You are on page 1of 17

Resistant Starch A Review

M.G. Sajilata, Rekha S. Singhal, and Pushpa R. Kulkarni

ABSTRACT CT: resistant starch evoked interest bioavailability starch ABSTRACT: The concept of resistant starch (RS) has evoked new interest in the bioavailability of starch and in its use as source dietary fiber, particularly adults. now considered pro proper operties a source of dietary fiber, particularly in adults. RS is now considered to provide functional properties and find applicaformation, formation, var ariety foods. tions in a variety of foods. Types of RS, factors influencing their formation, consequence of such formation, their methods prepar eparation, have briefly review eview. of preparation, their methods of estimation, and health benefits have been briefly discussed in this review. eywords: resistant starch functionality, formation, prepar eparation, determination, digestibility, Keywords: resistant starch (RS), functionality, formation, preparation, determination, digestibility, physiological effects, applications, commercial sources

Introduction
From the early years of emergence of nutritional science, it has been recognized that the ingested nutrients in the diet are not completely utilized in the body. An increasing volume of evidence suggests that with very few exceptions, only a proportion of the total ingested nutrients in a diet or food is available, and the term availability has come into use for this proportion (Southgate 1989). The nutrients measured by chemical analysis may not always be fully utilizable, mainly due to the indigestible cell walls, a bulky or dense structure, a low solubility, the presence of some compounds inhibiting the digestion, as well as components abundantly present in plant foods such as dietary fiber, phytic acid, and tannic acid, which may significantly reduce the absorption and utilization of some nutrients (Rosado and others 1987). During food processing, derivatization of nutrients and formation of cross linkages occur, thereby making the food inaccessible for digestion or/and metabolism. Such parts of nutrients are also unavailable (Erbersdobler 1989). Starch, which is the major dietary source of carbohydrates, is the most abundant storage polysaccharide in plants, and occurs as granules in the chloroplast of green leaves and the amyloplast of seeds, pulses, and tubers (Ellis and others 1998). The relatively recent recognition of incomplete digestion and absorption of starch in the small intestine as a normal phenomenon has raised interest in nondigestible starch fractions (Cummings and Englyst 1991; Englyst and others 1992). These are called resistant starches, and extensive studies have shown them to have physiological functions similar to those of dietary fiber (Asp 1994; Eerlingen and Delcour 1995). The diversity of the modern food industry and the enormous variety of food products it produces require starches that can tolerate a wide range of processing techniques and
MS 20050127 Submitted 2/28/05, Revised 8/2/05, Accepted 10/29/05. The authors are with Food Engineering and Technology Dept., Inst. of Chemical Technology, Matunga, Mumbai 400 019, India. Email: rekha@udct.org.

preparation conditions (Visser and others 1997). These demands are met by modifying native starches with chemical, physical, and enzymatic methods (Betancur and Chel 1997), which may lead to the formation of indigestible residues. The availability of such starches therefore deserves consideration. This review therefore focuses on the availability of the major nutrient, that is, the starch, with special reference to RS.

Starch and its classification


Chemically, starches are polysaccharides, composed of a number of monosaccharides or sugar (glucose) molecules linked together with -D-(1-4) and/or -D-(1-6) linkages. The starch consists of 2 main structural components, the amylose, which is essentially a linear polymer in which glucose residues are -D-(1-4) linked typically constituting 15% to 20% of starch, and amylopectin, which is a larger branched molecule with -D-(1-4) and -D-(1-6) linkages and is a major component of starch (BNF 1990). Amylose is linear or slightly branched, has a degree of polymerization up to DP 6000, and has a molecular mass of 105 to 106 g/mol. The chains can easily form single or double helices (Takeda and Takeda 1989). On the basis of X-ray diffraction studies on oriented amylase fibers, the presence of type A and type B amyloe is indicated (Figure 1, Galliard 1987). The structural elements of type B are double helices, which are packed in an antiparallel, hexagonal mode. The central channel surrounded by 6 double helices is filled with water (36 H2O/unit cell). Type A is very similar to type B, except that the central channel is occupied by another double helix, making the packing closer. In this type, only 8 molecules of water per unit cell are inserted between the double helices. Amylopectin (107 to 109 g/mol) is highly branched and has an average DP of 2 million, making it one of the largest molecules in nature. Chain lengths of 20 to 25 glucose units between branch points are typical. Its structure is often described by a cluster model (Figure 2). The cluster model gained greater credence when Hizukuri postulated that amylopectin
1

2006 Institute of Food Technologists

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


chains are either located within a single cluster or serve to connect 2 or more clusters (Hizukuri 1986; Thompson 2000). Short chains (A) of DP 12-16 that can form double helices are arranged in clusters. The clusters comprise 80% to 90% of the chains and are linked by longer chains (B) that form the other 10% to 20% of the chains. Most B-chains extend into 2 (DP about 40) or 3 clusters (DP about 70), but some extend into more clusters (DP about 110) (Figure 2 and 3) (www.lsbu.ac.uk). On the basis of X-ray diffraction experiments, starch granules are said to have a semicrystalline character, which indicates a high degree of orientation of the glucan molecules. About 70% of the mass of starch granule is regarded as amorphous and about 30% as crystalline. The amorphous regions contain the main amount of amylose but also a considerable part of the amylopectin. The crystalline region consists primarily of the amylopectin.
Various ways to classify native starches

X-ray diffraction. Three types of starches, designated as type A, type B, and type C, have been identified based on X-ray diffraction

patterns. These depend partly on the chain lengths making up the amylopectin lattice, the density of packing within the granules, and the presence of water (Wu and Sarko 1978). Although type A and type B are real crystalline modifications, type C is a mixed form. The important features of the types of starches are as follows. Type A. The type A structure has amylopectin of chain lengths of 23 to 29 glucose units. The hydrogen bonding between the hydroxyl groups of the chains of amylopectin molecules results in the formation of outer double helical structure. In between these micelles, linear chains of amylose moieties are packed by forming hydrogen bonds with outer linear chains of amylopectin. This pattern is very common in cereals. Type B. The type B structure consists of amylopectin of chain lengths of 30 to 44 glucose molecules with water inter-spread. This is the usual pattern of starches in raw potato and banana. Type C. The type C structure is made up of amylopectin of chain lengths of 26 to 29 glucose molecules, a combination of type A and type B, which is typical of peas and beans. An additional form, called type V, occurs in swollen granules. Xray diffraction diagrams of these starches are shown in Figure 4.
Based on the action of enzymes

According to Berry (1986), starches can be classified according to their behavior when incubated with enzymes without prior exposure to dispersing agents as follows. Rapidly digestible starch (RDS). RDS consists mainly of amorphous and dispersed starch and is found in high amounts in starchy foods cooked by moist heat, such as bread and potatoes. It is measured chemically as the starch, which is converted to the constituent glucose molecules in 20 min of enzyme digestion.

Figure 1Unit cells and arrangement of double helices (cross section) in A-amylose (top) and B-amylose (bottom) (Galliard 1987) 2

Figure 2Cluster model of amylopectin

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


Slowly digestible starch (SDS). Like RDS, SDS is expected to be completely digested in the small intestine, but for 1 reason or another, it is digested more slowly. This category consists of physically inaccessible amorphous starch and raw starch with a type A and type C crystalline structure, such as cereals and type B starch, either in granule form or retrograded form in cooked foods. It is measured chemically as starch converted to glucose after a further 100 min of enzyme digestion. Resistant starch. The term resistant starch was first coined by Englyst and others (1982) to describe a small fraction of starch that was resistant to hydrolysis by exhaustive -amylase and pullulanase treatment in vitro. RS is the starch not hydrolyzed after 120 min of incubation (Englyst and others 1992). However, because starch reaching the large intestine may be more or less fermented by the gut microflora, RS is now defined as that fraction of dietary starch, which escapes digestion in the small intestine. It is measured chemically as the difference between total starch (TS) obtained from homogenized and chemically treated sample and the sum of RDS and SDS, generated from non-homogenized food samples by enzyme digestion. RS = TS (RDS + SDS)
Based on the nutritional characteristics

Figure 3(a) Shows the essential features of amylopectin. (b) Shows the organization of the amorphous and crystalline regions (or domains) of the structure generating the concentric layers that contribute to the growth rings that are visible by light microscopy. (c) Shows the orientation of the amylopectin molecules in a cross section of an idealized entire granule. (d) Shows the likely double helix structure taken up by neighboring chains and giving rise to the extensive degree of crystallinity in granule (www.lsbu.ac.uk).

This classification is based on the extent of digestibility of the starch as follows. Digestible starches. These include the starches digestible by body enzymes, namely the rapidly digestible starches (RDS) and the slowly digestible starches (SDS). RDS consists mainly of amorphous and dispersed starch, found in high amounts in starchy foods cooked by moist heat. Like RDS, SDS is expected to be completely digested in the small intestine, but for 1 reason or another, it is digested more slowly. Resistant starch. RS is indigestible by body enzymes. It is subdivided into 4 fractions: RS1, RS2, RS3, and RS4. These are also called as type I, II, III, and IV starches. RS1 represents starch that is resistant because it is in a physically inaccessible form such as partly milled grains and seeds and in some very dense types of processed starchy foods. It is measured chemically as the difference between the glucose released by the enzyme digestion of a homogenized food sample and that released from a nonhomogenized sample. RS1 is heat stable in most normal cooking operations and enables its use as an ingredient in a wide variety of conventional foods. RS2 represents starch that is in a certain granular form and resistant to enzyme digestion. It is measured chemically as the difference between the glucose released by the enzyme digestion of a boiled homogenized food sample and that from an unboiled, nonhomogenized food sample. In raw starch granules, starch is tightly packed in a radial pattern and is relatively dehydrated. This compact structure limits the accessibility of digestive enzymes, various amylases, and accounts for the resistant nature of RS2 such as, ungelatinized starch. In the diet, raw starch is consumed in foods like banana. RS1 and RS2 represent residues of starch forms, which are digested very slowly and incompletely in the small intestine. RS3 represents the most resistant starch fraction and is mainly retrograded amylose formed during cooling of gelatinized starch. Most moist-heated foods therefore contain some RS3. It is measured chemically as the fraction, which resists both dispersion by boiling and enzyme digestion. It can only be dispersed with KOH or dimethyl sulphoxide (Asp and Bjorck 1992). RS3 is entirely resistant to digestion by pancreatic amylases. Therefore, RS1 = TS (RDS + SDS) RS2 RS3 RS2 = TS (RDS + SDS) RS1 RS3

Figure 4X-ray diffraction diagrams of starches: type A (cereals), type B (legumes), and type V (swollen starch, Va: water-free, Vh : hydrated) (Galliard 1987)

RS3 = TS (RDS + SDS) - RS2 RS1 RS4 is the RS where novel chemical bonds other than -(1-4) or
3

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


Table 1Classification of types of resistant starch (RS), food sources, and factors affecting their resistance to digestion in the colon (Nugent 2005) Type of RS RS1 RS2 RS3 RS4 Description Physically protected Ungelatinized resistant granules with type B crystallinity, slowly hydrolyzed by -amylase Retrograded starch Chemically modified starches due to cross-linking with chemical reagents Food sources Whole- or partly milled grains and seeds, legumes Raw potatoes, green bananas, some legumes, high amylose corn Cooked and cooled potatoes, bread, cornflakes, food products with repeated moist heat treatment Foods in which modified starches have been used (for example, breads, cakes) Resistance minimized by Milling, chewing Food processing and cooking Processing conditions Less susceptible to digestibility in vitro

Table 2In vitro digestibility of starch in a variety of foods (BNF 1990)a Foods Flour, white Short bread Bread, white Bread, whole meal Spaghetti, white Biscuits made with 50% raw banana flour Biscuits made with 50% raw potato flour Peas, chick, canned Beans, dried, freshly cooked Beans, red kidney, canned % RDS 38 56 94 90 55 34 36 56 37 25 % SDS % RS1 %RS 2 59 43 4 8 36 27 29 24 45 8 5 11 3 38 35 Traces 15 %RS 3 Traces Traces 2 2 1 Traces Traces 14 6 60

a Values are expressed as % of the total starch present in the food.

-(1-6) are formed. Modified starches obtained by various types of chemical treatments are included in this category. Table 1 outlines a summary of the different types of RS, their classification criteria, and food sources. In vitro digestibility of starch in a variety of foods (BNF 1990) is shown in Table 2.

Structure of RS
RS1 is the physically protected form of starch found in whole grains (www.cerestarhealthand nutrition.com). Figure 5 shows microscopic view of the physically inaccessible RS1 in cell or tissue structures of partly milled grains, seeds, and vegetables.
RS2 RS1

stranded helices. A type A crystalline structure can be obtained if RS is formed in gelatinized starch stored at high temperature (that is, 100 C) for several hours (Eerlingen and others 1993a). It has a dense structure and only few water molecules in the monoclinic unit cell. Upon further retrogradation, the double helices pack in a hexagonal unit cell. The B form with hexagonal symmetry is more open. Water molecules (36 to 42 molecules per unit cell) in the B structure are located in fixed positions within a central channel formed by 6 double helices. The degree of polymerization (DP) of amylose also affects the yield of RS3; it rises with DP up to 100 and thereafter remains constant (Eerlingen and others 1993b). A minimum DP of 10 and a maximum of 100 seems to be necessary to form the double helix (Gidley and others 1995). Schematic presentation of RS3 formed in aqueous amylose solutions depicted as miscelle and lamella model is shown in Figures 7 and 8. Structural features of in vivo RS (ingestion of retrograded highamylose maize starch, complexed high amylose maize starch, bean flakes, or potato flakes) were assessed using the ileal contents of 4 humans (Faisant and others 1993). For all samples, starch fractions, which escaped digestion in the small intestine, were composed of 3 populations of -glucans with proportions differing according to the substrate. Small quantities of oligosaccharides made up the 1st population, illustrating a limitation of absorption in the small intestine. The 2nd population, the main RS, consisted of retrograded amylose of mean degree of polymerization (DPn) of about 35 glucose units with a melting temperature of 150 C and exhibiting a type B pattern. Finally high-molecular-weight semicrystalline -glucans were attributed to fragments of starch. This study showed that some potentially digestible starch could reach the colon and that crystalline fractions constituted only part of the starch that escaped digestion in the human small intestine.
RS4

In raw starch granules, starch is tightly packed in a radial pattern and is relatively dehydrated. This compact structure limits the accessibility of digestive enzymes and accounts for the resistant nature of RS2 such as, ungelatinized starch. Figure 6 shows the RS granules, that is, raw potato, banana, and high-amylose starch (www.cerestarhealthand nutrition.com).
RS3

Structure of RS4 includes structures of modified starches obtained by chemical treatments like distarch phosphate ester (Figure 9).

RS3 represents retrograded starch. Thus, in the formation of RS3, the starch granule is completely hydrated. Amylose leaches from the granules into the solution as a random coil polymer. Upon cooling, the polymer chains begin to reassociate as double helices, stabilized by hydrogen bonds (Wu and Sarko 1978). The individual strands in the helix contain 6 glucose units per turn in a 20.8 A repeat. The models for the double helices are left-handed, parallel-

Figure 5Structure of resistant starch type I (RS1)

Figure 6Structure of resistant starch type II (RS2)

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


Functionality of RS
RS has a small particle size, white appearance, and bland flavor. RS also has a low water-holding capacity. It has desirable physicochemical properties (Fausto and others 1997) such as swelling, viscosity increase, gel formation, and water-binding capacity, making it useful in a variety of foods. These properties make it possible to use most resistant starches to replace flour on a 1-for1 basis without significantly affecting dough handling or rheology. RS not only fortifies fiber but also imparts special characteristics not otherwise attainable in high-fiber foods (Tharanathan and Mahadevamma 2003). The functional properties and advantages of commercial sources of RS 2 and RS3 (Nugent 2005) have been summarized as follows. They are natural sources, bland in flavor, white in color, with fine particle size (which causes less interference with texture). They have high gelatinization temperature, good extrusion and film-forming qualities, and lower water-holding properties than traditional fiber products. They allow the formation of low-bulk high-fiber products with improved texture, appearance, and mouth feel (such as better organoleptic qualities) compared with traditional high-fiber products; they increase coating crispness of products and the bowl life of breakfast cereals. They are functional food ingredients lowering the calorific value of foods and useful in products for coeliacs, as bulk laxatives and in products for oral rehydration therapy. Some of these properties of RS have been successfully used in a range of baked and extruded products as described subsequently.
RS in bread-making for DF fortification (www.foodinnovation.com/functresist.pdf)

The physical properties of RS, particularly its low water-holding capacity, allow it to be a functional ingredient that provides good handling in processing and crispness, expansion, and improved texture in the final product. Bread is commonly fortified with dietary fiber. However, dark color, reduced loaf volume, poor mouthfeel, and masking of flavor are all negative attributes that are often associated with high-fiber breads. A study was conducted at the American Inst. of Baking (AIB) to evaluate the effect of RS on bread characteristics and to compare their performance to traditional fibers. The study included cellulose, oat fiber, wheat fiber, and 2 commercially available RS with 23% (Hylon VII starch) and 40% TDF (Novelose 240 starch) and a blend of oat fiber with Novelose 240 starch in a 50/50 ratio based on TDF contribution. Compared with oat fiber, cellulose, and wheat fiber, both RS had a lower water-holding capacity, which is similar to that of flour. Although the water-holding capacity of the RS was lower than that of the other fiber sources, the quantity required to obtain the same level of TDF was greater. This raised the total water requirement of the RS dough. However, absorption of the RS doughs was less compared with the fiber doughs, despite a larger quantity of water used. The lower dough absorption consequently had less impact on dough rheology and was closest to the white pan bread dough. Bread containing 40% TDF RS had greater loaf volume and better cell structure compared with traditional fibers tested (Baghurst and others 1996).

Figure 7Schematic presentation of enzyme resistant starch type III (RS3) formed in aqueous amylose solutions. Micelle model. Double helices are ordered into a crystalline structure (C) over a particular region of the chain, interspersed with amorphous, enzyme degradable regions.

Figure 8Schematic presentation of enzyme-resistant starch type III formed in aqueous amylose solutions. Lamella model. Lamellar structures are formed by folding of the polymer chains. The fold zones are amorphous (A), while the center of the lamella is crystalline (C).

Figure 9Preparation of cross-bonded starch

Figure 10Behavior of amylose molecules during cooling of a concentrated aqueous solution Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY 5

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


RS as a texture modifier in baked goods

One way to ensure that the general population receives adequate amounts of fiber in the diet is to fortify good-tasting foods that normally do not come to mind with fiber fortification but are often eaten as breakfast items or snacks. RS were incorporated in a variety of baked goods, many of which include batter systems, such as in cakes, cake-like muffins, or brownies. In general, application tests showed that RS acts as texture modifier, imparting a favorable tenderness to the crumb. A low-fat, loaf cake was formulated with RS and various fibers to obtain approximately 3% TDF or 2.5 g of fiber per 80 g serving. These included a 40% TDF RS (Novelose 240 starch), oat fiber, a blend of oat fiber with Novelose 240 starch in a 50/50 ratio based on TDF contribution, and a 23% TDF RS (Hylon VII starch). The baked cakes made with RS were similar to that containing oat fiber and the control in the amount of moisture loss after baking, height, specific volume, and density. A panel rated the 40% TDF RS loaf cakes as the best for flavor, grittiness, moisture perception, and tenderness 24 h after baking.
RS as a crisping agent

Among other functional properties, RS can be used as an ingredient that improves crispness in foods where high heat is applied to a products surface during processing. French toast and waffles, especially frozen reheated types, represent foods in which surface crispness is desired. Tests were conducted to compare the functionality of RS and various fibers in a buttermilk waffle formulation. Based on the evaluation of the toasted waffles for initial crispness, crispness after 3 min, moistness, and overall texture by a trained sensory panel RS waffle indicated greater crispness than control or traditional fiber.
RS as a functional ingredient in other foods

Along with textural enhancement, RS can improve expansion in extruded cereals and snacks. Various cereals were formulated to contain 40% TDF RS (Novelose 240 starch) alone and in combination with oat fiber in ratios of 50/50 and 25/75 based on weight. The cereal with RS and no oat fiber had greater volumetric expansion than the control. In blends with oat fiber, the cereal containing 75% of RS had better expansion than the one containing only 50%. Dried pasta products containing up to 15% RS can be made with little or no effect on dough rheology during extrusion. Although the resultant pasta was lighter in color, a firm al dente texture was obtained in the same cooking time as a control that had no added fiber. RS may also be used in thickened, opaque health beverages in which insoluble fiber is desired. Insoluble fibers generally require suspension and add opacity to beverages. Compared with insoluble fibers, RS imparts a less gritty mouthfeel and masks flavors less.

Factors influencing the formation of RS


Several factors influence the formation of RS.
Inherent properties of starch

Crystallinity of starch. One of the causes of resistance to enzymes is the crystallinity of native type B starch granules as observed in the case of amylomaize starch and also the encapsulation of starch within plant cell or tissue structures. X-ray diffraction and differential scanning calorimetry studies on crystalline residues from amylomaize starch samples have suggested that chain fragments packed in a type B crystalline structure with a slightly enlarged crystal lattice contribute to formation of RS from amylomaize starch. Any treatment that eliminates starch crystallinity (that is, gelatinization) or the integrity of the plant cell or tissue structure (that is, milling) increases enzyme availability and reduces the content of RS, whereas recrystallization and chemical mod6

ifications tend to increase the RS. The modified food starches are partially resistant to enzymes as a result of chemical modifications induced intentionally (Englyst and Cummings 1986; Bjorck and others 1989; Schweizer and others 1990). Besides these, the cellular structure of plant foods influences the digestibility of starch in the small intestine as well as the intrinsic digestibility of a particular physical form of starch. Granular structure. A large variability in susceptibility to amylases shown by raw starch granules also influences RS formation. Potato starch and high amylose maize starch are known to be very resistant in vitro and incompletely absorbed in vivo, whereas most cereal starches are slowly but virtually completely digested and absorbed in vivo (Holm and others 1987). The smaller surface-to-volume ratio of the large potato granules is probably important. The nature of the granule surface also needs to be considered; an adsorbed layer of non-starch material would effectively impede the action of the enzyme (Ring and others 1988). Raw tepary starch is found to be more resistant to hydrolysis than maize starch, perhaps due to differences in granule structure and amylose content (Abbas and others 1987). Amylose:amylopectin ratio. A higher content of amylose lowers the digestibility of starch due to positive correlation between amylose content and formation of RS (Berry 1986; Sievert and Pomeranz 1989b). The importance of the amylose:amylopectin ratio in the postprandial glycaemic and insulinaemic responses to corn was studied in commonly consumed corn products (Granfeldt and others 1995). The meals containing high amylose (70%) corn flour had an RS of 20 g/100 g DM than that containing ordinary corn flour (25% amylose) that had RS of 3 g /100 g DM. Retrogradation of amylose. When heated to about 50 C, in the presence of water, the amylose in the granule swells; the crystalline structure of the amylopectin disintegrates and the granule ruptures. The polysaccharide chains take up a random configuration, causing swelling of the starch and thickening of the surrounding matrix such as, gelatinizationa process that renders the starch easily digestible. On cooling/drying, recrystallization (retrogradation) occurs. This takes place very fast for the amylose moiety as the linear structure facilitates cross linkages by means of hydrogen bonds. Figure 10 shows the formation of gel and micelle on cooling of a concentrated solution of amylose (Belitz and Grosch 1999). The branched nature of amylopectin inhibits its recrystallization to some extent and it takes place over several days. Retrograded amylose in peas, maize, wheat, and potatoes was found to be highly resistant to amylolysis (Ring and others 1988). The rate and extent to which a starch may retrograde after gelatinization essentially depends on the amount of amylose present. Repeated autoclaving of wheat starch may generate up to 10% RS. The level obtained appeared to be strongly related to the amylose content, and the retrogradation of amylose was identified as the main mechanism for the formation of RS that can be generated in larger amounts by repeated autoclaving (Berry 1986; Bjorck and others 1990). During storage, the dispersed polymers of gelatinized starch are said to undergo retrogradation to semicrystalline forms that resist digestion by pancreatic -amylase. It forms a major portion of RS in wheat bread and corn flakes (Englyst and Cummings 1985), whereas only 25% of the RS in cooked, cooled potatoes can be accounted for as retrograded amylose. The digestibility of legume starch is much lower than that of cereal starch, which is attributable to higher content of amylose in the former. The digestibility of high amylose cereal starch is reported to be significantly lower (Tharanathan and Mahadevamma 2003). Native high-amylose starch is known to be high in type II RS (RS 2) (Berry 1986), which is defined as starch in its native granular state that is resistant to digestion in the small intestine. This after cooking and cooling gives high yields of type III RS (Berry 1986; Sievert and Pomeranz 1989a) or retrograded starch (Englyst and

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


others 1992). Heating of RS preparation from amylomaize VII resulted in broad endothermic transition, which is ascribed to melting of amylose crystallites (Sievert and Pomeranz 1989, 1990). Exothermic transitions during controlled cooling of isolated potato amylose fractions have been attributed to amylose chain association. The formation of RS likewise has been attributed to the ordering of amylose chains (Sievert and others 1991). Based on previous studies of amylose behavior, it has been suggested that the exotherms observed during the cooling of either amylose or a thermally treated RS preparation reflect chain association, which may involve amylose aggregation and gelation dominated by formation and subsequent lateral aggregation of type B double helices in crystalline arrays (Gidley 1989; Gidley and Bulpin 1989; Sievert and Wursch 1993). Gelatinized waxy corn starch stored at varying temperatures from 6 C to 60 C for 1 to 29 d also showed reduced enzyme susceptibility to pancreatic -amylase and amyloglucosidase (Eerlingen and others 1994). Influence of amylose chain length. Influence of amylose chain length on enzyme RS formation was studied by Eerlingen and others (1993b) by hydrolyzing potato starch amylose to varying degrees by incubation with barley -amylase for different periods, and monitored by measuring the number of average chain lengths or degree of polymerization (DPn). The DPn of RS varied between 19 and 26 and was independent of the chain length of the amylose (DP n 40 to 610) from which it was formed. Results suggested that RS might be formed by aggregation of amylose helices in a crystalline -type structure over a particular region of the chain (about 24 glucose units). Linearization of amylopectin. Linearization of amylopectin occurs during the long low-temperature baking process due to the prolonged activity of intrinsic amylases in the dough, and is prominent in the presence of certain organic acids that is, in bread products baked with added lactic acid (Liljeberg and others 1996). It has been reported to significantly increase RS formation during wet-autoclaving (Berry 1986).
Heat and moisture

subjected to dry and wet heat treatment brought out higher RS contents in foods subjected to dry heat treatment compared with wet processed ones. Sorghum, green gram dhal, and green plantain showed highest RS content (5.51%, 5.81%, and 10.7%, respectively) (Platel and Shurpalekar 1994).
Interaction of starch with other components

Water content is an important factor that affects formation of RS. Repeated heat/moisture treatment is associated with a decrease in the hydrolysis limit of pancreatic -amylase and increased formation of RS. Maximum RS yield was obtained at a starch:water ratio of 1: 3.5 (w/w) (Sievert and Pomeranz 1989b) and a heat treatment at 18% moisture gave increased levels of the degree of crystallinity of normal and waxy starches and thus reduced enzyme susceptibility. However, at 27% moisture, starch degradation to some extent made areas of starch more accessible to enzyme attack. Thus, proper heat treatment could be used as a method of preparation of RS (Franco and others 1995). In addition, higher temperature and less water results in type A configuration, whereas lower temperature and high water content results in type B configuration (Wu and Sarko 1978). Malshick-Shin and others (2003) determined solubility, water vapor sorption, and swelling characteristics for RS prepared from wheat starch and linterized wheat starch by autoclaving and cooling and by cross-linking The experimental RS made from wheat starch contained 10% to 73% RS versus 58% and 40% in commercial sources, Novelose 240 and 330 respectively, produced from high-amylose maize (corn) starch. In excess water, the experimental RS starches (except for the cross-linked wheat starch) gained 3 to 6 times more water than the commercial RS starches at 25 C, and 2 to 4 times more at 95 C. All starches showed similar water vapor sorption and desorption isotherms at 25 C and aw < 0.8. At aw 0.84 to 0.97, the RS made from wheat starch (except cross-linked wheat starch) showed approximately 10% higher water sorption than the commercial RS. RS determined in several selected cereals, legumes, and tubers

Interactions of starch with different components present in the food system are known to influence the formation of RS as follows. Protein. Starch-protein interaction has been believed to reduce RS contents as observed in case of potato starch and added albumin when autoclaved and subsequently cooled at 20 C (Escarpa and others1997). Dietary fiber. Insoluble dietary fiber constituents such as cellulose and lignin have been shown to have minimal effects on RS yields compared with other constituents such as potassium and calcium ions and catechin (Escarpa and others 1997). Enzyme inhibitors. Polyphenols, phytic acid, and lectins present mainly in leguminous seeds, have been reported to inhibit in vitro starch hydrolysis and to lower the glycemic index (Thompson and Yoon 1984). Tannic acid significantly inhibits both amylases and intestinal maltase activity (Bjorck and Nyman 1987). Indigestible residues from black beans (Phaseolus vulgaris cv. Tacari gua), green beans (P. vulgaris), carrots (Daucus carota), and rice bran (Oryza sativa) are all reported to inhibit pancreatic amylase in vitro (Moron and others 1989). Since amylolysis is inhibited by phytic acid, a decrease in phytate content increases starch digestibility (Thompson and Yoon 1984). Contradictory information exists in the literature on this aspect. The autoclaving and subsequent cooling of potato starch and catechin was found to significantly reduce the yields of RS, whereas the addition of phytic acid to potato starch reduced the RS contents to a minor extent (Escarpa and others 1997) compared with the RS formed from potato starch with no added constituent. The reasons for the same are still not clear. Ions. The yields of RS in potato starch gels decrease in the presence of calcium and potassium ions compared with those with no added constituent (Escarpa and others 1997), presumably due to the prevention of formation of hydrogen bonds between amylose and amylopectin chains caused by adsorption of these ions. Sugars. The addition of soluble sugars such as glucose, maltose, sucrose, and ribose has been found to reduce the level of crystallization and subsequently reduce the yields of RS (Buch and Walker 1988; IAnson and others 1990; Kohyama and Nishinari 1991). The mechanism of retrogradation inhibition was considered as the interaction between sugar molecules and the starch molecular chains, which change the matrix of gelatinized starch (the sugars act as anti-plasticizers and increase the glass transition temperature). The role of sugars on the formation of RS in starch gels (RS type III) was studied by Eerlingen and others (1994). Sugars influenced the RS levels in starch gels only when added in high concentration (final starch-water-sugar ratio of 1:10:5 w/w). In wheat starch gels, the RS yields decreased from approximately 3.4% to 2.8% in the presence of sucrose or glucose, and to 2.5% in the presence of ribose or maltose. An increase in RS yield was observed with high-amylose corn starch. The experiments showed that the differences in gelatinization temperature, lipid content, and apparent amylose content of the 2 starches were not the main causes of the different impact of sugars on RS yields. Lipids, emulsifiers. In a study, amylomaize VII starch, autoclaved at 125 C, was reacted during cooling below 100 C with lysophosphatidyl choline (LPC), sodium stearoyl lactylate (SSL), and hydroxylated lecithin (OHL) (Czuchajowska and others 1991). Differential scanning calorimetry (DSC) peaks at around
7

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


95 C to 110 C indicated formation of amylose-lipid complexes, and at about 155 C indicated the presence of enzyme-resistant starch (RS). Yields of RS from complexed samples isolated by thermostable bacterial -amylase or amyloglucosidase were lower than yields of RS from the autoclaved and cooled control. Formation of complexes competes with amylose chains involved in generation of RS. Amylose-lipid complexes are enzyme-degradable, and an increase in complexed amylose reduced yields of RS. Amylose recrystallization in RS formation is competitively affected by complexation of amylose with LPC and SSL. Results of X-ray diffraction powder crystallography were in agreement with DSC measurements. Complexes of amylose with LPC, SSL, and OHL gave type V patterns; enzymic hydrolysis of the complexes yielded type B RS structures. However, the viewpoint differs among scientists working in this area. While some workers believe amyloselipid complex to reduce the formation of RS, others believe the amylose-lipid complex itself to be a form of RS. From studies on isolated barley starch autoclaved with sodium stearoyl lactylate (SSL), distilled monoglycerides, diacetyl tartaric acid esters of mono-diglycerides (DATEM), and ethoxylated monoglycerides (bakery additives), it is postulated that amylose crystallization (as measured by enthalpies of the 158 C endotherm) that is involved in the formation of RS is competitively affected by its complexation with lipids (Szczodrak and Pomeranz 1992). Effect of citric acid and 2 emulsifiers SSL and DATEM on formation and structure of RS during extrusion of cornstarch and guar gum were studied (Adamu 2001). X-ray diffraction of the extruded starches gave a V-diffraction pattern indicating the formation of amylose-lipid complexes. Purification of the isolated RS by size exclusion high-performance liquid chromatography (HPLC) indicated the additives to have a substantial effect on molecular weight; DATEM and SSL increased molecular weight of RS, while citric acid decreased it. The influence of endogenous lipids on the formation of RS from wheat starch showed defatting to decrease the RS content (Eerlingen and others 1994). When SDS was added to defatted wheat or amylomaize VII starch, RS yields decreased substantially. X-ray diffraction and DSC showed that amylose-lipid complexes were formed in the presence of both endogenous and added lipids (SDS). A similar behavior has been reported on addition of lipids (Eliasoson and others 1988) such as olive oil (Escarpa and others 1997). Thus, less amylose was available for interactions leading to the formation of double helices and RS. Adding SDS to the starch also caused a difference in RS quality. Amylose-lipid complexes can also be formed during food processing (autoclaving and cooling). Lecithin, palmitic acid, oleic acid, and soya bean oil affect retrogradation to a lower extent than monoglycerides. Nevertheless, these authors found that pure potato amylose and oleic acid formed complexes highly resistant to amylolysis (Mercier 1980).
Processing conditions

of RS. Cooking under conditions of high moisture and temperature can significantly lower the RS content by disrupting crystalline structure. Increasing the levels of RS can be done in other conditions, such as extrusion followed by cooling to induce crystallization (Haralampu 2000). The RS contents in various processed food samples have been reported (Siljestrom and Bjorck 1990; Parchure and Kulkarni 1997; Kavita and others 1998).
Thermal processing

Processing techniques may affect both the gelatinization and retrogradation processes, influencing RS formation. This fact is of great importance for the food industry since it offers the possibility of increasing the RS content of processed foods and foodstuffs. Baking, pasta production, extrusion cooking, autoclaving, and so forth are known to influence the yield of RS in foods (Siljestrom and Asp 1985; Bjorck and Nyman 1987; Siljestrom and others 1989; Muir and ODea 1992; Rabe and Sievert 1992). Highly processed cereal flours and foods made from the flours, such as pasta, contain much lower levels of RS, averaging only about 1.5% to 8% RS on a dry basis. Since the crystalline structure of starch in legumes (type C) is more stable compared with the crystal structure in cereal grains (type A) (Ring and others 1988), processing cereal grains results in a large decrease in RS content, while legumes are excellent sources
8

Steam cooking. Steam cooking helps in production of RS. Starches isolated from several steam-heated legumes were rich in indigestible RS (19% to 31%, DM basis), which was not observed in raw beans (Tovar and Melito 1996). Similarly, RS measured directly in conventionally and high-pressure steamed beans were 3 to 5 times higher than in the raw pulses, suggesting retrogradation to be mainly responsible for the reduction in digestibility. Prolonged steaming as well as short dry pressure heating decreased the enzymically assessed total starch content of whole beans by 2% to 3% (DM basis), indicating that these treatments may induce formation of other types of indigestible starch (Tovar and Melito 1996). Autoclaving. Autoclaving results in increase in RS. Autoclaved wheat starch has 9% RS compared with less than 1% in uncooked wheat starch (Siljestrom and Asp 1985). Autoclaved wheat starch contained 6.2% RS (of dm); this increased to 7.8% after 3 further reboiling/cooling cycles (Bjorck and others 1987). Quantitative and qualitative influence of incubation time and temperature of autoclaved starch on RS formation was studied by Eerlingen and others (1993a). In another study, white flour subjected to repeated autoclaving and cooling cycles showed an increase in total dietary fiber >3 times that of bread flours and 4 times that of pastry flours (Ranhotra and others 1991b). The increase was primarily due to the formation of RS. Investigations on the formation of enzyme-resistant starch (RS) during autoclaving and cooling by Sievert and Pomeranz (1989a) showed highest yield (21.3%) to be obtained from amylomaize VII starch (70% amylose). Formation of RS in amylomaize VII starch was affected by the starch/water ratio, autoclaving temperature and number of autoclaving-cooling cycles. The number of cycles exerted the most pronounced effect on RS; increasing the number of cycles to 20 raised RS level to >40%. Furthermore, the thermoanalytical data suggested that amylose-lipid complexes were not involved in the formation of RS. Yields in excess of 20% RS can be obtained from autoclaved amylomaize starch containing 70% amylose. They can be raised to levels of 40% by increasing the number of autoclaving-cooling cycles up to 20 (Eerlingen and Delcour 1995). The extent of RS formation in commercially available autoclaved corn, potato, and leguminous products and in autoclaved purees intended for consumption by infants aged 3 to 8 mo was investigated by Siljestrom and Bjorck (1990). RS levels found (g/ 100 g DM) were as follows: 0.8 to 2.4 in purees, 0.2 to 3.2 in canned legumes, 1.9 in canned potatoes, 0.5 and 0.9 in reconstituted dried potatoes, <0.1 in canned corn, and 1.1 in cornflakes. Native starch (NS) extracted from wheat and subjected to 5 autoclaving and cooling cycles was found to contain 11.5% RS, which was measured as insoluble fiber; NS contained 0.5% RS (Ranhotra and others 1991a). Heat-moisture treatment (autoclaving at 121 C) with subsequent cooling was used to produce amylase-resistant starch (RS) from purified high-amylose starch samples. The formation of RS in barley starch was strongly affected by the number of autoclaving-cooling cycles; increasing the number of cycles from 1 to 20 raised the RS yield from 6% to 26% (Szczodrak and Pomeranz 1991). Parboiling. Parboiling increases RS production. In studies on 5 rice varieties, differing in amylose content, the in vitro and in vivo RS levels were low and positively correlated with amylose content (Eggum and others 1993). Higher RS starch levels were found in

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


cooked and parboiled-cooked rice than in raw rice; waxy rice had very low values. Higher contents of RS have been reported in parboiled rice than raw white rice, which also increased by cooling or freezing (Marsono and Topping 1999). Baking. Baking increases RS content. In a study to evaluate the effect of baking on RS formation, white bread was baked and divided into 3 fractions (crumb, inner crust, and outer crust) (Westerlund and others 1989). Starch levels were found to be highest in dough and lowest in outer crust after baking for 35 min. RS levels were lowest in dough and highest in crumb after baking for 35 min. A low-temperature, long-time baked product contained significantly higher amounts of RS than bread baked under ordinary conditions (Liljeberg and others 1996). Addition of lactic acid increased RS recovery further whereas malt had no impact on RS yield. The highest level of RS was noted in long-time baked bread based on highamylose barley flour. RS isolated from wheat-based foods such as chapatti and phulka was structurally characterized as a linear 1, 4linked -D-glucan essentially derived from retrograded amylose fraction, which was dependent on the severity of the processing treatments as well as the levels of gluten and damaged starch in the wheat flour (Tharanathan and Tharanathan 2001). Extrusion cooking. Effect of extrusion cooking, at different temperatures (90, 100, 120, 140, or 160 C), moisture contents (20%, 25%, 30%, 35%, or 40%) and screw speeds (60, 80, or 100 rpm), was investigated on the formation of RS of type 3 (RS3) in hull-less barley flours from CDC-Candle (waxy) and Phoenix (regular). The RS3 content of the native flours, in general, decreased by extrusion cooking, but not significantly. Storage of extruded flour samples at 4 C for 24 h before oven drying slightly increased RS3 content (Faraj and others 2004). With pearl barley used as the primary material in tests designed to optimize the production of RS by extrusion an extrusion temperature of 150 C and a barley moisture content from 17.5% and 22.5% moisture, followed by cold storage at 18 C gave the best results (Gebhardt and others 2001). Corn starches with and without guar gum [10% (w/w)] and 2% (w/w) of diacetyl tartaric acid ester of monoglyceride, sodium stearoyl-2-lactylate or citric acid, respectively, were extrusioncooked in a twin-screw extruder at 18% moisture, 150 C, and 180 rpm screw speed (Adamu 2001). The formation of RS in extruded corn starch was found to be strongly affected by the addition of gum and the different food additives. X-ray diffraction of the extruded starches gave a V diffraction pattern indicating the effect of extrusion cooking and amylose-lipid complexes. Enzymatic digestion did not affect the V structure, which could apparently be attributed to extrusion cooking. Purification of the isolated RS by size exclusion-HPLC showed a dependence of molecular weight on the added additives. Results of differential scanning calorimetry and X-ray diffraction suggest that amylose-lipid complexes could also be involved in the formation of RS in extruded cornstarch. Pyroconversion. Pyroconversion of starch increases RS content. Lima bean (Phaseolus lunatus) starch was modified using pyroconversion, the optimum product being recovered from native starch treated with a 160:1 starch/HCl ratio, at 90 C for 1 h, resulting in starches containing 49.5% indigestible starch (Tester and others 2004). Starch pyrodextrinization decreased the amount of enzymically available starch through formation of atypical glycosidic bonds that are not digested by the amylases and maltooligosaccharidases in the small intestine of humans. Microwave irradiation. Microwave irradiation improves the digestibility of tuber starches, which could be accompanied by physicochemical and structure changes (www.man.poznan.pl). Microwave cooking of legumes such as chickpeas and common beans produced a redistribution of the insoluble nonstarch polysaccharides to soluble fraction, although the total nonstarch polysaccharides were not affected. This was evaluated by assessing the physicochemical, nutritional, and microstructural modifications in starch and nonstarch polysaccharides (Marconi and others 2000). The RS level decreased from 32.5% of total starch in raw chickpeas and beans, respectively, to about 10% in cooked samples with a concomitant increase in the level of rapidly digestible starch from 35.6% and 27.5% to about 80%. Studies on effects of different heat treatments (cooking, microwave cooking, pressure cooking) on the rate of hydrolysis, hydrolysis index, and glycaemic index values of kudzu starch and cornstarch showed increase in digestible starch and decrease in RS following heat treatment. The rate of hydrolysis of kudzu starch and cornstarch increased following heat treatment, especially after microwaving (Geng and others 2003).
Miscellaneous treatments.

Milling. Leguminous seeds, in which cell structures are preserved after cooking (that is, bread with whole seeds); bean flour with intact cells (Schweizer and others 1990); foods containing large particles such as bread with whole seeds have lower physical accessibility of starch to amylase action, and thereby contribute to higher RS contents. In some foods, physically inaccessible starch is likely to be an important fraction of the total starch that is resistant to digestion in vivo. Schweizer and others (1990) found evidence of 20% starch malabsorption from a diet containing bean flour with intact cells. About half of the malabsorbed starch was retrograded amylose. The precooked flours (PCF) prepared from dried lentils and beans, rich in intact cells filled with starch granules, indicated that they contained important quantities of RS, such as retrograded amylose (3% to 9%, DM). Germination. Germination is shown to decrease the RS content in bengal gram, field beans, cow pea, and green gram (Kavita and others 1998). Fermentation. Fermentation reduces RS content. Flour from sorghum cv. Tabat was mixed with water and previously fermented dough starter, and fermented at 37 C for a maximum of 36 h showed an increase in the in vitro starch digestibility and a decrease in the content of RS and total starch (Abd-Elmoneim and others 2004). RS formation has also been shown to decrease in the fermented products, idlis and dhoklas (Kavita and others 1998).
Storage conditions

Generally, RS increases on storage, especially low-temperature storage. Cold storage seems to support an increase in RS content. Whole corn bread and corn bread crumb, when stored at different temperatures (20 C, 4 C, or 20 C) for 7 d showed RS contents to reach a maximum between 2 and 4 d at all storage temperatures, after which they decreased (Niba 2003). Lowest RS levels in whole corn bread were found after storage at 20 C (2.18 g/100 g) for 7 d. A comparison of masa and fresh and stored tortillas from common and Costeno corn varieties showed that Costeno had higher digestible starch and total dietary fiber contents than its common counterpart. During storage of both types of tortilla, digestible starch contents decreased, whereas those of RS starch increased (Mora-Escobedo and others 2004). Studies on the influence of cold storage on in vitro starch digestibility of tortillas showed a decrease in available starch content in tortillas after 48 h of cold storage, which was concomitant with increased total RS levels (Agama-Acevedo and others 2004). These changes were due mainly to retrogradation, as indicated by increased retrograded resistant starch (RRS) levels which accounted for the major portion of the total RS. Although amylolysis patterns for fresh and 72 h stored tortillas were similar, lower digestion rates were observed for stored samples. RS contents of gelatinized samples of corn, ragi, rice, sago, and potato flours increased on low-temperature storage and de9

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


creased on reheating the samples. Cooked food samples of rice, unleavened bread, potato, Bengal gram, and green gram also showed increased RS formation on storage. There was no significant increase in RS content of stored decorticated legumes (bengal gram, green gram, and red gram), whereas horse gram and lentils showed lesser RS on storage in comparison with fresh samples. Increase in RS was reported for gelatinized samples of corn, wheat, ragi, rice, sago, and potato flours on low-temperature storage. A 27% increase in RS of cooked rice was observed when stored at 4 C (Johansson and Siljestrom 1984). The longer the duration of storage of gelatinized wheat flour, the greater was the formation of RS (Kavita and others 1998). Rice stored at 20 C retrograded more than rice stored in the refrigerator (Mitsuda 1993). ture treatment tended to increase yield of boiling-stable granular RS to the maximum of 63.2%. Selective heat treatment of high amylose starch in the presence of agents inhibiting the swelling of starch like alkali and alkaline earth metal salts of halides, sulfates, and phosphates yield granular RS with high dietary fiber. Recently, pyrodextrinization has been recognized as a way of producing a RS that is water-soluble and has non-starch linkages (Laurentin and Edwards 2004). Pyroconversion refers strictly to the modification of dry starch through heat treatments, with or without addition of acids. Acids used include hydrochloric acid at 0.15% (based on starch dry weight) and orthophosphoric or sulfuric acids at 0.17% (Wurzburg 1995). Commercial pyrodextrins are generally produced by heating dry, acidified starch in a reactor with agitation. Acid may be sprayed on the starch to facilitate hydrolysis and transglycosidation. Depending on reaction conditions, pyroconversion produces a range of products that vary in digestibility, available starch, viscosity, cold-water solubility, swelling power, color, and stability (Ohkuma and Wakabayashi 2001). The production of indigestible dextrins or pyrodextrins by heat-treating potato starch in the presence of an acid and then refining the product has been described (Ohkuma and others 1994, 1995).
Enzymic treatment

Preparation of RS
RS can be prepared by using heat treatment, enzyme treatment, combined heat treatment and enzyme treatment, and chemical treatment.
Heat treatment

Heat treatment of starch to various extents leads to formation of RS. RS can be obtained by cooking the starch above the gelatinization temperature and simultaneously drying on heated rolls like drum driers or even extruders. The gelatinization of starch granules by heat processing strongly influences their susceptibility to enzymatic hydrolysis. In a high-moisture environment, amylose leaches from the granules, increasing the solubility of starch and thereby its susceptibility (Holm and others 1988). Good yields of RS can be obtained by gelatinizing starch at 120 C for 20 min, followed by cooling to room temperature (Garcia-Alonso and others 1999). The starch gels are then frozen overnight at 20 C and dried at 60 C before milling. Many combinations of time and temperature treatments have been used to make type III RS from various sources of native starch. Even for starches with normal amylose levels, it is recognized that cooking at >100 C can increase the yield of type III RS. The temperature treatments have included autoclaving the starch at 110 C (Berry 1986), at 121 C (Berry 1986; Bjorck and others 1987; Sievert and Pomeranz 1989; Sievert and Wursch 1993b), at 127 C (Berry 1986; Bjorck and others 1987), at 134 C (Berry 1986; Bjorck and others 1987; Russell and others 1989; Sievert and Pomeranz 1989), or at 148 C (Sievert and Pomeranz 1989) for periods ranging from 30 min to 1 h. An enzyme-RS type III, which has a melting point or endothermic peak of at least about 140 C, as determined by differential scanning calorimetry (DSC) can be produced in yields of at least 25% by weight, based on the weight of the original starch ingredient (Haynes and others 2000). A gelatinization stage, nucleation/propagation stage, and preferably a heat-treatment stage are required to produce reduced calorie starch-based compositions that contain the enzyme-resistant starch. It is produced using crystal nucleation and propagation temperatures, which avoid substantial production of lower melting amylopectin crystals, lower melting amylose crystals, and lower melting amylose-lipid complexes. The nucleating temperature used is above the melting point of the amylopectin crystals. The propagating temperature used is above the melting point of any amylose-lipid complexes but below the melting point of the enzyme RS. The high melting point of the enzyme RS permits its use in baked good formulations. Partial acid hydrolysis (PAH) of a high-amylose corn starch (aeVII) enhances the effects of hydrothermal treatments used to produce granular RS, which is stable to further heat treatment at atmospheric pressure (Brumovsky and Thompson 2001). PAH of ae-VII starch involved heating 35% (w/v) starch suspensions with 1% (w/w) HCl at 25 C for up to 78 h. PAH followed by heat mois10

The possibility of preparing a RS concentrate from isolated pea starch was investigated, and sorption of hydrophobic substances (indicative of health-benefiting properties) by such a concentrate was studied by Soral and Wronkowska (2000). By use of a thermally stable -amylase, a preparation of up to 70% RS containing a mixture of mineral and organic N compounds was obtained. The pea RS concentrate had an affinity to bile acid, deoxycholic, and cholesterol; however, its affinity to cholesterol was not as efficient as that of native pea starch. The results concluded that the pea RS concentrate may be potentially used as a food component in special diets, or for preventive, prophylactic, and therapeutic purposes. Readily fermentable heat-stable RS of optimal chain length from poly-1,4- -D-glucan useful in various functional foods can be obtained by in vitro synthesis by adding an enzyme extract containing the amylosucrase of Neisseria polysaccharea to sucrose solutions, followed by incubation at 37 C over several hours (Buttcher and others 1997). A method has been discovered to produce an RS product that retains the same cooking quality as found in untreated rice starch or flour, but has a higher percentage of starch resistant to -amylase digestion (King and Tan 2005). This method uses a debranching enzyme, that is, pullulanase, to digest the starch, but does not require pretreating the starch source before enzymatic treatment. This method produced RS from low amylose starches, rice starch (24%), and rice flour (20%). Surprisingly the RS product formed by this method retained the pasting characteristics of the untreated flour or starch and was heat stable. This method may also be used to produce RS from other botanical sources, that is, corn, wheat, potato, oat, barley, tapioca, sago, and arrowroot.
Heat and enzyme treatment

Preparation of RS to be used as a food-grade bulking agent, by retrogradation of starch followed by enzymatic or chemical hydrolysis to reduce or remove the amorphous regions of retrograded starch (Iyengar 1991). RS can be prepared from high amylose starch by gelatinization followed by treating the slurry with debranching enzymes like pullulanase and isolating the starch product by drying/extrusion. Controlled heat treatment of starch so as to achieve swelling and at the same time retain its granular structure followed by enzymatic debranching (Haralampu and Gross 1998) and annealing at suitable temperature followed by drying produces RS. These RS find applications in a variety of foods and

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


beverage products. Purified RS products having at least 50% RS content can be produced by forming a water-starch suspension wherein the ratio of starch to water is approximately 1:2 to 1:20, heating the waterstarch suspension in an autoclave at temperatures above 100 C. to ensure full starch gelatinization and then cooling to allow amylose retrogradation to take place. It is reported that best results were obtained at a temperature of 134 C, with 4 heating and cooling cycles and a starch:water ratio of 1:3.5. The RS was purified by comminuting the starch gel and mixing it with an amylase to digest non-RS fractions, leaving RS. The amylase is inactivated by heat treatment above 100 C (Pomeranz and Sievert 1990). For the preparation of a fragmented starch precipitate for use in reduced-fat foods, a debranched amylopectin starch is precipitated and then fragmented. The debranched amylopectin starch may be derived from a starch that contains amylopectin, for example, common corn starch and waxy maize starch, by gelatinizing the starch, followed by treatment with a debranching enzyme, such as isoamylase or pullulanase, and precipitation of the debranched starch. To form the precipitate, the solution is cooled to ambient temperature, to reduce the solubility of the debranched starch. The precipitate may then be heated to about 70 C, while in contact with a liquid medium, to dissolve at least a portion of the precipitate. Reprecipitation by cooling of the suspension/solution may then be employed. Repetition of the dissolving and the reprecipitation tends to improve the temperature stability of the resulting aqueous dispersion as was observed on repeating the cycle of heating and cooling, a total of 8 times (Harris and others 1994, 1995). A process for increasing the amount of amylase-RS (to a minimum of 15%) in high amylose starch, such as Hylon V or Hylon VII consisted essentially of gelatinization of a starch slurry, enzymic debranching of the starch, and isolation of the starch product by extrusion or drying. A further increase in amylase-resistant starch was obtained by addition of an inorganic salt to debranched starch before isolation (Chiu and others 1994).
Chemical treatment

ered the most reproducible and repeatable measurement of RS in starch and plant materials, but it has not been shown to analyze all RS as defined (Champ and others 2003). It is based on the principle of enzymic digestion and measures the portions of starch resistant to digestion at 37 C that are typically not quantitated due to the gelatization at 100 C followed by digestion at 60 C. Two general methods specifically proposed to determine RS (Berry 1986; Englyst and others 1992) remove digestible starch using different amylases, and the residual fraction is quantified after solubilization in 2M KOH. The Siljestrom and Asp (Siljestrom and Asp 1985) procedure includes preparation and quantification of dietary fiber residue before RS determination. This is usually done by drying the samples at 105 C. As heating influences the RS content in foods, results may be modified by this step. A modified method for measuring RS in dietary fiber residues from various sources developed by Saura-Calixto and others (1993) involves mixing fiber residues with KOH, acetate buffer, and HCl. After incubation with amyloglucosidase samples are centrifuged and diluted with distilled water. RS is calculated as glucose (mg) 0.9. Advantages of the method are the use of small amount of sample, less reagents and elimination of drying.
In vivo methods

In type IV RS, the enzyme resistance is introduced by modifying the starch by crosslinking with chemical agents (Haynes and others 2000). Crosslinked starches are obtained by the reaction of starch with bi- or polyfunctional reagents like sodium trimetaphosphate, phosphorus oxychloride, or mixed anhydrides of acetic acid and dicarboxylic acids like adipic acid. Cross-linking carried out by sulphonate and phosphate groups between various starch molecules involves their hydroxyl group thus bringing resistance to amylolytic attack on the starch molecule. Figure 9 shows the preparation of distarch phosphate ester (Hamilton and Paschall 1967). Distarch phosphates with 0.4% to 0.5% phosphorus have been prepared and they contain both slowly digested starch (SDS) and RS4 (Woo and others 1999). The modified starches were obtained in quantitative yield, and provided 13% to 69% of SDS and 18% to 87% RS4. RS4 starches with low swelling power have also been prepared similarly from wheat, corn, waxy corn, high amylose corn, oat, rice tapioca, mung bean, banana, and potato starches. Phosphated di-starch phosphate, a modified RS made from high amylose maize starch, is currently used as food additive (E1413) in the EU.

Different methods are used to analyze RS in vivo. One of the ways to assay RS physiologically is to determine starch in the undigested ileal content. Terminal ileal samples can be recovered by intubation or from ileostomy bags. The classic way to substantiate starch digestion is by measuring the glycemic index as described by Jenkins and others (1981). This implies measuring the area under the curve (AUC) of the serum glucose concentration over the first 2 h after administering a starch and dividing this by the serum glucose response after consumption of an equal amount of glucose. Determination of breath hydrogen (breath tests) can also be used as a semiquantitative measurement for RS. In a study on effect of RS on human colon, increased fermentation was verified by elevated breath hydrogen excretion (Hylla and others 1998). From the different animal models, the antibiotic-treated rat model is the one commonly used (Gudmand-Hoyer 1991).

Digestibility, energy value, and RDA of RS


RS is highly resistant to mammalian enzymes. In cereal products, the RS fraction is not digestible both in vitro and in vivo (Bornet 1993). Four different RS fractions have been identified in cereal products: native starch, retrograded amylose, the amylo-lipid complex, and encapsulated gelatinized starch. After reaching the large intestine, the RS fractions are fermented by the colonic flora, resulting in short-chain fatty acids (SCFA). SCFA profiles derived from RS are lower in acetate and higher in butyrate than those of conventional fibers. The SCFA are an energy source for colonic cells (butyrate) and to the body as a whole (acetate and propionate). Maize starch acylated with acetic, propionic, or butyric anhydride are also RS and raise large bowel SCFA, apparently through bacterial release of the esterified fatty acid and fermentation of the residual starch (Annison and others 2003). Some sources of RS seem to be less available for fermentation, as has been observed in some chemically modified starches as well as RS in arepa (high corn meal bread). Feeding trials on Himalaya 292, a hulless barley cultivar with a higher RS content, also have resulted in high levels of SCFA in feces (Bird and others 2004). Most studies indicate that 30% to 70% of RS is metabolized (Ranhotra and others 1991a; Behall and Howe 1995; Behall and Howe 1996; Cummings and others 1996; Ranhotra and others 1996), while the balance is excreted in the feces. The variability is
11

Determination of RS
In vitro methods

The main step of any method to measure the content of RS in foods must first remove all of the digestible starch from the product using thermostable -amylases (McCleary and Rossiter 2004). At present, the method of McCleary and Monaghan (2002) is consid-

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


largely due to effects caused by the malabsorption of the ingested starch. In human subjects, replacement of 27 g of digestible starch by RS (raw potato starch) in a single meal lowered diet-induced thermogenesis by an average of 90 KJ/5 h (Heijnen and others 1995). A study was designed to compare the metabolizable energy of 2 starch sources, standard cornstarch and high amylose cornstarch (Behall and Howe 1996). Based on energy intake and fecal excretion from all subjects, the partial digestible energy value for the RS averaged 11.7 KJ/g RS, which was 67.3% of the energy of standard cornstarch. Control and hyperinsulinemic subjects differed in their ability to digest RS, averaging 81.8% and 53.2%, respectively. RS averaged 2.8 kcal/g for all subjects but only 2.2 kcal/g in the hyperinsulinemic subjects. This enables the use of RS in reducing the energy value of foods. Approximately 20 g/d is recommended to obtain the beneficial health benefits of RS. However, worldwide, dietary intakes of RS are believed to vary considerably. It is estimated that intakes of RS in developing countries with high starch consumption rates range from approximately 30 to 40 g/d (Baghurst and others 2001). Dietary intakes in India and China were recently estimated at 10 and 18 g/d (Platel and Shurpalekar 1994; Muir and others 1998). Intakes in the EU are believed to lie between 3 and 6 g/d (Dyssler and Hoffmann 1994). Dietary intakes of RS in the U.K. are estimated at 2.76 g/d (Tomlin and Read 1990) and are believed to range from 5 to 7 g/d in Australia (Baghurst and others 2001). In Sweden, the daily RS intake is estimated to be 3.2 g (Liljeberg 2002). In New Zealand, RS intakes have been approximated to be 8.5 g/d and 5.2 g/d in 15- to 18-year-old males and females, respectively (Baghurst and others 1996). ing in preventing colonic cancer (Asp and Bjorck 1992). Significant changes in fecal pH and bulking as well as greater production of SCFA in the cecum of rats fed RS preparations have been reported (Ferguson and others, 2000), which have been suggested to resemble the effects of soluble dietary fiber. However, when RS was combined with an insoluble dietary fiber like wheat bran, much higher SCFA levels, in particular butyrate was observed in the feces (Leu and others 2002). In rats, when RS was combined with psyllium, the site of RS fermentation was pushed more distally. As the distal colon is the site where most tumors arise it may be of additional benefit for cancer protection if fermentation is further enhanced within the distal colon (Morita and others 1999).
Hypoglycaemic effects

Beneficial physiological effects of RS


A number of physiological effects have been ascribed to RS (Nugent 2005), which have been proved to be beneficial for health.
RS as a component of DF

Foods containing RS moderate the rate of digestion. The slow digestion of RS has implications for its use in controlled glucose release applications. The metabolism of RS occurs 5 to 7 h after consumption, in contrast to normally cooked starch, which is digested almost immediately. Digestion over a 5- to 7-h period reduces postprandial glycemia and insulinemia and has the potential for increasing the period of satiety (Raben and others 1994; Reader and others 1997). A study using 10 healthy normal-weight males fed test meals containing either 50 g starch free of RS (0% RS), or 50 g starch containing a high level of RS (54% RS) proved the ability of high RS meals to significantly lower the postprandial concentration of blood glucose, insulin, and epinephrine (Raben and others 1994). Similarly, from a human study (Reader and others 1997), using a commercial RS3 ingredient (CrystaLean ), the maximum blood glucose level was found to be significantly lower than that of other carbohydrates (simple sugars, oligosaccharides, and common starch). The RS3-containing bar decreased postprandial blood glucose and may play a role in providing improved metabolic control in type II diabetes (non-insulin dependent).
As a prebiotic

RS appears to be highly resistant to mammalian enzyme and may be classified as a component of fiber on the basis of the recent definitions of dietary fiber given by AACC (2000) and NAS (2002). While part of the RS may consist of low-molecular-weight dextrins, the bulk consists of polymers, of which retrograded amylose often forms the major fraction (Ranhotra and others 1991a). Although not a cell wall component, RS is obviously nutritionally more similar to NSP than to digestible starch. There is ample justification that RS behaves physiologically like fiber. RS assays as insoluble fiber, but has the physiological benefits of soluble fiber. Additionally, RS exhibits a level of slow digestibility and can be used as a vehicle for the slow release of glucose. Also, like soluble fiber, it has a positive impact on colonic health by increasing the crypt cell production rate, or decreasing the colonic epithelial atrophy in comparison with no-fiber diets. There is indication that RS like guar, a soluble fiber, influences tumorigenesis, and reduces serum cholesterol and triglycerides. Overall, since RS behaves physiologically as a fiber, it should be retained in the TDF assay (Haralampu 2000).
Prevention of colonic cancer

RS has been suggested for use in probiotic compositions to promote the growth of such beneficial microorganisms as Bifidobacterium (Brown and others 1996). Since RS almost entirely passes the small intestine, it can behave as a substrate for growth of the probiotic microorganisms.
Reduction of gall stone formation

The digestibility of starch in rice and wheat is increased by milling to flour (Heaton 1988). Digestible starch contributes gall stone formation via a greater secretion of insulin and insulin in turn leads to the stimulation of cholesterol synthesis. RS is found to reduce the incidence of gallstones (Malhotra 1968). Gallstones are less in South India where, whole grains are consumed rather than as flour in Northern India. Notably, dietary intake of RS is 2- to 4fold lower in the United States, Europe, and Australia, compared with populations consuming high-starch diets, such as India and China, which may reflect in the difference in the number of gallstone cases in these countries (Birkett and others 2000).
Hypocholesterolaemic effects

Starch unabsorbed in the small intestine is fermented by the microflora of the large intestine. Generally, starch is not present in the feces of humans or experimental animals, indicating more or less complete fermentation. In vitro experiments with human fecal inocula have shown that the butyrate yield from starch is high. As butyrate is a main energy substrate for large intestinal epithelial cells and inhibits the malignant transformation of such cells in vitro; this makes easily fermentable RS fractions especially interest12

Hypocholesterolemic effects of RS have been amply proved. In rats, RS diets (25% raw potato) markedly raised the cecal size and the cecal pool of short-chain fatty acids (SCFA), as well as SCFA absorption and lowered plasma cholesterol and triglycerides. Also, there was a lower concentration of cholesterol in all lipoprotein fractions, especially the HDL1 and a decreased concentration of triglycerides in the triglyceride-rich lipoprotein fraction. Results of feeding trials on rats using RS from Adzuki starch (AS) and tebou starch (TS) suggested that AS and TS had a serum cho-

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


lesterol-lowering function due to the enhanced levels of hepatic SR-B1 (scavenger receptor class B 1) and cholesterol 7 -hydroxylase m RNA (Han and others 2003). The bean starches lowered the levels of serum total cholesterol and VLDL + IDL + LDL cholesterol, increased caecal concentration of short-chain fatty acids (in particular butyric acid concentration), and increased faecal neutral sterol excretion. From studies on hamsters fed diets containing cassava starch extrudated with 9.9% oat fiber or cassava starch extruded with 9.7% RS, hypocholesterolemic properties of both were demonstrated suggesting their use in foods to improve cardiovascular health (Martinez and others 2004).
Inhibition of fat accumulation

High amylose corn starch (HACS) coatings with RS at 20% to 24% showed excellent properties useful for tablet coatings (Dimantov and others 2004). Hydrolyzed starches (those which retain their granular structure and essentially behave like unmodified starches in undergoing gelatinization on heating), which are also referred to as thin boiling starches, are also a form of RS. The advantage of this starch is the high concentration, which can be used as a paste of low viscosity and its ability to set as a firm gel (Seib and Kyungsoo 1999). Cross-linked starches based on maize, tapioca, potato of RS4 type have been useful in formulations needing pulpy texture, smoothness, flowability, low pH storage, and high temperature storage (Sajilata and Singhal 2005).

Replacement of 5.4% of total dietary carbohydrates with RS in a meal could significantly increase postprandial lipid oxidation suggesting reduction in fat accumulation in the long term (Higgins and others 2004).
Absorption of minerals

Commercial sources of RS
In the commercial development of RS, it is advantageous to start with a native starch high in amylose. Nearly all patented processes are based on the propensity of high-amylose starch to retrograde, or form highly crystalline regions, which are resistant to enzymatic hydrolysis (Crosby 2001). High-amylose maize starches have high gelatinization temperatures, requiring temperatures that are often not reached in conventional cooking practices (154 C to 171 C) before the granules are completely disrupted. These starches offer an opportunity to manipulate the amount of RS present in food products. The first commercial RS was introduced as Hi-maize in Australia in 1993 by Starch Australasia, now part of Natl. Starch and Chemical Co. This product is a natural granular form of starch produced from a corn hybrid containing more than 80% amylose. Hi-maize analyses as 42% RS and has gained widespread use in Australia in breads and other baked goods. Commercial sources of RS (Ranhotra and others 1996), such as CrystaLean (Opta Food Ingredients, Inc., Bedford, Mass., U.S.A.), Novelose (Natl. Starch and Chemical Co., Bridgewater, N.J., U.S.A.) and Amylomaize VII (Cerestar Inc., Hammond, Ind., U.S.A.) are now available to increase the DF content in foods and provide other functional properties. CrystaLean is a commercial, highly retrograded RS3 based on the aeVII hybrid. It is produced by first fully hydrating and disrupting the starch granules, followed by an enzymatic debranching of the amylopectin to yield a low DE maltodextrin mixture, which is almost entirely a straight chain. Then, the mixture is treated through thermal cycles to achieve a high level of retrogradation before drying. CrystaLean containing 41% RS is digested slowly, at approximately half the rate of maltodextrin. The ingredient was introduced in the early 1990s and is now used in products for diabetics. Shortly after the launch of CrystaLean, Natl. Starch introduced a very similar product named Novelose 330. More recently, Natl. Starch Chemical Co. has developed processes for manufacturing granular forms of concentrated RS containing 47% to 60% RS by heating and cooling high-amylose corn starch under conditions of carefully controlled moisture and temperature. These products are marketed as Novelose 240 and 260. Novelose 240 is a thermally modified RS2 based on ae-VII hybrid of corn (Shi and Trzasko 1997). The modification renders the native granule more stable by holding the starch at elevated temperature (60 C to 160 C) in the presence of limited water (10% to 80%). Novelose 260 contains 60% TDF, the highest level available in a RS. It can be formulated into a broad range of foods such as pasta, cereals, and snack foods that can carry a rich-in-fiber labeling. Novelose 240 (RS2), 260 (RS2), and 330 (RS3) have melting temperatures of 99.7 C, 114.4 C, and 121.5 C, respectively. These products offer medium/high, high, and very high process tolerance and are therefore suitable for use in a variety of processed foods. Another player is Cerestar (a Cargill company), which has launched Ac13

A study to compare the intestinal apparent absorption of calcium, phosphorus, iron, and zinc in the presence of either resistant or digestible starch brought out that a meal containing 16.4% RS resulted in a greater apparent absorption of calcium and iron compared with a completely digestible starch (Morais and others 1996). Thus RS could have a positive effect on intestinal calcium and iron absorption.

Undesirable manifestations of RS
Some undesirable manifestations of RS are also seen in clinical studies (Reussner and others 1963). One of the most alarming influences is the cecal enlargement, especially in rats. But, in humans, it is said to be of little relevance because of the considerably smaller size and weight of cecum as well as its smaller role in the human physiological function. Pelvic nephrocalcinosis is another phenomenon observed in rats especially those fed with mono substituted cross-linked starches and mono substituted starches (Baley and others 1973).

Applications of RS
The industrial applications of RS are mainly in the preparation of moisture-free food products (Yue and Waring 1998). Bakery products such as bread, muffins, and breakfast cereals can be prepared by using RS as a source of fiber. The amount of RS used to replace flour depends on the particular starch being used, the application, the desired fiber level, and, in some cases, the desired structure-function claims. From a quality standpoint, some applications are more sensitive to flour replacement than others. For example, bread and rolls, which generally have a bland flavor, are low-fat and require a minimum amount of gluten for structure; the maximum flour replacement is typically 10% to 20% without noticeably changing the texture. A type III enzyme-RS with a melting point of at least 140 C, which could be used as a low-calorie flour replacer in bakery products exhibits baking characteristics (cookie spread, golden brown color, pleasant aroma, surface cracking) comparable to those achieved using conventional wheat flour (Haynes and others 2000). Excellent quality sponge cakes have been prepared by replacing 30% of flour with 4 cycled autoclaved-cooled RS3 corn starch (RS3), cross-linked maize starch (RS4) and annealed and crosslinked RS4 maize starch (ARS4) while for yellow layer cake, the replacement level was found to be at 12.5% (Po and others 1994; Myung-Hee-Kim and others 2001).

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


tistar containing 58% RS, made by crystallizing hydrolyzed tapioca starch (maltodextrins). High-amylose corn (amylomaize) is the generic name for corn that has an amylose content higher than 50%. The endosperm mutant amylose-extender (ae) increases the amylose content of the endosperm to about 60% in many dent backgrounds. Modifying factors alter the amylose contents as well as desirable agronomic characteristics of the grain. The amylose-extender gone expression is characterized by a tarnished, translucent, sometimes semi-full kernel appearance. U.S. production of high-amylose corn is forecasted at 50000 acres for 2001. 100% of this product is grown under contract in Indiana for Cerestar and Pioneer. Highamylose corn yields vary, depending upon location, but average only 65% to 75% of that of ordinary dents. Three types are produced commercially: Class V (50% to 60% amylose) and Class VII (70% to 80% amylose), and Class IX (90% amylose). C*ActiStar (Cerestar Food and Pharma Specialties) is made from native, partially hydrolyzed starch by generating microscopic structures that are not hydrolyzed by digestive enzymes in the small intestine (58% RS). The specific structure of C*ActiStar favors its fermentation to butyrate and helps to reach the lower parts of the large bowel, which is the most relevant segment of the colon for the gut health maintenance. It can be used in a variety of foodstuffs such as bread, cereal bars, biscuits/cookies, muesli, low-fat fermented milk, UHT flavored milk drinks, and ready-to-use powdered mixes such as instant soups and instant chocolate. With C*ActiStar, each serving can provide a significant amount of RS, which will contribute to an increase in the pre-existing daily intake of RS. MGP Ingredients, Inc. (Atchison, Kans.) and Cargill have announced a business alliance for the production and marketing of a new RS product called FibersymTM HA that is derived from high-amylose corn and ideal for use in a wide array of lower net carbohydrate food products. Delivering more than 70% dietary fiber, it greatly reduces net carbohydrate levels in foods. Applications cover a wide variety of products, including breads, tortillas, pizza crust, cookies, muffins, waffles, breakfast cereals, snack products, and nutritional bars. FibersymTM HA joins MGPIs other resistant starches, which include a wheat-based RS, FibersymTM 70 and a potato-based variety, FibersymTM 80 ST. Roquette (Roquette, Freres, France) recently developed Nutriose FB, a new dextrin that offers all the benefits of RS and, in addition, is a soluble fiber. Because it is only slightly digested in the small intestine and then slowly fermented in the colon, it fits the functional pattern of RS with a low glycemic response. Fibersol-2, offered through a joint venture of ADM and Matsutani, is a digestion-resistant maltodextrin. Resistant starches that test high in TDF such as Roquettes Nutriose FB 06 at 85% fiber content or ADM/Matsutanis Fibersol-2 at 90% make it possible to enrich products with fiber to optimum recommended levels and to support label claims. Research is intensifying regarding RS4 starches created using difunctional phosphate reagents, available for inclusion in foods. However, as yet there is a lack of information regarding their potential clinical and physiological effects (Brown 2004). As RS is included within the definitions of dietary fiber by the AACC (Anon 2000; Jones 2000), the Inst. of Medicine of the Natl. Academies (Inst. of Medicine 2002), and is measured within the remit of the AOAC method (Prosky and others 1985), which is used in the United States, U.K., Australia, and Japan, commercially manufactured sources of RS can be used as vehicles to increase the total dietary fiber content of foods and food products without affecting taste and texture (Liversey 1994). They may also be used to provide fiber in some commercially available low-carbohydrate foods marketed for those following low-carbohydrate dieting regimens. There are a number of advantages to using commercially manufactured RS in food products. Unlike natural sources of RS (that is, legumes, potatoes, bananas), commercially manufactured RS are
14

not affected by processing and storage conditions. For example, the amount of RS2 in green bananas decreases with increasing ripeness; however, a commercial form of RS2, Hi-maize, does not experience these difficulties. Among the newest developments in resistant starches is an RS2 that remains resistant after mild food processing (Novelose 240). Compared with conventional fibers, it has many advantageous features. It is white and has a bland flavor and a fine particle size between 10 and 15 m. It also has a reduced caloric content and may be used as a bulking agent to complement reduced sugar or reduced-fat formulations. With a TDF content of approximately 40%, this RS2 can be used alone or as a functional complement to other fiber sources and can be labeled simply as cornstarch. Most importantly, the commercial RS has a much lower water-holding capacity than do various traditional fibers. Because it absorbs less water, adjustments in product formulations and processing are substantially minimized. In general, RS has many properties, which outweigh the disadvantages posed by it. Therefore, for deriving maximum benefits from RS and minimizing undesirable effects of RS, it becomes essential to regulate the quantity of RS to be used. Process tolerance needs to be considered when selecting and using RS for a particular application (www.nstarch.com). The majority of commercial RS on the market will retain their dietary fiber through typical baking processes and even mild extrusion. However, extreme processing conditions may damage the RS, resulting in a loss of dietary fiber. Formulators should therefore analyze finished foods made with processes involving extremely high temperatures, pressure, and/or shear to verify the level of dietary fiber. RS is well-suited for low-moisture food systems where they can be used at very high levels compared to traditional starches. The majority of the commercially available RS rely on intact granules or compact crystalline regions with high melting temperatures to resist digestion. Therefore, these products generally will not swell or contribute viscosity during most processing. Essentially insoluble, RS would not replace viscosifying starch in liquid applications.

Conclusions
RS has received much attention for both its potential health benefits and functional properties. As a functional fiber, its fine particles and bland taste make possible the formulation of a number of food products with better consumer acceptability and greater palatability than those made with traditional fibers. RS shows improved crispness and expansion in certain products and better mouthfeel, color, and flavor over products produced with some traditional, insoluble fibers. It is ideal for use in RTE cereals, snacks, pasta/ noodles, baked goods, and fried foods and permits for easy labeling as simply starch, conferring additional nutraceutical benefits. Being nondigestible, RS can be used in reduced-fat and sugar formulations. RS has properties similar to fiber and shows promising physiological benefits in humans, which may result in disease prevention. Foods containing high levels of RS yield fewer calories and lower glycaemic loadsimportant formulation considerations for diabetics as well as the weight-conscious. It is classified as a fiber component with partial or low fermentation. Technically, it is possible to increase the RS content in foods by modifying the processing conditions such as pH, heating temperature and time, number of heating and cooling cycles, freezing, and drying. A number of commercially available RS preparations would make it possible for a wide range of applications with nutraceutical implications.

References
[AACC] American Assn. of Cereal Chemists. 2000. Approved methods of the AACC. 10th ed. Method 44-15A. St. Paul, Minn.: AACC.

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


Abbas IR, Scheerens JC, Berry JW. 1987. Tepary bean starch. Part III. In vitro digestibility. Starch/Starke 39(8):2804. Abd-Elmoneim-O-Elkhalifa, Schiffler B, Bernhard R. 2004. Effect of fermentation on the starch digestibility, resistant starch and some physicochemical properties of sorghum flour. Nahrung/Food 48(2):914. Adamu BOA. 2001. Resistant starch derived from extruded corn starch and guar gum as affected by acid and surfactants: structural characterization. Starch/ Starke 53(11):58291. Agama-Acevedo E, Rendon-Villalobos R, Tovar J, Parades-Lopez O, Islas-Hernandez JJ, Bello-Perez L. 2004. In vitro starch digestibility changes during storage of maize flour tortillas. Nahrung/Food 48(1):3842. Annison G, Illman RJ, Topping DL. 2003. Acetylated, propionylated or butyrylated starches raise large bowel short-chain fatty acids preferentially when fed to rats. J Nutr 133(11):35238. Anon. 2000. AACC Board holds midyear meeting. Cereal FoodsWorld 45:325. Asp NG, Bjorck I. 1992. Resistant STarch. Trends Food Sci Technol 3(5):1114. Asp NG. 1994. Nutritional classification of food carbohydrates. Am J Clin Nutr 59:S67981. Baghurst K, Baghurst PA, Record SJ. 2001. Dietary fiber, nonstarch polysaccharide and resistant starch intakes in Australia. In: GA Spiller, editor. CRC handbook of dietary fiber in human health. Boca Raton, Fla.: CRC Press. p 58391. Baghurst PA, Baghurst KI, Record, SJ. 1996. Dietary fiber, non-starch polysaccharides and resistant starcha review. Food Aust 48(3):S3S35. Baley DE, Cox GE, Morgarareidge K. 1973. Report 140. Maspeth, N.Y.: Food and Drug Research Laboratories. Behall KM, Howe JC. 1995. Contribution of fiber and resistant starch to metabolizable energy. Am J Clin Nutr 62 (Suppl): 1158S60S. Behall KM, Howe JC. 1996. Resistant starch as energy. J Am Coll Nutr 15(3):24853. Belitz HD and Grosch W. 1999. Polysaccharides. In: Food chemistry. 2nd ed. Berlin, Germany: Springer-Verlag. p 301. Berry CS. 1986. Resistant starch. formation and measurement of starch that survives exhaustive digestion with amylolytic enzymes during the determination of dietary fiber. J Cereal Sci 4: 30114. Betancur AD, Chel GL. 1997. Acid hydrolysis and characterization of Canavalia ensiformis starch. J Agric Food Chem 45:423741. Bird AR, Flory C, Davies DA, Usher S, Topping DL. 2004. A novel barley cultivar (Himalaya 292) with a specific gene mutation in starch synthase IIa raises large bowel starch and short-chain fatty acids in rats. J Nutr 134(4):8315. Birkett AM, Mathers JC, Jones GP, Walker KZ, Roth MJ, Muir JG. 2000. Changes to the quality and processing of starchy foods in a Western diet can increase polysaccharides escaping digestion and improve in vitro fermentation variables. Br J Nutr 84:6372. Bjorck I, Eliasson AC, Drews A, Gudmundsson M, Karlson R. 1990. Some nutritional properties of starch and dietary fiber in barley genotypes containing different levels of amylose. Cereal Chem 67:32733. Bjorck I, Gunnarsson A, Ostergard K. 1989. A study of native and chemically modified potato starch. Part II. Digestibility in the rat intestinal tract. Starke 41:12834. Bjorck I, Nyman M, Pedersen P, Siljestrom M, Asp NG, Eggum BO. 1987. Formation of enzyme resistant starch during autoclaving of wheat starch: Studies in vitro and in vivo. J Cereal Sci 6:15972. Bjorck IM, Nyman ME. 1987. In vitro effects of phytic acid and polyphenols on starch digestion and fiber degradation. J Food Sci 52:158894. [BNF] British Nutrition Foundation. 1990. Complex carbohydrates in foods: the report of The British Nutrition Foundations Task Force. London: Chapman & Hall. Bornet F. 1993. Technological treatments of cereals. Repercussions on the physiological properties of starch. Carbohydr Polym 21(2/3):195203. Brown IL, McNaught KJ, Ganly RN, Conway PL, Evans AJ, Topping DL, Wang X. 1996. Probiotic compositions. Intl. Patent WO 96/ 08261/ A1. Issued Mar 21, 1996; Univ New South Wales; Burns Philip and Co. Limited; Burns Philip Res and Dev Pty; Mauri Lab Pty Ltd; Commw Sci Ind Res Org; Arnotts Biscuits Ltd; Good man Fielder Ingredients Li; Goodman Fielder Ltd; Brown IL, McNaught KJ, Ganly RN, Conway PL, Evans AJ, Topping DL, Wang X. Brown IL, McNaught KJ, Moloney E. 1995. Hi-maize TM: new directions in starch technology and nutrition. Food Aust 47:2725. Brown IL. 2004. Applications and uses of resistant starch. J Assoc Off Anal Chem Int 87(3):72732. Brumovsky JO, Thompson DB. 2001. Production of boiling-stable granular resistant starch by partial acid hydrolysis and hydrothermal treatments of high-amylose maize starch. Cereal Chem 78(6):6809. Buch JS, Walker CE. 1988. Sugar and sucrose ester effects on maize and wheat starch gelatinization patterns by differential scanning calorimeter. Starke 40:3536. Buttcher V, Welsh T, Mitzer LS, Kossmann J. 1997. Cloning and characterization of the gene for amylosucrase from Neisseria polysaccharea: production of linear 1, 4-glucan. J Bacteriol 179:332430. Champ M, Langkilde AM, Brouns F. 2003. Advances in dietary fiber characterization. 2. Consumption, chemistry, physiology and measurement of resistant starch; implications for health and food labeling. Nutr Res Rev 16:14361. Chiu CW, Henley M, Altieri P. 1994. Process for making amylase resistant starch from high amylose starch. US 5 281 276; Jan 25, 1994. Wilmington, Del.: National Starch and Chem Investment Holding Corp. Crosby GA. 2001. Commercial processes for the manufacture of resistant starch. Opta Food Ingredients, Inc., Bedford, MA 01730. AACC Annual Meeting; Charlotte, N.C.; 2001 Oct. 148. Cummings JH, Englyst HN. 1991. Measurement of starch fermentation in the human large intestine. Can J Physiol Pharmacol 69:1219. Cummings, JH, Beatty, ER, Kingman S M, Bingham SA, Englyst HN. 1996. Digestion and physiological properties of resistant starch in human large bowel. Br J Nutr 75:73347. Czuchajowska Z, Sievert D, Pomeranz Y. 1991. Enzyme-resistant starch. IV. Effects of complexing lipids. Cereal Chem 68(5):53742. Dimantov A, Greenberg M, Kesselman E, Shimoni E. 2004. Study of high amylose corn starch as food grade enteric coating in a microcapsule model system. Innov Food Sci Emerg Technol 5(1):93100. Eerlingen RC, Crombez M and Delcour JA 1993a. Enzyme-resistant starch. I. Quantitative and qualitative influence of incubation time and temperature of autoclaved starch on resistant starch formation. Cereal Chem 70(3):33944. Eerlingen RC, Deceuninck M and Delcour JA 1993b. Enzyme-resistant starch. II. Influence of amylose chain length on resistant starch formation. Cereal Chem 70(3):34550. Eerlingen RC, Delcour JA. 1995. Formation, analysis, structure and properties of type III enzyme resistant starch. J Cereal Sci 22:12938. Eerlingen RC, Delcour JA. 1995. Formation, analysis, structure and properties of type III enzyme resistant starch. J Cereal Sci 21:18. Eerlingen RC, Van den Broeck I, Delcour JA, Levine H. 1994. Enzyme resistant starch. VI. Influence of sugars on resistant starch formation. Cereal Chem 70:345. Eggum BO, Juliano BO, Perez CM and Acedo EF. 1993. The resistant starch, undigestible energy and undigestible protein contents of raw and cooked milled rice. J Cereal Sci 18(2):15970. Eliasson AC, Finstad H, Ljunger G. 1988. A study of starch-lipid interactions for some native and modified maize starches. Starke 40:95100. Ellis RP, Cochrane MP, Dale MFB, Duffus CM, Lynn A, Morrison IM, Prentice RDM, Swanston JS, Tiller SA. 1998. Starch production and industrial use. J Sci Food Agric 77:289311. Englyst HN, Cummings JH. 1985. Digestion of the polysaccharides of some cereal foods in the human small intestine. Am J Clin Nutr 42:77887. Englyst HN, Cummings JH. 1986. Digestion of the carbohydrates of banana (Musa paradisiaca sapientum ) in the human small intestine. Am J Clin Nutr 44:4250. Englyst HN, Kingman SM, Cummings JH. 1992. Classification and measurement of nutritionally important starch fractions. Eur J Clin Nutr 46:S33S50. Englyst HN, Wiggins HS, Cummings JH. 1982. Determination of the non-starch polysaccharides in plant foods by gas-liquid chromatography of constituent sugars as alditol acetates. Analyst 107:30718. Erbersdobler FH. 1989. Factors affecting uptake and utilization of macro nutrients. In: Southgate DAT, Johnson IT, Fenwick GR, editors. Nutrient availability: chemical and biological aspects. Cambridge, U.K.: Royal Society of Chemistry. p 3309. Escarpa A, Gonzalez MC, Morales MD, Saura-Calixto F. 1997. An approach to the influence of nutrients and other food constituents on resistant starch formation. Food Chem 60(4):52732. Faisant N, Champ M, Colonna P, Buleon A, Molis C, Langkilde AM, Schweizer T, Flourie B, Galmiche JP. 1993. Structural features of resistant starch at the end of the human small intestine. Eur J Clin Nutr 47:28596. Faraj A, Vasanthan T, Hoover R. 2004. The effect of extrusion cooking on resistant starch formation in waxy and regular barley flours. Food Res Int 37(5):51725. Fausto FD, Kacchi AI, Mehta D. 1997. Starch products in confectionery. Bev Food World 24(4):416, 24. Ferguson LR, Tasman-Jones C, Englyst H, Harris PJ. 2000. Comparative effects of three resistant starch preparations on transit time and short-chain fatty acid production in rats. Nutr Cancer 36:230-7. Franco CML, Ciacco CF, Tavares DQ. 1995. Effect of heat-moist treatment on enzymatic susceptibility. Starke/Starch 47(6):2238. Galliard T. 1987. In starch: properties and potential. Chichester, U.K.: John Wiley and Sons. Garcia-Alonso A, Jimenez-Escrig A, Martin-Carron N, Bravo L, Saura-Calixto F. 1999. Assessment of some parameters involved in the gelatinization and retrogradation of starch. Food Chem 66:1817. Gebhardt E, Dongowski G, Huth M, Rabe E. 2001. Formation of resistant starch by extrusion and the completion of the dietary fiber analysis. Getreide-Mehl-undBrot. 55(6):36371. Geng Z, Zongdao C, Toledo R. 2003. Effects of different processing methods on the glycemic index of kudzu starch. J Chinese Cereals Oils Assoc 18(5):5. Gidley MJ. 1989. Molecular mechanisms underlying amylose aggregation and gelation. Macromolecules 22:3518. Gidley MJ, Bulpin PV. 1989. Agregation of amylose in aqueous systems: The effect of chain length on phase behavior and aggregation kinetics. Macromolecules 22:3416. Gidley MJ, Cooke D, Drake AH, Hoffman RA, Russell AL, Greenwell P. 1995. Molecular order and structure in enzyme-resistant retrograded starch. Carbohydr Polym 28:2331. GranfeldtY, Drews A, Bjoerck I. 1995. Arepas made from high amylose corn flour produce favorably low glucose and insulin responses in healthy humans. J Nutr 125(3):45965 Gudmand-Hoyer E. 1991. Methodological aspects of in vivo methods for measurement of starch digestibility. Report of a European Flair Concerted Action Workshop; Vedback, Copenhagen; 1991 Nov. 102. Denmark: Euresta. Hamilton RM, Paschall EF. 1967. Production and uses of starch phosphates. In: Whistler RL, Paschall EF, editors. Starch: chemistry & technology. Vol. II. New York and London: Academic Press. p 35165. Han KH, Fukushima M, Kato T, Kojima M, Ohba K, Shimada K, Sekikawa M, Nakano M. 2003. Enzyme-resistant fractions of beans lowered serum cholesterol and increased sterol excretions and hepatic mRNA levels in rats. Lipids 38(9):91924. Haralampu SG. 2000. Resistant starcha review of the physical properties and biological impact of RS3. Carbohyder Polym 41:28592. Haralampu SG, Gross A. 1998. Granular RS and method of making. U.S. Patent 58, 49, 090. Dec 15, 1998. Bedford, Ma.: Opta Food Ingredients, Inc. Harris DW, Little JA. 1994. Method of preparing reduced fat foods. U.S. patent 5,374,442. Dec 20, 1994. Decatur, Ill.: Stanley Manufacturing Co. Harris DW, Little JA. 1995. Method of preparing reduced fat foods. U.S. patent 5,395,640. March 7, 1995. Decatur, Ill.: Stanley Manufacturing Co. Haynes L, Gimmler N, Locke JP, Mee-Ra-Kweon, Slade L, Levine H. 2000. Process for making enzyme-resistant starch for reduced-calorie flour replacer. U.S. patent 6,013,299. Jan 11, 2000. Wilmington, Del.: Nabisco Technology Co. Heaton KW. 1988. Gall stone prevention. In Bile acids and diseases. Lancester: MTP Press. p 57169. Heijnen MA, Deurenberg P, Van A, Johan MM, Beynen AC. 1995. Replacement of

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

15

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


digestible by resistant starch lowers diet-induced thermogenesis in healthy men. Br J Nutr 73(3):42332. Higgins JA, HR Dana, Donahoo WT, Brown IL, Bell ML, Bessesen DH. 2004. Resistant starch consumption promotes lipid oxidation. Nutr Met 1:18. Hizukuri S. 1986. Polymodal distribution of chain lengths of amylopectins and its significance. Carbohydr Res 147:3427. Holm J, Asp NG, Bjorck I. 1987. Factors affecting enzymatic degradation of cereal starches in vitro and in vivo. In: Morton Ellis D, editor. Cereal in a European context. Bournemouth, Dorset: First European Conference on Food Science and Technology. New York, NY: VCH Publisher. Harwood, Chichester, U.K.: p 16987. Holm J, Bjorck I. 1992. Bioavailability of starch in various wheat bread products evaluation of metabolic responses in healthy subjects and rate and extent of in vitro starch digestion. Am J Clin Nutr 55:4209. Holm J, Lundquist I, Bjorck I, Eliasson AC, Asp NG. 1988. Relationship between degree of gelatinization, digestion rate in vitro and metabolic response in rats. Am J Clin Nutr 47:10106. IAnson KJ, Miles MJ, Morris VJ, Besford LS, Jarvis DA, Marsh RA. 1990. The effects of added sugars on the retrogradation of wheat starch gels. J Cereal Sci 11:2438. Iyengar R. 1991. U.S. Patent on starch-derived, food-grade, insoluble bulking agent. U.S. Patent 5, 051, 271. Sept 24, 1991. Cambridge, Ma.: Opta Food Ingredients Inc. Jaylin J, Robyt JF. 1984. Structure studies of amylose-V complexes and retrograded amylose by action of alpha amylases and a new method for preparing amylodextrins. Carbohydr Res 132:10518. Jenkins DJ, Wolever TM, Taylor RH. 1981. Glycemic index of foods: a physiological basis for carbohydrate exchange. Am J Clin Nutr 34:3626. Jenkins DJA, Jenkins AL, Wolever TMS, Rao AV, Thompson LU. 1986. Fiber and starchy foods: gut function and implication in disease. Am J Gastroenterol 81:92030. Jian-Huali, Gao-Qunyu, Liang-Shizhong. 2003. Study on crystallization property of resistant starch. Food Sci China 24(7):447. Johansson CG, Siljestrom M. 1984. Dietary fiber of bread and formation of RS on baking. Z Lebesnm Forsch 179:248. Jones J. 2000. Update on defining dietary fiber. Cereal Foods World 45:21920. Kavita V, Varghese S, Chitra GR, Jamuna P. 1998. Effects of processing, storage time and temperature on the resistant starch of foods. J Food Sci Technol 35(4):299304. King JM, Tan SY. 2005. Resistant starch with cooking properties similar to untreated starch. US Patent Application 20050089624. April 28, 2005. Kohyama K, Nishinari K. 1991. Effect of soluble sugars on gelatinization and retrogradation of sweet potato starch. J Agric Food Chem 39:140610. Laurentin A, Edwards CA. 2004. Differential fermentation of glucose-based carbohydrates in vitro by human faecal bacteria. A study of pyrodextrinised starches from different sources. Eur J Nutr 43(3):1839. Leu RK, Hu Y, Young GP. 2002. Effects of resistant starch and nonstarch polysaccharides on colonic luminal environment and genotoxin-induced apoptosis in the rat. Carcinogenesis 23(5):7139. Liljeberg EH. 2002. Resistant starch content in a selection of starchy foods on the Swedish market. Eur J Clin Nutr 56(6):5005. Liljeberg H, Akerberg A, Bjorck I. 1996. Resistant starch formation in bread as influenced by choice of ingredients or baking conditions. Food Chem 56(4):38994. Liversey G (1994) Energy value of resistant starch. In: Asp G, Van Amelsvoort JMM, Hautvast JGAJ, editors. Proceedings of the Concluding Plenary Meeting of EURESTA EURESTA. The Netherlands: Wageningen, p 5662. Malhotra SL. 1968. Epidemiological study of cholelithiasis among rail road workers in India. Gut 9:2905. Malshick S, Kyungsoo W, Seib PA. 2003. Hot-water solubilities and water sorptions of resistant starches at 25C. Cereal Chem 80(5):5646. Manners DJ. 1989. Recent developments in our understanding of amylopectin structure. Carbohydr Polym 11:87112. Marconi E, Ruggeri S, Cappelloni M, Leonardi D, Carnovale E. 2000. Related Physicochemical, nutritional, and microstructural characteristics of chickpeas (Cicer arietinum L.) and common beans (Phaseolus vulgaris L.) following microwave cooking. J Agric Food Chem 48(12):598694. Marsono Y, Topping DL. 1999. Effects of particle size of rice on resistant starch and SCFA of the digesta in caecostomised pigs. Indonesian Food Nut Prog 6(2):4450. Martinez-Flores HE; Yoon-Kil-Chang, Martinez-Bustos F, Sgarbieri V. 2004. Effect of high fiber products on blood lipids and lipoproteins in hamsters. Nutr Res 24(1):8593. McNaught KJ, Maloney E, Brown II, Knight AT. 1994. High amylose starch and resistant starch fractions. Intl. Patent WO 94/03049. Feb 17, 1994. Goodman Fielder Ingredients Li (AU); McNaught KJ (AU), Maloney E (AU), Brown IL (AU), Knight AT (AU). McCleary BV, Monaghan DA. 2002. Measurement of resistant starch. J Assoc Off Anal Chem Int 85:66575. McCleary BV, Rossiter P. 2004. Measurement of novel dietary fibres. J Assoc Off Anal Chem Internat 87(3):70717. Mercier C. 1980. Structure and digestibility alterations of cereal starches by twin screw extrusion cooking. In: Malkki LY, Olkku J, Larinkari J, editors. Food process engineering. Vol. 1. London: Applied Science Publishing. p 795807. Mitsuda H. 1993. Retrogradation of cooked rice. J Food Qual 16:3215. Mora-Escobedo R, Osorio-Diaz P, Garcia-Rosas MI, Bello-Perez A, Hernandez-Unzon H. 2004. Changes in selected nutrients and microstructure of white starch quality maize and common maize during tortilla preparation and storage. Food Sci Technol Int 10(2):7987. Morais MB, Feste A, Miller RG, Lifichitz CH. 1996. Effect of resistant starch and digestible starch on intestinal absorption of calcium, iron and zinc in infant pigs. Paediatr Res 39(5):8726. Morita T, Kasaoka S, Hase K, Kiriyama S. 1999. Psyllium shifts the fermentation site of high-amylose cornstarch toward the distal colon and increases fecal butyrate concentration in rats. J Nutr 129:20817. Moron D, Melito C, Tovar J. 1989. Effect of indigestible residue from foodstuffs on trypsin and pancreatic -amylase activity in vitro. J Sci Food Agric 47(2):1719. Muir JG, ODea K. 1992. Measurement of resistant starch. Factors affecting the amount of starch escaping digestion in vitro. Am J Clin Nutr 56:1237. Muir JB, Walker KZ, Kaimakamis MA. 1998. Modulation of fecal markers relevant to colon cancer risk: a high-starch Chinese diet did not generate expected beneficial changes relative to a Western-type diet. Am J Clin Nutr 68:3729. Myung-Hee-Kim, Jeong-Ok-Kim, Mal-Shick-Shin. 2001. Effects of resistant starches on the characteristics of sponge cakes. J Korean Soc Food Sci Nutr 30(4):6239. Niba LL. 2003. Effect of storage period and temperature on resistant starch and beta-glucan content in cornbread. Food Chem 83(4):4938. Nugent AP. 2005. Health properties of resistant starch. Br Nutr Foundation Nutr Bull 30:2754 Ohkuma K, Hanno Y, Inada K, Matsuda I, Katta Y. 1994. Indigestible dextrin. U.S. patent 5,364,652. Nov 15, 1994. Hyogo, Japan: Matsutani Chemical Industries Co. Ltd. Ohkuma K, Hanno Y, Inada K, Matsuda I, Katta Y. 1995. Indigestible dextrin. U.S. patent 5,472,732. Dec 5, 1995. Hyogo, Japan: Matsutani Chemical Industries Co. Ltd. Ohkuma K, Wakabayashi S. 2001. Fibersol-2: Soluble, non-digestible, starch-derived dietary fiber. In: McCleary BV, Prosky L, editors. Advanced dietary fibretechnology. Oxford, U.K.: Blackwell Science. p 50922. Parchure AA, Kulkarni PR. 1997. Effect of food processing treatments on generation of resistant starch. Intl. J Food Sci Nutr 48:25760. Platel K, Shurpalekar KS. 1994. Resistant starch content of Indian foods. Plant Foods Human Nutr 45(1):915. Po YL, Czuchajowska Z, Pomeranz Y. 1994. Enzyme-resistant starch in yellow layer cake. Cereal Chem 71(1):6975. Pomeranz Y, Sievert D. 1990. Purified resistant starch products and their preparation. WO 9015147. Dec 13, 1990. Univ of Washington. Prosky L, Schweizer TF, De Vries JW, Furba I. 1985. Determination of insoluble, soluble and total dietary fiber in food and food products: interlaboratory study. J Assoc Off Anal Chem 71:101723. Rabe E, Sievert D. 1992. Effects of baking, pasta production and extrusion cooking on formation of resistant starch. Eur J Clin Nutr 46(Suppl):S1056. Raben A, Tagliabue A, Christensen NJ, Madsn J, Holst JJ, Astrup A. 1994. Resistant Starch: the effect on postprandial glycemia, hormonal response and satiety. Am J Clin Nutr 60:54451. Ranhotra GS, Gelroth JA, Astroth K, Eisenbraun GJ 1991a. Effect of resistant starch on intestinal responses in rats. Cereal Chem 68(2):1302. Ranhotra GS, Gelroth JA, Eisenbraun GJ. 1991b. High-fiber white flour and its use in cookie products. Cereal Chem 68(4):4324 Ranhotra, GS, Gelroth JA, Glaser BK. 1996. Energy value of resistant starch. J Food Sci 61(2):4535. Reader D, Johnson ML, Hollander P, Franz M. 1997. Response of resistant starch in a food bar vs. two commercially available bars in persons with type II diabetes mellitus. Diabetes 46(1):254A. Reussner G, Andros J, Thiessen R. 1963. Studies on the utilization of various starches and sugar in the rat. J Nutr 80:291. Ring SG, Gee JM, Whittam M, Orford, P, Johnson I. 1988. Resistant Starch. Its chemical form in foodstuffs and effect on digestibility in vitro. Food Chem 28:97109. Rosado JL, Morales M, Allen LH. 1987. Energy and macronutrient bioavailability from rural and urban Mexican diets. In: Southgate DAT, Johnson I, Fenwick GR, editors. Nutrient availability: chemical and biological aspects. Cambridge, U.K.: Royal Society of Chemistry. p 3279. Sajilata MG, Singhal RS. 2005. Specialty starches for snack foods. Carbohydr Polym 59:13151. Saura-Calixto F, Goni I, Bravo L, Manas E 1993. Resistant starch in foods: modified method for dietary fiber residues. J Food Sci 58(3):6423. Schweizer, TF, Anderson H, Lankilde AM, Reimann S, Torsdottir I. 1990. Nutrients excreted in ileostomy effluents after consumption of mixed diets with beans or potatoes II. Starch, dietary fiber and sugars. Eur J Clin Nutr 44:56775. Seib PA, Kyungsoo W. 1999. Food grade starch resistant to -amylase and method of preparation. U.S. patent 5,855,946. Jan 5, 1999. Manhattan, Kans.: Kansas State Univ Research Foundation. Shi YC, Trzasko PT. 1997.Process for producing amylase resistant granular starch. U.S. patent 5, 593, 503. Jan 14, 1997. Wilmington, Del.: National Starch and Chem Investment Holding Corp. Sievert D, Czuchajowska Z, Pomeranz Y 1991. Enzyme-resistant starch. III. X-ray diffraction of autoclaved amylomaize VII starch and enzyme-resistant starch residues. Cereal Chem 68(1):8691. Sievert D, Pomeranz Y 1989a. Enzyme-resistant starch. I. Characterization and evaluation by enzymatic, thermoanalytical, and microscopic methods. Cereal Chem 66(4):3427. Sievert D, Pomeranz Y. 1989b. Enzyme-resistant starch II. Characterization and evaluation by enzymatic, thermoanalytical and microscopic methods. Cereal Chem 66(4):3427. Sievert D, Pomeranz Y. 1990. Enzyme-resistant starch. II. Differential scanning calorimetry studies on heat-treated starches and enzyme-resistant starch residues. Cereal Chem 67(3):21721. Sievert D, Wursch P. 1993b. Thermal behavior of potato amylase and enzyme-resistant starch from maize. Cereal Chem 70:3338. Siljestrom M, Asp NG. 1985. Resistant starch formation during baking. Effect of baking time and temperature and variation in the recipe. Z Lebensm Unters Forsch 4:118. Siljestrom M, Bjorck I. 1990. Digestible and undigestible carbohydrates in autoclaved legumes, potatoes and corn. Food Chem 38:14552. Siljestrom M, Eliasson AC, Asp NG. 1989. Characterisation of resistant starch from autoclaved wheat starch. Starch/Starke 41:14751. Soral SM, Wronkowska M. 2000. Resistant starch of pea origin. Zywnosc 7(2):20412. Southgate DAT. 1989. Conceptual issues concerning the assessment of nutrient bioavailability. In: Southgate DAT, Johnson IT, Fenwick GR, editors. Nutrient availability: chemical and biological aspects. Cambridge, U.K.: Royal Society of Chemistry. p 102.

16

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Resistant starch - a review


Szczodrak J, Pomeranz Y. 1991. Starch and enzyme-resistant starch from high-amylose barley. Cereal Chem 68(6):58996. Szczodrak J, Pomeranz Y. 1992. Starch-lipid interactions and formation of resistant starch in high-amylose barley. Cereal Chem 69(6):62632. Takeda C, Takeda Y,Hizukuri S. 1989. Structure of amylomaize amylase. Cereal Chem 66:225. Tester RF, Karkalas J, Qi X. 2004. Starch structure and digestibility. Enzyme-substrate relationship. World Poultry Sci J 60(2):18695. Tharanathan M, Tharanathan RN. 2001. Resistant starch in wheat-based products: isolation and characterisation. J Cereal Sci 34:7384. Tharanathan RN, Mahadevamma S. 2003. Grain legumes-a boon to human nutrition. Trends Food Sci Technol 14:50718. Thed ST, Phillips RD. 1995. Changes of dietary fiber and starch composition of processed potato products during domestic cooking. Food Chem 52(3):3014. Thompson DB. 2000. On the non-random nature of amylopectin branching. Carbohydr Polym 43:22339. Thompson LU, Yoon JH. 1984. Starch digestibility as affected by polyphenol and phytic acid. J Food Sci 49:12289. Tovar J, Bjoerck IM, Asp NG. 1990b. Analytical and nutritional implications of limited enzymic availability of starch in cooked red kidney beans. J Agric Food Chem 38(2):48893. Tovar J, Melito C. 1996. Steam-cooking and dry heating produce resistant starch in legumes. J Agric Food Chem 44(9):26425. Visser RGF, Suurs LCJM, Bruinenberg PM, Bleeker I, Jacobsen E. 1997. Comparison between amylose-free and amylose containing potato starches. Starch/Strke 49:43843. Westerlund E, Theander O, Andersson R, Aman P. 1989. Effects of baking on polysaccharides in white bread fractions. J Cereal Sci 10(2):14956. Woo KS, Shin MS, Seib PA. 1999. 49 Cross-linked, type RS (4) resistant starch: Preparation and properties. Seattle, Wash.: AACC Annual Meeting; 1999 Oct 31 Nov 3. Manhattan, Kans.: Dept. of Grain Science and Industry, Kansas State Univ. Wu HC, Sarko A. 1978. The double helical molecular structure of crystalline Aamylose. Carbohydr Res 61:7. Wurzburg OB.1995. Modified starches. In: Stephen AM, editor. Food polysaccharides and their applications. New York: Marcel Dekker Inc. p 6797. Yue P, Waring S. 1998. Functionality of resistant starch in food applications. Food Aust 50(12):61521.

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

17

You might also like