewani, M.K. Pervez, S.J. Mahboob, S.A. Soomro, Non-destructive
Science and Technology (CACTI). Ailish Maher translated the manu­ FT-IR analysis of mono azo dyes, Bulg. Chem. Commun. 48 (1) (2016) 71–77.
[30] G. Socrates, Infrared and Raman Characteristic Group Frequencies:Tables and
script and revised the English in a near-final version of the manuscript. Charts, 3rd edn., John Wiley & Sons, Baffins Lane, Chichester, 2001, p. 366p.
[31] P.A. Schroeder, Infrared spectroscopy in clay science, in: A. Rule, S. Guggenheim
References (Eds.), CMS Workshop Lectures. Vol. 11. Teaching Clay Science, The clay Mineral
Society, Aurosa CO, 2002, pp. 181–206.
[32] V. Šontevska, G. Jovanovski, P. Makreski, Minerals from Macedonia. Part XIX.
[1] C. Santabárbara, Street art conservation: beyond mural restoration. OPUS no2/
Vibrational spectroscopy as identification tool for some sheet silicate minerals,
2018, J. Hist. Architect. Conserv. Draw. (2019) 149–164. ISBN13 9788849237122.
J. Mol. Struct. 834–836 (2007) 318–327.
[2] EwaGlos, in: A. Weyer, P. Roig, D. Pop, J. Cassar, A. Özkôse, J.-M. Vallet, I. Srsa
[33] M. Al-Amin, S. Chandra Den, T. Ur Rashid, Md. Ashaduzzaman, S. Shamsuddin,
(Eds.), European Illustrated Glossary of Conservation Terms for Wall PaintIngs and
Solar assisted photocatalytic degradation of reactive azo dyes in presence of
Architectural Surfaces, Michael Imhof Verlag, Petersberg, Germany, 2015. ISBN:
anatase titanium dioxide, Int. J. Latest Res. Eng. Technol. 2 (3) (2016) 14–21.
978-3-7319-0260-7 450p.
[34] P.S.R. Prasad, K. Shiva Prasad, V. Krishna Chaitanya, E.V.S.S.K. Babu, B. Sreedhar,
[3] R. Lambourne, T.A. Strivens (Eds.), Paint and Surface Coatings. Theory and
S. Ramana Murthy, In situ FTIR study on the dehydration of natural goethite,
Practice, 2nd edition, Published by Woodhead Publishing Ltd, Abington Hall,
J. Asian Earth Sci. 27 (2006) (2006) 503–511.
Cambridge CBl 6AH, England, 1999. Woodhead Publishing Ltd.

14
E.M. Alonso-Villar et al. Progress in Organic Coatings 154 (2021) 106180

[35] M. Veneranda, J. Aramendia, L. Bellot-Gurlet, P. Colomban, K. Castro, FTIR [37] M. Chollet, M. Horgnies, Analyses of the surfaces of concrete by Raman and FT-IR
spectroscopic semi-quantification of iron phases: a new method to evaluate the spectroscopies: comparative study of hardened samples after demoulding and after
protection ability index (PAI) of archaeological artefacts corrosion systems, Corros. organic post-treatment, Surf. Interface Anal. 43 (2011) 714–725.
Sci. 133 (2018) (2018) 68–77. [38] A.R. Marrion (Ed.), The Chemistry and Physics of Coatings, 2nd edition, The Royal
[36] P. Makreski, G. Jovanovski, B. Kaitner, Minerals from Macedonia. XXIV. Spectra- Society of Chemistry, 2004. ISBN 978-0-85404-604-1.
structure characterization of tectosilicates, J. Mol. Struct. 924–926 (2009) [39] J. Zhang, P. Zhou, J. Liu, J. Yu, Photocatalytic activity among anatase, rutile and
413–419. brookite TiO2, Phys. Chem. Chem. Phys. 16 (2014) 20382–20386.

15

You might also like