You are on page 1of 16

Food Research International 43 (2010) 1959–1974

Contents lists available at ScienceDirect

Food Research International


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / f o o d r e s

Review

Resistant starch: A review of analytical protocols for determining resistant starch and
of factors affecting the resistant starch content of foods
A. Perera a,⁎, V. Meda a, R.T. Tyler b
a
Department of Agricultural and Bioresource Engineering, University of Saskatchewan, Canada
b
Department of Food and Bioproduct Sciences, University of Saskatchewan, Canada

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

Article history: Resistant starch (RS) has drawn considerable attention over the last two decades due to its demonstrated
Received 8 March 2010 and putative positive impacts on health. This has resulted in the development of a variety of analytical
Accepted 8 June 2010 protocols for its determination, hence a comprehensive review of methodologies for analyzing resistant
starch is warranted. Adding to the complexity, the RS contents of starchy materials vary widely and RS levels
Keywords:
are impacted by processing. This review compares commonly used methods for the determination of RS,
Resistant starch
Starch digestibility
briefly reviews the process of starch digestion in the human, and discusses the complications and challenges
Starch determination associated with determining RS in foods.
Starch hydrolysis © 2010 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1959
2. Starch digestion in the human . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1960
3. Demonstrated and putative health benefits of RS consumption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1960
4. The rat as a model for starch digestion in humans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1961
5. Methods of RS determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1962
6. Effect of starch composition on RS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1967
7. Effect of processing on RS levels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1969
8. Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1972
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1972

1. Introduction susceptible to hydrolysis by amyloglucosidase is referred to as resistant


starch (RS). This definition of RS was modified later to include ‘the sum
Polysaccharides in foods consist of starch and plant cell wall of starch and starch-degradation products that, on average, reach the
constituents. On the basis of its digestibility, starch was classified by human large intestine’ (Englyst, Kingman, Hudson, & Cummings, 1996).
Englyst and Cummings (1987a) into three groups, namely readily Partially resistant starch and RS as defined by Englyst and Cummings
digestible starch, partially resistant starch and resistant starch. Accord- (1987a) were subsequently reclassified as RS type 1, RS type 2 and RS
ing to this classification, starch that is not accessible to digestive type 3 (Englyst, Kingman, & Cummings, 1992; Sajilatha, Singhal, &
enzymes, such as that protected by hard to digest coatings in whole and Kulkarni, 2006). Enzyme inaccessible starch resulting from structural
coarsely-milled grains, starch in raw banana and potato exhibiting B- or rigidity is defined as RS1, and it is calculated as the difference between
C-type X-ray diffraction patterns, and retrograded amylopectin, is the enzyme-extractable glucose contents of a food determined with and
considered to be partially resistant starch. Retrograded amylose that can without an homogenization treatment prior to analysis (Sajilatha et al.,
be solubilized in potassium hydroxide and which subsequently becomes 2006). Englyst et al. (1992) reported RS1 as the difference between the
glucose contents of two samples of a food, one after a milling/
homogenization treatment to obtain 0.2-mm-sized particles and the
other after a mincing treatment, and subsequent hydrolysis with
⁎ Corresponding author. Tel.: + 1 306 966 2454. pancreatin and amyloglucosidase for 120 min. It seems that Englyst et al.
E-mail address: anula.perera@usask.ca (A. Perera). (1992) assumed that comminution (milling or homogenization) of

0963-9969/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.foodres.2010.06.003
1960 A. Perera et al. / Food Research International 43 (2010) 1959–1974

foods would yield a consistency and particle size similar to that of the epithelial cells of the small intestine, such as enterokinase, sucrase,
mastication. However, this may not be true for all foods and RS1 values maltase and lactase. These enzymes are responsible for the hydrolysis of
determined for comminuted samples may differ from those determined carbohydrates into monosaccharides. In addition, proteolytic enzymes
for otherwise identical masticated samples. For example, Åkerberg, such as trypsin, chymotrypsin, carboxypeptidase and elastase, lipolytic
Liljeberg, Granfeldt, Drews, and Björck (1998) evaluated RS in selected enzymes such as lipase, phospholipase and cholesterol esterase, and the
foods including RS1-rich raw banana. Foods were masticated (15 chews DNA and RNA hydrolyzing enzymes ribonuclease and deoxyribonucle-
in 15 s) prior to analysis. Åkerberg, Liljeberg, Granfeldt, et al. (1998) ase are secreted by the pancreas and mixed with the chyme (Cichoke,
reported RS values (72 g/100 g starch) that were higher than those 1998; DeSesso & Jacobson, 2001). Chyme that is not hydrolyzed into
reported (53.6 g/100 g starch) by Englyst and Cummings (1986). The absorbable molecules in the small intestine, e.g. starch which has
latter group employed the analytical method of Englyst, Wiggins, and escaped digestion, along with non-starch polysaccharides, moves into
Cummings (1982) where samples were ball-milled. Starch granules in the colon and undergoes bacterial fermentation. The resulting short-
raw foods which are resistant to digestion comprise RS2. Analytically, chain fatty acids are absorbed in the colon and provide 5–10% of the
RS2 is measured as the difference in the amount of glucose liberated by energy requirement of humans (Nordgaard & Mortensen, 1995). It has
pancreatin and amyloglucosidase, after 120 min of hydrolysis, from two been shown that of the total quantity of starch consumed by ileostomy
samples of the same food (having the same particle size), one having subjects, approximately 3% of freshly cooked starch and approximately
been subjected to a prior boiling treatment and one not. Retrograded 12% of cooked and cooled starch escape digestion in the small intestine
starch formed during the cooling of gelatinized starch is termed RS3 and can be recovered in the ileum. All of the RS and the non-starch
(Englyst et al., 1992). Based on the definitions above, it is clear that the polysaccharides in the diet were found in the ileum effluent of tested
RS1, RS2 and RS3 contents determined for a particular food sample may subjects (Englyst & Cummings, 1985, 1987b). Approximately 80–90% of
be affected by differences in sample processing (mastication, homog- the RS that passes into the human colon is fermented and the remainder
enization, grinding, etc.). Cross-linking of starches from wheat, corn, is released with the feces. Individual variations in the microbial
rice, potato, tapioca, mung bean and oat, using sodium trimetapho- population of the colon can influence significantly the degree of RS
sphate (STMP), sodium tripolyphospahate (STPP), epichlorohydrin or fermentation. The presence of RS in the diet decreases considerably the
phosphoryl chloride (POCl3), produced starch resistant to α-amylase bacterial fermentation of non-starch polysaccharides, which is useful as
hydrolysis, which was termed RS4 (Seib & Woo, 1999). These authors non-starch polysaccharides increase the bulk of the excreta and improve
demonstrated that RS levels in wheat starch cross-linked with 12% its water holding capacity (Cummings, Beatty, Kingman, Bingham, &
STMP/STPP, 2% POCl3 and 2% epichlorohydrin were 75.6, 85.6 and Englyst, 1996).
75.8 g/100 g starch, respectively. Other chemically-modified starches,
e.g. starch etherified with propylene oxide or cross-linked with 3. Demonstrated and putative health benefits of RS consumption
phosphorous oxychloride, acid-modified starch and starch oxidized
with sodium hypochlorite, also have been reported to be resistant to in RS is recognized as a significant contributor to gastrointestinal
vitro hydrolysis by α-amylase and amyloglucosidase (Wolf, Bauer, & health. Undigested starch is fermented by microorganisms in the large
Fahey, 1999). The classification of starch has been expanded to include bowel to produce short chain fatty acids, including acetate, propionate
rapidly digestible starch (RDS) and slowly digestible starch (SDS) in and butyrate (Bird, Brown, & Topping, 2000; Brouns, Kettlitz, & Arrigoni,
addition to RS. Starch that is hydrolyzed in vitro within 20 min with α- 2002; Wong, de Souza, Kendall, Emam, & Jenkins, 2006) and isobutyrate,
amylase and amyloglucosidase is designated RDS (Englyst, Veenstra, & valerate and isovalerate (Hylla et al., 1998). A high-amylose corn starch
Hudson, 1996). The category of SDS includes enzyme inaccessible starch diet containing 46–50% of RS2 significantly increased the colonic
and raw starch that is fully hydrolyzed in vitro during prolonged production of acetate, propionate and butyrate in rats. This diet also
incubation (20–120 min) (Englyst et al., 1992; Sajilatha et al., 2006). increased apoptosis in the colon subsequent to externally induced gene
Starch that is not hydrolysed after 120 min incubation with pancreatin damage, which suggests an ability of RS to control the initiation of
and amyloglucosidase is termed RS (Englyst et al., 1992). colonic cancer (Leu, Hu, Brown, & Young, 2009). Butyrate was reported
to increase the level of glutathione, an antioxidant, in colonic mucosa,
2. Starch digestion in the human improving colonic resistance to toxic agents in the diet (Hamer et al.,
2009). A significant decrease in bile acids (cholic acid, chenodeoxycholic
As described above, the concept of RS is based on the inability of acid, deoxycholic acid, lithocholic acid, isolithocholic acid and urso-
digestive enzymes to hydrolyze some physical and chemical forms of deoxycholic acid) and total neutral sterols (coprostanol, cholesterol, 4-
starch in foods in vivo or in vitro. Therefore, it is useful to review the cholesten-3-one, campesterol, stigmasterol and β-sitosterol) in the fecal
process of starch digestion in humans. Fig. 1 summarizes the major matter of healthy adults was reported after consuming experimental
organs that participate in the digestion of foods and the enzymes foods rich in RS for 4 weeks (Hylla et al., 1998). As the fecal matter of
secreted. Salivary glands that open into the oral cavity produce several colon cancer patients carries more bile acids and neutral steroids than
enzymes that initiate carbohydrate digestion, e.g. α-amylase, glycosi- does that of healthy controls (Reddy, Ernst, & Wynderm, 1977), it is
dases, glucose oxidase, lactate dehydrogenase and β-glucuronidase believed that RS has a positive effect on the control and prevention of
(Makinen, 1989). Food masticated and partially hydrolyzed in the colon cancer. Significant expression of a ‘cell cycle regulatory and cell
mouth passes through the pharynx and then through the esophagus and proliferation gene’ was observed in colorectal cancer patients after their
into the stomach. Submucosal glands of the esophagus secrete consumption of 40 g/day of RS for 2–4 weeks. The expression of the
bicarbonates and mucin. This is considered a mechanism to protect gene was inhibited in control subjects (Dronamraju, Coxhead, Kelly, &
the esophagus from HCl refluxed from the stomach (Abdulnour- Mathers, 2007).
Nakhoul et al., 2005). In the stomach, food is mixed with pepsin, gastric Consumption of RS has other benefits. A diet rich in dietary fibre and
lipase and HCl, which activates pepsin. No significant digestion of starch RS was reported to significantly increase the weight of the pancreas,
takes place in the stomach but contraction of muscles facilitates liver and intestine, and the length of the intestine, in rats (Correa,
comminution of foods (Correa, 1988). Food that is passed into the Angelina, Reis, Maria, & de Oliveira Costa, 2009). RS was reported to
small intestine is termed ‘chyme’ and there it is mixed with bicarbonates decrease the postprandial blood glucose level considerably within the
and mucin secreted by the Brunner's glands of the small intestine first 120 min (Kendall et al., 2008). Compared to a diet containing
(DeSesso & Jacobson, 2001). In the duodenum, the first part of the small autoclaved and cooled starch from waxy barley (which would not
intestine, chyme is blended with bile secreted by the gallbladder, undergo significant retrogradation), significantly lower levels of blood
pancreatic α-amylase, and amylolytic and other enzymes secreted by glucose (within the first 120 min) were reported after feeding
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1961

Fig. 1. Enzymes secreted by major organs participating in the digestion of foods in humans. The digestive system, an extension of the environment into the body, also secretes mucus,
acid, bicarbonate and salts to facilitate hydrolysis of foods and absorption of nutrients essential for life.

retrograded high-amylose barley (∼18% RS) to rats (Xue, Newman, & 2007). Consumption of common bean (Phaseolus vulgaris L.) rich in RS
Newman, 1996). RS3 in the diet has a laxative effect. A supplement of (12–13 g/100 g raw starch) and soluble dietary fibre introduced
25 g/day in healthy test subjects increased the daily fecal output above flatulence, bloating, abdominal noise and pain in test subjects, but
the normal level with minimal gastrointestinal discomfort (Maki et al., symptoms significantly decreased when fermented beans were con-
2009). A RS-rich green banana diet had beneficial effects on controlling sumed (Granito, Michel, Frías, Champ, & Guerra, 2005). In a review,
shigellosis in child patients by decreasing stool volume, number of stools Grabitske and Slavin (2009) explained that diarrhea is an extreme
per day and red blood cells in stools, and increasing levels of acetate, outcome of ingesting carbohydrates with very low digestibility,
propionate and butyrate in feces. Despite the expected negative impact including RS, when the capacity of colonic fermentation is exceeded
of antibacterial treatments (ciprofloxacin) on short chain fatty acid by consumption. They also noted that adverse effects are more
production in the colon, acetate, propionate and butyrate production in prominent with RS3 than with RS2, and that the severity of symptoms
child patients increased on the fifth day of feeding green banana differed with personal attributes such as age, weight, gender, genetics,
(Rabbani et al., 2009). A RS-supplemented diet significantly increased health and the composition of the gut microflora. Overall, the benefits of
populations of lactobacilli, bifidobacteria, staphylococci and streptococ- RS consumption were considered to outweigh any gastrointestinal
ci, decreased the enterobacteria population, and altered microbial discomfort.
enzyme metabolism in the colon of experimental rats (Silvi, Rumney,
Cresci, & Rowland, 1999). 4. The rat as a model for starch digestion in humans
For a balanced view of the effect on RS on health, it is important to
note that the consumption of high amounts of RS may have some As described above, animal models have been used in some RS
negative effects on gastrointestinal performance. These include bloat- research studies and thus it is important to discuss briefly the
ing, borborygmi (noise due to gas movement in the intestine), dependability and/or applicability of the results obtained. Animal
flatulence, colic and watery feces (Storey, Lee, Bornet, & Brouns, models used in biomedical research should be either analogous or
1962 A. Perera et al. / Food Research International 43 (2010) 1959–1974

homologous to humans considering the anatomy and physiology of current methods of analysis of RS and difficulties related to comparing
the organ or system under investigation. Other criteria include, but data generated by different methods.
are not limited to, the selected animal should be easily available in In vitro methods for the analysis of RS in foods are based
substantial numbers, should be easy to handle and manageable with conceptually on the gastrointestinal digestion of starch in foods.
the resources available, and the life span should be sufficiently long to Enzyme hydrolysis is a common feature of all methods, combined
carry out the research. In keeping with these criteria, omnivores such with chemical hydrolysis or gravimetric isolation. Differences in
as the rat and the pig are used to simulate the gastrointestinal tract sample preparation are significant and range from mechanical
and liver functions of the human (Chow, 2008). In a comprehensive methods (milling, grinding and homogenization) to mastication.
review, DeSesso and Jacobson (2001) compared the similarities and The type of RS quantified is dependant on the protocol. Most methods
differences between the gastrointestinal tracts of the human and the are focused on the determination of total RS, but specific methods
rat. Some significant differences are listed here. The human pharynx is have been developed to quantify RS1, RS2 and RS3.
a single tube that is the passageway for both food and air, whereas the Methods for the determination of RS proposed by Englyst et al.
rat pharynx is separated into two compartments for respiratory and (1982) and Englyst et al. (1992) are the basis for this discussion.
digestive purposes. The human stomach is a single chamber where Significant modifications proposed by other authors are compared and
food is mixed with enzymes and acid. The rat stomach, although a contrasted. As proposed by Englyst et al. (1982), RS and other non-
single chamber, is functionally divided into two areas. The forest- starch polysaccharides are first separated from the enzyme-hydrolyz-
omach is the site of bacterial digestion, whereas the glandular able starch. Then, RS is solubilized in alkali and thereby separated from
stomach secretes enzymes and acid. Bacterial digestion is absent in the other enzyme resistant polysaccharides. Even though Englyst et al.
the human stomach. Although the rat is used to imitate digestion in (1982) labelled this fraction as RS, the analytical protocol used reveals
the human, there is the likelihood of species differences in enzyme that it measures RS3 only. Briefly, in this method (Fig. 2) 100–200 mg of
activity. The small intestine has three functionally different regions, homogenized wet sample or ground dry sample is mixed with sodium
the duodenum, the jejunum and the ileum. Compared to the human, acetate buffer at pH 5.4 and heated for 1 h at 100 °C to gelatinize starch.
where the gallbladder secretes bile, the rat liver secretes twice as Having included homogenization and boiling steps, this protocol
much as bile into the duodenum per kg of body weight per day. eliminates the contribution of RS1 and RS2 to the final RS value. An
Nutrient absorption takes place in the duodenum and the forejejunum enzyme mixture containing α-amylase, pullulanase and amyloglucosi-
of both rats and humans through diffusion, active transport and dase is added and hydrolysis is carried out for 16 h at 40 °C. Absolute
solvent drag mechanisms. Compared to the total length of the ethanol is added to terminate enzyme activity and to precipitate non-
digestive tract, the jejunum of the rat is proportionally longer than hydrolyzed starch. The pellet is collected after centrifugation and
that of the human. The jejunum of the rat represents 90% of the length washed twice with 80% ethanol. The residue is dried with acetone and
of the small intestine, and that of the human, 38%. The transit time of then treated with 2 M potassium hydroxide for 30 min at room
food through the rat small intestine is slower than that in the human. temperature to solubilize retrograded starch (RS3). An aliquot of alkali
Despite the human small intestine being 5.5 times longer than that of digest is mixed with 2 M acetic acid and amyloglucosidase and the
the rat, the transit time of food through the small intestine of the contents are incubated at 65 °C for 1 h. After cooling and centrifuging,
human and the rat is similar, 3–4 h. The surface area of the human neutral sugars (glucose, galactose, mannose, xylose and arabinose) in
small intestine is 200 times greater than that of the rat, which is the supernatant are analyzed by gas–liquid chromatography (GLC). The
significant given the difference in their lengths. Water and nutrients quantity of glucose detected by GLC represents the amount of RS (RS3)
formed through bacterial metabolism are absorbed in the large in the sample.
intestine. The human large intestine is divided into the cecum, The above procedure was modified later by the authors (Englyst
ascending colon, transverse colon, descending colon, sigmoid colon, et al., 1992) to measure total glucose, free glucose, total RS, and RS1, RS2
rectum and anus. The sigmoid colon is absent in the rat. The cecum is and RS3 as separate fractions. The determination of total RS is carried
responsible for most of the microbial digestion of food in the rat and out in step-wise fashion. Total glucose (TG) and free glucose (FG)
represents up to 26% of the length of the large intestine. In contrast, contents are determined, as explained in Fig. 3.1, in order that total
the human cecum is only 5% of the length of the large intestine. In starch (TS) may be calculated as TS= (TG − FG) × 0.9. In this protocol
addition, the types of microorganisms naturally occurring in the for TS determination, RS1 and RS2 are eliminated by mincing/milling/
digestive systems of the human and the rat may be different (DeSesso homogenizing the food sample and subsequently heating the sample at
& Jacobson, 2001). The differences described above indicate that the 100 °C for 30 min, respectively. Any retrograded starch that contributes
rat model may not provide reliable data on RS digestion in humans, to RS3 is eliminated by treating the sample with 7 M potassium
but it may be useful in understanding the fate of RS within the hydroxide at 0 °C. The rapidly digestible starch (RDS) and slowly
digestive tract. digestible starch (SDS) contents of the sample are determined
separately and total RS is calculated as RS = TS − (RDS + SDS)
(Fig. 3.2). In this method, 0.8–4.0 g of sample is minced only to
5. Methods of RS determination minimize structural changes. The low viscosity of samples leads to the
release of considerably higher amounts of glucose during enzyme
Determination of RS in food ingredients and processed foods has hydrolysis, especially in the absence of other polysaccharides, than is
become vital to the provision of nutritional information to consumers the case with the in vivo digestion of starch. Thus, guar gum is added to
and others. To enable effective use of the research output on RS for food increase the viscosity. Glass beads are added to samples to ensure
processing and nutritional applications, analytical procedures for the proper mixing and to break cell walls during enzyme hydrolysis. For the
determination of RS need to be compared. At present, significant determination of RDS and SDS, samples are equilibrated with 20 mL of
differences exist among procedures with respect to sample preparation, acetate buffer at 37 °C and then starch is hydrolyzed with an enzyme
the enzymes used, and the establishment of experimental conditions mix containing pancreatin as a source of α-amylase (30,000 BPU/g),
that mimic gastrointestinal digestion of starch. Ongoing improvements amyloglucosidase (400 AGU/mL), and invertase. (3000 EU/mL). Englyst
in analytical methodology are essential, but frequent modifications in et al. (1992) proposed that enzyme hydrolysis be carried out at 37 °C
protocols reduce the availability of comparable data to assess the instead of at 40 °C, as previously used in the Englyst et al. (1982)
efficacy of methods and the nutritional quality of foods. Added to this, protocol, to better mimic human physiological conditions of digestion. A
food analytes used in the various procedures have differed in their significant reduction in the duration of enzyme digestion is used in the
genetic origin, composition and processing history. This paper reviews Englyst et al. (1992) protocol. Instead of 16 h of sample digestion with
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1963

Fig. 2. Main steps in the Englyst et al. (1982) protocol for the determination of enzyme resistant starch (RS3). The presence of an enzyme resistant fraction of starch in foods was first
observed by the authors during the measurement of non-starch polysaccharides in foods.

pancreatin, as per Englyst et al. (1982), the Englyst et al. (1992) protocol after 120 min (SDS) is determined using the procedure outlined in
uses sequential removal of aliquots of enzyme digest at 20 min (as the Fig. 3.2. RS2 is calculated as RS2 = [G120(C) − G120(D)] × 0.9.
rapid release of sugars ended at 20 min of digestion with pancreatin) Determination of RS3 is limited to samples containing retrograded
and at 120 min from the beginning of the digestion. The rationale to starch (Englyst et al., 1992) and the protocol is similar to that of
terminate enzyme hydrolysis at 120 min is that the release of glucose Englyst et al. (1982) in many ways. In contrast to the protocol for the
reached a plateau at 120 min. The glucose content of the 20-min determination of RDS and SDS for the estimation of total RS (Fig. 3.2),
enzyme digest (G20 fraction) represents RDS and is calculated as RDS= where α-amylase, amyloglucosidase and invertase are used, the
(G20 − FG) × 0.9. The difference in glucose contents between the 120- protocol for RS3 determination uses α-amylase and pullulanase at
min and 20-min digests represents SDS and is calculated as SDS = (G120 40 °C. After removing the enzyme-hydrolyzable starch from the
− G20 − FG)× 0.9. Instead of directly measuring glucose originating sample, retrograded starch is solubilized in 4 M potassium hydroxide
from RS as in Englyst et al. (1982), this method indirectly estimates total at room temperature, whereas Englyst et al. (1982) used 2 M
RS as RS = TS− (RDS + SDS). As 37 °C is the highest temperature at potassium hydroxide. However, this concentration is lower than
which a sample is treated in this protocol, any RS2 in the food sample that used (7 M) in the protocol for the determination of TG outlined in
would not be hydrolyzed by α-amylase. RS1 also is preserved in the Fig. 3.1. The temperature at which retrograded starch is solubilized in
unhydrolyzed fraction as the sample is only minced and no milling or potassium hydroxide also is different in the two procedures; 0 °C for
homogenization is employed. As a result, this indirect measurement of TG and room temperature for RS3 determination (Englyst et al., 1992).
RS includes any RS1, RS2 and RS3 in the sample. Goñi, García-Diza, Mańasb, and Saura-Calixto (1996) cautioned
Englyst et al. (1992) devised protocols for determination of RS1, about the possible effect of sample preparation in RS protocols as
RS2 and RS3. By definition, RS1 is the fraction of starch that is drying, cooling and conditions of storage can alter the level of RS in
inaccessible to enzymes due to physical impediments, e.g. large-sized foods. Their method of RS determination is summarized in Fig. 4. In
particles and cell walls. The analytical procedure for RS1 determina- this procedure, a 100-mg sample is milled or homogenized before use
tion (Fig. 3.3a) differs from that of Englyst et al. (1982) (Fig. 2) with and by doing so RS1 is eliminated. Sample pH is decreased to 1.5 with
respect to the method of reduction of particle size. One portion of the HCl–KCl buffer to simulate gastric pH. As a major modification to the
sample (A) is minced once and the other portion (B) is either milled or procedure of Englyst et al. (1992), hydrolysis with pepsin at 40 °C for
homogenized, as appropriate, to obtain particles b 0.2 mm in 1 h is introduced. Sample pH is adjusted to 6.9 with tris–maleate
diameter. Then, using the protocol for the determination of RDS and buffer to simulate conditions in the small intestine. Instead of using a
SDS (Fig. 3.2), glucose released after 120 min of enzyme hydrolysis cocktail of enzymes as in Englyst et al. (1992), samples are hydrolysed
(G120) is determined for each portion. RS1 is calculated as RS1 = [G120 with α-amylase only for 16 h at 37 °C. As this method is focused on
(B) − G120(A)] × 0.9. Starch in raw foods and foods cooked with little the determination of RS (not SDS and RDS), products from enzyme
water (preventing complete gelatinization of starch) contribute to hydrolysis are discarded at this point. As 40 °C is the highest
RS2. The method for the determination of RS2 (Fig. 3.3b) compares temperature to which the sample is exposed, any RS2 in raw foods
the amount of glucose released enzymatically from a cooked sample is included in the unhydrolyzed fraction. The pellet is dispersed in 4 M
to that from a raw sample (Englyst et al., 1992). One portion (C) of a potassium hydroxide at room temperature for 30 min to solubilize
raw food sample is heated in 0.1 M sodium acetate buffer, pH 5.2, in a RS3. Dextrin in the alkali digest is hydrolyzed to glucose with
boiling water bath for 30 min, while another portion (D) is used amyloglucosidase at 60 °C for 45 min. Glucose content is determined
directly without a heat treatment. The samples are hydrolysed with using a colorimetric assay. Based on the analytical procedure of Goñi
amylase, amyloglucosidase and invertase and the glucose released et al. (1996), the presumed definition for RS is starch that is not
1964 A. Perera et al. / Food Research International 43 (2010) 1959–1974

Fig. 3.1. Summary of the protocol for the determination of total starch (TS) as per Englyst et al. (1992). The free glucose (FG) and total glucose (TG) contents are measured to estimate
the total starch content of the sample.

hydrolyzed by α-amylase within 16 h at body temperature, which is employs a mixture of α-amylase (3 Ceralpha Units/mg) and
in contrast to the Englyst et al. (1992) definition where RS is starch amyloglucosidase (3300 U/mL) to hydrolyze starch in raw or
that is not hydrolyzed by pancreatin, amyloglucosidase and invertase processed food samples. Hydrolysis is carried out for 16 h at 37 °C
within 120 min. as proposed by Englyst et al. (1982). As a result, RS2 in raw foods
Other modifications to the Englyst et al. (1992) method have been would not be hydrolyzed during this procedure. After inactivation of
proposed but the essence of the method (to remove hydrolyzable starch enzymes with 99% ethanol, glucose in the sample is removed by two
with enzymes) remained unchanged. Chung, Lim, and Lim (2006) washings with 50% ethanol. Similar to Englyst et al. (1982), the pellet
modified the Englyst et al. (1992) protocol to eliminate redundancy and collected from digestion is treated with 2 M potassium hydroxide to
to introduce a reduction in sample size (to 0.5 g). Instead of adding guar extract RS3 from the fibre-rich matrix. Dextrins thus produced are
gum separately to each sample tube with subsequent dissolution in hydrolyzed to glucose with amyloglucosidase. The incubation time
acetate buffer, Chung et al. (2006) prepared a solution of guar gum in with amyloglucosidase is reduced to 30 min at 50 °C (Megazyme,
HCl (to increase its solubility) and then added aliquots to each sample. 2008) from 1 h at 65 °C (Englyst et al., 1982). The Megazyme®
The volume of sodium acetate is decreased to 5 mL by increasing its method uses the glucose oxidase-peroxidase colorimetric assay to
molar concentration to 0.5 M (20 mL of 0.1 M sodium acetate is used in determine the glucose concentration in the final hydrolysate, as did
the Englyst et al. (1992) protocol). Chung et al. (2006) deleted invertase Englyst et al. (1992). The Megazyme® protocol for RS does not enable
from the enzyme mix. The digests collected (0.5 mL) after 20 and the determination of SDS and RDS.
120 min were mixed with 4 mL of 80% ethanol, which replaced the Eerlingen, Crombez, and Delcour (1993) proposed enzymatic-
20 mL of 66% ethanol in the Englyst et al. (1992) protocol. RS was gravimetric isolation of RS from wheat starch pre-gelatinized at
calculated as the difference between total starch and starch not 121 °C and subsequently stored at 0, 68 and 100 °C. In this method,
hydrolysed at 120 min as in Englyst et al. (1992). samples are placed in boiling phosphate buffer, pH 6, and hydrolyzed
The Megazyme® kit for RS determination is widely used in with Termamyl® (heat stable α-amylase) at 100 °C for 30 min. After
analytical laboratories and is the basis of both AOAC method 2002.02 cooling to room temperature, the pH is adjusted to 4.5 with 2%
and AACC Method 32–40 (Megazyme, 2008). The procedure is phosphoric acid. Samples are hydrolyzed with amyloglucosidase at
summarized in Fig. 5. This method, in many ways, is similar to that 60 °C for 30 min and then centrifuged. The pellet is washed with
of Englyst et al. (1982). Samples are ground to pass a 1-mm sieve deionized water, dispersed in phosphate buffer at pH 7.5 and
which generates a coarse meal, thus RS1 is accounted for in this incubated with protease at 42 °C for 4 h. The residue after digestion
method. The Megazyme® protocol eliminates the initial boiling of is collected by filtration and oven-dried at 80 °C. The weight of the
samples in acetate buffer and the use of pullulanase. Instead, it enzyme-insoluble residue represents the RS content of the starch
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1965

Fig. 3.2. Englyst et al. (1992) protocol for the determination of rapidly digestible starch (RDS) and slowly digestible starch (SDS) as the preliminary step in the estimation of RS in a
food sample. The procedure for the estimation of total starch (TS) is explained in Fig. 3.1.

sample. Faraj, Vasanthan, and Hoover (2004) also isolated RS Sample pH is then adjusted to 2 and the sample is hydrolysed with
gravimetrically, but introduced a preliminary hydrolysis with lichi- pepsin for 30 min at 37 °C. After neutralizing the pH with sodium
nase and β-glucosidase. Then, samples are treated with a fungal hydroxide, pancreatin and amyloglucosidase are added to samples in
protease at 42 °C, Termamyl® at 100 °C and amyloglucosidase at acetate buffer at pH 5 and incubated for 6 h at 37 °C. According to the
50 °C, respectively. RS is collected by drying the enzyme resistant authors, this reduction in hydrolysis time (6 h instead of 16 h) is
residue. It is important to note that the gravimetric separation of RS is intended to represent better the movement of food through the small
acceptable only for analytes free of non-starch polysaccharides. intestine. The Muir and O'Dea (1992) protocol significantly deviates
Other modifications have been made to the sample preparation from the approaches already discussed as the pellet collected after
step of RS determination with the intent of more closely simulating in pancreatin hydrolysis (at 37 °C) is further treated with thermostable
vivo digestion. Åkerberg, Liljeberg, Granfeldt, et al. (1998) and α-amylase (Termamyl) at 100 °C, presumably to hydrolyse RS2. The
Åkerberg, Liljeberg, and Björck (1998) employed chewing of food pellet recovered after Termamyl hydrolysis is boiled with dimethyl
samples to initiate hydrolysis by salivary amylase. According to these sulfoxide (DMSO) for 1 h. The supernatant from Termamyl hydrolysis
authors, the number of times a food sample is chewed has a significant (RS2) and the digest from DMSO treatment (RS3) are mixed and the
effect on the end result. Briefly, the chewed food sample (15 chews in glucose content of the mixture is determined. Introduction of DMSO
15 s) and subsequent mouth wash are expectorated into a beaker in lieu of potassium hydroxide for solubilization of RS3 is a novel
containing pepsin in Na/K phosphate buffer. Then, the pH is adjusted feature of the Muir and O'Dea (1992) protocol. Saura-Calixto, Goňi,
to 1.5 with HCl and the sample is incubated for 30 min at 37 °C to Bravo, and Maňas (1993) determined RS in dietary fibre. Instead of
simulate gastric conditions. Sample pH then is adjusted to 5 with using pancreatin as the source of α-amylase, they used heat stable α-
sodium acetate buffer (0.5 M, pH 5) and sodium hydroxide. amylase at 100 °C to eliminate traces of starch. The pellet from
Subsequently, starch hydrolysis is carried out with a mixture enzyme digestion is sequentially hydrolyzed using a protease and
containing MgCl2, CaCl2, 2-propanol, pancreatin and amyloglucosi- amyloglucosidase for 35 min at 60 °C. The residue is collected and
dase for 16 h at 40 °C. Then, 95% ethanol at 60 °C is added in excess, dehydrated with ethanol and acetone, and subsequently dispersed in
leading to the formation of a precipitate at room temperature. The 2 M potassium hydroxide at room temperature. After adjusting the pH
precipitate, collected by filtration, is dehydrated with 95% and 99% to 4.75 with acetate buffer, dextrins in the solution are hydrolyzed
ethanol and solubilized in 2 M potassium hydroxide. The total starch with amyloglucosidase at 60 °C for 35 min. The glucose oxidase-
content of the alkali digest is determined and expressed as RS. Muir peroxidase system is used to measure glucose in the supernatant
and O'Dea (1992) also employed chewing for sample preparation. collected after centrifugation.
1966 A. Perera et al. / Food Research International 43 (2010) 1959–1974
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1967

Fig. 4. Goñi et al. (1996) protocol for the determination of resistant starch (RS) in foods.

It is important to note that very little information is available on an aliquot for color development. The availability of RS controls from
the in vivo digestibility of foods for which in vitro RS data is available. Megazyme® is useful for understanding experimental errors.
From a nutritional point of view, this is a drawback. Grinding and It is reasonable to expect that variations in analytical procedures,
homogenization would certainly produce food particles much smaller including differences in the enzymes used and in their concentration
than those resulting from chewing. Englyst et al. (1992) compared and sequence of application, and dissimilarities in the conditions of
chewing of samples against mincing and reported that differences in experiments, would significantly alter the levels of RS detected in
results were more related to the source of starch than to the method similar foods. RS values of selected foods as reported by some authors
of mincing. Individuals differ in their chewing habits and may not (Åkerberg, Liljeberg, & Björck, 1998; Chung et al., 2006; Englyst et al.,
chew the same number of times as in an experimental procedure. 1992; Faraj et al., 2004; Muir & O'Dea, 1992; Szczodrak & Pomeranz,
Therefore, available RS1 may vary from person to person as its level 1991; Tribess et al., 2009) are presented in Table 1. Englyst and Hudson
depends on particle size. (1996) reported RS values of a large number of different food items, but
It is important to consider possible variability in RS data associated the non-availability of information on handling and storage pre- and
with the technique employed to sample the enzyme hydrolysate, post-processing prevent any useful comparison. The analysis of resistant
because the colorimetric glucose assay is sensitive to inaccuracies in starch in Englyst et al. (1982) is limited to a few sources including white
sample volume. Englyst et al. (1992) proposed drawing of 0.5-mL bread, whole wheat bread and corn flakes (0.8%, 0.8% and 2.9% RS,
volumes of the hydrolyzate after 20 and 120 min to determine RDS respectively, on a dry weight basis) because the main focus of this study
and SDS, respectively. Then, RS is computed as the difference between was to propose a method to determine non-starch polysaccharides in
total starch and enzyme-hydrolyzed starch. Chung, Liu, and Hoover foods, not RS.
(2009) reduced the volume of sample drawn to 0.1 mL. It has been
observed that undigested food particles interfere with pipetting 6. Effect of starch composition on RS
(especially at the 20-min sampling point) more frequently than one
might expect. This would significantly influence the reproducibility of Amylose and amylopectin are the two major polysaccharides in
the colorimetric assay of glucose. The advantage of using gas–liquid starch, both having glucose as the fundamental building block. Linear
chromatography to measure glucose is such errors can be corrected by chains of glucose in both amylose and amylopectin are linked through α
using an internal standard, as did Englyst et al. (1982). In this (1→ 4) bonds. Amylopectin is heavily branched and chains are linked
connection, the Megazyme® protocol eliminated variability related to through α (1→ 6) bonds to form a complex structure. Amylose shows a
sample volume by employing the total volume of the digest after low degree of branching (Bertoft, 2004). Conventionally, amylose in
potassium hydroxide treatment for centrifugation and then drawing starch is recognized by the deep blue color of the complex formed with

Fig. 3.3. Summary of protocols for the determination of RS1 (a) and RS2 (b) according to Englyst et al. (1992). RS1 represents the fraction of starch inaccessible to enzymes due to
large particle sizes. RS2 represents the enzyme resistant starch in raw foods.
1968 A. Perera et al. / Food Research International 43 (2010) 1959–1974

Fig. 5. Summary of the Megazyme method of determination of resistant starch (RS) (Megazyme, 2008).

I2. In this structure, poly-iodine occupies the interior of the amylose the amylose to amylopectin ratio, are governed by the genome and
helix consisting of six glucose units per turn and in almost circular cross- their genetic potential to undergo mutations (Rahman et al., 2007).
section (Rundle, 1947). A steady blue color (at 570 nm) corresponds to a For example, mutations in wheat that lead to null genes for GBSSI
degree of polymerization (DP) ≥ 45 or a helix with 7–8 turns. Amylose resulted in amylose-free endosperm starch (Vrinten & Nakamura,
chains with 12 b DPb 45 produce a range of colors from red to purple, 2000). Similarly, inhibition of the expression of GBSSI in potato
respectively (Bailey & Whelan, 1961). In the amylopectin structure, produced amylose-free starch. On the other hand, rice varieties high
areas where branching begins and those having only linear chains are in GBSSI were reported to produce endosperm rich in amylose (Wang
clearly separated, producing amorphous and crystalline layers, respec- et al., 1995). Differences in the GBSSI gene have been observed to
tively. It is believed that amylose is confined mostly to the amorphous produce differences in the amylopectin levels in japonica and indica
layers (Donald, 2004). Two models are postulated for the biosynthesis of rice (Umemoto, Yano, Satoh, Shomura, & Nakamura, 2002).
amylose. Debranching of pre-amylopectin to produce pre-amylose and The influence of genetics on the RS contents of pulses is presented
subsequent elongation of glucan chains by granule bound starch in Table 2. Three cultivars of pea, 1674-13, 1215-33 and 1329-12,
synthase I (GBSSI) to form amylose is a widely accepted model. The contained different levels of RS when analyzed using the same
other model proposes the synthesis of amylose through elongation of protocol. Two cultivars of lentil, CDC Meteor and CDC Robin, had
short chains of malto-oligosaccharides by GBSSI (Chibbar, Ganeshan, & similar levels of RS (Chung, Liu, Hoover, Warkentin, & Vandenberg,
Båga, 2007; Smith, 2001; van de Wal et al., 1998). 2008). Similarly, three cultivars of common bean, Majesty, Red
German, Blumenfeld, Guenin, Yuryev, and Toistoguz (1992) dem- Kannern and AC Nautica, had similar levels of RS (Chung et al.,
onstrated that gelatinized amylose aggregates to form a gel network 2008). Significant differences in RS levels were observed among
during cooling, whereas amylopectin acts as an agent of aggregation chickpea cultivars (Myles, FLIP 97-101C, 97-Indian2-11) (Chung, Liu,
and a network filler. Hydrolysis of amylose gels is dependant on gel Donner, et al., 2008).
strength, since diffusion of α-amylase is decreased in strong gels that Genotype-specific differences in the amylose and amylopectin
form resistant starch (Cairns, Sun, Morris, & Ring, 1995). The ratio of ratios of sorghum and related differences in RS were reported.
amylose to amylopectin, their chain lengths and the presence of lipid According to Sang, Bean, Seib, Pedersen, and Shi (2008), starch from
influence the degree of retrogradation of starch from cereals and heterowaxy sorghum (14% amylose) had the highest amount of RS
pulses (Chung & Liu, 2009a). In lipid-rich starches, gelatinized (23.7%) compared to waxy sorghum starch (0% amylose; 8.4% RS) and
amylose preferentially combines with lipid at a relatively high normal sorghum starch (23.7% amylose; 17.9% RS), indicating the
temperature (during cooling) before retrogradation begins and significance of the amylose to amylopectin ratio with respect to the RS
impedes retrogradation of the remainder of the amylose. On the level. In another study of four different hybrids, aeae × aeae (high
other hand, small amounts of amylose-lipid complexes act as amylose); fl1fl1 × fl1fl1; aeae (female) × fl1fl1 (native) and aeae
nucleation sites for amylose chains. During retrogradation, long (pollen) × fl1fl1 (female), starch from self pollinated aeae
chains of amylose in potato and amylopectin in amylomaize form produced the highest amounts of RS (55.2%) compared to the other
weak associations at low temperatures (exothermic peak at ∼30 °C) hybrids (1–6% RS) (Yao, Paez, & White, 2009). Genotypic variations
perhaps due to the obstructions caused by branched chains (Chung & influenced RS levels in thermally processed low- (23-Am), and high-
Liu, 2009a). Variations in the composition of cereal starch, in terms of amylose (67-Am) pea (Pisum sativum L). Autoclaved (1.35 bar for 1 h)
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1969

Table 1
Comparison of resistant starch contents of foods as determined by different methods.

Food Description RS Description of hydrolysis Author

White bread 1% Pancreatin, amyloglucosidase Englyst et al. (1992)


2% Chewing, pepsin, pancreatin, amyloglucosidase, Åkerberg, Liljeberg, and Björck (1998)
MgCl2, CaCl2, 2-propanol
Wholemeal bread 1% Pancreatin, amyloglucosidase Englyst et al. (1992)
70% barley wholemeal and 30% 6% Chewing, pepsin, pancreatin, amyloglucosidase, Åkerberg, Liljeberg, Granfeldt, et al. (1998)
white wheat flour MgCl2, CaCl2, 2-propanol
Wheat based products Wheat flour 0.085% Termamyl, protease, amyloglucosidase Tharanathan and Tharanathan (2001)
(Chapati, Poori and Phulka) Wheat flour and water at 2.5:1; salt 0.67–1.0%
and ground nut oil
Cornflakes 3% Pancreatin, amyloglucosidase Englyst et al. (1992)
6.5% Pepsin, pancreatin, amyloglucosidase, Muir and O'Dea (1992)
Termamyl
Rice Whole grain 11.8% Pepsin, pancreatin, amyloglucosidase, Muir and O'Dea (1992)
Termamyl
Retrograded native rice starch 9.3% Pancreatin, amyloglucosidase Chung et al. (2006)
Gelatinized rice starch 8.6% Pancreatin, amyloglucosidase Chung et al. (2006)
Barley Extruded flour 0.05% Lichinase, β-glucosidase, fungal protease, Faraj et al. (2004)
(100 °C, 20% moisture) Termamyl, amyloglucosidase
Raw, pure starch 4.1% Heat stable α amylase, amyloglucosidase Szczodrak and Pomeranz (1991)
Autoclaved, high-amylose starch 7.1% Heat stable α amylase Szczodrak, and Pomeranz (1992)
Green banana Dried flour at 52–58 °C 41–59% Amyloglucosidase, pepsin and amylase Tribess et al. (2009)
(AOAC method)
Biscuit (1:1 banana and wheat flour) 15% Pancreatin, amyloglucosidase Englyst, Kingman, et al. (1996)
Green pea (65-Am) Boiled seeds 33.4% Chewing, α-amylase, amyloglucosidase Skrabanja et al. (1999)
Autoclaved seeds 32.6%
Pea Boiled and frozen 6.3% As per Englyst et al.(1992) Bravo (1999)
Canned (autoclaved) 4.9%

seed of 23-Am (7.6% RS on a starch basis) had a lower level of RS than potato starch, Eerlingen, Deceuninck, and Delcour (1993) demon-
did boiled seed (12.85% RS on a starch basis), whereas autoclaved strated that the DP of amylose (chain lengths from 40 to 610) was not
(32.6% RS) and boiled (33.4% RS) seed of 65-Am had similar levels of related to the DP of RS (19–26) formed. Escarpa, González, Maňas,
RS (Skrabanja, Liljeberg, Hedley, Kreft, & Björck, 1999). Three García-Diz, and Saura-Calixto (1996) adjusted amylose to amylopec-
genetically-modified potato lines with different amylose contents tin (extracted from potato starch) ratios to observe changes in RS
were compared for their RS contents after similar gelatinization and contents. In an autoclave (115 °C, 2 bar, 20 min), samples of amylose,
retrogradation treatments (Leeman, Karlsson, Eliasson, & Björck, amylopectin and their mixtures were gelatinized with water at a ratio
2006). RS contents increased with increasing amylose content. For of 1:20. An amylose to amylopectin ratio of 100:0 produced 36.5% RS,
example, line 527-1 contained 1% amylose and 0.1% RS, Prevalent, 23% and when the ratio was decreased to 50:50, the RS content decreased
amylose and 5.3% RS, and line 342, 64% amylose and 26% RS. According to 21.5%. When the ratio was changed to 0:100 (amylose: amylopec-
to Åkerberg, Liljeberg, and Björck (1998), bread made from barley rich tin), RS decreased further to 7.5%.
in amylopectin but low in amylose (3%) contained a lower amount of These data clearly show that the RS contents of pulse and cereal
RS (∼ 2.5%) than did bread made from high (44%) amylose barley starches, and of their starch pastes, vary with the genetic factors that
which contained ∼10% RS. The effect of amylose/amylopectin chain influence amylose and amylopectin content and their chain lengths.
length on the levels of RS formed is a complex phenomenon. Therefore, comparison of analytical protocols for determination of RS
Comparing wild-type rice with mutants rich in RS, Shu et al. (2007) should consider the genetic origin of analytes, as comparison of RS
concluded that the level of RS in cooked rice could not be explained on values of food products having similar starch ingredients but with
the basis of amylose content alone. Instead, RS content was different genetic make-ups may lead to invalid conclusions.
determined by the degree of polymerization (DP) of the side chains
of amylopectin. Rice with high levels of RS had increased amounts of 7. Effect of processing on RS levels
short chains (DP ≤ 12), and decreased amounts of intermediate
(13 ≤ DP ≤ 36) and long (DP ≥ 37) chains. Cooked waxy maize starch Available moisture and temperature influence starch gelatinization
with higher amounts of amylopectin long chains (DP N 13) than short and retrogradation (Evans & Haisman, 1982; Liu & Thompson, 1998;
chains (DP b 13) produced increasing amounts of RS over the cooling Eerlingen, Crombez, et al., 1993; Eerlingen, Deceuninck, et al., 1993).
phase at 4 °C, and after 160 h the RS level reached approximately 45 g/ Retrogradation of amylopectin occurs well at 2 °C forming elastic gels
100 g starch (Zhang, Sofyan, & Hamaker, 2008). Waxy maize starch which melt above 100 °C (Ring, 1985). Amylose gels retrograde better at
with higher amounts of amylopectin short chains (DP b 13) than long higher temperatures (Eerlingen, Crombez, & Delcour, 1993). Storage of
chains (DP N 13) produced very low levels of RS (b5 g/100 g starch) gelatinized, amylose-rich wheat starch at a low temperature (0 °C)
during the cooling phase. Ao, Simsek, Zhang, Venkatachalam, Reuhs favours nucleation of the gel network; however, propagation of the
and Hamaker (2007) reported that partial hydrolysis of α-1,4 linkages network is weak due to the low mobility of molecules. The reverse is
of side chains of pre-gelatinized maize starch using β-amylase and true during high temperature (100 °C) storage, where nucleation of the
maltogenic α-amylase and re-forming α-1,6 linkages using transglu- gel network is weak but propagation is favoured due to the increased
cosidase resulted in products that were higher in RS compared to mobility of molecules. During cooling an ‘ordered structure’ develops as
untreated maize starch. Cairns, Morris, Botham, and Ring (1996) double-helical regions of several amylose chains align parallel to each
studied the degree of polymerization (DP) of retrograded pea starch other, or single amylose chains fold to produce parallel areas, where
gels and reported three levels (on a number average basis), with DP folded regions are amorphous and the lamellae are crystalline. For
107–108 and 26–28 being the most common types, and 5, a minor example, the gel strength of potato starch is related to the storage
type, indicating the contribution of long chains to gel formation. Using temperature. Considering storage temperatures between 5 and 50 °C,
1970 A. Perera et al. / Food Research International 43 (2010) 1959–1974

Table 2
Resistant starch concentrations in pulses as influenced by genetics, processing method and RS analytical protocol.

Starch source Treatment RS Method of RS analysis Author

Pea Raw, flour Cultivar — 1674–13 13.3% AACC method 32–40 (2000) with modifications Chung, Liu, Hoover, et al. (2008)
Cultivar — 1215–33 14.7%
Cultivar — 1329–12 10.1%
Pea Starch Cultivar — 1674–13 12.6% AACC method 32–40 (2000) Chung, Liu, Donner, et al. (2008)
Cultivar — 1215–33 8.1%
Cultivar — 1329–12 9.7%
Pea Raw Cultivar — 1674–13 10% Englyst et al. (1992) with modifications Chung, Liu, and Hoover (2009)
Annealed (15 °C; 70% moisture; 24 h) 10.9%
Heat-moisture treatment 13.3%
(100 °C; 30% moisture; 2 h)
Heat-moisture treatment 14.5%
(120 °C; 30% moisture; 2 h)
Common bean Raw flour Cultivar — Majesty 36% AACC method 32–40 (2000) Chung, Liu, Pauls, et al. (2008)
Cultivar — Red Kanner 35.5%
Cultivar — AC Nautica 32.4%
Starch Cultivar — Majesty 21.3%
Cultivar — Red Kanner 21.9%
Cultivar — AC Nautica 17.2%
Lentils Raw flour Cultivar — CDC Meteor 14.9% AACC method 32–40 (2000) with modifications Chung, Liu, Hoover, et al. (2008)
Cultivar — CDC Robin 14.4%
Lentils Starch Cultivar — CDC Meteor 13% AACC method 32–40 (2000) Chung, Liu, Donner, et al. (2008)
Cultivar — CDC Robin 13.2%
Lentils Raw Cultivar — CDC Meteor 9.1% Englyst et al. (1992) with modifications Chung, Liu, and Hoover (2009)
Annealed (15 °C; 70% moisture; 24 h) 11.4%
Heat-moisture treatment 13.2%
(100 °C; 30% moisture; 2 h)
Heat-moisture treatment 14.7%
(120 °C; 30% moisture; 2 h)
Lentil starch-gelatinized Native Cultivar 1544-8 12.8% Englyst et al. (1992) with modification Chung et al. (2010)
Annealed with 70% moisture for 14.1%
24 h at 50 °C
Heat-moisture treatment 17.5%
(30% moisture + 24 h at ambient
temperature + 120 °C for 24 h)
Annealed and then heat-moisture 18.5%
treated
Heat-moisture treated and then 19.6%
annealed
Chickpea Raw flour Cultivar — Myles 6.4% AACC method 32–40 (2000) with modifications Chung, Liu, Hoover, et al. (2008)
Cultivar — FLIP 97-101C 4.7%
Cultivar — 97-Indian2-11 3.1%
Chickpea Starch Cultivar — Myles 8.4% AACC method 32–40 (2000) Chung, Liu, Donner, et al. (2008)
Cultivar — FLIP 97-101C 18.4%
Cultivar — 97-Indian2-11 13.3%
Moth bean Raw 12.2% Goñi et al. (1996) Bravo et al. (1998)
Freshly cooked 3.9%
Cooked; stored at 4 °C for 24 h 4.8%
Horse gram Raw 26.4% Goñi et al. (1996) Bravo et al. (1998)
Freshly cooked 5.2%
Cooked; stored at 4 °C for 24 h 5.8%
Black gram Raw 19.7% Goñi et al. (1996) Bravo et al. (1998)
Freshly cooked 3.4%
Cooked; stored at 4 °C for 24 h 4%
Navy bean starch-gelatinized Native 14.8% Englyst et al. (1992) with modification Chung et al. (2010)
Cultivar AC Compass Annealed with 70% moisture for 24 h at 16.4%
50 °C
Heat-moisture treatment 19.8%
(30% moisture + 24 h at ambient
temperature + 120 °C for 24 h)
Annealed and then heat-moisture 20.5%
treated
Heat-moisture treated and then 21.9%
annealed

the higher the storage temperature, the higher the gel strength, glucose molecules of an amylose chain participate in lamella formation
provided the starch content is N50% (Nakazawa, Noguchi, Takahashi, (Eerlingen, Crombez, et al., 1993; Eerlingen, Deceuninck, et al., 1993).
& Takada, 1985). X-ray diffraction studies showed that two different These helices are thought to be formed through hydrogen bonding (Wu
types of crystals were formed during retrogradation of wheat starch at & Sarko, 1978). As retrogradation proceeds, double helices are
40 and 95 °C. At higher temperatures, wheat starch retrograded to form organized into ‘hexagonal cells’ (Haralampu, 2000). Through their
structures with a large number of lamellae and having large particle influence on gelatinization and retrogradation, differences in the
sizes. Resistance to α-amylase increased as the number of lamellae in conditions of food processing, such as changes in moisture, temperature
crystals increased (Zabar, Shimoni, & Bianco-Peled, 2008). Lamellae do and duration of heating and subsequent cooling, influence the RS
not form along the entire length of the amylose chain, rather about 24 content of foods.
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1971

Recent publications describing the influence of processing method 1990), the presence of which may result in very high levels of resistant
on variations in the RS contents of pulses and pulse products are starch in assays of raw flours and meals.
summarized in Table 2. For three cultivars of pea, 1674-13, 1215-33 and The effect of processing on the RS content of some cereal products is
1329-12, refined starch contained less RS than did the meal (Chung, Liu, presented in Table 3. Han and Lim (2009) reported significantly higher
Hoover, et al., 2008), presumably due to the elimination of structural RS values for japonica brown rice as the moisture content of soaked rice
impediments to amylase hydrolysis, i.e. a reduction in RS1 levels, during reached 20%, compared to rice containing 30% moisture prior to cooking.
starch extraction. This observation also was applicable to starch isolated As reported by Chung et al. (2006), native waxy rice starch had the
from the common bean cultivars Majesty, Red Kanner and AC Nautica highest level of RS (∼9%), whereas gelatinized waxy rice starch (∼3%)
(Chung, Liu, Pauls, et al., 2008). For the chickpea cultivars Myles, FLIP 97- and retrograded, gelatinized waxy rice starch (3–3.8%) had the lowest
101C and 97-Indian 2-11, starch contained more RS (Chung, Liu, Donner, levels of RS. The observation of higher RS levels in native waxy starch
et al., 2008) than did the meal (Chung, Liu, Hoover, et al., 2008). In these than in gelatinized and retrograded waxy starches is expected, as all or
studies, pulse flours were prepared with a cyclone mill and starch was some of the RS1 and RS2 in native waxy starch may be converted to
isolated by the same procedure in each case. However, the analytical enzyme hydrolysable starch during heating, and the formation of RS in
procedures used to determine RS in chickpea starch and meal differed the high-amylopectin waxy starches would not be extensive (Eerlingen
slightly, making it difficult to draw conclusions from the observations. & Delcour, 1995; Russel, Berry, & Greenwell, 1989). This idea is
Raw moth bean, horse gram and black gram had considerably higher RS supported by the observation that waxy rice starch, partially gelatinized
values than did cooked beans. When the cooked pulses were stored at at 60 or 70 °C, had RS contents (8.6 and 7.7%, respectively) similar to that
4 °C for 24 h, RS contents increased slightly (Bravo, Siddhuraju, & Saura- of native waxy starch. The RS contents of retrograded gels of wheat, corn
Calixto, 1998). The cotyledons of some pulses, e.g. bean (Phaseolus and potato were increased significantly when starch was used instead of
vulgaris L.), contain mammalian and insect α-amylase inhibitors as part flour (as expected), but the opposite was reported for rice (García-
of the seeds' defence mechanism (Moreno, Altabella, and Chrispeels, Alonso, Jiménez-Escrig, Martín-Carrón, Bravo, & Saura-Calixto, 1999).

Table 3
Resistant starch concentrations in cereal starches as influenced by processing.

Starch source Treatment RS Method of RS analysis Author

Japonica brown rice Pre-soaked in water at 25 or 50 °C to reach 20% moisture + cooked 29.2–32.4% Englyst et al. (1992) Han, and Lim (2009)
Pre-soaked in water at 25 or 50 °C to reach 30% moisture + cooked 20.7–25.9%
Waxy rice Native waxy rice starch 9.3% Englyst et al. (1992) Chung et al. (2006)
Gelatinized starch 3.0%
Partially gelatinized starch at 60 °C for 5 min 8.6%
Partially gelatinized starch at 70 °C for 5 min 7.7%
Pastry wheat flour Extruded at 20% moisture; 150/200/250 rpm; 40–120 °C; stored at 4 °C/0 days 0.48–0.52% Megazyme® assay Kim et al. (2006)
Extruded at 20% moisture; 150/200/250 rpm; 40–120 °C; stored at 4 °C/7–14 days 1.21–1.35%
Extruded at 40% moisture; 150/200/250 rpm; 40–120 °C; stored at 4 °C/0 days 0.63–0.67%
Extruded at 40% moisture; 150/200/250 rpm; 40–120 °C; stored at 4 °C/7–14 days 1.52–1.86%
Extruded at 60% moisture; 150/200/250 rpm; 40–120 °C; stored at 4 °C/0 days 2.54–2.65%
Extruded at 60% moisture; 150/200/250 rpm; 40–120 °C; stored at 4 °C/7–14 days 3.55–4.25%
Corn Acid modified with 1.64 M HCl at 40 °C for 4 h + gelatinized 5% AOAC 991.43 (1998) Koksel, Masatcioglu,
+ autoclaved + lyophilized Kahraman, Ozturk, and
Basman (2008)
Acid modified with 1.64 M HCl at 40 °C for 4 h + gelatinized 12%
+ autoclaved + Stored at 95 °C for 48 h + lyophilized
Corn Normal corn starch 19.7% Englyst et al. (1992) modified as Chung, Hoover, and Liu
per Chung, Liu, et al. (2009) (2009)
Annealed with excess water for 24 h at 50 °C 18.3%
Heat-moisture treatment 16.9%
(30% moisture + 24 h at ambient temperature + 120 °C for 24 h)
Annealed with excess water for 24 h at 50 °C + Heat-moisture treatment 17.3%
(30% moisture + 24 h ambient temperature + 120 °C for 24 h)
Heat-moisture treatment 19.7%
(30% moisture + 24 h ambient temperature + 120 °C for 24 h)
+ annealed with excess water for 24 h at 50 °C
Corn Native corn starch 1.3% AOAC 991.43 (1985) Brumovsky, Brumovsky,
Fretes, and Peralta (2009)
Heat-moisture treatment 1.5%
— 30% moisture + 100 °C for 60 min
30% moisture + 120 °C for 60 min 4.2%
Wheat Native wheat starch 1.0%
Heat-moisture treatment 1.9%
— 30% moisture + 80 °C for 60 min
30% moisture + 100 °C for 60 min 1.6%
High-amylose corn Autoclaved at 120 °C + Stored at 4 °C for 24 h (repeated twice) 30% Goñi et al. (1996) Zhao and Lin (2009)
Autoclaved at 120 °C + Stored at 4 °C for 24 h (repeated twice) 39%
+ hydrolysed with 0.1 M citric acid
Corn (normal) Treated with acid (HCl) and methanol at 25 °C for 30 days ∼12% Englyst et al. (1992) Lin et al. (2009)
Treated with acid (HCl) and methanol treatment at 25 °C for 30 days ∼38%
+ annealed at 50 °C for 72 h with water
Corn (normal) Gamma irradiation (60Co) 0 kGy 19.7% Englyst et al. (1992) Chung and Liu (2009b)
10 kGy (2 kGy/h) 22.2%
10 kGy (0.67 kGy/h) 23.0%
10 kGy (0.40 kGy/h) 24.7%
1972 A. Perera et al. / Food Research International 43 (2010) 1959–1974

Green banana starch is a rich source of RS but the amounts varied to significant differences in the RS contents of foods and food
significantly (41–59%) with the method of drying employed during ingredients. Therefore, meaningful comparison of analytical proce-
starch processing (Tribess et al., 2009). dures for RS determination requires analysis of foods processed under
Heat-moisture treatments (30% moisture, 24 h at ambient temper- identical conditions. Comparison of analytical procedures for RS using
ature, heated at 100 or 120 °C for 2 h) were applied to corn (Table 3) and currently available data is difficult as the analytes used are often
lentil and pea (Table 2) starches, with and without a pre-gelatinization dissimilar.
treatment (Chung, Liu, & Hoover, 2009). The heat-moisture treatment
increased RS levels considerably. For example, native corn, lentil and pea 8. Conclusion
starches had RS contents of 5, 9 and 10% respectively; corresponding
values were increased to 12%, 15% and 15% after the 120 °C heat- Analytical procedures for RS vary with respect to the sample
moisture treatment. In another experiment, a heat-moisture treatment preparation method, particle size of samples, nature of the prelimi-
(30% moisture, 24 h at ambient temperature, heated at 120 °C for 24 h) nary heat treatments for raw analytes (if any), enzymes and buffers
decreased the level of RS in corn starch (16.9%) compared to that of employed, incubation temperatures, sampling of the enzyme digests,
native, normal corn starch (19.7%). A subsequent annealing treatment chemicals and conditions of hydrolysis, and the glucose determina-
(sample to water ratio of 3:7, 50 °C, 24 h) compensated for the lower RS tion method. The effects of these changes on the RS values obtained
content (Chung, Hoover, & Liu, 2009). Heating of corn starch at 30% generally cannot be compared as the analytes used were not identical
moisture for 40 or 60 min at 100 °C or 120 °C yielded products with 1.3 in most cases. Genetic differences with respect to the source of starch,
and 1.5% RS, respectively, at 100 °C and with 1.8 and 4.2% RS, type of processing (meal vs. starch) and conditions of processing
respectively, at 120 °C, illustrating the significance of the combined (partial vs. complete gelatinization), chemical modification of
time and temperature effect on RS levels (Brumovsky, Brumovsky, samples, and storage conditions significantly influence RS values.
Fretes, & Peralta, 2009). Annealed and/or heat-moisture-treated pulse Therefore, the concept of RS in a food material is a subjective
starches which were subsequently gelatinized produced higher parameter and any generalizations made for a particular food item or
amounts of RS than did gelatinization of native starches (Chung, Liu, & category must be made with caution.
Hoover, 2010). For example, gelatinized, native lentil and navy bean
starches contained 12.8 and 14.8 g RS/100 g starch, respectively. Lentil
starch annealed (70% moisture, 24 h, 50 °C) and subsequently heat- References
moisture treated (30% moisture, 120 °C, 24 h), and heat-moisture- AACC International (2000). Approved methods of the American association of cereal
treated lentil starch which was subsequently annealed, exhibited RS chemists. St. Paul, MN: AACC, Inc.
Abdulnour-Nakhoul, S., Nakhoul, N. L., Wheeler, S. A., Wang, P., Swenson, E. R., &
levels of 18.5 and 19.6 g/100 g starch, respectively, after gelatinization
Orlando, R. C. (2005). HCO3− secretion in the esophagal submucosal glands.
(Table 2). Corresponding values for navy bean after gelatinization were American Journal of Physiology, 22, G736−G744.
20.5 and 21.9 g/100 g starch (Table 2). The effect of extrusion pelletizing Åkerberg, A., Liljeberg, H., & Björck, I. (1998). Effects of amylose/amylopectin ratio and
on RS was determined with barley, wheat, oat, corn and rice. Dough was baking conditions on resistant starch formation and glycaemic indices. Journal of
Cereal Science, 28, 71−80.
cooked at 95 °C for 1 h before extrusion and then dried at 100 °C for 1 h. Åkerberg, A. K. E., Liljeberg, H. G. M., Granfeldt, Y. E., Drews, A. W., & Björck, I. M. E. (1998).
For all grain types used, with the exception of rice, a significant An in vitro method, based on chewing, to predict resistant starch content in foods
reduction in RS was reported after extrusion (Hernot, Boileau, Bauer, allows parallel determination of potentially available starch and dietary fiber. The
Journal of Nutrition, 12, 651−660.
Swanson, & Fahey, 2008). The RS level of extruded pastry wheat flour Ao, Z., Simsek, S., Zhang, G., Venkatachalam, M., Reuhs, B. L., & Hamaker, B. R. (2007). Starch
increased as the initial moisture level of the feed (20, 40 or 60%) and with a slow digestion property produced by altering its chain length, branch density, and
post-process storage time at 4 °C (0, 7 or 14 days) increased (Kim, crystalline structure. Journal of Agricultural and Food Chemistry, 55, 4540−4547.
AOAC (1985). Official methods of analysis. Arlington, VA: AOAC.
Tanhehco, & Ng, 2006) (Table 3). High-amylose corn starches (Hylon VII AOAC (1998). Official methods of analysis. Arlington, VA: AOAC.
and Gelose 80; 70 and 80% amylose and 60 and 46% RS, respectively) Bailey, J. M., & Whelan, W. J. (1961). Physical properties of starch I. Relationship
subjected to extrusion cooking (35 and 50% moisture, 100 and 140 °C, between iodine stain and chain length. Journal of Biological Chemistry, 236(4),
969−973.
150 and 750 s−1 shear) exhibited significantly reduced RS concentra-
Bertoft, E. (2004). Understanding starch structure and functionality. In A. Eliasson (Ed.),
tions (14 and 17%, respectively). This was thought to be due to enhanced Starch in food: Structure, function and applications (pp. 156−184)., 1st ed.
opportunities for amylase action after shear damage to starch granules Cambridge: Woodhead Publishing Limited.
Bird, A. R., Brown, I. L., & Topping, D. L. (2000). Starches, resistant starches, the gut
(Htoon et al., 2009). Both a pre-gelatinization treatment of corn starch
microflora and human health. Current Issues in Intestinal Microbiology, 1(1), 25−37.
with HCl at 40 °C (Koksel et al., 2008) and a post-retrogradation Bravo, L. (1999). Effect of processing on the non-starch polysaccharides and in vitro starch
treatment of high-amylose corn starch with citric acid (Zhao & Lin, digestibility of legumes. Food Science and Technology International, 5, 415−423.
2009) increased the level of RS (Table 3). Wheat and corn starch cooked Bravo, L., Siddhuraju, P., & Saura-Calixto, F. (1998). Effect of various processing methods
on the in vitro starch digestibility and resistant starch content of Indian pulses.
at 100 °C for 10 min during an in vitro digestibility assay had RS levels of Journal of Agricultural and Food Chemistry, 46, 4667−4674.
9 and 11.4%, respectively; similar levels of RS were seen in wheat Brouns, F., Kettlitz, B., & Arrigoni, E. (2002). Resistant starch and butyrate revolution.
(10.9%) and corn (13.3%) starch autoclaved three times at 121 °C for Trends in Food Science & Technology, 13, 251−261.
Brumovsky, L. A., Brumovsky, J. O., Fretes, M. R., & Peralta, J. M. (2009). Quantification of
15 min and then cooled within 15 min (Hickman, Janaswamy, & Yao, resistant starch in several starch sources treated thermally. International Journal of
2009). Starch autoclaved as above, and subsequently dried, when Food Properties, 12, 451−460.
subjected to β-amylase hydrolysis at pH 5.5 and 55 °C for 20 h exhibited Cairns, P., Morris, V. J., Botham, R. L., & Ring, S. G. (1996). Physicochemical studies on
resistant starch in vitro and in vivo. Journal of Cereal Science, 23, 265−275.
significantly higher levels of RS, 23.1 and 29.9% in wheat and corn starch, Cairns, P., Sun, L., Morris, V. J., & Ring, S. G. (1995). Physicochemical studies using amylose
respectively (Hickman et al., 2009). Commercially available canned as an in vitro model for resistant starch. Journal of Cereal Science, 21(1), 37−47.
beans had lower levels of RS than did canned bean flour prepared after Chibbar, R. N., Ganeshan, S., & Båga, M. (2007). In planta novel starch synthesis. In P.
Ranalli (Ed.), Improvement of crop plants for industrial end uses (pp. 181−208). The
drying at 50 °C (Osorio-Díaz et al., 2002). RS levels in corn starch
Netherlands: Springer.
subjected to acid–methanol treatment only or acid–methanol treatment Chow, P. (2008). The rationale for the use of animal models in biomedical research. In P.
combined with annealing at 50 °C for 72 h were 12 and 38 g/100 g K. H. Chow, R. T. H. Ng, & B. E. Ogden (Eds.), Using animal models in biomedical
research (pp. 2−10). Singapore: World Scientific Publishing Company Pte Limited
starch, respectively (Lin, Wang, & Chang, 2009). Gamma irradiation at
Available from: bhttp://www.worldscibooks.com/etextbook/6454/6454_chap01.
10 or 50 kGy (dose rate of 2 kGy/h at room temperature) increased the pdfN. Accessed October 08, 2009.
RS content of corn starch from 19.7% to 2.2 and 25.1%, respectively Chung, H., Hoover, R., & Liu, Q. (2009). The impact of single and dual hydrothermal
(Chung and Liu, 2009b). modifications on the molecular structure and physicochemical properties of normal
corn starch. International Journal of Biological Macromolecules, 44, 203−210.
It is apparent that even minor differences in the conditions under Chung, H., Lim, H., & Lim, S. (2006). Effect of partial gelatinization and retrogradation on
which starch gelatinization and retrogradation are achieved can lead the enzymic digestion of waxy rice starch. Journal of Cereal Science, 43, 353−359.
A. Perera et al. / Food Research International 43 (2010) 1959–1974 1973

Chung, H., & Liu, Q. (2009a). Impact of molecular structure of amylopectin and amylose on Goñi, I., García-Diza, L., Mańasb, E., & Saura-Calixto, F. (1996). Analysis of resistant
amylose chain association during cooling. Carbohydrate Polymers, 77, 807−815. starch: A method for foods and food product. Food Chemistry, 56(4), 445−459.
Chung, H. J., & Liu, Q. (2009b). Effect of gamma irradiation on molecular structure and Grabitske, H. A., & Slavin, J. L. (2009). Gastrointestinal effects of low-digestible
physicochemical properties of corn starch. Journal of Food Science, 74(5), 353−361. carbohydrates. Critical Reviews in Food Science and Nutrition, 49, 327−360.
Chung, H., Liu, Q., Donner, E., Hoover, R., Warkentin, T. D., & Vandenberg, B. (2008). Granito, M., Michel, C., Frías, J., Champ, M., & Guerra, M. (2005). Fermented Phaseolus
Composition, molecular structure, properties, and in vitro digestibility of vulgaris: Acceptability and intestinal effects. European Food Research and
starches from newly released Canadian pulse cultivars. Cereal Chemistry, 85(4), Technology, 220, 182−186.
471−479. Hamer, H. M., Jonkers, D. M. A. E., Bast, A., Vanhoutvin, S. A. L. W., Fischer, M. A. G.,
Chung, H., Liu, Q., & Hoover, R. (2009). Impact of annealing and heat-moisture treatment Kodde, A., et al. (2009). Butyrate modulates oxidative stress in the colonic mucosa
on rapidly digestible, slowly digestible and resistant starch levels in native and of healthy humans. Clinical Nutrition, 28, 88−93.
gelatinized corn, pea and lentil starches. Carbohydrate Polymers, 75, 436−447. Han, J., & Lim, S. (2009). Effect of presoaking on textural, thermal, and digestive
Chung, H., Liu, Q., & Hoover, R. (2010). Effect of single and dual hydrothermal properties of cooked brown rice. Cereal Chemistry, 86(1), 100−105.
treatments on the crystalline structure, thermal properties and nutritional fractions Haralampu, S. G. (2000). Resistant starch — A review of the physical properties and
of pea, lentil and navy bean starches. Food Research International, 43(2), 501−508. biological impact of RS3. Carbohydrate Polymers, 41, 285−292.
Chung, H., Liu, Q., Hoover, R., Warkentin, T. D., & Vandenberg, B. (2008). In vitro starch Hernot, D. C., Boileau, T. W., Bauer, L. L., Swanson, K. S., & Fahey, G. C. (2008). In vitro
digestibility, expected glycemic index, and thermal and pasting properties of flours digestion characteristics of unprocessed and processed whole grains and their
from pea, lentil and chickpea cultivars. Food Chemistry, 111, 316−321. components. Journal of Agricultural and Food Chemistry, 56, 10,721−10,726.
Chung, H., Liu, Q., Pauls, P., Fan, M. Z., & Yada, R. (2008). In vitro starch digestibility, Hickman, E., Janaswamy, S., & Yao, Y. (2009). Autoclave and β-amylosis lead to reduced
expected glycemic index and some physicochemical properties of starch and flour in vitro digestability of starch. Journal of Agricultural and Food Chemistry, 57,
from common bean (Phaseolus vulgaris L.) varieties grown in Canada. Food Research 7005−7012.
International, 41, 869−875. Htoon, A., Shrestha, A. K., Flanagan, B. M., Lopez-Rubio, A., Bird, A. R., Gilbert, E. P., et al.
Cichoke, A. J. (1998). The complete book of enzyme therapy.New York: Penguin Putnam (2009). Effects of processing high amylose maize starches under controlled
Inc. 10–16 pp. conditions on structural organisation and amylase digestibility. Carbohydrate
Correa, P. (1988). A human model of gastric carcinogenesis 1. Cancer Research, 48, Polymers, 75, 236−245.
3554−3560. Hylla, S., Gostner, A., Dusel, G., Anger, H., Bartram, H., Christl, S. U., et al. (1998). Effects
Correa, F., Angelina, T., Reis, P. M., Maria, S., & de Oliveira Costa, A. (2009). Increase in of resistant starch on the colon in healthy volunteers: Possible implications for
digestive organs of rats due to the ingestion of dietary fibre with similar solubility cancer prevention. American Journal of Clinical Nutrition, 6, 136−142.
to that of common bean. Archivos Latinoamericanos De Nutricon, 59(1), 47−53. Kendall, C. W. C., Esfahani, A., Hoffman, A. J., Evans, A., Sanders, L. M., Josse, A. R., et al.
Cummings, J. H., Beatty, E. R., Kingman, S. M., Bingham, S. A., & Englyst, H. N. (1996). (2008). Effect of novel maize-based dietary fibers on postprandial glycemia and
Digestion and physiological properties of resistant starch in the human large bowel. insulinemia. Journal of the American College of Nutrition, 27(6), 711−718.
British Journal of Nutrition, 75, 733−747. Kim, J. H., Tanhehco, E. J., & Ng, P. K. W. (2006). Effect of extrusion conditions on
DeSesso, J. M., & Jacobson, C. F. (2001). Anatomical physiological parameters affecting resistant starch formation from pastry wheat flour. Food Chemistry, 99, 718−723.
gastrointestinal absorption in humans and rats. Food and Chemical Toxicology, 39, Koksel, H., Masatcioglu, T., Kahraman, K., Ozturk, S., & Basman, A. (2008). Improving effect
209−228. of lyophilization on functional properties of resistant starch preparations formed by
Donald, A. M. (2004). Understanding starch structure and functionality. In A. Eliasson acid hydrolysis and heat treatment. Journal of Cereal Science, 47, 275−282.
(Ed.), Starch in food: Structure, function and applications (pp. 156−184). Cambridge: Leeman, A. M., Karlsson, M. E., Eliasson, A., & Björck, I. M. E. (2006). Resistant starch
Woodhead Publishing Limited. formation in temperature treated potato starches varying in amylose/amylopectin
Dronamraju, S. S., Coxhead, J. M., Kelly, S. B., & Mathers, J. C. (2007). Role of resistant ratio. Carbohydrate Polymers, 6, 306−313.
starch in colorectal cancer prevention: A prospective randomised controlled trial. Leu, R. K. L., Hu, Y., Brown, I. L., & Young, G. P. (2009). Effect of high amylose maize
American Journal of Gastroenterology, 102(S2), 556−557. starches on colonic fermentation and apoptotic response to DNA-damage in the
Eerlingen, R. C., Crombez, M., & Delcour, J. A. (1993). Enzyme-resistant starch. I. colon of rats. Nutrition & Metabolism, 6(1), 11−20.
Quantitative and qualitative influence of incubation time and temperature of Lin, J., Wang, S., & Chang, Y. (2009). Impacts of acid–methanol treatment and annealing
autoclaved starch on resistant starch formation. Cereal Chemistry, 70(3), 339−344. on the enzymatic resistance of corn starches. Food Hydrocolloids, 23, 1465−1472.
Eerlingen, R. C., Deceuninck, M., & Delcour, J. A. (1993). Enzyme resistant starch II. Liu, Q., & Thompson, D. B. (1998). Effects of moisture content and different
Influence of amylose chain length on resistant starch formation. Cereal Chemistry, gelatinization heating temperatures on retrogradation of waxy-type maize
70(3), 345−350. starches. Carbohydrate Research, 314, 221−235.
Eerlingen, R. C., & Delcour, J. A. (1995). Formation, analysis and properties of type III Maki, K. C., Sanders, L. M., Reeves, M. S., Kaden, V. N., Rains, T. M., & Cartwright, Y.
enzyme resistant starch. Journal of Cereal Science, 22, 129−138. (2009). Beneficial effects of resistant starch on laxation in healthy adults.
Englyst, H. N., & Cummings, J. H. (1985). Digestion of the polysaccharides of some cereal International Journal of Food Science and Nutrition, 60(S4), 296−305.
foods in the human small intestine. American Journal of Clinical Nutrition, 42, 778−787. Makinen, K. K. (1989). Salivary enzymes. In J. O. Tenovuo (Ed.), Human saliva: Clinical
Englyst, H. N., & Cummings, J. H. (1986). Digestion of the carbohydrates in banana chemistry and microbiology (volume II) (pp. 94−114). Florida: CRC Press.
(Musa paradisiaca sapientum) in the human small intestine. The American Journal of Megazyme (2008). Resistant starch assay procedure.http://secure.megazyme.com/
Clinical Nutrition, 44, 42−50. downloads/en/data/K-RSTAR.pdf Accessed July 25, 2009.
Englyst, H. N., & Cummings, J. H. (1987a). Digestion of the polysaccharides of potato in Moreno, J., Altabella, T., & Chrispeels, M. J. (1990). Characterization of α-amylase inhibitor,
the small intestine of man. American Journal of Clinical Nutrition, 45, 423−431. a lectin-like protein in the seeds of Phaseolus vulgaris. Plant Physiology, 92, 703−709.
Englyst, H. N., & Cummings, J. H. (1987b). Resistant starch, a ‘new’ food component: A Muir, J. G., & O'Dea, K. (1992). Measurement of resistant starch: Factors affecting the
classification of starch for nutritional purposes. In I. D. Morton (Ed.), Cereals in a amount of starch escaping digestion in vitro. Journal of Clinical Nutrition, 56, 123−127.
European context (pp. 221−233). New York: First European Conference on Food Nakazawa, F., Noguchi, S., Takahashi, J., & Takada, M. (1985). Retrogradation of
Science and Technology. gelatinized potato starch studied by differential scanning calorimetry. Agricultural
Englyst, H. N., & Hudson, G. L. (1996). The classification and measurement of dietary and Biological Chemistry, 49(4), 953−957.
carbohydrates. Food Chemistry, 57(1), 15−21. Nordgaard, I., & Mortensen, P. B. (1995). Digestive process in the human colon.
Englyst, H. N., Kingman, S. M., & Cummings, J. H. (1992). Classification and Nutrition, 11(1), 37−45.
measurement of nutritionally important starch fractions. European Journal of Osorio-Díaz, P., Bello-Pérez, L. A., Agama-Acevedo, E., Vargas-Torres, A., Tovar, J., & Paredes-
Clinical Nutrition, 46, S33−S50. López, O. (2002). In vitro digestibility and resistant starch content of some
Englyst, H. N., Kingman, S. M., Hudson, G. J., & Cummings, J. H. (1996). Measurement of industrialized commercial beans (Phaseolus vulgaris L.). Food Chemistry, 78, 333−337.
resistant starch in vitro and in vivo. British Journal of Nutrition, 75, 749−755. Rabbani, G. H., Ahmed, S., Hossain, I., Islam, R., Marni, F., Akhtar, M., et al. (2009). Green
Englyst, H. N., Veenstra, J., & Hudson, G. L. (1996). Measurement of rapidly available banana reduces clinical severity of childhood shigellosis: A double-blind, randomized,
glucose (RAG) in plant foods: A potential in vitro predictor of the glycaemic controlled clinical trial. The Pediatric Infectious Disease Journal, 28(5), 420−425.
response. British Journal of Nutrition, 75, 327−337. Rahman, S., Birda, A., Reginaa, A., Lia, Z., Rala, J. P., McMaugha, S., et al. (2007). Resistant
Englyst, H. N., Wiggins, H. S., & Cummings, J. H. (1982). Determination of the non-starch starch in cereals: Exploiting genetic engineering and genetic variation. Journal of
polysaccharides in plant foods by gas liquid chromatography of constituent sugars Cereal Science, 46, 251−260.
as alditol acetates. Analyst, 107, 307−318. Reddy, B. S., Ernst, L., & Wynderm, D. (1977). Metabolic epidemiology of colon cancer:
Escarpa, A., González, M. C., Maňas, E., García-Diz, L., & Saura-Calixto, F. (1996). Fecal bile acids and neutral sterols in colon cancer patients and patients with
Resistant starch formation: Standardization of a high-pressure autoclave process. adenomatous polyps. Cancer, 39, 2533−2539.
Journal of Agricultural and Food Chemistry, 44, 924−928. Ring, S. G. (1985). Observations on the crystallization of amylopectin from aqueous
Evans, D., & Haisman, D. R. (1982). The effect of solutes on the gelatinization solution. International Journal of Biological Macromolecules, 7(4), 253−254.
temperature range of potato starch. Starch, 34(7), 224−231. Rundle, R. E. (1947). The configuration of starch in the starch–iodine complex. V.
Faraj, A., Vasanthan, T., & Hoover, R. (2004). The effect of extrusion cooking on resistant Fourier projections from X-ray diagrams. Journal of American Chemical Society, 69
starch formation in waxy and regular barley flours. Food Research International, 37, (7), 1769−1772.
517−525. Russel, C. L., Berry, P. S., & Greenwell, P. (1989). Characterization of resistant starch from
García-Alonso, A., Jiménez-Escrig, A., Martín-Carrón, N., Bravo, L., & Saura-Calixto, F. wheat and maize. Journal of Cereal Science, 9(1), 1−15.
(1999). Assessment of some parameters involved in the gelatinization and Sajilatha, M. G., Singhal, R. S., & Kulkarni, P. R. (2006). Resistant starch: A review.
retrogradation of starch. Food Chemistry, 66, 181−187. Comprehensive Reviews in Food Science and Food Safety, 5, 1−17.
German, M. L., Blumenfeld, A. L., Guenin, Y. V., Yuryev, V. P., & Toistoguz, V. B. (1992). Sang, Y., Bean, S., Seib, P. A., Pedersen, J., & Shi, Y. (2008). Structure and functional
Structure formation in systems containing amylose, amylopectin and their properties of sorghum starches differing in amylose content. Journal of Agricultural
mixtures. Carbohydrate Polymers, 18, 27−34. and Food Chemistry, 56, 6680−6685.
1974 A. Perera et al. / Food Research International 43 (2010) 1959–1974

Saura-Calixto, F., Goňi, I., Bravo, L. E., & Maňas, E. (1993). Resistant starch in foods: van de Wal, M., D'Hulst, C., Vincken, J., Buléon, A., Visser, R., & Ball, S. (1998). Amylose is
Modified method for dietary fiber residues. Journal of Food Science, 58(3), 642−643. synthesized in vitro by extension of and cleavage from amylopectin. Journal of
Seib, P. A., and Woo, K. (1999). Food grade starch resistant to α-amylase and the Biological Chemistry, 273(35), 22,232−22,240.
method of preparing the same. US Patent Number 5,855,946. Available from bhttp: Vrinten, P. L., & Nakamura, T. (2000). Wheat granule-bound starch synthase I and II are
//www.freepatentsonline.com/5855946.pdfN. Accessed 30 April, 2010. encoded by separate genes that are expressed in different tissues. Plant Physiology,
Shu, X., Jiaa, L., Gao, J., Song, Y., Zhao, H., Nakamura, Y., et al. (2007). The influences of 122, 255−263.
chain length of amylopectin on resistant starch in rice (Oryza sativa L.). Starch, 59, Wang, Z., Zheng, F., Shen, G., Gao, J., Snusted, D. P., Li, M., et al. (1995). The amylose
504−509. content in rice endosperm is related to the post-transcriptional regulation of the
Silvi, S., Rumney, C. J., Cresci, A., & Rowland, I. R. (1999). Resistant starch modifies gut waxy gene. The Plant Journal, 7(4), 613−622.
microflora and microbial metabolism in human flora-associated rats inoculated Wolf, B. W., Bauer, L. L., & Fahey, G. C. (1999). Effects of chemical modification on in
with faeces from Italian and UK donors. Journal of Applied Microbiology, 86, vitro rate and extent of food starch digestion: An attempt to discover a slowly
521−530. digested starch. Journal of Agricultural and Food Chemistry, 47, 4178−4183.
Skrabanja, V., Liljeberg, H. G. M., Hedley, C. L., Kreft, I., & Björck, I. M. E. (1999). Influence Wong, J. M. W., de Souza, R., Kendall, C. W. C., Emam, A., & Jenkins, D. J. A. (2006).
of genotype and processing on the in vitro rate of starch hydrolysis and resistant Colonic health: Fermentation and short chain fatty acids. Journal of Clinical
starch formation in peas (Pisum sativum L.). Journal of Agricultural and Food Gastroenterology, 40(3), 235−243.
Chemistry, 47, 2033−2039. Wu, H. C. H., & Sarko, A. (1978). The double-helical molecular structure of crystalline B-
Smith, A. M. (2001). The biosynthesis of starch granules. Biomacromolecules, 2, amylose. Carbohydrate Research, 61, 7−25.
335−341. Xue, Q., Newman, R. K., & Newman, C. W. (1996). Effects of heat treatment of barley
Storey, D., Lee, A., Bornet, F., & Brouns, F. (2007). Gastrointestinal responses following starches on in vitro digestibility and glucose responses in rats. Cereal Chemistry, 73
acute and medium term intake of retrograded resistant maltodextrins, classified as (5), 588−592.
type 3 resistant starch. European Journal of Clinical Nutrition, 61, 1262−1270. Yao, N., Paez, A. V., & White, P. J. (2009). Structure and function of starch and resistant
Szczodrak, J., & Pomeranz, Y. (1991). Starch and enzyme-resistant starch from high- starch from corn with different doses of mutant amylase-extender and floury-1
amylose barley. Cereal Chemistry, 68(6), 589−596. alleles. Journal of Agricultural and Food Chemistry, 57, 2040−2048.
Szczodrak, J., & Pomeranz, Y. (1992). Starch–lipid interactions and formation of Zabar, S., Shimoni, E., & Bianco-Peled, H. (2008). Development of nanostructure in
resistant starch in high amylose barley. Cereal Chemistry, 69(6), 626−632. resistant starch type III during thermal treatments and cycling. Macromolecular
Tharanathan, M., & Tharanathan, R. N. (2001). Resistant starch in wheat-based Bioscience, 8, 163−170.
products: Isolation and characterization. Journal of Cereal Science, 34, 73−84. Zhang, G., Sofyan, M., & Hamaker, B. R. (2008). Slowly digestible state of starch:
Tribess, T. B., Hernández-Uribe, J. P., Méndez-Montealvo, M. G. C., Menezes, E. W., Bello- mechanism of slow digestion property of gelatinized maize starch. Journal of
Perez, L. A., & Tadini, C. C. (2009). Thermal properties and resistant starch content Agricultural and Food Chemistry, 56, 4695−4702.
of green banana flour (Musa cavendishii) produced at different drying conditions. Zhao, X., & Lin, Y. (2009). Resistant starch prepared from high-amylose maize starch
Food Science & Technology, 42, 1022−1025. with citric acid hydrolysis and its simulated fermentation in vitro. European Food
Umemoto, T., Yano, M., Satoh, H., Shomura, A., & Nakamura, Y. (2002). Mapping of a Research and Technology, 228, 1015−1021.
gene responsible for the difference in amylopectin structure between japonica-type
and indica-type rice varieties. Theoretical and Applied Genetics, 104, 1−8.

You might also like