You are on page 1of 15

Journal of the American Institute for Conservation

ISSN: 0197-1360 (Print) 1945-2330 (Online) Journal homepage: https://www.tandfonline.com/loi/yjac20

Sublimable layers for protection of painted pottery


during desalination. A comparative study

Hamada Sadek, Barbara H. Berrie & Richard G. Weiss

To cite this article: Hamada Sadek, Barbara H. Berrie & Richard G. Weiss (2018) Sublimable
layers for protection of painted pottery during desalination. A comparative study, Journal of the
American Institute for Conservation, 57:4, 189-202, DOI: 10.1080/01971360.2018.1537414

To link to this article: https://doi.org/10.1080/01971360.2018.1537414

Published online: 08 Nov 2018.

Submit your article to this journal

Article views: 82

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=yjac20
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION
2018, VOL. 57, NO. 4, 189–202
https://doi.org/10.1080/01971360.2018.1537414

Sublimable layers for protection of painted pottery during desalination.


A comparative study
Hamada Sadeka, Barbara H. Berrieb and Richard G. Weissc
a
Faculty of Archaeology, Conservation Department, Fayoum University, Al Fayoum, Egypt; bConservation Division, National Gallery of Art,
Washington, DC, USA; cDepartment of Chemistry, Institute for Soft Matter Synthesis and Metrology, Georgetown University, Washington,
DC, USA

ABSTRACT KEYWORDS
Cyclododecane (CDD) has been used for several years as a temporary consolidant for organic and Desalination; temporary
inorganic materials of interest in cultural heritage. In this work, a comparative study has been consolidation; sublimation;
conducted with CDD and three other sublimable compounds – menthol, camphene, and elemental surface analysis;
cyclododecanone – as protectants of water-sensitive paint during the desalination of painted soluble salts
pottery. The experiments include salination, consolidation, and desalination, and were conducted
on samples newly prepared in the laboratory. Optical microscopy, SEM-EDS, elemental mapping,
and FTIR data have been used to evaluate the relative efficiencies of the four consolidants. In
addition, the relative rates of sublimation of the consolidants from a glass surface have been
measured. Although all of the consolidants are able to protect the painted layers, there are
distinct differences in efficacy and rates of sublimation that should make some better than
others depending on the specific nature of the objects being treated.

RÉSUMÉ
Le cyclododécane (CDD) est utilisé depuis de nombreuses années pour la consolidation temporaire
des matériaux organiques et inorganiques du patrimoine culturel. Dans cette étude comparative, le
CDD et trois autres composés sublimables, le menthol, le camphène et le cyclododécanone ont été
évalués pour la protection des enduits sensibles à l’eau lors des traitements de dessalement des
poteries peintes. Des échantillons ont été fabriqués en laboratoire et ont été préparés à l’aide de
solutions salines; par la suite, des consolidants ont été appliqués sur les échantillons et un
traitement de dessalement a été effectué. L’efficacité relative des quatre consolidants a été
évaluée par microscopie optique, par microscopie électronique à balayage, par cartographie
chimique et par spectroscopie infrarouge à transformée de Fourier. De plus, des mesures des
taux de sublimation des consolidants déposés sur une surface de verre ont été effectuées. Même
si tous les produits testés permettent une protection des enduits, des différences existent dans
leur efficacité et leurs taux de sublimation, ce qui implique une meilleure performance de
certains selon la nature des objets traités. Traduit par André Bergeron.

RESUMO
O ciclododecano (CDD) é utilizado há vários anos como consolidante temporário de materiais
orgânicos e inorgânicos de interesse para o patrimônio cultural. Neste trabalho, um estudo
comparativo foi conduzido com CDD e três outros compostos sublimados - mentol, canfeno e
ciclododecanona - como protetores da tinta sensível à água durante a dessalinização de
cerâmica pintada. As experiências incluem salinização, consolidação, dessalinização e foram
realizadas em amostras recém preparadas no laboratório. Microscopia óptica, SEM-EDS,
mapeamento elementar e dados de FTIR foram usados para avaliar as eficiências relativas dos
quatro consolidantes. Além disso, as taxas relativas de sublimação dos consolidantes de uma
superfície de vidro foram medidas. Embora todos os consolidantes sejam capazes de proteger as
camadas pintadas, há diferenças distintas na eficácia e nas taxas de sublimação que tornam
alguns consolidantes mais apropriados que outros, dependendo da natureza específica dos
objetos a serem tratados. Traduzido por Tereza Lança, e revisado pelo Beatriz Haspo.

RESUMEN
El Ciclododecano (CDD por sus siglas en inglés) ha sido utilizado por varios años como consolidante
temporal para los materiales orgánicos e inorgánicos que son importantes para el patrimonio
cultural. En este proyecto se llevó a cabo un estudio comparativo con CDD y otros tres
compuestos sublimables — mentol, canfeno y ciclododecano — como protectores de pintura
soluble en agua durante la desalinización de la cerámica policromada. Se realizaron
experimentos de salinización, consolidación y desalinización en muestras nuevas preparadas en

CONTACT Hamada Sadek hsr00@fayoum.edu.eg Faculty of Archaeology, Conservation Department, Fayoum University, Al Fayoum, Egypt
© American Institute for Conservation of Historic and Artistic Works 2018
190 H. SADEK ET AL.

el laboratorio. Se utilizó la información obtenida de microscopia óptica, SEM-EDS, mapeo elemental


y FTIR para evaluar la eficacia relativa de los cuatro consolidantes. Además, se midió el índice
relativo de sublimación de los consolidantes desde una superficie vítrea. Aunque todos los
consolidantes protegen a las capas de pintura, hay diferencias en la eficacia y en el índice de
sublimación que hacen que unos sean mejores que otros, dependiendo de la naturaleza del
objeto que va a ser intervenido. Traducido por Mirasol Estrada; revisado por Amparo Rueda.

1. Introduction may indicate that crystallization is occurring below the


surface (i.e. sub- or crypto-efflorescence). This is a threat
Archaeological objects comprised of painted porous
to the stability and integrity of archaeological objects
materials often contain significant amounts of soluble
because it causes separation and spalling of outer layers
salts that may originate from production techniques or
and can result in damage to the entire structure (Bena-
as a result of environmental exposure over time (Zehn-
vente et al. 2004; Espinosa-Marzal and Scherer 2010;
der 1987, 2007). In that regard, the salts may be present
Scherer 2000).
within objects such as masonry, ceramics, and pottery
The fluctuation between crystallization, dissolution,
because the initial raw materials used in their construc-
and recrystallisation of soluble salts, such as halite, causes
tion contain salts or because soluble salts in the soil
deterioration to porous materials such as pottery,
have accumulated over time in buried objects (Charola
especially in uncontrolled environments. The deterio-
and Bläuer 2015). For example, the Tethys Sea, which
ration results from the changes in phase transformations
covered Egypt 40 million years ago, left salt deposits in
inside the porous materials. The formation of crystals
the sands that were used by many artisans to make cer-
leading to growth of salt crystals creates pressure on the
amic objects (Sampsell 2014).
pores walls. Even inside museums or storage areas for arti-
Salts also can be from indoor sources, including out-
facts, daily, seasonal, and annual variations of humidity
gassing materials used for housing and displaying the
and temperature can cause repeated cycles of crystalliza-
artifacts (Skipper and Rubinstein 2018). Several factors
tion and recrystallization. Thus, controlling indoor
affect the location of salt crystallization, including the
environmental conditions minimizes the damage caused
solubility of the salt, water evaporation rate, pore size
by crystallization cycles or during removal of salts from
and structure within the materials, the water content
the artifacts. Crystallization by efflorescence causes
and its transport rate, temperature, humidity and
deterioration phenomena by flaking, powdering of the
airflow (Charola 2000). The presence of salts in the por-
object components, and honeycomb weathering (Delgado
ous materials made from impure sand or other raw
et al. 2016). Furthermore, the damage caused by NaCl
materials can cause deterioration, especially when the
(halite) is irreversible because of salt dilation in pores
salts crystallize and cause cracks to form. Salts that crys-
(Desarnaud, Bertrand, and Shahidzadeh-Bonn 2011).
tallize on the surface (i.e. efflorescence) hide paint layers
Sodium chloride is one of the most common water-
or decorative patterns, as shown in Figure 1, and finally
soluble salts found in porous materials. It can cause
lead to paint spalling. The presence of salts on the surface
serious damage to materials because its high solubility
enables ions to move throughout an object as long as
moisture is present. Also, the hydrophilicity of materials
associated with pottery, stone objects, and even edifices
aids water absorption that can lead to cycles of re-crystal-
lization of halite and other minerals with which it can
interact. The re-crystallization of halite (and other sol-
uble salts) can lead to the development and propagation
of cracks, spalling of surface layers, and changes in the
appearance of a pottery substrate and painted surfaces.
The long-term preservation and reduction of poten-
tial damage of some objects from the presence of salts
requires that they be desalinated. The method used for
desalination depends on the components of an artifact
and its sensitivity to the desalination method, as well
as its initial condition. Various methods for desalination
Figure 1. Salt crystallization on a piece of Egyptian painted have been employed in cultural heritage preservation,
pottery in Cairo, Egypt, from the Greco-Roman period. including poultices, water baths, and electromigration
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION 191

(Skibsted et al. 2015). Poulticing involves the removal of consolidants, fixatives, masking materials, and barrier
soluble salts by application of damp materials on the arti- layers in different areas of conservation. The most attrac-
fact surfaces (Paterakis 1999). The effectiveness of the tive characteristic of these reversible consolidants is their
poultice materials depends on their adhesion to the sur- ability to be removed without additional treatment by
face of interest, shrinkage, and workability (Kamran et al. slow sublimation at room temperature (Anselmi et al.
2012). Traditional poultices are prepared using materials 2015).
such as bentonite, attapulgite, and paper pulp. They can In this way, they can stabilize an object during the
be effective also for removal of surface stains or paint treatment and be removed at a later date of the conserva-
(Sawdy et al. 2010). tor’s choosing; no chemical or additional physical treat-
Objects with painted surfaces present a special chal- ment steps are needed for their removal (Han, Huang,
lenge to be desalinated without causing significant ancil- and Zhang 2016; Rowe and Rozeik 2009; Rozeik 2009).
lary damage because paint films on pottery and other Thus, menthol and CDD are considered reversible bar-
porous substrates are very sensitive to water (Koob and riers and consolidants and they have been used for protec-
Ng 2000). However, the paint can be protected by conso- tion of fragile materials such as stone and ceramic objects,
lidation processes in which the cohesive strength paintings, manuscripts, and textiles. They can be applied
between the paint and the object is enhanced using as a melt, in solution, or as a spray (Keynan and Eyb-
organic or inorganic compounds (Bromblet et al. 2011; Green 2000; Stein et al. 2000). CDD has been employed
Pinto and Rodrigues 2008). In that regard, various as a temporary consolidant on marble, a porous substrate,
types of organic and inorganic compounds can be and its behavior and characteristics of sublimation and
employed. They include polymers and polymerizable penetration were investigated (Anselmi et al. 2011). Pre-
materials (such as polyacrylates, polyacetates, polysilox- liminary examination of the ability of CDD to protect
anes, ethyl silicate, and epoxy resins) and inorganic water-sensitive ink on a ceramic during desalination has
salts (such as calcium hydroxide, barium hydroxide, been undertaken (Muros and Hirx 2004).
sodium and potassium silicates, fluorosilicates, and alu- Removing or re-solubilizing traditional consolidants
minates) (Torraca 2009). Obviously, the effectiveness of requires either mechanical action or solvents, both of
a consolidant depends on the properties of the consoli- which can lead to secondary changes in artifacts. How-
dant and the surface of the object being treated. ever, the use of subliming materials comes with some
Light, chemical, and mechanical factors can also affect potential disadvantages that include a lack of data
the stability of a consolidant while it is applied to a sur- about the long-term effects of these materials on an
face. Therefore, the balance between beneficial mechan- object, concerns regarding their toxicity (Han et al.
ical effects and possible adverse chemical and optical 2014; Horie 2010), and, as noted previously, compli-
effects must be weighed in each case. For example, cations associated with different modes of application
some consolidants cannot be removed completely from for different compounds.
an object after treatment without irreversible changes Although cyclododecane and menthol had been used
to the surface being treated. To address the issue of previously in consolidation of artifacts at excavation sites
“reversibility” (i.e. optimization of the degree of remova- (Han 2014; Presciutti et al. 2015), our studies expand the
bility of the consolidant and minimization of the chemi- study of the use of sublimable compounds to other con-
cal and aesthetic changes that permanent consolidants solidants. In this work, we have used a number of
can cause), temporary consolidants (i.e. those that can analytical techniques to evaluate and compare the
be completely removed) have been used during desalina- degrees to which four sublimable compounds, applied
tion without mechanical or chemical methods. as melts, are able to protect water-sensitive painted sur-
The development of less invasive and less altering faces on pottery during its desalination (Chickos and
materials for the treatment of water-sensitive paint sur- Acree 2003; Shahidzadeh-Bonn et al. 2010b; Trefil et al.
faces are needed badly. Such materials and methods 2005). They are cyclododecane (CDD; C12H24), menthol
are emerging as a result of unrelated research in areas (C10H20O), cyclododecanone (C12H22O) which has simi-
of chemistry, nano-science, and materials. The challenge lar properties to CDD but sublimes more slowly, making
today is as much to learn how to apply these materials to it preferable when longer consolidation times are
conservation as it is to discover their specific interactions required, and camphene (C10H16) which has a consolida-
on different surfaces (Watters 2007). Examples of inter- tion function and sublimes quickly comparing to other
est are sublimable materials, such as menthol and cyclo- consolidants in this study. Their structures and some
dodecane (CDD). Sublimable compounds have been pertinent properties are included in Table 1. Safety and
used in the field of art and archaeological conservation health concerns about the consolidants in this study
since the 1990s. They have been applied as temporary are included in an appendix.
192 H. SADEK ET AL.

Table 1. Materials used in this study.


Cyclododecane (CDD), C12H24 Cyclododecanone C12H22O Menthol C10H20O Camphene C10H16
Structure

Melting point (°C) 58–61 59–69 43–45 40–50


Boiling point (°C) 243 277 212 159–160
Flash point (°C) 114 118 101 34
Purity (%) 98 99 99 80
Vapor pressure at 20°C 0.0098 0.00075 0.11 0.33
(kPa)
Source Kremer Pigments Acros Organics Aldrich Aldrich

2. Materials and methods of halite crystals and dissolution of the organic paint bin-
der as indicated by a loss of adhesion between the pig-
2.1. Materials
ment and the pottery. A separate group of samples, as
The pottery was made using Nile silt and sand added to a control, was not weathered.
clay and kneaded, then shaped into 2 × 2 × 0.8 cm blocks
before firing at 800°C. The paint was prepared by mixing
10 grams of green earth pigment, a fine-grained siliceous 2.1.3. Application of sublimable consolidants
earth (i.e. a complex of SiO2, Al2O3, Fe2O3, MgO and, Before desalination, one of the four sublimable com-
CaO3, Na2O, MnO, and K2O) and binder consisting of pounds was applied in melt form using a spatula to
10 mL of 10% (wt/vol) animal skin glue in water. Both form a layer ∼0.5 mm thick to the painted side of the
materials were obtained from the Conservation Labora- pottery samples. The melting temperatures of the conso-
tories, Faculty of Archaeology, Fayoum University. lidants are listed in Table 1. The melts cooled and
became solid very quickly after their application. The
entire experimental protocol was conducted in a fume
2.1.1. Preparation of test samples
hood because of the high vapor pressures of the tempor-
Test samples consisting of painted pottery that simulates
ary consolidants.
archaeological objects were prepared by painting the sur-
face of the pottery using a tempera technique in which
the paint was brushed onto the surface. The painted 2.1.4. Desalination procedure
samples were dried at room temperature for 30 days in To remove salt, the artificially weathered samples were
air. The thickness of the final paint layer, 200–300 µm, weighed and immersed in a 50 fold (wt:vol) excess of deio-
was measured using optical microscopy, observing the nized water. The sample weights were recorded in a for-
focal plane of the top of the pottery and the top of mat developed at the Arizona State Museum (White,
paint and scaling for the distance the stage moved for a Pool, and Carroll 2010). The time-dependent changes in
given objective. Also, glass slides were coated with the conductivity of the water were measured with an
equal weights of melts of the sublimable compounds Amber model 604 conductivity meter periodically over
and then allowed to cool to room temperature in order three days or until a time-invariant value was reached.
to track their rates of sublimation. The data were treated according to Equation (1). Knorm,
the normalized conductivity change per day, is the desali-
2.1.2. Artificial salt weathering nation rate; it is usually reported without units. ΔK is the
The painted test samples were subjected to artificial salt change in conductivity per unit of time in units of micro-
weathering using 10 wet/dry cycles. In each cycle, the Siemens (µS/cm), L is the water volume in liters, t is the
samples were submerged in 250 mL of an aqueous sol- time in days, and g is the sample weight in grams.
ution of 20% (wt/vol) NaCl for 16 hours, and then D KL
removed and kept in air for 8 hours at ∼ 25°C. During Knorm (1)
D tg
the weathering cycles, the samples suffered the growth
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION 193

2.2. Methods detectable level of residue was present, spectra were


obtained from six positions on each sample.
2.2.1. Optical microscopy
Optical microscopy, using a Leica DMRX microscope,
with 10x and 20x nPL objectives in bright and dark 3. Results
field reflected light illumination, was used to evaluate
As indicated by the results provided below, the salt
visually the condition of the salt-weathered pottery and
weathering cycles damaged the paint layer as well as
to compare the morphology of painted samples before
the pottery substrate of the consolidated and unconsoli-
and after applying the sublimable consolidants. Images
dated samples.
were captured using a Canon EOS 5D Mark III camera.
Because the samples were not flat, the collected images
were stack focused using ImageJ software (Schneider, 3.1. Weathered samples before desalination
Rasband, and Eliceiri 2012).
3.1.1. Optical microscopy
Before being subjected to the salination procedure, the
2.2.2. Scanning electron microscopy (SEM) film of paint on the surface of the pottery was homo-
SEM was performed using a Hitachi S-3400N variable geneous and continuous (Figure 2(a)). After salination,
pressure electron microscope (20 kV, 50 Pa) equipped damage from formation of halite crystals on the surface
with a Hitachi backscattered electron detector. SEM- of the pottery and the paint layers is visible (as manifested
EDS (SEM-energy-dispersive X-ray spectroscopy) was by cracking of the pottery and disruption of the film of the
very useful in determining the elements and their distri- paint layer). Cubic crystals of halite formed between paint
bution in the painted layer compositions. The SEM-EDS grains as the samples were dried (Al-Jibbouri and Ulrich
analyses were performed at 20 kV to obtain element 2002). The concentration and grain size of the crystals
maps and qualitative point analyses using an Oxford are known to influence the microstructure and degree of
X-Max 80 mm2 SDD detector and an Oxford Aztec damage (Desarnaud, Bertrand, and Shahidzadeh-Bonn
300 spectrometer. The working distance was 10 mm. 2013). A significant portion of the green paint was lost
as a result of salt crystal formation, leaving the matrix of
2.2.3. Attenuated total reflectance-Fourier fired pottery visible at the surface (Figure 2(b)).
transformed infrared spectroscopy (ATR-FTIR)
ATR-FTIR measurements were carried out on a Perkin 3.1.2. SEM-EDS
Elmer 100 spectrometer in the ATR mode in the wave- Back-scattered electron (BSE) and secondary electron (SE)
length range of 400–4000 cm−1 with a spectral resolution imaging in the SEM showed that cracks had developed in
of 8 cm−1, using a diamond cell sample holder. Thirty- the pottery from the crystallization of sodium chloride in
two scans were averaged and the spectra were normal- the ceramic body during the drying step in the salinization
ized with respect to the background of the clean diamond procedure. Formation of crystals in porous materials cre-
cell for each analysis. In order to determine whether a ates pressure on the pore walls, and stress leads to crack

Figure 2. Stack–focused photomicrograph (using the plug-in of ImageJ) of the bright-field image of the painted surfaces of a pottery
sample: (a) before salination, showing the intact paint layer; (b) after salination in the absence of prior treatment with consolidant
where the green paint has been displaced by the formation of cubic halite crystals, revealing the orange-red color of the fired pottery.
194 H. SADEK ET AL.

in section 2.1.4. The progress of the desalination was


monitored over three days and the samples were
removed when no more salt loss from the pottery
could be measured using Knorm (Unruh 2001).
The rate of desalination in water is dependent on the
NaCl concentration gradient within the pottery and
advection, which is influenced by pottery pore sizes. In
general, salts move from coarse pores into smaller
pores (Voronina et al. 2013; Voronina, Pel, and Kopinga
2014). Due to the presence of the very hydrophobic con-
solidants, the salts move primarily from the pottery into
the water through the surfaces that are not covered with
sublimable materials. The movement of the ions depends
on a number of factors, including the porosity of the pot-
Figure 3. BSE image of microcracks in the pottery resulting from tery, P = 26–27%, which can be calculated as
halite crystal formation after salination. W −D
P= × 100 %, where, D is the weight of dry
W −S
formation (Shahidzadeh-Bonn et al. 2010a). Usually, such sample, W is the weight of the sample when it is satu-
micro-cracks are invisible initially; they become observa- rated with deionized water (i.e. after being removed
ble when the salt crystals grow and cause networks of from the water and wiped lightly with a tissue), and S
micro-cracks around the crystals (Figure 3). In that is the sample weight when it is saturated by and sus-
regard, the SEM micrographs show the presence of pended in deionized water (Lubelli and De Rooij 2009).
sodium chloride in both cubic and needle-like whisker The extraction of salt is rapid over the short initial time
form, at and near the surface of the pottery (Figure 4). interval because the salt crystals on and near the surface of
The whisker-like crystals are known to grow one-dimen- the pottery dissolve almost immediately. Subsequently,
sionally on substrates in low humidity conditions or from salt inside the pores of the pottery dissolves more slowly
thin films (Lopez-Acevedo et al. 1997). A correlation and is transported to the surface of the sample. The
between elements in the maps shown in Figure 5 for the change in the conductivity of water in which the pottery
whisker needles shows that sodium and chlorine are dis- was immersed shows a rapid initial increase followed by
tributed in the same way throughout them. a slower continuous one (Figure 6). This behavior is con-
sistent with the model advanced for salt removal. The
rates of desalination (i.e. loss of halite), using each sublim-
3.2. Desalination process able compound as a covering layer, follow a similar quali-
tative temporal course for desalination. Note also that the
Pottery samples were treated with one of the four conso- rates of salt flow in samples with menthol and cyclodode-
lidants before desalination using the procedure described canone layers are higher than those with camphene and
cyclododecane layers. Additional research will be needed
to determine whether specific interactions between the
consolidant and the pottery surface are able to influence
the rate of flow of salt crystals from the pottery matrix.
Following the desalination process, the pottery
samples were washed with distilled water in order to
remove any remaining salts that may have remained
on the surfaces. Visually and with additional magnifi-
cation by optical microscopy at magnification 10x and
20x, it is clear that the rate of sublimation depends on
the consolidant used. At one atmosphere of pressure
and at room temperature, camphene appears to have
sublimed totally after 4–6 hours, whereas cyclododecane,
cyclododecanone and menthol require much longer
periods (∼20 days) before visual inspection suggests
Figure 4. SE image showing the presence of needle-shaped
whisker crystals on the surface of a painted pottery sample they are completely removed. We assume that little or
after salination. no consolidant was lost during the desalinization
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION 195

Figure 5. EDS element maps of an area of the pottery surface that includes whisker crystals formed after accelerated salination. The two
maps focus on halite whisker forms: (a) sodium and (b) chlorine.

procedures because the samples were completely sub- thicknesses were nearly the same, it is possible to compare
merged in water and the consolidants are insoluble in it. the rates of loss of the 4 consolidants. Camphene had the
highest rate of sublimation; it was lost completely within
24 hours. Menthol and cyclododecane required ∼237
3.3. The sublimation rates of consolidants on and ∼261 hours to sublime completely. The rate of subli-
glass slides mation of cyclododecanone was the slowest; it required
Sublimation rate here is the speed at which a consolidant ∼1300 hours to be removed completely. As expected,
is lost from a surface in contact with the air at room temp- this trend is qualitatively consistent with the vapor press-
erature. It is an important factor in the selection of a con- ures reported in Table 1.
solidant for a particular application. We monitored the
weight changes of the consolidants on glass slides as a
3.4. Analysis after desalination
function of time to calculate these rates. Measurements
were continued until the weights returned to those of The samples were analyzed after desalination using OM,
the uncoated glass slides. Based on the data in Figure 7, FTIR, and SEM-EDS to assess the efficacy of the four
and consistent with the fact that the areas of the films in materials for consolidating and protecting the paint
contact with the air remained constant throughout the film, and to determine whether residue could be detected
sublimation, the rates remain constant. Because the film on the pottery surfaces.

Figure 6. Plot of Knorm (conductivity change) versus time, showing the loss of salt from the pottery samples that were coated with the
different consolidants.
196 H. SADEK ET AL.

Figure 7. Degree of sublimation of the temporary consolidants on a glass slide versus time.

3.4.1. Visual investigations analyses performed, we conclude qualitatively that the


As noted, the accelerated saline weathering caused the consolidants did afford some protection to the paint
painted layers to be dissolved and damage to the pottery layers. However, although camphene protected the pig-
(Figure 7(B)). The post-desalination stage is most impor- ments on the surface, some of the paint did suffer damage
tant in the evaluation of the protection afforded by the in the water bath (Figure 8(D1)). Menthol, cyclododecane,
consolidants to the paint layer in water. Based on the and cyclododecanone appeared to afford better protection

Figure 8. Pottery samples (A) before salination, (B) after salt weathering, (C) after salt weathering followed by consolidation, and (D)
after desalination and sublimation of the consolidant. Samples that were not exposed to salt and dried (i.e. controls) are shown in
column 6. In (C), pottery tiles in column 1 were consolidated with camphene, in column 2 with menthol, in 3 with cyclododecane,
and in 4 with cyclododecanone. Note that the unconsolidated samples (shown in column 5) experienced almost complete loss of
their paint layer during desalination.
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION 197

(Figure 8(D2–4)). Regardless, the samples not coated with This creates a color change that can be detected when
a consolidant lost most of their paint layers (Figure 8 comparing samples before and after treatment.
(D5)). No residue from any of the consolidants was visible
by optical microscopy on the painted surfaces after the 3.4.2. FTIR
desalination and sublimation procedures. Due to the sali- Evaluation of the degree of final sublimation of the con-
nation and desalination steps, a small portion of the paint solidants was assessed using ATR-FTIR spectra. The
might have been lost, and the paint surface roughened. spectra of the samples treated with camphene show no

Figure 9. ATR-FTIR spectra of the surfaces of accelerated-weathered painted pottery after consolidation and desalination, indicating the
absence of camphene, menthol, and cyclododecane residues and the presence of some residue from cyclododecanone.
198 H. SADEK ET AL.

Figure 10. EDS maps of the painted side of a pottery sample after completion of the desalinization and sublimation procedures: (a) Cl;
(b) Ca. The arrows point to sites where concentrations of these elements are highest.

evidence for bands characteristic of camphene at The organic binder used in the paints for the samples
3066 cm−1 (for = C–H), and saturated C–H stretches is animal glue that consists of polymerized amino acids
and bending modes. FTIR spectra of samples treated through amide linkages. FTIR bands from amide bonds
with menthol did not show bands at 3240, 2855–2924, are found at three regions: amide I at 1600–1700 cm−1
1448, or 1025 cm−1 that are characteristic of O–H, includes C = O stretching, C–N and N-H bending;
CH3, −C–H, and C–O groups, respectively. Similarly, amide II at 1500–1600 cm−1 from C–N stretching
the FTIR spectra of samples treated with cyclododecane with N-H bending; and amide III at 1200–1350 cm−1
did not show bands at 2955–2870 cm−1 that are charac- from N-H bending, C–N stretching, and specific C–H
teristic of saturated C–H bond stretches. Thus, we con- modes (Singh et al. 1993). The decrease in intensity
clude that the levels of residues with these consolidants of the amide I, II, and III bands indicate that the
are very low. However, bands at 2850–2950 cm−1, due weathering process used for the pottery samples did
to aliphatic C–H stretching modes, were detected on cause the loss of some of the paint, probably due to
the surface of pottery consolidated with cyclododeca- the dissolution of the organic paint binder in the halite
none even after 10 days of sublimation (Figure 8). solution.

Figure 11. SEM-EDS: (a) micrograph of SEM on painted pottery treated with cyclododecanone; (b) spectrum of a quartz grain in (a)
showing the presence of carbon; (c) element map of carbon by SEM-EDS showing the presence of carbon, presumably due to the pres-
ence of cyclododecanone residue.
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION 199

3.4.3. SEM-EDS analyses indicate the absence of discernible camphene,


The success of the desalination procedures and protection menthol, or cyclododecane on the surface of the treated
of the sensitive paints from dissolution and spalling are samples after leaving them for several days at room
shown clearly by the absence of sodium in the SEM- temperature in the air (i.e. conditions under which the
EDS maps in Figure 9. However, the EDS elemental consolidants sublime); however, traces of cyclododeca-
maps (Figure 10) do indicate the presence of a new cal- none did appear on the surface for a very long period.
cium species that correlates with the position of chloride. In conclusion, the four sublimable consolidants pro-
We suspect that CaCl2 is produced in the pottery samples; tected the paint layer. Cyclododecane, cyclododecanone
the clay contains some calcium. Although CaCl2 is not a and menthol afforded similar degrees of protection to
hygroscopic and water-soluble salt (Winkler 2013), pre- the paint layer while camphene was less efficient, perhaps
caution should be taken to limit any alteration or chemical due to its more rapid rate of sublimation. Additional
reaction by it during conservation treatment. studies and research on mixtures of cylododecanone
Although elemental mapping of samples treated with and camphene will be necessary to optimize the
cyclododecanone is consistent with the disappearance of efficiency of their temporary consolidation of pottery
the consolidant from most of the surface (as indicated by and painted surfaces during desalination.
the absence of carbon in the sum spectrum (Figure 11)), Future studies will include additional tests of these
some carbon is located on the surface of the quartz par- and other sublimable consolidants to protect different
ticles in the pottery as residue of the cyclododecanone on types of painted surfaces on different types of materials.
the pottery samples. Those studies will include additional analytical investi-
gations of the properties of the object surfaces before
and after treatment, and measurements of sublimation
4. Conclusions
rates under different atmospheric conditions. Further
The ability of four commercially available, sublimable work includes measurements to determine the pen-
and relatively unreactive molecules, which are exten- etration depth of the consolidants into samples and cul-
sively available commercially and inexpensive (i.e. cyclo- tural heritage objects.
dodecane, menthol, cyclododecanone and camphene), to
protect paint on pottery surfaces during their aqueous Acknowledgments
desalination has been demonstrated and compared. We
HS and RGW thank the US National Science Foundation
have performed dynamic analyses of their behavior on (Grant CHE-1502856) for its support of the portion of this
experimental test samples and have compared the research conducted at Georgetown University. HS is grateful
efficiency of their consolidation during desalination. to the Fulbright Scholars Program, administered by the Coun-
The results demonstrate that cyclododecanone and cam- cil of the International Exchange of Scholars, for a fellowship.
phene afford good protection to a water-sensitive paint
on pottery during desalination. Although menthol,
Disclosure statement
cyclododecane, and cyclododecanone protected the
painted samples better than camphene, all afforded No potential conflict of interest was reported by the authors.
some degree of protection compared to not using a
consolidant.
Funding
Removal of salt from the samples occurs in two stages:
the more rapid one is the dissolution of salts at or near This work was supported by National Science Foundation
[grant number CHE-1502856]; Binational Fulbright Commis-
the surface; the slower step is removal of salts located in
sion in Egypt (AY 2015/2016).
interior pores. Some of the results indicate that the molecu-
lar structure and bulk properties of the consolidants influ-
ence the desalination rate despite there being little, if any, Notes on contributors
sublimation while the samples are submerged in water to Hamada Sadek holds a M.Sc. and Ph.D. in Archaeological Con-
remove the salt; visually, the consolidants remained on servation from Cairo University, Egypt. He is a staff member of
the surfaces throughout the desalination process. Also, the Conservation Department, Faculty of Archaeology, Fayoum
the desalination rates of samples treated with menthol or University, Egypt. He was a Fulbright postdoctoral fellow at
cyclododecanone were found to be higher than those trea- Georgetown University and at the National Gallery of Art,
Washington, DC. Sadek was Samuel H. Kress Fellow for
ted with camphene or cyclododecane.
Advanced Studies in Art Conservation 2010-2011, Arizona
Residues left as a result of the application of a protec- State Museum – the University of Arizona. He has been fellow
tive treatment on a work of art are highly undesirable at universities and institutes in France, Germany, Poland, Roma-
and an important concern. In that regard, FTIR spectral nia and USA. He is a Professional Associate and Sustainability
200 H. SADEK ET AL.

Committee member of the American Institute for Conservation Anselmi, C., F. Presciutti, B. Doherty, B. G. Brunetti, A.
of Historic and Artistic Works. Since 2016 he has been a mem- Sgamellotti, and C. Miliani. 2015. “A Non-invasive
ber of the conservation team of the University of Michigan Investigation of Cyclododecane Kinetics in Porous
Abydos Middle Cemetery (AMC) project. Mailing address: Matrices by Near-infrared Spectroscopy and NMR In-
Conservation Department, Faculty of Archaeology, Fayoum depth Profilometry.” Journal of Cultural Heritage 16 (2):
University, Po Box 63511, Al Fayoum, Egypt. Email: 151–158.
hsr00@fayoum.edu.eg. Benavente, D., M. G. Del Cura, J. Garcıa-Guinea, S. Sánchez-
Barbara H. Berrie holds a doctorate in chemistry from George- Moral, and S. Ordóñez. 2004. “Role of Pore Structure in
town University in Washington, DC. Before joining the National Salt Crystallisation in Unsaturated Porous Stone.” Journal
Gallery of Art in Washington, DC, she was an NRC postdoctoral of Crystal Growth 260 (3–4): 532–544.
fellow at the US Naval Research Laboratory. Berrie has published Bernson, V. S. M., and B. Pettersson. 1983. “The Toxicity of
widely in conservation science on topics as diverse as new Menthol in Short-term Bioassays.” Chemico-biological
materials for treatment of works, analytical methods for pigment Interactions 46 (2): 233–246.
identication, and history of artists’ use of materials. She is editor Bromblet, P. H., V. Vergès-Belmin., C. Franzen, S. Aze, and
of volume 4 of the series Artists’ Pigments: A Handbook of Their O. Rolland 2011. “Toward an Optimization of the
History and Characteristics. Berrie was awarded a Samuel Specifications for Water Bath Desalination of Stone
H. Kress Paired Fellowship (with Louisa C. Matthew) at the Cen- Objects.” Salt Weathering on Uildings and Tone Culpture,
ter for Advanced Study on the Visual Arts (2002) and in 2018 Proceedings from the International Conference, 397–404,
was awarded the John D. and Susan P. Diekman Fellowship at Limassol, Cyprus, October 19–22.
the Djerassi Resident Artists Program. She is head of the Scien- Charola, A. E. 2000. “Salts in the Deterioration of Porous
tific Research Department at the National Gallery of Art. Mailing Materials: An Overview.” Journal of the American Institute
address: 2000B South Club Drive, Landover, MD, 20785, USA. for Conservation 39 (3): 327–343.
Email: b-berrie@nga.gov. Charola, A. E., and C. Bläuer. 2015. “Salts in Masonry: An
Overview of the Problem.” Restoration of Buildings and
Richard G. Weiss received an ScB degree from Brown Univer- Monuments 21 (4–6): 119–135.
sity and MS and PhD degrees from the University of Connecti- Chickos, J. S., and W. E. Acree. 2003. “Enthalpies of
cut. He was an NIH Postdoctoral Fellow at California Institute Vaporization of Organic and Organometallic Compounds,
of Technology and a Visiting Assistant Professor and National 1880–2002.” Journal of Physical and Chemical Reference
Academy of Sciences Overseas Fellow at the Instituto de Qui- Data 32 (2): 519–878.
mica of the Universidade de São Paulo in Brazil before coming Delgado, J., A. Guimarães, V. De Freitas, I. Antepara, V. Kočí,
to Georgetown University in 1974, where he is a member of the and R. Černý. 2016. “Salt Damage and Rising Damp
Chemistry Department and the Institute for Soft Matter Syn- Treatment in Building Structures.” Advances in Materials
thesis and Metrology at Georgetown. He has been a visiting Science and Engineering 2016: 1–13.
professor at universities and institutes in Brazil, China, Costa Desarnaud, J., F. Bertrand, and N. Shahidzadeh-Bonn. 2011.
Rica, Germany, France, Italy, Japan, Slovakia, and Spain. He “Dynamics of Salt Crystallization.” SWBSS Proceedings,
is a Fellow of IUPAC, a member of the Brazilian Academy Limassol, Cyprus, 23–29.
of Sciences, and received a doctorate honoris causa from Uni- Desarnaud, J., F. Bertrand, and N. Shahidzadeh-Bonn. 2013.
versité de Bordeaux 1. He was a senior editor of the ACS jour- “Impact of the Kinetics of Salt Crystallization on Stone
nal Langmuir for 10 years, is currently a member of the Damage During Rewetting/Drying and Humidity
editorial advisory boards of the Journal of the Brazilian Chemi- Cycling.” Journal of Applied Mechanics-Transactions of the
cal Society and Gels, and is an Associate Editor for Polimeros ASME 80: 020911-1.
and a member of the Scientific Committee of Substantia. His Espinosa-Marzal, R. M., and G. W. Scherer. 2010. “Advances
research interests include investigations of photochemical, in Understanding Damage by Salt Crystallization.”
photophysical and thermal reactions of molecules in anisotro- Accounts of Chemical Research 43 (6): 897–905.
pic environments, and the development and application of Han, X. 2014. “The Use of Menthol on the Archaeological Site
ionic liquids, ionic liquid crystals, and new molecular and at the Mausoleum of the First Qin Emperor.” Studies in
polymer gels for various purposes, including chemical spill Conservation 59 (1): S223–SS24.
remediation and the conservation of objects of cultural heri- Han, X., X. Huang, and B. Zhang. 2016. “Morphological
tage. Mailing address: 240 Reiss Science Building, Georgetown Studies of Menthol as a Temporary Consolidant for
University, Washington, DC 20057-1227, USA. Email: Urgent Conservation in Archaeological Field.” Journal of
weissr@georgetown.edu. Cultural Heritage 18: 271–278.
Han, X., B. Rong, X. Huang, T. Zhou, H. Luo, and C. Wang.
2014. “The Use of Menthol as Temporary Consolidant in
References
the Excavation of Qin Shihuang’s Terracotta Army.”
Al-Jibbouri, S., and J. Ulrich. 2002. “The Growth and Archaeometry 56 (6): 1041–1053.
Dissolution of Sodium Chloride in a Fluidized Bed Horie, C.V. 2010. Materials for Conservation: Organic
Crystallizer.” Journal of Crystal Growth 234 (1): 237–246. Consolidants, Adhesives and Coatings. 2nd ed.
Anselmi, C., F. Presciutti, B. Doherty, B. Brunetti, A. Sgamellotti, Amsterdam; Boston: Butterworth-Heinemann.
and C. Miliani. 2011. “The Study of Cyclododecane as a Kamran, K., L. Pel, A. Sawdy, H. Huinink, and K. Kopinga.
Temporary Coating for Marble by NMR Profilometry and 2012. “Desalination of Porous Building Materials by
FTIR Reflectance Spectroscopies.” Applied Physics A: Electrokinetics: An NMR Study.” Materials and Structures
Materials Science & Processing 104 (1): 401–406. 45 (1): 297–308.
JOURNAL OF THE AMERICAN INSTITUTE FOR CONSERVATION 201

Keynan, D., and S. Eyb-Green. 2000. “Cyclododecane and Singh, B. R., D. B. Deoliveira, F.-N. Fu, and M. P. Fuller. 1993.
Modern Paper: A Note on Ongoing Research.” WAAC “Fourier Transform Infrared Analysis of Amide iii Bands of
Newsletter 22 (3): 18–21. Proteins for the Secondary Structure Estimation.” Proc.
Koob, S. P., and Won Yee Ng. 2000. “The Desalination of SPIE 1890, Biomolecular Spectroscopy III, 47–55.
Ceramics Using a Semi-automated Continuous Washing Skibsted, G., L. M. Ottosen, P. E. Jensen, and J. M. Paz-Garcia.
Station.” Studies in Conservation 45 (4): 265–273. 2015. “Electrochemical Desalination of Bricks - Experimental
Lopez-Acevedo, V., C. Viedma, V. Gonzalez, and A. La Iglesia. and Modeling.” Electrochimica Acta 181: 24–30.
1997. “Salt Crystallization in Porous Construction Materials. Skipper, L., and N. Rubinstein. 2018. “Detecting Chloride
Ii. Mass Transport and Crystallization Processes.” Journal of Contamination of Objects and Buildings–Evaluating a
Crystal Growth 182 (1–2): 103–110. New Testing Process.” Journal of Conservation and
Lubelli, B., and M. R. De Rooij. 2009. “NaCl Crystallization in Museum Studies 16 (1): 1–9.
Restoration Plasters.” Construction and Building Materials Stein, R., J. Kimmel, M. Marincola, and F. Klemm. 2000.
23 (5): 1736–1742. “Observations on Cyclododecane as a Temporary
Muros, V., and J. Hirx. 2004. “The Use of Cyclododecane as a Consolidant for Stone.” Journal of the American Institute
Temporary Barrier for Water-sensitive Ink on for Conservation 39 (3): 355–369.
Archaeological Ceramics During Desalination.” Journal of Torraca, G. 2009. Lectures on Materials Science for
the American Institute for Conservation 43 (1): 75–89. Architectural Conservation. Los Angeles: The Getty
Nichols, K., and R. Mustalish. 2002. “Cyclododecane in Paper Conservation Institute.
Conservation Discussion.” The Book and Paper Group Trefil, J. S., R. M. Hazen, A. J. Gaudin, and G. A. Steehler. 2005. The
Annual 21 ( ): 81–84. Sciences: An Integrated Approach. 6th ed. New York: Wiley.
Paterakis, A. B. 1999. “Those Evasive Salt Crystals.” In Unruh, J. 2001. “A Revised Endpoint for Ceramics
The 12th Triennial Meeting, ICOM Committee for Desalination at the Archaeological Site of Gordion,
Conservation, edited by J. Bridgland, 799–802. Lyon: Turkey.” Studies in Conservation 46 (2): 81–92.
James & James (Science Publishers Ltd). Voronina, V., L. Pel, and K. Kopinga. 2014. “Effect of Osmotic
Pinto, A. P. F., and J. D. Rodrigues. 2008. “Stone Pressure on Salt Extraction by a Poultice.” Construction and
Consolidation: The Role of Treatment Procedures.” Building Materials 53: 432–438.
Journal of Cultural Heritage 9 (1): 38–53. Voronina, V., L. Pel, A. Sawdy, and K. Kopinga. 2013. “The
Pool, M. 2006. “Some Chemical Things Considered: Influence of Osmotic Pressure on Poulticing Treatments
Cyclododecane.” American Institute for Conservation News for Cultural Heritage Objects.” Materials and Structures
31 (1): 14–15. 46 (1–2): 221–231.
Presciutti, F., B. Doherty, C. Anselmi, B. Brunetti, A. Watters, C. 2007. “Cyclododecane: A Closer Look at Practical
Sgamellotti, and C. Miliani. 2015. “A Non-invasive NMR Issues.” Anatolian Archaeological Studies 16: 195–204.
Relaxometric Characterization of the Cyclododecane– White, C., M. Pool, and N. Carroll. 2010. “A Revised Method to
Solvent System Inside Porous Substrates.” Magnetic Calculate Desalination Rates and Improve Data Resolution.”
Resonance in Chemistry 53 (1): 27–33. Journal of the American Institute for Conservation 49 (1): 45–52.
Rowe, S., and C. Rozeik. 2009. “The Uses of Cyclododecane in Winkler, E. 2013. Stone in Architecture: Properties, Durability.
Conservation.” Studies in Conservation 54 (2): 17–31. Berlin: Springer.
Rozeik, C. 2009. “The Treatment of an Unbaked Mud Statue Zehnder, K. 1987. “Monitoring Wall Paintings Affected by
from Ancient Egypt.” Journal of the American Institute for Soluble Salts, in the Conservation of Wall Paintings.”
Conservation 48 (1): 69–81. Proceedings of a symposium organized by the Courtauld
Sampsell, B. M. 2014. The Geology of Egypt: A Traveler’s Institute of Art and the Getty Conservation Institute, 103–
Handbook. Cairo: American University Press in Cairo. 136.
Sawdy, A., B. Lubelli, V. Voronina, and L. Pel. 2010. “Optimizing Zehnder, K. 2007. “Long-term Monitoring of Wall Paintings
the Extraction of Soluble Salts from Porous Materials by Affected by Soluble Salts.” Environmental Geology 52 (2):
Poultices.” Studies in Conservation 55 (1): 26–40. 353–367.
Scherer, G. W. 2000. “Stress from Crystallization of Salt in
Pores.” Proceedings of the 9th International Congress on
Deterioration and Conservation of Stone, 187–94.
Appendix
Schiffer, Thomas, and Georg Oenbrink. 2000.
“Cyclododecanol, Cyclododecanone, and Laurolactam.” Health and safety issues concerning the
Ullmann’s Encyclopedia of Industrial Chemistry. doi:10. sublimable consolidants
1002/14356007.a08_201.
Schneider, C. A., W. S. Rasband, and K. W. Eliceiri. 2012. Health and safety information about the use of the sublimable
“NIH Image to Imagej: 25 Years of Image Analysis.” consolidants is very limited. For that reason, storage containers
Nature Methods 9 (7): 671–675. should be well sealed and marked with appropriate warnings.
Shahidzadeh-Bonn, N., J. Desarnaud, F. Bertrand, X. Chateau, Until adequate studies on the effects of the consolidants have
and D. Bonn. 2010a. “Damage in Porous Media Due to Salt been performed on humans, conservators should avoid inhala-
Crystallization.” Physical Review E 81 (6): 066110. tion and skin contact. The section below offers information
Shahidzadeh-Bonn, N., J. Desarnaud, F. Bertrand, X. Chateau found in the safety data sheets for the materials.
and D. Bonn. 2010b. “Effect of the Kinetics of Cyclododecane: low level of toxicity; direct contact can
Crystallization on Salt Weathering Of Stones.” AIP cause eye and skin irritation (i.e. tearing discomfort and vision
Conference Proceeding, 1254 vols (1), 121–125, May 30. obscuring), and may cause long lasting harmful effects to
202 H. SADEK ET AL.

aquatic life (Nichols and Mustalish 2002). There are no Precautions for applying sublimable compounds: Based on
available data about ventilation system and air sampling tests the information above, conservators should avoid direct con-
(Pool 2006). tact with and breathing vapors of the compounds by applying
Cyclododecanone: no known effect on the eyes and mucous them in a fume hood or under an exhaust ventilator. In
membranes, but it may be slightly irritating. The oral LD50 is addition, the usual precautions with the use of any chemical
2.5 g/kg (Schiffer and Oenbrink 2000). It may cause long-last- should be followed: wearing safety glasses, chemical resistant
ing harmful effects to aquatic life. gloves, and lab coats.
Menthol: exposure may cause irritation to skin, eyes, mucous The authors recommend these procedures, noting that
membranes, and upper respiratory tract. It may cause lesions to most of them are followed normally by conservators when
biological membranes (Bernson and Pettersson 1983). they are applying other materials. The water washes contain
Camphene: inhalation causes irritation of the nose and virtually none of the consolidants (because they are not soluble
throat and direct contact with eyes or skin causes irritation. in aqueous media) and small amounts of salts; as such, they are
It is toxic to aquatic life. not considered to impose an environmental hazard.

You might also like