You are on page 1of 19

Maney Publishing

The Characterization of Metal Soaps


Author(s): Laurianne Robinet and Marie-Claude Corbeil
Source: Studies in Conservation, Vol. 48, No. 1 (2003), pp. 23-40
Published by: Maney Publishing on behalf of the International Institute for Conservation of
Historic and Artistic Works
Stable URL: http://www.jstor.org/stable/1506821 .
Accessed: 02/03/2014 06:17
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .
http://www.jstor.org/page/info/about/policies/terms.jsp
.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of
content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms
of scholarship. For more information about JSTOR, please contact support@jstor.org.
.
Maney Publishing and International Institute for Conservation of Historic and Artistic Works are collaborating
with JSTOR to digitize, preserve and extend access to Studies in Conservation.
http://www.jstor.org
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
23
The Characterization of Metal
Soaps
Laurianne Robinet and Marie-Claude Corbeil
To characterize more
fully
the metal
soaps found
in
paint films
or on metal
surfaces,
several metal
soaps
were
synthesized
and their
X-ray diffraction pattern
and Fourier
transform
infrared (FTIR)
and Raman
spectra
measured. Metal
soaps
were
obtained
from four different fatty
acids
found
in
drying
oils
-
two saturated
(palmitic
and stearic
acids)
and two
unsaturated
(oleic
and linoleic
acids)
-
and
from copper,
zinc and
lead,
three metals that are
typically found
in metal
alloys
and
paint systems.
Data are
reported for
the
following compounds: palmitic acid,
stearic
acid,
oleic
acid,
linoleic
acid,
zinc
palmitate,
zinc
stearate,
zinc
oleate,
zinc
linoleate, copper palmitate, copper stearate, copper oleate,
lead
palmitate,
lead stearate and lead oleate. Features that are characteristic
of specific compounds
were observed.
Soaps
obtained
from different fatty
acids with the same metal ion show
differences,
as do
soaps
obtained with the same
fatty
acid but with
different
metal ions.
Identification
is
key
to
understanding
how and
why
metal
soaps form
on actual
objects,
and this
may
lead to
preventive
measures.
INTRODUCTION
Metal
soaps
are water-insoluble
compounds
contain-
ing
alkaline earth or
heavy
metals combined with
monobasic
carboxylic
acids of 7-22 carbon atoms. The
most
important group
of metal
soaps
consists of driers
that
promote
or accelerate the
drying, curing
or
hardening
of oxidizable
coating
vehicles such as
paints;
metal
soaps
can also be used as
waterproofing agents
on substrates such as
fabric, paper, masonry
and
metals,
and as
grease
and lubricant thickeners
[1, 2].
Metal
soaps
are added
initially
to formulations in
some
applications,
but in other cases
they
form in situ.
They
can result from the reaction of oil used as
protective coating
or lubricant with a metal surface to
which it was
applied,
or from the interaction of a metal
support
or metal leaf with
drying
oil in a
painted
artwork. For
example, copper soaps
formed on the
brass wire
wrapped
around the socket
part
of African
iron
spears
that had been coated with animal fat
[3],
in
paintings
on
copper [4],
and in
paintings
where brass
Received November 2001
leaf had been used instead of
gold
leaf
[5].
Such
uncontrolled,
in situ reactions
may
often be detrimen-
tal to the artifact or work of
art,
as shown in the case
of Benin
copper-alloy objects
where
application
of oil
as a
protective coating
resulted in discoloured and
corroded surfaces
[6, 7].
The
negative impact
of
copper soaps
that formed on the surface of bronzes
was described
by
Burmester and Koller
[8].
Metal
soaps
can also form from the reaction of a
simple
metal
salt,
present
as a drier or
pigment,
with an
organic
medium. While such reaction
may
reinforce
the
physical properties
of the film and can also
play
a
role in the
development
of anti-corrosive
properties,
metal
soaps
can also have detrimental effects on
paint
films
[9].
For
example,
Bell observed loss of
gloss
in
alkyd gloss paints containing
zinc white and found it
to be due to the
appearance
of
protuberances resulting
from the formation of zinc
soaps,
the
hazing
effect
being
worst in the case of
alkyd
resins with
high
saturated and oleic acid ester content
[10].
Bell also
noted a reduction in
opacity
in
paint
films and
suggested
that it could be
explained by
a
layer
of zinc
soaps,
close to the zinc white
particles,
of refractive
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
24 L.ROBINETANDM.-C.CORBEIL
index intermediate between those of the resin and the
pigment.
Similar
protuberances
have been observed in
paint layers
in works of art
[11].
It is often difficult to
identify
metal
soaps, especially
in mixtures. In
theory, gas chromatography
would be
the ideal method.
However,
since
fatty
acid salts are
usually
converted to
methyl
ester derivatives in order
to be
analysed,
it is difficult to
specify
whether
they
were
originally present
as metal
soaps,
as free
fatty
acids,
or as
glycerides.
While metal
soaps
which form
as
degradation products
on the surface of metal
artifacts are often
crystalline, only
one
X-ray powder
diffraction
pattern
has been
reported
to date in the
International Centre for Diffraction Data
(ICDD)
database,
that of zinc stearate
(PDF 5-0079). Fragmen-
tary
information
concerning
the
X-ray
diffraction
patterns
of metal
soaps
is scattered
throughout
the
literature. The most extensive research so far has been
published by
Schrenk,
who
reported X-ray
diffraction
data for some
copper soaps
of
fatty
acids
[6, 7].
The
goal
of the
present study
was to
attempt
to
characterize metal
soaps by
the
techniques
that are
most
commonly
used in conservation science labora-
tories,
specifically powder X-ray
diffraction
(XRD),
Fourier transform infrared
(FTIR) spectroscopy,
and
Raman
spectroscopy. Although
some metal
soaps
are
available
commercially,
it was decided to
synthesize
them to ensure that the
compounds
to be character-
ized would be as
pure
as
possible.
The
study
was
limited to metal
soaps
obtained from four different
fatty
acids found in
drying
oils
-
two saturated
(palmitic
and stearic
acids)
and two unsaturated
(oleic
and linoleic
acids)
-
and from
copper,
zinc and
lead,
three metals that are
typically
found in metal
alloys
and
paint systems.
The
synthesis
of the metal
soaps
was
performed
as
described
previously [12],
and is summarized in the
Appendix. Copper, zinc,
and lead
palmitate
and
stearate were
synthesized.
Not all metal
soaps synthe-
sized from oleic and linoleic
acids,
which are them-
selves
liquids
at room
temperature,
were
solids;
only
the
copper,
zinc,
and lead oleate and zinc linoleate
were obtained as
crystalline
solids.
CHARACTERIZATION
X-ray
diffraction
D-spacings
and intensities are
reported
in Tables 1 to
12. Patterns have been
reported
in the ICDD database
for stearic acid
(ICDD 38-1923)
and
palmitic
acid
(ICDD 24-1853)
but were re-measured to allow
Table 1
X-ray powder
diffraction data for
palmitic acid,
C,~H3202
dobs
(nm)
1/10
(%)
dobs (nm)
1/10
(%)
3.573 100 0.3566 1.9
1.789 8.7 0.3471 0.3
1.1919 30.8 0.3406 0.7
0.8970 2.9 0.3292 0.5
0.7142 6.3 0.3233 0.7
0.5098 2.6 0.3146 0.2
0.4786 0.3 0.3082 0.6
0.4568 1.6 0.2960 2.3
0.4466 0.3 0.2907 0.2
0.4365 2.6 0.2852 0.4
0.4121 40.2 0.2838 0.3
0.3960 0.7 0.2709 0.6
0.3686 15.2
Table 2
X-ray powder
diffraction data for stearic acid,
C,,H3602
d
obs
(nm)
1/10
(%)
dobs
(nm)
1/10
(%)
3.955 100 0.4433 1.1
1.998 22.6 0.4133 2.1
1.3313 78.6 0.3988 1.2
0.9983* 4.8 0.3704 1.0
0.7983 25.5 0.3327* 1.1
0.5698 6.3 0.2849 0.9
0.4988* 0.6
*Peak
overlapping
with
peak from
internal standard
(mica).
Table 3
X-ray powder
diffraction data for zinc
palmitate,
(C,6H310)2Zn
d
obs
(nm)
1/10
(%)
dobs
(nm)
1/10
(%)
3.825 100 0.4044 4.2
1.913 24.1 0.3990 2.2
1.2743 27.8 0.3840 13.3
0.9554 6.2 0.3806 2.8
0.7637 7.2 0.3677 0.8
0.6361 1.3 0.3631 2.3
0.5452 2.3 0.3553 1.6
0.4773 1.8 0.3228 0.5
0.4544 4.3 0.3135 0.7
0.4501 1.4 0.3047 0.3
0.4460 0.5 0.2979 0.3
0.4385 1.8 0.2914 1.5
0.4339 3.6 0.2704 0.4
0.4238 0.7 0.2656 0.9
0.4170 2.5
comparison
with
patterns
of metals
soaps
which were
measured
using
cobalt radiation. After
measuring
the
two
patterns
and
examining
the data
reported
in the
ICDD
database,
it became clear that the
pattern
reported
for stearic acid
probably corresponded
to
that of a mixture of stearic and
palmitic
acid. The X-
ray
diffraction
patterns
of oleic and linoleic acids were
not measured since these
compounds
are
liquids
at
room
temperature. X-ray powder
diffraction data
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 25
Table 4
X-ray powder
diffraction data for zinc stearate,
(C,8H350~2Zn
dobs
(nm) 1/10
(%) dobs
(nm) 1/10
(%)
4.248 100 0.3878 5.2
2.1323 29.9 0.3800 2.3
1.4204 38.2 0.3736 16.8
1.0649 7.9 0.3671 0.8
0.8518 12.0 0.3579 1.4
0.7095 2.3 0.3551 3.2
0.6080 3.8 0.3327* 1.1
0.5319 0.7 0.3271 0.4
0.4726 1.9 0.3193 0.7
0.4544 5.3 0.3135 1.8
0.4493 2.0 0.3026 1.1
0.4396 5.9 0.2906 0.7
0.4260 5.9 0.2868 1.5
0.4099 2.5 0.2721 0.7
0.3920 10.4 0.2700 1.2
*Peak
overlapping
with
peak from
internal standard
(mica).
Table 5
X-ray powder
diffraction data for zinc oleate,
(C,8H330)2Zn
d
obs (nm) 1/10
(%) dobs
(nm)
1/10
(%)
4.266 100 0.3915 1.1
2.1362 25.7 0.3875 1.7
1.4233 47.8 0.3840 2.0
1.0667 7.0 0.3789 0.9
0.8530 14.8 0.3765 0.3
0.7107 2.5 0.3706 0.6
0.6091 4.8 0.3642 2.2
0.5330 0.5 0.3594 0.3
0.4737 2.2 0.3552 0.6
0.4680 0.9 0.3448 0.2
0.4630 1.7 0.3396 0.4
0.4529 1.8 0.3278 3.0
0.4357 3.3 0.3140 0.4
0.4264 4.4 0.3088 0.2
0.4229 1.1 0.3042 0.2
0.4180 2.7 0.2927 0.2
0.4118 0.8 0.2911 0.3
0.4045 4.4 0.2822 0.2
0.3989 1.7
Table 6
X-ray powder
diffraction data for zinc linoleate,
(C,8
H310)2Zn
dobs
(nm) 1/10
(%) dobs
(nm) / (%)
4.213 100 0.4007 5.3
2.107 25.0 0.3957 1.3
1.4056 37.0 0.3815 1.8
1.0533 9.4 0.3747 1.0
0.8418 8.0 0.3686 0.5
0.7011 1.5 0.3606 0.3
0.6009 3.4 0.3540 0.3
0.5260 0.4 0.3406 0.2
0.4763 1.5 0.3287 0.4
continued
Table 6 continued
do,
(nm) 1/10 (%) do1 (nm) I/10 (%)
0.4685 3.3 0.3233 0.4
0.4637 0.7 0.3156 0.3
0.4533 3.2 0.3074 0.3
0.4371 3.7 0.2998 0.5
0.4261 1.5 0.2960 0.2
0.4202 3.3 0.2918 0.3
0.4166 1.7 0.2884 0.2
0.4120 0.7
Table 7
X-ray powder
diffraction data for
copper palmitate,
(C,6H3,O)2Cu
dobs
(nm)
1/10
(%) dob
(nm)
1/10
(%)
4.252 100 0.4660 1.2
2.131 24.2 0.4286 4.9
1.4199 33.1 0.4166 6.7
1.0644 5.5 0.4035 1.4
0.8513 7.0 0.3898 6.7
0.8070 2.5 0.3796 3.7
0.7998 3.9 0.3676 2.7
0.7625 2.9 0.3566 0.3
0.7097 1.9 0.3392 0.2
0.6965 0.6 0.3240 0.3
0.6507 0.9 0.3143 0.4
0.6081 1.4 0.3085 0.4
0.5805 0.9 0.2985 0.3
0.5368 1.0 0.2926 0.3
0.4826 2.4 0.2910 0.2
0.4739 2.3 0.2792 0.7
Table 8
X-ray powder
diffraction data for
copper stearate,
(C8HO350)2Cu
dob
(nm)
1/10
(%)
dobs
(nm)
1/10
(%)
4.753 100 0.4681 0.7
2.375 24.0 0.4454 0.7
1.5827 36.0 0.4270 4.2
1.1859 6.7 0.4233 0.7
0.9484 11.4 0.4165 8.0
0.8054 2.9 0.4002 0.9
0.7981 1.6 0.3951 1.6
0.7907 1.7 0.3878 6.2
0.7692 2.0 0.3785 3.8
0.7580 2.9 0.3674 1.3
0.7250 4.1 0.3601 1.4
0.7113 0.7 0.3548 0.6
0.6769 3.5 0.3411 3.3
0.6604 1.3 0.3257 1.1
0.6209 0.8 0.3138 1.2
0.6080 0.2 0.3031 0.2
0.5919 0.3 0.2986 0.3
0.5693 0.2 0.2916 0.4
0.5591 0.5 0.2879 0.6
0.5264 1.2 0.2732 0.6
0.4816 2.7 0.2698 0.2
0.4763 2.2
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
26 L. ROBINETAND M.-C. CORBEIL
Table 9
X-ray powder
diffraction data for
copper oleate,
(C,8H33O2)2Cu
dobs
(nm)
1/10
(%)
dobs
(nm)
1/10
(%)
4.224 100 0.4777 5.6
2.162 24.8 0.4618 1.1
1.447 36.0 0.4449 9.2
1.085 5.3 0.4393 8.3
0.8808 14.1 0.4341 6.2
0.8662 3.1 0.4123 11.0
0.8484 6.3 0.3987 4.4
0.8059 20.6 0.3922 7.4
0.7344 7.5 0.3760 2.4
0.7197 1.3 0.3592 3.9
0.7026 0.8 0.3481 0.7
0.6674 0.3 0.3399 0.4
0.6155 1.6 0.3287 1.1
0.6018 1.6
Table 10
X-ray powder
diffraction data for lead
palmitate,
(C,6H310)2Pb
d
obs
(nm)
1/10
(%)
dobs
(nm)
I/10 (%)
4.474 100 0.3649 2.4
2.258 56.5 0.3614 0.8
1.5083 52.5 0.3572 0.4
1.1321 21.1 0.3541 1.1
0.9057 19.5 0.3486 2.4
0.7549 7.4 0.3424 0.5
0.6472 8.4 0.3397 1.5
0.5663 3.7 0.3327* 1.5
0.5037 5.1 0.3237 2.7
0.4535 3.3 0.3185 0.5
0.4120 2.0 0.3114 1.7
0.4069 7.1 0.3044 0.2
0.4001 1.4 0.3021 2.7
0.3963 0.6 0.2972 1.7
0.3915 2.1 0.2909 1.3
0.3862 0.7 0.2832 1.7
0.3803 1.3 0.2781 1.5
0.3778 0.9 0.2718 1.6
0.3743 0.4 0.2666 2.6
0.3677 1.7
*Peak
overlapping
with
peakfrom
internal standard
(mica).
have been discussed in detail elsewhere
[12].
Charac-
teristic features will be
briefly
discussed here.
The
patterns
of
palmitic
and stearic acid are
similar,
showing
five or six
sharp peaks corresponding
to
orders of the
long spacing1
followed
by
a cluster
Table 11
X-ray powder
diffraction data for lead stearate,
(C18H350)2Pb
d
obs
(nm)
1/10
(%)
dobs
(nm)
1/10
(%)
5.019 100 0.3604 2.3
2.523 54.8 0.3505 1.3
1.6831 59.0 0.3478 1.6
1.2623 25.5 0.3386 0.8
1.0103 32.9 0.3362 1.5
0.8413 11.1 0.3251 3.9
0.7211 11.9 0.3205 0.7
0.6310 5.5 0.3153 0.6
0.5609 6.5 0.3121 1.4
0.5049 4.4 0.3062 0.4
0.4587 4.2 0.2998 1.5
0.4204 1.3 0.2968 1.0
0.4078 5.8 0.2942 0.6
0.4030 1.7 0.2902 0.4
0.3987 1.6 0.2875 0.9
0.3952 1.9 0.2801 2.4
0.3881 4.9 0.2761 0.8
0.3750 1.1 0.2710 0.8
0.3654 4.9 0.2655 1.6
Table 12
X-ray powder
diffraction data for lead oleate,
(C,8H330)2Pb
d
obs
(nm)
1/10
(%)
dobs
(nm)
1/10
(%)
4.700 100 0.3891 2.6
2.332 28.3 0.3811 1.7
1.5545 36.1 0.3776 1.3
1.1654 12.6 0.3719 0.5
0.9320 13.0 0.3628 3.3
0.7767 6.3 0.3584 1.8
0.6656 6.7 0.3457 1.9
0.5825 3.3 0.3327* 1.9
0.5175 2.6 0.3288 2.7
0.4654 3.7 0.3128 2.4
0.4560 5.4 0.3104 1.9
0.4450 0.8 0.3056 0.8
0.4389 0.3 0.2976 2.7
0.4319 2.0 0.2912 1.4
0.4232 1.6 0.2830 2.1
0.4160 1.6 0.2780 0.9
0.4052 0.5 0.2735 0.5
0.3984 1.3 0.2702 1.2
*Peak
overlapping
with
peak from
internal standard
(mica).
comprising
several close
peaks
between 20 and
30',
20.
The stearic acid
peaks
are shifted to lower
angles
compared
to those of
palmitic acid,
as
expected
for a
compound
with a
longer
carbon chain. Another
difference is the relative
intensity
of the main
peaks
below 20' and the
presence
of intense
peaks
between
20 and
30',
20 in the case of
palmitic
acid.
The four zinc
soaps (zinc palmitate,
stearate,
oleate
and
linoleate)
have similar
patterns,
with seven well-
resolved
peaks
below
20', 20,
the relative
intensity
of
these
peaks being comparable
in all
patterns,
and
'Long-chain compounds
such as
fatty
acids and their
soaps usually
crystallize
with the
long
chains
packed
and
aligned along
the c
axis,
the
longest
axis
ofthe
lattice. The term
'long spacing' refers
to the
interplanar
distance
along
the c
axis;
it does not
correspond
to the exact chain
length
unless the lattice
angle
beta is
equal
to 900.
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 27
5 10 15 20 25 30 35 40
2-i
Theta(
Figure
1 XRD
patterns
of
(a)
zinc
palmitate, (b)
zinc stearate, (c)
zinc oleate and
(d) zinc linoleate from 1 to 400, 20.
including
also a cluster of weak
peaks
between 20 and
300,
20
(Figure 1).
There is a noticeable shift towards
the low
angles
when
going
from C16
palmitate
to one
of the C18
compounds;
the
presence
of one or two
double bonds does not seem to have
any impact
on
the
pattern
in the
region
below
200,
20.
Therefore,
the
patterns
of the C18 zinc
soaps
are
indistinguishable
in
this
region. However,
the
shape (but
not so much the
position)
of the
peak
cluster between 20 and
300,
20
changes considerably
when one or two double bonds
are introduced
(Figure 2).
The same trends are observed in the
copper
and
lead
soaps
series. One
striking
difference, however,
is
that the
peaks
at low
angles
in the oleate
pattern
are
shifted back to
higher angles
with
respect
to the
stearate.
The
patterns
of several of the zinc and
copper
metal
soaps
were found to be
very
similar in the
region
below
200, 20,
with their five orders of
long spacing
(corresponding
to the five most intense
peaks)
at
similar
d-spacings (around 4.2, 2.1, 1.4,
1.1 and
0.9nm)
and relative
intensity
values.
Corresponding peaks
are
found at smaller
d-spacing
values in the case of zinc
palmitate
and at
higher d-spacing
values in the case of
copper
stearate.
Therefore,
zinc
stearate,
oleate and
linoleate can
only
be
distinguished
from each other
and from
copper palmitate
and oleate based on
peaks
of
considerably
lower
intensity
in the
region 20-300,
id
b
A
J
tiI
20 22 24 26 28 30 32 34 36 38 40
2-Theta()
Figure
2
Enlarged
XRD
patterns
of
(a)
zinc
palmitate, (b)
zinc
stearate, (c)
zinc oleate and
(d) zinc linoleate from 20 to 400, 20.
20. Additional low
intensity peaks
are also observed in
the case of
copper palmitate
in the
region
around 12-
180,
20. On the other
hand,
the
patterns
of the lead
soaps
can
easily
be
distinguished
from those of zinc and
copper soaps,
since
they
are
very
different in both d-
spacings
and relative intensities.
The
experimental pattern
obtained for zinc stearate
in the
present study
was found to be different from
that
reported
in the ICDD database
(5-0079).
This is
because low
angle
data were
missing
from the
original
pattern (PDF 5-0079)
and
peaks
that were
reported
in
other
angular regions
consisted of more
peaks
that
were not resolved due to
broadening.
Schrenk
[6] reported
the
X-ray
diffraction
patterns
of
copper
stearate and oleate obtained
commercially.
The
copper
stearate was not
pure, containing
about
25%
copper palmitate.
Since the
X-ray patterns
re-
ported by
Schrenck were measured
using
a Gandolfi
camera,
the most intense
peaks
at low
angle
were
missing.
As a
result,
the relative intensities are
wrong,
the 100
intensity being
attributed to the
peak
with a
d-spacing
of 1.5nm in the case of
copper stearate,
for
example.
As the
X-ray long spacing
is related to the
length
of
the carbon
chain,
it would be
expected
to increase
when
going
from C16 to C18 or when
inserting
metal
atoms of
increasing
radius. Stearic acid has a
larger long
spacing
value
(3.940nm)
than
palmitic
acid
(3.573nm).
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
28 L. ROBINETANDM.-C.CORBEIL
When
soaps
are
formed,
the
long spacing
increases,
with the
largest
value observed in the case of the lead
soaps.
However,
the
copper soaps
do not follow this
trend. Since
copper
atoms have a
slightly
smaller radius
than zinc
atoms,
long spacing
values
slightly
lower than
those observed for the zinc
soaps
would be
expected;
however,
considerably larger
values were observed.
Copper
salts of
carboxylic
acids
adopt
a structure in
which two
copper
atoms are
bridged by
four
carboxy-
late
groups,
as in
copper
acetate
[13].
A similar
structure has been
proposed
for
copper palmitate
and
stearate
[14, 15].
The
copper soaps
are therefore
quite
different from the zinc and lead
soaps,
which would
explain why they
divert from the trend observed for
the latter.
The
long spacing
value does not
change
much
when
going
from saturated C18 stearic acid to unsatu-
rated C18 oleic acid or from zinc stearate to zinc oleate.
However,
it decreases when
going
from the
copper
stearate to the
copper
oleate and when
going
from the
lead stearate to the lead oleate. This is
likely
to be due
to conformational
changes
in the carbon chain
arising
from the double bond. This effect is much more
pronounced
in the case of the
copper
salt,
which
again
may
be due to the
bridged
structure of
copper
carboxylate compounds.
Fourier transform infrared
spectroscopy
The infrared
absorptions
of the
fatty
acids and of the
metal stearates and oleates are listed in Tables 13 and
14. As can be
expected,
the infrared
spectra
of
palmitic
and stearic acids are
very
similar,
as are those of oleic
and linoleic
acids,
and the two
pairs
of
spectra
differ
b
a
.. ............. ...
a v
,,
e
r.
um s7f
Figure
3 IR
spectra
of
palmitic
acid
(a)
and stearic acid
(b)
in the
1400-1100 cm-1
region.
from each other. The main difference between the
infrared
spectra
of
palmitic
and stearic acid is that there
are seven bands between about 1310 and
1185cm-1
in
the
spectrum
of
palmitic acid,
while there are
eight
in
the
spectrum
of stearic acid
(Figure 3).
These
multiple
bands are due to the
CH2 wagging
modes
coupled
with the
carboxyl
vibrations
[16].
The
spectrum
of
oleic acid has a few extra bands in the 1200-1000cm-'
region compared
to that of linoleic acid.
The formation of a
soap
is marked
by
the
disappear-
ance of the
very
broad band attributed to the O-H
stretch in the 3300-2500cm-'
region
and the
replace-
ment of the bands
assigned
to the C=O and
C-O
stretch around 1700 and
1300cm-',
respectively,
with
the bands attributed to the COO-
asymmetric
and
WaeImbr
cm.
C
,
i
Wavenumber (cm-
Figure
4 IR
spectra
of stearic acid
(a), copper
stearate
(b),
zinc stearate
(c)
and lead stearate
(d)
in the 1800-1100 cm-
region.
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 29
Table 13 Infrared
frequencies (cm-)
for
palmitic
and stearic acid and metal stearates
Palmitic acid Stearic acid
Copper
stearate Zinc stearate Lead stearate Attribution*
3300-2500 br 3300-2500 br v OH
2954 m 2954 m 2953 s 2952 s 2955 s
va CH3
2917 vs 2917 vs 2916 vs 2918 vs 2919 vs
va CH2
2871 m 2871 m
vs CH3
2849 vs 2850 vs 2850 vs 2848 vs 2850 vs
vs CH2
2677 m, br 2676 m, br v OH
1703 vs 1703 vs v C=O
1586 vs 1540 vs 1541 s, 1513 vs
va
COO
1509 w
1471 m 1472 s 1473 s 8
CH2
1463 m 1463 m 1468 m 1465 s 1462 s 8
CH2
1448 m 6
CH2
1432 m 1432 m 1441 m 8
CH2
1422 m 1398 s 1420 s
vs
COO
1410 m 1411 m 1407 m 8
CH2
1372 vw 1372 vw 1366 w 8
CH3
1358 w
1348 vw 1347 vw 1340 vw 1347 m
1329 w 1331 w 1334 m
1311 m 1313 m 1316 m 1324 vw 1318 m
1296 m 1298 s 1306 vw 1299 m 6
CH2
1272 m 1279 m 1286 vw 1282 m
1260 m 1266 vw 1263 m
1250 m
1241 m 1247 vw 1244 m
1228 m 1223 m 1227 vw 1225 m
1207 m 1204 m 1207 vw 1206 m
1188 m 1187 m 1187 vw 1187 m
1116 w 1114 vw, sh 1117 w v C-C + 8 C-C-C
1099 w 1103 w 1105 w 1104 w vC-C + 8 C-C-C
1075 vw v C-C + 8 C-C-C
1033 w 1031 vw
942 m 942 m 949 w 8
CH2
929 w
911 w 913 w v C-C
(carboxyl)
891 w 892 w 891 vw 8
CH3+vC-C
876 vw 872 w 879 w
811 vw 807 w 813 w
782 w 780 vw
762 vw 758 w, sh 761 w
745 m 731 m
720 m 719 m 722 m 723 m 719 m
707 m
688 w 688 w 684 w 8 COO
627 w Cu-O
593 vw
579 w
549 w 548 w 550 vw 540 w 8 COo
499 w 505 w 503 vw
483 w 475 vw
440 w 449 w 459 vw 456 vw
428 w 436 w
vs
=
very strong,
s
=
strong,
m
=
medium,
w
=
weak,
vw
=
very weak,
sh
=
shoulder,
br
=
broad,
V
=
stretching,
5
=
bending
*Based on
Yokoyama
et al.
[16]forpalmitic
acid
(IR
and
Raman),
Ahn and Franses
[17]for
stearic acid and lead stearate
(IR),
and Hu et al.
[18]for copperstearate (Raman).
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
30 L. ROBINET AND M.-C. CORBEIL
Table 14 Infrared
frequencies (cm )
for oleic and linoleic acid and metal oleates
Oleic acid Linoleic acid
Copper
oleate Zinc oleate Lead oleate Attribution*
3300-2500 br 3300-2500 br v OH
3006 s 3011 s 3004 m 3002 m 3003 m v CH
2955 sh 2959 sh 2955 s 2955 m, sh 2954 s
va CH3
2926 vs 2927 vs 2920 vs 2922 vs 2919 vs
v. CH2
2855 s 2855 s 2851 vs 2851 s 2851 vs
v. CH2
2676 m 2675 m v OH
1712 vs 1711 vs v C=0
1655 sh 1656 sh v C=C
1589 m
1584 vs 1548 vs, 1526 vs 1505 sh, 1489 vs
Va
COO
1508 w
1465 s 1463 s 1468 m 1467 s, 1455 s 1471 vs
8CH2
1435 sh 1431 sh 1439 m 8
CH2
1426 m 1407 s, 1398 s 1423 sh, 1406 s
vs
COO
1413 s 1413 s 1416 m
1377 m 1376 m
1349 vw 1346 w
1317 w 1320 vw 1312 w
1291 vw 8
CH2
1285 m 1285 m 1281 w 1281 vw 1276 w
1246 sh 1245 sh 1240 vw 1252 vw
1225 sh 1219 sh
1199 vw 1195 w
1162 w 1160 vw
1117 w 1114 w 1119 vw 1114 w v C-C
1089 w 1094 w 1093 w v C-C
1069 vw
1024 vw 1013 vw
940 m 938 m 932 m
896 vw
826 w 827 w
794 vw 795 vw
776 w 766 w 763 vw
746 m
724 m 723 m 721 m 722 m 717 m
691 m 684 w
651 m
622 w
604 vw 606 vw
582 vw
550 vw
511 vw
493 vw
484 vw 481 vw 475 vw
456 vw
*Based on Tandon et al.
[19]for
oleic acid
(Raman).
symmetric
stretch around 1550 and
1400cm-', respec-
tively (Figure 4).
In the case of the lead
soaps,
the
COO-
asymmetric
stretch consists of a doublet
around 1540 and
1510cm-1 [17].
The doublet is
shifted to lower
frequencies
in the case of the lead
oleate. The band of the COO-
asymmetric
stretch
of all three
copper soaps
is located at a
higher
frequency
than those of the zinc and lead
soaps,
at
around
1590cm-',
as observed in the case of 12-
hydroxyoctadecanoate soaps [20].
This is close to the
frequency
of
1604cm-1
observed for
copper
acetate
[21],
a
compound
where two
copper
atoms are
bridged by
four
carboxylate groups,
as mentioned
earlier;
the
higher frequency
of the COO-
asymmet-
ric stretch reflects the covalent character of the
Cu-O bond. The
higher frequency
of the COO-
asymmetric
stretch observed in the case of the
copper
soaps
indicates that
they
have a similar
bridged
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 31
-.-.-.. . -. . -.-. --.-. -------- ----- ----- ------ ------. ----................ ................ .....
NX !1-(- i81 SO 4r> l -
w NO G 00 C
a
b
c d
Wavenumber
(cm-)
Figure
5 IR
spectra
of
(a)
oleic acid, (b) copper oleate, (c)
zinc oleate and
(d)
lead oleate in the 1800-600cm1
region.
structure,
as has been
proposed
for
copper palmitate
and stearate
[14, 15].
The same trends observed for the infrared
spectra
of the
fatty
acids are observed for those of
soaps
obtained with the same metal: the
spectra
of the
palmitate
and stearate are
very similar,
as are those of
the oleate and linoleate
(although
in the latter case this
could be observed
only
for salts obtained with
zinc)
which in turn are
very
different from those of the
palmitate
and stearate.
Consequently, only
the infrared
frequencies
of the stearate and oleate series are
given
in Tables 13 and 14. The
palmitate
and stearate of the
same metal can be
distinguished
from each other based
on differences in the 1380-1150cm-1
region,
where
vibrations associated with the
CH2 groups
occur
[17];
the oleate and linoleate
spectra
show differences in the
1380-600cm-1
region (Figure 5).
Metal
soaps
obtained
from the same
fatty
acid but with different metals can
be differentiated
primarily
based on the
positions
of
the COO-
asymmetric
and
symmetric
stretch and on
differences in the 1600-1100cm-1
region (Figure 4).
In
addition,
the
spectra
of the lead
palmitate
and lead
stearate are characterized
by
much more intense
multiple
bands due to the
CH2 wagging
modes be-
tween 1400 and 1180cm-'
compared
to those of the
corresponding
zinc or
copper soap (Figure 4d).
Raman
spectroscopy
As the Raman
spectra
for the
compounds
examined
were
very
similar or identical above
1700cm-1,
this area
was not considered.
Only
the
range
between 200 and
1700cm-1,
which showed
differences,
is discussed
below. Raman
frequencies
for the acids and metal
soaps
are listed in Tables 15 to 17. Since
only
a few of
these
compounds
have been studied
by
Raman
spectroscopy,
the
peaks
in the
spectra reported
here
were not
fully assigned.
The
spectra
of
palmitic
acid and stearic acid are
almost
identical,
that of stearic acid
being
weaker than
that of
palmitic acid,
however
(Figure 6).
This is also
the case for the
spectra
of the stearate
soaps compared
to those of the
palmitate soaps,
which is
why
the latter
are
reproduced
in the
present
article. Two weak
peaks
below 600cm-' can be used to differentiate the two
acids;
these are located at 370 and 567cm-' in the case
of
palmitic
acid and at 340 and
570cm1
in the case of
wivenumner em-I
Figure
6 Raman
spectra
of
(a)
stearic acid and
(b) palmitic acid.
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
32 L. ROBINETAND M.-C. CORBEIL
Table 15 Raman
frequencies (cm1)
for
palmitic
acid and metal
palmitates
Palmitic acid
Copper palmitate
Zinc
palmitate
Lead
palmitate Attribution*
1627 w
1540 w
1514 w
1480 w 1481 w
1463 m 1463 s 1468 s 8
CH2
1455 m 1458 s 1456 s 1458 m
1437 s 1433 vs 1439 vs 1440 vs 8
CH2
1419 m 1415 s
1411 w 8
CH2
1406 w
1398 m
1370 w 1368 w 1373 w 1370 vw 8
CH3
1348 vw
1333 vw
1316 vw
1294 vs 1295 vs 1294 vs 1294 vs 8
CH2
1275 w 1273 w
1253 vw
1231 vw
1210 vw
1188 vw
1172 m 1172 m 1173 m
8CH2
1128 s 1127 s 1130 vs 1128 vs vC-C + 8 C-C-C
1100 m 1095 m 1100 m 1097 m v C-C + 8 C-C-C
1062 vs 1060 vs 1061 vs 1062 vs v C-C + 8 C-C-C
1052 sh 1049 w 1048 w
1027 w 1027 vw 1027 vw
1008 vw
975 vw 973 vw
946 m 950 m 923 m
930 w
910 w v C-C
(carboxyl)
891 m 887 m 888 m 888 m 6
CH3
+ v C-C
873 m
696 vw
680 m COO
590 vw
581 w
567 vw 8 COO
501 w
461 w 460 w 457 w
449 w
424 w
390 m 398 w
370 w 370 w 375 m 378 w
304 s 270 w M-O
253 s 233 w M-O
*Based on
Yokoyama
et al.
[1 6]for palmitic
acid
(IR
and
Raman)
and Hu
et
al.
[18]for
stearic acid and
copper
stearate
(Raman).
stearic acid.
However,
these
peaks
are so weak in the
spectrum
of stearic acid that
they
can be difficult to
observe. A
peak corresponding
to skeletal vibrations
around 1100cm-'
might
be used to differentiate
pal-
mitic acid from stearic acid and
palmitate soaps
from
stearate
soaps,
as the
frequency
is
always
lower for the
latter.
However,
the difference in
frequency (only
2 or
3cm-') may
not be
large enough
to allow differentia-
tion of one acid from the
other,
depending
on the
instrument resolution.
The
palmitate
and stearate
soaps
can
easily
be
differentiated from the
corresponding
acid. While the
spectra
of the
soaps
and the acids are almost identical
between 880 and
1700cm-',
where
peaks
associated
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 33
Table 16 Raman
frequencies (cm-)
for stearic acid and metal stearates
Stearic acid
Copper
stearate Zinc stearate Lead stearate Attribution*
1541 vw 1544 w
1520 vw 1518 vw 1515 w
1488 vw 1487 w 1488 w
1465 s 1467 sh 8
CH2
1459 s 1458 s 1460 s
1438 vs 1435 vs 1438 vs 1440 vs 8
CH2
1417 m 1415 s
1410 w 1406 w 8
CH2
1397 m
1368 w 1368 w 1370 w 1368 w
8CH3
1346 vw
1333 vw
1317 vw
1294 vs 1293 vs 1293 vs 1294 vs 8
CH2
1280 sh
1264 m 1261 vw
1243 vw
1224 vw
1205 vw
1185 vw
1172 m 1175 w 1170 m 1171 m
8CH2
1129 s 1127 s 1129 m 1127 vs vC-C+SC-C-C
1102 m 1101 m 1104 m 1102 m vC-C + 8 C-C-C
1061 s 1060 s 1061 vs 1061 vs v C-C + 8 C-C-C
1037 w 1037 vw 1039 w
945 m 948 m 922 m
CH2
919 w
908 w v C-C
(carboxyl)
891 m 889 m 889 m 887 m
8 CH3
+ v C-C
873 w
758 w
694 vw
681 m 8 COO
626 w Cu-O
570 vw
547 w
502 vw
454 vw
427 m 426 vw 418 w
393 w 399 vw 386 vw
360 m
340 vw 341 w 340 w
285 s M-O
242 m 243 w M-O
*Based on Hu et al.
[18]for
stearic acid and
copper
stearate
(Raman).
with vibrations of the
hydrocarbon
chain are
found,
additional
peaks
are observed in the
spectra
of the
soaps
below
880cm-1 (Figure 7).
Differences are ob-
served in this
region
between the
palmitate
and
stearate of the same metal and between
palmitates
or
stearates obtained with different metals. In
addition,
a
peak
located between 920 and 950cm-'
appears
in the
spectrum
of
every
metal
soap (Figure 8).
A
peak
at
946cm-' was
reported
for stearic acid and
assigned
to
CH2 bending
modes
[18],
but this
peak
was not
observed in the
spectrum
of stearic acid obtained in
the course of the
present study.
It is
possible
that this
peak
is weak in the case of the acid and becomes more
intense in the case of the
soaps.
The
frequency
of the
peak
is the same for the
palmitate
and stearate of the
same
metal,
but
changes
with the metal ion.
However,
since the difference is
quite
small when
going
from
zinc to
copper (about 4cm-'),
other
peaks
must be used
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
34 L. ROBINET AND M.-C. CORBEIL
Table 17 Raman
frequencies (cm1)
for metal oleates and zinc linoleate
Copper
oleate Zinc oleate Zinc linoleate Lead oleate Attribution*
1655 s 1652 m 1657 vs 1634 m v C=C
1458 vs 1458 vs 8
CH2
1439 vs 1438 vs 1445 vs 1444 vs 8
CH2
1412 sh 1408 m
1396 m 1396 m
1371 vw 1370 vw
1298 s 1295 vs 1292 s 1294 s
1271 m 1265 m 1268 s
1242 w 1241 w
1120 m 1119 m 1109 m v C-C
1087 vs 1093 vs 1095 vs v C-C
1062 s 1061 vs 1060 s
1041 w
969 w 963 m 971 m
955 m 949 m 950 m
926 m
886 m 888 s 886 w
781 s
726 w
709 m
687 w
544 w 543 vw
479 w
457 m
441 m
412 vw
374 w
360 m
318 w 321 w
294 m
247 vw
236 m
204 m
*Based on Tandon et al.
[19]for
oleic acid
(Raman).
Figure 7 Raman
spectra
of
(a) copper palmitate, (b)
zinc
palmitate
and
(c)
lead
palmitate.
2 400 oo 600 o
oC-
1 400
PVavenumnbr cm-1
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 35
to confirm the identification. Raman
spectra
of lead
stearate and lead
palmitate
are characterized
by
multi-
ple
bands between 1180 and
1400cm-';
these bands are
absent from the
spectra
of the
corresponding copper
and zinc
soaps (Figure 9).
It is
likely
that these bands
correspond
to
CH2 wagging
modes
which,
as noted
OC\
b
rOr
,-
A
91000
WVavenumber c m-i1
Figure
8 Raman
spectra
of
(a) copper palmitate, (b)
zinc
palmitate
and
(c)
lead
palmitate
in the 850-1050cm1
region.
a /en, cm
- 1'
0
A
,w-'n,.1e {m.
Figure
9 Raman
spectra
of
(a)
lead
palmitate
and
(b)
lead stearate
in the 1150-1400cm
-
region.
previously, produced
much more intense bands in the
infrared
spectra
of lead stearate and lead
palmitate
compared
to those in the infrared
spectra
of the
corresponding copper
and zinc
soaps.
This
hypothesis
is
supported by
the fact that more bands are observed
in the Raman
spectrum
of lead stearate than in the
Raman
spectrum
of lead
palmitate,
as observed in the
infrared
spectrum
of the two
compounds.
The
spectra
of the zinc
palmitate
and stearate and lead
palmitate
and stearate show an additional
peak
around 1398cm-'
in the case of the zinc
soaps
and around 1415cm-' in
the case of the lead
soaps compared
to the
spectra
of
the
corresponding copper soaps.
This
peak probably
corresponds
to the COO-
symmetric stretch,
a vibra-
tion which
produced
a much more intense band in the
infrared
spectra
of the zinc and lead
soaps
than in the
infrared
spectra
of the
copper soaps.
Having synthesized only
one linoleate
soap,
zinc
linoleate,
it is more difficult to
identify
trends. Based
on the
spectra
obtained for the zinc
soaps,
oleate and
linoleate soaps
have similar
spectra,
which are in turn
very
different from those of
palmitate
or stearate
soaps.
Zinc oleate can
easily
be
distinguished
from zinc
linoleate
by looking
at the
region
between about
900-1120cm-'
(Figure 10).
As observed in the case of
the
palmitates
and
stearates,
the
spectra
of zinc oleate
Cw
850 900 950 1000 1050 1100
Wavenumber
(cm-1)
FIgure
10 Raman
spectra
of
(a)
zinc oleate and
(b)
zinc linoleate in
the 850-1150cm-1
region.
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
36 L. ROBINET AND M.-C. CORBEIL
and linoleate both show an additional
peak
around
1396cm-' while the
spectrum
of lead oleate shows an
additional
peak
at 1408cm-' with
respect
to the
spec-
trum of
copper
oleate. It must be noted that the
spectrum
obtained for lead oleate was
very poor.
APPLICATION TO CASE HISTORIES
The diffraction
patterns
and FTIR
spectra
obtained
for
pure
metal
soaps
were used to
try
to determine the
composition
of metal
soaps
or metal
soap
mixtures
found on artifacts to see if the information
gained
from
the characterization of the
pure synthetic compounds
could be
applied
to actual
samples.
The first case
study
concerns white
crystals
that were
observed on a wind
pipe
encased in leather that was
treated at the Canadian Conservation Institute
(CCI)
in the
early
1990s
[22].
The metal of the wind
pipe
was
identified as zinc. An
X-ray
diffraction
pattern
of the
white
crystals
was obtained that matched the ICDD
pattern
5-0079 for zinc
stearate,
but extra lines were
observed.
Comparison
with
patterns
obtained as
part
of the current
study
showed that all lines could in fact
be attributed to zinc
stearate,
the ICDD
pattern
5-0079
being incomplete [12].
The
X-ray long spacing
at
4.248nm was
missing
from the
pattern
of the white
crystals
because the
pattern
was measured with the
microdiffractometer which does not measure below
30,
20.
However,
considerable line
broadening
was
observed which
precluded using
the
region
between
20 and
300,
20 to determine if zinc oleate or linoleate
were also
present,
since,
as discussed
previously,
the
patterns
of these three zinc
soaps
are identical below
200,
20.
Surprisingly,
the infrared
spectrum
of the
white
crystals
showed that zinc
palmitate
was
present
in addition to zinc
stearate,
and was in fact the
major
component.
This result was confirmed
by gas
chroma-
tography-mass spectrometry (GC-MS),
which also
showed that zinc oleate was
present
in minor amounts.
The white
powder covering
the zinc
pipe
resulted
from the reaction of an oil that was used on the leather
with the metal of the
pipe.
It has been shown that
fatty
acids can
readily evaporate
from oil
paint
films to
condense on
glass
as
ghost images [23],
this reaction
being
much more
rapid
in the case of
palmitic
acid than
stearic acid.
Therefore,
it is
possible
that
palmitic
acid
released
by
the oil on the leather reacted much more
rapidly
with
zinc, forming
an
amorphous precipitate
(i.e.,
one that would not
yield
an
X-ray pattern),
while
the slower reaction between stearic acid and zinc
allowed for a
crystalline compound
to be formed.
A second
example
is a white inclusion that was
observed in the
paint layer
of a
painting
entitled
L'Annonciation
by Jean-Antoine Aide-Crequy,
dated
1776
(parish
of Notre-Dame de Bonsecours de
l'Islet)
[24].
A cross-section
prepared
from a
sample
taken
from a dark
green
area showed a white inclusion with
a
waxy
texture
(Figure 11). X-ray
diffraction
analysis
of
the inclusion was
performed directly
on the cross-
section
using
the microdiffractometer attachment with
a
50?tm
collimator
[25]
and revealed the
presence
of
lead white
only.
The infrared
spectrum
of
scrapings
of
the white substance was also dominated
by absorp-
tions attributable to lead white at
3537, 1404,
1046 and
681cm-';
a trace of
drying
oil was
present,
as indicated
by
a weak
absorption
around 1700cm-'. A
strong
band
centred at
1521cm-'
was characteristic of lead
palmitate
and/or stearate.
However,
the broad carbonate
absorp-
tion centred at
1404cm-'
masked the 1380-1150cm-'
region
where weaker
peaks
that are characteristic of
lead stearate and
palmitate
would be located. There-
fore in this
particular
case it was
impossible
to differ-
entiate between lead
palmitate
and stearate because of
the
presence
of lead white.
A third
example
concerns an efflorescence that was
covering large
areas of a
painting by
Alfred Bastien
entitled Over the
Top,
Neuville-
Vitasse,
dated 1918
(Canadian
War
Museum) [26].
It was found
by
FTIR
spectroscopy
to be a mixture of lead
palmitate,
lead
stearate and free
fatty
acids. The
presence
of lead
soaps
was indicated
by strong absorptions
at 1541 and
1513cm-1',
while the
presence
of free
fatty
acids was
indicated
by
an
absorption
around 1711cm-'. In this
case,
characteristic
peaks
for lead
palmitate
and lead
stearate were observed in the 1380-1150cm-'
region,
Figure
11
Photomicrograph
of a cross-section
showing
the white
inclusion found in the
painting
entitled L'Annonciation
by
Aide-
Cr6quy.
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 37
those of lead
palmitate being
more intense than those
of lead stearate.
Finally,
it was
possible
to
identify
white
crystals
found on an
Egyptian mummy
and an unrelated coffin
from the 25th or 26th
Dynasty (700 BC)
in the
collection of the British Museum as
primarily palmitic
acid with
possibly
a trace of stearic
acid,
based on the
infrared
spectroscopy
data
acquired during
the
present
study [27, 28].
The
1280-1180cm-1 region
showed
peaks coinciding
with those of
palmitic acid;
two
much weaker
peaks
could
correspond
to those of
stearic acid.
CONCLUSION
Synthesis
of
pure
metal
soaps (zinc palmitate,
zinc
stearate,
zinc
oleate,
zinc
linoleate, copper palmitate,
copper
stearate,
copper
oleate,
lead
palmitate,
lead
stearate and lead
oleate)
and their characterization
by
X-ray diffraction,
infrared
spectroscopy
and Raman
spectroscopy
made it
possible
to document features
that are characteristic of
specific compounds. Soaps
obtained from different
fatty
acids with the same metal
ion show differences
by
all
techniques,
as do
soaps
obtained with the same
fatty
acid but with different
metal ions.
Characterization
by X-ray
diffraction
requires
that
the
soap
be in the
crystalline
form.
Also,
since the most
intense characteristic lines of the various metal
soaps
are at low
angles, patterns
must be measured with
equipment
that can resolve lines in this
region.
Lines
at
higher angles
are also characteristic to a lesser extent
but are much weaker.
FTIR
spectroscopy provides
useful information
and does not
impose
limitations due to the
type
of
instrument used. The
position
of the
carboxylate
absorption
is related to the metal
present
in the
soap,
while weaker
absorptions
in the
1300-1000cm-1 region
can be used to determine the
fatty
acid of the
soap.
Raman
spectroscopy
also
provides
information that
makes it
possible
to differentiate metal
soaps.
Metal
soaps
found on artifacts and works of art
were characterized based on the data obtained in the
current
study.
However, accretions,
efflorescence
or inclusions in
paint
film often consist of mixtures
of
degradation products
and
undegraded original
materials, making
a
precise
identification of all com-
ponents
more difficult. Infrared
spectroscopy
seems
to be best suited for
analysing
mixtures of metal
soaps;
this
technique
is
very
sensitive,
as
long
as
absorptions
from
pigments
or other
compounds
do not mask
characteristic features
as,
unfortunately,
often
hap-
pens.
It is
hoped
that the work
presented
here will
help
in the characterization of metal
soaps
found on actual
objects.
Identification is
key
to
understanding
how
and
why they
form and this
may
lead to
preventive
measures.
APPENDIX: SYNTHESIS AND ANALYTICAL
PROCEDURES
Synthesis
The
synthesis
of metal
soaps
has been
reported
in
detail elsewhere
[12]
and will therefore
only
be
summarized here.
The
starting
chemicals were
supplied by Sigma-
Aldrich
(palmitic
and stearic acid:
purity grade
>
99.5%;
oleic and linoleic acids:
purity grade - 99%;
zinc,
lead and
copper
chloride or
sulphate).
The
purity
of the
fatty
acids was verified
by
GC-MS
analysis,
since it was essential to obtain
pure
metal
soaps.
The
appropriate quantities
of
acid,
metal salt
(in excess,
dissolved in
water)
and sodium
hydroxide
solution
(0.5M)
were heated in the same flask to
700C. At that
temperature,
the
appearance
of the
solution
changed suddenly,
the acid
crystals being
replaced
with a flocculent
precipitate
which
agglom-
erated at the surface of the solution. The solution was
heated at 70-80'C for about 20 minutes. The
pre-
cipitate
was filtered and washed with ethanol and
warm
water,
and dried around 1000C for one to two
hours. The
purity
of the
synthesized
metal
soaps
was
assessed
through
instrumental characterization car-
ried out as
part
of the
present study
and
through
elemental
analysis
for carbon.
X-ray powder
diffraction
The
X-ray
diffraction
patterns
of the metal
soaps
synthesized
as
part
of the
present study
as well as those
of
palmitic
and stearic acids were obtained
using
a
Rigaku
RTP 300 RC
generator equipped
with a
rotating
anode and a cobalt
target operated
at 45kV and 160mA.
Diffractograms
were measured
at room
temperature (250C)
in
step
mode from 1
to
400, 20,
using
a
0-20
vertical
goniometer (step
size:
0.020,
count time:
1s).
Monochromatic Co Ka
(0.17890nm)
radiation was
produced by filtering
the diffracted beam
(Fe).
Mica
(NIST
SRM
675)
was
used as an internal
d-spacing
calibration standard.
Since cobalt radiation was used instead of
copper
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
38 L. ROBINETAND M.-C. CORBEIL
radiation,
reference
intensity
ratios were not deter-
mined.
Fourier transform infrared
spectroscopy
The
samples
were mounted in a
Spectra-Tech
low
pressure
diamond anvil cell and
analysed by
Fourier transform infrared
spectroscopy (FTIR)
using
a Bomem Michelson MB-100
spectrometer
equipped
with a wide band
mercury
cadmium telluride
(MCT)
detector and a microbeam
compartment.
Spectra
were
acquired
in the 4000 to
400cm-1 range
at a resolution of
4cm-1 by summing
200
interferograms.
Raman
spectroscopy
Raman
spectra
were obtained
using
a
Jobin
Yvon
Infinity
micro-Raman
spectrometer equipped
with
a
green Nd:Yag
laser
operating
at 532nm and an
1800
groove per
mm
grating,
with the laser
power kept
below
500ltW.
The
spectra
were recorded between
200 and 1700cm-' with a resolution of
?+3cm-1
and an
exposure
time of 300s
(two
or three
acquisitions).
ACKNOWLEDGEMENTS
The authors would like to thank Elizabeth
Moffatt,
Senior Conservation
Scientist, CCI,
for
sharing
her
knowledge
of metal
soaps
and infrared
spectroscopy.
Laurianne Robinet's
internship
at the Canadian Con-
servation Institute was
financially supported by Jean-
Philippe
Baur and the
Rotary
Club Orlans Val de
Loire.
REFERENCES
1
Licata, F.J.,
'Some metallic
soaps
used in the
protective
coatings industry', Official Digest
388
(1957)
485-491.
2
Rinse, J.,
'Metal
soaps',
American
PaintJournal (March 13, 1967)
22-28.
3 Tilbrooke, D.,
'The
fatty
acid corrosion of
copper alloys
and
its
treatment', ICCM Bulletin
6(3-4) (1980)
46-52.
4
Horovitz, I., 'Paintings
on
copper supports: techniques,
deterioration and conservation', The
Consewrvator
10
(1986)
44-48.
5 Gunn, M.,
and Martin, E.,
'M&canisme
d'alteration d'un
alliage
cuivreux en
presence
d'un liant huileux' in Art et
Chimie,
la
Couleur, ed.
J. Goupy
and
J.-P. Mohen,
CNRS
Editions, Paris
(2000)
141-145.
6 Schrenk, J.L.,
'Corrosion and
past "protective"
treatments of
the Benin "bronzes" in the National Museum of African Art'
in Materials Issues in Art and
Archaeology
II, ed. P.B.
Vandiver,
J.
Druzik and G.S. Wheeler,
Materials Research
Society,
Pittsburgh (1991)
805-812.
7
Schrenk, J.L.,
'The
royal
art of Benin: surfaces, past
and
present'
in Ancient and Historic Metals:
Consewrvation
and
Scientific
Research, Getty
Conservation
Institute,
Marina del
Rey (1994)
51-62.
8
Burmester, A.,
and
Koller, J.,
'Known and new corrosion
products
on bronzes: their identification and assessment,
particularly
in relation to
organic protective coatings'
in
Recent Advances in the
Consewrvation
and
Analysis of Artifacts,
Summer Schools
Press,
London
(1987)
97-104.
9
O'Neill, L.A.,
and
Brett, R.A.,
'Chemical reactions in
paint
films', Journal of
the Oil and Colour Chemists' Association 52
(1969)
1054-1074.
10
Bell, S.H., 'Controlling
loss of
gloss',
Paint and Varnish
Production 60
(1970)
55-60.
11
Noble, P., Wadum, J., Groen, K., Heeren, R.,
and van den
Berg, Kj., 'Aspects
of 17th
century binding
medium:
inclusions in Rembrandt's
Anatomy
Lesson
of
Dr Nicolaes
Tulp'
in Art et
Chimie,
la
Couleur,
ed.
J. Goupy
and
J.-P. Mohen,
CNRS
Editions,
Paris
(2000)
126-129.
12
Corbeil, M.-C.,
and
Robinet, L., 'X-ray powder
diffraction
data for selected metal
soaps',
Powder
Difraction 17(1) (2002)
52-60.
13 van
Niekerk, J.N.,
and
Schoening, F.R.L.,
'A new
type
of
copper complex
as found in the
crystal
structure of
cupric
acetate,
Cu2(CH3COO)4-2H20',
Acta
Crystallographica
6
(1953)
227-232.
14
Herron, R.C.,
and
Pink, R.C., 'Magnetic properties
of some
metal
carboxylates', Journal of
the Chemical
Society (1956)
3948-3952.
15 Stommen, D.P., Giroud-Godquin, A.-M., Maldivi, P.,
Marchon, J.-C.,
and
Marchon, B.,
'Vibrational studies
of some
dicopper tetracarboxylates
which exhibit a
thermotropic
columnar
mesophase', Liquid Crystals 2(5)
(1987)
689-699.
16
Yokoyama, I., Miwa, Y., Machida, K., Umemura, J.,
and
Hayashi, S.,
'Extended molecular mechanics simulation of
thermodynamic quantities,
structures,
and vibrational
spectra
of
fatty acids',
Bulletin
of
the Chemical
Society ofJapan
66
(1993)
400-413.
17 Ahn, D.J.,
and
Franses, E.I., 'Orientations of chain axes and
transition moments in
Langmuir-Blodgett monolayers
determined
by polarized
FTIR-ATR
spectroscopy', Journal
of Physical Chemistry
96
(1992)
9952-9959.
18 Hu, Z.-S., Hsu, S.L.,
and
Wang, P.S.,
'Tribochemical
reaction of stearic acid on
copper
surface studied
by
surface
enhanced Raman
spectroscopy', Tribology
Transactions
35(3)
(1992)
417-422.
19
Tandon, P., F6rster, G., Neubert, R.,
and
Wartewig, S.,
'Phase transitions in oleic acid as studied
by X-ray
diffraction
and FT-Raman
spectroscopy', Journal of
Molecular Structure
524
(2000)
201-215.
20 Tachibana, T., Kayama, K.,
and Takeno, H.,
'Studies of
helical
aggregates
of molecules. III. The bivalent metal
soaps
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
THE CHARACTERIZATION OF METAL SOAPS 39
of
optically
active
12-hydroxyoctadecanoic
acid',
Bulletin
of
the Chemical
Society of Japan
45
(1972)
415-422.
21
Grigor'ev, A.L.,
'Infrared
absorption spectra
of
copper
and
zinc acetates and their
hydrates
and
ammines',
Russian
Journal
of Inorganic Chemistry
10
(11) (1965)
1358-1361.
22
Corbeil, M.-C., Moffatt, E.,
and
Miller, D.,
'Characterization
of the metal
component
of a wind
pipe', unpublished
report
ARS 3069,
Canadian Conservation Institute,
Ottawa
(1992).
23
Schilling, M.R., Carson, D.M.,
and
Khanjian, H.P.,
'Gas
chromatographic
determination of the
fatty
acid and
glycerol
content of
lipids.
IV.
Evaporation
of
fatty
acids and the
formation of
ghost images by
framed oil
paintings'
in ICOM
Committee
for Conservation,
12th Triennial
Meeting, Lyon,
Preprints,
ed.
J. Bridgland, James
&
James,
London
(1999)
242-247.
24
Corbeil, M.-C.,
and Moffatt, E., 'Analyse
des materiaux de
L'Annonciation, par Aide-Crequy', unpublished report
ARL
3880,
Canadian Conservation Institute,
Ottawa
(2000).
25 Corbeil, M.-C.,
and
Sirois, P.J., 'Applications
of
microdiffractometry
to the
study
of artistic and historic
objects'
in
Proceedings,
4th International
Conference
on Non-
destructive
Testing of
Works
of Art,
Deutsche Gesellschaft
ffir
Zerst6rungsfreie Priifung
e.V.,
Berlin
(1994)
278-284.
26 Corbeil, M.-C.,
and Moffatt, E., 'Analysis
of the
efflorescence and surface
coating
on a
painting by
Alfred
Bastien', unpublished report
ARL
3802,
Canadian
Conservation Institute,
Ottawa
(1999).
27
Robinet, L.,
and
Thickett, D., 'Analysis
of white
deposit
from
mummy EA15654C', unpublished report
CA2000/
33,
British Museum,
London
(2000).
28
Robinet, L.,
and
Thickett, D., 'Analysis
of white
deposit
on
Dunrobin coffin', unpublished report CA2000/36,
British
Museum,
London
(2000).
AUTHORS
LAURIANNE ROBINET obtained a
degree
in
chemistry
from the Institut Universitaire de
Technologie
in
Orsay,
France in 1997. She then
specialized
in
organic
and
analytical chemistry
and
completed
a master's
degree
in 1999 and a DESS in instrumentation and
physico-chemical analytical
methods in 2000 at the
Universit&
d'Orsay. Following internships
at the
Laboratoire de Recherche des Musees de France
(1997),
the Laboratoire de
Dynamique,
Interaction et
Reactivit&
(1999)
and the Canadian Conservation
Institute
(2000),
she worked at the British Museum
from 2000 to 2002 as a conservation scientist. She then
joined
the Conservation and
Analytical
Research
Department
of the National Museums of Scotland
where she is
responsible
for
inorganic analysis.
Address:
National Museums
of Scotland,
Conservation and
Analytical
Research
Department,
Chambers
Street, Edinburgh
EH1
1JF,
UK. Email:
l.robinet@nms.ac.uk
MARIE-CLAUDE CORBEIL received a BSc in
chemistry
from Universit& de
Montr&al.
She then
specialized
in
inorganic chemistry
and
crystallography
and com-
pleted,
at the same
university,
a
master's program
in
1984 and a PhD
program
in 1987. She has worked in
the
Analytical
Research
Laboratory
of the Canadian
Conservation Institute since
1988,
performing analyses
of museum
objects,
and is
currently
a Senior Conser-
vation Scientist. Address:
Analytical
Research
Laboratory,
Canadian Conservation
Institute, Department of
Canadian
Heritage,
1030 Innes
Road, Ottawa, Ontario,
Canada
KIA
OM5. Email:
marie-claudecorbeil@pch.gc.ca
Risumi
-
Afin
de mieux caracteriser les savons
metalliques
trouves dans les couches
picturales
ou sur les
surfaces
metalliques, plusieurs
savons
mitalliques
ont
it6 synthetises
et leurs
patrons
de
diffraction
des
rayons
X et
spectres infrarouge
et Raman mesures. Les savons
metalliques
ont
6t6
obtenus a'
partir
de
quatre
acides
gras qu'on
trouve dans les huiles
siccatives
-
deux acides
gras
satures
(acide palmitique
et acide
stiarique)
et deux acides
gras
insatures
(acide
olique
et acide
linoldique)
-
et de trois
metaux,
le
cuivre,
le zinc et le
plomb, qu'on
rencontre
fri6quemment
dans les
alliages
et les
peintures.
Les donnies sont
rapporties pour
les
composes
suivants: acide
palmitique,
aide
stearique,
acide
oleique,
acide
linoleique, palmitate
de
cuivre,
stearate de
cuivre,
oleate de
cuivre, palmitate
de
zinc,
stearate de
zinc,
oleate de
zinc,
linoleate de
zinc, palmitate
de
plomb,
stiarate de
plomb,
oleate de
plomb.
Les
patrons
et
spectres
des
composes
montrent des
particularitis qui permettent
de
diffrencier
les savons obtenus a'
partir
du meme metal mais d'acides
gras diffirents
ou les
savons obtenus a'
partir
du meme acide
gras
mais de
diffirents
metaux.
L'identification
des savons
metalliques
est necessaire
si l'on veut
comprendre
comment et
pourquoi
ces
composes seforment
sur les
objets,
ce
qui peut
mener a' l'?tablissement de
mesures
preventives.
Zusammenfassung
-
Um
Metallseifen
wie sie in Malschichten oder
auf Metallobeflachen gefunden
werden besser
charakterisieren zu
kdnnen,
wurde eine Anzahl von ihnen
synthezisiert
und durch
Rontgendiffraktometrie,
Fouriertransforminfrarotspektroskopie (FTIR)
und
Ramanspektroskopie analysiert.
Die
Metallseifen
wurden aus
typisch
in
trocknenden Olen
aufiretenden
Fettsduren
-
zwei
ungesdttigte (Palmitin-
und
Stearinsdure)
und zwei
gesdttigte (01-
und
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions
40 L. ROBINETAND M.-C. CORBEIL
Linolsaure)
-
und den Metallen
Kupfer,
Zink und Blei
hergestellt.
Die Messdaten werden
fuiir
die
folgenden
Verbindungen angegeben: Palmitinsdure, Stearinsdure, Olsdure, Linolsdure, Zinkpalmitat, Zinkstearat, Zinkoleat,
Zinklinolat, Kupferpalmitat, Kupferstearat, Kupferoleat, Bleipalmitat,
Bleistearat und Bleioleat.
Fiir
alle
Verbindungen
konnten charakteristische
Eigenschaften gefunden
werden. Sowohl die
Seifen
mit verschiedenen Metallionen und
gleicher
Fettsdure wie auch die mit
gleichem
Metallion und verschiedenen Fettsduren
konnen
voneinander unterschieden werden. Die
Identifizierung
ist ein
Schliissel
zum Verstandnis der
Entstehung
von
Metallseifen.
Resumen
-
Con
elfin
de caracterizar de una manera
mrds
completa
los
jabones metdlicos que
se detectan en
peliculas
pict6ricas
o en
superficies metdlicas,
se sintetizaron varios
jabones metdlicos,
midiendose
y estudidndose
los
perfiles
de
difracci6n
de
rayos
X
y
los
espectros
Raman e
infrarrojos transformado
de Fourier
(FTIR).
Los
jabones metdlicos
se
obtuvieron de cuatro acidos
grasos que
se encuentran habitualmente en aceites secos
-
dos saturados
(6cidos palmitico y
estedrico) y
dos insaturados
(oleico y linoleico)
-
y
de los elementos
cobre,
cinc
y plomo,
tres metales
que
son encontrados a
menudo en aleaciones
metdlicas y
sistemas de
capas pict6ricas.
Se relatan los datos obtenidos en los casos de los
siguientes
compuestos:
acido
palmitico,
acido
estedrico,
acido
oleico,
acido
linoleico, palmitato
de
cinc,
estearato de
cinc,
oleato de
cinc,
linoleato de
cinc, palmitato
de
cobre,
estearato de
cobre,
oleato de
cobre, palmitato
de
plomo,
estearato de
plomo y
oleato de
plomo.
Se observaron las caracterfsticas
que
son
especfficas
de los
compuestos
en
cuesti6n.
Los
jabones
obtenidos de los
diferentes
acidos
grasos
con el mismo ion metclico mostraron
diferencias,
asi como en el caso de los
jabones
obtenidos con el
mismo acido
graso pero
con
diferentes
iones metalicos. La
identificaci6n
es clave
para comprender c6mo y por que
se
forman
jabones metdlicos
en los
objetos, pudiendo
llevarnos a
plantear
medidas
preventivas.
STUDIES IN CONSERVATION 48
(2003)
PAGES 23-40
This content downloaded from 193.227.1.43 on Sun, 2 Mar 2014 06:17:35 AM
All use subject to JSTOR Terms and Conditions

You might also like