You are on page 1of 22

G. Doxastakisand V.

Kiosseoglou(Editors)
Novel Macromolecules in Food Systems
9 2000 Elsevier Science B.V. All rights reserved. 309

Konjac Glucomannan

K. Nishinari

Department of Food and Nutrition, Faculty of Human Life Science, Osaka City University
3-3-138 Sugimoto, Sumiyoshi-ku, Osaka 558-8585, Japan

I. INTRODUCTION

Konjac glucomannan (KGM) is a polysaccharide which is extracted from the tuber of


Amorphophallus Konjac C.Koch. It forms a thermally stable gel in the presence of an alkaline
coagulant, and has been used in Japanese traditional dishes for a longtime [1]. Konjac gels are
boiled with vegetables, mushrooms, meat, surimi etc. A konjac gel has been a popular food
since it has a unique texture, and KGM is a dietary fibre and believed to be good for health.
It has recently attracted much attention because it forms a gel by mixing with xanthan or ~-
carrageenan in the absense of alkaline coagulant [2]. As a result of the fact that this gel can be
made at low pH, many kinds of dessert jellies containing various fruit juices have appeared in
the market in Japan. Finally, konjac gum has also been used as binding material in pet foods in
Europe.
In the present work, the physico-chemical properties and functionality of konjac
glucomannan are reviewed.

2. STRUCTURE AND M O L E ~ WEIGHT OF KONJAC GLUCOMANNAN

The konjac gtucomannan molecule consists of D-glucose and D-mannose linked by 13-1,4
linkage, and the ratio ofmannose to glucose has been reported as 1.6:1 [3], while there is some
branching at the C-3 of the mannose unit [4]. An acetyl group is attached to one per 19 sugar
residues [5].

2.1. The Man/Glc ratio


The ratio of mannose (Man) to glucose (Glc) for KGM samples with different molecular
weights, prepared by enzymatic degradation, was recently determined by HPLC [6].
M an/Glc ratio was calculated from the peak areas of mannose and glucose detected with a
refractive index detector, and was approximately 2.0 for all the fractions, which was higher
than 1.6 reported previously [3]. It was close to the ratio 2.1 reported for glucomannan
extracted from Scotch pine [7]. The experimental fact that the M an/Glc ratio was not much
different for a non-degraded specimen and for enzymatically degraded specimens suggests that
there is no block structure in konjac glucomannan because glucose and mannose residues do
310

not exhibit a difference in enzyme reactivity.

2.2. Fractionation
To obtain fractions with different molecular weights, konjac glucomannan powder was
dissolved in water. Methanol as a precipitant was added, and the solution was kept at 30~
Although a clear phase separation could not be achieved as has been done for man!r solutions
of synthetic polymers in organic solvents, cloud-like flakes appeared in the solution after one
day. The flakes were removed and methanol was added again. By repeating this procedure,
four fractions (F1-F4) were obtained [8].
Fractions with different molecular weights were also prepared by Shimizu Chemical Co.
(Hiroshima, Japan) using an enzymatic degradation method. The native and non-degraded
konjac glucomannan (hiD) were treated with an enzyme (SP-249, Novo Nordisk A/S,
Copenhagen, Denmark) for different reaction times at ambient temperature and four fractions
with low molecular weights LM 1-LM4 were obtained [9].

Table 1 Molecular weight of various KGM fraction samples


Sample Temp(~ Solvent Mw( • 105) Mw/Mn Mz/Mw
LM1 23 Cadoxen 2.56 6.50 5.04
LM2 23 Cadoxen 4.38 7.72 5.06
LM3 23 Cadoxen 4.44 7.09 4.30
LM4 23 Cadoxen 5.96 5.68 3.85
ND 23 Cadoxen 6.89 8.02 4.44

2.3. Molecular weight


The molecular weight of each KGM fraction was determined by (]PC at room temperature.
KGM was dissolved in cadoxen to obtain a dilute solution because of its very low solubility in
aqueous solutions [1 ]. Cadoxen is known to dissolve cellulose and to give a clear, coloufless
and stable solution [10]. A colourless solution allows the use of an RI detector [ 11 ], and the
much lower viscosity of the cadoxen solution compared to that of aqueous solution is a great
advantage for GPC analysis. The peak corresponding to the highest molecular weight fraction
(8.53 X105) of pullulan appeared at 11.73min and that corresponding to the lowest one (5.80
• 103) at 15.07min. The standard pullulan samples [12] gave a good linear calibration curve
(data not shown). The logarithm of molecular weight calculated from a regression analysis is
shown in Figure 1 (lower horizontal axis). The weight average molecular weight (Mw, the
fourth column of Table 1) for five KGMs decreased in the order ofND, LM4, LM3, LM2 and
LM 1 which was the same order of the reaction time with the enzyme. The molecular weight
distribution was not narrow according to the chromatograms, as shown in Figure 1, and the
ratios of Mw to number average molecular weight (Mn) and that of z-average molecular weight
(Mz) to Mw are shown in the fifth and sixth columns of Table 1, respectively. The elution
curve for ND in Figure 1 showed the second peak in the lower molecular weight region. The
sample did not flow homogeneously through the columns, due probably to the high viscosity.
This behaviour was typically observed in a highly viscous solution and the true average
311

molecular weight could be higher. Even in cadoxen, which forms a KGM solution with far
lower viscosity than in water, it was difficult to prepare a suitable solution for GPC analysis,
especially in the case of the highest molecular weight fraction. Native (non-degraded) KGM is
known as a gum with very high viscosity and it is difficult to determine the molecular weight
of native KGM by GPC. As shown in Figure 1, the ND sample contains molecules with

0.6
....... LM1 /~
..... u~ /",;.~7"',,
- .... LM3 /,," \~, ',

~" 0.4 ------ND ..~~ ', ',,

/i,.'/ i V,~"
"~
~,
0

"i 0.2 ',"

, . ~\l

0.0 , , L~;.,
10 12 14 1"6
Elution Time (min)
1()7 1'0+ 1~ I'0 4.
log mw

Figure 1. Elution curves for various K GM solutions in cadoxen. The lower horizontal axis
shows the logarithm of molecular weight calibrated by eight pullulan samples. A relative
amount was calculated as the ratio of the RI intensity at each time to the accumulative value.
(From ref.9)

molecular weights higher than 1.0 X 10 6, but no lower than 1.0 X 10 4 molecular weight fraction.
A similar wide distribution of molecular weights was also reported for aqueous konjac mannan
[13]. It ranged from 4.0 X 10 4 to more than 1.0 X 10 6 and was almost the whole range
separable for their GPC gels (4.0 X 104 - 2.0 • 107). The enzyme treatment increased the ratio
of lower molecular weight polymers, however, the fraction of higher than 1.0 X 10 6 molecular
weight still remained even in LM 1. The exclusion limit for the GPC column was assumed to
be 5.0 X 107, because there are are no standard polymers with higher than 1.0 X 106 molecular
weight. It is difficult to assign a molecular weight for the earlier eluted fraction. High accuracy
for the molecular weight obtained from GPC, therefore, should not be expected. It is certain
however, that the KGM fractions have different molecular weights.

3. SOLUTION PROPERTIES OF KONJAC GLUCOMANNAN

3.1. I n t r i n s i c v i s c o s i t y
The intrinsic viscosity was determined in water and in 4M urea for the four F l-F4 fractions,
and the results are shown in Table 2 [8].
A fraction with a high molecular weight began to deviate from the straight line in the plot of
312

the reduced viscosity vs konjac glucomannan concentration at a lower concentration than a


fraction with a lower molecular weight. This has also been observed for puUulan [12] and
many other polymers.

Table 2 The intrinsic viscosity ofkonjac gl_u_comann_an_at 25.4~ . . . . . . . . .

Sample I in water - ] ...... in. 4M


4. urea
bF .[ . J . gJ 15.8
F1 19.2 21.8
F2 18.6 20.4
F3 13.6 16.4
F4 6.7 8.0

The intrinsic viscosity observed in 4M urea was higher than that observed in water for all
the fractions. The conformation of konjac glucomannan molecules may be more expanded in
urea because hy drogen bonds between hydroxyl groups in konjac glucomannan molecules are
broken.

3.2. Zero shear specific viscosity


The relationship between the logarithm of the zero shear specific viscosity (~sp0) and the
logarithm of KGM concentration (c) is shown in Figure 2. Log r/~poincreased linearly with

o
o

9 unF
9 unF o
o
oF2 oF2
*F3
~F3
o Ft.
6 o Ft.
o 9

o
o
o %
o
o
o
%
oo
o o
I o~ I
O o O
e o
o
o o*ae
o $ o

D
o
o
oa~ s o
o~,
o 9
o
oa
a a e
Ae ~~
o

caD

9 | , , _ L . . . . . I

-1.0 -0.5 0 -o~ - o o.s ~.o


log c log ctrl]

Figure 2. Concentration dependence of the Figure 3. Dependence of the zero shear


zero shear specific viscosity (rhpo) of konjac specific viscosity (r/sp0) of konjac gluco-
glucomannan solution. UnF(unfractionated mannan solution on the coil overlap
material, O), F2(O), F3(A), F4(~). (From parameter c[ r/]. Symbols are the same as
ref.8) Figure 2. (From ref.8)
313

increasing log c when the value of log 79~p0was lower than about 1. The double logarithmic
plots of the zero shear specific viscosity 79sp0against the coil overlap parameter (c[ 79]) are
shown in Figure 3. For dilute solutions, slopes of the plots were close to 1.4 for all the
fractions as observed for many polysaccharide solutions. It was not possible to obtain a clear
inflection point of the curves in Figure 3, because the low solubility of konjac glucomannan
made it difficult to prepare solutions which exhibited large c[ 79 ] values. However the
inclination of the slope of the curves increased gradually with increasing log c[ 79], suggesting
that significant coil overlap and entanglement had already started when c[ r/]> 1. The onset of
coil overlap occurs at lower concentrations for KGM molecules than for other polysaccharides.

3.3. Dynamic viscoelasticity of KGM dispersions


The frequency dependence of the storage shear modulus, G', and the loss shear modulus, G",
for KGM dispersions of different concentrations is shown in Figure 4. The behaviour of KM

103 I ' '


c.(%)

10 z . t KM plus water
~9 0 . 3 5 % G'
---0-- 0.35% 6"
0 . 7 0 % G'
-0- 0.70*,6 G"
--9 1.05% G'
t~ 10~
&. I .O5% G"
.1" 1.40% G'

1.40% G"
-1o o

10 "I

10. Z l - - I ! I I
1 0 "~ 1 0 .2 . 1 0 -1 10 o 101

co / r a d . s ~

Figure 4. Frequency dependence of G' and G" of KGM aqueous dispersion of various
concentrations. (From ref.46)

plus water is typical of a concentrated polymer solution. In concentrated polymer solutions


the response is liquid- like, i.e. G" is larger than G' and both moduli increase with increasing
frequency at lower frequencies, whilst the behaviour approaches that of solid-like materials, i.e.
G' is larger than G", and both moduli become frequency independent at higher frequencies.
The molecular chains can disentangle and rearrange during the long period of oscillation, so G"
is larger than G' at lower frequencies in concentrated polymer solutions. At higher frequencies,
molecular chains cannot disentangle during the short period of oscillation, so G' is large than
G", because the entanglement points play the role of a temporary cross-linkingjunction zone.
314

Moreover, the crossover frequency of G' and G" shifted, as expected, to lower frequencies
with increasing KM concentration.

3.4. Other solution properties


The partial specific volume of konjac glucomannan was determined as a function of pH by
density measurements [14] and showed a steep rise at around pH=l 1.5 and 3. This su/~gests
that the conformational change is necessary for gel formation of konjac gtucomannan. Light
scattering measurements were carried out for methylated konjac glucomannan [ 15].

4. GELATION OF KGM

4.1. Gelation kinetics of KGM with different molecular weights [16]


Figure 5 shows the time evolution of the storage shear modulus, G', and the loss shear

nt• 3
o ,2

c~
Figure 5. Time dependence of G', G" and
13_ tan 6 for 2% aqueous K GM dispersions
~2
(5 with different molecular weights, r-l:
c

o
LM1, ~ : LM2, A : LM3, • :LM4.
1 Measurement temperature:60 ~
Symbols represent the experimental
-4
values and the solid lines represent the
0 10 20 30 40 50 60 calculated curves. (From ref. 16)
Time/rain

~tP
%
o %
• ~
~o ~ 1 7 6 1 7 6 1 7 6 1 7 6 1~~176176
76176176176176176176176
C 0.1

0.01
~b 2'o a6 4"o s'o
Time / rain
315

Table 3 Parameters of the first order kinetic model for the gelation of 2% KGM dispersions
with different molecular weights (G',~t the plateau value of G' after a long time; k, the rate
constant of gelation of KGM; to, the gelation time; r, correlation coefficient)
Sample G',., (Pa) k(min) to (min) r
LM1 41 0.121 11 0.978
LM2 453 0.248 7 0.976
LM3 562 0.228 6 0.973
LM4 2300 0.214 5 0.942

modulus, G", for 2 % dispersions of KGM with different molecular weights in the presence of
20/z 1 of 1M sodium carbonate at 60~ and at a constant frequency 1 rad/s. G' and G"
increased with time and attained a plateau value, while the mechanical loss tangent tan 6
=G"/G' value decreased. The plateau value of tan 6 for LM 1 was higher than 0.1, and the 2%
LM 1 dispersion formed a weak gel, while the values for 2 % LM 2, 3 and 4 dispersions were
lower than 0.1 and these 2 % KGM dispersions formed elastic gels. The plateau values of G'
and G" at time t=60 min increased with increasing molecular weight. A number of gelation
processes at a constant temperature have been treated by an equation of first order kinetics or
other modified equations [17]. These curves could be well approximated by the following
equation: G(t)=G'sat[1-e'k(t't~ where G'sat is the plateau value of G' after a longtime, k is the
rate constant of gelation of KGM, and to is the gelation time. The gelation time to is taken as
the time at which G' begins to deviate from the baseline. The baseline is defined as 10~ Pa of
G' since it is the lower limit value for G' detection and the reproducibility of G' is satisfactory
above this baseline. G'sat and k were determined from the best fitting of the experimental and
calculated values of G'. Symbols represent the experimental values. The solid lines represent
calculated curves. The gelation time to became shorter and the rate constant k and the plateau
value G'sat increased with increasing molecular weight (Table 3). Since the molecular weights
of LM2 and LM3 were not much different, the rate constant k and the plateau value G'sat
were not different either. Generally, the rheological properties of gels depend on the molecular
structure of the gelling agent. However, the dependence of elastic modulus of gels on
molecular weight has not been studied extensively. The dependence of the elastic modulus of
gels on the molecular weight has been studied for gelatin [18], agarose [19], kappa-carrageenan
[20], casein micelles [21], and methyl cellulose [22]. For some polymer gels, the elastic
modulus was found to increase with increasing molecular weight and to become independent
above a certain molecular weight [23]. The probability for junction zone formation increases
and the flexible molecular chains connecting the junction zones will become longer with
increasing molecular weight [24]. Therefore, the number of elastically active chains increases
and hence the elastic modulus of gels increases with increasing molecular weight. If this
picture is valid for konjac gels, both the breaking stress and breaking strain will also increase
with increasing molecular weight. Because of the limited quantity of the specimens with
different molecular weights, it was not easy to perform the large deformation rheology,
however, this should be clarified in the near future.
316

4.2. Gelation kinetics of KGM at different concentrations


Figure 6 shows the time evolution of G' and G" of LM1 dispersions of various
concentrations in the presence of 20/1. I of 1M sodium carbonate at 65~ and at a constant

0
0 10 20 30 40 50 60 70 80 90

Time/min

Figure 6. Time dependence of G' for aqueous LM 1 dispersions of different concentrations.


~" 1%, [S]'1.5%, A:2%, )<:3%. Measurement temperature:65~ Symbols represent the
experimental values and the solid lines represent the calculated curves. (From ref. 16)

Table 4 Parameters of the first order kinetic model for the gelation of LM 1 dispersions of
various concentrations (abbreviations as in Table 3)
....Sample(%) G'~, (Pa) k(min) to (min) r
1.0 19 0.019 47 0.798
1.5 27 0.133 10 0.976
2.0 185 0.149 7 0.995
3.0 229 0.296 4 0.970

frequency 1 rad/s. The gelation time, to, became shorter and the rate constant k and the
plateau value G'sat increased with increasing concentration (Table 4). Many researchers
reported that G' of biopolymer gels (e.g gelatin, a ~ , casein micelle) showed a square power
law in concentration dependence [17]. The concentration dependence of the elastic modulus
ofbiopolymer gels was discussed by Clark and Ross-Murphy [25] using a cascade treatment
and by Oakenfull [26] using a modified theory of rubber elasticity. Figure 7 shows a double
logarithmic plot of G'sat versus the concentration, c. The exponent m in the relation between
G'sat and the concentration, c, G'~c-cm, for konjac glucomannan in the concentration range
studied, was about 3. Since both theories of Clark and Ross-Murphy [25] and Oakenfull [26]
predict a larger exponent at a lower concentration range and a smaller one at a higher
concentration range, this should be further studied for a wide concentration range. The
exponent m will also depend on the molecular weight. It is necessary to examine this point
using many fractions with different molecular weights.
317

~4
~3
0 .......... 0
2

0
0 0.1 0.2 0.3 0.4 0.5

log1 o C/%

Figure 7. Double logarithmic plot of G'sat of the LM l dispersion and the concentration, c.
(From ref. 16)

4.3. Gelation kinetics of KGM at different heating temperatures


Figure 8 shows the time dependence of G' and G" for 1% LM2 dispersions at different
heating temperatures in the presence of 20 lz I of 1M sodium carbonate at a constant

3
t~
s

(5 z
o

o r"~ ~.xa, ~ , , L ~
0 50 100

Time / min

Figure 8. Time dependence of G', G" for 1% aqueous LM2 dispersions at different heating
temperatures. ?-1:50~ A:60~ X:70~ *:75~ O:80~ Measurement temperature:60~
Symbols represent the experimental values and the solid lines represent the calculated curves.
(From ref. 16)
318

Table 5 Parameters of the first order kinetic model for the gelation of 1% LM2 dispersions at
different heating t, mperatures (abbreviations as in Table 3)
TemPerature(~ G'~ (Pa) k(min) t9 (min) r
50 35 0.082 38 0.981
60 311 0.120 12 0.970
70 420 0.098 6 0.969
75 910 0.158 1 0.975
80 ,,
115 0.691 1 0.992

frequency 1 rad/s. The gelation time, to, became shorter and the rate constant, k, increased
with increasing heating temperature. The plateau value, G'sat, increased with increasing heating
temperature up to 75~ and then decreased at 80~ (Table 5). Figure 9 shows G' and G" for
2 % LM2 dispersions as a function of time at different heating temperatures from 45 to 80~
The gelation time became shorter and the rate constant k increased with increasing heating
temperature. The plateau value, G'sat, increased with increasing heating temperature up to
50~ and then decreased. The value of G' increased quickly and decreased at 75~ and 80~
(Table 6). Generally, the gelation proceeds faster at higher temperatures for heat-setting
4

3E' C
o

4 ~ - L . . . . . 9 L

_.o

0
0 50 100
Time / rain

Figure 9. Time dependence of G', G" for 2% aqueous LM2 dispersions at different heating
temperatures. O :45 ~ I--1:50~ A :60~ X :70 ~ *:75 ~ O'80~ Measurement
temperature:60~ Symbols represent the experimental values and the solid lines represent the
calculated curves. (From ref. 16)
319

Table 6 Parameters of the first order kinetic model for the gelation of 2% LM 2 disp ersions at
different heating t, mperatures (abbre.viati0,ns as in Table 3)
Temperature(~ G',., (Pa) k(min) to (rain) r
45 2370 0.090 27 0.'969
5O 4418 0.110 8 0.992
60 453 0.248 7 0.976
70 171 0.532 5 0.908
75 1522 0.641 2 0.953
80 614 1.260 1 0.964

systems such as methyl cellulose [22,27,28], xyloglucan [29,30], curdlan [31,32], gly cinin, and
/3-conglycinin [33-35]. The reason why the plateau value, G'sat, decreased with increasing
temperature beyond a certain temperature may be attributed to one of the following
possibilities: 1)The KGM dispersions form gels so rapidly at higher temperatures that a
disordered network is formed and weak gels result. 2) The slippage between the plate and
sample solutions may occur during the measurement because of the exuded water from KGM
gel. This should be explored in the future.

4.4. Gelation kinetics of KGM at different alkaline concentrations


Figure 10 shows G' and G" for a 1% LM 1 dispersion as a function of time at 60~ in the
presence of alkaline coagulant with different concentrations at 65~ and at a constant
frequency 1 rad/s. The pH value of 1% LM 1 dispersion was 5.84 at the time t=0. When 20
/1. 1Na2CO3 (0.5 M, 1.0 M, 1.5 M, and 2.0 M) was added to KGM dispersions, the values of
pH increased gradually and became 10.56, 10.75, 10.92, and 10.98, respectively, after 30 mins.
The G' and G" values for KGM dispersions with 1.0 M, 1.5 M, and 2.0 M Na2CO3 increased
with time and attained plateau values faster than those for a KGM dispersion with 0.5 M
Na2CO3.

Table 7 Parameters of the first order kinetic model for the gelation of 1% LM 1 dispersions
with alkaline co. ant of different concentrations (abbreviations as in Table 3)..
Sample G'~, (Pa) k(min) to (min) . r
0.5M 251 0.059 13 0.964
1.0M 520 0.167 4 0.985
1.5M 634 0.171 4 0.976
2.0M 98 0.206 2 0.994

The gelation time, to, became shorter with increasing alkaline concentration. The p lateau value,
G'sat , reached a maximum at 1.5 M Na2CO3 (Table 7). The reason for that may also be
attributed to one of the two possibilities mentioned above. In addition to this, the possibility
of molecular degradation at high p H should also be taken into account. The fact that the
rapidly formed gels show a smaller elastic modulus than those formed slowly has also been
observed for many other gels such as rennet-induced casein gels [36,37], gellan gum gels [38],
and amylose gels [39]. If this phenomenon is induced by the slippage in the shear mode
320

viscoelastic measurements, this should be checked by adopting indentation or other methods


which are free from slippage. If it is caused by the formation of the disordered structure in a
rapid gelling process, the structural observation should be useful.

z
m oO o o o ~
fl_
o~

~9
o
o3'
o

It
9/- P ,
,/o..
L~ I l i

I1.
=--
(.9 -o~176176176176162
o
oO~ o oov
o | 66 r - ~ ~ ~176176176
~• ~^^ o.... ~
x o o~
xX o oo
x ~Qo ooo

o~ oo
~x.-~ . ~
0 20 4'0 50 60
Time / min

Figure 10. Time dependence of G', G" for 1% aqueous LMldispersions with alkaline
coagulant of different concentrations. ~:0.5M, V]- 1.0M, A:I.5M, X :2.0M. Measurement
temperature:60~ Symbols represent the experimental values and the solid lines represent the
calculated curves. (From ref. 16)

5. MIXTURE OF KONJAC WITH OTHER POLYSACCHAR]DES

5.1. Konjac/xanthan mixtures


Since the discovery of the synergistic effect of xanthan and locust bean gum, there have been
many investigations on the interaction between two different polysaccharides. Individually
none of either konjac or xanthan aqueous solution forms a gel at a neutral p H and room
temperature, however, on mixing a gel can be produced. Williams and his co-workers studied
the interaction of these polymers, and found that the mixture on heating above 55~ formed a
gel. This temperature is lower than the conformational transition temperature of xanthan
molecules (Figure 11). It was concluded, therefore, that the junction zones in mixed gels are
formed by the association of konjac and the helical form of xanthan [40]. The effects of salts
were also examined, and it was found that their addition shifted the gelation temperature of the
321

mixture to lower temperatures. This was attributed to the self-association of xanthan


molecules rather than the association of konjac and xanthan (Figure 12)[41 ].

e--~-.e."

2130
if,
Et

"""~K3."h. "*'--4~
0
1 . . . . . . . 1 '0 20 3O aO 50 6O 7O aO

Frequency / Hz Temperature fC

Figure 11. Storage modulus as a function of Figure 12. G' as a function of temperature
frequency for xanthan in 0.04 mol/dm3 N aCI for xanthan-KM mixtures (1:1, 0.6 wt%) in
in the presence of gtucomannan measured at the presence and absence of monovalent
25~ after heating to 25 ~ (C)), 35 ~ (+), cations (0.04mol/dm3). H20(O); NaCI(C));
45~ 55~ 65~ (Cooling rate, KCI(I--1); CsCI(A); NH4CI(~ ) (]2rom ref.
approx. 1~ (From ref. 40) 41)

5.2. Konjac/~ -carrageenan mixtures


Although ~ -carrageenan can form a gel, the mixture of konjac and ~ -carrageenan has been
used to produce dessert jellies containing fruit juices in Japan. The texture of the mixed gel is
more rubber-like compared to that of ~-carrageenan gels. Williams et al observed two
exothermic peaks in cooling D SC curves for mixtures of konjac and ~-carrageenan with the
mixing ratio KGM/CAR from 0.1/0.5 to 0.2/0.4 (total polysac~haride concentration 0.6wt%)
[41 ]. Only one exothermic peak was observed for mixtures with CAR content below 0.3/0.3,
and this was attributed to the formation of an ordered structure by the interaction between
konjac glucomannan and ~-carrageenan. When the CAR content became higher than 0.45%,
the lower temperature peak began to appear in addition to the exothermic peak originating
from the formation of the ordered structure by the interaction between konjac gtucomannan
and ~-carrageenan. Exothermic enthalpy per unit mass of ~-carrageenan accompanying
gelation of the mixtures was significantly smaller than that for ~;-carrageenan alone, indicating
that the gel formation of ~-carrageenan is strongly affected by the presence of konjac gluco-
mannan. ESR spectra for mixtures and konjac glucomannan alone showed that the segmental
motion ofkonjac glucomannan is reduced by the interaction with ~c-carrageenan molecules.
322

O.02mW

(a)

30 40 50 60 70 80

Temperature/~ (b)

Figure 13. DSC cooling curves for various Figure 14. A model for the mixed gel of
ratios of CAR/KGM mixtures in the presence KGM/CAR. Thick curves represent CAR
of 50 mM KCI (0.6% total polymer while thin stand for K GM. Closed ovals
concentration). Scanning rate was 0.1 ~ represent CAR junction zones and open
CAR/KGM: (A)0.1/0.5,(B)0.2/0.4,(C)0.3/0.3, ovals stand for weak C AR-KGM junction
(D)0.4/0.2,(E)0.45/0.15,(F)0.5/0.1,(G)0.6/0 zones. (a) low Mw KGM/CAR, (b)/high
(From ref. 42) Mw KGM/CAR (From ref. 6)

Kohyama et al [6]carried out a large deformation extension measurement on mixed gels of


konjac gtucomannan with different molecular weights and ~-carrageenan molded into a ring
shape, and found that the breaking stress increased with increasing molecular weight of KGM
indicating that KGM chains contribute to the network structure. Since the gel-to-sol
transition temperature of mixed gels did not depend so much on the molecular weight of KGM,
it was suggested that the KGM chains interact weakly with CAR, and this can contribute to
strengthening of the network mechanically but does not affect its thermal stability to a great
extent (Figure 14).

5.3. Konjac/gellan mixtures


The rheological properties of mixtures of konjac glucomannan with different molecular
weights and gellan (GELL) were also examined [43]. The elastic modulus of mixtures of geUan
323

with a lower and a medium molecular weight KGM as a function of mixing ratio exhibited a
maximum at a certain KGM content whilst that with higher molecular weight KGM increased
with increasing KGM content. The relaxational strength of mixtures, defined as the difference
between the loss shear moduli as a function of temprature at a lower temperature and at a
higher temperature, was smaller than that of gellan gum alone, indicating that the gel formation
of gellan is strongly affected by the presence of KGM. The decrease in the relaxational
strength was more significant in mixtures with higher molecular weight KGM indicating that
the higher molecular weight KGM inhibits the gel formation of gellan. The storage modulus of
mixtures of gellan and KGM with a medium molecular weight showed a maximum at the
mixing ratio GELL/KGM=0.3/0.5. The effect of sodium chloride and calcium chloride on this

(a) i Co)

~ 2s

'T
15
0
CaCI~ C o n c . / nxmol-I "x
8 0 2 4
CaCI 2 C o n c .
6
/ mmol-I -t
8

Figure 15. Dependence of the midpoint temperature of transition TM (a) and the relaxation
strength A G (b) for mixtures with GELL/KGM = 0.3/0.5 (total poly saccharide concentration
0.8%) or 0.3% or 0.8% GELL solutions on the CaC12 concentration: (O) mixture; (A) 0.3%
GELL; (7-1) 0.8% GELL. (From ref.44)

mixture was examined by rheology and DSC. The midpoint transition temperature shifted to
higher temperatures with increasing concentration of sodium chloride or calcium chloride,
whilst the relaxational strength showed a maximum at a certain concentration of calcium
chloride [44]. The sodium or calcium ions shield the electrostatic charge of the carboxyl
groups of gellan, and promote the self a~egation of gellan and network formation by the
attachment of KGM molecules on the surface of gellan aggregates. Excessive salt addition
promotes the self aggregation of gellan and hinders the attachment of KGM thus leading to
phase separation [45].

5.4 Konjac/starch mixtures


When the konjac was added to a starch dispersion, both the elastic modulus and the breaking
stress of the resulted gel increased. However, following storage for a long time (two weeks),
324

the starch gel containing konjac showed smaller values for these parameters. The
retrogradation ratio A H2/AH], the re-gelatinisation enthalpy observed in the 2nd run DSC
heating curve divided by the gelatinisation enthalpy observed in the 1st run D SC heating curve,
showed the same tendency to rheological observation. Therefore, the addition of a small
amount ofkonjac to starch promotes the retrogradation for short storage, however, it prevents
the retrogradation after long storage [46]. The addition of a small amount of konjac to the

0.6 " ' " ' " I . . . . I . . . . I ' ' " ' I ' ' " ' 0.6 . . . . ! 9 9 "" I 9 ~ ' 9 i . . . . 'l-" 9 9 9

CS and KM mixture
C S plus water
O.S O.S
CS/KM
---0-- 3.5096 9.9/0.1
O.4 _E 0.4 3.S096 9.7S/0.2S
.E ~. 3.S0% 9 / I , 812, 7/3, 6/4
..C
v
<~
.~ 0.3 0.3
tr)
o~ .__

9
L~
CD 0.2 I,,_
0.2
c-
c.(%) t--
CO o.~ ;3.50
>"
C/') o.1
.15
--e-- 2.80
0 2.45
v 2.10
. . . . ! . . . . i . . . . i . . . . I . . . .
. . . . !. . . . . I . L i . I . . j . I . . . .

0 5 lO 15 zo 25 5 10 15 20 25

Storage time /day Storage time / day

Figure 16. Syneresis of a corn starch (CS) dispersion and a CS/KM mixture. C, concentration
of corn starch: 0 , 3.50% wt%; II, 3.15 wt%; O, 2.80 wt%; A, 2.45 wt%; V, 2.10 wt%.
CS~M mixing ratio: O, 3.50 wt% 9.9/0.1; D, 3.50 wt% 9.75/0.25; O, 3.50 wt% 9/1, 8/2, 7/3,
6/4. (From ref.47)

gelatinised dispersion of starch was found to be effective in preventing the syneresis and the
consequent phase separation as shown in Figure 16 [47].

6. DIELECTRIC, VISCOEIASTIC AND BROAD-IANE NMR S T U D I ~ OF KONJAC


GLUCOMANNAN FILMS

6.1. Biodegradable films


The widespread use of synthetic polymer films has caused serious polution problems and as
a result of this, biodegradable films have attracted much attention. KGM films were prepared
by a conventional casting method and the dielectric, viscoelastic and broad-line NMR
measurements were carried out to elucidate the relationship between the chemical structure of
KGM and its physico-chemical properties. The molecular motion of solid KGM is discussed
in comparison with amylose [48], pullulan [49], dextran [50] and cellulose derivatives [51].
325

6.2. Material
KGM isolated from the tuber of Amorphophallus konjac K. Koch was kindly supplied by
Dr Maekaji (Hiroshima Food Research Institute). It was dispersed in water and reprecipitated
by methanol. The ratio of D-mannose to D-glucose of this sample as determined by gas
chromatographic analysis on acid hydrolysate of KGM was 63:37, which agreed fairly well
with previous reports [3]. The content of acetyl groups must be negligible because the signal
of the carbonyl carbon was not perceived at 180 ppm in the CP/MAS NMR spectrum. KGM
was dissolved in water and cast into a film of about 100m thickness for dielectric, viscoelastic
and NMR measurements.

6.3. Measurements
Didectric studies :Electric fields of 10Hz and 7 V were applied to the silver electrodes,
evaporated onto the film (1.0 X 2.0cm). The film was dried in nitrogen gas by heating at 100,
120, 140, and 160~ for 40min after reaching the respective temperature. The temperature
was raised from the liquid nitrogen temperature at the rate of 2~ The complex dielectric
constant e *= e '- i e " was measured using a Rheolograph-Solid from Toy o-Seiki-Seisakusho
Ltd (Tokyo) by a conventional method described previously [47].
Viscoelasticity : The complex viscoelastic constant c* = c' + ic" at 10Hz was determined by
detecting the sinusoidal strain and stress at both ends of the film (0.15 X 1.8cm). The heating
rate and drying conditions were the same as in the dielectric measurements.
NMR : The powdered sample was put in a glass tube, which was then evacuated at 120~
for lh and sealed. Broad-line 1H-NMR derivative spectra were obtained by means of a
continuous wave method with a JEOL-JNM-W-40 spectrometer operated at 40 MHz as a
function of temperature from-150~ to 100~ and the second moments were calculated from
the spectra [52]. The field modulation frequency was 35 Hz and in most cases the amplitude
was kept low enough to neglect the modulation broadening A radio frequency fidd below the
limit of saturation was chosen by comparing the curves at different power levels.

6.4. Temperature dependence of the dielectric coefficient


Figure 17 shows the real and imaginary parts of the complex dielectric coefficient, ~' and ~ ",
for the glucomannan film as a function of temperature for various moisture levels. At low
temperatures, ~ 'was almost independent of the moisture level. The real part, ~ ' increased
gradually with increasing temperature, while the imaginary part, ~ " of glucomannan film,
heated at 100~ for 40 min (solid curve), showed a large peak at about -50~ This peak
became smaller and shifted to higher temperatures with decreasing moisture level. The peak
disappeared when the film was further dried at 100~ Therefore, it is considered to be due to
the presence of water. When the film was further dried at 120~ for 40 min. a new peak for
" appeared at about -100~ The intensity of this peak increased with decreasing moisture
level. This peak was attributed to the rotational motion of the hydroxymethyl groups
attached to the C-5 atom in glucose and mannose residues. The dielectric relaxation strength
A ~ was determined from the difference between the asymptote at the lower temperatures
and ~' of thoroughly dried sample at the higher temperature. The A ~ (ca. 2.2) of KGM is
326

smaller than that of amylose (3.0) [48] but larger than those of pullulan (1.5) [49] or dextran
(0.83) [50](Table 8). It is considered that the magnitude of A a is mainly determined by the
rotational motion of hydroxymethyl groups because the sample has no acetyl group. The
number of hydroxymethyl groups in KGM and in amylose is considered the same, while the
number in pullulan is about two thirds of that. Dextran has very few hydroxymethyl groups.

16-
12-

O,__ I I I I I I

_~L, (~ (~ , ,
00-150-1 0'-50 50 100
Temperature /~

Figure 17. Temperature dependence of the dielectric coefficients ~' and ~ " for various
moisture levels: ( ~ ) heated at 100~ ( - e o o - - ) heated at 120~ (.......... ) heated at 140~
(--AAA--) heated at 160~ Heating time: 40 min. The asymptote at the lower temperature in
the upper graph is drawn to estimate the dielectric relaxation strength.

Table 8 Dielectric and viscoelastic characteristics of KGM and other poly saccharides in solid
state
Polysaccharide Temp. of max ~ " A~ Temp. of max c" C'/109 at -150~
KGM -100 2.2 -100 7.2
Dextran -120 0.83
Pullulan -90 1.5 -100 7.8
Amylose -75 3.0 -75 8.4
HEC MS2.5 -128 -132 4.8
HPC MS2.4 -122 -126 . .
4.4. .
327

Therefore, the order of magnitude of the dielectric relaxation strength for these three
polysaccharides, amylose, pullulan and dextran, can be understood in terms of the content of
this group as discussed previously [50] (Table 8). According to these experimental results,
the rotational motion of the hydroxymethyl groups in KGM is slightly hindered in
comparison to the motion of the hydroxymethyl groups in amylose

6.5. Temperature dependence of the viscoelastic coefficient


The temperature dependence of the viscoelastic constants of KGM is shown in Figure l 8.

10
Q_

t::, 4

I I I I I I I
..
.." ".

1.5

o .." /
n 1.0

-"-0.5

0
-2oo -1 o -l&O -io & 5tO 160 "i,50
Temperoture /~

Figure 18. Temperature dependence of the elastic coefficients c' and c" at various moisture
levels: ( ..... ) non-dried; ( - - ) heated at 100~ ( - o o o - - ) heated at 120~ ( ) heated at
140~ (--AAA--) heated at 160~ Heating time: 40 min

The real part decreased monotonically with increasing temperature. The value of c" for the
non-dried sample showed a large peak at about -50~ and this peak decreased in height when
the sample was dried just as in the case of the dielectric coefficient results. For dry KGM, a
peak appeared at about- 100~ which is attributed to the rotation of hydroxymethyl groups
attached to the C-5 atom in mannose and glucose residues. Kimura et al. [53] also found a
mechanical loss peak at -100~ for pine glucomannan using a torsion pendulum method and
attributed this to the rotational motion of the hydroxymethyl groups. As the moisture
content decreases, the value of c' became larger at temperatures below about -20~ but it
became smaller at higher temperatures The water plays the role of a plasticizer at
temperatures higher than the ambient temperature
328

6.6. Temperature dependence of the second moment of NMR


The narrow component appeared superimposed on the broad one at all the temperatures and,
in addition to the other narrow components, it appeared at higher temperatures in the
observed derivative spectra of KGM. The narrow component which appeared in amylose,
pullulan, and dextran was attributed to a small amount of water since specimens left in air
showed a steep rise in the narrow component [51 ]. However, the narrow component in the
spectra of glucomannan was not influenced so much by the drying conditions. This must be
attributed to other factors than the presence of water, but the reason is not dear at present.
Second moments of KGM together with those of amylose, pullulan and dextran are shown in
Figure 19 as a function of temperature. The second moment of KGM at lower temperatures is
smaller than that of the other three polysaccharides. The second moment decreased rapidly at

22
O o

20 ..O
\
.x
18 .~. o 0"%,

16
0
'-,t.r-O ~ &l~'"4p. \ ,,0
t~ 14
o~o~o \,...~.o o
I ~

i | . l_ I I i_

- 150 -100 - 50 0 50 1130

Temperature I~

Figure 19. Temperature dependence of the second moment of various polysaccharides in the
solid state. (F-l--) Konjac glucomannan; (O-----) amylose; (O .... ) pullulan; (q~----) dextran.

about 0 and -60~ The temperatures at which the second moment decreases rapidly are
higher than those in the case of the other p oly saccharides. This suggests that the movements
of hydroxymethyl groups of KGM are slightly hindered in comparison with the other three
poly saccharides. Further investigations are urgently required in order to clarify whether this is
due to the different modes of glucosidic linkages, to the presence of branching or to other
reasons.

REFERENCES
1. K.Nishinari, In G.O.Phillips, D.J.Wedlock and P.A.Williams (eds), Gums and Stabilisers
for the Food Industry 4, IRL Press at Oxford University Press, Oxford, (1987) 373.
2. K.Nishinari, P.A.Williams and G.O.Phillips, Food Hydrocoll., 6 (1992) 199.
3. K.Kato and K.Matsuda, Agric.Biol. Chem., 33 (1969) 1446.
329

4. M.Maeda, H.Shimahara and N.Sugiyama, Agric. Biol. Chem., 44 (1980) 245.


5. K.Maekaji, Agile. Biol. Chem., 38 (1974) 315.
6. K.Kohyama, H.Iida and K.Nishinari, Food Hydrocoll., 7 (1993) 213.
7. H.Chanzy, A.Grosrenaud, J.P.Joseleau, M.Dube and R.H.Marchessault, Biopolymers, 21
(1982) 301.
8. K.Kohyama and K.Nishinari, JARQ, 31 (1997) 301.
9. K.Kohyama, Y, Sano and K.Nishinari, Food Hydrocoll., 10 (1996) 229.
10. D.Henley, Arkiv. Kemi., 18 (1961) 327.
11. W.Schwald and O. Bobleter, J.Appl. Polym. Sci., 35 (1988) 1937.
12. K.Nishinari, K.Kohyama, P.A.Williams, G.O.Phillips, W.Burchard and K. Ogino,
Macromolecules, 24 (1991) 5590.
13. S.M.Clegg, G.O.Phillips and P.A.Williams, In G.O.Phillips, D.J.Wedlock and
P.A.Williams (eds), Gums and Stabilisers for the Food Industry 5, IRL Press at Oxford
University Press, Oxford, (1990) 463.
14. K.Kohyama and K Nishinari, In G.O.Phillips, D.J.Wedlock and P.A.Williams (eds),
Gums and Stabilisers for the Food Industry 5, IRL Press at Oxford University Press,
Oxford, (1990) 459.
15. N. Kishida, S. Okimasu and T.Kamata, Agric.Biol.Chem., 42 (1978) 1645.
16. M Yoshimura and K Nishinari, Food Hydrocoll., 13 (1999) 227.
17. K. Nishinari, Colloid & Polym. Sci., 275 (1997) 1093,
18 A.G.Ward and P.R.Saunders, Nature, 176 (1955) 26,
19. M.Watase and K.Nishinari, Rheol. Acta., 22 (1983) 580,
20. C.Rochas, M.Rinaudo and S.Landry, Carbohydr. Polym., 12 (1990) 255,
21 R.Niki, K.Kohyama, Y.Sano and K.Nishinari, Polymer Gels & Netw., 2 (1994) 105,
22. K.Nishinari, K.Hofmann, H.Moritaka, K.Kohyama and K.Nishinari, Macromol. Chem.
Phys., 198 (1997) 1217,
23. J.R.Mitchell, J Texture Stud., 11 (1980) 315,
24. K.Nishinari, S.Koide and K.Ogino, J. Phys. (France), 46 (1985) 793.
25. A.H.Clark and S.B.Ross-Murphy, Brit. Polym. J., 17 (1985) 164,
26. D. Oakenfull, J. Food Sci., 49 (1984) 1103, 1104, 1110.
27. A.Haque and E.R.Morris, Carbohydr. Polym., 22 (1992) 161.
28. M.Vigouret, M.Rinaudo and J.J.Desbrieres, J. Chim. Phys., 93 (1996) 858.
29. J.S.G.Reid, M.Edwards and I.C.M.Dea, In G.O.Phillips, D.J.Wedlock and P.A.Williams
(eds), Gums and Stabilisers for the Food Industry 4, IRL Press at Oxford University Press,
Oxford, (1998) 391.
30. M.Shirakawa, M.Yamatoya and K.Nishinari, Food Hydrocoll., 12 (1998) 25.
31 M.Hirashima, T.Takaya and K.Nishinari, Thermochim. Acta., 306 (1997) 109.
32. T.Funami et al. Food Hydrocoll, in press
33 K.Kohyama and K Nishinari, J. Agric. Food Chem., 41 (1993) 8.
34. T.Nagano, T.Akasaka and K Nishinari, Biopolymers, 34 (1994) 1303.
35 T.Nagano, H.Mori and K.Nishinari, J. Agric. Food Chem., 42 (1994) 1415.
36. R.Niki and T.Sasaki, Jpn. J. Zootech. Sci., 58 (1987) 1032.
37. P.Zoon, T.van Vliet and P.Walstra, Neth. Milk. Dairy J., 43 (1989) 17.
330

38. K.Nakamura, K.Harada and Y.Tanaka, Food Hydrocll., 7 (1993) 435.


39. T.Amari andM.Nakamura, Nihon Reorogi Gakkaishi, 6 (1977) 28.
40. P.A.Williams, P.Annable, G.O.Phillips and K.Nishinari, in K.Nishinari and E.Doi (eds),
Food Hydrocolloids;Structures, Functions, and Properties, Plenum Press, New York
(1994) 435.
41. P.Annable, P.A.Williams and K.Nishinari, Macromolecules, 27 (1994) 4204.
42. P.A.Williams, S.M.Clegg, M.J.Langdon, K.Nishinari and L.Piculell, Macromolecules, 26
(1993) 5441.
43. K.Nishinari, E.Miyoshi, T.Takaya and P.A.Williams, Carbohydr. Polym., 30 (1996) 193.
44. E.Miyoshi, T.Takaya and K.Nishinari, J.Agric.Food Chem., 44 (1996) 2486.
45. E.Miyoshi, T.Takaya and K.Nishinari, Polymer Gels & Netw., 6 (1999) 273.
46. M.Yoshimura, T.Takaya and K.Nishinari, J.Agric.Food Chem., 44 (1996) 2970.
47. M.Yoshimura, T.Takaya and K.Nishinari, Carbohydr. Polym., 35 (1998) 71.
48. K.Nishinari and E.Fukada, J. Polym. Sci. Polym. Phys. Ed., 18 (1980) 1609.
49. K.Nishinari, H.Horiuchi and E.Fukada, Rep. Prog. Polym. Phys. Jpn., 23 (1980) 759.
50. K.Nishinari, N.Shibuya and K.Kainuma, Makromol. Chem., 186 (1985) 433.
51. K.Nishinari, K.Kohyama, N.Shibuya, K.Y.Kim, N.H.Kim, M.Watase and A.Tsutsumi,
in W.Glasser and H.Hatakeyama (eds), Viscoelasticity ofBiomaterials, ACS Symposium
Series 489(1992) 357.
52. K.Nishinari and A.Tsutsumi, J. Polym. Sci. Polym. Phys. Ed., 22 (1984) 95.
53. M.Kimura, M.Usuda and T.Kadoya, Sen-i Gakkaishi, 30 (1974) T-221.

You might also like