You are on page 1of 8

Journal of Cultural Heritage 10 (2009) 501508

Case Study
Noninvasive physicochemical characterization of two 19th
century English ferrotypes
Emiliano Carretti , Marco Milano , Luigi Dei

, Piero Baglioni
Department of Chemistry and CSGI Consortium, University of Florence, via della Lastruccia, 3, 50019 Sesto Fiorentino (Firenze), Italy
Received 8 February 2008; accepted 25 February 2009
Abstract
The present work was one of the rst attempts to analyze the conservation status of two ferrotypes, ancient photographic plates realized on
a support made of iron. The photographic material was constituted of collodion as binder for the photosensitive silver halides grains. The two
ferrotypes studied belonged to a private collection of a family from Durham, UK, and were made at the end of the 19th century. The analytical
techniques used for the morphological and physicochemical characterization were noninvasive. The surface morphology was studied by means
of optical microscopy (OM) and environmental scanning electron microscopy (ESEM) coupled with an energy dispersive X-rays (EDX) system
for the elemental analysis. These techniques, together with microreectance Fourier transform infrared spectroscopy (-FTIR) and contact angle,
allowed to obtain information on both the chemical elemental composition of the materials constituting the ferrotypes, and the conservation
status of these photographic plates. The study showed that the physicochemical diagnostics allowed to characterize the two ferrotypes that, despite
their similar age and provenance, showed different conservation status, surface properties, and elemental composition.
2009 Elsevier Masson SAS. All rights reserved.
Keywords: Ferrotypes; Collodion; Photographs; Physicochemical diagnostics
1. Introduction
The photographic process called ferrotype or also tintype
was patented in the United States on February 19th, 1856 by
Hamilton Smith, professor at Kenyon College, in Ohio [1] and
became soon one of the most diffused processes until the intro-
duction of the gelatin based systems and the invention by Kodak
of the commercial photocameras [2]. In particular, with respect
to the ambrotypes, in the ferrotypes the support glass plate was
replaced by an iron plate coated with a surface layer of wet col-
lodion. As in the case of ambrotypes, in all the ferrotypes the
photosensitive grains are formed on the upper layer of the col-
lodion [3,4]. In these photographic techniques, the rst step is
the preparation of the support (metal or glass) which dimensions
should be decided as a function of the cameras size, taking into
account that the maximum possible size of the nal image is
exactly the one of the plate. Then, in order to produce a big
image (i.e. 13 18 cm), a camera of compatible dimensions

Corresponding author. Fax: +39 055 4573036.


E-mail address: dei@csgi.uni.it (L. Dei).
is needed. Once dened the size of the picture, the collodion
layer is deposed directly onto the support surface as a concen-
trated solution of nitrocellulose in alcohol plus ether. Then, the
obtained plate, not completely dried, is immersed into a solution
of AgNO
3
in order to precipitate the silver halide AgX from the
salts (usually binary mixtures of alkaline and/or earth-alkaline
bromides or iodides) which are homogeneously dispersed into
the wet collodion layer. The precipitation of the photosensitive
grains occurs within the upper layer of the plate. While still wet,
the ferrotype is placed in the camera and an exposure is made
(usually the exposition time varies from 1 to 30 seconds). This
is the reason why this process is known as the wet collodion
method, the same that is used for the production of ambrotypes
[5]. During the exposition, the interaction between the light and
the silver halide grains induces a photoreduction of the AgX,
giving the formation of clusters like Ag
+
n
and Ag
0
n
(usually n >
10) containing free metal silver atoms. The reaction mechanism
is described below:
AgX(crystal) + h(radiation) Ag
+
+X + e

(1)
Ag
+
+e

Ag
0
(2)
1296-2074/$ see front matter 2009 Elsevier Masson SAS. All rights reserved.
doi:10.1016/j.culher.2009.02.002
502 E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508
The elemental silver atoms produced constitute the latent
image that is not visible yet; nevertheless, the latent image
contains all the information for the formation of the nal pic-
ture. To reveal the latent image, the exposed plate is placed into a
solution of gallic acid +silver nitrate that produces rounded par-
ticles on the upper collodion layer (physical development). In
the following step, all the not reacted AgX crystals are removed
by means of a solution of KCN or Na
2
S
2
O
3
that solubilize AgX
thanks to the formation of the complexes ([Ag(S
2
O
3
)
2
]
3
and
[Ag(CN)
4
]
3
) [6] easily washed by water. This procedure allows
the complete elimination of all the residues of the photosensitive
material that can continue to react if exposed again to the light
[7]. Finally, a coating layer is applied onto the nal surface for
protective purposes. It is worthwhile to recall that the ferrotype
procedure generates directly the positive, instead of the negative
as in the case of traditional photography. In fact, according to this
technique one exploits the contrast betweenthe iron(plate) back-
ground that appears dark due to the transparency of the collodion
not impressed by the light, and more or less grey regions of the
collodion impressed by light and not longer transparent. Being
the ferrotypes simpler and faster to prepare than the ambrotypes
and daguerrotypes, (the best photographers were able to prepare
the plate, expose it, and reveal the image in few minutes), in the
second half of the 19th century, especially in the United States,
they gained a lot of popularity [8]. Indeed, the new materials
highly reduced the cost of production and were characterized
by a high durability. From the rst picture obtained by Niepce
in 1827, many different technologies allowed the production of
more and more versatile photographic plates. In this case, the
use of the bitumen of Judea deposed on a pewter plate required a
very long exposition time (up to eight hours) without warranting
a good resolution. Due to these reasons, this kind of technology
did not have large diffusion. On the contrary, the daguerrotypes
(Daguerre obtained the rst a fewyears before the ofcial inven-
tion in 1839) were the rst highly expensive (the photographic
layer was made mainly by silver and silver salts) photographic
documents that, in the second half of the nineteenth century, had
a large diffusion, especially in the high classes of the western
countries. Due to the drastic reduction of the cost, only with the
successive introduction of the ferrotypes the photography had a
wide achievement also in the middle class, playing a very impor-
tant role also fromthe historical point of view. All the categories
of photographs between 1839 and the rst decade of 1900, due
to the variety of materials used for their production, present dif-
ferent chemical and physicochemical characteristics [9]. Also
the degradation phenomena are strictly related to the composi-
tion of the layers constituting the structure of the photographic
plate [1012].
To obtain useful information about the conservation status, in
view of dening possible conservation protocol or conservation
conditions to be adopted, a chemical diagnostic investigation
plays a primary role in the preliminary phase of the work [13].
Even if in the last 30 years, the interest towards ancient photo-
graphic materials grew up, the scientic literature on ferrotypes
characterization and conservation is still scarce [13,14].
The aim of the present work was to present and compare the
results obtained during the physicochemical characterization of
Fig. 1. Image of the ferrotype F1 analyzed. The dimensions are 6.5 8 cm.
the conservation status of two 19th centurys ferrotypes from
England (Figs. 1 and 2 and labeled as F1 and F2, respectively) by
means of noninvasive techniques like microreectance Fourier
transform infrared spectroscopy (-FTIR), optical microscopy
Fig. 2. Image of the ferrotype F2 analyzed. The dimensions are 5 8.5 cm.
E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508 503
Fig. 3. Microreectance -FTIR spectrum of a nitrocellulose standard in the
range 8004000 cm
-1
(A) and magnication of the ngerprint region (B).
(OM) and environmental scanning electron microscopy (ESEM)
coupled with EDXanalysis, all techniques usually employed for
the characterization of several artifacts [1520].
2. Materials and methods
Diethylether, ethanol (for both purity > 99.0%) and nitrocel-
lulose were purchased from Aldrich and used as received, water
was puried with a Millipore MilliRO-6 and MilliQ (Organex
System) apparatus: the resistance of the water was greater than
15 Mcm.
ESEM images were collected by means of a FEI Quanta 200
microscope coupled with an EDX(SUTWdetector). The dimen-
sions of the microscope chamber were 20 15 cm, allowing the
introduction of the full plates and avoiding the necessity of sam-
pling. The working distance was 10 mm and the acceleration
potential 30 KV. For the morfological analysis, ESEM was cho-
sen because it was possible to collect the electronic micrographs
without graphitizingor gildingthe sample surface. This noninva-
sive technique is commonly used for biological samples [21,22]
and also for the study of artistic [23] and photographic mate-
rials like albumen photographs [24]. The EDX spectra were
collected using a microprobe coupled with the ESEM micro-
scope.
-FTIRspectra were obtained in the micro-reectance mode
using a BioRad FTS-40 spectrometer equipped with a BioRad
UMA500 microscope (MCT detector) with 8 cm
1
resolution
and 512 scans. Contact angle measurements [25,26] were per-
formed onto the ferrotype surface by means of a PC-controlled
NRLRam-Hart Inc. apparatus interfacedtoa PC. Water droplets
(5 l) were deposed onto the solid surface using a Hamilton
microsyringe and the contact angle was determined 5 s after the
deposition. The contact angles reported were the average over
10 measurements and the errors were the standard deviations.
All the optical micrographs were collected in the reectance
mode by means of a Reichert Zetopan optical microscope,
equipped with 5.5, 11, 28objectives and a 8ocular. All
the images collected in polarized light were in crossed nicols
modality.
Fig. 4. Microreectance -FTIR spectra collected from the front side of the
ferrotype F1 (A, spectrum1) and F2 (B, spectrum1) compared with the spectrum
of the standard of nitrocellulose (spectrum 2) in the ngerprint region.
504 E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508
Fig. 5. Optical micrographs of the ferrotype F1 (A and B) and F2 (C and D). B and D are magnications of A and C, respectively.
3. Results and discussion
For the FTIR analysis, rst a standard of not aged nitrocellu-
lose was prepared by solubilization of this polymer (7%w/w) in
a solution of ether plus alcohol. A portion of this solution was
put onto a Teon plate and, once dried, the obtained lm was
analyzed by means of micro-reectance -FTIR. The spectrum
(Fig. 3) was used as reference and compared with the ones col-
lected from the ferrotype surface. The diagnostic peaks for the
identication of nitrocellulose were the ones below 1800 cm
1
(Fig. 3B). In particular, the peak at 1730 cm
1
, due to the stretch-
ing of the C O bond, the peak at 1650 cm
1
(stretching of the
C Nbond), andthe peakat 1280 cm
1
(stretchingof the O NO
2
group). The band at 3500 cm
1
is due to the stretching of the
OH; the peaks between 3100 cm
1
and 2800 cm
1
are due to
the stretching of aliphatic CH
2
and CH
3
(Fig. 3A).
Fig. 4 reports the microreectance -FTIR spectra collected
from the front side of the two ferrotypes (spectra 1A for F1
and 1B for F2). By comparison between them and the stan-
dard spectrum of nitrocellulose (spectra 2A and 2B), it was
evident that, as expected for ferrotypes, collodion has been used
as binder for the photosensitive grains that allow the forma-
tion of the image. In both cases, a correspondence of the FTIR
signals with the ones of the standard was observed in the n-
gerprint region. The two surfaces of F1 and F2 (front side) were
very similar, with the presence of collodion in both cases. It
was interesting to notice that both the spectra showed the pres-
ence of the peak at around 1560 cm
1
attributable to the C O
amide II typical of proteinaceous materials. This could be due
to the presence of some protein-based protective coatings, base
for retouching/pre-coatings [3]. No natural resins as protective
varnishes were detected by -FTIR: this could be due to the pres-
ence of the collodion whose strong peaks matches the typical
spectral features of natural resins.
In order to deepen the investigation, morphological analysis
of the two ferrotypes was performed by means of optical and
ESEM microscopy. As indicated in the optical micrographs in
Fig. 5A and C, at a microscopic level, the morphology of the
two ferrotypes was totally different. While the surface of the
ferrotype F1 was characterized by a roughness (in the order of
micrometers) similar in all the regions of the plate, the one of
F2 appeared completely at. This nding could be attributed to
the presence of a morphologically different protective surface
layer deposed onto the F2 ferrotype. This was also conrmed by
contact angle measurements. While the contact angle value for
the ferrotype F2 was almost the same in both the front and the
back sides (around 60 7

), in the plate F1 this parameter varied


from a value close to 60
o
in the back and in some regions of the
image where the grazing light analysis showed the presence of
a shiny surface layer, to a much higher value (around 75 5

)
in the regions where this layer was not detectable. From these
data, it was possible to conclude that some regions of the front
side of F1 were less hydrophilic than its backside and than both
sides of F2. This was in agreement with the presence on both
the surfaces of a material more hydrophilic than collodion (i.e.
proteinaceous), except in the degraded regions of the front side
of F1 where the coating was lost and the wettability showed a
behaviour typical of more hydrophobic substances (i.e. collo-
dion). These ndings, associated with the morphology shown in
Fig. 5, indicated a worse state of conservation of the ferrotype
E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508 505
Fig. 6. ESEM micrographs obtained from the surface of the ferrotype F1 obtained with secondary (A) and back-scattered electrons (B). (C) Image obtained with
back-scattered electrons indicating the two points where elemental analysis was performed. (D) EDX spectrum collected in the Point 1.
F1. Then it was possible to conclude that the conservation sta-
tus of the two plates was different and particularly F2 showed a
surface texture better than F1.
The pattern seen in the optical micrographs reported in
Fig. 5A and B were extraordinary similar to the well-known
dewetting gures [27]. This effect could be due to two different
alternative mechanisms of dewetting of the collodion from
the support, leading in certain cases, to the appearance of true
holes. (i) The amount of solvent used for its preparation was too
large causing the formation of a low viscous collodion-based
layer inducing a rapid local dewetting from the plate; as a
consequence, after the exposition to the light, the solvent
rapidly evaporated causing the formation of the nitrocellulose
layer embedding the reduced Ag
0
clusters, the residual AgX
crystals, and characterized by a morphologically irregular
surface. (ii) An extremely slow dewetting process on the time
scale (years) typical of photographic collodion deterioration.
Unfortunately, we did not have experimental evidences to
discriminate between these two possibilities.
In the optical micrographs in Fig. 5A and B, it was also evi-
dent the presence of small circular features (Point 1inFig. 5A). A
simple morphological investigation suggested that they could be
constituted of crystals or solid particles that precipitate as impu-
rities into the collodion layer. In order to verify this hypothesis,
the investigation by means of the ESEM/EDX microscope was
particularly useful. Fig. 6 shows the micrographs obtained with
secondary and back-scattered electrons (Fig. 6A and B, respec-
tively) that did not show any difference in the tone of the grey in
correspondence of these spots. This nding was also conrmed
by the elemental analysis carried out both in a dark spot and also
in a point of the surrounding area (Fig. 6CPoints 1 and 2, respec-
tively) and reported in Table 1. This indicated that the chemical
elemental composition was almost the same. This result was
a further conrmation that the dark circular features were sim-
ply surface imperfections and that the drying mechanism of the
wet collodion layer could be considered as the main cause of the
particular morphology of the ferrotype surface. The presence of
the chloride element (between 4 and 5%) could be attributed to a
washing process carried out using Cl

-contaminated water, or to
the contamination coming from other sources. The presence of
S came fromthe xing process made by S
2
O
3
2
. EDX(Table 1)
indicated also the presence of Fe, K, and Mg. The rst was prob-
ably due to the presence of Fe-compounds originated from the
Table 1
Elemental composition from EDX (expressed as atomic %) of the Points 1 and
2 in Fig. 6C.
Element Atomic % Point 1 Atomic % Point 2
Mg 11.6 11.5
K 1.7 1.9
S 5.0 4.5
Cl 4.9 4.2
Ag 72.3 73.2
Fe 4.5 4.6
506 E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508
Fig. 7. (A) Image of the ferrotype F1; in the inset the region where a lacuna of collodion was present. (B) ESEM micrograph obtained with secondary electrons, the
arrow indicates the region where the elemental analysis was performed (C is the magnication of B in proximity of the arrow and X is the point were EDX spectrum
was collected). (D) SEM micrograph obtained with secondary electrons of a naturally aged nitrocellulose lm.
degradation of the plate; the second ones were due to remaining
of impurities of potassium and magnesium halides used for the
precipitation of the photosensitive materials.
Fromthe ESEM/EDXanalysis, it was also possible to achieve
information about the composition elemental of the two pho-
tographic plates. As it is shown in Fig. 7A and indicated by the
arrow in Fig. 7B, in the F1 plate there was a region where part
of the collodion layer was lost, but -FTIR indicated that a
residual amount of collodion was still present. The elemental
analysis performed in that area (Fig. 7C and Table 2) showed
that the amount of Ag was much smaller than the one previously
detected onto the surface of the same plate (Table 1). This effect
was not surprising due to the fact that for ferrotypes Ag is con-
centrated onto the supercial layer of the collodion lm where
the photosensitive grains of AgXare precipitated. Also the mor-
phology of F1 was completely different from that of F2 (vide
infra, Fig. 8) and it was similar to the one of a naturally 80 years
Table 2
Elemental composition from EDX (expressed as atomic %) of the Points 1 in
Fig. 7C.
Element Atomic %
Na 4.9
Si 5.3
S 70.1
Ag 2.6
K 5.8
Fe 11.3
aged nitrocellulose lmwhich ESEMmicrographs is reported in
Fig. 7D. This sample was obtained from a portion of a naturally
aged Italian motion-picture lm of the 1920s made by cellulose
nitrate. Particularly, the pattern of the fractures observed in the
ESEM micrographs (Fig. 7C and D) was similar, indicating that
their formation was mainly due to the aging of the collodion
layer. Finally, Na was also detected because the xing agent
S
2
O
3
2
was used as sodium salt. Finally, the presence of Si
could be attributed to the deposition of powder onto the region
were the collodion lacuna was present.
Fig. 8. ESEM micrographs collected from the surface of the ferrotype F2. 1 and
2 indicate the points where the elemental analysis was performed.
E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508 507
Table 3
Elemental composition from EDX (expressed as atomic %) of the Points 1 and
2 in Fig.8.
Element Atomic % Point 1 Atomic % Point 2
Ag 28.0 72.2
Fe 72.0 22.2
S / 5.6
ESEM analysis was also performed onto the F2 surface (see
Fig. 8 and Table 3): this investigation highlighted that, with
respect to F1, there were three main differences. First, the sur-
face appeared much smoother and more homogeneous; second,
the amount of Ag was less homogeneous in passing from a
point to another zone of the surface; third, the concentration
of iron was quite higher. This difference in Ag and Fe con-
centrations could be ascribable to a thinner collodion layer for
F2.
4. Conclusions
The multi-technique approachpresentedinthis paper allowed
to study for the rst time the chemical elemental compo-
sition and the conservation status of two ferrotypes made in
England at the end of the 19th century, without the collection of
any samples, i.e. by means of noninvasive techniques. Micro-
reectance FTIR indicated that some proteinaceous materials
were present, ascribable to protective coatings for retouching
and/or pre-coating and no presence of natural varnishes was
found.
The morphological investigation, performed both by means
of optical and ESEMmicroscopy, allowed to establish the differ-
ent conservation status of the two ferrotypes, indicating that F2
sample surface was homogeneous and not meaningfully dam-
aged, while F1 surface roughness was clearly high. Contact
angle measurements further conrmed this feature: indeed, only
F1 surface was characterized by the presence of a texture
attributable to the surface dewetting process of the collodion
caused by the kinetics of the evaporation of the solvents used
for the preparation of the collodion layer. Moreover, contact
angle measurements clearly indicated that for F1 the wetta-
bility was altered, probably by a not homogenous protective
layer.
The analysis carried out by EDX spectra showed that even
the chemical elemental composition was different comparing
the two plates. In fact, F2 plate presented Ag and Fe amounts
in agreement with a quite thin collodion layer, whereas F1 plate
showed lower concentration of Fe indicating a thicker collodion
layer. This was quite interesting since we were able to conclude
that the thickness of the collodion stratum and concentration of
silver in its upper layers are parameters not directly associated
to the conservation status of the ferrotypes.
In conclusion, our study allowed to compare the morphology,
elemental composition, and surface wettability of two ferrotypes
of the same period enabling to underline the differences of these
parameters associated to the different status of conservation,
despite their similar age and provenance.
Acknowledgements
The Authors wish express their gratitude to the Reviewer
whose comments and criticisms were fundamental for improv-
ing the paper. Thanks to Mr. Leopoldo Morandi, who gave us
the opportunity to study the ferrotypes and some very important
historical indications about the two ferrotypes analyzed, and to
Dr. Giovanna Poggi for useful suggestions. Authors also thank
the Cineteca Comunale, Bologna, Italy, for giving the cellu-
lose nitrate motion-picture lm sample. Thanks are also due to
MIUR, Italy (PRIN 2006), to the Consorzio Interuniversitario
per lo sviluppo dei Sistemi a Grande Interfase (CSGI-Firenze)
for nancial support, and to the University of Florence (Fondi
dAteneo ex-60%).
References
[1] WilliamWelling, PhotographyinAmerica, Crowell, NewYork, USA, 1978.
[2] L. Scaramella, Materiali della cultura artistica, in: Fotograa. Storia e
riconoscimento dei procedimenti fotograci, Edizioni De Luca, Roma,
1999.
[3] E.M. Eastbrooke, The Ferrotype and Howto Make it, Linsday Publications,
USA, 2007.
[4] G. Berkhofer, Wet Collodion Photography A Short Manual, Lulu.com
Publications, USA, 2007.
[5] C.B. Neblette, Fundamentals of Photography, Van Norstrand Reinhold Co,
Princeton, 1970.
[6] The reaction that allows the solubilization of AgX is:
AgX+2S
2
O
3
2
[Ag(S
2
O
3
)
2
]
3
+X
-
. Then, the equilibrium constant is:
K
eq
=

[X

][Ag(S
2
O
3
)
3
2
]

/[(S
2
O
3
)
3
2
]
2
=Kps
(AgX)
Kst
[Ag(S
2
O
3
)
3
2
]
Kps
(AgCl)
= 1.8 10
10
; Kst
[Ag(S
2
O
3
)
3
2
]
= 2.9 10
13
; Kps
(AgBr)
=
5.0 10
13
If X corresponds to Cl, then the Keq =5.2 10
3
, if X is Br,
Keq is 14.4. In both cases the solubilization of AgX is complete. A similar
behaviour is also observed if CN
-
is used as xing agent instead of S
2
O
3
2
Kst
[Ag(CN)
3
4
]
= 4.0 10
20
. All the constant values are from: J.A. Dean,
Langes Handbook of Chemistry, McGraw-Hill, New York, 1985.
[7] C.E.K. Mees, T.H. James, The Theory of the Photographic Process, third
ed., The Macmillan Co, New York, 1966.
[8] W. Crawford, Let del collodio, Cesco Capanna Editore, Milano, 1981.
[9] B. Cattaneo, D. Chelazzi, R. Giorgi, T. Serena, C. Merlo, P. Baglioni,
Physico-chemical characterization and conservation issues of photographs
dated between 18901910, Journal of Cultural Heritage 9 (2008) 277284.
[10] K.B. Hendriks, The stability and preservation of recorded images, in: J.
Sturge, al. et (Eds.), Imaging Processes and Materials, Van Nostrand Rein-
hold, New York, 1989, pp. 637684.
[11] R.W. Henn, D.G. Wiest, Microscopic spots in processed microlms: their
nature and prevention, Photographic Science and Engineering 7 (1963)
253261.
[12] G. Di Pietro, Silver mirroring on silver gelatin glass negatives, PhDThesis,
University of Basel, 2002.
[13] N. Davis, Tintypes: preliminary research and testing, in: Proceedings of
the art conservation training programs conference, Coopers Town, 1984,
pp. 1328.
[14] Bovis M., Les anciens procds photographiques : V. Amphitypie, fer-
rotypie, collodion sec, Photo-Cine-Revue, October 1971, pp. 450453.
[15] S. Masanori, Y. Sasaki, Studies on ancient Japanese silk bres using FTIR
microscopy, in: Postprints of the First Annual Conference Scientic Anal-
ysis of Ancient and Historic Textiles: Informing Preservation, Display and
Interpretation, 1315 July 2004, Archetype Publications, London, 2005,
pp. 4447.
[16] J. Van der Weerd, A. Van Loon, J. Boon, FTIR studies of the effects of
pigments on the aging of oil, Studies in Conservation 50 (2005) 322.
508 E. Carretti et al. / Journal of Cultural Heritage 10 (2009) 501508
[17] S.A. Centeno, M.I. Guzman, A. Yamazaki-Kleps, C.O. Vdova, Char-
acterization by FTIR of the effect of lead white on some properties of
proteinaceous binding media, Journal of the American Institute for Conser-
vation 43 (2004) 139150.
[18] A. Langley, A. Burnstock, The analysis of layered paint samples from
modern paintings using FTIR microscopy, in: Preprints of the 12th ICOM
Triennial Meeting, Lyon, 29 August3 September 1999, James & James,
London, 1999, pp. 234241.
[19] D. Magaloni, M. Aguilar, V. Casta no, Electron and optical microscopy of
Prehispanic mural paintings, in: P.B. Vandiver, J. Druzik, G.S. Wheeler
(Eds.), Proceedings of the Symposium Materials Issues in Art and
Archaeology II, San Francisco, April 1721, Materials Research Society,
Pittsburgh, 1991, pp. 145150.
[20] H.B. Berg, Characterization of printing papers by optical and electron
microscopy, Tappi 49 (1966) 554562.
[21] L. Muscariello, F. Rosso, G. Marino, A critical overview of ESEM appli-
cations in the biological eld, Journal of Cellular Physiology 205 (2005)
328334.
[22] V.A. Edward, V.L. Pillay, P. Swart, Localisation of Thermomyces lanug-
inosus SSBP xylanase on polysulphone membranes using immunogold
labelling and environmental scanning electron microscopy (ESEM), Pro-
cess Biochemistry 38 (2003) 939943.
[23] A. Masic, E. Badea, R. Ceccarelli, G. Della Gatta, S. Coluccia. Compara-
tive DSC and SEM/EDX study of ancient and articially aged parchment,
Proceedings of the II National Conference Lo stato dellArte 2: Conser-
vazione e Restauro, Confronto di Esperienze, Genova, 2729 settembre
2004, Il Prato, Saonara, pp 5261.
[24] P. Messier, T. Vitale, Cracking in albumen photographs: an ESEM investi-
gation, Microscopy Research and Technique 25 (1993) 374383.
[25] J. Drelich, J.L. Wilbur, J.D. Miller, G.M. Whitesides, Langmuir 12 (1996)
19131922.
[26] J. Drelich, J.D. Miller, A. Kumar, G.M. Whitesides, Colloids and Surfaces
A: Physicochemical and Engineering Aspects 93 (1994) 113.
[27] C. Neto, K. Jacobs, R. Seemann, R. Blossey, J. Becker, G. Grn, Satellite
hole formation during dewetting: experiment and simulation, Journal of
Physics: Condensed Matter 15 (2003) 33553366.

You might also like