You are on page 1of 7

Food Chemistry 164 (2014) 317–323

Contents lists available at ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Pea starch (Pisum sativum L.) with slow digestion property produced
using b-amylase and transglucosidase
Miaomiao Shi a, Zhiheng Zhang a, Shujuan Yu a, Kai Wang b,c, Robert G. Gilbert b,c,⇑, Qunyu Gao a,⇑
a
Carbohydrate Laboratory, College of Light Industry and Food Sciences, South China University of Technology, Guangzhou 510640, PR China
b
Tongji School of Pharmacy, Huazhong University of Science and Technology, Wuhan, Hubei 430030, PR China
c
The University of Queensland, Queensland Alliance for Agriculture and Food Innovation, Brisbane, Qld 4072, Australia

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

Article history: Starches extracted from wrinkled (WP) and smooth (SP) peas were treated using b-amylase (B) alone and
Received 20 December 2013 also with a combination of b-amylase and transglucosidase (BT). After enzymatic treatment, the propor-
Received in revised form 7 April 2014 tions of slowly digested starch in WP-B, WP-BT, SP-B and SP-BT samples were increased by 6%, 9%, 9% and
Accepted 7 May 2014
12%, respectively. Starches treated by a combination of b-amylase and transglucosidase exhibited a smal-
Available online 20 May 2014
ler amount of longer amylopectin chains, a larger amount of short amylopectin chains, and higher
branching fraction. The branching fraction was significantly increased, with an increase of 8%, 10%, 13%
Keywords:
and 14% for WP-B, WP-BT, SP-B and SP-BT, respectively. The maximum absorbance and iodine binding
Pea starch
Slowly digestible starch
of enzyme-treated starches were reduced compared with their native starch parents. The C-type crystal-
b-Amylase line structure completely disappeared after enzymatic treatment. The results support previous findings
Transglucosidase that increases in the amount of shorter amylopectin chains and branch fraction are likely to contribute
Chain length distribution to the slow digestion of starch.
1
H NMR Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Cummings, 1992). Different preparations, physicochemical investi-


gations, fine structure analyses and functional properties of RS
Starch is the main carbohydrate material in human nutrition have been reviewed elsewhere (Cai & Shi, 2010; Kumari, Urooj, &
and also has a wide range of industrial applications (Lehmann & Prasad, 2007; Ozturk, Koksel, Kahraman, & Ng, 2009). However,
Robin, 2007). The digestion rate strongly depends on processing the structure of SDS and its potential health benefits are only par-
and the state of the starch (Lehmann & Robin, 2007; Wolever, tially understood.
2003). Sources of starch with a range of content of rapidly digest- Pea (Pisum sativum L., also known as field pea, garden pea and
ible starch (RDS), slowly digestible starch (SDS) and resistant common pea) is one of the oldest domesticated food crops, and is
starch (RS) are of great interest (Pongjanta, Utaipattanaceep, grown worldwide as a cool-season grain legume that provides a
Naivikul, & Piyachomkwan, 2009). Differences in glycemic and ins- good source of dietary protein and energy for humans and live-
ulinemic responses to dietary starch are directly related to the rate stock. The starch and protein contents of the grains range
of starch digestion (O’Dea, Snow, & Nestel, 1981). RDS induces a between 30–50% and 20–25%, respectively, of the dry matter. In
fast increase in blood glucose and insulin levels, whereas SDS pro- the species P. sativum L., two different seed phenotypes exist:
vides for an extended release of glucose, with a low glycemic smooth (with a smooth seed surface) and wrinkled pea (wrinkled
response (Lehmann & Robin, 2007). Foods containing high levels seed surface). The two types are genetically different and produce
of SDS have slow digestion rates, making them beneficial to health starches with different granular morphologies and characteristics.
(Han & BeMiller, 2007). The digestibility of starch in the human Smooth pea is a common commercial legume and is widely
small intestine can be modified from rapidly digestible starch researched. Wrinkled pea is a wild variety cultivated in China,
hydrolysis products, to ‘‘indigestible’’ RS (Englyst, Kingman, & and little is known about the structure of its starch. Some prop-
erties of wrinkled and smooth pea starches are the subject of
the present study.
⇑ Corresponding authors. Tel.: +86 13660261703; fax: +86 2087113848 (Q. Gao).
Tel.: +61 7 3365 4809; fax: +61 7 3365 1188 (R.G. Gilbert).
Starch is a homopolymer of anhydroglucose, with a-D-glucopyr-
E-mail addresses: chengzi3090@126.com (M. Shi), b.gilbert@uq.edu.au anosyl monomeric units linearly extended by a-(1 ? 4) linkages
(R.G. Gilbert), qygao@scut.edu.cn (Q. Gao). and branches formed by a-(1 ? 6) branching points. Starch

http://dx.doi.org/10.1016/j.foodchem.2014.05.045
0308-8146/Ó 2014 Elsevier Ltd. All rights reserved.
318 M. Shi et al. / Food Chemistry 164 (2014) 317–323

comprises basically two types of macromolecules: amylose, with a 2.3. Preparation of starch samples with slow digestion properties
few long-chain branches and relatively low molecular weight
(105–106), and amylopectin, a highly branched molecule with Smooth or wrinkled pea starch (50 g, dry weight) was mixed
much higher molecular weight (107–109). The molecular struc- with sodium acetate buffer (0.1 M, pH 5.0) to a 10% suspension
ture of starch has been shown to affect the digestion rate of the by weight. The suspension was heated at 95 °C for 30 min with
starch. One way to increase SDS is by increasing the ratio of short stirring followed by incubation in a 58 °C water bath with b-amy-
chains (degree of polymerisation, DP, < 13) to longer chains lase, with or without transglucosidase, for 12 h. The enzyme
(DP P 13) of amylopectin (Zhang, Sofyan, & Hamaker, 2008). amounts used were 130 U/mL and 12 kU/mL for b-amylase and
Another way is to increase the relative amounts of a-(1 ? 6) link- transglucosidase, respectively. Then 1 volume of 95% ethanol was
ages by using branching enzyme (Backer & Saniez, 2005; Shin, added followed by centrifugation at 1500g for 10 min. The precip-
Simsek, Reuhs, & Yao, 2008) giving smaller chains and a higher itated starch was then washed with deionized water and collected
branching fraction. by centrifugation; this step was repeated twice. The collected
In this study, the structure of pea starch was modified using starch was dried in an oven at 45 °C (air stream) for 2 days. The
b-amylase and transglucosidase. b-Amylase, an exo-amylase, dried starch was then ground into powder (100-mesh) and stored
reduces the chain length of starch by catalysing the successive in a desiccator for further analysis. These enzymatically hydrolysed
removal of maltose from the non-reducing ends of the starch samples were prepared and termed as follows: b-amylase-treated
chains (accompanied by inversion of the anomeric configuration) smooth pea starch (SP-B); b-amylase-treated wrinkled pea starch
(Derde, Gomand, Courtin, & Delcour, 2012). Transglucosidase catal- (WP-B); b-amylase-and transglucosidase-treated smooth pea
yses hydrolytic and transfer reactions to form new a-(1 ? 6) link- starch (SP-BT); and b-amylase-and transglucosidase-treated wrin-
ages and thus increases the branch density (Ao et al., 2007). kled pea starch (WP-BT).
The physicochemical properties of modified starches from
wrinkled peas have been characterised (Ratnayake, Hoover, & 2.4. In-vitro digestion with porcine pancreas a-amylase
Warkentin, 2002), such as cross-linking and hydroxypropylation
(chemical modification), annealing and heat moisture treatment The digestion properties were analysed using the method of
(physical modification). However, information on enzymatic Zhang and Hamaker (1998) with some modifications. Starch
modification of wrinkled pea starch is still lacking. In this study, (50 mg) with 10 mL phosphate buffer (0.2 M, pH 6.9) was cooked
starches from smooth and wrinkled peas were enzymatically in a boiling water bath for 30 min. The solution was equilibrated
modified by b-amylase or with a combination of b-amylase at 37 °C for 10 min, and 300 U of porcine pancreas a-amylase
and transglucosidase to produce starches with shorter and was added. Enzyme digestion was carried out at 37 °C, and
more highly branched chains. The structural features of these 0.1 mL aliquots of hydrolysed solution were collected at 20, 40,
modified starch products, which have slow digestibility, were 60, 80, 120, 180 and 240 min. The aliquots were immediately
investigated. mixed with 3 mL 95% ethanol to deactivate the enzyme. 3 mL of
dinitrosalicylic acid (DNS) reagent was then added. Samples were
then placed in boiling water for 15 min. The absorbance at
2. Materials and methods
550 nm of the solution was then evaluated using a spectrometer
(Model 722sp, Shanghai LengGuang Tech. Ltd., China). The maltose
2.1. Materials
content was determined from a calibration curve obtained using
known amounts of maltose against their corresponding absor-
Smooth pea starch was purchased from Yantai Dongfang
bance. Maltose is not the only product produced by a-amylase:
Protein Science and Technology Co., Ltd., China. Wrinkled pea (a
Jane and Robyt (1984) showed that the products of starch hydroly-
wild variety cultivated in China) was obtained from Dingxi Gansu
sis of porcine pancreatic a-amylase are mainly maltose, maltotri-
Province. b-Amylase from barley and transglucosidase L ‘‘Amano’’
ose and maltotetraose. In this study, the content of the dominant
from Aspergillus niger were obtained from Amano Enzyme Inc.
product, maltose, was used to represent the digestibility profiles.
(Nagoya, Japan). Pancreatin from porcine pancreas (P7545),
a-amylase from porcine pancreas (A3176) and amyloglucosidase 2.5. In-vitro digestion with pancreatin and amyloglucosidase
from A. niger (A7095) were purchased from Sigma–Aldrich Chem-
ical Co. (St. Louis, USA). Glucose oxidase–peroxidase (GOPOD)
The digestion properties of starch samples were analysed using
assay kit and isoamylase from Pseudomonas sp. were from Mega-
the method of Englyst et al. (1992) with modifications. The enzyme
zyme International Ireland Ltd. (Wicklow, Ireland). Panose was
solution was prepared by suspending pancreatin (2.25 g, 8USP) in
obtained from Hayashibara Biochemical Laboratories, Inc. (Oka-
sodium acetate buffer (7.5 mL, 0.1 M, pH 5.0) with magnetic
yama, Japan). A series of pullulan standards were purchased from
stirring for 30 min, and then centrifuged for 10 min at 1500g. The
Polymer Standards Service GmbH (Mainz, Germany). Chemicals
supernatant was transferred into a beaker and mixed with
and solvents were of analytical grade.
0.75 mL of amyloglucosidase (300 U/mL) before use. Starch
(300 mg) with 10 mL of sodium acetate buffer (0.1 M, pH 5.0)
2.2. Starch isolation was cooked in a boiling water bath for 30 min, followed by an
equilibration at 37 °C for 10 min, and then a 0.75 mL mixture of
Starch was extracted using the method of Beta, Corke, Rooney, pancreatin and amyloglucosidase was added. Enzyme digestion
and Taylor (2001). Wrinkled pea (1000 g) was steeped in was carried out in a 37 °C water bath at 150 rpm, and 0.5 mL
2000 mL of water at room temperature for 24 h. The steeped pea aliquots of hydrolysed solution were collected at 20 and 120 min.
was washed and ground with an equal volume of water. The slurry Then 20 mL of ethanol (95%) was added to the aliquots to deacti-
was filtered through a 200-mesh screen. The material remaining vate the enzyme. After centrifugation (1500g, 10 min), the glucose
on the sieve was rinsed twice with deionized water. The filtrate content was determined using the glucose oxidase/peroxidase
was subsequently washed several times with NaOH (0.2% w/v) (GOPOD) assay kit. The percentage of hydrolysed starch was calcu-
until the grey, top protein-rich layer was removed. The starch lated by multiplying the glucose content by a factor of 0.9, which is
was washed with water to remove residual NaOH and dried for the molar mass conversion from glucose to anhydroglucose (the
24 h at 45 °C. starch monomer unit). The values of RDS, SDS and RS were
M. Shi et al. / Food Chemistry 164 (2014) 317–323 319

obtained by combining the values of G20 (glucose released at The relative crystallinity was estimated by calculating the ratio
20 min), G120 (glucose released at 120 min), FG (free glucose) of the crystalline area compared to the total diffractogram area.
and TS (total starch) using the following formulae:
2.9. Nuclear magnetic resonance spectroscopy (NMR)
RDS ð%Þ ¼ 0:9ðG120  FGÞ=TS
1
H NMR analyses of starch samples were performed using a
SDS ð%Þ ¼ 0:9ðG120  G20Þ=TS
Varian Unity Inova 300 MHz NMR spectrometer (Varian Inc.,
USA) using the method of Tizzotti, Sweedman, Tang, Schaefer,
RS ð%Þ ¼ 1  RDS%  SDS% and Gilbert (2011) with some modifications. Panose, containing
50.74% a-(1 ? 6) linkages, was used as a reference. Starch (6 mg)
2.6. Chain length distributions (CLDs) of starch samples using size- was dissolved in a DMSO-d6 (0.7 mL)/trifluoroacetic acid-d1
exclusion chromatography (SEC) (20 lL) mixture, and the 1H NMR spectra measured at 80 °C. The
DB (degree of branching) of starch is calculated using:
All starch samples were debranched using isoamylase in an Iað1!6Þ
acetate buffer solution (0.1 mL, 0.1 M, pH 3.5) and freeze dried DBð%Þ ¼ 100
Iað1!4Þ þ Iað1!6Þ
overnight following a method described elsewhere (Hasjim,
Lavau, Gidley, & Gilbert, 2010; Tran et al., 2011). The dried deb- where Ia-(1 ? 4) and Ia-(1 ? 6)are the 1H NMR integrals of internal
ranched starch was then dissolved overnight in DMSO/0.5% wt LiBr a-(1 ? 4) and a-(1 ? 6) linkages, respectively.
solution in a thermomixer (Eppendorf, Hamburg, Germany) at
80 °C with shaking at 350 rpm, followed by centrifugation to 2.10. Statistical analysis
remove insoluble components.
The SEC weight CLDs of debranched starch samples were deter- The differences between the mean values of multiple groups
mined using an Agilent 1100 Series SEC (Agilent Technologies, were analysed by one-way analysis of variance (ANOVA) with
Waldbronn, Germany) coupled with an isocratic pump, a series Duncan’s multiple range tests. ANOVA data with P < 0.05 were
of separation columns (GRAM precolumn, GRAM 30, and 1000 ana- classified as statistically significant. SPSS 17.0 software and Origin
lytical columns, Polymer Standard Services, Mainz, Germany), and 8.0 was used to analyse and report the data. Mean values were
a refractive index detector (RID; ShimadzuRID-10A, Shimadzu from triplicate experiments.
Corp., Japan) following a method described elsewhere (Cave,
Seabrook, Gidley, & Gilbert, 2009; Vilaplana & Gilbert, 2010). The
3. Results and discussion
DMSO/LiBr solution was used as mobile phase after being filtered
through a 0.45 lm hydrophilic Teflon membrane filter (Millipore,
3.1. In-vitro digestion of starch samples with porcine pancreas
Billerica, MA, USA). Pullulan standards with peak molecular
a-amylase and the combination of pancreatin and amyloglucosidase
weights ranging from 342 to 2.35  106 were used for calibration
to convert SEC elution volume to the molecular size (hydrody-
The Englyst assay was used to study the digestion of starch
namic volume, Vh, or equivalently hydrodynamic radius, Rh) using
samples in vitro, because it has been correlated with human stud-
the Mark–Houwink equation (Cave et al., 2009). Note that SEC
ies (Englyst et al., 1992) and thus can be used as a reference to
gives the weight distribution of the debranched chains as a func-
in vivo digestion behaviour. Fig. 1 shows the digestibility profiles
tion of Rh, w(logRh): the weight of chains containing X monomer
of treated wrinkled and smooth pea starches compared with the
units in the size increment d(log Rh). This is related to the CLD,
native starches. Native wrinkled and smooth pea starches showed
the number distribution Nde(X) (the relative number of chains con-
similar digestibilities. Starch samples treated with b-amylase and
taining X monomer units) by Nde(X) = X2 w(logRh) (Castro, Ward,
the combination of b-amylase and transglucosidase showed much
Gilbert, & Fitzgerald, 2005).
lower digestibilities compared with the untreated starches. For
wrinkled pea starch samples, the digestibility of WP-B was slightly
2.7. Iodine binding analysis higher than that of WP-BT. Smooth pea starches showed the same
trend, with SP-BT having a reduced digestibility compared with
Iodine binding analysis was performed using a UV/visible spec- SP-B. Thus the digestibility of starches treated with b-amylase
trophotometer (I5, Hanon Instrument Co., Ltd., Guangzhou, China) and transglucosidase were lower than those of starches treated
following a method published elsewhere (Shin et al., 2007). An with b-amylase alone, suggesting that the addition of transglucos-
iodine reagent was prepared by adding 2 mg of I2 and 20 mg of idase could further decrease the digestibility of starches modified
KI to1 mL of deionized water. Starch (80 mg) was dissolved in by b-amylase (Ao et al., 2007). In addition, the increased maltose
20 mL of DMSO. An aliquot of the starch in DMSO solution (1 mL) content of SP-B and SP-BT prior to 120 min was lower than that
was diluted to 100 mL using deionized water. 2.8 mL of the diluted of WP-B and WP-BT, respectively, indicating that the digestibility
starch solution was placed into a 1 cm cuvette and mixed with of enzymatically-treatment smooth pea starch is lower than
75 lL of iodine reagent. The absorbance spectra and wavelength that of treated wrinkled pea starch.
of maximum absorption (kmax) were analysed over a scan of wave- To further evaluate the release of glucose from enzymatically
lengths from 450 to 800 nm. treated starch, samples were incubated with a combination of
pancreatin and amyloglucosidase. Pancreatin a-amylase itself has
2.8. Wide-angle X-ray diffraction patterns no debranching activity and can only cleave a-(1 ? 4) linkages at
random locations; amyloglucosidase, on the other hand, produces
A D/Max-2200 X-ray diffractometer (Rigaku Denki Co., Tokyo, b-glucose by hydrolysing both a-(1 ? 4) and a-(1 ? 6) linkages
Japan) was used to obtain X-ray diffractograms. Cu Ka radiation at a slower rate (Lee et al., 2013; Weill, Burch, & Van Dyk, 1954).
at 44 kV and 26 mA was applied. Before analysis, starch samples The contents of RDS, SDS, and RS are summarised in Table 1.
were equilibrated in a sealed desiccator with water at room tem- Cooked WP starch had 88.99% RDS, 6.94% SDS and 4.07% RS, while
perature for 12 h. The diffractogram scanning was run between cooked SP showed relatively higher RDS content (91.29%) and
4° and 35° (2h) at a rate of 5°/min (Shi, Chen, Yu, & Gao, 2013). lower SDS and RS contents (5.13% and 3.58%, respectively).
320 M. Shi et al. / Food Chemistry 164 (2014) 317–323

25 DP
(A)
Increased maltose content (mg/mL)
5 10 25 100 1000 5000 10000
WP
20 (A) 17

Normalized w(logRh) (arb. units)


WP-B
15 WP-BT
WP
10 SP
44

5 324
804

0
0 60 120 180 240 300 360
Incubation time (min)
0.1 1 10 100
Rh (nm)
25
(B)
Increased maltose content (mg/mL)

SP DP
20 5 10 25 100 1000 500010000
SP-B (B) 14 32

Normalized w(logRh) (arb. units)


15
SP-BT
WP-B
10 4 14 26 WP-BT

5
186
0
0 60 120 180 240 300 360
164
Incubation time (min)
2
Fig. 1. Digestion profiles of native and enzymatically treated starches: (A) wrinkled 0.1 1 10 100
pea starches, (B) smooth pea starches. WP = native wrinkled pea starch, SP = native Rh (nm)
smooth pea starch, WP-B = b-amylase treated wrinkled pea starch, WP-BT =
b-amylase and transglucosidase treated wrinkled pea starch, SP-B = b-amylase
treated smooth pea starch, SP-BT = b-amylase and transglucosidase treated smooth DP
pea starch. 5 10 25 100 1000 5000 10000

(C) 16
Normalized w(logRh) (arb. units)

32
Table 1
Contents (%, w/w, dry weight) of RDS, SDS and RSa.
29 177 SP-B
Samples RDS (%) SDS (%) RS (%) 4 16 SP-BT
WP 88.99 ± 2.01d 6.94 ± 0.66a 4.07 ± 2.67a
WP-B 79.62 ± 0.67c 12.82 ± 0.53b 7.56 ± 0.14b
WP-BT 66.72 ± 1.33a 15.70 ± 1.06 cd 17.58 ± 2.39d
SP 91.29 ± 0.98d 5.13 ± 1.67a 3.58 ± 0.69a 199
SP-B 72.82 ± 1.26b 14.04 ± 1.23bc 13.14 ± 0.03c
SP-BT 65.17 ± 1.44a 17.26 ± 0.71d 17.57 ± 0.73d
a
Values with different letters in the same column are significantly different
2
(P < 0.05).
0.1 1 10 100
Rh (nm)
Compared to their counterpart native starches, rapidly digested
starch contents of WP-B, WP-BT, SP-B and SP-BT were decreased Fig. 2. Chain length distributions (CLDs) of native and enzymatically treated
by 9.37%, 22.27%, 18.47% and 26.12%, respectively; slowly starches: (A) native wrinkled pea starch (WP)and smooth pea starch (SP), (B) WP-B
digestible starch contents of these samples were increased by and WP-BT, (C) SP-B and SP-BT. Numbers above the curves are degree of
5.88%, 8.76%, 8.91% and 12.13%, respectively. The resistant starch polymerisation (DP) values.
contents of enzymatically treated starches were also increased.
These results show that enzymatically treated starches are signifi- amylopectin, trans-lamellar amylopectin and amylose chains.
cantly different from their native starches, with slower digestion Regarding the CLDs of the two native starches, SP, compared with
properties. WP, has more shorter amylose chains with DP ranging from
approximately 100 to 800, but less longer amylose chains with
3.2. CLDs DP larger than 800. After being treated by b-amylase alone or the
combination of b-amylase and transglucosidase, amylose chains
The SEC weight CLDs of native pea starches and their deriva- were degraded into much shorter ones, with the peak maximum
tives subjected to b-amylase and the combination of b-amylase shifted to much smaller DP values. Amylose chains with
and transglucosidase treatments are presented in Fig. 2. All DP > 3000 were not visible in the SEC debranched distributions of
distributions were normalised to yield the same height of the enzyme-treated starches, meaning that those chains were totally
highest peak maximum to avoid the influences of different sample degraded into chains shorter than DP 3000.
concentrations. Typical SEC weight CLDs are observed for all starch After enzymatic treatment, the two amylopectin peaks were
samples, with three peaks corresponding to single-lamellar shifted from DP 17 and 44 to DP 14–16 and 26–32, respectively
M. Shi et al. / Food Chemistry 164 (2014) 317–323 321

(Fig. 2B and C), suggesting that, compared with the native starches, 0.6
higher amounts of shorter amylopectin chains were produced by (A)
these enzymes. New peaks ranging from DP  2 to 6 were observed 0.5
WP
for starches treated with b-amylase; however, these peaks
0.4
became much smaller for starches treated with the combination

Absorbance
of b-amylase and transglucosidase, with only two small bumps
0.3
observed at DP about 2 and 4. Note that these values of the DP WP-B
are inferred using the Mark–Houwink relation, rather than mea- 0.2
sured directly, and are subject to some uncertainty (Cave et al., WP-BT
2009; Vilaplana & Gilbert, 2010). These results demonstrate that 0.1
the effects of b-amylase and transglucosidase treatment are signif-
icant. b-Amylase trims off amylose chains and amylopectin 0.0
500 600 700 800
external chains by hydrolysing a-(1 ? 4) linkages from the
non-reducing ends of chains, releasing maltose. Following a multi- Wavelength (nm)
ple attack mechanism, it degrades starch successively until there 0.6
are 2 or 3 glucose units left on the ‘‘stub’’. b-Amylase cannot bypass (B) SP
a-(1 ? 6) linkages; therefore it cannot hydrolyse internal chains of 0.5
amylopectin, because it cannot bypass a-(1 ? 6) linkages SP-B
(Bijttebier, Goesaert, & Delcour, 2010; Yao, Zhang, & Ding, 2003). 0.4

Absorbance
Peaks covering DP  2–6 in the CLDs of WP-B and SP-B are attrib- SP-BT
uted to the very short chains originating directly from the ‘‘stubs’’ 0.3
left over. Shen, Bertoft, Zhang, and Hamaker (2013) also reported
0.2
that large changes in the CLDs of waxy and amylose-extender
waxy starches occurred after starches being hydrolysed using 0.1
b-amylase, with new peaks emerged at DP 2–4. The b-amylolysis
of amylopectin proceeds in two stages: a rapid progression up to 0.0
52% hydrolysis, in which chains are shortened to maltotetraosyl 500 600 700 800
residues (DP 4) and then a slow progression to completion (Lee, Wavelength (nm)
1971). With the combination of b-amylase and transglucosidase,
the maltose molecules released by b-amylase are transferred to Fig. 3. Iodine binding curves of native and enzyme treated starches: (A) wrinkled
pea starches, (B) smooth pea starches.
suitable acceptor substrates by transglucosidase, forming new
a-(1 ? 6) linkages on starch chains. As a result, more internal amy-
lopectin chains that cannot be hydrolysed by b-amylase formed, length of starch decreased after b-amylase treatment or the combi-
which leads to the smaller proportions of very short chains with nation of b-amylase and transglucosidase treatment, there were a
DP 2–6 in the CLDs of WP-BT and SP-BT, compared with WP-B reduced number of helical turns, leading to a lower iodine binding
and SP-B. On the other hand, the removal of maltose by transglu- capacity. This confirmed that the occurrence of iodine complexes
cosidase increases the reaction speed of b-amylase hydrolysis, with internal or external chains involves loss of iodine binding
resulting in higher proportions of shorter chains. This explains ability after shortening of external chains by b-amylase (Shen
the phenomena observed in Fig. 2B and C that the maximum DPs et al., 2013).
of each peak are smaller for WP-BT and SP-BT than those of
WP-B and SP-B. The amylopectin structural changes are consistent
with results with maize starch reported by Miao et al. (2014) and 3.4. X-ray diffraction
Ao et al. (2007). However, those two groups only focused on
amylopectin chains but not amylose components. This is the first The diffraction patterns of native and treated WP and SP starch
study reporting the impacts of b-amylase and transglucosidase samples are shown in Fig. 4. It is noted that there is some residual
on amylose fine structure. crystallinity for SP-B, although the sample was gelatinized before
treatment with enzyme. This effect has been seen elsewhere: Ao
3.3. Iodine binding analysis et al. (2007) and Casarrubias-Castillo, Hamaker, Rodriguez-
Ambriz, and Bello-Pérez (2012) saw this with normal maize and
Iodine can complex with amylose and some amylopectin chains plantain starches treated with b-amylase. These authors gave the
to produce a characteristic colour (Zheng et al., 2013). The tested reasonable explanation that starch hydrolysed with b-amylase
colour intensity and the wavelength of maximum absorbance can form new structures with more exposed long interior chains
(kmax) are affected by the DP of the starch chains and thus can be that are more easily retrograded. The retrograded fraction can
used to explore the structural change of starches subject to partly crystallize. In the case of native WP and SP, the diffraction
enzymes. As shown in Fig. 3, the iodine binding curves of native pattern is mainly composed of five characteristic profiles at
starches and enzymatically modified starches shows different kmax 2h = 5.6°, 11.5°, 15.4°, 17.6° and 23.6°, respectively. These are
and absorptivity values. The maximum absorbance values of WP indicative of C-type starch diffraction, which is a mixture of
and SP were at 621 and 620 nm, respectively; however, those of A- and B-type crystallinity within the granule. Ratnayake et al.
WP-B, WP-BT, SP-B and SP-BT were at significantly lower wave- (2002) reported that a wrinkled pea starch from Canada displayed
lengths, ranging from 595 to 599 nm. This decreased maximum a B-type crystalline pattern. The differences in starch crystalline
wavelength is caused by the decreased average chain length of structure reflect the variations in the molecular structure of these
enzymatically modified starch. The iodine binding reduced sub- two wrinkled pea starches, attributed to the differences in the
stantially from 0.52 to 0.25 (WP-B) and 0.23 (WP-BT) for wrinkled genetic background or growth conditions between those two vari-
pea starch and was reduced from 0.56 to 0.48 (SP-B) and 0.47 eties. The relative crystallinity for WP and SP is 30.7% and 29.2%,
(SP-BT) for smooth pea starch. The iodine binding values of smooth respectively. Untreated WP showed a lower crystallinity level than
pea samples were higher than those of wrinkled pea. As the chain untreated SP, indicating a structural difference in a range of starch
322 M. Shi et al. / Food Chemistry 164 (2014) 317–323

Table 2
(A) Degree of branching (DB) of unmodified and enzyme treated wrinkled (WP) and
smooth (SP) pea starches.

Samples Panose WP WP- WP- SP SP-B SP-


Relative intensity

WP-BT B BT BT
DB from 1H NMR 50.74 4.76 13.04 16.67 5.66 13.79 15.25
WP-B (%)
DB from SEC (%) – 4.03 10.26 7.61 4.07 7.85 5.74

WP = native wrinkled pea starch, SP = native smooth pea starch, WP-B = b-amylase
WP
treated wrinkled pea starch, WP-BT = b-amylase and transglucosidase treated
wrinkled pea starch, SP-B = b-amylase treated smooth pea starch, SP-BT = b-amy-
lase and transglucosidase treated smooth pea starch.
5 10 15 20 25 30 35
Diffraction angle (2θ )
R1
1 Nde ðXÞdX
(B) DB ¼  ¼ R 10
Xde ðRh Þ 0
XN de ðXÞdX

Ratnayake, Hoover, Shahidi, Perera, and Jane (2001) and


Relative intensity

SP-BT
Silverio, Fredriksson, Andersson, Eliasson, and Åman (2000)
reported amylopectin Nde(X) of five pea starch cultivars/varieties
characterised using high-performance anion exchange chromatog-
SP-B raphy (HPAEC). The calculated DBs obtained here using their data
are 4.39%, 4.57%, 4.69%, 4.47% and 5.48%, respectively, which are
SP in good agreement with DBs values obtained using 1H NMR. It
should be noted that the DBs of starch are generally lower than
that of the counterpart amylopectin, because amylose molecules
5 10 15 20 25 30 35
contribute a significant proportion of a-(1 ? 4) linkages but few
Diffraction angle (2θ )
a-(1 ? 6) linkages. The DBs of native and enzymatically treated
Fig. 4. X-ray diffraction patterns of native and enzyme treated starches: (A)
smooth pea and winkled pea starches were also calculated follow-
wrinkled pea starches, (B) smooth pea starches. ing the method described using Nde(X) obtained from SEC. The
calculated results are shown in Table 2. The DB values obtained
from SEC and 1H NMR are similar for WP and SP; however, they
are significantly different for enzymatically treated starch samples,
components in the granule (Millan-Testa, Mendez-Montealvo, with DBs from SEC showing systematically smaller values. This
Ottenhof, Farhat, & Bello-Pérez, 2005). means that SEC and 1H NMR give similar DBs for native starches
b-Amylase and transglucosidase treatments have pronounced (Syahariza, Sar, Hasjim, Tizzotti, & Gilbert, 2013); however, SEC is
effects on the crystalline structure of starch samples, with the not reliable for characterising DBs of enzymatically treated
C-type crystalline structure being completely removed after treat- starches, because the presence of a large amount of very short
ment. Moreover, a strong peak at 17.0° is observed for both chains produced by enzymes causes high uncertainty in the calcu-
starches treated with a combination of b-amylase and transglu- lated DB. SEC relies on Mark–Houwink relation to convert elution
cosidase, and was lower than for those treated with b-amylase volume to the chain length (DP) of individual chains. This relation
alone. The relative crystallinity of wrinkled and smooth pea is not accurate for very short chains, as obtained here. However,
starches treated with b-amylase and transglucosidase were lower the DB values for the native starches, where there are not nearly
than that of starches with b-amylase by 0.8% and 5.5%, respec- so many short chains, will be relatively accurate. The DBs of
tively. Ao et al. (2007) reported the situation in normal maize unmodified pea starches are noticeably smaller than those typical
starch was similar to that in pea starch. The level of crystallinity of grains, but the present values are consistent with those calcu-
is largely controlled by the chain lengths of the amylopectin mol- lated from pea-starch data given by other workers.
ecules (Hizukuri, 1985; Witt, Doutch, Gilbert, & Gilbert, 2012). Both starches treated with b-amylase alone had a higher DB
High crystallinity is related to a higher amount of short chains than the untreated starches. This was expected as b-amylase
(Fig. 2B and C) (Bello-Pérez, Roger, Baud, & Colonna, 1998). How- directly decreases the amount of a-(1 ? 4) linkages, without
ever, when chains are so short that they are unable to form a affecting a-(1 ? 6) linkages. Treatment with b-amylase and
double helix, they are unable to contribute to the crystallinity transglucosidase further increased the DB, as expected (Ao et al.,
of the sample. All enzyme-treated starch samples had a weak 2007; Casarrubias-Castillo et al., 2012). The increase of the DB
peak at 19.7°, which indicates an amylose-ethanol complex due was 8.28%, 11.91%, 8.13% and 9.59% for WP-B, WP-BT, SP-B and
to some ethanol remaining after ethanol precipitation (Ao et al., SP-BT, respectively. These data combined with the CLDs suggest
2007). These results confirm the observation that the resulting that the enzymatically-synthesised, highly-branched products are
SDS fraction might consist of less perfect crystallites and amor- good candidates for slowly digestible carbohydrates (Lee et al.,
phous components (Lehmann & Robin, 2007). 2013). A difference was also found between both smooth and
wrinkled pea starches: smooth pea starch had a higher DB than
3.5. Degree of branching of native and enzymatically modified starches wrinkled pea starch after b-amylase treatment. This difference
may due to their different fine structures. It has been reported that
The degrees of branching (DBs) from 1H NMR spectroscopy are the hydrolysis of a-(1 ? 6) linkages is slower than that of
presented in Table 2. The DB values of WP and SP starches are a-(1 ? 4) linkages, and may be the rate-limiting step for glucose
4.76% and 5.66%, respectively. The DB of starch is the inverse of release, with a reduction in digestibility (Kerr, Cleveland, &
the number-average chain length, X  de ðRh Þ, and can be calculated Katzbeck, 1951). The increase in DB through b-amylase and trans-
from Nde(X) with the following equation: glucosidase treatments reduced the overall starch digestion rate.
M. Shi et al. / Food Chemistry 164 (2014) 317–323 323

4. Conclusions Lee, E. (1971). The action of sweet potato b-amylase on glycogen and amylopectin:
Formation of a novel limit dextrin. Archives of Biochemistry and Biophysics,
146(2), 488–492.
This study shows that b-amylase alone and the combination of Lee, B. H., Yan, L., Phillips, R. J., Reuhs, B. L., Jones, K., Rose, D. R., et al. (2013).
b-amylase and transglucosidase treatment can be used to enhance Enzyme-synthesized highly branched maltodextrins have slow glucose
generation at the mucosal alpha-glucosidase level and are slowly digestible
the slow digestion properties of starch. After enzymatic treatment,
in vivo. PLoS One, 8(4), 1–10.
the CLD, crystalline structure and in vitro digestibility of wrinkled Lehmann, U., & Robin, F. (2007). Slowly digestible starch–Its structure and health
and smooth pea starches were significantly changed. For amylo- implications: a review. Trends in Food Science & Technology, 18(7), 346–355.
Miao, M., Xiong, S., Jiang, B., Jiang, H., Cui, S. W., & Zhang, T. (2014). Dual-enzymatic
pectin with a higher proportion of short chains, the amylopectin
modification of maize starch for increasing slow digestion property. Food
structure itself slows enzymatic digestion due to its higher branch Hydrocolloids, 38, 180–185.
density and short chain length, which are more difficult for the Millan-Testa, C., Mendez-Montealvo, M., Ottenhof, M.-A., Farhat, I., & Bello-Pérez, L.
amyloytic enzymes to digest. (2005). Determination of the molecular and structural characteristics of okenia,
mango, and banana starches. Journal of Agricultural and Food Chemistry, 53(3),
495–501.
Acknowledgments O’Dea, K., Snow, P., & Nestel, P. (1981). Rate of starch hydrolysis in vitro as a
predictor of metabolic responses to complex carbohydrate in vivo. The American
Journal of Clinical Nutrition, 34(10), 1991–1993.
The authors gratefully acknowledge financial support from the Ozturk, S., Koksel, H., Kahraman, K., & Ng, P. (2009). Effect of debranching and heat
State Key Program of National Natural Science of China (Grant No. treatments on formation and functional properties of resistant starch from
31230057), and Guangdong Province Program of China (Grant No. high-amylose corn starches. European Food Research and Technology, 229(1),
115–125.
2012B091100291). Careful reading and editing of the manuscript Pongjanta, J., Utaipattanaceep, A., Naivikul, O., & Piyachomkwan, K. (2009).
by Mitchell A Sullivan is greatly appreciated. R.G.G. gratefully Debranching enzyme concentration effected on physicochemical properties
acknowledges the support of an Australian Research Council and alpha-amylase hydrolysis rate of resistant starch type III from amylose rice
starch. Carbohydrate Polymers, 78(1), 5–9.
Discovery Grant, DP130102461.
Ratnayake, W., Hoover, R., Shahidi, F., Perera, C., & Jane, J. (2001). Composition,
molecular structure, and physicochemical properties of starches from four field
References pea (Pisum sativum L.) cultivars. Food Chemistry, 74(2), 189–202.
Ratnayake, W. S., Hoover, R., & Warkentin, T. (2002). Pea starch: Composition,
Ao, Z., Simsek, S., Zhang, G., Venkatachalam, M., Reuhs, B. L., & Hamaker, B. R. (2007). structure and properties—A review. Starch – Stärke, 54(6), 217–234.
Starch with a slow digestion property produced by altering its chain length, Shen, X., Bertoft, E., Zhang, G., & Hamaker, B. R. (2013). Iodine binding to explore the
branch density, and crystalline structure. Journal of Agricultural and Food conformational state of internal chains of amylopectin. Carbohydrate Polymers,
chemistry, 55(11), 4540–4547. 98(1), 778–783.
Backer, D., & Saniez, M.-H. (2005). Soluble highly branched glucose polymers and Shi, M. M., Chen, Y., Yu, S. J., & Gao, Q. Y. (2013). Preparation and properties of RS III
their method of production. US Patent 2005/0142167A1. from waxy maize starch with pullulanase. Food Hydrocolloids, 33(1), 19–25.
Bello-Pérez, L., Roger, P., Baud, B., & Colonna, P. (1998). Macromolecular features of Shin, S. I., Lee, C. J., Kim, D.-I., Lee, H. A., Cheong, J.-J., Chung, K. M., et al. (2007).
starches determined by aqueous high-performance size exclusion Formation, characterization, and glucose response in mice to rice starch with
chromatography. Journal of Cereal Science, 27(3), 267–278. low digestibility produced by citric acid treatment. Journal of Cereal Science,
Beta, T., Corke, H., Rooney, L. W., & Taylor, J. (2001). Starch properties as affected by 45(1), 24–33.
sorghum grain chemistry. Journal of the Science of Food and Agriculture, 81(2), Shin, J.-E., Simsek, S., Reuhs, B. L., & Yao, Y. (2008). Glucose release of water-soluble
245–251. starch-related a-glucans by pancreatin and amyloglucosidase is affected by the
Bijttebier, A., Goesaert, H., & Delcour, J. A. (2010). Hydrolysis of amylopectin by abundance of a-1,6-glucosidic linkages. Journal of Agricultural and Food
amylolytic enzymes: Structural analysis of the residual amylopectin population. chemistry, 56(22), 10879–10886.
Carbohydrate Research, 345(2), 235–242. Silverio, J., Fredriksson, H., Andersson, R., Eliasson, A.-C., & Åman, P. (2000). The
Cai, L. M., & Shi, Y. C. (2010). Structure and digestibility of crystalline short-chain effect of temperature cycling on the amylopectin retrogradation of starches
amylose from debranched waxy wheat, waxy maize, and waxy potato starches. with different amylopectin unit-chain length distribution. Carbohydrate
Carbohydrate Polymers, 79(4), 1117–1123. Polymers, 42(2), 175–184.
Casarrubias-Castillo, M. G., Hamaker, B. R., Rodriguez-Ambriz, S. L., & Bello-Pérez, L. Syahariza, Z., Sar, S., Hasjim, J., Tizzotti, M. J., & Gilbert, R. G. (2013). The importance
A. (2012). Physicochemical, structural, and digestibility properties of enzymatic of amylose and amylopectin fine structures for starch digestibility in cooked
modified plantain and mango starches. Starch – Stärke, 64(4), 304–312. rice grains. Food Chemistry, 136(2), 742–749.
Castro, J. V., Ward, R. M., Gilbert, R. G., & Fitzgerald, M. A. (2005). Measurement of Tizzotti, M. J., Sweedman, M. C., Tang, D., Schaefer, C., & Gilbert, R. G. (2011). New 1H
the molecular weight distribution of debranched starch. Biomacromolecules, NMR procedure for the characterization of native and modified food-grade
6(4), 2260–2270. starches. Journal of Agricultural and Food Chemistry, 59(13), 6913–6919.
Cave, R. A., Seabrook, S. A., Gidley, M. J., & Gilbert, R. G. (2009). Characterization of Tran, T. T., Shelat, K. J., Tang, D., Li, E., Gilbert, R. G., & Hasjim, J. (2011). Milling of rice
starch by size-exclusion chromatography: The limitations imposed by shear grains. The degradation on three structural levels of starch in rice flour can be
scission. Biomacromolecules, 10(8), 2245–2253. independently controlled during grinding. Journal of Agricultural and Food
Derde, L., Gomand, S., Courtin, C., & Delcour, J. (2012). Characterisation of three Chemistry, 59(8), 3964–3973.
starch degrading enzymes: Thermostable b-amylase, maltotetraogenic and Vilaplana, F., & Gilbert, R. G. (2010). Two-dimensional size/branch length
maltogenic a-amylases. Food Chemistry, 135(2), 713–721. distributions of a branched polymer. Macromolecules, 43(17), 7321–7329.
Englyst, H. N., Kingman, S. M., & Cummings, J. H. (1992). Classification and Weill, C., Burch, R., & Van Dyk, J. (1954). An a-amyloglucosidase that produces b-
measurement of nutritionally important starch fractions. European Journal of glucose. Cereal Chemistry, 31, 150–158.
Clinical Nutrition, 46, S33–S50. Witt, T., Doutch, J., Gilbert, E. P., & Gilbert, R. G. (2012). Relations between
Han, J.-A., & BeMiller, J. N. (2007). Preparation and physical characteristics of slowly molecular, crystalline, and lamellar structures of amylopectin.
digesting modified food starches. Carbohydrate Polymers, 67(3), 366–374. Biomacromolecules, 13(12), 4273–4282.
Hasjim, J., Lavau, G. C., Gidley, M. J., & Gilbert, R. G. (2010). In vivo and in vitro starch Wolever, T. M. S. (2003). Slow digestible carbohydrates. Danone Vitapole Nutritopics,
digestion: Are current in vitro techniques adequate? Biomacromolecules, 11(12), 28, 1–17.
3600–3608. Yao, Y., Zhang, J., & Ding, X. (2003). Partial b-amylolysis retards starch
Hizukuri, S. (1985). Relationship between the distribution of the chain length of retrogradation in rice products. Journal of Agricultural and Food Chemistry,
amylopectin and the crystalline structure of starch granules. Carbohydrate 51(14), 4066–4071.
Research, 141(2), 295–306. Zhang, G., & Hamaker, B. R. (1998). Low a-amylase starch digestibility of cooked
Jane, J.-L., & Robyt, J. F. (1984). Structure studies of amylose-V complexes and retro- sorghum flours and the effect of protein 1. Cereal Chemistry, 75(5), 710–713.
graded amylose by action of alpha amylases, and a new method for preparing Zhang, G. Y., Sofyan, M., & Hamaker, B. R. (2008). Slowly digestible state of starch:
amylodextrins. Carbohydrate Research, 132(1), 105–118. Mechanism of slow digestion property of gelatinized maize starch. Journal of
Kerr, R., Cleveland, F., & Katzbeck, W. (1951). The action of amylo-glucosidase on Agricultural and Food Chemistry, 56(12), 4695–4702.
amylose and amylopectin. Journal of the American Chemical Society, 73(8), Zheng, J., Li, Q., Hu, A., Yang, L., Lu, J., Zhang, X., et al. (2013). Dual-frequency
3916–3921. ultrasound effect on structure and properties of sweet potato starch. Starch –
Kumari, M., Urooj, A., & Prasad, N. N. (2007). Effect of storage on resistant starch and Stärke, 65(7–8), 621–627.
amylose content of cereal-pulse based ready-to-eat commercial products. Food
Chemistry, 102(4), 1425–1430.

You might also like