You are on page 1of 8

Carbohydrate Polymers 134 (2015) 285–292

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Effects of pH and temperature on colloidal properties and molecular


characteristics of Konjac glucomannan
Wenjie Jian a,b , Ka-Chai Siu a , Jian-Yong Wu a,∗
a
Department of Applied Biology & Chemical Technology, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong
b
Department of Pharmacy, Xiamen Medical College, Xiamen 361000, China

a r t i c l e i n f o a b s t r a c t

Article history: The hydrocolloidal characteristics of Konjac glucomannan (KGM) are important for its application as a
Received 18 May 2015 thickening and gelling agent for liquid foods. In this study, the rheological behavior and molecular proper-
Received in revised form 8 July 2015 ties such as molar mass, hydrodynamic radius and chain conformation of KGM in water were determined
Accepted 11 July 2015
at various pH levels (4.0–10.0) during heating from 20 to 80 ◦ C. Acidic and neutral conditions (pH 4.0–7.0)
Available online 31 July 2015
promoted the dispersion of KGM, and alkaline condition at pH 10 favored its aggregation in water, while
KGM maintained a random coil conformation in the whole pH range. Associated with the pH effects were
Keywords:
changes in the rheological behavior during heating from 20 to 80 ◦ C. The significant differences in the
Konjac glucomannan
Colloidal solution
colloidal and rheological characteristics were mainly attributed to alteration of intermolecular interac-
pH tion (attractive or repulsive) rather than deacetylation at various pH levels. Deacetylation occurred in
Temperature both acidic and alkaline condition. The second virial value was positive in acidic and negative in alka-
Light scattering analysis line condition. The results showed that hydrocolloidal characteristics of KGM in water were significantly
Molecular interaction affected by pH.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction & Nishinari, 2002; Nishinari & Takahashi, 2003; Solo-de-Zaldivar,


Tovar, Borderias, & Herranz, 2014). Through the formation of a
Konjac glucomannan (KGM), a plant polysaccharide derived three-dimensional network of molecular chain by junction zones
from the tuber of Amorphophallus konjac C. Koch, has been widely after deacetylation, the aqueous solution of KGM is transformed
applied in food industry as a gelling and thickening agent due into a thermally stable gel (Williams et al., 2000). Exposure to alka-
to its ability of thermal-irreversible gelling and effective viscos- line and heat treatment may cause the deacetylation of KGM and
ity enhancement (Herranz et al., 2012; Tomczynska-Mleko et al., significant reduction of solubility in an aqueous solution.
2014). Many recent studies have shown its various biological The afore-stated gelation mechanism suggests that thermal and
activities including anti-tumor, immunomodulation, wound heal- alkaline conditions are important factors affecting the gelling prop-
ing (Al-Ghazzewi, Elamir, Tester, & Elzagoze, 2015; Bo, Muschin, erties of KGM. Although many previous studies have evaluated the
Kanamoto, Nakashima, & Yoshida, 2013; Zhang, Chen, & Yang, rheological properties of KGM in aqueous solutions, most of these
2014). Although the exact molecular composition may vary with have been carried out at neutral pH and room temperature (Jin
the source and production process, purified KGM has a polysaccha- et al., 2014; Xu, Sun, Yang, Ding, & Pang, 2007). The characteristics
ride backbone composed of glucose and mannose at 1:1.5–1:1.6 of KGM in acidic and alkaline pH conditions and at higher tem-
molar ratio, with 5–10% acetyl substitution (Dave & McCarthy, peratures are important for their applications in these conditions.
1997). It is generally believed that acetyl group on the KGM The evaluation and quantification of the related characteristics and
polymer structure is the most important for its solubility and parameters is helpful to understanding the dynamic behavior and
gelling properties in water (Huang, Takahashi, Kobayashi, Kawase, interactions of KGM polymer chains, and their relationship to the
textural properties of liquid foods (Nishinari and Takahashi, 2003).
This study aims to examine the hydrocolloid characteristics of
∗ Corresponding author. Fax: +852 2364 9932.
KGM in water at various pH levels in response to heating, and
E-mail address: jian-yong.wu@polyu.edu.hk (J.-Y. Wu).
to elucidate the related gelling mechanisms. The parameters of

http://dx.doi.org/10.1016/j.carbpol.2015.07.050
0144-8617/© 2015 Elsevier Ltd. All rights reserved.
286 W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292

molecular properties, and chain conformation, and aggregation will strains in the range of 0.1–100% to find the linear viscoelastic region
be derived through light scattering analysis as useful references for for pH 4.0, 7.0, 9.0 and 10.0 corresponding to the strains of 30%,
the application of KGM in food processing. 35%, 20% and 1.5%, respectively. Accordingly, the shear storage
modulus (G ) and loss modulus (G ) were measured at angular fre-
2. Materials and methods quency of 6.28 rad/s and 1% strain as a function of temperature
varying from 20 ◦ C to 80 ◦ C at 2 ◦ C/min and data collection every
2.1. Purification and composition analysis of KGM 0.25 min. All tests were performed in triplicate, and the results were
averaged.
The initial Konjac glucomanan (KGM) material used for this
study was purchased from Shaotong Shanai Kojac Development 2.4. Infrared and Raman spectral measurements
Co. (Yunnan, China). Before use, the KGM powder was purified by
alcohol precipitation followed as described in our previous study Infrared (IR) measurement was performed on a Fourier trans-
(Jian, Yao, Wang, Guan, & Pang, 2010). Its weight-average molecu- forms IR spectrometer (Nicolet 6700, Thermo Fisher, USA) at a
lar weight was about 1.14×103 kDa. This purified KGM was used resolution of 4 cm−1 in the range of 400–4000 cm−1 (Xu et al.,
in this study and its composition was analyzed as follows. The 2007). Freeze-dried KGM samples were prepared as KBr slices and
moisture content was determined after drying in an oven (105 ◦ C) scanned against air background for 32 times with each sample to
to constant weight, and the ash content determined in a muffle attain an average. Raman spectra of KGM samples (freezed-dried
furnace (550 ◦ C) according to methods from International Organi- powder) were recorded from 100 cm−1 to 3000 cm−1 on a HR800
zation for Standardization (ISO 2291:1980, ISO 936:1998). The total Raman microscope (Horiba Jobin Yvon), equipped with a 488 nm
carbohydrate content was determined by phenol sulphuric acid laser operated at 180 mW and spot size of excitation laser about
method using glucose as a standard (Chaplin & Kennedy, 1994) and 1 ␮m (Synytsya, Copikova, Matejka, & Machovic, 2003).
the total protein content by the Kjeldah method (Crisan & Sands,
1978). Monosaccharide constituents were analyzed by HPLC after
2.5. Dynamics light scattering measurement
acid hydrolysis and 1-phenyl-3-methyl-5-pyrazolone (PMP) reac-
tion as reported previously (Chen, Siu, Cheung, & Wu, 2014). Acetyl
Dynamics light scattering (DLS) measurement was performed
content was determined by titration using hydrochloride solution
to attain the hydrodynamic radius (Rh) and to estimate molar
according to the method of Laignel, Bliard, Massiot, & Nuzillard,
mass (Mw-R) with a DynaPro NanoStar instrument (Wyatt, USA)
(1997).
(Sagou, Rotureau, Thomas, & Duval, 2013), Before loading, the KGM
sample solution in water was filtered through a Millex-HN nylon
2.2. Preparation of KGM solution
clarification kit with a pore size of 0.2 ␮m (Fisher Scientific). The
temperature was varied from 25 ◦ C to 80 ◦ C at 0.5 ◦ C/min ramp
All experiments in this part were carried out at room temper-
rate. The signal acquisition period was set to 10 s and the aver-
ature as well as the storage of prepared solution. KGM powder
aged result of 10 acquisitions was taken as a measurement. The
(0.1 g) was dissolved in super pure water (200 ml) by constantly
refractive index increment for KGM was set to 0.138. The molar
stirring for 24 h at room temperature. The KGM solution pH was
mass Mw-R was estimated from Rh by the model of linear polymer
adjusted to 4.0, 7.0, 9.0, and 10.0 with 0.1 M and 1 M NaOH or HCl.
and random coil conformation (Prawitwong, Takigami, & Phillips,
The volume of NaOH or HCl added for to the KGM solution for the
2007).
pH adjustment was very small, having a negligible effect on the
solution concentration. The solution was kept at room tempera-
ture for 12 h to allow the KGM to swell, followed by centrifuging 2.6. GPC-MALLS analysis
at 18,000 rpm (∼1814 × g) for 15 min to remove insolubles (aggre-
gates and gels formed at high concentrations, if any). The solution Gel permeation chromatography with on-line multi-angle laser
(0.5 mg/ml) was directly used for measurement of dynamic light light scattering (GPC-MALLS) was applied to determine the weight-
scattering. For IR analysis, the solution was dialyzed with ultrafil- average molecular weight (Mw ) and conformation characteristics
tration membrane made of regenerated cellulose (1.0 kDa MWCO, of KGM dissolved in water (Prawitwong et al., 2007). The GPC-
Millipore, USA), and then freeze-dried. As the KGM sample was MALLS system consisted of a Waters e2695HPLC system and a
isolated by ethanol precipitation of the commercial KGM, dialy- Waters 2414 refractive index detector, and a Wyatt DAWN HELEOS
sis is a common step for the removal of low molecular weight II detector (18 angles). Sample solution (0.2 mg/mL) was injected
impurities such as small salts which may be attached to the KGM into the system at 100 ␮L. Two GPC columns were used in a series
polymer chains. Because of the high viscosity of KGM solution, for the separation, TSK-GEL G4000PWxl and G5000PWxl, (Tosoh
a low concentration solution (0.2 mg/ml) was prepared for the Bioscience, Tokyo, Japan). Sodium nitrate solution (50 mM with
GPC-MALLS measurement to assure complete passage of the dis- 0.02% w/v sodium azide) was used as the mobile phase at a flow
solved KGM through the GPC-MALLS system. On the other hand, rate of 0.4 ml/min and 25 ◦ C. Normalization was performed with
a much higher concentration of KGM solution (5 mg/ml) was pre- dextran standards (Sigma Aldrich, Mw 30 kDa). Both GPC data and
pared in the same way but without centrifuging for the rheological MALLS data were collected and analyzed using Astra6.0 software
measurement. package.

2.3. Measurement of dynamic rheological behavior 2.7. Determination of second virial coefficient by MALLS analysis

The rheological behavior of KGM was measured in various con- The second virial coefficient was determined through MALLS
ditions according to a reference (Herranz, Borderias, Solas & Tovar, analysis with a Wyatt Dawn Heleos II detector linked with
2012) with minor modifications. Dynamic rheology measurement a Wyatt Calypso automating delivery system on batch mode
was performed of the KGM solution on an Anton Paar MCR302 (Ma et al., 2015). Before the measurement, the initial KGM
Rheometer (Bohlin Instruments, Inc. Cranbury, NJ) with parallel- solution at 0.5 mg/ml was diluted into a series of gradient con-
plate pp50 (20 mm diameter and 1 mm gap). Preliminary tests centrations (0.1, 0.2, 0.3 and 0.4 mg/ml) with the respective pH
were performed at 25 ◦ C with angular frequency of 6.28 rad/s and buffer solution. Meanwhile, 1 mg/ml solution of dextran standard
W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292 287

Fig. 1. Change of storage modulus and loss modulus of KGM solution (5 mg/ml) with temperature at various pH levels.

(30 kDa) was prepared with each respective pH buffer solu- 1.53% ash (all by mass) and of mannose and glucose in 2.3:1 molar
tion, and used as the normalization standard for the following ratio. As KGM is extracted from the tuber of Amorphophallus konjac
measurement. C. Koch, the purity and composition depend on the source plant
The instrument system was first flushed with a pH buffer solu- material as well as the extraction process (Nishinari, 2000), and its
tion to obtain the initial baseline light scattering intensity, followed monosaccharide composition varies with the plant species (Pang,
by injection of 1 mg/ml dextran solution to normalize the system. Lin, Zhang, Tian, & Sun, 2003). Most of the above composition values
KGM sample solution of gradient concentrations (0.1–0.5 mg) was of KGM were within the ranges commonly reported in the litera-
then injected, followed by a buffer flush to retain the baseline. The ture (Kato & Matsuda, 1969; Koroskenyi & McCarthy, 2001; Smith
light scattering intensity reading was taken 5 min after each injec- & Srivastava, 1959), except for the relatively higher molar ratio of
tion. Zimm plot was generated with the light scattering data based mannose to glucose, which is commonly in the range of 1.5:1–1.6:1
on the Rayleigh theory, from which the second virial coefficient (Koroskenyi & McCarthy, 2001; Maeda, Shimahara, & Sugiyama,
was determined by the slope of the concentration versus the excess 1980). Therefore, the KGM sample used in this study was repre-
Rayleigh ratio plot. sentative of those widely used in the food industry and its property
data can be useful references for general situation.
2.8. Zeta-potential measurement

Zeta-potential of the KGM solutions (0.5 mg/ml) was measured 3.2. pH and temperature effects on rheological behavior of KGM
by dynamic laser light scattering (DLS) analysis on a Malvern in water
Zetasizer Nano-ZS90 instrument using a folded capillary cuvette
DTS1060 at room temperature, with five repeating measurements Fig. 1 shows the changes viscoelastic modulus of KGM solu-
taken for each sample (Hunter, 2013). tion with temperature increase (20–80 ◦ C) at various pH levels.
The rheological measurements showed a high reproducibility, with
2.9. Statistical analysis errors less than 3%. KGM presented significantly different rheolog-
ical behaviors with the pH changes as shown by Fig. 1. At pH 4
All experiments were carried out in triplicate and the results (Fig. 1a) and pH 7.0 (Fig. 1b), the storage modulus (G ) was close to
were represented by means and standard deviations. Analysis the loss modulus (G ) at 20 ◦ C, and as the temperature increased,
of variance (ANOVA) was computed using SPSS 16.0 (SPSS Inc., G became higher than G . This tend suggests that KGM at 20 ◦ C was
Chicago, IL). in the colloidal solution-to-gel transition state and remained in a
colloidal solution as the temperature exceeded 20 ◦ C in acidic and
3. Results and discussion neural conditions. Another trend was the linear decline of G and G
with temperature increase to a very low level at 80 ◦ C, implying the
3.1. Chemical composition of KGM weakening of inter- and intra-molecular attraction forces rapidly
by heating, which may involve the degradation of polymer chain
According to analysis, KGM was composed of 93.4% total car- and dissociation of polymer aggregates. At pH 9.0 (Fig. 1c), the ini-
bohydrate, 1.87% total protein, 1.62% acetyl, 8.49% moisture, and tial moduli at 20 ◦ C were significantly higher, and dropped sharply
288 W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292

100 100
(a) (b)

Transmittance (%)
Transmittance (%)
80 80

60 60
Acetyl
2930
40 1650 40

3430
20 20
3600 2800 2000 1200 400 3600 2800 2000 1200 400
-1
Wavenumber (cm-1) Wavenumber (cm )
100 100
(c) (d)
Transmittance (%)

Transmittance (%)
80 80

60 60

40 40

20 20
3600 2800 2000 1200 400 3600 2800 2000 1200 400
-1 -1
Wavenumber (cm ) Wavenumber (cm )

Fig.2. FT-IR spectra: (a) Native KGM; freeze-dried KGM dissolved in water at (b) pH 4.0, (c) pH 9.0, and (d) pH 10.0.

as the temperature was increased to 21 ◦ C, and then followed the


similar trends to those at the lower pH.
Most interesting and dramatic changes were observed at pH 10.0
(Fig. 1d). The initial G and G at 20 ◦ C were much lower than those
at the lower pH, the difference between G and G was relatively
large. G and G followed very different trends with the tempera-
ture increase from 20 to 80 ◦ C. G declined gradually in the whole
temperature range except at 60 ◦ C, it jumped to a notably higher
level. In contrast, G showed a small decline with the temperature
increase from 20 to about 40 ◦ C and remained steady till 60 ◦ C; at
about 60 ◦ C, it jumped to a much higher level, and then increased
sharply with further increase of temperature to 80 ◦ C. These trends
were indicative of a colloidal solution undergoing gelation, i.e. at
pH 10.0, the KGM solution turned to a gel as the temperature was
increased to 60 ◦ C and the gel strength increased rapidly with fur-
ther temperature increase. The gelation process at pH 10.0 due to
heating was consistent with the alkaline-induced gelation of KGM
reported previously (Herranz, Borderias, Solas et al., 2012).
Therefore, the influence of temperature on the rheological
behavior of KGM solution was strongly dependent on the pH range.
In the acidic and weakly alkaline range (pH 4.0–9.0), lower tem-
peratures (below 21 ◦ C) favored the gelling of KGM, and higher
temperatures (above 21 ◦ C) promoted its dispersion in water.
In highly alkaline pH range, lower temperatures favored solu-
tion stability, and higher temperatures (above 60 ◦ C) promoted
gelling. Besides the rheological measurement, further analysis and
quantification of the structural, conformation and aggregation
properties of KGM in solution will be helpful for characterization
and elucidation of the changes in different pH ranges with temper-
ature.

3.3. Infrared spectral characteristics

Fig. 2 shows the FT-IR spectra of KGM samples from var-


ious pH solutions. On the spectrum at neutral pH 7.0 (Fig. 2a), Fig. 3. Raman spectra: (a) Native KGM; (b) freeze-dried KGM dissolved in water at
the strong vibration at 3430 cm−1 is ascribed to the hydroxyl ( OH) various pH levels.
W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292 289

Fig. 4. Temperature sweep of hydrodynamics radius (a) and weight-average molecular mass (b) of KGM solution at various pH levels.

group of polysaccharide, peak at 2930 cm−1 to the vibration IR spectra except for the absence of acetyl (1730 cm−1 ) and methyl
of methyl ( CH3 ), and the small peak at 1730 cm−1 to the acetyl (2930 cm−1 ) peaks. This implied that acidic and alkaline treatment
group (CH3 CO ) as the characteristic substituted group in KGM caused the deacetylation but no change in the backbone of KGM.
(Zhang et al., 2001). The KGM in acidic solution (pH 4, Fig. 2b) and Fig. 3a shows the Raman spectrum of native KGM. As the acetyl
alkaline solutions (pH 9, Fig. 2c; pH 10 Fig. 2d) showed the similar group (CH3 CO ) of KGM is attached with hydroxyl at C-6 posi-
290 W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292

tion to form acetic acid ester linage, the weak peak at 1740 cm−1
is attribute to the vibration of carbonyl(C O) and the peak at
1473 cm−1 to the vibration of methyl (CH3 ) in the acetyl group (De
Gelder, De Gussem, Vandenabeele, & Moens, 2007; Synytsya et al.,
2003). In comparison of the Raman spectra before (Fig. 3a) and after
treatment (Fig. 3b), there was a significant decrease in Raman inten-
sity after acidic and alkaline treatment and also the elimination of
the characteristic for acetyl peaks at 1740 cm−1 and 1473 cm−1 .
The findings from the Raman spectra confirmed the results of FT-IR
analysis.
Similar to the changes observed from the above IR and Raman
spectra of KGM, deacetylation has also occurred to KGM as a
result of combined alkaline and heat treatment in previous studies
(Herranz et al., 2012; Solo-de-Zaldivar et al., 2014; Williams et al.,
2000). Deacetylation in KGM could significantly affect its solubility
and gelling properties (Du, Li, Chen, & Li, 2012). However, previ-
ous studies have mostly focused on alkaline deacetylation of KGM
and ignored acidic conditions. Our present study has shown that
deacetylation also occurred in acidic conditions, similar to the acid
hydrolysis of xylan-associated acetyl groups in dilute sulfuric acid
solution (Maloney, Chapman, & Baker, 1985).
As the acetyl group of KGM can be hydrolyzed in both acidic
and alkaline conditions, deacetylation is not an essential condition
for gelation of KGM. Meanwhile, although the degree of acetyla-
tion has been shown to affect the solubility of KGM in neutral
water (pH 7.0) (Du et al., 2012; Huang et al., 2002), it may be much
less important than temperature and pH in affecting the colloidal
and rheological properties KGM. In other words, the changes of
rheological behaviors with pH and temperature can be associated
with alternation of other molecular properties than deacetylation.
The findings are in agreement with the view of Williams et al.
(2000) that gelation of native KGM solution may also occur without
deacetylation.

3.4. Changes of hydrodynamic radius and molar mass

As shown in Fig. 4a, hydrodynamic radius (Rh) presented differ-


ent trends with temperature increase at different pH levels. At pH
4.0 and 7.0, Rh decreased sharply from about 120 nm to 40 nm with
temperature increasing from 25 ◦ C to 50 ◦ C; at pH 10.0 Rh increased Fig. 5. Chromatograms of KGM at various pH levels obtained from GPC-MALLS anal-
ysis: (a) GPC-MALLS spectra; (b) RMS radius; (c) molar mass (0.2 mg/ml KGM in
sharply at 60 ◦ C; at pH 9.0, Rh was not affected by temperature
water).
change. The molar mass (Mw-R) of KGM showed the same trends
with temperature as Rh at various pH ranges (Fig. 4b). Both Rh and
Mw-R were significantly affected by the pH change (P < 0.01). The
2035 kDa with the pH increase to 10.0. The number-averaged molar
results suggested that the changes in Rh with temperature were
mass (Mn) as well as the dispersity index (Mw/Mn) followed the
mainly attributed to the aggregation or disaggregation of the poly-
similar trends of change with temperature (Table 1). The Mw and
mer chains. These results have shown the significance of pH level
Rm values at pH 7.0 obtained in this study were close to the those
on the thermal stability and aggregation of KGM. At pH 4.0 and 7.0,
(1170 kDa and 120.3 nm reported by Prawitwong et al. (2007)), but
increasing temperature favored its dispersion and colloidal stability
significantly different from those (Mw 247.6 kDa and Rm 100 nm)
in solution, while at pH 10, temperature had an opposite effect. Such
by Xu et al. (2013) due probably to the different raw material prop-
trends were consistent with the above rheological changes with pH
erties and preparation methods.
and temperature, indicating that pH and temperature affected KGM
The changes in Mw and other related molecular properties
rheological behavior through variation of intermolecular aggrega-
of KGM could be attributed to different mechanisms. At pH
tion or disaggregation.
4, the decrease in Mw was probably caused by acidic degra-
dation of glycoside linkage; at pH 9.0, the slight decrease in
3.5. Molecular characteristics from GPC-MALLS analysis Mw was probably a result of deacetylation. The significant Mw
increase at pH 10.0 was attributed to the aggregation of KGM
Besides the Rh and Mw-R obtained from DLS, the weight- in alkaline conditions. The changes of the root mean square
average molar mass (Mw), root-mean-square radius (Rm) and chain radius (Rm) with pH also followed the similar trends with Mw
conformation are also important molecular characteristics which (Table 1).
are derived from GPC-MALLS analysis (Fig. 5). The profile at pH In spite of the significant changes in Mw and particle (aggregate)
10.0 (Fig. 5a) was significantly different from those at other pH lev- size with pH, the conformation of KGM remained in the random coil
els, indicating the Mw changes at pH 10. The Mw decreased from as seen from the relatively small change in the slope of conforma-
1463 to 1318 kDa with the pH decrease from pH 7.0 to pH 4.0, and tion plot (Rm vs Molar mass) from 0.41 to 0.49 (Table 1). Therefore,
to 1385 kDa with pH increase from pH 7.0 to 9.0, but increased to it is concluded that pH range had a significant effect on the molec-
W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292 291

Table 1
Molecular characteristics of KGM solution (0.2 mg/ml) at pH 4.0, pH 7.0, pH 9.0 and 10.0 obtained from GPC-MALLS.

pH Mn (kDa) Mw (kDa) Rm (nm) RMS plot slope

4.0 1024.9( ± 3.10%) 1318(±1.60%) 123.6(±0.70%) 0.49(±0.09%)


7.0 1103.2( ± 2.80%) 1463.3(±2.60%) 127.2(±0.80%) 0.44(±0.12%)
9.0 1155.6( ± 1.60%) 1385.2(±1.50%) 123.3(±0.60%) 0.44(±0.11%)
10.0 1345.8( ± 3.90%) 2035.1(±4.90%) 129.9(±2.70%) 0.41(±0.27%)

tend to form aggregates or gels at pH 10.0. This pH-dependent


dispersion or aggragation of KGM in water is in close agreement
with that dervied from the results of rehological and molecular
weight measurements.
The colloidal and gelling properties of biopolymers in solution
are strongly dependent on short-range inter-molecular inter-
actions such as hydrgen bonding and hydrophobic interaction
(Nishinari & Zhang, 2004; Yin, Zhang, Huang, & Nishinari, 2008). The
second virial coefficient value (A2 ) is a useful index for the degree of
intermolecular interactions of biopolymers in dilute solutions due
to several molecular forces including ionic, hard-sphere, van der
Waals and other short-range interactions (e.g., hydrophobic inter-
actions) (Lehermayr, Mahler, Mader, & Fischer, 2011). Negative A2
values are indicative of attactive intermolecular interactions for
aggregate formation while positive A2 values represent repulsive
intermolecular interactions for colloidal dispersion and stability.
Zeta potential as a measure of electrostatic charges is another
useful parameter for molecular interactions through electrostatic
force (Saito et al., 2013). As shown in Fig. 6c, the KGM solu-
tion exhibited negative zeta potentials (negatively charged) at pH
3.0–10.0, and the negative zeta potential increased from −4.12
to −6.12 mv with the pH increase from pH 7.0 to 10.0. However,
according to the DLVO theory, the absolute magnitude of zeta
potential between 10 and 30 mv indicates an unstable dispersion,
and the value below 5 mv a strong aggregation (Bhattacharjee,
Elimelech, & Borkovec, 1998). In view of the low absolute zeta
potential of KGM, the electrostatic potential was not sufficient to
maintain its stable disepersion. It furter suggests that electrostatic
interactions in KGM molecules were in much less significant than
other molecuar intreactions in affecting its colloidal stability and
gelation behavior.

4. Conclusions

The results have shown that the colloidal and rheological


behaviors of KGM in water were significantly affected by pH and
temperature. At room or relatively low temperatures, the pH level
strongly affected the average molecular weight and particle size
of KGM but not the chain conformation. Acidic conditions pro-
Fig. 6. Zimm plot of KGM at pH 7.0 and 25 ◦ C (a) and the corresponding second
moted the degradation and disaggregation of molecular chain
virial coefficient (b) and zeta potential (c) of KGM solution (0.5 mg/ml) at various whereas alkaline conditions favored its aggregation. KGM remained
pH levels. in random coil conformation in all pH conditions. The pH range sig-
nificantly affected the rheological behaviors of KGM upon thermal
treatment. In relatively acidic conditions, thermal treatment pro-
ular weight and particle size but not on the chain conformation of
moted the dispersion of KGM or the formation of a stable solution in
KGM in water, and acidic condition (pH 4.0) promoted the degra-
water. In alkaline conditions, KGM presented the typical gelation
dation of KGM molecular chain, and alkaline condition favored its
process at temperature over 60 ◦ C. These effects of pH and tem-
aggregation.
perature on the colloidal and rheological behaviors were mostly
attributed to the changes in intermolecular interactions instead of
3.6. Molecular interactions deacetylation.

Fig. 6a is the Zimm plot of KGM solution at pH 7.0 and 25 ◦ C


obtained by static laser scattering measurement (Zimm plots at Acknowledgements
other pH levels not shown). The second virial coefficient values
(A2 ) derived from the Zimm plots are positive at pH 3.0–9.0 and This work was supported by Hong Kong Scholar Program
negative at pH 10.0 (Fig. 6b). The results indicate that KGM retains (XJ2013031) and The Hong Kong Polytechnic University, Hong
a stable dispersion in acidic and slightly alkaline conditions and Kong Government GRF grant (PolyU 5033/11P), China Postdoctoral
292 W. Jian et al. / Carbohydrate Polymers 134 (2015) 285–292

Science Foundation (2012M510199), and the Natural Science Foun- Ma, Y. F., Acosta, D. M., Whitney, J. R., Podgornik, R., Steinmetz, N. F., French, R. H.,
dation of Fujian Province (2012J05046). et al. (2015). Determination of the second virial coefficient of bovine serum
albumin under varying pH and ionic strength by composition-gradient
multi-angle static light scattering. Journal of Biological Physics, 41(1), 85–97.
References Maeda, M., Shimahara, H., & Sugiyama, N. (1980). Studies of mannan and
related-compounds.5. detailed examination of the branched structure of
konjac glucomannan. Agricultural and biological chemistry, 44(2), 245–252.
Al-Ghazzewi, F., Elamir, A., Tester, R., & Elzagoze, A. (2015). Effect of depolymerised Maloney, M. T., Chapman, T. W., & Baker, A. J. (1985). Dilute acid-hydrolysis of
konjac glucomannan on wound healing. Bioactive Carbohydrates and Dietary paper birch – kinetics studies of xylan and acetyl-group hydrolysis.
Fibre, 5(2), 125–128. Biotechnology and Bioengineering, 27(3), 355–361.
Bhattacharjee, S., Elimelech, M., & Borkovec, M. (1998). DLVO interaction between Nishinari, K., & Takahashi, R. (2003). Interaction in polysaccharide solutions and
colloidal particles: beyond Derjaguin’s approximation. Croatica Chemica Acta, gels. Current Opinion in Colloid & Interface Science, 8(4–5), 396–400.
71(4), 883–903. Nishinari, K., & Zhang, H. (2004). Recent advances in the understanding of heat set
Bo, S., Muschin, T., Kanamoto, T., Nakashima, H., & Yoshida, T. (2013). Sulfation and gelling polysaccharides. Trends in Food Science & Technology, 15(6), 305–
biological activities of konjac glucomannan. Carbohydrate Polymers, 94(2), 312.
899–903. Nishinari, K. (2000). Konjac glucomannan. Developments in Food Science, 41,
Chaplin, M. F., & Kennedy, J. F. (1994). Carbohydrate analysis: a practical approach. 309–330.
IRL Press Limited. Pang, J., Lin, Q., Zhang, F. S., Tian, S. P., & Sun, Y. M. (2003). Progress in the
Chen, X., Siu, K. C., Cheung, Y. C., & Wu, J. Y. (2014). Structure and properties of application and studies on functional material of konjac glucomannan. Chinese
a(1(3)- ␤-d-glucan from ultrasound-degraded exopolysaccharides of a Journal of Structural Chemistry, 22(6), 633–642.
medici-nal fungus. Carbohydrate Polymers, 106, 270–275. Prawitwong, P., Takigami, S., & Phillips, G. O. (2007). Effects of gamma-irradiation
Crisan, E., & Sands, A. (1978). Nutritional value. Academic Press: New York. on molar mass and properties of konjac mannan. Food Hydrocolloids, 21(8),
Dave, V., & McCarthy, S. P. (1997). Review of konjac glucomannan. Journal of 1362–1367.
Environmental Polymer Degradation, 5(4), 237–241. Sagou, J.-P. S., Rotureau, E., Thomas, F., & Duval, J. F. L. (2013). Impact of metallic
De Gelder, J., De Gussem, K., Vandenabeele, P., & Moens, L. (2007). Reference ions on electrohydrodynamics of soft colloidal polysaccharides. Colloid Surface
database of Raman spectra of biological molecules. Journal of Raman A, 435, 16–21.
Spectroscopy, 38(9), 1133–1147. Saito, S., Hasegawa, J., Kobayashi, N., Tomitsuka, T., Uchiyama, S., & Fukui, K.
Du, X. Z., Li, J., Chen, J., & Li, B. (2012). Effect of degree of deacetylation on (2013). Effects of ionic strength and sugars on the aggregation propensity of
physicochemical and gelation properties of konjac glucomannan. Food monoclonal antibodies: influence of colloidal and conformational stabilities.
Research International, 46(1), 270–278. Pharmaceutical Research, 30(5), 1263–1280.
Herranz, B., Borderias, A. J., Solo-de-Zaldivar, B., Solas, M. T., & Tovar, C. A. (2012). Smith, F., & Srivastava, H. (1959). Constitutional studies on the glucomannan of
Thermostability analyses of glucomannan gels. Concentration influence. Food konjak flour1. Journal of the American Chemical Society, 81(7), 1715–
Hydrocolloids, 29(1), 85–92. 1718.
Herranz, B., Borderias, A. J., Solas, M. T., & Tovar, C. A. (2012). Influence of Solo-de-Zaldivar, B., Tovar, C. A., Borderias, A. J., & Herranz, B. (2014). Effect of
measurement temperature on the rheological and microstructural properties deacetylation on the glucomannan gelation process for making restructured
of glucomannan gels with different thermal histories. Food Research seafood products. Food Hydrocolloids, 35, 59–68.
International, 48(2), 885–892. Synytsya, A., Copikova, J., Matejka, P., & Machovic, V. (2003). Fourier transform
Huang, L., Takahashi, R., Kobayashi, S., Kawase, T., & Nishinari, K. (2002). Gelation Raman and infrared spectroscopy of pectins. Carbohydrate Polymers, 54(1),
behavior of native and acetylated konjac glucomannan. Biomacromolecules, 97–106.
3(6), 1296–1303. Tomczynska-Mleko, M., Brenner, T., Nishinari, K., Mleko, S., Szwajgier, D.,
Hunter, R. J. (2013). Zeta potential in colloid science: principles and applications. Czernecki, T., et al. (2014). Rheological properties of mixed gels: gelatin, konjac
Academic Press. glucomannan and locust bean gum. Food Science and Technology Research,
Jian, W.-J., Yao, M.-N., Wang, M., Guan, Y.-G., & Pang, J. (2010). Formation 20(3), 607–611.
mechanism and stability study of konjac glucomannan helical structure. Williams, M. A. K., Foster, T. J., Martin, D. R., Norton, I. T., Yoshimura, M., &
Chinese Journal of Structural Chemistry, 29(4), 543–550. Nishinari, K. (2000). A molecular description of the gelation mechanism of
Jin, W. P., Xu, W., Li, Z. S., Li, J., Zhou, B., Zhang, C. L., & Li, B. (2014). Degraded konjac konjac mannan. Biomacromolecules, 1(3), 440–450.
glucomannan by gamma-ray irradiation assisted with ethanol: preparation Xu, Z., Sun, Y., Yang, Y., Ding, J., & Pang, J. (2007). Effect of gamma-irradiation on
and characterization. Food Hydrocolloids, 36, 85–92. some physiochemical properties of konjac glucomannan. Carbohydrate
Kato, K., & Matsuda, K. (1969). Studies on the chemical structure of konjac Polymers, 70(4), 444–450.
mannan: part I. Isolation and characterization of oligosaccharides from the Xu, M., Li, D.-S., Li, B., Wang, C., Zhu, Y.-P., Lv, W.-P., & Xie, B.-J. (2013). Comparative
partial acid hydrolyzate of the mannan. Agricultural and biological chemistry, study on molecular weight of konjac glucomannan by gel permeation
33(10), 1446–1453. chromatography-laser light scattering-refractive index and laser
Koroskenyi, B., & McCarthy, S. P. (2001). Synthesis of acetylated konjac light-scattering methods. Journal of Spectroscopy.
glucomannan and effect of degree of acetylation on water absorbency. Yin, W. C., Zhang, H. B., Huang, L., & Nishinari, K. (2008). Effects of SDS on the
Biomacromolecules, 2(3), 824–826. sol–gel transition of konjac glucomannan in SDS aqueous solutions. Colloid and
Laignel, B., Bliard, C., Massiot, G., & Nuzillard, J. M. (1997). Proton NMR Polymer Science, 286(6–7), 663–672.
spectroscopy assignment of -glucose residues in highly acetylated starch. Zhang, H., Yoshimura, M., Nishinari, K., Williams, M. A. K., Foster, T. J., & Norton, I. T.
Carbohydrate Research, 298(4), 251–260. (2001). Gelation behaviour of konjac glucomannan with different molecular
Lehermayr, C., Mahler, H. C., Mader, K., & Fischer, S. (2011). Assessment of net weights. Biopolymers, 59(1), 38–50.
charge and interactions of different monoclonal antibodies. Journal of Zhang, C., Chen, J. D., & Yang, F. Q. (2014). Konjac glucomannan, a promising
Pharmaceutical Sciences, 100(7), 2551–2562. polysaccharide for OCDDS. Carbohydrate Polymers, 104, 175–181.

You might also like