Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Carbohydrate Polymers 250 (2020) 116840

Contents lists available at ScienceDirect

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

Changes in microstructure and rheological properties of konjac T


glucomannan/zein blend film-forming solution during drying
Chong Lia,b, Fei Xianga,b, Kao Wua,b, Fatang Jianga,b,c, Xuewen Nia,b,*
a
Glyn O. Philips Hydrocolloid Research Centre at HUT, Hubei University of Technology, Wuhan, 430068, China
b
School of Bioengineering and Food Science, Hubei University of Technology, Wuhan, 430068, China
c
Department of Architecture and Built Environment, Faculty of Engineering, University of Nottingham, University Park, Nottingham, NG7 2RD, UK

A R T I C LE I N FO A B S T R A C T

Keywords: During film formation at 60 °C, the microstructure and rheological properties of konjac glucomannan (KGM)
Film-forming process film-forming solution and KGM/zein blend film-forming solution were investigated. The drying process of film-
Microstructure forming solutions was divided into two stages according to the drying curves. Scanning electron microscopy
Rheology showed that KGM chains in the blend solution aggregated into thicker chains and formed a molecular network
Konjac glucomannan
with larger pores. Zein particles grew larger but were homogeneously distributed during drying as observed by
Zein
confocal laser scanning microscopy. The addition of zein improved the thermal stability of the film-forming
solution. As the drying proceeded (up to 8 h), KGM solution exhibited a typical concentrated solution behavior
due to molecular entanglement; whereas the blend solution gradually formed a weak gel after 2 h. Complex
viscosity data for the film-forming solutions were well-fitted by the power-law model. The information obtained
from the study is important for understanding the film-forming mechanism.

1. Introduction assembly, a transition from a rubbery to a vitreous phase, a phase se-


paration, and so on. For example, Liu and Han (2005) found that the
With the increasing concern about global environmental protection film formation of dilute starch solutions included helical formation,
and the rising consumer need for safe and healthy foods, edible or aggregation or gelation, and reorganization of aggregates. The high-
biodegradable films have received great interest as an alternative to amylose (55 %) starch film showed a heterogeneous structure with both
petroleum-derived plastics on packaging usage. Edible or biodegradable amylopectin-rich and amylose-rich phases, and there were no visible
films perform functions as a barrier against oxygen, moisture and solute boundaries existed due to molecular and supramolecular interaction
migration within the food itself or between the food and environment between amylopectin and amylose. Subirade, Kelly, Guéguen, and
(Bourlieu, Guillard, Vallès-Pamiès, Guilbert, & Gontard, 2009; Hassan, Pézolet (1998) reported that the conformational changes of glycinin
Chatha, Hussain, Zia, & Akhtar, 2018). Generally, they are prepared occurred upon the film-forming process, and intermolecular hydrogen
from renewable resources, such as natural polysaccharides, proteins bonding between segments of the β-sheet may act as junction zones
and lipids (Dehghani, Hosseini, & Regenstein, 2018; Hassan et al., forming a network. Ghoshal, Mattea, Denner, and Stapf (2010) found
2018). In recent times, extensive research works from edible or bio- that the spatial heterogeneous distribution of solutes occurred with
degradable films have focused on their composition, structure, thermal, evaporation proceeding during gelatin film formation, and the final film
mechanical and barrier properties (Bourtoom & Chinnan, 2008; Denavi still had structural heterogeneity. Understanding the formation of films
et al., 2009; Hassan et al., 2018; Łupina et al., 2019; Mohajer, Rezaei, & is critical to predict or control their structure and properties.
Hosseini, 2017; Niu, Shao, Chen, & Sun, 2019; Tong, Xiao, & Lim, 2008; Konjac glucomannan (KGM) is a natural water-soluble poly-
Wang et al., 2017), while only a few of them studied the film formation saccharide derived from the tuber of the Amorphophallus konjac plant C.
process, e.g. KGM/ethyl cellulose composite films (Xiao et al., 2016). Koch. It is mainly composed of β-1, 4-linked D-mannosyl and D-glucosyl
The film formation process could highly affect the film characteristics, residues in a ratio of approximately 1.6:1 (Kato & Matsuda, 1969), or
and involves various complex physical or chemical processes. When the 1.5:1 (Smith & Srivastava, 1959), actually, the ratio varies with konjac
solvent is removed due to evaporation, various phenomena may occur, breeds (Kohyama, Iida, & Nishinari, 1993). KGM has bright application
such as solute migration, molecular interaction, polymer chains prospects for edible or biodegradable packaging preparation due to its


Corresponding author at: Glyn O. Philips Hydrocolloid Research Centre at HUT, Hubei University of Technology, Wuhan, 430068, China.
E-mail address: nixuewen@126.com (X. Ni).

https://doi.org/10.1016/j.carbpol.2020.116840
Received 17 December 2019; Received in revised form 28 July 2020; Accepted 28 July 2020
Available online 15 August 2020
0144-8617/ © 2020 Elsevier Ltd. All rights reserved.
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

excellent film-forming capability (Jia, Fang, & Yao, 2009). But pure 2.3. Drying curves
KGM film does not behave well as moisture barriers because it is highly
hydrophilic in nature. Some researches have been conducted to en- The drying curves were obtained by measuring the mass loss during
hance the water barrier properties of KGM film by addition of other the drying period. The weight of the samples was recorded accurately
film-forming components (Li et al., 2015; Rhim & Wang, 2013; Wang using the electronic balance (JE1002, Puchuang metrology Co., Ltd.,
et al., 2017; Wei et al., 2015; Wu et al., 2018). Zein is the major storage Shanghai, China) at the designated time intervals until the film formed.
protein of corn with abundant sources and low prices. It is insoluble in The endpoint of drying was considered to be reached when the moisture
pure water due to the high proportion of non-polar amino acid residues, content of the film was 9.0 % (w.b.). The oven was placed in a con-
and it has great potential in film formation (Panchapakesan, Sozer, trolled environment of 25 °C and 68 % relative humidity to maintain
Dogan, Huang, & Kokini, 2012; Wang & Padua, 2010). Pure zein film consistent experimental conditions. Three replications of each sample
presents good water vapor barrier properties, however, it lacks elasti- were performed, and the data given was an average of these results.
city and is very brittle (Scramin et al., 2011). Using two or more film-
forming polymers of different categories to prepare composite films by 2.4. Scanning electron microscopy (SEM)
physically blending is an effective method to improve the properties of
films. According to our previous research (Li et al., 2019; Ni et al., The microstructures of the samples were observed with a high-re-
2018; Wang et al., 2017), KGM/zein blend films could be formed by solution field emission scanning electron microscope (Su8010, JEOL,
solution casting, which showed better mechanical, thermal, water Co., Ltd., Tokyo, Japan). The film-forming solutions at different drying
vapor and oxygen barrier properties compared to pure KGM film and times were frozen instantaneously with liquid nitrogen and then va-
zein film. The addition of zein to KGM may change the formation of cuum freeze-dried. Then they were cut into small pieces for surface and
KGM molecular network. Further investigation on the KGM/zein film- cross-section observation. The cross-section of pure KGM film and
forming process could contribute to understandings of film structure KGM/zein blend film samples were prepared by breaking samples after
and properties. freezing in liquid nitrogen. Before testing, all samples were coated with
The objective of this work was to understand the microstructural a thin layer of gold powder. The sputtered time was about 15 s and the
development and the change of rheological properties of KGM/zein accelerating voltage was 2 kV.
blend film-forming solution during drying at 60 °C. This information is
essential for explaining the film-forming mechanism, as well as pro- 2.5. Confocal laser scanning microscopy (CLSM)
ducing the films with desirable microstructures that determine their
performance. CLSM (TCS SP8, Leica, Co., Ltd., Germany) was used to observe the
morphology and distribution of zein in KGM/zein blend samples.
Protein staining was based on the method of Li et al. (2019). The zein
2. Materials and methods was stained with rhodamine B. First, 4 mg of rhodamine B was dis-
solved in 1 mL of water to prepare the dye solution. Then 20 μL of the
2.1. Materials dye solution was added into 20 mL of zein solution and mixed for 15
min at 150 rpm at 25 °C in the dark to ensure that the solution was
Zein (Mw 2.5 × 104 Da; protein mass fraction was 98 %) was pro- homogeneous, and also to give time for the dye to bind to the protein.
vided by Beijing J & K Technology Co., Ltd. (Beijing, China). Konjac The dyed zein solution was mixed with the KGM solution to prepare the
glucomannan (KGM) was obtained from Li Cheng Biological dyed blend film-forming solution. The upper layer of the blend solution
Technology Co., Ltd. (Hubei, China). The KGM sample was character- at different drying times was taken out and placed on the microscope
ized by a Mw of 1.012 × 106 Da, a molar ratio of mannose:glucose of slide for testing. The samples were excited by a red laser beam at 638
1.6:1, and a degree of acetylation of 1.85 %. Ethanol (AR, purity ≧99.5 nm.
%) and glycerol (AR, purity ≧99 %) were purchased from Sinopharm
(Chemical Reagent Co., Ltd., Shanghai, China). Rhodamine B (AR, 2.6. Differential scanning calorimetry (DSC)
purity ≧99 %) was purchased from Aladdin Bio-Chem Technology Co.,
Ltd. (Shanghai, China). The thermal properties of the samples were analyzed using a DSC
device (Mettler Toledo, Zurich, Switzerland) referring to the method of
Liu et al. (2020). The upper layer of the sample at different drying times
2.2. Samples preparation for the film-forming process was taken out, frozen instantaneously with liquid nitrogen, vacuum
freeze-dried, and then placed on a Mettler aluminum pan for testing.
Zein (0.1 g) was dissolved in 20 mL of 80 % (v/v) ethanol/water The test conditions were designed under N2 atmosphere, with an empty
solvent under stirring for 15 min at 25 °C. KGM (0.9 g) and glycerol (15 pan as a reference. The test temperature was from 20 to 250 °C, and the
% based on the total amount of KGM and zein, w/w) were completely heating rate was 10 °C min−1.
dissolved in 100 mL water with a continuous stirring electric mixer
(OS20-Pro SCILOGEX Co., Ltd., USA) at 600 rpm for 1.0 h at 60 °C. 2.7. Rheological measurements of film-forming solutions
Afterward, the zein solution was slowly dropped into the KGM solution
to prepare KGM/zein blend film-forming solution at 60 °C for 30 min at Dynamical rheological properties of the film-forming solutions dried
stirring speed 1000 rpm. Then, the blend film-forming solution was at different drying times were measured with a rheometer (RS6000,
poured into the glass plates (14.0 cm × 14.0 cm × 1.5 cm), and dried Hake Ltd., Germany) equipped with a parallel plate (35 mm diameter).
in an oven (DNG-9031A, Jing Hong Co., Ltd., Shanghai, China) at 60 °C. The gap was set as 1.0 mm and the temperature was set at 60 °C. For
During the drying process, the samples were taken out at 0 h, 2 h, 4 h, 6 each measurement, the upper layer of the sample was taken out and
h, 8 h, and 12 h (completely dried to films), respectively. According to placed on the parallel plate instrument. A thin layer of silicon oil was
drying time, the samples were coded as 0 h KZ, 2 h KZ, 4 h KZ, 6 h KZ, 8 placed around the sample to prevent the solvent evaporation during the
h KZ and KZ film, respectively. As the reference sample, pure KGM film measurement. The values of the strain amplitude were checked to en-
was prepared by drying 100 mL KGM solution (1.0 %, w/v) on glass sure that all measurements were set as 1.0 %, which was within the
plates (14.0 cm × 14.0 cm × 1.5 cm) and the amount of glycerol added linear viscoelastic zone. The frequency sweeps were performed from 0.1
was 15 % (w/w) based on the total solid content. to 10 Hz. The storage modulus (G') and loss modulus (G") were mea-
sured as a function of angular frequency. Complex viscosity (η*, Pa·s)

2
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

was calculated as follows Eq. (1). endpoint of drying was 12 h, which was shorter than that of pure KGM
film-forming solution. This phenomenon was explained by that the in-
(G′)2 + (G″)2 teraction between zein and water molecules in the blend solution was
η* =
ω (1) very weak due to the hydrophobic nature of zein, and the interaction
Where ω is the angular frequency. between zein and KGM (Wang et al., 2017) may weaken the interaction
The relationship between η*and ω was analyzed using a power-law between KGM and water. In addition, the small amount of ethanol in
model Eq. (2). the blend solution evaporated easier than water.

η* = Kωn (2) 3.2. Structure and morphology analysis


Where K is the consistency index (Pa s ), ω is the angular frequency and
n
The SEM micrographs of the solution surface and cross-section
n is the power-law index.
during the formation of KGM film are shown in Fig. 2. KGM exhibited a
random coil conformation with molecular entanglement in water (Jian,
2.8. Statistical analysis
Siu, & Wu, 2015; Ni et al., 2018). At the beginning of drying (0 h), KGM
chains were fully expanded and entangled in water, and the entangled
All samples were prepared and tested at least in triplicate. Statistical
molecular chains were overlapped in layers. With the increase of KGM
analysis was conducted using Origin 2017 (originLab corporation,
concentration (4 h, 8 h), the density of the molecular chain increased,
Northampton MA). Analysis of variance was performed using ANOVA
and molecular chains crossed and intertwined with each other, forming
procedures of the IBM SPSS software (version 20.0, IBM Inc., USA,
more junction zones, and ultimately leading to a strong entanglement
2012) by Tukey’s multiple range test at significance level p < 0.05.
network at higher concentrations. At the end of drying (KGM film),
KGM chains developed into a continuous network structure, and the
3. Results and discussion movement of the chains was limited, forming a dense solid film. It was
observed from the cross-section that the closer to the solution surface,
3.1. Drying curves the denser the network, the smaller the pore and the thicker the pore
wall. Therefore in the drying mode of heat conduction, the closer to the
Plots of the solvent evaporation ratio versus drying time are pre- solution surface, the faster the water evaporates.
sented in Fig. 1. Typically, the drying of high moisture materials pre- The microstructure change during the formation of KGM/zein blend
sents a constant drying rate period and one or two falling rate periods film (Fig. 3) indicated a significant difference from KGM film. Zein in
(de Moraes, Scheibe, Augusto, Carciofi, & Laurindo, 2015). The drying ethanol/water solution (80 %) formed spherical particles due to the
curves showed two stages: a first stage with linear solvent loss variation aggregation of molecules (Guo, Liu, An, Li, & Hu, 2005; Kim & Xu,
from 0 to 89 % at a constant rate and a second stage in which the drying 2008). The KGM/zein blend solution formed a stable homogeneous
rate decreased. This should be due to that the gradual solidification of system due to the balance between KGM, zein and solvents in the
the surface of the film-forming solution prevented evaporation of the system (Ni et al., 2018). Zein particles were filled in the KGM molecular
solvent in the second stage. Similar tendencies have been reported entanglement network during drying, indicating the existence of in-
during the drying of soy protein films (Ortiz, de Moraes, Vicente, termolecular interactions between zein and KGM. This was in agree-
Laurindo, & Mauri, 2017) and starch-fiber films (de Moraes et al., ment with previous FTIR spectra result supporting that hydrogen bond
2015). Using conductive drying, Karapantsios (2006) found that the interactions and Maillard reaction occurred between zein and KGM
falling rate periods of the drying curves of starch films started at below molecules (Wang et al., 2017). It was found that KGM chains in the
50 % (w/w) water content for films less than 1 mm, which was inter- blend solution were thicker than that in the pure KGM solution (0 h KZ
preted by that the evaporation loss was prevented when the starch/ in Fig. 3), and a loose network with a lot of holes was formed. This may
water viscous paste began to transform into a solid-like gel. Throughout be due to that the hydrophobicity of zein promoted the approach and
the drying period, the drying rate of KGM/zein blend film-forming so- overlap of hydrophilic KGM molecules, leading to the aggregation of
lution was greatly higher than that of pure KGM film-forming solution. KGM chains into thicker chains, and increasing of the distance between
The time for KGM/zein blend film-forming solution to reach the thicker molecular chains, which was conducive to the formation of the
macroporous network. With drying proceeded (4 h, 8 h), the network in
the blend solution was gradually strengthened and became dense. The
network was irregular and had larger pores compared with that in pure
KGM solution. The surface of the final blend film was rough and the
cross-section was not compact compared with pure KGM film. The
molecular network structure of the blend film was different from that of
pure KGM film, which would affect their physical and chemical prop-
erties. Previous reports have also confirmed their significant differences
in performance (Li et al., 2019; Wang et al., 2017).
Further CLSM microstructure observation (Fig. 4) was performed to
visualize zein association states during the drying process of KGM/zein
blend film-forming solution. The zein particles grew to larger sizes and
the particle density increased significantly with the extension of drying
time. The reason was that solvent evaporation led to an increase in zein
concentration, resulting in easy aggregation of zein particles. Zein
particles in all samples were homogeneously distributed in KGM con-
tinuous matrix without phase separation. The high viscosity formed by
KGM water phase could help to prevent large-scale agglomeration of
zein particles. The rough surface of the KGM/zein blend film (Fig. 3)
can be related to zein aggregation. KGM/zein blend films showed better
Fig. 1. Drying curves of KGM film-forming solution and KGM/zein blend film- moisture resistance compared to pure KGM film (Wang et al., 2017),
forming solution at 60 °C. which was ascribed to that the distribution of zein particles created

3
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

Fig. 2. SEM micrographs of surface and cross-section of KGM samples at different drying times.

impermeable obstacles to permeating water vapor molecules and ex- film exhibited endothermic peaks at 75.4 °C and 92.7 °C respectively,
tended the water diffusion path. The control of the drying process will corresponding to the respective glass transition temperature (Tg),
also affect the degree of zein aggregation. Li et al. (2019) reported that consistent with reported results (Li et al., 2019). The higher Tg of the
different drying temperatures could affect the size of zein aggregation blend film in comparison to that of KGM film indicated better thermal
particles in the film, thus changing the mechanical and barrier prop- stability, which was caused by the interaction and the good compat-
erties of the film. ibility between KGM and zein molecules(Wang et al., 2017). As the
drying progressed, the Tg of 0 h KGM, 4 h KGM, 8 h KGM were 62.1,
3.3. Thermal properties 65.9 and 71.4 °C, respectively; 0 h KZ, 4 h KZ, 8 h KZ showed Tg at
73.5, 81.1 and 85.0 °C, respectively. The Tg of KGM samples increased
The DSC heating curves of the samples at different drying times are with drying, and the Tg of KZ samples also showed the same trend.
presented in Fig. 5. It was found that KGM film and KGM/zein blend Moreover, under the same drying time, the Tg of the KZ sample was

4
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

Fig. 3. SEM micrographs of surface and cross-section of KGM/zein blend samples at different drying times.

5
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

Fig. 4. CLSM images of KGM/zein blend samples at different drying times. The red area represents zein particles and the black area represents KGM components.

significantly higher than that of the KGM sample. Changes in Tg in- begins to form. Therefore, the rheological properties at five drying
dicated that the molecular interaction in film-forming solutions was times (0 h, 2 h, 4 h, 6 h, 8 h) were investigated.
strengthened during drying, and the addition of zein improved the For KGM film-forming solution dried at 60 °C, both G' and G" of the
thermal stability of film-forming solutions. samples at different drying times increased with frequency and showed
a crossover (Fig. 6). G" values were higher than G' at lower frequencies
but the G' values became larger than G" at higher frequencies. This is a
3.4. Rheological properties
typical concentrated solution behavior (Nishinari, 1997). G' is related to
the elastic energy stored in viscoelastic material during deformation,
3.4.1. Dynamic viscoelastic properties
reflecting the elastic solid-like characteristics of polymers; while G" is
In order to obtain the information of molecular entanglement and
related to the dissipated energy as heat during deformation, reflecting
molecular network formation during drying, the dynamic viscoelastic
the viscosity of the material. KGM samples showed more liquid-like
properties of the film-forming solutions were evaluated. Frequency
properties at lower frequencies (G" > G') and more elastic solid-like
dependence of storage modulus G' and loss modulus G" was observed at
behavior at higher frequencies (G' > G"). This could be explained by
different drying times (Figs. 6 & 7 ). Above 8 h drying time, it is difficult
that the molecular chains had more time to relax and disentangle
to test and explain the rheological properties because the solid film

Fig. 5. DSC curves of KGM samples and KGM/zein blend samples at different drying times.

6
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

Fig. 6. Storage modulus (G′) and loss modulus (G″) for KGM samples at different drying times.

during a long period oscillation at low frequencies, and the solution continuously strengthened with drying.
behaved as a viscous liquid. However, the molecular chains could not Fig. 7 shows the change in G' and G" as a function of angular fre-
disentangle during a short period oscillation at high frequencies and quency for KGM/zein blend film-forming solution. At the initial stage of
formed a temporary three-dimensional network structure, and thus the drying (0 h), the G' and G" values of the blend solution increased with
solution exhibited an elastic solid-like state (Nishinari, 1997; frequency. Compared with 0 h KGM solution, G' and G" of the blend
Yoshimura, Takaya, & Nishinari, 1998). As drying proceeded, the KGM solution were lower and their crossover point shifted to high frequency.
concentration increased due to the evaporation of water, and both G' This was due to the low content of KGM in the blend solution, which
and G" of the samples were increased. This may be explained as higher reduced the entanglement of molecular chains. Entanglement effects
KGM content resulting in an increased chance for macromolecular play an important role in the rheological behavior of flexible linear
chain entanglements, as reflected by the increasing modulus (Wang polymers, but they do not exist in colloidal suspensions (Tanaka,
et al., 2014). The crossover point moved toward low frequencies. At 0 Nishikawa, & Koyama, 2005). As drying proceeded (2 h–8 h), the blend
h, 2 h, 4 h, 6 h and 8 h, the crossover frequency of samples occurred at solution showed elastic solid-like behavior (G' > G") and no crossover
1.0 Hz, 0.31 Hz, 0.23 Hz, 0.17 Hz, and 0.13 Hz, respectively. The lower was observed over the frequency range. It indicated that the average
crossover frequency indicated the stronger effects of entanglement and chain relaxation time of the blend solution was greater than that of the
hydrogen bonding (Luo, He, & Lin, 2013). It was consistent with the 0 h blend solution. Furthermore, with the increase of drying time, G'
result of the SEM (Fig. 2) that the entanglement network was and G" gradually became parallel, and both moduli showed slight

7
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

Fig. 7. Storage modulus (G′) and loss modulus (G") for KGM/zein blend samples at different drying times.

frequency dependence. This might be attributed to the gradual forma- increasing frequency. In order to quantitatively evaluate the viscoe-
tion of a weak gel (Nishinari, 1997, 2009). These observations re- lastic nature of samples at different drying times, a power-law model
presented that zein particles had a great effect on the entanglement of had been applied to the complex viscosity and frequency data (Tables 1
KGM molecules, and the blend solution was easier to form a weak gel and 2), which showed high R2 (all > 0.96). The power-law index (n)
than KGM solution. The overlap and entanglement of KGM chains decreased, indicating a more pseudoplastic behavior in samples and the
caused by the addition of zein promoted the transition from solution to weak dependence on frequency (Kim & Yoo, 2009). The n value of 2 h
weak gel. Luo et al. (2013) reported that the sodium hydroxide solution KZ (3.61 % solid content) was almost equal as that of 8 h KGM (6.98 %
could restrain the expansion of KGM chains and increase the rate of solid content), which was obviously smaller than that of 6 h KGM (4.69
gelation. % solid content). This may be attributed to the enhanced molecular
entanglement force and improved molecular network with the addition
of zein into KGM solution. Our previous studies (Wang et al., 2017)
3.4.2. Complex viscosity showed that KGM/zein blend films had good mechanical properties,
The complex viscosity (η*) for film-forming solutions at different which was related to the enhanced network structure shown in this
drying times is plotted against the frequency in Fig. 8. As drying pro- study.
gressed, the η* values at a given frequency were observed to increase
significantly. Moreover, KGM samples and KZ samples revealed shear-
thinning behavior as their η* values decreased gradually with

8
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

Fig. 8. Complex viscosity (η*) of KGM samples (a) and KGM/zein blend samples (b) at different drying times.

Table 1 gel. The η* data for KGM solution and KGM/zein blend solution were
Power-law parameters (K, n) of KGM samples. fitted by the power-law model. As a result, the n value of the blend
Samples Power-law parameters solution became smaller than that of KGM solution with the drying
process, which could be attributed to the enhancement of molecular
K (Pa sn) n R2 entanglement force.
0 h KGM 12.84 ± 1.42a −0.44 ± 0.014a 0.98
2 h KGM 50.18 ± 1.49b −0.48 ± 0.007b 0.98
CRediT authorship contribution statement
4 h KGM 112.47 ± 14.45c −0.53 ± 0.009c 0.98
6 h KGM 153.64 ± 8.80d −0.55 ± 0.012d 0.98 Chong Li: Methodology, Data curation, Software, Writing - original
8 h KGM 225.45 ± 31.74e −0.59 ± 0.008e 0.97 draft. Fei Xiang: Validation, Formal analysis, Visualization. Kao Wu:
Writing - review & editing, Supervision, Data curation. Fatang Jiang:
Different superscript letters (a–e) within the same column indicate significant
Writing - review & editing. Xuewen Ni: Supervision, Funding acquisi-
differences between formulations (p < 0.05).
tion, Conceptualization, Methodology, Visualization, Writing - review &
editing.
Table 2
Power-law parameters (K, n) of KGM/zein blend samples.
Declaration of Competing Interest
Samples Power-law parameters

K (Pa sn) n R2 The authors report no declarations of interest.

0 h KZ 3.138 ± 0.80a
−0.39 ± 0.02a
0.98 Acknowledgements
2 h KZ 36.63 ± 1.79b −0.60 ± 0.007b 0.99
4 h KZ 60.99 ± 2.86c −0.64 ± 0.018c 0.99
6 h KZ 109.58 ± 7.30d −0.73 ± 0.03d 0.99 This work was financially supported by National Natural Science
8 h KZ 353.35 ± 72.60e −0.62 ± 0.25bc 0.97 Foundation of China (Grant No. 31901655) and the technical support
program of Hubei University of Technology (Grant No. CPYF2018004).
Different superscript letters (a–e) within the same column indicate significant
differences between formulations (p < 0.05).
References

4. Conclusions
Bourlieu, C., Guillard, V., Vallès-Pamiès, B., Guilbert, S., & Gontard, N. (2009). Edible
moisture barriers: How to assess of their potential and limits in food products shelf-
During film formation, the drying process of KGM solution and life extension? Food Science and Nutrition, 49, 474–499.
Bourtoom, T., & Chinnan, M. S. (2008). Preparation and properties of rice starch-chitosan
KGM/zein blend solution at 60 °C was divided into two stages: a con-
blend biodegradable film. LWT-Food Science and Technology, 41, 1633–1641.
stant drying rate period and a falling rate period. The gradual solidifi- de Moraes, J. O., Scheibe, A. S., Augusto, B., Carciofi, M., & Laurindo, J. B. (2015).
cation of the solution surface reduced the rate of solvent evaporation Conductive drying of starch-fiber films prepared by tape casting: Drying rates and
with drying. The addition of zein affected the entanglement network of film properties. LWT-Food Science and Technology, 64, 356–366.
Dehghani, S., Hosseini, S. V., & Regenstein, J. M. (2018). Edible films and coatings in
KGM in the blend solution and increased the entanglement force. seafood preservation: A review. Food Chemistry, 240, 505–513.
Compared with KGM solution, KGM chains in the blend solution were Denavi, G. A., Pérez-Mateos, M., Añón, M. C., Montero, P., Mauri, A. N., & Gómez-Guillén,
thicker and formed a molecular network with larger pores. Zein parti- M. C. (2009). Structural and functional properties of soy protein isolate and cod
gelatin blend films. Food Hydrocolloids, 23, 2094–2101.
cles grew to larger sizes due to aggregation but were homogeneously Ghoshal, S., Mattea, C., Denner, P., & Stapf, S. (2010). Heterogeneities in gelatin film
distributed in KGM continuous matrix without phase separation formation using single-sided NMR. The Journal of Physical Chemistry B, 114(49),
throughout the entire drying process. The blend solution showed better 16356–16363.
Guo, Y., Liu, Z., An, H., Li, M., & Hu, J. (2005). Nano-structure and properties of maize
thermal stability than KGM solution. As the drying proceeded (up to 8 zein studied by atomic force microscopy. Journal of Cereal Science, 41, 277–281.
h), G' and G" of KGM solution increased with frequency and showed a Hassan, B., Chatha, S. A. S., Hussain, A. I., Zia, K. M., & Akhtar, N. (2018). Recent ad-
crossover. The crossover point moved toward low frequencies with vances on polysaccharides, lipids and protein based edible films and coatings: A re-
view. International Journal of Biological Macromolecules, 109, 1095–1107.
drying, indicating the stronger effects of entanglement and hydrogen
Jia, D., Fang, Y., & Yao, K. (2009). Water vapor barrier and mechanical properties of
bonding. The blend solution showed elastic solid-like behavior konjac glucomannan-chitosan-soy protein isolate edible films. Food and Bioproducts
(G' > G") after drying for 2 h. G' and G" gradually became parallel with Processing, 87(1), 7–10.
slight frequency dependence, due to the gradual formation of a weak Jian, W., Siu, K., & Wu, J. (2015). Effects of pH and temperature on colloidal properties
and molecular characteristics of konjac glucomannan. Carbohydrate Polymers, 134,

9
C. Li, et al. Carbohydrate Polymers 250 (2020) 116840

285–292. preservation. Carbohydrate Polymers, 208, 276–284.


Karapantsios, T. D. (2006). Conductive drying kinetics of pregelatinized starch thin films. Ortiz, C. M., de Moraes, J. O., Vicente, A. R., Laurindo, J. B., & Mauri, A. N. (2017). Scale-
Journal of Food Engineering, 76, 477–489. up of the production of soy (Glycine max L.) protein films using tape casting:
Kato, K., & Matsuda, K. (1969). Studies on the chemical structure of konjac mannan. Part Formulation of film-forming suspension and drying conditions. Food Hydrocolloids,
I. Isolation and characterization of oligosaccharides from the partial acid hydrolyzate 66, 110–117.
of the mannan. Agricultural and Biological Chemistry, 33, 1446–1453. Panchapakesan, C., Sozer, N., Dogan, H., Huang, Q., & Kokini, J. L. (2012). Effect of
Kim, S., & Xu, J. (2008). Aggregate formation of zein and its structural inversion in different fractions of zein on the mechanical and phase properties of zein films at
aqueous ethanol. Journal of Cereal Science, 47, 1–5. nano-scale. Journal of Cereal Science, 55, 174–182.
Kim, W., & Yoo, B. (2009). Rheological behaviour of acorn starch dispersions: Effects of Rhim, J., & Wang, L. (2013). Mechanical and water barrier properties of agar/κ-carra-
concentration and temperature. International Journal of Food Science and Technology, geenan/konjac glucomannan ternary blend biohydrogel films. Carbohydrate Polymers,
44, 503–509. 96, 71–81.
Kohyama, K., Iida, H., & Nishinari, K. (1993). A mixed system composed of different Scramin, J. A., de Britto, D., Forato, L. A., Bernardes-Filho, R., Colnago, L. A., & Assis, O.
molecular weights konjac glucomannan and kappa carrageenan: Large deformation B. G. (2011). Characterisation of zein-oleic acid films and applications in fruit
and dynamic viscoelastic study. Food Hydrocolloids, 7, 213–226. coating. International Journal of Food Science and Technology, 46, 2145–2152.
Li, C., Wu, K., Su, Y., Riffat, S. B., Ni, X., & Jiang, F. (2019). Effect of drying temperature Smith, F., & Srivastava, H. C. (1959). Constitutional studies on the glucomannan of konjak
on structural and thermomechanical properties of konjac glucomannan-zein blend flour. Journal of American Chemical Society, 81, 1715–1718.
films. International Journal of Biological Macromolecules, 138, 135–143. Subirade, M., Kelly, I., Guéguen, J., & Pézolet, M. (1998). Molecular basis of film for-
Li, X., Jiang, F., Ni, X., Yan, W., Fang, Y., Corke, H., et al. (2015). Preparation and mation from a soybean protein: Comparison between the conformation of glycinin in
characterization of konjac glucomannan and ethyl cellulose blend films. Food aqueous solution and in films. International Journal of Biological Macromolecules, 23,
Hydrocolloids, 44, 229–236. 241–249.
Liu, Z., & Han, J. H. (2005). Film-forming characteristics of starches. Journal of Food Tanaka, H., Nishikawa, Y., & Koyama, T. (2005). Network-forming phase separation of
Science, 70, E31–E36. colloidal suspensions. Journal of Physics: Condensed Matter, 17, 143–153.
Liu, Z., Liu, C., Sun, X., Zhang, S., Yuan, Y., Wang, D., et al. (2020). Fabrication and Tong, Q., Xiao, Q., & Lim, L. (2008). Preparation and properties of pullulan-alginate-
characterization of cold-gelation whey protein-chitosan complex hydrogels for the carboxymethylcellulose blend films. Food Research International, 41, 1007–1014.
controlled release of curcumin. Food Hydrocolloids, 103, Article 105619. Wang, Y., & Padua, G. W. (2010). Formation of zein microphases in ethanol-water.
Luo, X., He, P., & Lin, X. (2013). The mechanism of sodium hydroxide solution promoting Langmuir, 26(15), 12897–12901.
the gelation of konjac glucomannan (KGM). Food Hydrocolloids, 30, 92–99. Wang, K., Wu, K., Xiao, M., Kuang, Y., Corke, H., Ni, X., et al. (2017). Structural char-
Łupina, K., Kowalczyk, D., Zięba, E., Kazimierczak, W., Mężyńska, M., Basiura-Cembala, acterization and properties of konjac glucomannan and zein blend films. International
M., et al. (2019). Edible films made from blends of gelatin and polysaccharide-based Journal of Biological Macromolecules, 105, 1096–1104.
emulsifiers-a comparative study. Food Hydrocolloids, 96, 555–567. Wang, S., Zhan, Y., Wu, X., Ye, T., Li, Y., Wang, L., et al. (2014). Dissolution and rheo-
Mohajer, S., Rezaei, M., & Hosseini, S. F. (2017). Physico-chemical and microstructural logical behavior of deacetylated konjac glucomannan in urea aqueous solution.
properties of fish gelatin/agar bio-based blend films. Carbohydrate Polymers, 157, Carbohydrate Polymers, 101, 499–504.
784–793. Wei, X., Pang, J., Zhang, C., Yu, C., Chen, H., & Xie, B. (2015). Structure and properties of
Ni, X., Wang, K., Wu, K., Corke, H., Nishinari, K., & Jiang, F. (2018). Stability, micro- moisture-resistant konjac glucomannan films coated with shellac/stearic acid
structure and rheological behavior of konjac glucomannan-zein mixed systems. coating. Carbohydrate Polymers, 118, 119–125.
Carbohydrate Polymers, 188, 260–267. Wu, K., Zhu, Q., Qian, H., Xiao, M., Corke, H., Nishinari, K., et al. (2018). Controllable
Nishinari, K. (1997). Rheological and DSC study of sol-gel transition in aqueous disper- hydrophilicity-hydrophobicity and related properties of konjac glucomannan and
sions of industrially important polymers and colloids. Colloid and Polymer Science, ethyl cellulose composite films. Food Hydrocolloids, 79, 301–309.
275(12), 1093–1107. Xiao, M., Wan, L., Corke, H., Yan, W., Ni, X., Fang, Y., et al. (2016). Characterization of
Nishinari, K. (2009). Some thoughts on the definition of a gel. Progress in Colloid & konjac glucomannan-ethyl cellulose film formation via microscopy. International
Polymer Science, 136, 87–94. Journal of Biological Macromolecules, 85, 434–441.
Niu, B., Shao, P., Chen, H., & Sun, P. (2019). Structural and physiochemical character- Yoshimura, M., Takaya, T., & Nishinari, K. (1998). Rheological studies on mixtures of
ization of novel hydrophobic packaging films based on pullulan derivatives for fruits corn starch and konjac-glucomannan. Carbohydrate Polymers, 35, 71–79.

10

You might also like