You are on page 1of 8

Food Hydrocolloids 48 (2015) 189e196

Contents lists available at ScienceDirect

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

The effect of starch concentration on the gelatinization and


liquefaction of corn starch
Zhaofeng Li a, b, c, Wenjing Liu b, Zhengbiao Gu a, b, c, *, Caiming Li b, Yan Hong a, b, c,
Li Cheng a, b
a
State Key Laboratory of Food Science and Technology, Jiangnan University, Wuxi 214122, China
b
School of Food Science and Technology, Jiangnan University, Wuxi 214122, China
c
Synergetic Innovation Center of Food Safety and Nutrition, Jiangnan University, Wuxi, Jiangsu 214122, China

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

Article history: Starch liquefaction is a key step in the industrial production of syrups, including those used in microbial
Received 12 July 2014 fermentation processes. As part of an effort to improve the efficiency of this process, the effects of starch
Received in revised form concentration on the gelatinization and liquefaction of corn starch were investigated. The results
2 February 2015
demonstrated that high starch concentrations in the slurry make it difficult to gelatinize the starch.
Accepted 23 February 2015
Available online 3 March 2015
Elevated starch concentrations inhibited swelling and disruption of starch granules and resulted in the
retention of starch crystallinity after heat treatment. It became very difficult to destroy the granular and
crystalline structure of the starch by heat treatment when the starch concentration exceeded 45%.
Keywords:
Corn starch
Furthermore, elevated starch concentration had negative effects on the liquefaction of corn starch
Starch concentration catalyzed by a thermostable a-amylase. The dextrose equivalent (DE) values of starch hydrolysates
Gelatinization decreased as the starch concentration increased. Nevertheless, increasing the starch concentration from
Liquefaction 30 to 45% caused only a slight reduction in the DE value of the hydrolysate, indicating that starch
DE value liquefaction at 45% starch could be comparable with that at 30%. Starch concentrations exceeding 45%
severely retarded starch liquefaction. The relatively low degree of starch liquefaction seen at high starch
concentrations, especially above 45%, was related to incomplete gelatinization of the starch, and low
mobility of the water molecules and polymeric chains in the reaction system. These results provide
additional insight into the feasibility of high-temperature liquefaction at high corn starch concentrations.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction generally involves gelatinization of the starch either before or


during the addition of the a-amylase, while gelatinization is
The enzymatic hydrolysis of starch is one of the most important accompanied by swelling and disruption of the starch granules and
enzymatic reactions (Presecki, Blazevic, & Vasic-Racki, 2013). melting of their crystalline structure (Bogracheva, Morris, Ring, &
Liquefaction is a key step in the starch hydrolysis process used to Hedley, 1998; Mandala & Bayas, 2004). Since native starch is hy-
produce syrups, including those used in the industrial production drolyzed more slowly than gelatinized starch whose crystallinity
of microbial fermentation products such as alcohol, organic acids, has been largely destroyed (Blazek & Gilbert, 2010), the degree of
amino acids, and antibiotics. Liquefaction is commonly achieved starch liquefaction is closely related to starch gelatinization. The
through the dispersion of insoluble starch granules in an aqueous other factors affecting starch liquefaction include the source and
solution, followed by partial hydrolysis at a relatively high tem- concentration of the starch; the source and activity of the a-
perature using thermostable a-amylases, which are endogluca- amylase; the concentration of calcium ions, which are related to the
nases that catalyze the hydrolysis of internal a-1,4-glycosidic activation and stability of a-amylase; and other reaction conditions,
linkages in starch (Presecki et al., 2013). Starch liquefaction such as temperature and pH (Ariff, Asbi, Azudin, & Kennedy, 1997;
Baruque, Baruque, & Sant'Anna, 2000; Khedher, Bressollier, Urdaci,
Limam, & Marzouki, 2008; Presecki et al., 2013). Even though
substantial efforts have been devoted to the optimization of starch
* Corresponding author. School of Food Science and Technology, Jiangnan Uni-
versity, Wuxi, Jiangsu 214122, China. Tel./fax: þ86 510 85329237. liquefaction, such as use of high performance a-amylase, the
E-mail address: zhengbiaogu@jiangnan.edu.cn (Z. Gu). addition of calcium, optimization of process parameters, and starch

http://dx.doi.org/10.1016/j.foodhyd.2015.02.030
0268-005X/© 2015 Elsevier Ltd. All rights reserved.
190 Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196

pretreatment (Nikoli c, Mojovi c, Rakin, Pejin, & Savi c, 2008; starch, was calculated according to the method shown by Tester
Presecki et al., 2013; Richardson et al., 2002), there is still a need and Morrison (Tester & Morrison, 1990).
to improve this step, since the processes of syrup production and
fermentation can be made more efficient. 2.3. Scanning electron microscopy analysis
Native starch slurries derived from conventional corn wet
milling contain approximately 40% dry solids. They are generally Slurries containing 10e60% (w/w, dry basis) starch were pre-
diluted to 25e35% dry solids before heating above the liquefaction pared for scanning electron microscopy (SEM) by incubation at
temperature (Bhargava, Frisner, & Johal, 2008; Lee, Shetty, & 90  C for 1 h, cooling, and then freeze-drying. The freeze-dried
Strohm, 2013). However, increased amounts of energy are samples were coated with goldepalladium by using a sputter
required to heat starch slurries of relatively low solid content and to coater (Denton Vacuum, LLC, Moorestown, NJ), and then viewed at
evaporate the excess water when concentrating the mash after 2000  resolution with a scanning electron microscope (S-3800N,
subsequent saccharification. In addition, efficient production of Hitachi Science Systems, Ltd., Japan) operating at an accelerating
fermentation products may require the culture medium to be voltage of 20 kV.
supplemented with an appropriate carbon source, which is often a
syrup containing a high concentration of fermentable saccharides 2.4. Light microscopy analysis
(above 40%). Furthermore, increasing the initial concentration of
starch slurry during liquefaction can yield higher productivity, and Slurries containing 10e60% (w/w, dry basis) starch were incu-
higher enzyme stability (Baks, Kappen, Janssen, & Boom, 2008; de bated at 90  C for 1 h. After cooling, the diluted samples were
Cordt, Hendrickx, Maesmans, & Tobback, 1994). Thus, it is desirable promptly viewed using an optical microscope (Model BX51,
to use more concentrated starch slurries. Olympus Co., Japan) fitted with a polarized light filter and a digital
Although the effects of the initial concentration on starch camera. Observations were conducted under cross-polarized light
liquefaction have been mentioned in previous reports (Baks et al., using a 40 objective.
2008; Konsula & Liakopoulou-Kyriakides, 2004; Yankov, Dobreva,
Beschkov, & Emanuilova, 1986), the degree of liquefaction has not 2.5. X-ray diffraction analysis
been measured over a wide range of starch concentrations. In the
present study, the effects of starch concentration on the gelatini- Slurries containing 10e60% (w/w, dry basis) starch were incu-
zation of corn starch were examined. Then, the high temperature bated at 90  C for 1 h. After cooling, the samples were freeze-dried
liquefaction of corn starch slurries was investigated over a wide and pulverized to a powder. X-ray diffractograms of these starch
concentration range (10e60%, w/w, dry basis), with the expectation powders were obtained with a copper anode X-ray tube using a
that increasing the initial concentration of the starch slurry would Rigaku D-Max- 2200 X-ray diffractometer (Rigaku Denki Co. Tokyo,
enhance hydrolysate productivity. The relationship between starch Japan). X-ray diffraction patterns were acquired at room tempera-
gelatinization and liquefaction was also analyzed. The results pro- ture over the 2q range of 4e36 with a step size of 0.02 . Crystal-
vide additional insight into the feasibility for enzymatic liquefac- linity was quantified by integrating the area under the fitted
tion of corn starch at high concentrations. crystalline peaks using MDI Jade 5.0 software.

2. Materials and methods 2.6. Differential scanning calorimetry (DSC) analysis

2.1. Materials The thermal properties of slurries containing 10e60% (w/w, dry
basis) starch were determined from DSC curves obtained using a
Corn starch was obtained from Zhucheng Xingmao Corn Pyris 1-DSC (PerkinElmer Corp., Norwalk, CT, USA). These samples
Developing Co., Ltd (Shandong, China). The thermostable a- (approximately 6 mg) were accurately weighed and sealed in
amylase from Bacillus subtilis was obtained from Genencor Inter- aluminum DSC pans, and then equilibrated at 4  C for 24 h. This
national (18,100 U/mL; Palo Alto, CA, USA). Blue dextran, which was equilibration resulted in an even distribution of water without
prepared from dextran with an average molecular weight of microbial fermentation, making the starch slurries more uniform.
approximate 2.0  106 g/mol, was purchased from Sigma (St. Louis, Scans were performed from 30  C to 90  C at a constant rate of
MO, USA). Soluble starch, dinitrosalicylic acid (DNS), hydrochloric 10  C/min. Enthalpy changes (DH) were evaluated based on the area
acid (HCl), sodium hydroxide, acetic acid, sodium acetate, iodine of the main endothermic peak and expressed in terms of J/g of dry
(I2), and potassium iodide (KI) were purchased from Sinopharm starch. A sealed, empty crucible was used as a reference. All mea-
Chemical Reagent Co. (Shanghai, China). surements were performed in triplicate.

2.2. Determination of the swelling factor 2.7. Determination of water molecule mobility

The swelling factor of the starch was measured using the blue The measurement of spinespin relaxation time (T2) was per-
dextran dye exclusion method of Tester and Morrison (Tester & formed using an MQC-23 nuclear magnetic resonance (NMR)
Morrison, 1990), with some modifications. Slurries containing spectrometer (Oxford Instruments LTD., England) operating at
10e60% (w/w, dry basis) starch were incubated at 90  C for 1 h. 25  C and a resonance frequency of 20.9 MHz. Slurries containing
After cooling, the samples originally containing 100 mg of dry 10e60% (w/w, dry basis) starch were incubated at 90  C for 1 h.
starch were weighed into 10-ml screw-cap tubes, distilled water After cooling to 25  C, the samples (3 g) were poured into a 10 mm
was added to a total of 5 mL, 0.5 mL of blue dextran solution (5 mg/ diameter NMR tube that was then sealed and immediately trans-
mL) was added, and then the contents were mixed by gently ferred to the NMR probe. The T2 values of water protons were
inverting the closed tubes several times. After centrifugation at measured using the Carr-Purcell-Meiboom-Gill (CPMG) pulse
5000  g for 5 min, the absorbance of the supernatant was sequence. The pulse gap between the 90 and 180 pulses was
measured at 620 nm. The absorbance of reference tubes that con- 200 ms. Eight scans, each containing 1024 echoes, were averaged to
tained no starch was also measured. The swelling factor, reported obtain the final signal. All measurements were performed in trip-
as a ratio of the volume of swollen granules to the volume of dry licate. Relaxation curves were obtained by fitting the data to the
Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196 191

exponential model described in Eq. (1), where T2 is spinespin Differences resulting in values of p < 0.05 were considered statis-
relaxation time and A0 is the relative measure of the amount of tically significant.
water fractions corresponding to T2.
3. Results and discussion

t
T2
A ¼ A0 e (1) 3.1. Effects of starch concentration on the swelling and disruption of
starch granules
Fitting was accomplished using 1st Opt software, version 1.5
(7D-Soft High Technology Inc., Beijing, China).
3.1.1. Swelling factor analysis
Starch granules are generally insoluble in cold water. When
2.8. Liquefaction process of starch starch is heated in water, the starch granules absorb water and
swell (Ratnayake & Jackson, 2006). The extent of granule swelling
Starch liquefaction was performed in the sample cup of a Bra- can be determined by measuring the swelling factor (Contreras
bender Viscoamylograph (C.W. Brabender Instruments, Inc., South pez, Role
Lo e, & Le Meste, 2004). After slurries containing
Hackensack, NJ). Four hundred grams of slurries containing 10e60% different concentrations of corn starch (no enzyme added) were
(w/w, dry basis) starch were prepared in distilled water, and incubated at 90  C for 1 h, the swelling factors of the starch were
adjusted to pH 6.0. The temperature of the starch slurry was analyzed. As shown in Fig. 1, the swelling factors decreased grad-
increased to 60  C at a rate of 3.0  C/min, and then a thermostable ually with increasing starch concentration, suggesting that the
a-amylase (12 U/g of dry starch) was added to the reaction mixture. starch concentration had an inhibitory effect on granule swelling.
Subsequently, the temperature was increased to 90  C at a rate of When the slurry contained 10% starch, the swelling factor reached
1.0  C/min. Liquefaction was continued for 2 h at a constant approximately 15, which was about 12% lower than that for the
agitation speed of 200 rpm. At different time intervals, 5 mL of slurry containing 2% (v/w, dry basis) starch (Li, Cai, Gu, & Shi, 2014).
starch hydrolysate was withdrawn for analysis. Liquefaction was Increasing the starch concentration from 10% to 45% resulted in a
immediately stopped by chilling to 50  C followed by acidification decrease of approximately 70% in the swelling factor. When the
at pH 3.5 with 0.1 N HCl. slurries reached starch concentrations of 50e60%, the swelling
factors were very low, indicating very little swelling of starch
2.9. a-Amylase activity assay granule during heat treatment.

Soluble starch (0.4 g) was dissolved in 25 mL of 50 mM sodium 3.1.2. SEM analysis


acetate buffer (pH 6.0). The temperature of starch solution was When heated in water, starch undergoes a transition process
adjusted to 70  C in a water bath, 1 mL of appropriately diluted during which the granules break down (Ratnayake & Jackson,
enzyme solution was added, and the mixture was incubated at 2009). Thus, heat treatment of a starch slurry is associated with
70  C for 5 min. Then, 1 mL of reaction mixture was added to a the disruption of granular structure, causing starch molecules to
mixture containing 5 mL of iodine reagent (0.01% I2 in 4% KI) and dissolve in water. After slurries containing different concentrations
0.5 mL of 0.1 M HCl, which stopped the reaction. A complete assay of starch were incubated at 90  C for 1 h, freeze-dried starch
mixture without enzyme was used as a control. The degradation of samples were observed by SEM. As shown in Fig. 2, when the starch
starch by the enzyme was determined by comparing the absor- concentration was 10e20%, heat treatment led to complete granule
bance of the sample or the control at 660 nm with the absorbance disruption. Although increasing the starch concentration to
of a blank prepared by treating 1 mL of water with the same iodine/ 30e45% resulted in a viscous starch paste, almost no intact starch
HCl mixture. One unit of a-amylase activity was defined as the granules appeared in the SEM images, indicating that most of the
amount of enzyme that caused a 10% reduction in the starch granules were damaged by heat treatment to a certain
starcheiodine color under the assay conditions. extent. A few intact starch granules could be observed by SEM
when the starch concentration reached 50e55%, indicating
2.10. Dextrose equivalent (DE) of the starch hydrolysate incomplete granule disruption by heat treatment. When the slurry

The reducing sugar content of the starch hydrolysate was


determined using the DNS method (Bruner, 1964). The degree of
liquefaction is expressed as dextrose equivalent (DE), which is
defined as the total reducing sugars expressed as dextrose and
calculated as a percentage of the dry substance.

2.11. Changes in viscosity during starch liquefaction

During starch liquefaction, a Brabender Viscoamylograph (C.W.


Brabender Instruments, Inc., South Hackensack, NJ) was used to
determine the change in viscosity as a function of time.

2.12. Statistical analysis

Experimental results about swelling factor, T2 value, DE value,


and enthalpy are reported as the average of triplicate measure-
ments, along with their associated standard deviations. Differences
between treatments were analyzed for statistical significance using Fig. 1. The effect of starch concentration on the swelling factor. Each value represents
the StudenteNewmaneKeuls (SNK) procedure, as implemented in the mean of three independent measurements. Bars with different letters on the top
the statistical software package from SPSS Inc. (Chicago, IL, USA). are significantly different (p < 0.05).
192 Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196

Fig. 2. SEM images of samples containing different concentrations of corn starch, after heat treatment at 90  C for 1 h, and then freeze drying. Magnification, 500.

containing 60% starch was analyzed, most of the starch granules starch molecules within the granule to cleavage by amylase,
remained intact. resulting in the generation of amorphous material within the starch
Continued heating generally causes starch granules to swell and granule and an associated increase in granule porosity (Wang &
be disrupted to a greater extent. However, the starch concentration Copeland, 2013). Thus, the inhibited swelling and disruption of
had a substantial effect on the swelling and disruption of starch starch granules, as characteristics of poor starch gelatinization,
granules, indicating varying degrees of starch gelatinization. The would have negative effect on starch liquefaction.
increase in starch concentration, particularly to concentrations
above 45%, resulted in reduced extents of granule swelling and 3.2. Effects of starch concentration on the crystalline structure of
disruption. It has been shown that starch granules are quite resis- starch granules after heat treatment
tant to enzymatic hydrolysis and are predominantly hydrolyzed by
surface abrasion (Asare et al., 2011; Siswoyo & Morita, 2003; 3.2.1. Light microscopy analysis
Srichuwong, Isono, Mishima, & Hisamatsu, 2005). Granule Corn starch molecules are densely packed in starch granules,
swelling and disruption could greatly increase the susceptibility of resulting in a semi-crystalline state that gives rise to birefringence
Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196 193

under cross-polarized light (Chen, Huang, Tang, Chen, & Zhang, birefringence. When slurries containing 30e45% starch were heated,
2011). When starch is heated in water, the absorption of water almost no birefringence was observed in polarizing microscope
into the granules can destabilize their crystalline structure, images, indicating that the crystalline structure of most starch
resulting in the loss of birefringence (Parker & Ring, 2001; granules were destroyed, to some extent, by the heat treatment. As
Ratnayake & Jackson, 2006). Therefore, a loss of birefringence is a slurries containing even higher concentrations of starch were heat-
good indicator of the destruction of the crystalline structure ed, many starch granules retained their birefringence, indicating that
of starch granules. Polarized light micrographs of diluted heat treatment could not completely destroy the crystalline struc-
starch samples, obtained after slurries containing different starch ture of the starch granules present at high starch concentrations.
concentrations were incubated at 90  C for 1 h, are shown in Fig. 3.
The results demonstrated that the loss of birefringence was 3.2.2. X-ray diffraction analysis
controlled by the availability of water. Heat treatment of slurries The crystalline properties of starch granules can be studied
containing 10e20% starch led to the complete loss of starch granule using X-ray diffraction (Eliasson, 2006; Ratnayake & Jackson,

Fig. 3. Photomicrographs of samples containing different concentrations of corn starch, after heat treatment at 90  C for 1 h. Images were obtained at 400 using a polarizing
microscope.
194 Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196

2008). The crystalline structures of starches can be classified into gelatinization (Tester & Sommerville, 2001). Thus, the retention of
three forms (A, B and C) on the basis of their X-ray diffraction starch crystallinity can be evaluated by the residual enthalpy after
spectra (Cheetham & Tao, 1998). The X-ray diffraction patterns of heat treatment (Tester & Sommerville, 2001). It is apparent that the
freeze-dried starch samples obtained after slurries containing slurries containing higher concentrations of starch retained greater
different concentrations of starch were incubated at 90  C for 1 h crystallinity after heat-treatment. When the slurry contained >45%
are shown in Fig. 4. Native corn starch showed the strongest starch, the retention of crystallinity reached >60%, suggesting that
diffraction peaks, at 2q values of 15.3 , 17.1, 18.2 and 23.5 more than half of the crystalline starch could not be destroyed by
(Fig. 4a), indicating that this starch had a characteristic A-type heat treatment.
crystalline form, which is consistent with previous reports (Koo, Starch gelatinization can be reflected in the melting of starch
Lee, & Lee, 2010; Sun, Zhao, Zeng, Li, & Li, 2010; Uthumporn, crystallites (Zobel, Young, & Rocca, 1988). Increasing the starch con-
Zaidul, & Karim, 2010). Compared with those of native starch, the centration, especially to concentrations above 45%, severely reduced
characteristic diffraction peaks of heat-treated starch decreased in the loss of crystallinity, suggesting incomplete gelatinization of the
intensity, or disappeared (Fig. 4bei). During heating in excess starch. Previous reports have also shown that the crystalline lamellae
water, individual crystalline regions within starch molecules can in starch granules were resistant to enzymatic erosion, and that
melt within a specific narrow temperature range (Jackson, 2003). crystalline starch was much less susceptible to enzymatic hydrolysis
Heat treatment of a slurry containing 10% starch led to a complete than amorphous material (Oates, 1997; Oates & Powell, 1996). Thus,
loss of characteristic diffraction peaks (Fig. 4i), indicating that the the retention of crystallinity retards starch liquefaction. In addition,
heated and freeze-dried starch sample had almost no crystallinity. heat not only melts crystallite structures during gelatinization, but
When the slurries contained 20e45% starch, some of the charac- also facilitates new molecular rearrangements or formation of new
teristic diffraction peaks were combined (Fig. 4eeh), indicating bonds among molecules before gelatinization (Ratnayake & Jackson,
that the heated and freeze-dried starch samples had relatively low 2007). As a result, incomplete gelatinization of starch at high con-
crystallinity. As slurries containing higher starch concentrations centrations might be accompanied by molecular rearrangements,
were heated, the starch samples displayed A-type X-ray diffraction which are also unfavorable for starch liquefaction.
patterns with different amounts of decrease in the intensity of the
characteristic diffraction peaks. These starch samples had a rela- 3.3. Effect of starch concentration on the mobility of water
tively high degree of crystallinity. When the slurries contained 55% molecules
or 60% starch, the relative crystallinities of starch samples reached
16.5% and 18.3%, respectively, which were comparable to that of Water molecules play an important role in the gelatinization of
native corn starch (21.2%) Thus, the results obtained using X-ray starch. NMR spectroscopic techniques can be used to probe the
diffraction are largely consistent with those obtained using light mobility of water molecules in aqueous solution. Constant spine-
microscopy. spin relaxation times (T2s) are closely related to the mobility of
water molecules (Baranowska, Sikora, Kowalski, & Tomasik, 2008;
3.2.3. DSC analysis Ritota, Gianferri, Bucci, & Brosio, 2008; Wang, Gu, Li, Hong, &
The thermal properties of slurries containing different concen- Cheng, 2013). In general, aqueous solutions with higher T2 values
trations of starch were analyzed using DSC. The DSC enthalpy contain higher fractions of mobile water molecules.
represents the net sum of all endothermic processes that take place To investigate the effect of starch concentration on the mobility
during heating (Ratnayake, Otani, & Jackson, 2009). As shown in of water molecules, the slurries containing different concentrations
Table 1, the enthalpy change upon starch gelatinization decreased of corn starch were incubated at 90  C for 1 h, and then the T2
as the starch concentration increased. This resulted in an increase values of these starch pastes were determined. As shown in Fig. 5,
in residual enthalpy, which is defined as the difference between the with the increase in starch concentration, the T2 values decreased
enthalpic value for the restricted water sample and that for the high gradually, suggesting that the pastes with a high starch concen-
moisture condition when gelatinization is complete. For slurries tration had lower water molecule mobilities. An increase in starch
containing high concentrations of starch, the residual enthalpy concentration from 10% to 45% resulted in a >90% decrease in the T2
existed because the starches contained too little moisture to permit value of the starch paste. Although further increases in starch
concentration did not cause a noticeable decrease, the T2 values of
starch pastes at 50e60% were very low (<30 ms), indicating very
poor mobility of water molecules in the pastes.

Table 1
Thermal properties of slurries containing different concentrations of corn starch.

Starch concentration Enthalpy (J/g)c Residual enthalpyb


(%, w/w, dry basis)
(J/g) (%)

10 a e e
20 12.2 ± 0.5g 0 0
30 10.6 ± 0.4f 1.6 13.1
40 8.6 ± 0.3e 3.6 29.5
45 6.9 ± 0.3d 5.3 43.4
50 4.5 ± 0.2c 7.7 63.1
55 2.3 ± 0.1b 9.9 81.1
60 1.2 ± 0.1a 11.0 90.2
a
Not determined, primarily because of excessive moisture content.
b
Calculated using the difference in enthalpy between the endotherm at 20% and
Fig. 4. X-ray diffraction patterns of samples containing different concentrations of other starch concentrations.
corn starch, after heat treatment at 90  C for 1 h a: native starch, b: 60%, c: 55%, d: 50%, c
Values are means ± SD (n ¼ 3). Means with different online letters within the
e: 45%, f: 40%, g: 30%, h: 20%, i: 10%. same column are significantly different (p < 0.05).
Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196 195

from a slurry with starch concentration of 10%. By comparison,


increasing the starch concentration from 30 to 45% caused an
approximately 7% reduction in the DE values of starch hydrolysate
after 2 h of liquefaction, indicating that starch liquefaction at an
initial concentration of 45% was comparable with that at 30%.
However, further increases in the initial concentration resulted in
much lower degrees of starch liquefaction. The DE values of starch
hydrolysates obtained by 2 h liquefaction of slurries containing 50
or 55% starch were approximately 25% or 45% lower, respectively,
than that of the hydrolysate produced from a slurry with starch
concentration of 45%. When the starch concentration reached 60%,
the starch paste could not be easily stirred during gelatinization,
resulting in a very low DE value of the starch hydrolysate after 2 h of
liquefaction (data not shown). Taken as a whole, these results
indicate that starch slurries that are less well liquefied correspond
to those that were less well gelatinizated by heat treatment and had
lower water molecule mobilities.
Fig. 5. Dependence of the spinespin T2 value on the concentration of corn starch. Each
value represents the mean of three independent measurements. Bars with different
letters on the top are significantly different (p < 0.05). 3.5. Effect of starch concentration on the viscosity profiles of starch
hydrolysate
Previous reports have shown that the rates of enzymatic re-
actions are closely related to diffusion in high viscosity aqueous During starch liquefaction, the viscosity of the starch paste
solutions (Cicerone & Soles, 2004; Neri, Pittia, Bertolo, Torreggiani, initially increased rapidly, primarily due to starch gelatinization
& Sacchetti, 2010). However, diffusion would be impeded by rela- and shear-thickening, which results from the break-up of highly
tive low mobility of water molecules in the reaction system. Thus, concentrated gel-like starch clusters and resultant increase in the
an increase in starch concentration would reduce the effective effective starch concentration (Kim, Willett, Carriere, & Felker,
diffusion of enzyme and substrate, which could negatively affect 2002). Despite the relatively low degree of gelatinization, slurries
the liquefaction of starch catalyzed by a thermostable a-amylase. containing 40e50% starch might exhibit obvious shear-thickening
behaviors at a rapid agitation speed (200 rpm). The viscosity then
decreased with the liquefaction time, primarily due to the degra-
3.4. Effect of starch concentration on the DE values of starch
dation of starch molecules by the thermostable a-amylase. As
hydrolysate
shown in Fig. 7, different initial starch concentrations in the slurry
resulted in distinct viscosity profiles. The higher the initial starch
During liquefaction, starch is converted to shorter chains and
concentration, the higher the viscosity of the starch pastes. If the
less viscous dextrins (Henderson & Doyal, 2011). In this study, corn
starch concentration increased from 10% to 40%, the peak viscosity
starch was liquefied by the thermostable a-amylase at 12 U/g of dry
of the starch paste during liquefaction increased significantly.
starch. As shown in Fig. 6, the DE values of starch hydrolysates
When the starch concentration was controlled at 45% or 50%, the
increased gradually with liquefaction time. When viewed at the
Brabender viscoamylograph did not produce smooth viscosity
same liquefaction time, slurries with higher initial starch concen-
curves during liquefaction, probably because very high viscosity led
trations had lower DE values. As the starch concentration increased
to difficulty in stirring the starch paste. A further increase in starch
from 10% to 30%, the degree of starch liquefaction decreased
concentration, to above 50%, resulted in a completely irregular
noticeably. After 2 h of liquefaction, slurries with starch concen-
viscosity curve during gelatinization and liquefaction (data not
trations of 20% or 30% yielded starch hydrolysates with DE values 10
shown). When the starch concentration is 60%, a slurry with a very
or 20% lower, respectively, than that seen in a hydrolysate produced
low degree of gelatinization might exhibit a solid-like behavior
(Crawford et al., 2013).

Fig. 6. The effect of corn starch concentration on the DE values of starch hydrolysate
during liquefaction. Each value represents the mean of three independent measure- Fig. 7. Changes in the viscosity profile during liquefaction of slurries containing
ments. Deviations from the mean were below 5% for all values displayed. different starch concentrations. a: 10%; b: 20%; c, 30%; d, 40%; e, 45%; f, 50%.
196 Z. Li et al. / Food Hydrocolloids 48 (2015) 189e196

4. Conclusions Jackson, D. S. (2003). STARCH j functional properties. In B. Caballero (Ed.), Ency-


clopedia of food sciences and nutrition (2nd ed.). (pp. 5572e5575). Oxford: Ac-
ademic Press.
The starch concentration in the slurry had a substantial effect on Khedher, I. B. A., Bressollier, P., Urdaci, M. C., Limam, F., & Marzouki, M. N. (2008).
starch gelatinization. High starch concentrations inhibited swelling Production and biochemical characterization of Sclerotinia sclerotiorum a-
and disruption of starch granules and caused retention of starch amylase ScAmy(1): assay in starch liquefaction treatments. Journal of Food
Biochemistry, 32(5), 597e614.
crystallinity after heat treatment, to varying degrees. When the Kim, S., Willett, J. L., Carriere, C. J., & Felker, F. C. (2002). Shear-thickening and shear-
starch concentration was above 45%, it was very difficult to destroy induced pattern formation in starch solutions. Carbohydrate Polymers, 47(4),
the granule and crystalline structure of starch using heat treatment. 347e356.
Konsula, Z., & Liakopoulou-Kyriakides, M. (2004). Hydrolysis of starches by the action
The starch concentration also substantially influenced the high- of an a-amylase from Bacillus subtilis. Process Biochemistry, 39(11), 1745e1749.
temperature liquefaction of corn starch catalyzed by a thermo- Koo, S. H., Lee, K. Y., & Lee, H. G. (2010). Effect of cross-linking on the physico-
stable a-amylase. The relatively low degree of liquefaction seen at chemical and physiological properties of corn starch. Food Hydrocolloids,
24(6e7), 619e625.
high starch concentrations was intimately related to the incom- Lee, S. H., Shetty, J..,K., & Strohm, B..,A. (2013). Liquefaction and saccharification of
plete gelatinization of the starch and low mobility of water mole- granular starch at high concentration. International Patent, WO 2013048700 A1.
cules in the reaction system. To improve the liquefaction of slurries Li, Z. F., Cai, L. M., Gu, Z. B., & Shi, Y. C. (2014). Effects of granule swelling on starch
saccharification by granular starch hydrolyzing enzyme. Journal of Agricultural
containing high concentrations of starch, it will be necessary to and Food Chemistry, 62(32), 8114e8119.
increase the mobility of water molecules, to swell or damage starch Mandala, I. G., & Bayas, E. (2004). Xanthan effect on swelling, solubility and vis-
granules, and to destroy starch crystallinity by pretreatment. cosity of wheat starch dispersions. Food Hydrocolloids, 18(2), 191e201.
Neri, L., Pittia, P., Bertolo, G., Torreggiani, D., & Sacchetti, G. (2010). Influence of
water activity and molecular mobility on peroxidase activity in salt and sorbi-
Acknowledgments tolemaltodextrin systems. Journal of Food Engineering, 101(3), 289e295.
Nikoli c, S., Mojovic, L., Rakin, M., Pejin, D., & Savi c, D. (2008). A microwave-assisted
liquefaction as a pretreatment for the bioethanol production by the simulta-
This work was financially supported by the Twelfth Five-Year neous saccharification and fermentation of corn meal. Chemical Industry and
National Key Technology Research and Development Program of Chemical Engineering Quarterly, 14(4), 231e234.
the Ministry of Science and Technology of China (No. Oates, C. G. (1997). Towards an understanding of starch granule structure and hy-
drolysis. Trends in Food Science & Technology, 8(11), 375e382.
2012BAD34B07), and the National Natural Science Foundation of Oates, C. G., & Powell, A. D. (1996). Bioavailability of carbohydrate material stored in
China (No. 31371787). tropical fruit seeds. Food Chemistry, 56(4), 405e414.
Parker, R., & Ring, S. G. (2001). Aspects of the physical chemistry of starch. Journal of
Cereal Science, 34(1), 1e17.
References Presecki, A. V., Blazevic, Z. F., & Vasic-Racki, E. (2013). Mathematical modeling of
maize starch liquefaction catalyzed by alpha-amylases from Bacillus lichen-
Ariff, A. B., Asbi, B. A., Azudin, M. N., & Kennedy, J. F. (1997). Effect of mixing on iformis: effect of calcium, pH and temperature. Bioprocess and Biosystems En-
enzymatic liquefaction of sage starch. Carbohydrate Polymers, 33(2e3), 101e108. gineering, 36(1), 117e126.
Asare, E. K., Jaiswal, S., Maley, J., Baga, M., Sammynaiken, R., Rossnagel, B. G., et al. Ratnayake, W. S., & Jackson, D. S. (2006). Gelatinization and solubility of corn starch
(2011). Barley grain constituents, starch composition, and structure affect starch during heating in excess water: new insights. Journal of Agricultural and Food
in vitro enzymatic hydrolysis. Journal of Agricultural and Food Chemistry, 59(9), Chemistry, 54(10), 3712e3716.
4743e4754. Ratnayake, W. S., & Jackson, D. S. (2007). A new insight into the gelatinization
Baks, T., Kappen, F. H. J., Janssen, A. E. M., & Boom, R. M. (2008). Towards an optimal process of native starches. Carbohydrate Polymers, 67(4), 511e529.
process for gelatinisation and hydrolysis of highly concentrated starchewater Ratnayake, W. S., & Jackson, D. S. (2008). Chapter 5 starch gelatinization. In
mixtures with alpha-amylase from B. licheniformis. Journal of Cereal Science, L. T. Steve (Ed.), Advances in food and nutrition research (Vol. 55, pp. 221e268).
47(2), 214e225. Oxford: Academic Press.
Baranowska, H. M., Sikora, M., Kowalski, S., & Tomasik, P. (2008). Interactions of Ratnayake, W. S., & Jackson, D. S. (2009). Starch gelatinization. Advances in Food and
potato starch with selected polysaccharide hydrocolloids as measured by low- Nutrition Research, 55, 221e268.
field NMR. Food Hydrocolloids, 22(2), 336e345. Ratnayake, W. S., Otani, C., & Jackson, D. S. (2009). DSC enthalpic transitions during
Baruque, E. A., Baruque, M. D. A., & Sant'Anna, G. L. (2000). Babassu coconut starch starch gelatinisation in excess water, dilute sodium chloride and dilute sucrose
liquefaction: an industrial scale approach to improve conversion yield. Bio- solutions. Journal of the Science of Food and Agriculture, 89(12), 2156e2164.
resource Technology, 75(1), 49e55. Richardson, T. H., Tan, X., Frey, G., Callen, W., Cabell, M., Lam, D., et al. (2002).
Bhargava, S., Frisner, H., & Johal, M. (2008). Liquefaction and saccharification process. A novel, high performance enzyme for starch liquefaction. Discovery and
US Patent, US 20080121227 A1. optimization of a low pH, thermostable a-amylase. Journal of Biological Chem-
Blazek, J., & Gilbert, E. P. (2010). Effect of enzymatic hydrolysis on native starch istry, 277(29), 26501e26507.
granule structure. Biomacromolecules, 11(12), 3275e3289. Ritota, M., Gianferri, R., Bucci, R., & Brosio, E. (2008). Proton NMR relaxation study of
Bogracheva, T. Y., Morris, V. J., Ring, S. G., & Hedley, C. L. (1998). The granular swelling and gelatinisation process in rice starchewater samples. Food Chem-
structure of C-type pea starch and its role in gelatinization. Biopolymers, 45(4), istry, 110(1), 14e22.
323e332. Siswoyo, T. A., & Morita, N. (2003). Thermal properties of partially hydrolyzed
Bruner, R. L. (1964). Determination of reducing value: 3,5-Dinitrosalicylic acid starch-glycerophosphatidylcholine complexes with various acyl chains. Journal
method. In R. L. Whistler, R. J. Smith, J. N. BeMiller, & M. L. Wolform (Eds.), of Agricultural and Food Chemistry, 51(10), 3162e3167.
Methods in carbohydrate chemistry (Vol. 4, pp. 67e71). New York: Academic Srichuwong, S., Isono, N., Mishima, T., & Hisamatsu, M. (2005). Structure of lintn-
Press. erized starch is related to X-ray diffraction pattern and susceptibility to acid and
Cheetham, N. W. H., & Tao, L. (1998). Variation in crystalline type with amylose enzyme hydrolysis of starch granules. International Journal of Biological Mac-
content in maize starch granules: an X-ray powder diffraction study. Carbohy- romolecules, 37(3), 115e121.
drate Polymers, 36(4), 277e284. Sun, J., Zhao, R., Zeng, J., Li, G., & Li, X. (2010). Characterization of destrins with
Chen, Y., Huang, S., Tang, Z., Chen, X., & Zhang, Z. (2011). Structural changes of different dextrose equivalents. Molecules, 15(8), 5162e5173.
cassava starch granules hydrolyzed by a mixture of a-amylase and glucoamy- Tester, R. F., & Morrison, W. R. (1990). Swelling and gelatinization of cereal starches.
lase. Carbohydrate Polymers, 85(1), 272e275. I. Effects of amylopectin, amylose, and lipids. Cereal Chemisrty, 67, 551e557.
Cicerone, M. T., & Soles, C. L. (2004). Fast dynamics and stabilization of proteins: Tester, R. F., & Sommerville, M. D. (2001). Swelling and enzymatic hydrolysis of
binary glasses of trehalose and glycerol. Biophysical Journal, 86(6), 3836e3845. starch in low water systems. Journal of Cereal Science, 33(2), 193e203.
Contreras Lo pez, E., Role e, A., & Le Meste, M. (2004). Study of starch granules Uthumporn, U., Zaidul, I. S. M., & Karim, A. A. (2010). Hydrolysis of granular starch at
swelling by the blue dextran method and by microscopy. Starch e Sta €rke, sub-gelatinization temperature using a mixture of amylolytic enzymes. Food
56(12), 576e581. and Bioproducts Processing, 88(1), 47e54.
de Cordt, S., Hendrickx, M., Maesmans, G., & Tobback, P. (1994). The influence of Wang, S., & Copeland, L. (2013). Molecular disassembly of starch granules during
polyalcohols and carbohydrates on the thermostability of a-amylase. Biotech- gelatinization and its effect on starch digestibility: a review. Food Function,
nology and Bioengineering, 43(2), 107e114. 4(11), 1564e1580.
Crawford, N. C., Popp, L. B., Johns, K. E., Caire, L. M., Peterson, B. N., & Wang, Z., Gu, Z., Li, Z., Hong, Y., & Cheng, L. (2013). Effects of urea on freeze-thaw
Liberatore, M. W. (2013). Shear thickening of corn starch suspensions: does stability of starch-based wood adhesive. Carbohydrate Polymers, 95(1), 397e403.
concentration matter? Journal of Colloid and Interface Science, 396, 83e89. Yankov, D., Dobreva, E., Beschkov, V., & Emanuilova, E. (1986). Study of optimum
Eliasson, A. C. (2006). Carbohydrates in food (2nd ed.). Boca Raton, FL: CRC/Taylor & conditions and kinetics of starch hydrolysis by means of thermostable a-
Francis. amylase. Enzyme and microbial technology, 8(11), 665e667.
Henderson, J. M. R., & Doyal, R. J. (2011). Ethanol yields in fermentation from an Zobel, H., Young, S., & Rocca, L. (1988). Starch gelatinization: an X-ray diffraction
improved liquefaction process. EP Patent, EP 2430169 A2. study. Cereal Chemistry, 65(6), 443e446.

You might also like