Sie sind auf Seite 1von 12

Animal glues: a review of their key properties relevant to conservation

Nanke C. Schellmann
Abstract

in changing ambient environment, and ageing


characteristics.

Collagen-based animal glues are widely used in


the conservation of artefacts, serving as adhesives,
binders and consolidants for organic and inorganic
materials. With a variety of different animal glues on
the market, such as hide and bone glues, fish glues,
isinglass and gelatin, their individual properties
need to be well understood in order to choose a
glue fit for a specific purpose. This paper reviews
a wide range of publications on currently available
animal glues, with respect to their specific physical,
chemical and mechanical properties.

Types of commercially available animal glue


Hide glues are primarily derived from bovine skins
and those of smaller mammals, although connective
tissue may also be used. Bone glues are predominantly
prepared from fresh (green) bones or sometimes
extracted bones (degreased and demineralised, known
as ossein) from cattle and pigs. Hide and bone glues are
produced and sold as coarse powders, pearls, cubes,
and cakes or plates, though the latter two appear to be
increasingly rare [9]. Commercial gelatin, the purified
active ingredient of any collagen-derived glue (pure
denatured collagen), may be obtained from either skin
or bone sources [9, 10] and is supplied in the form of
thin sheets, plates or powder.

Introduction
Animal glues are natural polymers derived from
mammalian or fish collagen the major structural
protein constituent of skins, connective tissue, cartilage
and bones. These glues may exhibit varied physical,
chemical and mechanical properties depending on their
origin and method of preparation. In the manufacture of
objects and artefacts, an extensive traditional knowledge
exists on which animal glues are most suitable for
specific purposes. However, conservators sometimes
lack the confidence to make informed choices between
the different collagen-based glues available when
conserving objects.

As the name suggests, rabbit skin glues should be


produced purely from rabbit skins [10, 11], though
collagenous waste from various small mammals may
also be used [12]. Some suppliers sell rabbit skin glue
that is mixed with bovine hide glue to alter its properties
[13]. The information on the source, pre-treatment,
or additives provided by suppliers may not always
be reliable, as they may not have been given accurate
information by the manufacturers. It is generally
assumed that most animal glues contain preservatives
of some kind (e.g. sulphur dioxide) [9, 14]. Even rabbit
skin compressed into cubes, a by-product from the felt
industry commercially sold as a raw (and thus usually
thought to be a pure) form of rabbit skin glue [10], has
recently been found to contain preservatives [9]. Some
traditional glues, such as the deer glue used in Japan as
a binder for some inks, are now made from bovine or
porcine gelatin manufactured to match the properties of
the traditional genuine material [15].

The selection and preparation of glues are discussed in


patent descriptions, woodworking and artists manuals,
as well as conservation literature and product details
from suppliers [1, 2]. There is also a large amount of
technical research on the properties of collagen and
gelatin published in scientific journals on polymerand bio-technology, medical science and the food and
brewing industry. However, much of this literature
is not readily accessible to conservators and it can
be ambiguous or contradictory. This paper seeks to
provide a review of the literature and to identify which
properties of glue need to be considered when making
decisions about conservation treatments.

The skins of non-oily types of fish [16, 17], as well as


their bones [10, 12, 18, 19], are used to manufacture fish
glues which are sold in liquid form. The swim bladders
of various species are the source for isinglass [2024],
which is available either in the form of complete dried
bladders or membranes, thin plates or fine strips.
In recent years, fish skin and bone gelatin has also
become available in the food industry as a substitute for
mammalian gelatin [2527].

The applications of collagen-based glue in the


conservation field are diverse, ranging from its use as an
adhesive, consolidant or binding medium for pigments
and filler particles [38]. Generally, the following key
properties need to be considered:

chemical structure and denaturation of the protein


molecules.

gelling properties: gelling temperature (Tgel), gel


strength and setting times.

properties of the glue solution: viscosity, surface


tension and pH.

properties of the dried film: cohesion, adhesion


and final bond strength, mechanical behaviour

A number of industrially manufactured cold liquid


animal glues are available that have modified properties
and a long shelf life. These glues usually contain
additives that alter their natural behaviour, extending
the working time at room temperature, or decreasing
the propensity for biodeterioration and reducing the
dried films sensitivity to moisture. However, the exact
composition of industrially tailored collagen-derived
glues and their overall performance may be difficult
to judge, as manufacturers tend to keep their recipes
55

REVIEWS IN CONSERVATION NUMBER 8

2007

low temperature yields gelatinous matrices containing


protein fractions of long chain length and high
molecular weight (MW) [38, 46]. As a general rule,
gentle processing is appropriate for the hides of young
mammals, as well as all fish skin and swim bladders,
because they are rich in collagen and the collagen is not
so strongly stabilised by the additional chemical bonds
that develop in older mammals. Furthermore, glues that
are derived from fish cleave more easily on extensive
heating than those of mammalian origin owing to their
chemical structure [40, 46]. Conservators should thus
be aware that when preparing a collagen-based solution,
mild procedures should be employed [4, 46]. Preparation
temperatures for collagen-based glues are generally
recommended to be around 5563C. However, there is
little loss of gel strength on heating at high temperatures
(e.g. 8090C), even in the case of isinglass, but only if
the solution is kept at these temperatures for no more
than a few minutes [46, 47].

secret. Furthermore, conservation requirements such


as long-term stability and resolubility are unlikely to
be a priority for commercial manufacturers. Given the
range of additives that may be present in industrial glue
formulations, minimally modified glues represent the
safest option for conservation.

Chemical structure and properties


Chemical structure and denaturation
Collagen consists of long protein molecules composed
of naturally occurring amino acids that are linked in a
specific sequence by covalent peptide bonds. Due to
the spatial conformation of some amino acid groups
(notably proline and hydroxyproline) and the many
ionisable and polar functional groups in the protein
chain, the individual chains form triple-stranded helical
coils that are generally believed to be internally stabilised
by hydrogen-bonding [7, 25, 2933].
Collagen is insoluble in cold water [30, 34] and is
transformed into soluble gelatin by denaturation, a
process of critical importance for the performance
of the resulting glue. This is achieved by hot water
extraction (hydrolytic breakdown) [9, 3439]. Pretreatment (either acidic or basic) is necessary for
most skin and bone collagen, but is not required for
the extraction of isinglass from fish bladders, which
contain less cross-linkage within the collagen. During
extraction, the bonds (predominantly H-bonds) in the
triple-helix structures of the collagen are broken so
that it separates into disordered random coils of single
protein chains, thus completing the transition to gelatin
[40]; in perfect conditions gelatin is pure denatured
collagen. The temperature (Td) at which denaturation
occurs is dependent on the chemical structure of the
proteins in the particular collagen source, notably on
the content of the amino acid derivatives proline (Pro)
and hydroxyproline (Hyp). These are supposed to be
largely responsible for the stabilising H-bonded water
bridges in the triple helix [41] and are present more
abundantly in mammalian collagen than in marine
species [42]. Thus, adult mammalian collagen denatures
at 4041C [31, 33], while isinglass and other fish
collagens denature at lower temperatures. The Td of
fish collagens ranges from approximately 15C for deep
cold water fish (such as cod used for fish glue) [10, 43]
up to 29C for most warm water species [31, 33, 44],
which are the preferred source of isinglass produced for
commercial clarification of alcoholic beverages. There
are also a few tropical fish species that reach Td levels of
up to 36C [31, 44].

Gelation and gelling temperature (Tgel)


Although the process of denaturation, with the loss of
the triple helix arrangement of the protein molecules,
is irreversible, some helical structure can be restored
during gelling and drying. On gelling the single
random protein coils undergo partial rearrangement
(renaturation) back into collagen-like triple helices [7,
26, 46, 48-50]. However, the misalignment of the single
strands means that renaturation causes nodes (junction
zones) involving only part of certain strands. The
remainder of these strands may form further nodes so
that a continuous three-dimensional network structure
emerges. The degree of renaturation is dependent on
the chemical composition (Pro and Hyp content), the
chain length of the molecules (molecular weight, MW),
concentration in solution and temperature [42, 49, 51].
High Pro and Hyp content, high MW, high solution
concentrations and slow drying at a low temperature
promote a high degree of renaturation and the
development of a highly ordered network structure [34,
37, 48, 52]. The number of nodes that are established
by the formation of H-bonds (and probably also by
electrostatic interaction [42]) within and between the
molecules determines gel strength and the rigidity and
elasticity of the glue matrix [7, 46, 51].
The ability to form a rigid gel on cooling, which can be
repeatedly reliquefied by reheating, is one of the unique
properties of collagen-based glues. The temperature at
which gelation of the glue solution occurs (Tgel) depends
mainly on the collagen source, but is also affected by
the degree of protein cleavage. Gelation temperatures
decrease with lower denaturation temperature (Td)
and also with increasing cleavage of the molecules.
Mammalian gelatin gels at around 3035C, and cold
water fish gelatin remains liquid down to around 8C
[14, 43, 53]. However, this temperature will be lowered
if the preparation temperature of the glue is significantly
exceeded.

The process of denaturation is necessary for collagen


to convert to gelatin, which can be used as a glue.
Cleavage of the single protein molecule may also occur
during pre-treatment, extraction and dissolution, and
will significantly affect the properties of the gelatinous
glue. The more vigorous the extraction process (i.e.
the more extreme the pH, the longer the treatment
and the higher the temperature during extraction), the
more bonds within the protein molecule are randomly
cleaved, leading to ever decreasing molecular weights
[34, 37, 39, 45]. Mild extraction at moderate pH and

Gel strength
Gel strength is a measure of the gel rigidity of gelatinous
glues, and is strongly influenced by the molecular weight
56

ANIMAL GLUES: A REVIEW OF THEIR KEY PROPERTIES RELEVANT TO CONSERVATION

than alkaline pre-treated collagen derivatives (type B


gelatins), whose MW distribution is skewed towards
lower MW fractions [9, 34, 46].

of the constituent proteins [34, 54]. According to


several authors [35, 39], the average molecular weight
(AMW) of animal glues can range from around 20000 to
250000 g.mol-1. It is thought that permanent gelling
does not take place below an AMW of 20000 g.mol-1
[38, p. 43]. Isinglass from sturgeon, if prepared under
mild conditions, reaches average molecular weight values
of well over 150000 g.mol-1 [4, 33, 46], while liquid fish
glue has AMW values of around 60000 g.mol-1 [10, 14],
placing it at the lower end of the range. For most other
commercial collagen-based adhesives, information on
AMW is not readily available.

Open (gelling) time, tack and drying


The setting time of animal glues depends primarily on
Tgel and gel strength. The lower the Tgel and gel strength,
the longer the open time of the solution (i.e. the longer
it takes for the glue to gel). High Bloom hot hide glues
tend to gel rapidly, as gelation occurs at comparatively
high temperatures [10, 11, 14, 39, 59]. Gelatinous
glues derived from fish, which have low Tgel due to
their chemical structure [42, 43, 58], and cold-set liquid
hide glues are convenient to use when long open times
are required. Commercial fish glues usually contain
preservatives [60] and, sometimes, small amounts of
other additives such as colour brightener, deodorizing
agents or fragrance [10]. Liquid hide glues generally
have further additives to inhibit gelation at room
temperature [17, 28]. These are typically salts (e.g.
urea, thiourea) or phenols that extend the setting time
by inhibiting renaturation of the gelatinous matrix [28,
52]. Some manufacturers claim that their liquid hide
glue does not contain gelling inhibitors [17], in which
case the gelatinous matrix must be considerably affected
by molecular cleavage to achieve the comparatively
low MW that is necessary for the glue to be in a liquid
state.

Characterisation by AMW is only common for fish


glues, which are liquid at room temperature. Most
other gelatinous glues are usually characterised by their
gel strength, as AMW does not describe the molecular
weight distribution and therefore may not always
correlate reliably with the physical and mechanical
properties of a glue [34, p. 60] (Table 1). However, it
would be expected that high AMW adhesives, such as
skin glues, have higher gel strength and viscosity, gel
more rapidly and produce stronger bonds.
Gel strength is strongly influenced by AMW but also
shows a linear correlation with the degree to which
the protein solution renatures during gelation [55],
i.e. the higher the degree of formation of helical
structures, the higher the gel strength. The presence of
salts also influences gel strength, which decreases with
an increasing concentration of ions in solution [42,
56].

The ability of collagen-based glue to develop tack upon


gelation is a unique property. In general, glues of higher
Bloom strength develop tack faster than lower Bloom
glues. The tack strength of glue can be empirically
tested by conservators between two fingertips. Isinglass
solutions may appear to be less tacky than equivalent
concentrations of mammalian gelatin or hide glue, as
they take longer to set at room temperature, since their
lower gelation temperature delays the development of
tack.

Gel strength, also known as Bloom strength, is measured


in grams (g), or Bloom grams (gB), and equals the force
required to make a specified depression into a gel sample
prepared under standard conditions [25, 35, 37, 39].
Manufacturers commonly distinguish between grades
of glue by their Bloom strength, which usually covers
a wide range, being as low as 30 g for weak bone glues
and rigorously extracted hide glues, and up to around
500 g for very strong hide glue [10, 11, 35, 37, 57].
Gelatins derived from tropical fish have significantly
lower Bloom values than mammalian gelatins [58],
since the degree of stabilisation of the triple helix by
H-bonding is lower. Gelatins extracted from cold water
fish do not have specified gel strengths as they are liquid
at room temperature [42].

Drying time generally depends on the ambient


temperature and relative humidity (RH). After gelation,
the glue matrix dries by evaporation of water and this
process can be accelerated by elevating the temperature.
However, collagen-based adhesives should be allowed to
dry as slowly as possible, as a longer period of molecular
mobility after gelation and during drying encourages the
development of highly ordered network structures [52].
This maximises the elasticity and strength (toughness) of
the resulting glue film. Isinglass naturally develops highly
stable and elastic films if dried at room temperature,
being slightly above its Tgel [9].

As gel strength is dependent on the structural


conformation of the gelatinous matrix, it is useful
for estimating the toughness, strength and resilience
of the resulting bond. Furthermore, Bloom strength
also correlates with the water-sorption capacity of the
glue (in gel and solid state), viscosity (at least to some
degree), and gelling temperature (Tgel), which generally
all increase with rising Bloom value. High Bloom glues
require a lower solid content in solution than glues with
a lower Bloom rating to be effective as an adhesive, as
they offer many sites for intermolecular bonding in a
given volume [35, 56]. Mammalian skin glues are usually
considered to have the highest AMW and produce the
strongest gels and films [10], particularly those extracted
by acid pre-treatment. Generally, acid pre-treated glues
(type A gelatins) contain larger fractions of high MW

Properties of gelatinous glue solutions


Viscosity
The viscosity of the glue solution is primarily dependent
on the molecular weight distribution [51]; the greater
the proportion of molecules of higher MW the higher
the viscosity [2, 35]. For a given MW distribution,
the viscosity increases with increasing solution
concentration and decreasing temperature [39, 51, 61].
The degree to which collagen-like helices [62, p. 128]
57

58

96000196000

60000

n.a.

fish gelatin
(from fish
skin, bone and
cartilage)

liquid fish glue

cold liquid
hide glue

n.a. data not available

c.150000 and higher up


to 300000

isinglass (from
fish swim
bladders)

[14, 57]

[26, 45,
54]

[4, 33,
46]

[9, 26,
34, 45,
46, 54]

rabbit skin glue high

110000168000 (type A
gelatin achieves higher
values than type B
gelatin)

[10]

high

hide glue

mammalian
gelatin

[10]

low to medium

bone glue

GLUE TYPES

low to medium

medium to high

medium to high (but can


be produced to achieve
Bloom values as low as
75 gB)

high (up to 500 gB)

high (up to 500 gB, hide


glue pearls produce lower
Bloom values than hide
glue grains)

low to medium (down to


50 gB)

[58]

[9, 10,
14]

[911,
14, 35,
37, 57,
69]

[911,
14, 35,
37, 57,
69]

[Ref.]

decreases with rigorous pretreatment increases with higher MW and


and with excessive/prolonged heating increasing helicity

gel / Bloom strength [gB]

Factors
influencing
property

[Ref.]

molecular weight (MW)

PROPERTY

[Ref.]

medium

low to medium

medium

medium to high

medium to high

high to very high

medium to very
high

low to medium

increases with higher MW,


higher Pro and Hyp content
and increasing solution
concentration

degree of helicity

[Ref.]

pH (approximate values)

[10, 14,
17, 30,
34, 35,
51, 56,
75]

[10, 14,
69]

[4, 22,
46, 65]

[56, 61]

high (4000 mPa.s


at manufactured
concentration)

high (4000
6000 mPa.s at
manufactured
concentration)

[10]

[10, 14,
53]

medium to high (min. [43, 54]


viscosity between pH
79)

highest

medium to high
(type B gelatin
comparatively more
viscous than type A
gelatin)

high (min. viscosity at [9, 10,


around pH 7.0 9.0) 14]

medium to high (min.


viscosity of alkaline
pretreated glue at
around pH 4.55.5,
and of of acid
pretreated glue at pH
7.0 9.0)

low to medium (min.


viscosity around pH
4.5 5.5)

[10, 43,
54, 58]

[19, 61,
71, 79]

[10, 61]

[911,
61]

[911,
16, 35,
68]

[5, 10,
35, 37,
39]

[Ref.]

6.5

[10]

4.06.0 (higher pH values [10, 14]


may be possible)

3.55.0

6.07.5

5.06.5

5.07.5 (wider variations


are possible)

6.57.4 (wider variations


are possible)

5 7

increases with increasing Bloom, influences the viscosity


dependent on isoelectric point
(pI) and pH

viscosity

Table 1 Comparison of the properties of different glue types. The glues are qualitatively ranked relative to one another for each property, i.e. within each individual column. Numerical data is only
referred to in those cases where information was consistent in the literature

REVIEWS IN CONSERVATION NUMBER 8


2007

high (tensile strength


typically around 39
Mpa)

hide glue

59

high

medium

medium

n.a.

isinglass (from
fish swim
bladders)

fish gelatin
(from fish
skin, bone and
cartilage)

liquid fish glue

cold liquid
hide glue

n.a. data not available

high (low Bloom value


gelatin will achieve lower
mechanical strength)

mammalian
gelatin

rabbit skin glue high, but lower than


other hide glue

low to medium

bone glue

GLUE TYPES

[26, 27,
53]

[26, 27,
53]

[4]

[23, 39,
78]

[76]

increases with increasing content of


helical structures

Factors
influencing
property

[Ref.]

mechanical strength

PROPERTY

Table 1 (contd.)
[Ref.]

n.a.

more elastic than hide


glue (but more brittle)

more elastic than


mammalian gelatin (but
more brittle)

more elastic than hide


glue

less elastic (stiffer) than


gelatin from aquatic
sources

more elastic than hide


glue

less elastic (stiffer) than


bone glue and gelatin
from aquatic sources

more elastic than hide


glue (but more brittle)

[26, 58]

[23, 84]

[26, 58]

increases with increasing molecular


weight, helicity and solution
concentration

elasticity
[Ref.]

n.a

medium

medium to high

very high

medium to high

high

high

medium

[4]

[76, 86]

[6, 28]

increases with increasing


helicity

stress development
in fluctuating RH
resolubility with age

[Ref.]

as stable as hide glue


(after 6 months RH
and temperature
cycling)

less stable than cold


liquid hide glue

n.a.

higher than
mammalian gelatin

less stable than


isinglass

less sensitive to
moisture than hide
glue

[17]

[17]

[79]

[5, 39]

more stable than bone [6, 28]


glue, less sensitive
than cold liquid hide
glue

less stable than hide /


rabbit skin glue

resoluble (after 6 months


RH and temperature
cycling)

resoluble (after 6 months


RH and temperature
cycling)

generally thought to be
resoluble

(contradictory data)

generally thought to be
resoluble

generally thought to be
resoluble

generally thought to be
resoluble

more resoluble than hide


glue

[17]

[17]

[23, 79]

[90]

increases with increasing helicity decreases with lower original solution


concentration

stability in fluctuating [Ref.]


environment

ANIMAL GLUES: A REVIEW OF THEIR KEY PROPERTIES RELEVANT TO CONSERVATION

REVIEWS IN CONSERVATION NUMBER 8

2007

and gelatins contain less than 1% fat because of modern


manufacturing methods [9, 10, 54, 58, 69] and may
require additives to reduce the surface tension.

and intermolecular bonds have developed within the


network (gel/Bloom strength) further contributes to
higher viscosity [63, 64]. Strongly denatured gelatinous
solutions (such as bone glues) or those affected by a
high degree of molecular cleavage will normally have a
comparatively low viscosity. At a given Bloom strength,
alkaline pre-treated (Type B) gelatins are generally more
viscous than acid pre-treated (Type A) gelatins [56]
(Table 1).

Ethanol is commonly added to lower the surface tension


and improve the wetting abilities of collagen-based glues
[21, 70, 71]. In one case beer containing 9% alcohol
was added to fish glue that was used in the conservation
of Boulle-marquetry, and was shown to improve the
wetting properties leading to stronger joints between
the wood and brass components [70]. However, alcohol
may also raise the gelling temperature, speeding up the
gelation and decreasing the time for which the glue
is workable [28, p. 102, 110], and may also promote
swelling of the substrate. Alternatively, surfactants can
be added to lower the surface tension [3, 8, 28, 72,
p. 123].

Viscosity is an important factor in the choice of adhesive


for bonding or consolidation, as it will affect the degree
of penetration into a substrate. If the viscosity is too
low the glue may penetrate too far into the substrate,
leaving a joint starved of adhesive. For consolidation of
porous materials, high viscosity may prevent adequate
penetration and cause stress to develop at the interface
between consolidated and unconsolidated areas.
Unfortunately, the viscosity values for animal glues given
in the literature and by suppliers vary widely and are
not easily compared. Measurements were often taken
under different experimental conditions and at different
degrees of cleavage in the protein molecules [4, 21, 35,
37, 46].

Animal glues can have an undesirable tendency to foam,


developing small air bubbles in the glue matrix which
can disrupt the uniformity of the dried glue film and
weaken bonds [5, 59]. Natural fats or free fatty acids
present in glues play a vital role in reducing foaming [5,
68, 73], although some authors still express some doubt
that there is a direct correlation between fat content
and tendency to foam [9]. Nevertheless, Skans [73,
p. 66] suggested that a natural fat content of above 5%
would inhibit the development of pinholes in gesso for
gilding. Sauer and Aldinger [68] have demonstrated an
unambiguous dependency of the degree of foaming on
fat content, whereas no direct relationship could be
established with surface tension. They also could not
find any influence of protein degradation products
on foaming, while pH was established to have an
inconsistent effect.

Isinglass has a much higher viscosity than hide glue at


an equivalent solution concentration and temperature
(above Tgel), which can be explained by its comparatively
high proportion of high molecular weight fractions
(which, in the following paragraphs, will be referred to
as high Molecular Weight Distribution (MWD)) [4, 22,
46, 65]. This is contrary to what is often stated in the
literature and to the traditional beliefs about the handling
properties of isinglass [20, 21, 61]. However, where low
viscosity values have been obtained for isinglass, it is
likely that the particular preparation procedure of the
glue used for the tests resulted in greater cleavage of
the protein molecules [46]. Despite isinglass having a
large fraction of high MW compounds, its low gelling
temperature compensates for this by allowing more
time for the glue to penetrate porous substrates at room
temperature, therefore improving its penetration ability
in comparison to gelatin and rabbit skin glue of similar
high MW fractions, which will gel faster [8, 66].

pH
For conservation applications, the choice of adhesive
may be dependent on the pH sensitivity of the substrate
[71]. Collagen-based glues can display varying pH
values that are difficult to predict purely on the basis
of the glue type or treatment during manufacture. The
assumption that glues which undergo alkaline pretreatment display a slightly alkaline pH and acid-treated
ones have an acidic pH [39, p. 171] is incorrect. It is
stated in the literature that hide and fish glue solutions
often have a fairly neutral pH in the range of 6.5 to
7.4, although wider variations are possible [911, 16,
35, 68]. In general, bone glues tend to be slightly more
acidic [5, 10, 39], with pH levels between 5 and just
below 7 [35, 37]. Pure gelatins from mammals and fish
range between pH 5.06.5 and 3.55.0 respectively [10,
53, 54, 58, 61]. Isinglass yields solutions with a pH in
the neutral range [19, 61, 71]. Conservators should test
the pH value of the chosen glue before use if sensitivity
of the substrate is of potential concern.

In order to obtain glue solutions of low viscosity, it is


not always advisable to dilute viscous high Bloom glues
excessively. The use of an over-diluted glue may result
in swelling, leaching or staining of the substrate if it is
water sensitive [67]. In such cases, a glue with a lower
gel strength would be preferable.
Surface tension
Slow gelation and lower viscosity promote uniform film
formation as the glue is able to spread evenly, providing
adequate wetting of the surface. Wetting is improved
with a decrease in the surface tension of the glue solution.
Sauer and Aldinger [68] confirm that a decrease in
surface tension of a gelatinous solution is directly linked
to the presence of fats. Free fatty acids and neutral
fats are regarded as particularly effective in reducing
surface tension even in small concentrations. With the
exception of rabbit skin glue (which has comparatively
high fat levels of around 5% [9, 11]), most animal glues

Apart from being a relevant aspect to consider in


conjunction with the sensitivity of the substrate, pH
values also have an influence on the properties of the
glue, as the viscosity increases when the pH of the
solution shifts away from its isoelectric point (pI) [1, 37,
61, 74]. Since proteins and amino acids are amphoteric
in nature (i.e. containing both acidic and basic functional
60

ANIMAL GLUES: A REVIEW OF THEIR KEY PROPERTIES RELEVANT TO CONSERVATION

78]. This is thought to be due to its high fat content


[9, 23].

groups), they have an isoelectric point, which is the


pH at which all positive and negative charges within
the molecule are balanced and the molecule carries no
net electrical charge. If the electrical potential of the
ions is unbalanced, solution viscosity and Tgel increase,
as well as the capacity for water-sorption and swelling
ability, while gel strength decreases [9, 46, 52, 61, 62].
Commercial animal glues extracted by alkaline pretreatment (most hide and bone glues, type B gelatins)
usually have a pI of approximately 4.5 to 5.5, whereas
glues derived from acid pre-treated collagen sources
commonly display pI values of between 7.0 to 9.0 [17,
30, 34, 35, 51, 56, 75]. For practical purposes, this
means that glues having a pH near their pI value (such
as bone glues and type B gelatins) will already be at the
lowest possible viscosity, as opposed to those which
have pH values different from their pI, where to achieve
the lowest possible viscosity the pH would have to be
modified to take it closer to the pI (Table 1). The effect
of the pH of a glue solution on its surface tension is
inconsistent [28, p. 75, 68].

Elasticity, resistance to impact (toughness) and creep


As for many of the other physical properties of gelatinbased glue films, the elasticity and stiffness are greatly
dependent on their MW distribution [63], the degree
to which helical structures reform on gelling and the
intra-/intermolecular bonding [7, 26]. The stiffness of
the glue (elastic modulus, known as Youngs modulus
E, mathematically calculated from the ratio of stress to
strain values) increases with a higher ratio of high MW
fractions, higher solution concentrations and with a
greater renaturation level in the network [26, 45, 55].
Stabilisation of the gel network by increased electrostatic
bonding induced by pH levels above or below the pI
also increases the stiffness [42]. Mammalian gelatin
generally has a higher modulus and therefore greater
stiffness than fish gelatin due to its higher network
stabilisation by intra- and intermolecular bonding [26,
58]. Isinglass is also more elastic than mammalian
gelatin [79].

Mechanical properties of the dried film

The moisture content of an animal glue has an important


effect on the mechanical properties. Under normal
ambient conditions (50% RH and room temperature)
gelatin-based glue films contain 1214% of structural
water bound to the polar groups of the protein
macromolecules [52, p. 654]. This water contributes to
the stabilisation of the helical structures within the glue
and a specific amount of water is needed to maintain
structural stability. Above around 25% moisture content
the glue turns from a glassy to a rubbery state at room
temperature [52]. Excessive dehydration of gelatinous
films below a moisture content of 0.2% leads to the
development of covalent cross-links between the protein
molecules, which ultimately renders the glue insoluble
in water [80, p. 509].

Cohesion, adhesion and bond strength


The cohesive strength of the gelatinous matrix of a glue is
determined by its molecular structure and intermolecular
bonding, as expressed by the Bloom value. To produce an
animal glue film that is as strong as possible in the dried
state, the same rules apply as for obtaining a high gel
strength (i.e. high MW distribution/minimum cleavage
of protein molecules, maximum renaturation/content of
collagen-like triple-helices, high intra-/intermolecular
stabilisation). The cohesion strength of animal glues can
be improved by the addition of a suitable amount of an
alcohol, such as ethanol or glycerine [28, p. 108]. To
achieve strong bonding, chemical adhesion between the
glue and the substrate is as important as high cohesion
within the glue matrix.

In general, gelatinous glue films with a low moisture


content are very brittle regardless of the collagen
source and molecular structure [34, p. 63, 52]. Even
at a normal (1214%) water content, gelatinous films
undergo brittle fracture under impact. Randomly coiled
structures exhibit much lower resistance to impact
(greater brittleness) than helical glue matrices in the glassy
state. Glue recipes often contain additives such as sugar
alcohols (e.g. glycerine, sorbitol) and polysaccharides
(e.g. dextrins) to improve elasticity and toughness
[28, 34, 81, 91]. One traditional recommendation for
achieving elastic and resilient glue films is the addition
of honey [4, 18, 21, 22, 61, 82]. Sugars are hygroscopic
and so stabilise the protein molecules by introducing
additional hydrogen bonds involving water [25, 83],
inducing an increase in gel strength and viscosity.
Although these additives do not actually plasticise the
glue matrix, they are often referred to as plasticisers in
the literature. A high proportion of fat also improves
elasticity, although it simultaneously reduces the gel
strength of the glue and final bond strength [23, 84].
A higher water content or an excess of hygroscopic
additives generates a reduction in the glass transition
temperature of the glue [61, 81], which can promote an
unwanted tendency to creep (elongation with time).

Hide glues generally have greater cohesive strength than


the bone glues with highly cleaved molecules, which
display a lower tensile strength and are much more
brittle (Table 1). The tensile strength of hide glues is
typically around 39 megapascals (MPa) (5700 pounds
per square inch, psi) [76]. Mammalian collagen tends
to yield stronger glues than most aquatic sources, owing
to the reduced number of stabilising inter- and intramolecular bonds in fish collagen [33, 49]. Cold water
fish gelatins in particular have a lower propensity to
reform helical structures due to their small proportion of
the amino acid derivatives, Hyp and Pro, and therefore
show a comparatively low tensile strength of around 22
MPa (3200 psi) [26, 27, 53]. This value is comparable
to the strength of bovine bone gelatin [54].
A high tensile strength similar to that of hide glue
has been reported for mildly prepared isinglass from
sturgeon [4], making it a useful adhesive for bonding
wooden joints. The literature confirms that isinglass
has often been used for structural woodwork in the Far
East [24, 77]. Although rabbit skin glue has a high gel
strength, it has been stated as having lower cohesion
and bonding strength than other hide glues [23, 39,
61

REVIEWS IN CONSERVATION NUMBER 8

2007

for improving the glue films hardness and resistance


to water are the addition of tanning agents, such as
aluminium trisulphate (alum), disodium triborate
(borax), sodium acetate or formaldehyde [9, 28, 35, 91].
These salts remove a certain amount of bound water
from the proteinaceous matrix by covalently bonding
to the hydrophilic sites in the glue, thus inducing the
formation of numerous new cross-links between the
protein molecules.

Stability in ambient environment and sensitivity to


fluctuating levels of moisture and heat
Drying of collagen-derived glue films leads to the
development of high internal stress and tensile forces
within the glue matrix, while increasing humidity
generally causes progressive loss of tension [4, 85]. This
behaviour is dependent on the physical and chemical
structure of the glue. A high degree of collagen-like
triple-helix arrangement in a gelatin film has been
shown to result in a reduced tendency to swell [34, 55],
but is also responsible for increasing stress values due
to stronger cohesion. Isinglass from sturgeon, which
contains a high proportion of helical structures (due to
its high MWD, despite its lower Hyp and Pro content),
develops particularly high stress levels, which it is
suggested are twice as high as in hide glue [4].

Mechanical properties of animal glues used as gap fillers


Although the excessive shrinkage and brittleness of
animal glues at low RH [88] makes them inferior gap
fillers on their own, modification with plasticisers and
bulking agents can alter their properties, improving
their suitability for this application [39, 81]. Hard films
with a minimum tendency to distort can be achieved
by the addition of fillers such as magnesium sulphate
or mineral clays together with sugars and dextrins [28,
35].

If kept under moderate relative humidity conditions over


a long period of time, initial stresses within gelatinous
films relax owing to the absence of covalent cross-links
[86]. Under fluctuating environmental conditions, the
mechanical properties of collagen-based glues are subject
to continual change [85]. Considerable development of
internal stresses will affect the glues elasticity, strength
and physical stability and may lead to significant damage
to the substrates [48, 8587].

The addition of an inert filler dramatically changes


the physical and mechanical performance of animal
glues, depending on the proportion of glue present.
A high pigment concentration significantly reduces
intermolecular bonding within the glue medium [76,
86] and thus impedes dimensional changes of the
matrix in response to relative humidity changes [76]. In
addition, with the lack of chemical adhesion between a
proteinaceous binder and inert filler particles, the glue
is substantially weakened and this leads to low tensile
strength [7]. Therefore high MW glues, with their long
protein strands and ability to develop stabilising Hbonds, are appropriate for fillers and gesso with a high
pigment concentration.

At high RH levels (above 85%) animal glue films undergo


a continuing reformation of helical structures. This
will result in new, higher stresses on subsequent drying
and can lead to severe shrinkage due to contraction of
the glue matrix [48, 85, 86]. Cycling of RH can cause
further strain for rabbit skin glue, non-permanent total
dimensional changes of up to 6% have been reported
as the result of a single RH cycle [76, 86], which are
only partly recoverable. According to Zumbhl [48],
contraction mechanisms compete with plastic relaxation
processes above 65% RH. However, even at high RH
levels plastic relaxation may not sufficiently compensate
for these stresses and continuous cycling further reduces
the ability for stress relaxation [48]. This will result in
permanent shrinkage of the glue matrix (up to 5% for
rabbit skin glue [76]) and, in this case, loss of tension is
only possible by substrate deformation or mechanical
destruction (embrittlement) of the glue film [48].

Ageing characteristics
Whilst substantial research has been published on the
behaviour of collagen-derived glues in a fluctuating
environment, information on the ageing mechanisms
and behaviour on exposure to light seems to be more
limited. According to Michel et al., isinglass from
sturgeon, of all animal glues, best retains its mechanical
properties with thermal and ultraviolet (UV) light ageing
and RH cycling [79, p. 271]. It shows markedly less
change in strength and stiffness than pure mammalian
gelatin. Mammalian gelatin increases in tensile strength
but becomes stiffer and more brittle upon artificial
ageing under UV light, fluctuating RH and temperature.
Isinglass from sturgeon remains much tougher and more
elastic than gelatin [79, p. 274]. It also develops the least
permanent dimensional change, whereas gelatin films
swell or creep slightly during ageing, and other animal
glues show an even more marked effect.

At low RH levels, more randomly coiled gelatinous


structures (such as bone glues), which have comparatively
low tensile strength and low resistance to stress, induce
relaxation at an early stage by developing cracks in the
glue matrix, thereby preventing high stresses on the
substrate. These glues also show a greater tendency to
creep under stress at high RH levels [61, p. 14]. Although
animal glues containing a high degree of helical structure
exhibit comparatively high stress when exposed to
extreme and fluctuating environmental conditions, they
still display greater stability in their strength properties
than more randomly coiled structures. The strength
properties of hot hide glue have been shown to be less
sensitive to fluctuating RH and temperature than those
of cold liquid hide glue [6, 28]. Liquid fish glues are even
less stable than cold liquid hide glues under fluctuating
conditions [17]. It has also been suggested that a high fat
content, such as in rabbit skin glue, accounts for better
stability in moist conditions [5, 39]. Common methods

Resolubility
Collagen-derived glues, unless they have been modified
by the addition of tanning agents which causes them
to become relatively resistant to water, generally swell
readily when exposed to water and redissolve when
heated, even after centuries [23, 39]. Neher [89]
established that the Bloom strengths of hide and rabbit
62

ANIMAL GLUES: A REVIEW OF THEIR KEY PROPERTIES RELEVANT TO CONSERVATION

low refractive index, when compared with mammalian


gelatin, causes the least change in appearance of the
pigments after drying [71, 79].

skin glues are not correlated to their water-resolubility


and that all tested samples were completely and equally
successfully reversible after one month of natural
drying. Wooden joints bonded with fish glue or cold
liquid hide glue have also been shown to be detachable
with water after six months of natural ageing or RH and
temperature cycling [17]. An effect of the tannic acids of
oak wood and walnut on their resolubility could not be
established in this study.

Conclusions
This review of the different types of currently available
animal glue has shown that collagen-derived adhesives
vary in their chemical, physical and mechanical
properties. Being a natural polymer, performance is
partly dependent on the original collagen source, which
determines the glues chemical composition, but is also
strongly affected by the extraction and preparation
procedures. Molecular weight distribution is an
important factor which directly influences the protein
solution viscosity and contributes to gel strength and
Tgel. The degree of stabilisation of the protein matrix
by hydrogen and other chemical bonding is determined
by amino acid composition, preparation procedures
and drying time. This has an even greater impact on
the performance of the glue, and significantly affects its
strength, mechanical behaviour, sensitivity to ambient
environment and stability with age. Changes in pH
and the addition of hygroscopic additives (plasticisers)
and salts can alter many of these properties. However,
manipulation of one individual factor cannot necessarily
be realised without simultaneously changing a whole
range of other properties. As most of the properties are
dependent on each other, selection of the appropriate
glue should be based on a correct balance rather than on
individual properties.

The dependence of water-resolubility on original


solution concentration has been demonstrated for aged
and UV-irradiated hide and bone glues at concentrations
of between 2.5 and 20% [90, p. 302]. This research
showed that the lower the original concentration, the
lower the resolubility of the glue film. Bone glues were
more resoluble than hide glues, supposedly because
of their more pronounced molecular cleavage in the
protein matrix (Table 1).
Przybylo tested isinglass from sturgeon obtained from
different suppliers [23], and found that the source,
origin and preparation temperature have no significant
effect on the resolubility of the glue in water after
natural and artificial ageing, as all the films in the test
series remained resoluble. In contrast, Michel et al. [79]
report that their artificially-aged sturgeon isinglass films
were insoluble in water, even though no significant
molecular changes within the protein were detected.
The contradictory results of these two studies may be
due to different preparation procedures and artificial
ageing conditions, which varied in the type of light
source as well as cycles of exposure time, temperature
and RH.

It has become evident that much important data that


would allow comparison of the properties of the different
types of glues is still missing. Very few gelatinous
glues have been prepared and tested under the same
conditions, and insufficient characterisation of these
glues makes it difficult to draw exact conclusions for a
general glue type, as physical and mechanical properties
can vary substantially. However, a summary of the data
does reveal general qualitative trends that can be used
by conservators to make well-informed decisions on the
suitability of a particular collagen-based glue for a given
application.

Resolubility of animal glues may be reduced in cases


where the protein has come into contact with metal
ions (e.g. metal foils, tools, pigments), or with certain
organic pigments and tannins, either before, during or
even after their application [12, 23, 78]. Resolubility
of collagen-derived glue containing no additives is thus
very much dependent on the environment to which it
has been exposed, rather than being predetermined by
the type of glue. Cold liquid hide and fish glues, the
ingredients of which are often unknown to the supplier
and end user, may already contain additives that promote
cross-linking and, therefore, increase insolubility.

Acknowledgements
The author would like to thank Shayne Rivers, Senior
Furniture Conservator at the Victoria and Albert
Museum, London, and Dr Ambrose C. Taylor, Imperial
College, London, for their ongoing support in discussing
this paper and their valuable advice.

Colour changes on ageing


Hide and bone glues are generally much more strongly
coloured (amber to brown) and less transparent than
gelatin or isinglass because of their higher impurity
content. Higher levels of denaturation and molecular
cleavage also intensify the colour of gelatinous solutions
[47]. This phenomenon may be responsible for the
general observation that the higher the Bloom value, the
less yellow the gelatin [56]. Gelatin and isinglass appear
clear and virtually colourless if dried to thin films, even
though they yield slightly yellow or whitish solutions
[10, 14, 43, 56, 58, 61]. They are also very light fast
and show hardly any discolouration or yellowing with
age [8, 30, 79], which is why they are the only collagenderived glues suitable for pigment consolidation.
Isinglass is particularly popular for this purpose, as its

References

63

Greber, J.M., Lehmann, E., and van der Werth, A., Die
Tierischen Leime. Geschichte Herstellung Untersuchung,
Verwendung und Patentbersicht, Heidelberg (1950).

Willers, H., Herstellung von tierischem Leim und seine


Verwendung im Bereich der Tafel- und Fassmalerei nach
Angaben deutschsprachiger Quellenliteratur des 16. bis Mitte
des 19. Jhds., diploma thesis, Akademie der Bildenden Knste
Stuttgart (1980).

von Reventlow, V., The treatment of gilded objects with


rabbit-skin glue size as consolidating adhesive, in Gilded

REVIEWS IN CONSERVATION NUMBER 8

8
9

10

11

12
13
14
15

16

17

18
19

20

21

22

23

Wood Conservation and History, ed. D. Bigelow, Sound View


Press (1991) 269275.
Luybavskaya, E.A., Investigation of properties of protein
glues, in ICOM-CC 9th Triennial Meeting, Dresden, Preprints,
ed. K. Grimstad, International Council of Museums, Los
Angeles (1990) Vol. 1 4750.
Skans, B., Analysis and properties of old animal glues,
in 7. IADA International Congress of Restorers of Graphic
Art, Uppsala, Sweden, ed. K.J. Palm and M.S. Koch,
Internationale Arbeitsgemeinschaft der Archiv-, Bibliotheksund Grafikrestauratoren (1991) 4350.
Buck, S.L., A study of the properties of commercial liquid
hide glue and traditional hot hide glue in response to changes
in relative humidity and temperature, in AIC Wooden
Artifacts Group Session, American Institute for Conservation,
Richmond, Virginia (1990).
von Endt, D.W., and Baker, M.T., The chemistry of filled
animal glue systems, in Gilded Wood Conservation and
History, ed. D. Bigelow, Sound View Press (1991) 155162.
Nicolaus, K., Handbuch der Gemlderestaurierung,
Knemann Verlag, Kln (n.d.) 230.
Wilde, A., Zur heutigen Herstellung von Glutinleim,
Zeitschrift fr Kunsttechnologie und Konservierung 20(2)
(2006) 379406.
Kremer-Pigmente, Bone glue (63000), hide glue (63010
63020), rabbit skin glue (63025, 63028, 23052), gelatin
(63040), isinglass (63100), Salianski-isinglass 63110,
fish glue (63550), Franklin Hyde Glue (63500 63512),
product data sheets, Kremer Pigmente GmbH & Co. KG,
Aichstetten, Germany (January 2007).
Natural-Pigments, Rabbit skin glue, http://www.
naturalpigments.com/details.asp?PRODUCT_ID=51021RSGLU (accessed 7 February 2007).
Doerner, M., Malmaterial und seine Verwendung im Bilde,
18th edn, Enke Verlag, Stuttgart (1994) 98, 180.
Hellmann, A., Hellmann Leim GmbH, Memmingen,
Germany, personal communication (February 2007).
Deffner and Johann, Leime, in Gesamtkatalog 2007/2008
Deffner & Johann GmbH, Schweinfurt (2006) 186188.
Morita, T., Nikawa traditional production of animal
glue, in Adhesives and Consolidants, Preprints of the
contributions to the Paris Congress, 28 September 1984,
ed. N.S. Brommelle, E.M. Pye, P. Smith and G. Thomson,
International Institute for Conservation, London (1984)
121122.
Walsh, H.C., Fish glue, in Handbook of Adhesives, 4th edn,
ed. I. Skeist, Reinhold Publishing Corp., New York (1965)
126128.
Coerdt, A., Zum Leimen zu gebrauchen Untersuchungen
zu kaltflssigen Glutinleimen Teil 1, Restauro 113(1)
(2007) 3238 and Teil 2, Restauro 113(1) (2007) 191197.
Nicolaus, K., Handbuch der Gemldekunde, DuMont Verlag,
Kln (2003) 226.
Petukhova, T., Potential applications of isinglass adhesive
for paper conservation, The Book and Paper Group Annual
of the American Institute of Conservation 8 (1989).
Foskett, S., An investigation into the properties of isinglass,
Scottish Society for Conservation and Restoration Journal 5
(4) (1994) 1114.
Petukhova, T., and Bonadies, S.D., Sturgeon glue for painting
consolidation in Russia, Journal of the American Institute for
Conservation 32 (1) (1993) 2331.
Habel-Schablitzky, A., Fischblasenleim Geschichte und
Eigenschaften sowie Anwendung in der Holzrestaurierung,
diploma thesis, Fachhochschule Kln (1992).
Przybylo, M., Langzeit-Lslichkeit von Strleim Tatsache
oder Mrchen?, VDR Beitrge zur Erhaltung von Kunst- und
Kulturgut 1 (2006) 117123.

64

2007

24

Lin, S.Y., Fish air-bladders of commercial value in China,


The Hong Kong Naturalist 9 (3) (1939) 108118.

25

Fernndez-Daz, M.D., Montero, P., and Gmez-Guilln,


M.C., Gel properties of collagen from skins of cod (Gadus
morhua) and hake (Merluccius merluccius) and their
modification by their coenhancers magnesium sulphate,
glycerol and transglutaminase, Food Chemistry 74 (2001)
161167.

26

Simon, A., Grohens, Y., Vandanjon, L., Bouseau, P., Balnois,


E., and Levesque, G., A comparative study of the rheological
and structural properties of gelatin gels of mammalian and
fish origin, Macromolecular Symposia 203 (2003) 331
338.

27

Gilsenan, P.M., and Ross-Murphy, S.B., Rheological


characterisation of gelatins from mammalian and marine
sources, Food Hydrocolloids 14 (2000) 191195.

28

Pitzen, C., Die Modifizierung von Glutinleimen


Mglichkeiten der Anpassung an objektspezifische und
verarbeitungstechnische Bedingungen nach Literaturangaben
des 18. bis 19. Jahrhunderts, diploma thesis, Fachbereich
Konservierung
und
Restaurierung,
Mbel/Holz,
Fachhochschule Kln (1991).

29

Woodhead-Galloway, J., Collagen: the Anatomy of a Protein,


London (1980) 2324.

30

Fuchs, R., Pergament Material, Geschichte, Restaurierung,


in Pergament: Geschichte Material Konservierung
Restaurierung, ed. R. Fuchs, C. Meinert and J. Schrempf,
Fachbereich Restaurierung / Konservierung, Fachhochschule
Kln (2001) 1012.

31

Burjanadze, T., Thermodynamic substantiation of waterbridged collagen structure, Biopolymers 32 (8) (1992) 941
949.

32

Holmgren, S.K., Taylor, K.M., and Bretscher, L.E., Code for


collagens stability deciphered, Nature 6677 (392) (1998)
666667.

33

Hickman, D., Sims, T.J., Miles, C.A., Bailey, A.J., de Mari,


M., and Koopmans, M., Isinglass/collagen: denaturation and
functionality, Journal of Biotechnology 79 (2000) 245257.

34

Rose, P.I., Gelatin general properties, in The Theory of the


Photographic Process, 4th edn, ed. T.H. James, Macmillan
Publishing Co., New York (1977) 5167.

35

Hubbard, J.R., Animal glues, in Handbook of Adhesives,


4th edn, ed. I. Skeist, Reinhold Publishing Corp., New York
(1965) 114125.

36

Johns, P., and Courts, A., The relationship between collagen


and gelatin, in Food Science and Technology of Gelatin, ed.
A.G. Ward and A. Courts, Academic Press, London (1977)
137177.

37

Hull, W.Q., and Bangert, W.G., Animal glue a staff-industry


collaborative report, Industrial and Engineering Chemistry
44 (10) (1952) 22752284.

38

von Endt, D.W., Technological modifications of protein


materials and the effect on stability. Protein adhesives, in
Protein Chemistry for Conservators, ed. C.L. Rose and D.W.
von Endt, American Institute for Conservation, Washington
D.C. (1984) 3946.

39

Rivers, S., and Umney, N., Conservation of Furniture,


Butterworths-Heinemann, Oxford/Auckland (2003) 156
173, 442.

40

Privalov, P.L., Stability of proteins, Advances in Protein


Chemistry 33 (1979) 167241.

41

Ramachandran, G.N., Modification of collagen and gelatin


by chemical reagents, in Recent Advances in Gelatin and
Glue Research, ed. G. Stainsby, Pergamon Press, London/
New York (1957) 32.

42

Haug, I.J., Draget, K.I., and Smidsrd, O., Physical and


rheological properties of fish gelatin compared to mammalian
gelatin, Food Hydrocolloids 18 (2004) 203213.

ANIMAL GLUES: A REVIEW OF THEIR KEY PROPERTIES RELEVANT TO CONSERVATION

43

Norland-Products, Fish glue, technical abstract, http://www.


norlandproducts.com/techrpts/fishgelrpt.html (accessed 7
February 2007).

63

Normand, V., Muller, S., Ravey, J.C., and Parker, A., Gelation
kinetics of gelatin: A master curve and network modeling,
Macromolecules 33 (2000) 10631071.

44

Ogawa, M., Moody, M.W., Portier, R.J., Bell, J., Schexnayder,


H.A., and Losso, J.N., Biochemical properties of black
drum and sheepshead seabream skin collagen, Journal of
Agricultural and Food Chemistry 51 (2003) 80888092.

64

Yannas, I.V., and Huang, C., Viscoelastic distinction between


helical and coiled macromolecules, Macromolecules 5 (1)
(1972) 99100.

65

45

Badii, F., and Howell, N.K., Fish gelatin: Structure, gelling


properties and interaction with egg albumen proteins, Food
Hydrocolloids 20 (5) (2006) 630640.

Leuenberger, B.H., Investigation of the viscosity and gelation


properties of different mammalian and fish gelatins, Food
Hydrocolloids 5 (1991) 353361.

66

46

Haupt, T., Zubereitung von Strleim Auswirkungen


der Zubereitungstemperatur und -zeit auf Viskositt,
Gelierverhalten und Molekulargewicht, Zeitschrift fr
Kunsttechnologie und Konservierung 18 (2) (2004) 318
328.

Webb, M., Lacquer Technology and Conservation,


Butterworth-Heinemann, Oxford (2000) 8081, 165.

67

Kato, H., The restoration of urushiware for export


with animal glue and urushi, in Japanese and European
Lacquerware, ed. M. Khlenthal, Mnchen (2000) 8184.

68

Sauer, E., and Aldinger, W., Oberflchenspannung und


Schaumbildung bei Glutinlsungen, Kolloid-Zeitschrift 88
(3) (1939) 329340.

69

Amstel-Products, Product description for bone glue and


technical gelatin/hide glue, http://www.amstelproducts.
nl/technicalProt/Specification/bone_glue.htm; http://www.
amstelproducts.nl/technicalProt/Specification/technical_
gelatin.htm (accessed 13 February 2007).

70

Triboulot, M.-C., Restauration de panneaux dcoratifs


premire partie: Amlioration du collage bois-laiton, ENSTIB
project report, Universit Henri Pointcar Nancy (1998).

71

Fuchs, R., Neue Methoden in der Papierrestaurierung:


Fixierung loser Partikel, in 8th IADA International Congress
of Restorers of Graphic Art, Tbingen, 1923 September 1995,
Internationale Arbeitsgemeinschaft der Archiv-, Bibliotheksund Grafikrestauratoren (1995) 163166.

72

Sandner, I., Bnsche, B., Meier, G., Schraum, H.P., and


Voss, J., Konservierung von Gemlden und Holzskulpturen,
Deutscher Verlag der Wissenschaften, Berlin (1990).

47

Gelita, Dissolution of gelatine, http://www.gelita.com


(accessed 17 February 2007).

48

Zumbhl, S., Proteinische Leime ein vertrauter Werkstoff?


Aspekte zum feuchtephysikalischen Verhalten von Gelatine,
Zeitschrift fr Kunsttechnologie und Konservierung 17 (1)
(2003) 95103.

49

Gmez-Guilln, M.C., Turnay, J., Fernndez-Daz, M.D.,


Ulmo, N., Lizarbe, M. A., and Montero, P., Structural and
physical properties of gelatin extracted from different marine
species: a comparative study, Food Hydrocolloids 16 (2002)
2534.

50

Hrmann, H., and Schlebusch, H., Reversible and irreversible


denaturation of collagen fibres, Biochemistry 10 (6) (1971)
932937.

51

Alleavitch, J., Turner, W.A., and Finch, C.A., Gelatin, in


Ullmanns Encyclopedia of Industrial Chemistry, 5th edn, ed.
B. Elvers, S. Hawkins, M. Ravenscroft, J.F. Rounsaville and
G. Schulz, VCH-Verlag, Weinheim (1989) 307317.

52

Kozlov, P.V., and Burdygina, G.I., The structure and properties


of solid gelatin and the principles of their modification,
Polymer Reviews 24 (1983) 651666.

73

Skans, B., and Michelsen, P., Die Bedeutung von Fett in


Tierleim fr Malzwecke, Maltechnik-Restauro 92 (2) (1986)
6371.

53

Norland-Products, High tack fish glue, product details,


http://www.norlandprod.com/msds/hightackmsd.html
(accessed 7 February 2007).

74

Stainsby, G., Viscosity of dilute gelatin solutions, Nature


169 (4303) (1952) 662663.

75

Fujita, S., Organic Chemistry of Photography, Springer,


Berlin/Heidelberg (2004).

76

Mecklenburg, M.F., Some mechanical and physical properties


of gilding gesso, in Gilded Wood Conservation and History,
ed. D. Bigelow, Sound View Press (1991) 163170.

77

Burgio, L., Rivers, S., Higgitt, C., Spring, M., and Wilson,
M., Spherical copper resinate on Coromandel objects:
analysis and conservation of matt green paint, Studies in
Conservation 52 (2007) 241254.

54

Muyonga, J.H., Cole, C.G.B., and Duodu, K.G., Extraction


and physico-chemical characterisation of Nile perch (Lates
niloticus) skin and bone gelatin, Food Hydrocolloids 18
(2004) 581592.

55

Bigi, A., Panzavolta, S., and Rubini, K., Relationship between


triple-helix content and mechanical properties of gelatin
films, Biomaterials 25 (2004) 56755680.

56

Gelita, GELITA Gelatine fr Speise, Pharma und Foto,


product brochure, Gelita The Gelatine Group, (n.d.).

78

57

Kremer-Pigmente, Natrliche Leime und Aquarellbinder,


in Kremer Pigmente Preisliste 2005/2006, Kremer Pigmente
GmbH & Co. KG, Aichstetten (2005) 5861.

Wehlte, K., Werkstoffe und Techniken


Neubearbeitung, Ravensburg (2000) 55.

79

58

Cheow, C.S., Norizah, M.S., Kyaw, Z.Y., and Howell, N.K.


Preparation and characterisation of gelatins from the
skin of sin croaker (Johnius dussumieri) and shortfin scad
(Decapterus macrosoma), Food Chemistry 101 (2007) 386
391.

Michel, F., Geiger, T., Reichlin, A., and Teoh-Sapkota, G.,


Funori, ein japanisches Festigungsmittel fr matte Malerei,
Zeitschrift fr Kunsttechnologie und Konservierung 16 (2)
(2002) 257275.

80

Yannas, I.V., and Tobolsky, A.V., Cross-linking of gelatine by


dehydration, Nature 215 (1967) 509510.

81

Thornton, J., A brief history and review of the early practice


and materials of gap-filling in the West, Journal of the
American Institute for Conservation 37 (1998) 322.

82

Petukhova, T., A history of fish glue as an artists material:


Applications in paper and parchment artifacts, The Book and
Paper Group Annual of the American Institute of Conservation
19 (2000) 1929.

83

Choi, Y.H., Lim, S.T., and Yoo, B., Measurement of dynamic


rheology during ageing of gelatine-sugar composites,
International Journal of Food Science and Technology 39
(2004) 935.

59

Klinger, J., and Thomas, R., Die Kunst zu Vergolden. Beispiele


Techniken Geschichte, Callwey, Mnchen (1989).

60

Lee Valley, High Tack Fish Glue, material safety data sheet,
Lee Valley Tools Ltd., customerservice@leevalley.com,
Ottawa (January 2007).

61

Haupt, M., Dyer, D., and Hanlan, J., An investigation into


three animal glues, The Conservator 14 (1990) 1016.

62

Courts, A., Citrate-promoted helix formation in gelatin,


Biochemical Journal 83 (1962) 124129.

65

der

Malerei,

REVIEWS IN CONSERVATION NUMBER 8

84

Author

Djagny, K.B., Wang, Z., and Xu, S., Conformational changes


and some functional characteristics of gelatin esterified with
fatty acid, Journal of Agriculture and Food Chemistry 49
(2001) 29872991.

85

Hedley, G., Relative humidity and the stress/strain response


of canvas paintings: uniaxial measurements of naturally aged
samples, Studies in Conservation 33 (1988) 133148.

86

Mecklenburg, M.F., Tumosa, C.S., and Erhardt, D.,


Structural response of painted wood surfaces to changes
in ambient relative humidity, in Painted Wood History
and Conservation, ed. V. Dolge and F.C. Howlett, Getty
Conservation Institute, Los Angeles (1998) 464483.

87

Mecklenburg, M.F., The effects of atmospheric moisture


on the mechanical properties of collagen under equilibrium
conditions, in 16th AIC Annual Meeting, New Orleans, 15
June 1988, American Institute for Conservation, Washington
D.C. (1988) 231244.

88

Grattan, D.W., and Barclay, R.L., A study for gap-fillers


for wooden objects, Studies in Conservation 33 (1988)
7186.

89

Neher, A.L., Investigation into the Reversibility of Distinct


Strengths of Animal Glues and Five Different Methods of
Reversing in Wooden Joints, final year research project, Royal
College of Art and Victoria and Albert Museum (RCA/V&A)
Joint MA Conservation Programme, London (1993).

90

Fiedler, I., and Walch, K., Fluoreszenzunterschiede von


Leimen an Furnieren, in Lacke des Barock und Rokoko/
Baroque and Rococo Lacquers Arbeitsheft 81 des Bayerischen
Landesamtes fr Denkmalpflege, ed. K. Walch and J. Koller,
Mnchen (1997) 297304.

91

Ulmer, R., and Westebbe, P.H., Modifizierte Glutinklebstoffe,


term paper, Studiengang Restaurierung, Kunsttechnologie
und Konservierungswissenschaften, Technische Universitt
Mnchen, 2002.

2007

Nanke Schellmann trained as a violin maker in


Mittenwald (Bavaria) before undertaking several years
of internships in the conservation departments of the
National Gallery (Frames) and the Wallace Collection
in London, the Bavarian National Museum, Munich
and the Germanic National Museum, Nuremberg. In
2003, she received an MA in Furniture Conservation
from the Royal College of Art/Victoria and Albert
Museum (RCA/V&A) Joint Conservation Programme,
London, UK. On finishing, she joined the workshop
of Clemens von Schoeler, Munich as a conservator for
furniture and historic wooden interiors. Since 2005
she has attended additional courses in natural sciences
at the Ludwig-Maximilians-University, Munich and is
currently undertaking a PhD at the University of Fine
Arts Dresden, together with the V&A Mazarin Chest
Project and Imperial College, London, in the field of
oriental lacquer conservation.
Correspondence can be sent to:
Nanke Schellmann
Mazarin Chest Project
Furniture, Textiles and Frames Conservation
Section
Victoria and Albert Museum
South Kensington
London SW7 2RL
UK
Email: n.schellmann@vam.ac.uk

66

Das könnte Ihnen auch gefallen