Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Food Hydrocolloids 57 (2016) 209e216

Contents lists available at ScienceDirect

Food Hydrocolloids
journal homepage: www.elsevier.com/locate/foodhyd

Comparative studies of konjac flours extracted from Amorphophallus


guripingensis and Amorphophallus rivirei: Based on chemical analysis
and rheology
Qing Huang a, Weiping Jin a, Shuxin Ye a, Ying Hu a, Yuntao Wang a, Wei Xu a, Jing Li a,
Bin Li a, b, c, *
a
College of Food Science and Technology, Huazhong Agricultural University, Wuhan 430070, China
b
Hubei Collaborative Innovation Centre for Industrial Fermentation, Hubei University of Technology, Wuhan 430068, China
c
Key Laboratory of Environment Correlative Dietology (Huazhong Agricultural University), Ministry of Education, China

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

Article history: Konjac glucomannan (KGM) has been widely used in food industry, which is commonly extracted from
Received 10 May 2015 Amorphophallus rivirei. In this paper, Amorphophallus guripingensis was found to be another good
Received in revised form resource of extracted KGM. The comparative studies, based on chemical composition analysis and
15 January 2016
rheological measurements, between those two different resources were investigated. Results showed
Accepted 20 January 2016
Available online 21 January 2016
that refined A. guripingensis flour (AGF) contained 61.35% KGM, which was lower than that of refined ARF
(72.80%). But, the viscosity-increased speed of AGF was faster than ARF. To be specific, even the viscosity
of 3.5 wt% AGF solution was half of ARF, while they had similar viscosity values at a concentration of
Keywords:
Amorphophallus rivirei
4.5 wt%. Besides, viscosity change of the refined AGF with pH was more pronounced under strong acid
Amorphophallus guripingensis condition. Meanwhile, thermo-gelling temperature of 4.0 wt% AGF solution was 85  C, lower than that of
Refined flour ARF. These results demonstrated that AGF solutions were sensitive to processing factors, such as pH and
pH-sensitivity temperature as compared to A. rivirei. Those properties endows AGF more valuable applications in food
Rheological properties industry and it has great potential to be a pH-sensitive matrix.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction China is the first country to study and utilize konjac (Chua,
Baldwin, Hocking, & Chan, 2010). In the past, konjac flour has
Refined konjac flour (refined KF), a kind of powder that is pu- been consumed as medicinal material and traditional food in the
rified from the tuber of amorphophallus, is mainly farmed in form of konjac jelly, tofu, noodles (Joyce Keithley and Swanson,
Southeast Asia and India (Li, Xia, Wang, & Xie, 2005). After har- 2005). After the methodology for purifying konjac was proposed
vesting, fresh konjac tuber is washed, sliced, dried and grounded. by Mashiko, the application of refined KF was greatly broadened
Then the crude powder is further purified by mechanical ways or/ (Chua et al., 2010). The refined konjac flour occupies the majority of
and alcohol wash so as to remove impurities and thus the refined the konjac market. Every year, about 1,200,000 tons refined konjac
konjac flour was obtained (Liu, 2004, chap. 19). Refined KF con- flour is produced in the world (Sheng & Teng, 2008). Due to its high
tained a large amount of konjac glucomannan (KGM), a natural water-absorbing capacity, konjac flour is commonly used as thick-
neutral polysaccharide consisting of mannose and glucose with b- ening and gelling agent (Saha & Bhattacharya, 2010). Recently,
1-4 linkage. Its molar ratio as well as acetyl groups along the KGM much more attention has been paid to refined KF as functional
backbone might differ depending on the original source of KGM products for its potential value, being a dietary fiber with functions
(An, Thien, Dong, Dung, & Du, 2010; Chen, Zong, & Li, 2006; of lowering blood sugar, removing blood fat and cholesterol, as well
Nguyen, Do, Nguyen, Pham, & Nguyen, 2011). as regulating intestinal flora (Jimenez-Colmenero et al., 2012; Joyce
Keithley and Swanson, 2005; Li et al., 2005; Sood, Baker, &
Coleman, 2008). Moreover, application in the field of material and
* Corresponding author. College of Food Science and Technology, Huazhong
medicine also achieved some degree of development, such as ex-
Agricultural University, Wuhan 430070, China. cipients for floating system, films for wound dressing and
E-mail address: libinfood@mail.hzau.edu.cn (B. Li).

http://dx.doi.org/10.1016/j.foodhyd.2016.01.017
0268-005X/© 2016 Elsevier Ltd. All rights reserved.
210 Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216

microspheres for low-cost desalting (Huang, Chu, Huang, Wu, & amount of amylase was added at room temperature till enzymatic
Tsai, 2015; Ji & Deng, 2011; Xiong et al., 2014). These recent de- hydrolysis was finished for removing amylum. Besides, protein
velopments in food, material and medical fields give rise to the was removed by the enzymatic hydrolysis, combining Sevag
increasing demand for konjac and it is urgent to fully characterize method with centrifugation for five times, and then water phase
the properties of konjac in order to fulfill its market potential. was collected. After that, the sample was deposited by adding
Konjac is an Araceae perennial herb, and there are about 170 same volume 95% ethanol (V/V) and then washed with absolute
kinds of konjac in the globe as previously reported (Hetterscheid & ethanol and aether. Finally, KGM was obtained after vacuum
Ittenbach, 1996). But konjac flours extracted form Amorphophallus drying.
rivirei and A. albus dominated the konjac market in the world. In The molecular weight of KGM was determined by the method of
order to satisfy the growing demand for konjac flour, exploring new (Jin et al., 2014). In brief, 0.06 wt% purified KGM was prepared by
varieties, such as Amorphophallus guripingensis, became an inevi- mechanical stirring and diluted into a series of concentration with
table trend. Compared with the commercial varieties (A. rivirei and 0.2 M NaCl solution. Before determination, the sample was filtered
A. albus), A. guripingensis has several outstanding advantages, such through a 0.45 mm Millipore filter to remove insoluble impurities.
as high propagation coefficient, disease resistance and adaptability Then viscosity-average molecular weight of purified KGM was
to hot or humid environment, which qualify it a potential com- measured using an Ubbelohde viscometer (Liangjing Co., Shanghai,
modity. But the rarely research about A. guripingensis seriously China) at 25 ± 0.1  C. All the measurements were performed in six
limited the application and development of new food ingredients. replicates.
Although differences in chemical composition and KGM structure The acetyl content was quantified by the modified means of (T.
between species have been reported previously (An et al., 2010; Zhang, Xue, Li, Wang, & Xue, 2015). Four grams dried purified KGM
Fang & Wu, 2004), rheological properties between varieties was was placed in a 150 ml Erlenmeyer flask with a stopper and 40 ml
scarcely conducted and they directly influenced practical usage of 75% ethanol was added. The flask was shaken in a thermostatic
konjac flour in the source of A. guripingensis. oscillator (its agitation speed was 150 rpm) for 30 min at 40  C.
The present study was aimed to understand the characteristics After that, 5 ml 0.5 M NaOH was added and the flask was shaken
of A. guripingensis refined flour, enrich the basic knowledge of its continually at 40  C for 24 h. The excess alkali was titrated back
rheological properties and broaden the application. In the with 0.1 M HCl using phenolphthalein as an indicator. The blank, no
comparative studies, chemical compositions and rheology are two KGM sample added, was titrated in parallel. All experiments were
major aspects for investigation. This study would provide theo- carried out three times at least.
retical and practical information on developing new konjac based
food ingredients.
2.4. Rheological properties
2. Materials and methods
2.4.1. Samples preparation
2.1. Materials To investigate the effect of concentration on rheological
behavior, two kinds of refined KF of various concentrations
Two kinds of refined konjac flour that extracted from including 3.5 wt%, 4.0 wt%, 4.5 wt%, 5.0 wt%, 5.5 wt% (w/w)
A. guripingensis and A. rivirei were supplied by QiangSen Biotech were completely dissolved in distilled water with mechanical
Co (Wuhan, China). In chemical analysis and rheological testes, stirring at 30  C for 2 h. Then the solution was filtered with
they were used directly without any further purification. In the 300 mesh gauze to remove insoluble impurities, and degassed by
KGM molecular structure measurements, KGM was purified from vacuum.
two kinds of refined konjac flour until the purity reached at least To study the effect of pH on rheological behavior, the concen-
90%. All of other chemicals used were of analytical grade and tration was fixed at 4.0 wt%. The solution was prepared with similar
purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, methods stated before. Different pH values including 2, 4, 6, 8, 10
China). were adjusted with 0.1 M sodium hydroxide and 0.1 M hydrochloric
acid.
2.2. Chemical compositions analysis

For the two kinds of refined KF, the contents of moisture, ash, 2.4.2. Rheological measurements
protein, and soluble sugar were analyzed according to AOAC All rheological measurements were performed using TA AR
methods (Horwitz, 2000). The dry basis content of KGM was 2000ex controlled strain rheometer with peltier plate system.
measured as previously reported (Xu et al., 2014). Freshly prepared samples were equilibrated at 25  C for 20 min
before rheological measurements. Steady-state flow curves were
2.3. Molecular parameters of konjac glucomannan obtained using 2 cone plates (40 mm diameter and 0.052 mm gap)
at 25  C, over the shear rate range of 0.1e300 s1.
Preparation and purification of KGM were performed accord- Dynamic oscillatory measurements were performed using par-
ing to (Jian et al., 2013). Briefly, 10 g refined KF was washed with allel plates (40 mm diameter, 1 mm gap). Frequency sweeps were
50 ml 50% ethanol containing 0.1% sodium azide for three times conducted in the 0.1e500 rad/s range and the strain was fixed at 2%
so that water-soluble impurities were removed. Then the flour in order to assure the working conditions within the linear visco-
was dried in air. After that, it was placed in 50 ml mixed solvent elastic region. Temperature sweeps were carried out from 25  C to
of absolute aether/absolute ethanol [2:1(V/V)] at 40  C with 90  C at a constant heating rate of 2  C/min. Meanwhile, the strain
mechanical stirring for 8 h in order to remove grease. Secondly, and frequency were fixed at 2% and 6.283 rad/s respectively. After
the degreased sample was dissolved in distilled water to be a 0.6% the sample was placed into the measuring device, a thin layer of
(W/V) solution with mechanical stirring at 30  C water bath for paraffin oil was spread over the outer edge of the sample to prevent
2 h. The fibrin and other insoluble compositions were separated moisture evaporation. The measurements were conducted in trip-
by over-speed freezing centrifuge at a rate of 16,000 r/min for licate. Viscosity, storage modulus (G0 ) and loss modulus (G00 ) were
20 min. Then the supernatant was collected and appropriate recorded.
Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216 211

3. Results and discussion Table 2


Main parameters of konjac glucomannan.

3.1. Composition analysis A. guripingensis A. rivirei

Molecular weight of KGM 1.236  106 1.669  106


The physicochemical properties were significantly affected by Acetyl content (%) 1.342 ± 0.033 2.285 ± 0.026
compositions of refined KF. Their main compositions were pre-
sented in Table 1, including the content of KGM, moisture, ash,
protein and soluble sugar. It was suggested that the content of KGM Wilson, 2014). Moreover, the viscosity of the refined KF solution
in refined AGF (61.347 ± 2.293%) was lower than that of ARF increased obviously with increase of concentration. This phenom-
(72.801 ± 1.945%). And the other content (not determined), espe- enon was mainly attributed to intermolecular interactions between
cially insoluble polysaccharides, such as starch, cellulose (the main konjac glucomannan (the main component of refined KF), such as
component of the cell wall), in refined AGF was higher compared hydrogen bonds, caused by massive hydroxyl groups. With
with refined ARF. Besides, the parameters of konjac glucomannan increasing concentration, the molecular chains within the unit
from two species were shown in Table 2. It was observed that the space increased, resulting in obvious stronger intermolecular
molecular weight of KGM in refined AGF and ARF were 1.236  106 hydrogen bonds and then more entangled points between molec-
and 1.669  106, respectively. Furthermore, the molecular weight ular chains appeared. Finally, the more entangled three-
was directly associated with solution viscosity (Jin et al., 2014) and dimensional network formed, which led to the higher viscosity
various viscosities were needed in different foods so that the above (Hua et al., 2015; Wu et al., 2015). When the concentration was
data might contribute to guide the industrial usage of konjac flour. lower than 4.5 wt%, the shear viscosity of refined AGF was lower
What's more, refined AGF possessed lower acetyl groups compared than that of refined ARF, while at 4.5 wt% (or above) concentration,
with ARF and acetyl content was closely related to gel-forming the shear viscosity of both KF were equivalent, which might be due
property (Du, Li, Chen, & Li, 2012; Gao & Nishinari, 2004a, to the difference between the plant cell wall of two species. At low
2004b). It was commonly considered that KGM with lower acetyl concentration, the friction force between the A. guripingensis par-
groups was easier to form a gel. These differences in composition ticles cell was so small that some of the particles did not swell and
and parameters of KGM were attributed to species, regions, soil extend completely. Thus small parts of undissolved KGM molecules
environment and so on (Liu, 2004, chap. 19). were also removed by filtration along with impurities together.
However, A. rivirei particles were in the opposite situation.
Accordingly, the varieties exhibited differences on the viscosity.
3.2. Rheological analysis The frequency dependence of AGF and ARF solutions' G0 and G00
with various concentrations from 3.5 wt% to 5.5 wt% at 25  C was
3.2.1. Effect of concentration on rheological properties presented in Fig. 2. It was found that both of G0 and G00 rose with
The flow curves of the refined AGF and ARF at the different increasing frequency and a crossover could be detected, which
concentrations were studied and the results were presented in could be explained by assuming that the entanglement of molec-
Fig. 1. The sample showed a remarked shear-thinning behavior at ular chain was transient instead of permanent. At low frequencies,
low shear rate and the viscosity finally remained constant at high KGM molecular chains disentangled during a long period oscilla-
shear rate, a typical pseudo-plastic fluid. Similar rheological tion, so that G00 was higher than G0 , indicating a liquid-like behavior.
observation had also been reported in previous study about puri- Whereas at higher frequencies, molecular chains had no enough
fied KGM (Wang et al., 2012). This type of behavior was expected for time to disentangle during a short period of oscillation. Therefore,
most polymer solutions (Hua, Wang, Yang, Kang, & Zhang, 2015; G0 was higher than G00 , showing a solid-like behavior (Du et al.,
Wu, Ding, Jia, & He, 2015). The phenomenon could be explained 2012; Jin et al., 2014).
by that shear force disturbed the polysaccharides' entangled In Fig. 2, with an increase of concentration, the crossover points
molecule networks and decreased the effective size of polymers of the refined AGF and ARF shifted to lower frequency from
(Bhandari, Singhal, & Kale, 2002). But molecular chains were 20.508 rad/s to 5.290 rad/s and from 2.746 rad/s to 1.721 rad/s,
rearranged and aligned with the orientation of the shear force, respectively, which indicated the enhanced entangled network (Jin
resulting in the unchanged viscosity (Hua et al., 2015). et al., 2014). Similar phenomenon was observed for other hydro-
The steady shear viscosity of the two kinds of refined KF as a colloids, such as tamarind seed gum (Khounvilay and Sittikijyothin,
function of shear rate for different concentrations varying from 2012) and Tara gum (Wu et al., 2015). It was contributed to that
3.5 wt% to 5.5 wt% was presented in Fig. 1a and b. It was demon- molecular chains of KGM got closer to each other at higher con-
strated that the shear-thinning behavior was more remarkable at centration and that the probability of the formation of a junction
25  C with the increase of concentration. This phenomenon was in zone was higher (Cai & Zhang, 2006). Moreover, the crossover
good agreement with those found for other solutions (Gong et al., points of refined ARF (Fig. 2a) were lower than AGFs' (Fig. 2b) at the
2012; Khounvilay & Sittikijyothin, 2012; Torres, Hallmark, & same concentration. The larger molecular weight and higher con-
tent of KGM in the refined ARF compared with AGF might account
Table 1 for these phenomena (Cai and Zhang, 2006; Jin et al., 2014; Zhong
Main compositions of refined konjac flour (%). et al., 2015). As the number of KGM molecules and length of the
KGM chain increased, their relaxation capacity decreased, resulting
A. guripingensis A. rivirei
in easier entanglement.
a
KGM (%) 61.35 ± 2.293 72.80 ± 1.945 Temperature sweeps were used to explain behavior upon
Moisture (%) 11.69 ± 0.009 11.97 ± 0.034
Ash (%) 4.262 ± 0.002 4.765 ± 0.008
heating in tested system. The effect of temperature on G0 and G00 of
Protein (%) 1.458 ± 0.006 4.685 ± 0.013 the refined AGF and refined ARF at various concentrations upon
Soluble sugar (%) 1.878 ± 0.015 3.310 ± 0.041 heating from 25  C to 90  C was shown in Fig. 3a and Fig. 3b,
Othersb e e respectively. In Fig. 3a, with increasing temperature, G0 and G00
e Not determined. decreased and departed. Meanwhile, G0 was lower than G00 ,
showing a fluid-like behavior and it was more pronounced at
a
Dry basis content.
b
Starch, cellulose (the main component of the cell wall) and so on.
212 Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216

Fig. 1. Viscosity-shear rate profiles of different concentrations from two species at shear rate from 0.1 s1 to 300 s1 (T ¼ 25  C). (a) the refined AGF; (b) the refined ARF.

Fig. 2. Frequency dependence of G0 and G00 for the refined KF solution at various concentrations from 3.5 wt% to 5.5 wt%. (a) the refined AGF; (b) the refined ARF. The data were
shifted along the vertical axis by 10a to avoid overlapping.

Fig. 3. Temperature dependence of G0 and G00 for the refined KF solution at various concentrations from 3.5 wt% to 5.5 wt%. (a) the refined AGF; (b) the refined ARF. The data were
shifted along the vertical axis by 10a to avoid overlapping.

higher temperature. It might be supported by the weakening of the might be due to the low acetyl content and strong hydrophobic
hydrogen bond, regarded as primary molecular force (T. Zhang interactions (predominated molecular force at an increasing tem-
et al., 2015) and the declining of chain entanglements caused by perature), giving rise to the aggregation of molecular chains of KGM
thermal disruption (Arancibia et al., 2015). As heating continued, G0 and the subsequent formation of junction zones or molecular chain
got closer to G00 and finally became higher than G00 when the con- clusters (Du et al., 2012; T. Zhang et al., 2015). Nevertheless, the
centration was above 4.5 wt%. It was demonstrated that the temperature-dependent G0 and G00 of the refined ARF were different
behavior of the refined AGF solution was shifted from fluid-like to (Fig. 3b). Both G0 and G00 decreased with increase of temperature
solid-like. In addition, fluids-like behavior was maintained over the from 25  C to 90  C. At low temperature, G0 was higher than G00 at all
whole tested temperature interval (25e90  C) at low concentra- investigated concentrations, showing a solid-like behavior, since a
tions, while the crossover points (G0 ¼ G00 ) were detected at con- weak gel network was already formed by hydrogen bonds. How-
centration higher than 4.5 wt% at particular temperatures (86.2  C, ever, when the temperature rose unceasingly, hydrogen bonds
84.7  C, 79.9  C, respectively). And there was a tendency to decrease were destroyed, resulting in the disruption of network and the
in temperature (crossover happened) as concentration increased. It weak gel melt to sol (G0 < G00 ).
Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216 213

3.2.2. Effect of pH on rheological properties Most importantly, the magnitude of overall forces was slightly
KGM was a neutral polysaccharide so that the rheological changed. Whereas the intersection value of each frequency sweep
properties of the sample should not be affected by pH among the for refined ARF was almost unchanged (Fig. 5b), probably because
tested range (2e10) obviously. But the results of the refined KF pH-insensitive KGM possessed the vast majority of the refined ARF.
were different. Viscosity-shear rate profiles of the refined KF were The thermo-rheological behaviors of the refined KF at different
presented in Fig. 4. It was worth noting that the samples still kept a pH were explored (Fig. 6). It was suggested that the influence of pH
shear-thinning behavior at various pH values as a result of the on rheological behavior of refined AGF was more remarkable than
breakup of intermolecular entanglements. This phenomenon was that of ARF. As it was clearly shown, in Fig. 6a, during the initial
similar to the effect of concentration. In Fig. 4a, the viscosity of the heating process, G0 and G00 at pH varying from 4 to 10 decreased
refined AGF was highly dependent on pH. However, there was with increasing temperature and G0 was lower than G00 , indicating a
almost no effect of pH on viscosity of refined ARF except for strong- viscous behavior. However, as the temperature rose continually, G0
acid condition (Fig. 4b). As for the refined AGF, the viscosity un- and G00 increased and the crossover point was observed, showing an
derwent a sharp increase from 348.8 Pa.s to 1522.0 Pa.s when pH elastic behavior at high temperature. Therefore, there was a solegel
decreased from 4 to 2 and increased obviously from 69.7 Pa.s to transition in the whole range of investigated temperature. This
594.5 Pa.s with pH increasing from 8 to 10. Besides, the viscosity phenomenon could be explained by the nature of thermal-induced
changed markedly at pH among 4 to 8. Since there was still a small gelation of konjac glucomannan (H. Zhang et al., 2001). Moreover,
portion of other ingredients and the extent affected by the external the crossover point was pH-dependent, owing to the different level
environment were also different. At acidic condition, lowing pH of entanglements and associations at pH ranging from 4 to 10.
could cause the protonation of carboxyl groups on polymers, just However, when pH value was low to 2, the weak gel was formed
like protein in the refined KF, facilitating the formation of inter- already and the gelesol transition happened until the temperature
molecular hydrogen bonding and leading to the increase of vis- rose to 45.9  C (Gigli, Garnier, & Piazza, 2009; Sittikijyothin,
cosity (Hua et al., 2015). On the other hand, KGM molecular chains Sampaio, & Gonçalves, 2007). Under strong acid condition, elec-
were broken by strong acid, resulting in the decrease of viscosity trostatic repulsion was suppressed and hydrogen-bonding in-
(Kato & Matsuda, 1969). At alkaline condition, the hydroxyl groups teractions dominated, which promoted the formation of a three-
on the polymer chains were ionized (Wu et al., 2015), giving rise to dimensional network structure, i.e. a weak gel was formed. When
the disruption of intra and intermolecular hydrogen bonding, thus the temperature rose to 45.9  C, hydrogen-bonding interactions
viscosity of sample solution decreased. For another, deacetylation were disrupted, the network structure was destroyed and the weak
occurred (Du et al., 2012; T. Zhang et al., 2015), resulting in the self- gel was transformed into sol.
association of KGM chains and giving rise to the higher viscosity. As for refined ARF (Fig. 6b), when the temperature rose gradu-
Consequently, the difference of viscosity at various pH values might ally within the tested temperature (from 25  C to 90  C), G0 and G00
be the combined results of above described factors. Nevertheless, of the sample at every pH decreased and an intersection was
the exact reasons for these behaviors were not clear so far and detected at about 50  C and gelesol transition happened, which
much more studies were needed. was similar to the refined AGF at pH 2. Moreover, the temperature
G0 and G00 of two refined KF solutions at different pH as a of the crossover detected was higher than that of refined AGF at pH
function of frequency were shown in Fig. 5. For all refined KF so- 2. This might due to the higher KGM content and larger molecular
lutions, G0 and G00 increased with an increase of frequency. More- weight in refined ARF, thus entanglements and associations
over, frequency at the intersection of refined AGF was larger than occurred even at weak acidic or alkaline conditions. Much more
that of ARF at same pH value, except under pH 2. Compared with ordered network was formed and therefore more energy to destroy
refined ARF, lower KGM content and smaller molecular weight in this network structure was needed.
AGF (Tables 1 and 2tbl2) might explain this phenomenon. In Fig. 5a, In fact, it was universally acknowledged that the addition of
the crossover point of G0 and G00 shifted to lower frequency at low alkali results in the deacetylation of KGM and finally leads to
pH (0.571 rad/s at pH 2), owing to the enhanced intermolecular gelation at elevated temperature, which was a time-independent
hydrogen bonds, intra and intermolecular entanglements and as- process. However, during this process, konjac concentration, mo-
sociations (Bu, Kjøniksen, & Nystro € m, 2005). Nevertheless, the lecular weight, degree of acetylation, alkali concentration, tem-
crossover point changed slightly at high pH values. This might be perature and heating time played an important role (Gao &
attributed to the reason that electrostatic interactions and Nishinari, 2004a, 2004b; Hirai, 1954; Huang, Takahashi,
hydrogen bonding interactions changed with the media pH values. Kobayashi, Kawase, & Nishinari, 2002; Zhang et al., 2001). These

Fig. 4. Viscosity-shear rate profiles of two kinds of refined KF solution (4.0 wt%) under different pH at shear rate from 0.1 s1 to 300 s1 (T ¼ 25  C). (a) the refined AGF; (b) the
refined ARF.
214 Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216

Fig. 5. Frequency dependence of G0 and G00 of two kinds of refined KF solution (4.0 wt%) at different pH. (a) the refined AGF; (b) the refined ARF. The data were shifted along the
vertical axis by 10a to avoid overlapping.

Fig. 6. Temperature dependence of G0 and G00 for two kinds refined KF solution (4.0 wt%) under different pH from 2 to 10. (a) the refined AGF; (b) the refined ARF. The data were
shifted along the vertical axis by 10a to avoid overlapping.

factors finally determined the gelation time. Besides, there are could be explained by the increased molecular distance (Kim & Yoo,
distinct differences both in the main composition of the refined 2009; Ma, Lin, Chen, Zhao, & Zhang, 2014; Xiao, Tong, & Lim, 2012)
konjac flour and the main parameters of konjac glucomannan and reduced hydrogen bonding force (Du et al., 2012) as a result of
extracted from two species (Tables 1 and 2tbl2). Hence, the gelation elevated temperature. As for hydrophilic polymer solutions, it was a
behaviors might vary from species to species. In Fig. 6a, the gels of common phenomenon that viscosity dropped with raising tem-
refined AGF (pH ¼ 4e10) were formed primarily by hydrophobic perature, such as carboxymethyl guar gum (Gong et al., 2012;
interactions at elevated temperature within the short time, which Shafiei-Sabet, Hamad, & Hatzikiriakos, 2012). When the tempera-
were thermo-irreversible gels (“real gel”). Furthermore, due to its ture reached up to 85  C, the viscosity of ARF also decreased
relative low acetyl content, hydrophobic interaction was strength- continually. Conversely, viscosity of AGF increased sharply, reach-
ened (Du et al., 2012), thus the gels were also formed in the tested ing 477.8 Pa.s, and even exceeded that of ARF. This might due to the
time even at weak acidic condition without completely deacetyla- relative low content of acetyl groups along the backbone of KGM in
tion. As for the refined ARF (in Fig. 6b) and refined AGF (pH ¼ 2), the refined AGF so that the hydrophobic interaction between molecules
gels were formed by hydrogen bonding interactions at low tem- enhanced significantly at 85  C, resulting in the increased viscosity.
perature, which were thermo-reversible gels (“fake gel”). Once the The dynamical oscillatory frequency sweep tests were carried
gel was subject to the high temperature, the weak gel was dis- out to determine the viscoelastic behavior at angular frequencies
rupted. But with the increasing temperature within the tested ranging from 0.1 to 500 rad/s. G0 and G00 for 4.0 wt% refined KF at
range, the thermo-irreversible gels were not formed within the different temperatures were found to increase with increasing
short time for the sake of high acetyl content. angular frequency (Fig. 8). Furthermore, in Fig. 8a, at temperature
below 85  C, G00 was larger than G0 at low frequencies, showing the
fluid-like behavior. There was a tendency for G0 and G00 to approach
3.2.3. Effect of temperature on rheological properties each other and the crossover was detected at high frequency. When
To study the effect of temperature on rheological behavior of the frequency increased continually, G00 was lower than G0 , indi-
refined KF, steady shear viscosity was measured at different tem- cating that the solid-like behavior dominated in the tested system
peratures and the results were presented in Fig. 7. For the 4.0 wt% (Du et al., 2012; Jin et al., 2014). Interestingly, G0 and G00 increased
refined KF solution (Fig. 7a and b), a typical pseudo-plastic behavior rapidly and showed a significant frequency-independent plateau at
was also observed within the investigated temperatures at shear 85  C. Besides, G0 was greater than G00 , showing an elastic behavior
rate ranging from 0.1 s1 to 300 s1. Besides, it was obvious that at all tested frequencies. This implied that the stable gel network
when the temperature was below 65  C, the viscosity of refined AGF structure was formed (Cai and Zhang, 2006). In Fig. 8b, the cross-
were less than that of ARF and both of them decreased with over frequency of G0 and G00 shifted gradually to a lower value with
increasing temperature from 1692.0 Pa.s to 239.6 Pa.s and an increase of temperature, from 1.184 to 17.241 rad/s, indicating an
968.9 Pa.s to 197.1 Pa.s, respectively. These declination in viscosity
Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216 215

Fig. 7. Viscosity-shear rate profiles of two kinds of refined KF solutions (4.0 wt%) at different temperature with shear rate ranging from 0.1 s1 to 300 s1. (a) the refined AGF; (b) the
refined ARF.

Fig. 8. Frequency dependence of G0 and G00 of two kinds of refined KF solution (4.0 wt%) at different temperatures. (a) the refined AGF; (b) the refined ARF. The data were shifted
along the vertical axis by 10a to avoid overlapping.

enhancement of temporary network structure of polymer chains Environment Correlative Dietology of Huazhong Agricultural Uni-
(Du et al., 2012). versity for offering many conveniences.

4. Conclusions References

An, N. T., Thien, D. T., Dong, N. T., Dung, P. L., & Du, N. V. (2010). Characterization of
The refined AGF (61.35 ± 2.293%) possessed lower content of
glucomannan from some Amorphophallus species in Vietnam. Carbohydrate
KGM in comparison with that of refined ARF (72.80 ± 1.945%). The Polymers, 80(1), 308e311.
content of acetyl groups along the backbone of KGM in AGF Arancibia, M. Y., Lopez-Caballero, M. E., Go  mez-Guille
n, M. C., Fern
andez-García, M.,
(1.342 ± 0.033%) was also lower than that of ARF (2.285 ± 0.026%). Fern andez-Martín, F., & Montero, P. (2015). Antimicrobial and rheological
properties of chitosan as affected by extracting conditions and humidity
But the molecular weights of KGM from two species were similar, exposure. LWT e Food Science and Technology, 60(2), 802e810.
1.236  106 and 1.669  106 respectively. Moreover, the viscosities Bhandari, P. N., Singhal, R., & Kale, D. (2002). Effect of succinylation on the rheo-
of refined AGF at low concentrations (below 4.5 wt%) were lower logical profile of starch pastes. Carbohydrate Polymers, 47(4), 365e371.
Bu, H., Kjøniksen, A.-L., & Nystro €m, B. (2005). Effects of pH on dynamics and
than that of refined ARF, while at concentration of 4.5 wt% (or rheology during association and gelation via the Ugi reaction of aqueous algi-
above), viscosities of both species were equivalent. These had nate. European Polymer Journal, 41(8), 1708e1717.
important implications for guiding the practical application of Cai, J., & Zhang, L. (2006). Unique gelation behavior of cellulose in NaOH/urea
aqueous solution. Biomacromolecules, 7(1), 183e189.
refined KF. Furthermore, rheological properties of refined AGF were Chen, Z.-G., Zong, M.-H., & Li, G.-J. (2006). Lipase-catalyzed acylation of konjac
susceptible to pH, especially at pH 2, giving rise to the potential glucomannan in organic media. Process Biochemistry, 41(7), 1514e1520.
value as pH-sensitive material. Nevertheless, Properties of refined Chua, M., Baldwin, T. C., Hocking, T. J., & Chan, K. (2010). Traditional uses and po-
tential health benefits of Amorphophallus konjac K. Koch ex N.E.Br. Journal of
ARF were nearly not influenced by the external environment,
Ethnopharmacology, 128(2), 268e278.
which contributed to a wide range of application in various fields. Du, X., Li, J., Chen, J., & Li, B. (2012). Effect of degree of deacetylation on physico-
Furthermore, the thermo-gelling temperature of refined AGF was chemical and gelation properties of konjac glucomannan. Food Research Inter-
national, 46(1), 270e278.
lower than that of refined ARF, which would be a pronounced
Fang, W., & Wu, P. (2004). Variations of Konjac glucomannan (KGM) from amor-
advantage for A. guripingensis to be widely used in food industry. phophallus konjac and its refined powder in China. Food Hydrocolloids, 18(1),
167e170.
Gao, S., & Nishinari, K. (2004a). Effect of deacetylation rate on gelation kinetics of
Acknowledgments
konjac glucomannan. Colloids and Surfaces B: Biointerfaces, 38(3), 241e249.
Gao, S., & Nishinari, K. (2004b). Effect of degree of acetylation on gelation of konjac
This work was financially supported by Key Project of Chinese glucomannan. Biomacromolecules, 5(1), 175e185.
Ministry of Education (Grant No. 113047A) and the National Natural Gigli, J., Garnier, C., & Piazza, L. (2009). Rheological behaviour of low-methoxyl
pectin gels over an extended frequency window. Food Hydrocolloids, 23(5),
Science Foundation of China (Grant No. 31371841). The authors 1406e1412.
expressed greatly thank to colleagues of Key Laboratory of Gong, H., Liu, M., Chen, J., Han, F., Gao, C., & Zhang, B. (2012). Synthesis and
216 Q. Huang et al. / Food Hydrocolloids 57 (2016) 209e216

characterization of carboxymethyl guar gum and rheological properties of its Ma, J., Lin, Y., Chen, X., Zhao, B., & Zhang, J. (2014). Flow behavior, thixotropy and
solutions. Carbohydrate Polymers, 88(3), 1015e1022. dynamical viscoelasticity of sodium alginate aqueous solutions. Food Hydro-
Hetterscheid, W., & Ittenbach, S. (1996). Everything you always wanted to know colloids, 38, 119e128.
about Amorphophallus, but were afraid to stick your nose into. Aroideana, 19, Nguyen, T. A., Do, T. T., Nguyen, T. D., Pham, L. D., & Nguyen, V. D. (2011). Isolation
7e131. and characteristics of polysaccharide from Amorphophallus corrugatus in
Hirai, N. (1954). Gelation behaviour of konjac glcomannan with different concen- Vietnam. Carbohydrate Polymers, 84(1), 64e68.
tration. Nippon Kagaku Zasshi, 75, 65. Saha, D., & Bhattacharya, S. (2010). Hydrocolloids as thickening and gelling agents
Horwitz, W. (2000). AOAC official methods of analysis (Vol. 1, pp. 9e10). in food: a critical review. Journal of Food Science and Technology, 47(6), 587e597.
Huang, Y. C., Chu, H. W., Huang, C. C., Wu, W. C., & Tsai, J. S. (2015). Alkali-treated Shafiei-Sabet, S., Hamad, W. Y., & Hatzikiriakos, S. G. (2012). Rheology of nano-
konjac glucomannan film as a novel wound dressing. Carbohydrate Polymers, crystalline cellulose aqueous suspensions. Langmuir, 28(49), 17124e17133.
117, 778e787. Sheng, D. X., & Teng, J. X. (2008). Present status and future of konjac industry. China
Huang, L., Takahashi, R., Kobayashi, S., Kawase, T., & Nishinari, K. (2002). Gelation Agricultural Information, 7, 39e40.
behavior of native and acetylated konjac glucomannan. Biomacromolecules, Sittikijyothin, W., Sampaio, P., & Gonçalves, M. P. (2007). Heat-induced gelation of b-
3(6), 1296e1303. lactoglobulin at varying pH: effect of tara gum on the rheological and structural
Hua, X., Wang, K., Yang, R., Kang, J., & Zhang, J. (2015). Rheological properties of properties of the gels. Food Hydrocolloids, 21(7), 1046e1055.
natural low-methoxyl pectin extracted from sunflower head. Food Hydrocol- Sood, N., Baker, W. L., & Coleman, C. I. (2008). Effect of glucomannan on plasma lipid
loids, 44, 122e128. and glucose concentrations, body weight, and blood pressure: systematic re-
Jian, W., Sun, Y., Huang, H., Yang, Y., Peng, S., Xiong, B., et al. (2013). Study on view and meta-analysis. The American journal of clinical nutrition, 88(4),
preparation and separation of konjac oligosaccharides. Carbohydrate Polymers, 1167e1175.
92(2), 1218e1224. Torres, M. D., Hallmark, B., & Wilson, D. I. (2014). Effect of concentration on shear
Ji, Y., & Deng, Y. (2011). In vitro evaluation of Konjac glucomannan as novel ex- and extensional rheology of guar gum solutions. Food Hydrocolloids, 40, 85e95.
cipients for floating systems. Journal of Controlled Release, 152, e34ee36. Wang, C., Xu, M., Lv, W.-P, Qiu, P., Gong, Y.-y., & Li, D.-S (2012). Study on rheological
Jime nez-Colmenero, F., Cofrades, S., Herrero, A. M., Ferna ndez-Martín, F., Rodríguez- behavior of konjac glucomannan. Physics Procedia, 33, 25e30.
Salas, L., & Ruiz-Capillas, C. (2012). Konjac gel fat analogue for use in meat Wu, Y., Ding, W., Jia, L., & He, Q. (2015). The rheological properties of tara gum
products: comparison with pork fats. Food Hydrocolloids, 26(1), 63e72. (Caesalpinia spinosa). Food Chemistry, 168, 366e371.
Jin, W., Xu, W., Li, Z., Li, J., Zhou, B., Zhang, C., et al. (2014). Degraded konjac glu- Xiao, Q., Tong, Q., & Lim, L.-T. (2012). Pullulan-sodium alginate based edible films:
comannan by g-ray irradiation assisted with ethanol: preparation and charac- rheological properties of film forming solutions. Carbohydrate Polymers, 87(2),
terization. Food Hydrocolloids, 36, 85e92. 1689e1695.
Joyce Keithley, D., & Swanson, B. (2005). Glucomannan and obesity: a critical re- Xiong, Z., Zhou, W., Sun, L., Li, X., Zhao, D., Chen, Y., et al. (2014). Konjac gluco-
view. Alternative Therapies, 11(6), 30e34. mannan microspheres for low-cost desalting of protein solution. Carbohydrate
Kato, K., & Matsuda, K. (1969). Studies on the chemical structure of konjac mannan: Polymers, 111, 56e62.
part I. isolation and characterization of oligosaccharides from the partial acid Xu, W., Wang, Y., Jin, W., Wang, S., Zhou, B., Li, J., et al. (2014). A one-step procedure
hydrolyzate of the mannan. Agricultural and Biological Chemistry, 33(10), for elevating the quality of konjac flour: azeotropy-assisted acidic ethanol. Food
1446e1453. Hydrocolloids, 35, 653e660.
Khounvilay, K., & Sittikijyothin, W. (2012). Rheological behaviour of tamarind seed Zhang, T., Xue, Y., Li, Z., Wang, Y., & Xue, C. (2015). Effects of deacetylation of konjac
gum in aqueous solutions. Food Hydrocolloids, 26(2), 334e338. glucomannan on Alaska Pollock surimi gels subjected to high-temperature
Kim, W.-W., & Yoo, B. (2009). Rheological behaviour of acorn starch dispersions: (120  C) treatment. Food Hydrocolloids, 43, 125e131.
effects of concentration and temperature. International Journal of Food Science & Zhang, H., Yoshimura, M., Nishinari, K., Williams, M., Foster, T., & Norton, I. (2001).
Technology, 44(3), 503e509. Gelation behaviour of konjac glucomannan with different molecular weights.
Liu, P. Y. (2004). Konjac (1th ed.). Beijing: China Agriculture Press. Biopolymers, 59(1), 38e50.
Li, B., Xia, J., Wang, Y., & Xie, B. (2005). Structure characterization and its antiobesity Zhong, K., Zhang, Q., Tong, L., Liu, L., Zhou, X., & Zhou, S. (2015). Molecular weight
of ball-milled konjac flour. European Food Research and Technology, 221(6), degradation and rheological properties of schizophyllan under ultrasonic
814e820. treatment. Ultrasonics Sonochemistry, 23, 75e80.

You might also like