Sie sind auf Seite 1von 14

Food Hydrocolloids Yol.5 no.4 pp.

339-352, 1991
Large deformation properties of 13-lactoglobulin gel structures
Mats Stading and Anne-Marie Hermansson
SIK-The Swedish Institute for Food Research, PO Box 5401, 5-40229
Goteborg, Sweden
Abstract. Different gel structures formed by 13-lactoglobulin dissolved in distilled water (12% w/w) at
pH 3.0-7.5 have been investigated using tensile measurements at large deformations. Gels formed at
pH 4-6 were opaque, whereas at pH values below or above this range they were transparent. The
fracture properties showed large variations over the pH range studied . Gels formed at low pH were
brittle with low strain and stress at fracture, as opposed to those formed at high pH, which were
rubber-like with high strain and stress at fracture . Gels formed at intermediate pH (pH 4-6) had an
intermediate, near-constant, strain at fracture . The fracture stress was, however , higher at pH 5.5-
6.0 than at pH 4.0--5.2. The specific fracture energy resembled the stress at fracture, with a
maximum of 6 J/m
2
at pH 6.0. Gels formed at pH 4.5, 5.5, 6.5 and 7.5 were all notch-sensitive. The
opaque gels were defined as aggregate gels and the transparent gels were defined as fine-stranded
gels. The fracture properties clearly showed there are differences between the fine-stranded gels
formed at high pH and those formed at low pH. The fracture stress demonstrated that there are
structural differences within the pH range in which the aggregate gels are formed. The non-linearity
of the stress-strain curve was the same for all fine-stranded gels, which had r-shaped stress-strain
curves. The stress-strain curves of the aggregate gels were almost linear. The non-linearity did not
influence the fracture properties. Young's modulus showed two peaks , at pH 3.5 and 6.0, coinciding
with the range where the structure changes between aggregated and fine-stranded. The stress at
fracture also has a maximum at pH 6.0. The high elasticity and fracture stress may depend on strong,
elastic areas in the network structure. Apart from the two peaks , Young' s modulus shows the same
behaviour as the storage modulus , G' , measured at small deformations, but Young's modulus is
slightly larger than 3G'.
Introduction
Many food biopolymers are able to form gels and therefore have an important
effect on the texture of food systems. The texture depends on the mechanical
properties, which in turn depend largely on the structure of the gel network. The
mechanical properties can be divided into small deformation properties and
large deformation properties. The small deformation properties are measured
using non-destructive methods, whereas testing by large deformations includes
fracture properties of the material. The large deformation measurements are
usually performed in compression or, as in this paper, tension.
I3-Lactoglobulin was chosen as a model system in this study because it forms
both aggregated and fine-stranded gels according to the pH in question. Gels
formed at pH 4-6 in distilled water are opaque having a network formed by
aggregates and are defined as ' aggregate gels' (1,2). Gels formed below or above
this intermediate pH range are transparent and are defined as 'fine-stranded
gels'. Figure l(a) shows an example of an aggregate gel and Figure l(b) an
example of a fine-stranded gel. Both gels are formed by 13-lactoglobulin and the
differing property is the pH value . The aggregate gel is formed at pH 5.25 and
the micrograph is obtained by scanning electron microscopy (SEM). The fine-
stranded gel is a 60 nm thick section of a 13-lactoglobulin gel at pH 7.5
photographed in a transmission electron microscope (TEM) . The experimental
339
M.Stading and A.-M.Hermansson
Fig. 1. Electron micrographs of J3-lactoglobulin gels. (a) 10% (w/w) , pH 5.25, (SEM); (b) 12%
(w/w), pH 7.5, (TEM).
technique and full details of the gel structure will be presented in
another paper (M.Langton and A.-M.Hermansson, in preparation). In Figure 1
the magnification factor is -25 times larger in b than in a.
We have earlier studied the small deformation behaviour of the different gel
structures formed by (3). The aggregate gels behaved differently
from the fine-stranded gels in many ways: the aggregate gels had a higher
storage modulus, G', lower critical gel concentration, were more frequency-
dependent and formed structures far below the denaturation temperature. The
fine-stranded gels formed a high pH behaved in the same way as those formed at
low pH, but the critical gel concentration was higher for the former.
The tensile strength of whey protein gels at pH 6.5 and 8.0 has been studied by
Langley et al. who used samples with varying contents of and a-
lactalbumin (4). Their purest sample contained 94% w/w
lactoglobulin and no a-lactalbumin. The whey powders were dissolved in
deionized water to give 15% (w/v) whey solutions, which were heated for 5 min
at 80C. The stress at fracture and the elastic modulus were related by power law
relations to the a-lactalbumin or fraction. The
content was found to contribute more than the a-lactalbumin content to the
tensile strength and to the clastic modulus.
Large deformation testing is a commonly used method for texture evaluation.
Both compressive and tensile tests are utilized but, in the case of biopolymer
gels, large deformation testing has, to date, almost exclusively been synonymous
with compressive testing, even though tensile tests give a clearer picture of the
stresses in the sample. The total stress in compressive testing is the sum of both
tensile stresses and shear stresses, whereas the shear stresses in tensile testing
are negligible. Other advantages of tensile testing are: the energy used in a
tensile test is used only for deformation and not for friction, the start of fracture
340
Deformation properties of [3-lactoglobulin gels
can be determined more exactly and it is possible to study the notch sensitivity of
the material (5).
The main drawback of tensile measurements of biopolymer gels is the
experimental difficulties. The first difficulty encountered is to attach the sample
to the measuring system. Attempts to clamp the gel usually cause damage unless
it is a very firm gel. There are, however, results reported on clamped samples of
cross-linked alginate, but only a small strain (strain <0.08) was used (6).
The second difficulty is to control the location of the fracture i.e. not to get
fracture at the ends. Dumb-bell-shaped samples solve this problem, but the
strain is not constant throughout the sample. Ring-shaped samples pulled
between two dowel pins is another solution which has been used with e.g. gelatin
and agarose gels (7). Myers and Wenrick have compared tensile measurements
of dumb-bell, ring, and oval samples and concluded that the dumb-bell samples
give higher stress and strain at fracture than ring and oval samples due to stress
concentration at the dowel pins (8). They also concluded that it is difficult to
obtain accurate low-strain data with dumb-bell-shaped samples. Straight bars
can be used instead of dumb-bell-shaped samples to obtain constant strain in the
material. The samples can then be glued to the instrument platens, as suggested
by Bagley (9). When a straight sample is used, the location of the fracture can be
directed by making a notch. Since a notch causes stress concentration, its size
must be carefully controlled.
The aim of this study is to show the large deformation behaviour of different
gel structures formed by This paper describes a novel technique
for making tensile measurements of biopolymer gels. The stress and strain
during the deformation are measured and the fracture energy is calculated. The
elastic modulus, the shape of the deformation curve, and the notch sensitivity
are also studied. Comparisons with small deformation measurements will also be
made.
Theory
Notch sensitivity
Materials vary in their sensitivity to notches. The notches may be artificially
made or they may be natural, such as cracks and other inhomogeneities. The
stress at the tip of a notch is always higher than elsewhere due to stress
concentration. It may even be high enough to cause fracture of the material. An
expression for the stress (Tn in a linear elastic material at the tip of a notch with
depth In and radius r
n
was derived early by Inglis (10):
(Tn = (T (1 + 2
r
(1)
where (T is the stress far away from the crack.
Unfortunately equation 1 is not valid for viscoelastic materials as they exhibit
flow.
Griffith has derived an expression for the notch dependence of the fracture
341
M.Stading and A.-M.Hermansson
stress based on the fracture energy (11). The expression is valid for linear elastic
materials and can be expressed as (12):
(2)
Purslow has improved this theory so that it is also valid for viscoelastic materials
with a stress-strain curve of the form (13):
where E is the strain and k and p are constants. The stress then becomes:
1
(J'CC -
r' n
(3)
(4)
where m = p/(P + 1).
The stress of a material which is not sensitive to notches decreases linearly
with increasing notch depth (14):
(

(J' = (J'fO 1
a
(5)
where (J'fO is the stress at fracture for a sample without an applied notch. The
decrease in stress here depends only on the decreasing area. Instead of using the
notch depth, a dimensionless variable a is used
In
a=-
a
(6)
where a is the breadth of the sample (see Figure 4a).
The notch sensitivity, i.e. stress (J' as a function of a, provides information on
the size of the largest structure element, x, causing fracture (15). The stress at
fracture, af, of a notch-sensitive material obeys equation (4) for l.> but for
In::5X the stress obeys equation (5). The smallest notch depth that obeys equation
(4) is then equal to the largest structure element (see Figure 2).
Material and methods
Material
was supplied by the Swedish Dairies Association (SMR) and
produced at Laiterie Triballat, Noyal-sur-Vilaine, France, after an industrial-
scale fractionation process developed at Institute National de Recherche
Agronomique (INRA) , Rennes, France (16). In the INRA process the lipid
fraction of whey is removed by heat precipitation and subsequent micro-
filtration. The batch is then ultrafiltered and diafiltered, rendering a 95% pure
342
Deformation properties of 13-lactoglobulin gels
stress at fracture
x
a
Fig. 2. Stress at fracture as a function of notch size for a materi al that is (a) notch-sensitive and (b)
not not ch-sensitive . The size of the largest structure element is shown by 'x' .
whey protein concentrate. a-Lactalbumin is removed by precipitation at its
isoelectric point followed by centrifugation. The supernatant is diafiltered and
spray-dried, rendering the powder.
The powder had the approximate composition: 95.8% dry
matter, 84.7% protein, 3.5% ash. HPLC, FPLC and gel electrophoresis showed
that the also contained small amounts of a-lactalbumin.
The sodium, potassium and calcium contents of the powder
have been measured using atomic absorption spectroscopy and X-ray fluor-
escence. The powder contained 1% w/w Na, 0.D3% w/w K and D.04% w/wCa.
from Sigma Chemical Co. (L-D13D, lot no. 106F-812D) was
also used as a comparison for small deformation measurements of the storage
modulus G'. These samples are denoted Sigma, as in a previous publication (3).
Sample preparation
The was dissolved in degassed, distilled water at 12% w/w and
pH was adjusted with D.l mol/drrr' HCl or D.l mol/dar' KOH. At extreme pH
and high concentration of 1 mol/drrr' HCI or 1 rnol/dnr' KOH
was used. The samples were then degassed . The solution was
poured into square moulds made of aluminium tubes, inner dimensions 13 x 13
x 60 mm, sealed with heat-proof tape (Scotch 425). The cylinders had been
greased with bearing grease to prevent the samples from sticking to the mould.
Moulds made of glass and acrylic plastic were also tested but the
gels stuck to the walls of the mould. Sigmacote was tested instead of the grease
but it promoted air bubble formation on the walls of the mould. The cylinders
were heated in a water-bath with the heating profile shown in Figure 3.
Tensile measurements
The gels were taken out of the moulds and cut into 2D-23 mm
long test-pieces with a scalpel. A notch was cut along one side of the test-piece
perpendicular to the direction of elongation as illustrated in Figure 4(a). The
notch was 1 mm deep unless otherwise stated. The samples that were fractured
without a notch were moulded in a separate mould , see Figure 4(b). Each test-
343
lClmin
Temperature
M.Stading and A.-M.Hermansson
C 100
80
60
40
20
O ~
o 20 40 60 80 100 120 140 160
time (minutes)
Fig. 3. Heat treatment of the 13-lactoglobulin samples.
Original
area
PneumatIc
grip
Sample
Notch
Sand
paper
Fig. 4. Experimental set-up for the tensile tests (a) and sample shape for measurements without
notch (b).
piece was attached to the measuring instrument using a method where the test-
pieces were glued to plates covered with sand paper, using a cyanoacrylate glue,
Loctite 401 or 406. The sandpaper was applied to increase the gluing area
between the sample and the plate. The plates were then clamped with two
pneumatic grips. The samples were fractured in tension using an Instron 1122,
and the force and the deflection were monitored on a personal computer. A
constant cross-head speed of 10 mm/min was used which corresponds to an
initial strain rate of E = 8.3 X 10-
3
S-1 for a 20 mm long sample.
Calculations
Strain. The strain is the dimensionless deformation and can be defined as:
344
Deformation properties of J3-lactoglobulin gels
I
EH = In ~ (7)
where La is the original length of the sample and I is the present length (17). EH is
called Hencky strain and is used in this study because it is additive. EH will be
referred to as E.
An alternative to Hencky strain is Cauchy strain, E
c
:
L- L
o
E
c
= - -
f
o
(8)
At small strains, when approaches L
o
, Hencky and Cauchy strains are
approximately equal.
Stress. The stress, rr , is the force F per unit area A:
F
cr=-
A
(9)
The easiest way to calculate the stress is to divide the measured force by the
original area, which is a sufficient measure for small strains. The sample will,
however, get narrower as the strain increases, since the volume of the sample is
approximately constant. It is therefore more relevant to use the smallest area.
The shape of a sample at pH 4.5 and 6.5 was photographed during a tensile
measurement, and the area at fracture was calculated from the photographs.
The samples at pH between 4.0 and 6.0 were assumed to narrow in the same way
as at pH 4.5 (see Figure 5). The samples at pH >6.0 were assumed to narrow in
the same way as the sample at pH 6.5, and the narrowing of the sample at pH <
4.0 was negligible. This measured area is used in the stress calculations unless
otherwise specified.
ELasticity. The elasticity was measured by Young's modulus E which is defined
as:
(10)
A 5th-order polynomial was fitted to the stress-strain curve using non-linear
regression, and Young's modulus was calculated as the derivative of the
polynomial at E = O. This method gives a more accurate measure of E than using
a ruler.
Energy. When a material is strained the applied energy is partly stored and
partly dissipated due to flow. The strain energy is the elastic energy stored in the
material during deformation. The maximum strain energy W is equal to the total
area under the stress-strain curve (18,19), that is:
345
M.Stading and A.-M.Hermansson
W = fade (11)
The fracture resistance R is the energy required to propagate a crack by unit
area (18,19). The strain energy stored in the material during deformation makes
the material fracture when it exceeds the fracture resistance. In a viscoelastic
material, however, a part of the applied energy is dissipated due to flow. The
specific fracture energy, R" which is the part of the resistance used to propagate
the crack, is therefore used instead of R. R, can be approximated by (5):
(12)
(13)
where In is the notch depth.
The specific fracture energy can be calculated from the stress-strain curve by
combining equations (11) and (12).
Statistics. The results presented in this paper are mean values ii of at least four
independent measurements, and the error bars denote:
s
u t Yn
where n is the number of measurements, S the estimated standard deviation and
t is taken from the Student's t distribution. This corresponds to a 95% confidence
interval when the measured property is assumed to be normally distributed. The
stress at fracture for the samples without a notch was estimated using the
jacknife method, which estimates the maximum stress at fracture (20).
Results and discussion
Fracture properties
The stress-strain curves for 12% w/w gels are shown in Figure 5.
The gel formed at pH 3.0 is brittle, whereas the gel formed at pH 7.5 can be
extended to large deformations. The gels formed at pH 4.5,5.5 and 6.0 show a
similar strain at fracture, but the stress at fracture differs. The stress and strain at
stress
10
5
pH6.5
pH 7.5
0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
strain
Fig. 5. Stress-strain curves in tension for 12% 13-1actoglobulingels.
346
Deformation properties of 13-lactoglobulin gels
fracture, the specific fracture energy and the non-linearity are extracted from
the stress-strain curves and are described below.
Strain. The strain at fracture Ef as a function of pH can be divided into three
regions, as shown in Figure 6(a). The strain is approximately constant at
intermediate pH (pH 4-6), i.e. in the range where aggregate gels are formed.
The strain for the transparent gels differs, however, between low and high pH.
The gels at low pH are brittle with low Ef' while gels at high pH have a large Ef.
Transparent, fine-stranded gels, both at low and high pH, have the same small-
deformation behaviour (3). Large deformation measurements, however,
indicate a difference in the structure between the fine-stranded gels formed at
low and at high pH.
In small-deformation measurements, the whole structure contributes equally
since no parts of the structure are broken by the small deformations. In large-
deformation measurements, the structure is broken and, more important, the
0.45
Strain at fracture
0.4
!
0.35
!
0.3
0.25
!
0.2
0.15
!
Ii i-1
0.1
1
0.05
0
2 3 4 5 6 7 8
a pH
kPa
20
Stress at fracture
18
!
16
14
i
12
10
i
8
II
!
6
4
I
I

0
2 3 4 5 6 7 8
b
pH
Fig. 6. Strain at fracture (a) and stress at fracture (b) for 12% 13-1actoglobulin gels.
347
M.Stading and A.-M.Hermansson
weaker areas of the structure are broken before the stronger parts. The weaker
parts may very well be highly elastic even if they fracture at low stress or at low
strain.
Stress . The stress at fracture ar shows a behaviour similar to the strain at fracture
(compare a and b in Figure 6); at is low at low pH and increases stepwise up to a
maximum at high pH. The decrease in at at pH> 6.0 coincides with an increased
critical gel concentration at high pH (3). The least concentration required to
form a gel at pH 7.5 was reported to be 11.5% wlw for a similar preparation of B.
lactoglobulin (3) . The deviations at pH 6.0 are larger than the surrounding
values, which shows that this gel structure is extra-sensitive to small variations in
the conditions during preparation.
The aggregate gels at pH 4.0-5.2 fractured at lower stresses than those at pH
5.5-6.0. This means there are differences in the gel structure within the pH
range in which aggregate gels are formed. This was neither demonstrated by the
strain at fracture nor by the small-deformation behaviour of the J3-lactoglobulin
gels (3). It is notable that the increase in fracture stress occurs close to the
isoelectric point of J3-lactoglobulin.
Langley et al. have studied the large-deformation properties of whey proteins
and found the fracture stress of a sample containing 14.1% wlw J3-lactoglobulin
to be 23 kPa at pH 6.5 and 26 kPa at pH 8.0 (4). The fracture stress of their
sample at pH 6.5 is about 60% higher than in the present study which could be
explained by the higher concentration. The fracture stress of their sample at pH
8.0 is, however, stronger than that at pH 6.5, contrary to the results in the
present study. An explanation of this is that the concentration of 14.1% wlw is
sufficiently higher than the critical gel concentration at pH 8.0 to form a strong
gel.
Energy. The specific fracture energy , R" shows a behaviour similar to both ar
and lOt. At pH < 4 R; = 0.1 J/m
2
, at pH between 4 and 5.2 R; = 0.6 J/m
2
and at
pH > 5.2 R; increases up to 6 J/m
2
The increase in R; at high pH starts at the
same pH as the stress at fracture does because of the influence of the stress in
equation (11).
Gel characteristics
Non-linearity. The stress-strain curve can be described by equation (3) as a =
ke
P
(21,22). A value ofp <1 gives an r-shaped curve andp >1 gives a J-shaped
curve . A linear or Hookean material has p = 1 and k equal to Young's modulus.
Gordon (18) suggested that materials exhibiting J-shaped stress-strain curves
resist cracks better than do linear elastic materials (18). This has been further
studied by Kendall and Fuller, who concluded that the non-linearity compared
to test geometry has little influence on the fracture properties (21). The
influence of the non-linearity on the fracture properties in tension tests depends
on the parameter RJ(lnE), where R; is the specific fracture energy, In is the
notch depth and E is Young's modulus. For lowvalues of R/(lnE), materials with
348
Fig. 7. The value of the exponent pin <T = ke" (equation 3).
a J-shaped stress-strain curve fracture at lower stresses than materials with r-
shaped stress-strain curves, whereas the situation is reversed for high values
(21 ).
Equation (3) has been fitted to the experimental results using linear
regression, and the value of the exponent, p, is shown in Figure 7. The results
show that the aggregate gels at intermediate pH have lower values of p, i.e. the
stress-strain curves are r-shaped. The fine-stranded gels both at low and high
pH are closer to being linear. The value of the exponent p is characteristic of the
type of gel structure. It does not , however, significantly influence the stress at
fracture of the gels in this study. This can be exemplified by the stress at fracture ,
which is much higher for the gels at pH > 6 than for the gels at pH < 4, although
they have similar curve shape (see Figure 6b).
The decreases of p at high pH depends on the sigmoidal shape of the stress-
strain curve, which does not fit equation (3) (ct . Figure 5). The sigmoidal curve
resembles that of rubber, and the properties are also similar to those of rubber
with high strain at fracture and high specific fracture energy.
Notch dependence. The notch dependence as described by equation (4) with e cc
l/a
m
is fitted to experimental results and shown in Figure 8. The samples with
pH 3.0 were too fragile for the notch-sensitivity to be measured. All the
measured fracture stresses of the samples with applied notches fitted well to
equation (4) whereas equation (5) predicted too high fracture stresses for the
samples with applied notches. The gels at all pH can consequently be concluded
to be notch-sensitive.
The largest structure element extracted from the notch-sensitivity curves by
far exceed the dimensions of the molecules or aggregates known from
microscopy (Langton and Hermansson, to be published). This could depend
on defects such as small air bubbles acting as notches, but the fracture surfaces
show few defects. Bubbles occur almost exclusively in the unnotched samples.
349
M.Stadlnl: and A.-M.Hermansson
kPa
50 Stress
40
\
\
30
\
20
10
0.05 0.1 0.15 0.2
Fig. 8. The notch dependence of 12% gels at pIl4.5 (+), 5.5 ( Li), 6.5 (0) and 7.5
(0) . The lines show the fitted curves of equ ation 4.
The radius of the bubbles was -1 mrn, and the reason for their occurrence was
probably the geometry of the sample moulds , where the bubbles were trapped
on the walls of the mould at the relatively narrow midsection (see Figure 4b) .
The bubbles cause stress concentration and decrease the measured fracture
stress, which in turn leads to a high value for the size of the largest structure
element. It is not possible to calculate the effect of the bubbles as the gels are
viscoelastic; the Inglis equation for the stress concentration, equation (1), can
only be used for elastic materials. If the gels are assumed to be elastic , equation
(1) gives about twice as high stre ss as the measured value. The visoelasticity
causes flow, which probably decreases the stress concentration and blunts the tip
of the notch (5) . Both these effects decrease the influence of the notch and lead
to smaller corrections of the stress at the notch than predicted by equation (1).
There could also exist density fluctuations in the gel network. The largest
structure element would then be equal to the size of the density fluctuations. The
technique has to be improved to measure the possible density fluctuations
accurately; shallower notches have to be made and all unnatural defects have to
be removed.
Elasti city. Young's modulus E of 12% w/w 13-lactoglobulin samples is shown in
Figure 9, where the storage modulus G' of 12% 13-lactoglobulinsamples (Sigma)
is also shown. Please note that E corresponds to 3G' on the scales. The
experimental details for the measurements of the storage modulus are presented
in reference 3.
The storage modulus G' was higher for the aggregate gels than for the fine-
stranded gels with a maximum at pH 5.5, as shown by the dashed line in Figure
9. Young's modulus E was slightly larger than the theoretically predicted 2(1 +
e)G (e is Poisson's ratio) even if the two extremes at pH 3.5 and 6.0 were
350
Deformation properties of 13-lactoglobulin gels
kPa
E
300
250
200
150
100
50
0
D-
2 3
kPa
100
80
60
40
20
0
4 5 6 7 8
pH
Fig. 9. Young's modulus E and the storage modulus G' of 12% 13-1actoglobuli n gels. G' is measured
as in reference 3.
excluded (18). Poisson's ratio is normally e = 0.5 for biological materials for
which a constant volume can be assumed. e > 0.5 would decrease the
discrepancy in this case but is unusual and not probable (18). The discrepancies
are more likely to have been caused by differences between the two 13-
lactoglobulin preparations and by the differing measuring techniques. Dynamic
rheological measurements only deform the samples within the linear limit, while
the samples in tensile tests are subject to higher deformations than in the
dynamic measurements even during preparation of the sampl es.
Young's modulus did not show clear differences between aggregate gels and
fine-stranded gels; instead E has two maximum values at pH 3.5 and 6.0, both
coinciding with the transitions between an aggregate and a fine-stranded gel
structure. The stress at fracture shows a similar behaviour (cf . Figure 6b). The
large deformation elasticity as expressed by Young 's modulus probably
measures the elasticity of the strong parts of the structure, since breakdown of
the weakest parts starts at as Iowa strain as 8 x 10-
3
(3). The storage modulus at
the small deformation G' , was measured within the linear limit , i.e . the structure
is intact and all parts contributes to G' . This hypothesi s leads to the conclusion
there are some highly elastic areas in the network of the gels at pH 6.0 and 3.5,
which are responsible for the high Young's modulus. These gels are also formed
at pH values that are intermediate between an aggregated and a fine-stranded
gel structure where fluctuations in the network structure may occur.
Conclusions
Small differences in the gel structure which are not seen by small-deformation
measurements can be detected by large-deformation measurements. The
transparent fine-stranded gel structure of exhibits completely
different large-deformation properties at low pH values than at high ones. They
are brittle at low pH and rubber-like at high pH. The behaviour of the opaque
aggregate gels also varies with the pH region in which they are formed. The
strain at fracture is constant , but the stress at fracture is higher above the
isoelectric point than below.
351
M.Stading and A.-M.Hermansson
The non-linearity of the stress-strain curve docs not significantly influence the
fracture properties of 13-lactoglobulin gels. It does, however , differentiate
between aggre gate gels and strand-like gels.
Gel s formed at a pH which lies in the transition regi on betwe en where
aggregate and fine- str anded gels ar e formed show a high Young' s modulus. At
the upper pH transition a maximum in fracture str ess also occurs. The Young's
modulus at other pH values is only slightly higher than 3G' in shear . The high
elasticity and fracture stress may depend on strong, elastic areas in the network
st ructure.
Acknowledgements
The authors thank Mr S.- E.Trank and A.Karlsson for their skilled technical
assistance in making accessories to the tensile test. We also want to thank the
Swedish Dair ies Associat ion for supplying us wit h the 13-laetoglobulin powder
and the Dani sh Government Res earch Institute for Dairy Industry, Hillered,
Denmark and Mats Obl ad , Chalmers University of Technology for analysing it.
The financial support from SJFR, The Swedi sh Council for Forestry and
Agricultural Research, is gratefully acknowledged.
References
1. Hennansson ,A.-M. (1988) In Mitchell,l . and Blanshar d,J. (cds), Food structure- its Creation
and Evaluation. Butterworths, Boston , p. 25.
2. Harwalker,A.M. and Kalab.M. (1985) Milchwissenschaft; 40, 665.
3. Stading,M. and Hennansson ,A.-M. (1990) Food Hydrocoll. , 4, 121.
4. Langley, K.R., Millard,D. and Evans, E. W. (1986) J, Dairy Res. , 53, 285.
5. Luyten,H. (1988) The Rheological and Fracture Behaviour of Gouda Cheese, PhD thesis,
Wageningen Agricultural University, Wageningen, The Netherl ands.
6. Yones e,M. , Baba.K. and Kishimoto,H. (1988) Bull. Chern. Soc. Jpn, 61, 1857.
7. McEvoy,H. , Ross-Murphy,S.B. and Clark ,A.H . (1985) Polymer, 26, 1483.
8. Myers,F.S. and Wenrick,J.D. (1974) Rubber Chern. Technol. , 47,1213.
9. BagleY,E.B. (1987) In Jowitt ,R., Escher,F. , Kent,M. , McKenna, B. and Roques,M.A. (eds) ,
Physical Propert ies of Foods . Elsevier, London , p. 345.
10. Inglis,C.E. (1913) Proc. Inst , Naval Architects London , 60, 219.
11. Gtiffith,A.A. (1921) Phil. Trans. Roy. Soc. London , A221, 163.
12. Wainwright ,S.A ., Biggs. W.D. , Currey,J. D. and Gosline,l.M. (1976) Mechanical Design in
Organisms. Ed ward Arn old, Lond on. .
13. Purslow,P.1. Mater. Sci. (in press).
14. Kelly.A. (1966) Strong Solids . Clarendon Press, Oxford.
15. Birchall ,J .D., Howard,A.J. and Kendall ,K. (1981) Nature, 289, 388.
16. Maub ois,J .L. , Pierre, A. , Fauquant,J . and Piot,M. (1987) Bull. IDF, 212,154.
17. Peleg,M. (1984) l. Text. Stud. , 15,317.
18. Gord on ,J. E. (1978) Structures. Pitman, Lond on.
19. Atk ins,G.A. and Mai,Y.-W. (1988) Elastic and Plastic Fracture. Ellis Horwood , Chichester.
20. Wonnacott ,R.I. and Wonnacott ,T.H. (1985) Introductory Statistics, 4th edn. John Wiley and
Sons, New York.
21. Kcndall ,K. and Full er ,K.N.G. (1987) 1. Phys. D: Appl. Phys. , 22,48.
22. Pur slow. P, (1989) l. Phys. D: Appl. Phys. , 22, 854.
Received on February 1, 1991; accepted on April 22, 199/
352

Das könnte Ihnen auch gefallen