The Noli me Tangere: study and conservation of a

advertisement
Technical Research Bulletin
VOLUME 5
2011
The Noli me Tangere: study and conservation of a
Cretan icon
Lynne Harrison, Janet Ambers, Rebecca Stacey, Caroline
Cartwright and Angeliki Lymberopoulou
Summary A seventeenth-century icon of the Noli me Tangere (1994,0501.3), in the collections of the Department of Prehistory and Europe at the British Museum, has been the subject of detailed technical examination
from scientific, conservation and historical perspectives. The aims were to investigate its original materials
and manufacture, its treatment history and, if possible, to look for indications of function and use in Orthodox
practice, with the ultimate purpose of informing the conservation required to stabilize the icon’s structure and
address the damaged condition of the image. Research into the icon’s original context was also undertaken
to strengthen its provenance. The wooden panel was identified as cypress and the original palette was found
to comprise lead white, gypsum, carbon-based black, red and yellow earths, a basic copper carbonate green,
verdigris, indigo, azurite, vermilion, red lead and a red lake. The painting was found to have been originally
coated with an oil and resin varnish and two campaigns of restoration were also identified. The results of
the study enabled a suitable conservation protocol to be devised and applied. This involved removal of the
degraded restoration varnish and the unstable restoration from the early twentieth century to reveal the fine
original painted surface. Those areas of restoration thought to date from the painting’s early history, including
repairs to the edges and the complete regilding of the background, were left in place. The results of the study
also support a Cretan origin for the icon.
INTRODUCTION
A seventeenth-century icon of the Noli me Tangere
(1994,0501.3), currently held by the Department of Prehistory and Europe at the British Museum, was purchased
in Chania, Crete (probably in 1894) and donated to the
National Gallery, London in 1924, before being transferred
to the British Museum in 1994 [1; Note 3]. Painted on a
wooden panel measuring 606 × 472 mm and coated with
a glossy and darkened varnish (Figure 1), the image shows
the announcement of Christ’s resurrection organized in six
successive scenes accompanied by Greek inscriptions.
A detailed technical examination of the icon was carried
out, the aims of which were to investigate the original materials and manufacture, the treatment history and, if possible,
to look for indications of function and use in Orthodox
practice, while informing the conservation needed to stabilize the icon’s structure and address the damaged condition
of the image. Research was also undertaken into the icon’s
original context in an attempt to strengthen its provenance
and to investigate a possible link to the post-Byzantine
Cretan artist Michael Damaskinos (1530/35–1592/93)
[2; p. 458 No. 100].
TECHNICAL EXAMINATION
Technical examination was carried out using X-radiography, emission radiography, ultraviolet (UV) and infrared
(IR) examination and photography, optical microscopy,
Raman spectroscopy, scanning electron microscopy with
energy dispersive X-ray analysis (SEM-EDX) and gas
chromatography-mass spectrometry (GC-MS), see the
experimental appendix for details. Layer structures were
investigated and the materials used in the original production and later interventions were identified. Unfortunately
it was not possible to sample all areas of the painting so
the results given below, while as comprehensive as possible,
cannot be viewed as exhaustive.
Large quantities of data were generated in the process
of this study and it is not possible to include them all here.
25
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
figure 1. Icon of the Noli me Tangere (1994,0501.3) before conservation. Note the highly glossy varnish
seen clearly at the bottom edge
Instead, the most significant results have been selected for
inclusion in Tables 1 and 2, and for discussion and interpretation below. Full details from the study are contained
in a British Museum internal report that can be accessed
through the online catalogue of British Museum icons [3].
Original materials
A single wooden panel of Cupressus sempervirens L.,
cypress (sometimes called Mediterranean or Italian
cypress) was cut in the radial longitudinal axis and
prepared so that it was smooth on the front surface, with
a rougher finish (with traces of adze work visible) on the
reverse [4; p. 33]. Two well-fitting wooden battens, also
of cypress, were then attached horizontally to the reverse
26
of the panel, at equal distances from the top and bottom
edges. No nails are visible on the reverse, but X-radiography revealed that short, flat-headed nails had been
inserted from the front of the panel, equally spaced along
the length of the battens.
A layer of plain (tabby) weave cloth was glued onto
the smoothed front surface of the panel. While it was not
possible to sample or identify either the cloth or glue used,
linen soaked in animal glue was usually employed for this
purpose [5; p. 28]. A white ground of gypsum bound in a
proteinaceous medium was then applied to the whole front
surface of the panel and burnished smooth. No ground was
applied on the reverse of the icon.1
Evidence from the IR reflectograms, which reveal
elements of the underdrawing, suggests that the layout of
the image, including the positions of the figures, drapery
Raman + optical microscopy
GC-MS
Physical tests
Paint (pigments)
Paint (binder)
Ground (binder)
Raman + optical microscopy
GC-MS
GC-MS
GC-MS
Pigments
Paint (binder)
Indigo overpaint
(binder)
Ground (binder)
SEM
Raman
Optical microscopy
Raman + optical microscopy
GC-MS
Raman + optical microscopy
Staining tests
Visual assessment
Optical microscopy
Gold
Bole
Mordant
Paint (pigments)
Paint (binder)
Ground
Ground (binder)
Cloth
Wood
Cupressus sempervirens L., cypress
Plain/tabby weave
Gypsum
Protein
Vermilion; lead white; carbon-based black; azurite; indigo; copper carbonate green (verditer?);c red lead; goethite, hematite; and gypsum
Verdigris; earth pigments; and red laked
Conifer resin; drying oil
Thick, pigmented and organic-rich (pigments and medium not identified)
Gold alloy (c.93% Au, 4% Ag, 3% Cu; c.22.4 carat equivalent)
Goethite- and hematite-containing red earth
Conifer resin and drying oil
Animal glue (gelatine)
Red lead; hematite; goethite; carbon-based black; vermilion; indigo; gypsum; and red lake
Fat (egg?)/oil; conifer resin; and sugars (gum?)b
Conifer resin and oil
Conifer resin; drying oil; and Pistacia resin (tr)
Water soluble
Cadmium yellow; vermilion;a Prussian blue; and carbon-based black
Conifer resin; drying oil; and Pistacia resin
Pistacia resin; drying oil; and conifer resin (tr)
Results
‘tr’ indicates a trace amount was found.
a
The small particle sizes suggested that the vermilion was produced using the wet process.
b
The presence of sugars is not understood at this time.
c
This may be natural malachite, but the regular, small and spherical particles make the use of artificial green verditer more likely.
d
Given the techniques available it was not possible to determine the organic colorant in the red lake.
Notes
GC-MS
Varnish
Original layers
GC-MS
Varnish
Early restoration materials
GC-MS
Analytical method
Varnish
Later restoration materials
Material
table 1. Summary of the main analytical results
THE NOLI ME TANGERE: STUDY AND CONSERVATION OF A CRETAN ICON
27
28
CS10 Angel’s grey robe, with old varnish residues. Ground layer missing
CS9 Green of angel’s wing with red underpaint. Upper varnishes removed
CS7 Overgilding and overpaint over original paint from edge of cliff to left
Restoration
10. Varnish, which shows white luminescence under UV illumination; to the left it is over layer 8 and penetrates through a crack to the
level of the original paint layers;
9. Paint layer comprising finely ground pigments (over original layers at right) – earths, red lake, carbon-based black (Raman);
8. Paint layer between two thin UV-luminescent coatings (at the left of the sample); and
7. Thick varnish showing cream luminescence under UV illumination
CS2 Christ’s left ankle, area of gilded decoration. Sample includes all layers of
varnish and possibly recent overpaint
Original
2. Dark grey thin paint layer – carbon-based black, lead white (both by Raman); and
1. Light grey thicker paint layer – carbon-based black, lead white and azurite (all by Raman)
Restoration
5. Some fragmentary particles on surface, possibly overpaint;
4. Non-luminescent surface layer – possibly calcium oxalate (visual examination, unable to confirm by Raman);
3. Remains of varnish layer above and below the oxalate layer showing cream luminescence under UV illumination;
Original
4. Varnish, possibly original, showing faint white luminescence under UV illumination;
3. White highlight paint – lead white (Raman);
2. Green paint – manufactured green copper carbonate and azurite (both by Raman); and
1. Red paint – vermilion, lead white (both by Raman) and red lake
Restoration
6. Remains of two surface coatings to the right. Some particles show luminescence under UV illumination; and
5. Dark layer penetrating into crack
Original
5. Dark layer – dirt(?);
4. Thin varnish layer showing luminescence under UV illumination;
3. Original dark green paint – azurite, lead white and carbon-based black (all by Raman) and yellow (unidentified);
2. Original light yellow-green paint (unidentified); and
1. White paint – lead white (Raman) plus fragment of gold leaf
Restoration
10. Two upper varnish layers showing white luminescence under UV illumination;
9. Gold leaf;
8. Red ‘bole’ type layer;
7. Dark overpaint, medium showing luminescence under UV illumination; and
6. Lower thick varnish that shows cream luminescence under UV illumination and penetrates into crack in original paint
Original
6. Metal leaf – gold (SEM-EDX);
5. Thick, pigmented organic-rich mordant;
4. Greenish-yellow layer of flesh paint – earth pigments and verdigris (optical examination only) topped with thin pinkish-red layer of
flesh paint – vermilion and lead white (Raman);
3. Warm white paint – lead white, carbon black and yellow ochre (Raman);
2. Dark layer; and
1. Gesso – gypsum (Raman) [protein]
Results (layer structure from surface down)
Sample
table 2. Selected paint cross-sections showing stratigraphy
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
Pigment identifications given are based on the method given in brackets and nature of the organic medium / varnish is based on microchemical staining tests.
Note
Restoration
6. Upper varnish showing white luminescence under UV illumination;
5. Pigment and gold overpaint – black and red particles (unidentified); and
4. Thick brittle varnish showing cream luminescence under UV illumination
CS13 Gold overpaint layers from area of damage, top centre of gilded
background and including original layers beneath
Original
3. Gold leaf;
2. Red ‘bole’ layer – hematite (Raman); and
1. Gesso
Original
4. Crust – degradation visible on surface; probably a calcium oxalate crust (unable to confirm by Raman);
3. Red glaze – red lake;
2. Red paint – vermilion (Raman) and red lake; and
1. Brownish paint – earth pigments (goethite and hematite) and carbon-based black (all by Raman)
CS12 Damaged red paint from kneeling figure to left. Restoration varnish
removed
THE NOLI ME TANGERE: STUDY AND CONSERVATION OF A CRETAN ICON
29
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
figure 2. Details from the IR reflectogram in the area of Christ’s
drapery showing carbon-based spots from pouncing of an original
anthibolon. The image to the right is an enlargement of the region
bounded by the rectangle in the image to the left
and architecture, was transferred to the ground layer from
an anthibolon (a cartoon or pattern used for tracing the basic
designs of works) by pricking and pouncing through the
pinholes with a carbon-containing material. The resulting
dots were then joined by drawn lines, Figure 2 [6; pp. 56–60
and 76–79, 7; pp. 169–170, 8]. From the X-radiographs it
is clear that some of the lines were then further incised
into the ground with a sharp implement so that they
could be seen throughout the painting process, Figure
3. Interestingly, the trees and rocks of the background,
the angels’ wings and the faces of the figures were not
incised, Figure 4. Further evidence of underdrawing,
without accompanying incised lines, can be seen in the IR
reflectogram, for example in the branches of the trees and
in the complete drapery of the left figure and folds of the
sleeve of the middle figure in the scene to the upper right
showing the Myrrophoroi (unguent bearers), Figure 5 [1;
p. 191]. Only slight changes in design (pentimenti)
between the underdrawing and the finished image are
evident, as in the hem of the drapery of the middle figure
in the Myrrophoroi scene, Figure 5.
A dark, patchy wash of colour on the faces and hands is
also apparent in the IR reflectograms and is interpreted as
an underpainting or proplasmos [9; p. 234].
Following the production of the underdrawing, the areas
to be gilded (the background and halos) were first coated
with a bole based on red earth, burnished and then covered
figure 3. Details from the X-radiograph in the area of Christ’s drapery
showing incised lines. The image to the right is an enlargement of the
region bounded by the rectangle in the image to the left
30
figure 4. Diagram indicating the positions of incised lines as deduced
from the X-radiographs
with thin metal leaf of soft, high-purity gold (see Table 1),
which was further burnished.
Examination of the layer structure revealed details of the
composition of the original paint layer and two additional
restoration layers; in some places the ground preparations
were also present, see below. The original painting technique was distinct from that used in the restoration layers in
pigment choice, preparation and colour mix, making it easy
to distinguish original workmanship from later additions.
As is common in icons, the use of restoration to maintain
a complete and functional image for worship has produced
a complex and confused layer structure with penetration
of solvents and media from layer to layer making it impossible to identify the original paint medium with certainty [3;
Section 4.1]. Conifer resin and drying oil were found in the
original paint layers but no proteins or fats were identified,
although their presence, perhaps only in small quantities,
may have been masked by other materials.
The original palette consisted of lead white, gypsum,
carbon-based black, red and yellow earths (coloured by
goethite and hematite), a basic copper carbonate green,
verdigris, indigo, azurite, vermilion, red lead and a red lake,
see Table 1. The image was built up in layers, with simple
paint mixtures of two or three pigments.
Selected passages of the painting have been studied in
greater detail; see Table 2 for details of the cross-sections
taken from these areas. The flesh tone of the large figure
of Christ was painted with a yellow-green paint containing
a mixture of yellow earth, verdigris and an unidentified
white pigment, CS2: Table 2. Highlights of lead white were
applied over this layer with the addition of small amounts of
vermilion for areas of warm flesh tones. A similar yellowishgreen paint mixture was used for some of the background
THE NOLI ME TANGERE: STUDY AND CONSERVATION OF A CRETAN ICON
cliffs (CS7: Table 2), where a lighter layer was laid down first
and then darker passages applied on top, with the addition of
azurite and carbon-based black to the paint mixture. A fragment of gold leaf was found within the lowest paint layers.
The grey robe of the angel to the left of the empty tomb
was painted using a mixture of a carbon-based black, lead
white and azurite with the darker folds added on top using
a mixture comprising only lead white and carbon-based
black, CS10: Table 2. The angel’s wings were first painted
with red (a mixture of vermilion, lead white and a red
lake) and then edged with greenish wing tips of a mixture
of malachite (or perhaps its artificial analogue green verditer) and azurite, with white highlights on top, CS9: Table
2. To the lower left, in the scene of the Chairete (‘All Hail’),
the red drapery of the kneeling figure of the Virgin Mary is
quite distinct from the other reds used in the image. This
effect was produced using three paint layers, CS12: Table 2.
The first brownish-red layer was coloured with earths and
carbon-based black and this was coated with a second thin
layer of semi-opaque red (a mixture of vermilion and red
lake) to which a comparatively thick third layer of a translucent red lake was finally applied.
After painting, mordant gilding was used to decorate Christ’s cloak and sandals, and the angels’ wings. The
mordant from an area of Christ’s sandal, visually identified
as an organic binder bulked out with inorganic pigment,
was applied to the painted surface and then coated with gold
leaf that was left unburnished, CS2: Table 2.2 The surface
of the painting was then coated with a varnish containing
a mixture of drying oil and conifer resin, the remains of
which were located in the paint cross-sections during analysis and confirmed during conservation.
Later changes (damage and restoration)
At some point, possibly relatively soon after completion,
the icon suffered extensive damage to the wood panel and
painted surface by wood-boring insects, probably attracted
to the glue-impregnated cloth layer as a potential food
source. In general the damage was restricted to the uppermost surface of the panel, with only few areas of activity
visible on the reverse, which are discussed below. As a
result, original paint and ground were damaged or lost and
restoration was clearly undertaken. The painted surface was
cleaned to remove the original varnish, resulting in some
damage to the paint (particularly the Virgin’s red robe in
the Chairete scene) and etching of the surface of the gilded
halos and mordant gilding, Figure 6. The lost areas – particularly at the edges – were replaced with a white ground
layer bound in animal glue and an attempt was made to
complete the detail of the lost image using pigments that
were probably bound in egg or an egg/oil mixture, Table
1. An exception to this was the indigo paint used to cover
areas of loss in the trees, which was oil-based. Curiously,
the restoration layer was not built up to the same level
as the original surface, but sat just below, forming a step
figure 5. Detail from the IR reflectogram showing underdrawing in:
(a) the tree branches; and (b) the scene of the Myrrophoroi
around the edges of the damages. Some of the insect flight
holes were also plugged with white fill and overpainted.
The early restoration palette included red lead, vermilion,
indigo, ochres, a red lake and gypsum, and differs from
the original palette both in pigment mixtures and the size
of the pigment particles. For example, restoration of the
red coffin in the Myrrophoroi scene, originally painted in
vermilion, was carried out in a mixture of red lead and
a red lake. Larger areas of loss of original ground in the
gold background were also replaced and the whole of the
background was regilded. Of the original gilding, only the
halos and traces in the background remain (see below). The
profiles of the mountains and architecture were repainted
and a black inscription added on top of the gold background between the mountains to the right. The original
crosses were completely overpainted. A comparatively thick
coating of an oil and conifer resin varnish was then applied
across the whole surface including the restorations and the
original paint.
figure 6. Photomicrograph of the mordant gilding on Christ’s sandal
strap showing loss of original mordant gilding beneath the cracked
and brittle restoration varnish. Image size 12 × 9 mm
31
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
figure 7. Icon during conservation showing those areas (shaded in pink) that comprise earlier restorations
or regilding, plus the black inscription; all of these were left in place
A later, probably early twentieth century, restoration
campaign has also been identified, which concentrated on
the lower left edge and left side of the icon, but included
other scattered areas of retouching across most of the
surface. The losses were replaced with white, water-soluble
filler and inpainted using pigments that were probably
bound in oil. The palette included traditional pigments
such as vermilion for the overpainted red border and
some post-eighteenth-century materials such as cadmium
yellow and Prussian blue; the latter was used to retouch
the Magdalene’s blue robe that had originally been painted
with a mixture containing indigo and gypsum, Table 1. This
paint was applied directly on top of the older degraded oil
and conifer resin varnish from the previous treatment and
concealed areas of original paint and earlier restoration.
Unlike the earlier repairs, no attempt was made to recreate
32
the lost parts, using instead simple blocks of colour to fill
the losses. A further, thinner, layer of oil and mastic resin
varnish was then brushed over the whole surface.
CONSERVATION TREATMENT
Condition before conservation
As a result of its history of construction and change, the
painting was in urgent need of conservation, both to stabilize it and to clarify the heavily restored and damaged
image. The icon was structurally unstable and visually
compromised by the previous restoration treatments. The
THE NOLI ME TANGERE: STUDY AND CONSERVATION OF A CRETAN ICON
two oil and resin varnish layers from previous restorations were brittle and flaking and original paint was often
attached to these flakes. There were localized areas of recent
paint loss, some showing cleavage between the paint layers
while others included the ground and exposed the wood
support. Two rigid paint blisters in the centre were raised
out of plane and hollow to the touch. The later restoration
was particularly unstable along the left edge and was separating from the wood.
Insect damage had caused areas of the original surface
to become soft and hollow. These had collapsed in places,
causing surface undulations and loss. The painted layers
were riddled with insect flight holes, many of which had
been plugged with restoration material. X-radiography
revealed the extent, depth and severity of a crack running
vertically through the centre of the panel; this was particularly evident in the electron emission radiograph, which
also provided a very clear view of other panel damage
and the restoration of the painted surface. The panel had
remained relatively flat with only a minor twist from top to
bottom, although shrinkage and expansion could readily be
observed in the behaviour of one of the paint blisters in the
upper centre right, which altered in height in response to
changes in relative humidity (RH).
Treatment
Conservation treatment was undertaken to render the icon
structurally stable and improve the clarity of the image.
The outermost oil and mastic resin varnish was removed
with a solvent mixture comprising propan-2-ol and white
spirits in a ratio of 2:3 (v/v), applied on small cotton wool
swabs rolled across the surface; this revealed the lower,
older varnish and the most recent restoration. Where these
restorations extended over the original surface they were
removed mechanically with a scalpel. The lower varnish
was then removed with propanone (acetone) applied in the
manner described above, uncovering the remains of a relatively insoluble coating that had been applied directly onto
the painted surface and which was, therefore, interpreted as
an original material.
The remainder of the most recent restoration and associated fills, along the left edge and where present on the
other edges, were removed mechanically using a scalpel,
revealing the wood beneath and exposing the insect
damage. The remains of a cloth layer between the wood
and ground were discovered at this point; its presence was
not previously known, as it had not been revealed by any
of the imaging methods applied to the painting. Traces of
earlier restoration were also uncovered along the left edge,
matching those present on the right, including the step in
level. It appears that prior to the later restoration the panel
had lost most of its previously restored left edge, together
with more of the original paint at the bottom and top left
corners, and that this may be the reason for the later treatment to repair the left side.
The earlier restorations were left in place except where
they covered original material and obscured the original
surface, where they were removed mechanically. The
regilding and the added black inscription between the
mountains in the background were left, as neither the
extent of any surviving original gilding nor the presence of
an earlier inscription could be determined, Figure 7.
Localized areas of flaking original paint were consolidated with an acrylic dispersion (Lascaux® 4176) applied
beneath the lifting flake with a small brush. The treated area
was then warmed with a heated spatula to approximately
40°C through release layers of lens tissue (closest to the paint
layer) and lightweight Melinex® polyester film to relax and
reattach the paint. The area was weighted with sandbags until
the adhesive dried. Localized areas of cleavage between the
ground and the cloth or wood were reattached by injecting
a warm solution of gelatine (approximately 10% w/v in
de-ionized water) through losses in the original surface with
a small syringe and then weighting with sandbags placed
over release layers as described above. The large rigid blister
discussed earlier was not treated, as the long-term stability of
this reattachment could not be guaranteed unless the icon is
henceforth stored permanently at a raised RH to prevent any
panel shrinkage. Were the panel to be subjected to a period
of low or fluctuating RH there would be increased risk of the
blister lifting again or, more worryingly, of the paint layers
becoming compressed, resulting in active flaking and loss.
For this reason the blister, which was considered stiff enough
to support itself, was not further treated.
The reverse of the panel was cleaned with ‘smoke sponge’,
a vulcanized rubber molecular trap that contains a minuscule percentage of a mild soap (<0.006%, or 0.06 grammes
per kilogramme).
Where access was possible, voids and soft areas of the
original painted surface (comprising paint and ground
layers), particularly around clusters of flight holes and at the
edges of paint losses, were strengthened with a 5% w/v solution of Paraloid® B72 (methyl acrylate/ethyl methacrylate
copolymer) in a 1:1 v/v mixture of acetone and industrial
methylated spirit (IMS). This was injected beneath the
figure 8. A detail of the upper left side of the icon during conservation showing the condition of the original surface after cleaning and
the newly applied white surface fills before inpainting
33
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
figure 9. Icon of the Noli me Tangere (1994,0501.3) after conservation
painted surface and through flight holes to consolidate the
wood and frass. Where access was possible to voids directly
beneath the original paint, these were filled by injecting a
suspension of fine silica microballoons in a 10% solution of
Paraloid B72 in the same solvent mixture through holes in
the surface.
In order to unify damaged and undamaged elements of
the image, new fills were also applied to losses in the original painted image and used to plug selected flight holes,
Figure 8. The original paint layer around the area to be
filled was first protected with a (temporary) brushed application of a ketone-based picture varnish. A filler comprising
chalk in a 10% w/v solution of gelatine in de-ionized water
was applied to the losses and, once dry, was carved with a
scalpel blade and smoothed with cotton wool swabs dampened with de-ionized water to mimic the original surface
34
topography. The temporary varnish was then removed with
white spirits. The larger losses along the edges were not
replaced. The whole painted surface of the icon was then
coated with a brushed application of a 12% w/v solution
of Paraloid B72 in dimethylbenzene (xylene) and the fills
were inpainted with raw pigments ground and mixed in a
20% w/v solution of Paraloid B72 in methoxypropan-2-ol.
Selected areas of wear, considered to be detrimental to the
integrity of the image, were also inpainted. A final thin coat
of a 10% w/v solution of Paraloid B72 in xylene was sprayed
onto the surface to reduce the gloss and to give a more even
appearance across the surface of the painting, Figure 9. As
the icon is generally displayed within a conditioned case,
only a thin protective coating, sufficient to saturate the
colours adequately and provide some protection against
dirt, was used.
THE NOLI ME TANGERE: STUDY AND CONSERVATION OF A CRETAN ICON
FINDINGS IN A HISTORICAL CONTEXT
While the need to conserve the painting was the primary
driving force in this investigation, the results have broader
implications, particularly with regard to provenance and
authorship.
The icon is recorded as having been purchased in
Crete and the findings described here certainly accord
with it having originated in this region. The subject of the
Noli me Tangere was common in Cretan painting from
the mid-fifteenth century onwards [2; p. 407 No. 50, 10;
p. 44 No. 370, 11; p. 187 No. 58, 12; p. 92 No. 19]. The
style of this icon is similar to other works from this period
that show Cretan or Venetian influences. The island of
Crete came under Venetian domination in 1211 and by
the mid-fifteenth century its hybrid society, consisting of
native Greek Orthodox Cretans and Catholic Venetians,
was experiencing widespread and fertile cross-cultural
interactions [13; pp. 194–217, 14; pp. 351–370]. The target
audience for icons such as this Noli me Tangere was both
Orthodox and Catholic and the image combines subjects
favoured by both Christian traditions, the resurrected
Christ being the focal point of the Greek Orthodox faith
and the penitent Mary Magdalene of great importance in
Catholic theology.
Analysis confirmed that the construction of the icon
seems to follow post-Byzantine Orthodox painting practice
as established in the literature [5, 8; pp. 54–56], employing
techniques used in earlier Italian panel paintings and
documented in painters’ manuals [7; pp. 152–192, 15;
pp. 11–48, 16]. Cypress wood was the traditional choice for
panel painting in Crete [5; p. 26], as it was widely available
and its inherent qualities of strength, hardness, fine grain,
resistance to splitting or warping and ease of cutting and
carving were ideal for this purpose. Furthermore, as cypress
retains its fragrance, it can be resistant to some insects [17].
The method of attachment of the horizontal battens on the
reverse of the icon, with nails too short to penetrate the
thickness of the structure, is more unusual. Longer nails
were more commonly used, with the exposed nail points
clinched (bent) at right angles into the wood to prevent the
batten easing off should the panel warp or develop a twist
[4; p. 34, 5; p. 26, 18; pp. 122–125].
The icon also shows great similarity with a considerably
larger depiction of the same subject in Saint Catherine’s
Monastery at Herakleion, Crete, which bears the signature
of the post-Byzantine Cretan artist Michael Damaskinos
(1530/1535–1592/1593) [2; p. 458 No.100]. It is possible
that the British Museum icon could be based on a scaleddown anthibolon of the painting by Damaskinos. Anthibola
were commonly used by Cretan painters to create quick and
efficient reproductions [19; p. 181]. Changing the scale of
a cartoon was certainly a common practice in Renaissance
Italy, the most common method being the use of a ‘proportional squaring grid’ [6; pp. 51 and 131], and it may be that
similar processes were applied to anthibola. Such were the
reputation of Damaskinos and the quality of his works that
his anthibola were in great demand after his death [20;
pp. 255–256 and 269–271]. His icons were copied by known
Cretan artists in the seventeenth and eighteenth centuries
[2; pp. 453, 457, 465 No. 107, colour plate 107], although it
should be acknowledged that stylistically they often differ
substantially from the originals [1, 2; p. 465 No. 107 and
pp. 455–457 No. 99]. As discussed above, investigations of
the underdrawing revealed omissions to the incised image
in the background, angels’ wings and faces of the figures that
could explain slight variations in position and appearance
between this icon and the painting by Michael Damaskinos.
Analysis of the original paint media was not conclusive
and neither proteins nor fats were identified, Table 1. In
traditional practice, egg is generally described as being used
in this context, either alone or mixed with oil in the form
of a tempera grassa medium [9; pp. 202–203 and 234, 16;
p. 11]. But as described earlier, the identification of proteins
in paintings such as this is difficult, particularly if they are
only present in small proportions or if an interaction has
occurred between pigment and binder [21]. Additionally,
research is increasingly demonstrating the complexities
of the use and identification of the binding media in early
paintings [22]. Despite these caveats it is worth noting that
the appearance of the paint surface is certainly consistent
with egg tempera painting practice.
The use of tempera grassa was previously indicated in
a group of late fifteenth-century Florentine panels [23;
p. 30]. It has been suggested that it was sometimes employed
by the fifteenth-century Cretan artist Angelos, as it was
detected in two of six panels analysed [5; p. 40], and later by
Damaskinos [24; p. 187]. The addition of drying oil would
certainly have enhanced the glazing effect of the indigo and
red lake in the composition. Darkening of the paint layer as
a result of ageing of the oil would also explain the darkened
look of the green trees and perhaps the change in appearance of the red lake glaze on the Virgin’s robe.3
The pigments identified, including the organic lakes,
are consistent with those found on Greek and Cretan icons
from the fifteenth century onwards [5; pp. 40–64, 25], and
would have been available as a consequence of vigorous
trade routes with Venice [5; p. 94 Note 43, 7; p. 183, 26;
p. 247]. The technique used to paint the Virgin’s robe is
thought to date from after the mid-fifteenth century and is
distinct from earlier practice employed for Virgins’ robes
[5; pp. 54, 92, Note 40 and 130]. Unfortunately neither the
palette nor the techniques employed are distinctive enough
to give a definitive date for the production of the British
Museum icon, although a seventeenth-century origin seems
most likely [1].
The icon certainly seems to have been used within a
liturgical context. Wax spots were identified on the original
surface above the earliest confirmed varnish and are interpreted as accretions accumulated during use. In Orthodox
practice icons are venerated through the act of Proskynesis
[27; p. 8], in which they are kissed, handled and exposed to
candles and incense [9; p. 48, 28; p. 173, 29; pp. 38–39]. From
the time of the earliest restoration, and predating its purchase
35
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
in the late nineteenth century, it is likely that repairs were
made to facilitate continued use in liturgy. The later restoration to complete the image and apply a coat of mastic varnish
was probably undertaken at the National Gallery in 1924 [30].
CONCLUSIONS
Using a range of analytical and conservation approaches
it has been possible to identify the original materials and
methods of construction of the icon depicting the Noli me
Tangere, together with the materials used in later conservation interventions. The interpretation of these findings was
only possible in combination with research into the history
of the production and use of icons and previous restoration
practices. Research has also helped establish a likely chronology for previous interventions. The latest conservation
treatment has rendered the icon structurally stable and to a
certain extent reintroduced clarity to the image, revealing
the fine quality of the painting technique. The icon is now
included in a permanent display at the British Museum
alongside other Cretan icons, helping to highlight the existence of this important part of the collection and stimulate
interest in its study.
Technical examination has shown that the icon is
complex and well made, following traditional practice.
Identification of the materials and techniques employed,
together with art historical evidence [1], has helped to
strengthen a Cretan provenance. The painting technique
and, to a certain extent, the panel preparation suggest the
work of a skilled artisan. The character of the underdrawing
suggests that the work is a copy of a prototype, possibly by
the Cretan artist Michael Damaskinos.
This icon, together with the work by Damaskinos in Saint
Catherine’s Monastery at Herakleion, are representative
examples of the style of art created within a cosmopolitan
and multicultural urban environment on Crete, which
contains elements designed to appeal to a hybrid audience
from different Christian traditions.
EXPERIMENTAL APPENDIX
Surface examination under magnification was carried out
using a Leica S8 APO stereomicroscope with an APO ×0.63
WD 100 mm lens. The images were captured with a Leica
DFC320 camera.
Ultraviolet (UV) examination was undertaken using
two UV fluorescent lamps and captured with a Hasselblad
503CW camera fitted with a PhaseOne H25 digital back
and an 80 mm Carl Zeiss lens fitted with an UV-absorbing
filter. The exposure times were generally long (in the range
of c.5–10 minutes).
X-radiographs were produced using typical exposure
conditions of 60–70 kV for 25 mA minutes on Kodak
36
Industrex film and then scanned using an Agfa RadView
digitizer with a 50 μm pixel size and 12-bit resolution to
allow digital manipulation and enhancement of the images.
On some areas, electron emission imaging using a heavily
filtered X-ray beam at 300 kV was carried out [31; p. 101].
IR reflectograms (IRR) were produced using tungsten
Elinchrom 500 lights with an Osiris infrared camera fitted
with an InGaAs sensor and a six-element 150 mm focal
length f/5.6–f/45 lens. A Schott RG830 glass filter with a
cut-on transmission of 50% at 830 nm was placed in front
of the lens.
For wood identification, small (< 1.5 × 1.5 mm) samples
were fractured to expose transverse, radial longitudinal
and tangential longitudinal surfaces for identification
using a Leica Aristomet biological optical microscope.
Reflected light with dark field mode was used at magnifications ranging from ×50 to ×520. Polarized light was
selected as required. Standard techniques of wood identification and terminology were used as set out by the
International Association of Wood Anatomists (IAWA)
for the identification of modern wood as exemplified by
Wheeler et al. [32, 33].
Paint cross-section samples were mounted in clear
casting AM polyester resin and dry ground using MicroMesh® abrasive and polishing cloths to avoid disturbing
any water-soluble layers. The samples were examined
under reflected visible and UV light at magnifications of
×400 and ×600. Any inorganic materials present were identified using Raman spectroscopy and SEM-EDX analysis.
Raman spectroscopy of dispersed samples and mounted
cross-sections was carried out using a Horiba Infinity spectrometer with green (532 nm) and near infrared (785 nm)
lasers, with a maximum power of 4 mW at the sample.
Samples for SEM-EDX analysis were carbon coated and
then examined in a JEOL JSM-840, equipped with an EDX
accessory (Oxford Instruments, ISIS with Si(Li) detector)
for elemental analysis. Sequential microchemical staining
tests were also carried out to indicate types of media. Amido
Black AB2A was used as a general stain for protein and
Rhodamine B as a general stain for oil [34, 35]. The crosssections were examined under magnification as above.
Samples for GC-MS analysis were collected by swabbing
or from surface scrapes. Methods of sample preparation
and analysis were selected according to sample type. A
lipid method was used principally for varnishes and the
characterization of paint media. Samples were extracted
using dichloromethane (DCM) and then derivatized
prior to analysis with bis(trimethylsilyl)trifluoroacetamide (BSTFA) + 1% trimethylchlorosilane (TMCS) to
form trimethylsilyl (TMS) derivatives. A protein method
was used principally for ground layers and glues or adhesives. Samples were prepared as amino acid extracts by
hydrolysation with hydrochloric acid, then derivatized
prior to analysis with N-(tertbutyldimethylsilyl)-N-methyl
trifluoroacetamide (MTBSTFA) + 1% tertbutyldimethyl
silyl chloride (TBDMSC). For further details of the analytical methods, see [3].
THE NOLI ME TANGERE: STUDY AND CONSERVATION OF A CRETAN ICON
ACKNOWLEDGEMENTS
The authors would like to thank Chris Entwistle, curator of the
icon collection at the British Museum; Marika Spring and Hayley
Tomlinson of the National Gallery, London; Aviva Burnstock of the
Courtauld Institute of Art, London; and colleagues from the British
Museum: Catherine Higgitt, Duncan Hook, Kevin Lovelock, Nigel
Meeks, Antony Simpson, Trevor Springett and Giovanni Verri.
MATERIALS AND SUPPLIERS
t
t
t
t
t
t
t
AM Polyester resin: Alec Tiranti Ltd, www.tiranti.co.uk
Lascaux 4176 and Paraloid B-72: AP Fitzpatrick, www.
apfitzpatrick.co.uk
Lens tissue: Falkiners fine papers, www.falkiners.com
Silica microballoons: Conservation by Design Ltd, www.
conservation-by-design.co.uk
Micro-Mesh: Craft Supplies Ltd, The Mill, Millers Dale,
Derbyshire SK17 8SN, UK.
Melinex and smoke sponge: Preservation Equipment Ltd, www.
preservationequipment.com
Solvents: VWR International Ltd, Magna Park, Hunter Boulevard, Lutterworth, Leicestershire LE17 4XN, UK, uk.vwr.com
AUTHORS
Lynne Harrison (lharrison@thebritishmuseum.ac.uk) is a conservator
and Janet Ambers (jambers@thebritishmusem.ac.uk), Rebecca
Stacey (rstacey@thebritishmuseum.ac.uk) and Caroline Cartwright
(ccartwright@thebritishmuseum.ac.uk) are scientists, all in the
Department of Conservation and Scientific Research at the British
Museum. Angeliki Lymberopoulou (a.lymberopoulou@open.ac.uk) is
a lecturer at the Open University.
REFERENCES
1. Lymberopoulou, A., Harrison, L. and Ambers, J., ‘The Noli me
Tangere icon at the British Museum: vision, message and reality’,
in Images of the Byzantine World: visions, messages and meanings. Studies presented to Leslie Brubaker, ed. A. Lymberopoulou,
Ashgate, Farnham (2011) 185–214.
2. Μπορμπουδάκης, M. (Borboudakis, M.) (ed.), Εικόνες της
Κρητικής Τέχνης (Από τον Χάνδακα ως την Μόσχα και την Αγία
Πετρούπολη), University of Crete, Herakleion (1993). [Images of
Cretan art (from Candia to Moscow and St Petersburg)]
3. Harrison, L., Ambers, J., Cartwright, C.R., Stacey, R. and Hook, D.,
Establishing an approach to the care and conservation of Orthodox
icons at the British Museum, Report No. 7449/1 (forthcoming),
www.britishmuseum.org/research/online_research_catalogues.
4. Papadopoulou, A., ‘Traditional wood technology and problems
relating to wooden supports’, in The conservation of late icons, ed.
N. Jolkkonen, A. Martiskainen, P. Martiskainen and H. Nikkanen,
Valamo Art Conservation Institute, Finland (1998) 31–40.
5. Milanou, K., Vourvopoulou, C., Vranopoulou, L. and Kalliga, A.L.,
Icons by the hand of Angelos: the painting method of a fifteenthcentury Cretan painter, Benaki Museum, Athens (2008).
6. Bambach, C.C., Drawing and painting in the Italian Renaissance
workshop: theory and practice, 1300–1600, Cambridge University
Press, Cambridge (1999).
7. Dunkerton, J., Foister, S., Gordon, D. and Penny, N., Giotto to
Durer: early European painting in the National Gallery, National
Gallery Company, London (1991).
8. Bouras, L., ‘Working drawings of painters in Greece after the fall
of Constantinople’, in From Byzantium to El Greco: Greek frescoes
and icons, ed. M. Acheimastou-Potamianou, Greek Ministry of
Culture and Byzantine Museum of Athens, Athens (1987) 54–56.
9. Sendler, S.J.E., The icon: image of the invisible, 2nd edn, translated
by S. Bingham, Oakwood Publications, California (1988).
10. Chatzidakis, N., Icons of Cretan School (15th–16th century),
Benaki Museum, Athens (1983).
11. Acheimastou-Potamianou, M. (ed.), From Byzantium to El Greco:
Greek frescoes and icons, Greek Ministry of Culture and Byzantine
Museum of Athens, Athens (1987).
12. Chatzidakis, N., Venetiae quasi alterum Byzantium: Candia to
Venice. Greek icons in Italy, 15th–16th centuries, Foundation for
Hellenic Culture, Athens (1993).
13. Lymberopoulou, A., The Church of the Archangel Michael at Kavalariana: art and society on fourteenth-century Venetian-dominated
Crete, Pindar Press, London (2006).
14. Lymberopoulou, A., ‘Late and post-Byzantine art under Venetian rule: frescoes versus icons and Crete in the middle’, in
A companion to Byzantium, ed. L. James, Wiley, Oxford (2010).
15. Bomford, D., Dunkerton, J., Gordon, D. and Roy, A., Art in the
making: Italian painting before 1400, National Gallery Company,
London (1989).
16. Hetherington, P. (tr.), The ‘painter’s manual’ of Dionysius of
Fourna: an English translation with commentary, of Cod. Gr. 708
in the Saltykov-Shchedrin State Public Library, Leningrad, 2nd edn,
Oakwood Publications, California (1989).
17. Charas, C., Revolon, C., Feinberg, M. and Ducauze, C., ‘Preference
of certain Scolytidae for different conifers’, Journal of Chemical
Ecology 8 (1982) 1093–1109.
18. Uzielli, L., ‘Historical overview of panel-making techniques in
central Italy’, in The structural conservation of panel paintings, ed.
K. Dardes and A. Rothe, Getty Trust Publications, Los Angeles
(1998) 110–135.
19. Lymberopoulou, A., ‘The painter Angelos and post-Byzantine Art’,
in Locating Renaissance art, ed. C.M. Richardson, Yale University
Press, New Haven (2007) 174–210.
20. Κωνσταντουδάκη-Κιτρομηλίδου, M. (Constantoudaki-Kitromilidou, M.), ‘Παραγγελίες Πινάκων, Εργαστήριο, Κυκλοφορία
Σχεδίων του Μιχαήλ Δαμασκηνού στο Χάνδακα: Ανέκδοτα
Έγγραφα (1585–1593)’, Θησαυρίσματα/Thesaurismata 34 (2004)
253–272. [Commissions for paintings, workshops and the circulation of drawings by Michael Damaskinos in Candia: unpublished
documents (1585–1593)]
21. Spring, M. and Higgitt, C., ‘Analyses reconsidered: the importance
of the pigment content of paint in the interpretation of the results
of examination of binding media’, in Medieval painting in Northern
Europe: techniques, analysis, art history, ed. J. Nadolny with
K. Kollandsrud, M.-L. Sauerberg and T. Frøysaker, Archetype
Publications, London (2006) 223–229.
22. Higgitt, C. and White, R., ‘Analysis of paint media: new studies of
Italian paintings of the fifteenth and sixteenth centuries’, National
Gallery Technical Bulletin 26 (2005) 88–104.
23. Dunkerton, J. and Roy, A., ‘The materials of a group of late
fifteenth-century Florentine panel paintings’, National Gallery
Technical Bulletin 17 (1996) 20–31.
24. Moshos, T., ‘The conservation work on the Michael Damaskinos
icons in the Saint Catherine Sinaiton collection in Heraklion’, in
The conservation of late icons, ed. N. Jolkkonen, A. Martiskainen,
P. Martiskainen and H. Nikkanen, Valamo Art Conservation
Institute, Finland (1998) 187–192.
25. Milanou, K., ‘The techniques of post-Byzantine icons of the 15th
century: observations on works of the Benaki Museum collection’, in Changes in post–Byzantine icon painting techniques, ed.
N. Jolkkonen and H. Nikkanen, Valamo Art Conservation Institute, Finland (1999) 5–6.
26. Mathew, L. and Berrie, B., ‘“Memoria de colori che bisognino torre
a vinetia”: Venice as a centre for the purchase of painters’ colours’,
37
LYNNE HARRISON, JANET AMBERS, REBECCA STACEY ET AL.
27.
28.
29.
30.
31.
32.
33.
38
in Trade in artists’ materials: materials, markets and commerce in
Europe to 1700, ed. J. Kirby, S. Nash and J. Cannon, Archetype
Publications, London (2010) 245–252.
Weitzmann, K., The icon, Evans Brothers, London (1982).
Lymberopoulou, A., ‘Audiences and markets for Cretan icons’,
in Viewing Renaissance art, ed. C.M. Richardson, K.M. Woods
and A. Lymberopoulou, Yale University Press, New Haven (2007)
171–206.
Tarasov, O., Icon and devotion: sacred spaces in imperial Russia,
translated by R.M. Gullard, Reaktion Books, London (2002).
Tomlinson, H., National Gallery, London, personal communication (January 2011).
Daniels, V. and Lang, J., ‘X-rays and paper’, in Radiography of
cultural material, 2nd edn, ed. J. Lang and A. Middleton, Elsevier,
Oxford (2004) 101–103.
Wheeler, E.A., Pearson, R.G., La Pasha, C.A., Zack, T. and Hatley,
W., Computer-aided wood identification, Bulletin 474, North Carolina Agricultural Research Service, Raleigh (1986).
Wheeler, E.A., Baas, P. and Gasson, P.E. (ed.), ‘IAWA list of microscopic features for hardwood identification’, IAWA Bulletin 10(3)
(1989) 219–332.
34. Martin, E., ‘Some improvements in techniques of analysis of paint
media’, Studies in Conservation 22 (1977) 63–67.
35. Wolbers, R.C. and Landrey, G., ‘The use of direct reactive fluorescent dyes for the characterization of binding media in cross
sectional examinations’, in 15th Annual Meeting, American Institute for Conservation, American Institute of Conservation, Washington (1987) 168–202.
NOTES
1. The term ‘ground’ is used here to describe the smooth white preparation layer between the cloth and the painted and gilded layers.
2. It was not possible to characterize this mordant fully because
of sampling difficulties; for further discussion see [3; Section
3.3.1.2].
3. It should be noted that the oil and resin mixture found may not
be intentional and could be the result of the accidental migration
of surface varnish layers as a consequence of traditional cleaning
and repair treatments [3; Section 4.1].
Download