2,400
Views
2
CrossRef citations to date
0
Altmetric
Review Article

Effects of Different Processing Methods on Pulses Phytochemicals: An Overview

, , , , , & show all

ABSTRACT

Pulses are edible seeds belonging to the Leguminosae family for human consumption and consist of various species such as common beans (Phaseolus vulgaris L.), peas (Pisum sativum), lentils (Lens culinaris), chickpeas (Cicer arietinum L.), and faba beans (Vicia faba). Pulses are sustainable sources of nutritional compounds, especially containing almost twice the protein content compared to cereal grains. In addition to becoming an excellent source of macronutrients and micronutrients, they are abundant in phytochemicals, containing bioactive compounds with potential health benefits resulting from various phenolic compounds, as well as antinutritional compounds (e.g. phytic acid, enzyme inhibitors, lectins, saponins), which have received widespread concern by researchers. More essentially, various processing approaches for consumption purposes will result not only in enhanced nutritional and sensory characteristics in pulses but also in affected phytochemicals contents and their bioavailability. In this review, the nutritional and phytochemical compositions of pulses will be first introduced, followed by different common-applied processing methods (thermal and non-thermal), along with their impacts and pulse storage effects on the content of corresponding phytochemicals. Furthermore, through food processing and digestion, the bioaccessibility and bioavailability of pulse phytochemicals will be improved, thus releasing more health benefits, and expressing pharmacological functions in the human body.

Introduction

Pulses are a subclass of legumes, which are included in the Leguminosae family that yield edible seeds for human and animal consumption.[Citation1] Pulses consist of various species, including the common beans (Phaseolus vulgaris L.), peas (Pisum sativum), chickpeas (Cicer arietinum L.), cowpeas (Vigna unguiculata), lentils (Lens culinaris), faba beans (Vicia faba), pigeon peas (Cajanus cajan), as well as some others, in terms of worldwide production and consumption, are dominant.[Citation2,Citation3] Typically, their consumption forms involve whole seeds, split grains, dehulled split grains, and flour, thus meeting different needs for consumers.[Citation2]

According to FAO statistics, the average annual consumption of pulses per person was projected to be between 2.7 and 10.8 kg in 2011.[Citation4] The share of different regions globally in the production of significant pulses is different (). Even though pulses are grown everywhere in the world, almost half of it is concentrated in South Asia and sub-Saharan Africa.[Citation2] Besides, the dominant importing countries of different classes of pulses are different. While the pulse production scale is relatively tiny compared to cereal output, their annual average yield has been rising for years.[Citation7] The annual global production of pulses had increased to almost 77 million metric tonnes in three years ending in 2014, amounting to over 2.8% of global cereal production. From the perspective of global pulse production (according to 2014 statistics), common beans ranked first among all major pulse species (31.2%) and chickpeas ranked second (16.9%), subsequently followed by peas (14.3%), cowpeas (9.1%) and other species.[Citation2]

Table 1. Pulse classification, production, consumption, and diets.

Pulses are generally recognized as a vital ingredient for human nutrition and health.[Citation5] Due to their high protein content (considerably higher than that of cereals), dietary fibre, starch, minerals, and vitamins, pulses have become a staple food and inexpensive source of protein in developing nations where protein-energy malnutrition (PEM) is common. Low-income groups of people especially prefer pulses because of their high protein quality. Due to the large regional differences, in several countries, pulses have become an essential dietary constituent, including Kenya (17.0 per person annually), India (15.4 per person annually), and Turkey (13.6 per person annually).[Citation7] In addition, in terms of the amino acid profile, cereal proteins could supply their abundant methionine and cysteine, which pulses lack. In contrast, pulses could contribute with their richer lysine which cereals lack.[Citation7] Therefore, especially for vegetarians, a complete complement of amino acids (or plant-based diets), and balanced nutrition, can be achieved through the combination of pulse and cereal proteins.[Citation18] Moreover, pulses and pulse products such as pulse flour can be utilized as gluten-free diets to improve nutrition for populations with unique demands (e.g., celiac disease).[Citation17]

Pulse is not only rich in nutritional compounds but also various phytochemicals, including phenolic compounds, alkaloids, phytosterols, saponins, enzyme inhibitors, and others. These phytochemicals are also related to health benefits, including anti-obesity, antidiabetic, anti-inflammatory, and antioxidant potential.[Citation19–21] In addition, various processing methods have been applied to produce pulse and pulse-based products due to storage and nutrient enhancement needs. However, individual processing methods have different specific influences on pulse phytochemicals, including their contents, bioaccessibility, and bioavailability. Therefore, this review aims to provide pulse nutritional composition along with a phytochemical profile, including bioactive compounds and antinutrient compounds. Meanwhile, different processing effects on pulse phytochemicals, the bioaccessibility, bioavailability, as well as health benefits on the human body, will be summarized.

Nutritional composition of pulses

As demonstrated in , pulses are abundant in various nutritional compounds. Primarily, they provide a rich source of protein with contents of 19–30%.[Citation18] The protein composition of the pea, lentil, and chickpea is comparable to that of cowpeas and beans.[Citation18] Albumins (<20%) and globulins (>50%) are the two dominant types of proteins found in pulses. Depending on the pulse species, lower amounts of glutelin and prolamins are found. Pulses also contain necessary amino acids like lysine, cysteine, and methionine.[Citation18] Pulses are rich in the amino acid lysine; however, cereal grains are frequently deficient in this amino acid. At the same time, pulses have low levels of the vital amino acids methionine and tryptophan, which can be presented in products made from grains.[Citation15] As a result, pulses and grain-based products complement each other when consumed with cereal to offer a higher quality protein.

Table 2. Nutritional composition of different pulse sources (raw/unprocessed seeds).

Relatively low in energy density, pulses after the cooking process provide a 1.3 kcal/g average.[Citation15] They have a higher carbohydrate content (up to 50%–65%) but are absorbed more slowly, giving them a lower glycemic index (GI) than other carb-rich meals like rice, white bread, or potatoes.[Citation56] Insoluble and soluble dietary fibre are both abundant in pulses, which offer about 7 g per half-cup serving, while the amount varies depending on the pulse species.[Citation41]

According to Hall, et al.,[Citation18] pulses typically have a lipid content less than 3%, whereas chickpea and lupin can have as high as 7% and 13%, respectively. In terms of the lipid fraction of pulses, it is characterized by a high level of triglycerides, along with a higher concentration of monounsaturated fatty acids (oleic acid, C 18:1n-9) and polyunsaturated fatty acids (linoleic acid, C 18:2n-6, and alpha-linoleic acid C 18:3n-3) that are both beneficial components.[Citation57] Recent research has revealed an intriguing n-6/n-3 PUFA ratio, emphasizing lentils and azuki beans with values of 4.0 and 3.2, respectively, which provide health benefits like lowered CVD risk.[Citation58]

Further, pulses are a good source of micronutrients, containing various minerals and vitamins.[Citation59] They include B-group vitamins, such as thiamine, niacin, folate, riboflavin, and pyridoxine, and are an excellent source of selenium.[Citation15] Moreover, pulses are a good origin of iron and zinc. For instance, a half-cup portion of beans contains over 10% of the daily recommended amount of iron, even though the amount in different types of beans might vary significantly.[Citation60]

Phytochemical composition of pulses

According to Neilson, et al.,[Citation61] plant bioactive compounds are dominantly secondary metabolites, which means they are not necessary for plant organisms’ survival. In detail, secondary metabolites include specific substances that give the plant advantages in survival and competition. Besides, “natural products” or “phytochemicals” are other names for plant secondary metabolites. Alkaloids, terpenoids, as well as phenolics/polyphenols, are the three main types of phytochemicals found in plants. Minor classes possess polyacetylenes, miscellaneous pigments, cyanogenic glucosides, glucosinolates, and nonprotein amino acids. As summarized in , pulses enrich various bioactive compounds, such as phenolic compounds, phytosterols, and non-digestible carbohydrates, all of which have essential physiological and metabolic effects.[Citation103] Many recent reviews also address the main phytochemical compounds in different pulses.[Citation103,Citation104]

Table 3. Contents of phytochemicals in different pulse sources.

As one of the dominant groups of pulses’ bioactive compounds, the aromatic rings of phenolic compounds have one or more hydroxyl substituents, resulting in simple to highly complex polymers.[Citation105] They have a broad therapeutic potential attributed to their antioxidant capabilities, which are due to their ability to act as reducing agents, hydrogen donors, and neutralizers of free radicals according to their ox-reduction potential.[Citation106] The primary phenolic components found in pulse seeds are condensed tannins, phenolic acids, and flavonoids, which are classified as flavanols, flavan-3-ols, and anthocyanins.[Citation103] Besides, phenolic acids discovered in pulses are either cinnamic acid (e.g., caffeic, ferulic, sinapic, and p-coumaric acid) or benzoic acid (e.g., gallic, syringic, and vanillic acid) derivatives, found as esters of caffeic and quinic acids.[Citation42] Flavonoids share the diphenylpropanes (C6-C3-C6) structure, consisting of two aromatic rings connected by three carbons and typically forming an oxygenated heterocycle.[Citation107] Isoflavones, in which the B ring of the flavone molecule is connected to carbon 3, are particularly common in pulses. While tannins are highly hydroxylated, forming insoluble complexes with proteins and carbohydrates.

Field bean (Vicia faba) hulls have been measured to contain condensed tannins (proanthocyanidins), which are also found in the seeds of coloured-flowered cultivars of peas.[Citation108] Further, the phenolic compositions in pulses are primarily found in the seed coat, while the cotyledon part has comparatively minor concentrations.[Citation103] Regarding the common bean, phenolic compounds are primarily found in the seed coat. The presence of flavonol glycosides, anthocyanins, and condensed tannins is principally responsible for the seed colour. Hou, et al.[Citation109] also concluded that the abundance of phenolic compounds present in pulse seed coats was highest in common beans, followed by lentils and peas, while least in mung bean and lupin. Additionally, pulses with a deep colour and a lot of pigment, including black gram (Vigna mungo) and red kidney beans (Phaseolus vulgaris), are indicated with higher phenolic concentrations.[Citation110,Citation111] The total phenolic content (TPC) also varies among different pulse varieties. For instance, in Xu, et al.[Citation112] study, the lentils and kidney bean samples (47.6 and 45.7 mg/g DW, respectively) showed the highest phenolic content, while lima bean showed the lowest (9.5 mg/g DW).

A group of incredibly diverse substances known as alkaloids typically consist of a heterocycle with a nitrogen atom inside the cycle. This confirmation gives the molecule a primary character; in an aqueous solution, the molecule tends to pick up a proton unless the nitrogen atom is adjacent to an electron acceptor (e.g. ricinine).[Citation42] Lupins possess the majority of them (i.e., lupanine, lupinine, sparteine, and multiflorine), but developing alkaloid-free variants has boosted the number of lupins in fodder for domestic cattle.[Citation113] Additionally, alkaloids were reported to be present in peas as well as in tiny amounts in chickpeas and lentils.[Citation57] Lupanine has recently been identified to increase insulin secretion, as reported by Wiedemann, et al.[Citation114]; as a result, it may be relevant for the supportive treatment of diabetes mellitus.

In addition to being significant bioactive components of pulses, phytosterols, plant sterols, and stanols are the equivalent ingredients to cholesterol present in mammals.[Citation115] These compounds attach to bile acids and stop them from being reabsorbed.[Citation103]

Pulses contain trace amounts of phytosterols, the most prevalent of which are β-sitosterol, campesterol, and stigmasterol.[Citation116] The pulses have a total phytosterol concentration that ranges from 134 mg/100 g (kidney bean) to 242 mg/100 g (pea), as Ryan, et al.[Citation86] studied. Total β-sitosterol content ranged between 85 mg/100 g in butter beans and 160 mg/100 g in chickpeas.[Citation86]

Many other bioactive phytochemicals have been reported in pulses. For instance, pulses contain higher oligosaccharides compared to other crops; among them, α-galactosides are the dominant ones, including raffinose, stachyose, and verbascose.[Citation117] Based on Han and Baik[Citation118] investigation, total oligosaccharides in different pulses ranged between 70.7 mg/g DW (in yellow pea seeds) and 144.9 mg/g DW (in chickpea seeds).

Additionally, they include more significant levels of inaccessible or indigestible carbohydrates, such as resistant starch and dietary fiber, leading to a lower glycemic index (GI). Furthermore, pulses contain tocopherol with higher antioxidant potential (up to 52 mg/100 g in lupine seeds as reported).[Citation18]

Anti-nutritional compounds

Pulses contain various antinutritional compounds, including protease inhibitors, lectins, phytic acid, saponins, condensed tannins, and others, which can reduce the bioavailability and digestibility of other nutritional compounds and exceptionally, could be the cause of diseases.[Citation119] Protease inhibitors (such as trypsin and chymotrypsin) and amylase inhibitors are the two most common enzyme inhibitors of pulses (). Proteases, amylases, lipases, glycosidases, and phosphatases can be inhibited by protein hydrolases found in pulses.[Citation120] Common beans (Phaseolus vulgaris) are the second-largest group of plant seeds after cereals that have been identified as natural sources of alpha-amylase inhibitors, which decreases the utilization of dietary starch and protein, potentially acting as a treatment for diabetes while also reducing the digestibility of dietary proteins.[Citation121] In addition, protease inhibitors also suppress the digestibility and bioavailability of nutrients.[Citation120]

Most plant foods include lectins or haemagglutinins, and pulses are the primary source of lectins in daily meals for humans. There have been reports of kidney beans showing a high level of lectins (840 × 10−5 hemagglutinating activity units (HU)/kg) and lupin seeds having a relatively low level (3 × 10−5 HU/kg).[Citation122] Due to some lectins disrupting the integrity of the intestinal epithelium, which affects the absorption and utilization of nutrients, some pulses’ lectins decrease the growth of experimental animals and diminish the digestibility and biological value of dietary proteins.[Citation123]

According to Martín-Cabrejas,[Citation57] pulses store phosphorus mainly in the form of phytic acid (IP6), which can account for 3–5% of their total phosphorus content. Since phytic acid is a potent chelating agent and decreases the bioavailability of primarily divalent cations like Zn2+, Fe2+, Mn2+, Mg2+, and Ca2+, it is regarded as a non-nutritive component. Insoluble metal-phytic complexes are generated under the digestive tract, decreasing their digestive use and subsequent assimilation in both animals and people. Along with interfering with digestive enzymes like pepsin, trypsin, and α-amylase, IP6 can also interact with proteins to produce phytate-protein complexes that impair the solubility and digestibility of dietary proteins.[Citation124] Based on Oomah, et al.,[Citation125] phytic acid makes up an average of 83% of all inositol phosphates, with the percentage varying from 77% in chickpea to 88% in black beans. Additionally, it tends to be present in larger concentrations in raw dry beans, black eye pea, and pigeon pea than in lentil, green split pea, and chickpea. On the other hand, pulses also contain significant amounts of phytate, the salt form of phytic acid found in the endosperm’s protein bodies.

Saponins are another typical group of antinutritional compounds widely distributed in various pulses, such as chickpea, soybean, lentil, phaseolus bean, and the alfalfa sprout.[Citation42] The primary sources of saponins in the human diet are indicated as soybean and chickpea.[Citation125] Moreover, saponins can reduce the availability of nutrients and limit enzyme activity, which can have a growth-retarding effect on animal growth.[Citation126]

Furthermore, other antinutritional components in pulses include oxalates (which can lower mineral bioavailability like phytates), phenolic compounds (including isoflavones), condensed tannins (which decrease food intake in animals), lignans (phyto-estrogens), and lignins (which reduce fermentability of dietary fibres), among others.[Citation42]

Effect of processing on pulse phytochemicals

Through various thermal and non-thermal processing methods, the phytochemical compounds in pulses can be influenced, particularly in terms of phenolic compounds and antinutritional compounds. Thermal processing can quickly eliminate partial heat-labile antinutritional factors, such as protease inhibitors and lectins.[Citation119] In contrast, other antinutritional compounds, such as phytic acid, tannins, and saponins, withstand heat. Coda, et al.[Citation127] concluded that dehulling, soaking, air classification, extrusion, and cooking are the primary technological methods for reducing the adverse effects of antinutritional compounds on pulse consumption. However, biological techniques tend to be more effective, including germination, enzyme treatments, and particularly fermentation. In the following sub-sections, various thermal and non-thermal processing methods and their effects on phytochemical profiles and contents in pulse seeds will be introduced ().

Table 4. Effect of different processing methods on pulse phytochemical compounds.

Soaking

As one basic processing approach, soaking can soften the cell walls of the pulse seeds, thus facilitating the subsequent cooking process.[Citation177] The concentration of relevant phytochemical compounds will hence be influenced by soaking processing. Primarily, due to pertinent chemicals seeping into the soaking water during this stage, polyphenols will become soluble and lose their potency, thus causing a reduced general content of total phenolic contents (TPC).[Citation177] For example, Xu and Chang[Citation76] reported a 4.9–11.5% reduction in TPC in green peas and 9.5–37.8% in lentils after the soaking process (50, 70, 85, 100% hydration). Similar decreasing trends in TPC in five soaked different faba bean cultivars were also observed by Siah, et al..[Citation167] Secondly, soaking can remarkedly influence the levels of certain phenolic compounds. Soaking pulses can effectively reduce the tannin content due to the diffusion effect into the water.[Citation178] Alonso, et al.[Citation129] reported a reduction of 47.7% and 24.2% in condensed tannin content in faba bean and kidney bean, respectively; Khandelwal, et al.[Citation136] also reported around 22% reduction in total tannins in red gram and Bengal

However, some studies showed the reverse effects of soaking on pulse phenolic compounds. Lentils were shown to contain more dihydroxybenzoic acid after being soaked, rising from 3.68 μg/g to 39.20 μg/g.[Citation135] And it was also demonstrated by Aguilera, et al.[Citation179] that flavonols in Sinaloa chickpea were increased by the soaking process. Not only soaking alone but also soaking with cooking on the different phenolic compounds has been studied and reported. Luthria and Pastor-Corrales[Citation180] observed that while overnight soaking samples showed a considerable loss in total phenolic acids, following cooking of bean samples retained more than 83% of the total phenolic acids.

Soaking of pulse seeds causes the reduction of phytate and phytic acid in general, mainly attributed to its water solubility, seeping into the soaking water, as well as the endogenous phytase activation.[Citation181] Khattab and Arntfield[Citation130] investigated the phytic acid of studied pulses (Canadian and Egyptian cowpea, kidney bean, and pea) reduced significantly after soaking. Shimelis and Rakshit[Citation131] illustrated that compared to the raw control, soaked kidney beans had a 14–18% reduction in phytic acid.

Moreover, soaking can significantly influence the level of other antinutrient and bioactive compounds. The soaking process results in the aqueous extraction of saponins because of their high water solubility.[Citation182] For instance, as Sharma and Sehgal[Citation137] reported, two of the faba bean cultivars that were studied had a 20–23% drop in saponins after soaking for 12 hours at 37°C. In addition, as for pulses, soaking can lower the concentration and activities of enzyme inhibitors such as trypsin, α-amylase, and chymotrypsin inhibitors.[Citation93] This phenomenon might be resulted from relevant inhibitors leaking into soaking water during treatment.[Citation183] In Shi, et al.[Citation93] study, after soaking, the level of α-amylase inhibitor in common beans was reduced by 4–10%, and that in soybean decreased by 4%. Simultaneously, trypsin inhibitor was significantly reduced by soaking in studied pea (17.34–30.74%), lentil (5.57–19.35%), faba bean (12.73–22.59%), chickpea (9.39–25.27%), common bean (4.88–9.09%), and soybeans (18.58%). Moreover, soaking also causes reduction of other phytochemicals, such as oligosaccharides (sucrose, stachyose and raffinose) in soybean, cowpea and ground bean samples.[Citation184]

Extrusion

Utilizing the mix of moisture, pressure, temperature, and mechanical shear, extrusion processing creates food products with distinctive physical and chemical properties, along with shortened cooking time and enhanced textural, nutritional, dietary, and sensory attributes of food ingredients at the same time.[Citation185] Tas and Shah[Citation186] indicated that incorporating pulse-based flours into extruded cereal-based snacks exerts better nutritional benefits (e.g., high protein and dietary fibre, low fat, elevated resistant starch levels) and functional properties (e.g., sensory appeal, textural attributes).

Extrusion can not only enhance the digestibility of starch and protein but also significantly impact the elimination of antinutritional factors, including trypsin inhibitors, lectins, phytic acid, and tannin.[Citation187,Citation188] As shown in , Alonso, et al.[Citation129] indicated that the best treatment for removing trypsin, chymotrypsin, α-amylase inhibitors, and hemagglutinin activity without changing the protein composition was extrusion processing. Rathod and Annapure[Citation189] extruded lentils (18% moisture, 160°C, 200 rpm), and discovered that the final product had trypsin inhibitor, phytate, and tannin concentrations that had been reduced by up to 99.54, 99.30, and 98.83%, respectively. Besides, according to Pasqualone, et al.,[Citation190] observed a remarkable decrease in trypsin inhibitor levels after extrusion and attributed this result to heat and intense mechanical stress. Moreover, trypsin inhibitors can be significantly inactivated and destructed by extrusion at higher temperatures or by lengthening the residence period when extruding at lower temperatures.[Citation191]

The impact of extrusion cooking on different groups of phenolic compounds may vary. Several studies have proved the significant reduction of total phenolic compounds in extruded pea, kidney bean, faba bean, and chickpea.[Citation129,Citation139,Citation192] The excessively high temperatures may cause the unfavorable effects of extrusion cooking on relevant phenolic compounds.[Citation193] On the other hand, some experiments reported increasing reverse impacts. In Arribas, et al.[Citation194] study, an increase in total phenolics and anthocyanins was observed in the extrudates made from pea, rice, and carob flour. It was demonstrated that if phenolics are mostly bound to dietary fibre in the cell walls in the starting flour, as they are in carob, then the extrusion process will tend to partially disrupt the fibre, thus releasing relevant phenolic compounds.[Citation195,Citation196]

Furthermore, some other phytochemical compounds can also be affected during the extrusion process. For instance, when compared to raw materials, extrusion cooking led to a considerable increase in the concentration of total α-galactosides of up to 85%.[Citation141] Lentil extrudates produced at 160°C had a greater total α-galactoside concentration than those produced at 140°C.[Citation141] This might be related to the fact that oligosaccharides are relatively heat-stable.[Citation197] Therefore, during the extrusion procedure, the cell walls undergo mechanical-structural changes, which may enhance their availability in the final extrudates.[Citation190] However, few studies indicate that extrusion processing affects vitamins’ stability in pulses. Lorenz, et al.[Citation198] have suggested that extrusion processing can successfully retain up to 80% of the vitamin C in the studied pulse sample.[Citation199]

Germination

Germination is usually applied to effectively reduce the anti-nutrient content of grains and pulses and simultaneously alters the food products’ nutritional level, biochemical properties, and physical characteristics.[Citation200,Citation201] The enzyme phytase in pulse and other plant seeds is often activated during germination, which destroys phytate and lowers the content of phytic acid in the final products.[Citation202] Different studies demonstrated that germination is a more efficient processing method than thermal processing for eliminating phytic acid in pulses, including chickpeas, peas, and faba beans.[Citation147,Citation203] For instance, in Khalil,[Citation147] the losses of phytic acid were up to 75% in germinated faba beans, significantly higher than in soaked (27%), autoclaved (49%), or cooked (40%) samples.

The TPC, tannin content, and other phenolic compounds can also be significantly influenced during germination processing. For instance, Guo, et al.[Citation145] reported a marked increase in TPC, free phenolics, bound phenolics, and Quercetin-3-O-glucoside in smung bean after germination. This might have occurred because, in comparison to raw seeds, germination can lead to the gradual accumulation of soluble phenolics in pulse sprouts and seeds.[Citation204] Moktan and Ojha[Citation146] also reported increased TPC (from 46.53 to 52.33 mg GAE/g) and antioxidant capacities in germinated horse grams. Moreover, in Mamilla and Mishra[Citation205] experiment, germination altered the TPC in chickpea and red lentil samples; while the TPC in kidney beans showed non-influence, which may be due to the reduced water absorption caused by increased polyphenol oxidase activity and a thicker seed coat.

Germination can also affect other antinutritional and certain bioactive substances in pulses. Although not as severe as thermal treatment, germination can considerably diminish enzyme inhibitor activity in general.[Citation206] In addition, the saponin content of pulses may also be influenced within the germination stage. Guajardo-Flores, et al.[Citation150] demonstrated that the concentration of total saponins in sprouts and cotyledons approximately doubled after the 1-day germination; however, the amount of the most abundant soybean saponins in germinated black bean sprouts decreased as germination days went on. Moktan and Ojha[Citation146] also claimed the reduced oxalate (from 3.18 to 1.73 mg/g) in horse grams due to germination. Furthermore, Shohag, et al.[Citation142] reported a remarkably increased total folate content in germinated soybean and mung bean samples. And Mubarak[Citation207] found the mineral content was effectively retained in mung bean seeds with decreased levels of stachyose and raffinose after the germination process.

Roasting

Roasting is a traditional thermal processing method that can suppress the adverse effects of antinutritional factors in pulses. Roasting is effective in removing total phenolic compounds (from 1450 to 1440 mg/100 g d.w.) but less effective than the pressure-cooking method (940 mg/100 g d.w.) in mangrove legume Canavalia cathartica, as Seena, et al.[Citation154] reported. Boateng, et al.[Citation151] also found a significant reduction in TPC in studied black eye pea, kidney bean, and pinto bean. However, some reverse results have been reported. Bolek[Citation156] discovered that the roasting process increased TPC and the antioxidant capacity of adzuki beans (Vigna angularis). The rise in phenolic levels during the roasting process may be caused by heat-induced changes in the molecule structures of substances related to the corresponding phenolic compounds.

On the other hand, roasting can significantly reduce other antinutrients. As Seena, et al.[Citation154] indicated, roasting effectively diminishes the hemagglutinin activity in raw seeds. In Aremu, et al.[Citation155] study, the oxalate, tannin, phytate, chemotrypsin, and saponin in roasted bambara groundnut, scarlet runner bean, and lima bean exhibited marked reduction individually compared to their raw samples; while the reducing effects through roasting are not pronounced as boiling. In addition, roasted adzuki beans showed a significant decrease in phytic acid content, saponin content, and trypsin inhibitor activity.[Citation156] Similar reducing effects on trypsin inhibitor activity were detected in Sacha inchi and Tamarindus indica seeds, which may be owing to the denaturation of this inhibitor at an excessive roasting temperature.[Citation208,Citation209]

Moreover, some other bioactive phytochemical compositions will be affected via roasting. For instance, alkaloid and cyanide in the study by Aremu, et al.[Citation155] exhibited a significant reduction after the roasting process in three pulses. Besides, Khattab and Arntfield[Citation130] along with Udensi, et al.[Citation210] noted a significant reduction (24% and 22.63% respectively) in α-galactosides in roasted cowpea and kidney bean seeds. And roasting treatment of faba beans at 120°C for 10 minutes caused a certain extent the elimination of vicine (1–12%) and convicine (3–30%), although not as effective as boiling processing.[Citation211]

Steaming

Steaming is a typical wet thermal processing method applied in food production. Steam processing has several benefits, including a shortened processing time, better retention of antioxidant compounds and activities, and preservation of the appearance and texture of cooked pulses.[Citation134]

Steaming shows significant effects on pulse phenolic compounds. Xu and Chang[Citation134] reported that all the studied pulse samples (green & yellow pea, chickpea, and lentil) exhibited a significant reduction in TPC values after conventional or pressure steaming processing. Simultaneously, the highly retained TPC, DPPH, and ORAC levels in all studied samples were found when compared to the boiling processing. Besides, in another study by Xu and Chang,[Citation76] the conventional and pressure steaming treatments both caused significant (p < 0.05) reduction in procyanidin content for all tested pulses. Meanwhile, the two used steaming methods caused reductions in catechin and flavan-3-ols in individual pulse samples. Nevertheless, Duan, et al.[Citation70] indicated that both faba bean seeds and leaves underwent steaming treatments for 30, 45, and 60 minutes with no discernible changes in TFC, which can be attributed to the destruction of heat-sensitive flavonoid compounds, while the comparatively heat-stable components persisted.

In addition, other pulses’ antinutritional and bioactive compounds can be affected during steaming processing. Ojo, et al.[Citation160] applied atmospheric steaming and pressure steaming on tested Vigna racemose samples and detected a significant decrease in phytic acid, saponin, and tannin. Significantly, the trypsin inhibitor activity diminished by 100% after two steaming methods. As “moin-moin” (a steamed cowpea paste with seasonings) was being produced, the tannins and phytic acid concentrations decreased by 19.6–24.7% and 7.8–14.0%, respectively.[Citation161] As a result of their thermal degradation, the phytate was decreased after heat treatment; this is because phytic acid is naturally thermolabile.[Citation210]

Autoclaving

Autoclaving is one high-pressure cooking approach that employs rapid cycles of high temperature (often operating at 121°C) and pressure (typically operating at 1.8 to 2.0 bar).[Citation212] It is a high-intensity thermal processing method employed in industry to produce canned pulses.[Citation212]

The phenolic compounds in pulses can markedly be influenced by autoclaving treatment. The TPC values in studied beans and peas showed a significant decrease after autoclaving.[Citation166] Jood, et al.[Citation168] reported that autoclaving had the most pronounced effect on phenolic content in chickpea and black gram samples, similar to germination. Significant reductions in total free phenolics (71%) and tannins (82%) were also detected in autoclaved Vigna unguiculata, as Kalpanadevi and Mohan[Citation169] reported. Moreover, Siah, et al.[Citation167] found that while TPC and TFC in autoclaved faba bean samples declined significantly, the total proanthocyanidins in Rossa faba bean showed a slight increase after autoclaving.

Bioactive substances and antinutritional elements, including tannins, phytic acid, and trypsin inhibitors are reduced or eliminated by autoclaving treatment.[Citation212] This method activates the phytase enzyme and enhances acidity in grains, pulses, and other plant-based diets, causing a reduction in phytic acid after autoclaving.[Citation213] In Khattab and Arntfield[Citation130] study, the autoclaving treatment led to a remarkable decrease in tannin and phytic acid content in six pulses. The phytic acid or phytate in faba bean and chickpeas also exhibited a significant reduction in content after autoclaving.[Citation164] In addition, the enzyme’s inhibitory activities may be partially or totally removed by autoclaving due to exposure to excessively high temperature. As Alajaji and El-Adawy[Citation28] reported, autoclaved chickpeas had a substantial decrease in trypsin inhibitor activity. And the trypsin inhibitor activity decreased to zero in autoclaved cowpea, pea, and kidney bean seeds as Khattab and Arntfield[Citation130] studied. Moreover, the α-amylase inhibitor activity in common beans showed a significant decreasing trend under different autoclaving conditions, as Batista, et al.[Citation163] illustrated.

Furthermore, other phytochemicals in pulses can be influenced by autoclaving treatment. Autoclaving caused the highest reduction in oligosaccharides in Khattab and Arntfield[Citation130] research. And the reduction of hydrogen cyanide by up to 64% was reported in Vigna unguiculata subsp. Unguiculata after autoclaving.[Citation169]

Boiling & cooking

Boiling is a typical cooking method and affects the phytochemicals in pulses. Typically, pulses are boiled in water at 100°C for a short period of time. The boiling approach improves the sensory qualities of pulse seeds, tenderizes the seed, and increases customer acceptability. This treatment also aids in the elimination of heat-sensitive antinutritional compounds. Khalil and Mansour[Citation31] indicated that in faba bean seeds, the boiling treatment removed their hemagglutinins. Carlini and Udedibie[Citation173] also found that jack bean required 3 hours of boiling to become lectin-free and 2 hours to eradicate trypsin inhibitor activity. Over 90% of the trypsin inhibitor activity in common beans can be eliminated by cooking them for 60 minutes at 100°C.[Citation211]

Boiling and cooking can significantly influence the phenolic compounds in pulses. Bressani and Elías[Citation176] noted that heating removes 30–40% of the polyphenols from common beans. Tungmunnithum, et al.[Citation170] reported a significant reduction in TPC, TFC, and monomeric anthocyanin content in peas and pigeon peas. In Aguilera, et al.[Citation79] study, some phenolic compounds, such as the hydroxybenzoic in cannellini bean, as well as catechins, procyanidins, and flavonols in pinto bean, all showed a significant decrease after cooking treatment. Moreover, tannin content was reported to decline significantly by cooking in several common beans and chickpeas in Wang, et al.[Citation48] study.

Other bioactive compounds can be influenced during boiling and cooking. For instance, boiling in water for 10 hours was effective in eliminating vicine and convicine in faba beans, as Abd Allah, et al.[Citation214] reported. Besides, cooking treatment generally leads to a lower tocopherol content in pulses compared to the uncooked controls.[Citation78] Moreover, the marked elevation in the carotenoid range was detected in lentils after cooking, along with the production of 13’-cis-lutein and 15-cis-lutein.[Citation215]

Effect of pulse storage on phytochemicals

Post-harvest storage of the pulses can alter their chemical composition, including the phytochemicals, and consequently, their nutritional and functional value will be changed.[Citation216]

Pulse storage has significant effects on their phenolic compounds. Granito, et al.[Citation216] stored common beans under three different conditions: 30°C, 11% humidity (C1), 50°C, 11% humidity (C2), and 50°C, 80% humidity (C3). And under C2 and C3 conditions, the TPC content within the first 90 day-storage increased significantly. The deamination of L-phenylalanine and L-tyrosine, catalysed by the enzyme phenylalanine and tyrosine ammonium lyase, might have caused the rise in TPC. One of phenolic compound derivatives during this process is hydrocinnamic acid, namely ferulic acid.[Citation216] While as for tannin content, it consistently decreased, which may be due to their condensation, migration, and bonding to the cell wall; after 90 days of storage, the TPC started to reduce. However, an inverse result was obtained from Nakitto, et al.[Citation217] study, where the hard-to-cook beans presented greater concentrations of tannin in all studied cultivars compared to the fresh controls, although their TPC showed a decrease after storage. And there may be a connection between these enhanced tannin compounds and low-molecular-weight non-tannin material, indicating post-harvest biochemical activity within the storage process.

Moreover, some other phytochemical compounds can also be influenced during pulse storage. For example, as Amarowicz and Pegg[Citation218] reported, the total ascorbic acid content in peas stored at 4°C remarkably reduced within 14 days; while after 7 days of storage at 20°C, the ascorbic acid content decreased by 72%. In Bento, et al.[Citation219] study, the lignin content significantly increased in all common bean cultivars during adverse storage.

Bioaccessibility, bioavailability, and digestibility of phytochemicals in pulses

According to the literature, the phenolic compounds’ bioaccessibility, i.e., the percentage of a compound liberated from its food matrix in the gastrointestinal tract and subsequently made available for intestinal absorption, determines their bioavailability.[Citation220] Bioaccessibility can be primarily influenced by chemical structure and matrix interactions.[Citation221] According to Nicolás-García, et al.,[Citation177] phenolic compounds are included in cell wall components, in which phenolic acids combine with lignin to produce complexes when the hydroxyl groups on the aromatic ring are bound together. In contrast, insoluble phenolic compounds can interact or bind via covalent bonds with cell wall structural compositions such as cellulose, hemicellulose (including arabinoxylans), lignin, pectin, and structural proteins. Additionally, cellulose and pectin exhibit greater anthocyanin absorption and interaction potential.[Citation177] Other elements that significantly affect the bioaccessibility and bioavailability of phenolic compounds include the substance concentration within the cell wall, changes in the structure of the cell wall, and the position of glycosides within cells.[Citation222]

Most polyphenols found in dietary components are seen as glycosides, esters, or polymers that are commonly difficult to absorb.[Citation223] Before glycosylated phenols are absorbed, the sugar moiety must be eliminated by digestive enzymes (in gastrointestinal mucosa or colonic microflora) or through various food processing methods.[Citation223] The bioaccessibility and bioavailability of phenolic compounds in original pulse seeds are limited (). According to Sancho, et al.,[Citation226] the bioavailability of total phenols in black bean coat and small red bean coat was only 24% and 49%, respectively, and the tannins in each sample were detected as 6% and 7%, respectively. During in vitro digestion, while the free and conjugated phenolic substances are available to be absorbed in human small and large intestines, those covalently bound to indigestible polysaccharides might be absorbed after being released from cell structures via digestive enzymes or microorganisms present in the intestinal lumen.[Citation224] Therefore, the bioaccessibility and bioavailability of relevant phenolic compounds in pulse ingredients will be increased in the digestion process. Chen, et al.[Citation227] noticed a rise in TPC in the intestinal phase compared to the gastric phase during in vitro digestion, indicating that the intestinal phase is primarily responsible for releasing phenolics from cooked lentils. Zhang, et al.[Citation225] noted a similar pattern, finding higher total phenolic, flavonoid, and tannin levels during intestinal digestion than during gastric digestion.

Table 5. Bioaccessibility and bioavailability of phenolics in various pulse samples.

The bioaccessibility and bioavailability of phenolic compounds in pulses will generally improve during various food processing approaches as can be observed in . Lafarga, et al.[Citation221] conducted the in vitro gastrointestinal digestion experiment of several pulses, confirming the enhancement impact of cooking treatment on the bioaccessibility of pulse phenolic compounds. Based on the results, cooked soybeans had the highest concentration of bioaccessible polyphenols, followed by faba beans and lentil samples (p < 0.05).[Citation221] Hithamani and Srinivasan[Citation228] also found that the roasting process significantly increased the bioaccessible polyphenol content in green gram (11%), and the open pan boiling also had an increasing effect. However, Akillioglu and Karakaya[Citation229] reported that soaking caused a significant reduction (25.61% for hot-water soaking and 38.63% for cold-water soaking) in the total phenols bioavailability of common cooked beans, along with decreased total flavonoid bioavailability, which could be a result of relevant phenolic compounds being released into the soaking water.[Citation229]

The bioavailability of some other bioactive phytochemical compounds, such as carotenoids, may be influenced by various pulse processing. McInerney, et al.[Citation230]

demonstrated that very little of the lutein as carotenoids in green beans was digested in vitro (≤14%); besides, the bioavailability of lutein in green beans was raised by pressure cooking at 600 MPa. Oghbaei and Prakash[Citation231] tested the carotenoids’ bioaccessibility in mixed green gram (Vigna radiata) and amaranth greens (Amaranthus caudatus) and found that when temperature decreased with increased moisture (−17°C, 78.4%), the bioaccessible carotenoids was significantly reduced to 0.18 mg/100 g compared to 0.56 mg/100 g in the control sample (28°C, 3.6%).

Health impacts of pulse consumption

Pulse consumption provides various health benefits. They are highly nutritious, containing roughly twice as much protein as cereals, and providing different essential or non-essential amino acids.[Citation18] Starch is the primary source of easily absorbed carbohydrates in pulse seeds and is present in the highest amounts (22–45%), coupled with oligosaccharides (1.8–18%), and dietary fibre (4.3–25%).[Citation232] Pulse consumption also serves various essential vitamin (e.g., vitamin B) and mineral (e.g., Fe, Ca, Zn, Se, Mg, P, Cu, and K) compounds for humans and animals.[Citation233]

Pulse exerts a strong influence in radical scavenging and detoxifying activities; and demonstrates chemo-preventive benefits due to the high level and efficacy of the phytochemical substances. summarizes the biological potential and health impacts of phytochemicals of pulses according to the literature. The phenolic compounds in pulses have been reported to exert various health benefits, which are associated with their antioxidant potential as well as preventive impacts on chronic diseases brought on by free radicals.[Citation243] As the most widely distributed phenolics in pulse seeds, the phenolic acids and flavonoids exhibit significant antioxidant capacities.[Citation244] Also, the pulses’ phenolics are confirmed to be related to other favorable properties such as anti-inflammatory and anti-microbial potential.[Citation19] Pulse can help manage obesity by reducing lipid levels when regularly consumed, due to the lipase inhibitory activity by relevant phenolic compounds.[Citation20] They also provide antidiabetic potential due to α-amylase and α-glucosidase inhibition by suitable phenolic compounds.[Citation235] In Hou, et al.[Citation109] review, pulse seed coats’ prevention effects on the chronic disease were concluded, including antioxidant, anti-obesity, anti-diabetic, anti-inflammatory, anticancer, and intestinal prebiotic capacities confirmed by a series of in vitro and in vivo studies. On the other hand, other bioactive and antinutritional compounds in pulses also exhibit health benefits. For instance, the saponin present in pulses shows potential anticarcinogenic and hypocholesterolemic activity.[Citation233] Additionally, phytates in pulses exhibit beneficial health effects, play a valuable antioxidant role, and provide protection from several cancers, coronary heart disease, renal stones, and diabetes mellitus.[Citation245]

Table 6. Biological potential and health impacts of phytochemicals in pulses.

Conclusion

Pulses, an inexpensive and sustainable source of nutritional compounds, have been utilized as an essential part of the daily human diet in many countries worldwide. Pulses are a source of nutritional compounds such as proteins, carbohydrates, minerals, and vitamins. In addition, they present several phytochemicals, such as bioactive compounds (e.g., phenolic compounds), and antinutritional compounds. Several listed processing methods, including soaking, germination, and thermal cooking, can influence pulse phytochemicals differently, as it depends on the technological processing and the pulse varieties. Generally, each processing can effectively reduce pulse antinutritional compounds (e.g., phytate, enzyme inhibitors, saponins, lectins, and others), hence improving the bioaccessibility, bioavailability as well as digestibility of nutritional compounds in pulses. Simultaneously, relevant phytochemicals can be positively affected under specific processing treatments. Germination helps increase the level of total phenolics and certain flavonoid compounds.

Moreover, relatively mild thermal processing methods, such as steaming, are more effective in retaining antioxidant compounds and activities in addition to inactivating antinutrients, compared to extrusion and autoclaving. In addition, the consumption of pulses exhibits huge health potential for the human body. The presence of bioactive compounds in pulses such as phenolics, phytosterols, and dietary fibre, are correlated with the prevention of chronic diseases, along with their antioxidant, anti-obesity, anti-diabetic, anti-inflammatory, and anticancer capacities. Thus, it is essential for industries as well as researchers to find more effective methods for improving the bioaccessibility of bioactive constituents in pulses nowadays. The effect of different processing methods on the bioaccessibility of phenolic compounds in various pulses through in vitro digestion and colonic fermentation should be further researched, providing supportive information for the commercial utilization of these pulses as functional food ingredients and their pharmaceutical development. Furthermore, epidemiological research and clinical trials can be conducted for more specific information about the role of pulses’ bioactive constituents in human health.

Acknowledgments

We would like to thank The Future Food Hallmark Research Initiative at the University of Melbourne, Australia. We would like to thank researchers of the Dr. Hafiz Suleria group from the School of Agriculture, Faculty of Science, the University of Melbourne for their incredible support.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The work was supported by the Australian Research Council [ARC-DECRA—DE220100055]; University of Melbourne [Future Food Hallmark Research Initiative Funds (grant no. UoM-21/23)]; University of Melbourne [Collaborative Research Development Grant;UoM-21/23]

References

  • FAO. Definition and Classification Commodities, 4. Pulses and Derived Products. 1994. http://www.fao.org/es/faodef/fdef04e.htm
  • Rawal, V., and Navarro, D. K.; Eds. Introduction. In The Global Economy of Pulses, Rome, Italy: FAO, 2019; pp 1–8. DOI: 10.4060/I7108EN.
  • Calles, T. The International Year of Pulses: What are They and Why are They Important. Agric. Dev. 2016, 26, 40.
  • Faostat, F. FAOSTAT Statistical Database; FAO (Food and Agriculture Organization of the United Nations): Rome, Italy, 2016.
  • Siddiq, M.; Uebersax, M. A. Dry Beans and Pulses Production and Consumption—An Overview. In Dry Beans and Pulses Production, Processing and Nutrition, Hoboken, New Jersey, United States: Wiley, 2012; pp 1–22. DOI: 10.1002/9781118448298.
  • Merga, B.; Haji, J.; Yildiz, F. Economic Importance of Chickpea: Production, Value, and World Trade. Cogent Food Agric. 2019, 5(1), 1615718. DOI: 10.1080/23311932.2019.1615718.
  • Sozer, N.; Holopainen-Mantila, U.; Poutanen, K. Traditional and New Food Uses of Pulses. Cereal Chem. J. 2017, 94(1), 66. DOI: 10.1094/CCHEM-04-16-0082-FI.
  • Coşkuner, Y.; Karababa, E. Leblebi: A Roasted Chickpea Product as a Traditional Turkish Snack Food. Food Rev. Int. 2004, 20(3), 257. DOI: 10.1081/FRI-200029424.
  • Chatzikamari, M.; Kyriakidis, D.; Tzanetakis, N.; Biliaderis, C.; Litopoulou-Tzanetaki, E. Biochemical Changes During a Submerged Chickpea Fermentation Used as a Leavening Agent for Bread Production. Eur. Food Res. Technol. 2007, 224(6), 715. DOI: 10.1007/s00217-006-0363-4.
  • Kebede, E.; Bekeko, Z.; Tejada Moral, M. Expounding the Production and Importance of Cowpea (Vigna Unguiculata (L.) Walp.) in Ethiopia. Cogent Food & Agriculture. 2020, 6(1), 1769805. DOI: 10.1080/23311932.2020.1769805.
  • Akibode, C. S.; Maredia, M. K. Global and Regional Trends in Production, Trade and Consumption of Food Legume Crops. Department of Agricultural, Food, and Resource Economics,Staff Paper Series 2012-10. Michigan, United States: Michigan State University, 2012. DOI: 10.22004/ag.econ.136293.
  • Affrifah, N. S.; Phillips, R. D.; Saalia, F. K. Cowpeas: Nutritional Profile, Processing Methods and Products—A Review. Legume Sci. 2021, 4(3), e131. DOI: 10.1002/leg3.131.
  • Dhull, S. B.; Kidwai, M. K.; Noor, R.; Chawla, P.; Rose, P. K. A Review of Nutritional Profile and Processing of Faba Bean (Vicia Faba L.). Legume Sci. 2021, 4(3), e129. DOI: 10.1002/leg3.129.
  • Ganesan, J.; Chelladurai, V. Pigeon Pea. In Pulses, New York, United States: Springer Cham, 2020; pp 275–296. DOI:10.1007/978-3-030-41376-7_15.
  • Dilis, V.; Trichopoulou, A. Nutritional and Health Properties of Pulses. Mediterr. J. Nutr. Metab. 2009, 1(3), 149. DOI: 10.3233/s12349-008-0023-2.
  • Boye, J.; Zare, F.; Pletch, A. Pulse Proteins: Processing, Characterization, Functional Properties and Applications in Food and Feed. Food Res. Int. 2010, 43(2), 414. DOI: 10.1016/j.foodres.2009.09.003.
  • Niewinski, M. M. Advances in Celiac Disease and Gluten-Free Diet. J. Am. Diet. Assoc. 2008, 108(4), 661. DOI: 10.1016/j.jada.2008.01.011.
  • Hall, C.; Hillen, C.; Garden Robinson, J. Composition, Nutritional Value, and Health Benefits of Pulses. Cereal Chem. J. 2017, 94(1), 11. DOI: 10.1094/CCHEM-03-16-0069-FI.
  • Liu, R.; Zheng, Y.; Cai, Z.; Xu, B. Saponins and Flavonoids from Adzuki Bean (Vigna Angularis L.) Ameliorate High-Fat Diet-Induced Obesity in ICR Mice. Front. Pharmacol. 2017, 8, 687. DOI: 10.3389/fphar.2017.00687.
  • Tan, Y.; Chang, S. K.; Zhang, Y. Comparison of α-Amylase, α-Glucosidase and Lipase Inhibitory Activity of the Phenolic Substances in Two Black Legumes of Different Genera. Food Chem. 2017, 214, 259. DOI: 10.1016/j.foodchem.2016.06.100.
  • Zhang, B.; Deng, Z.; Ramdath, D. D.; Tang, Y.; Chen, P. X.; Liu, R.; Liu, Q.; Tsao, R. Phenolic Profiles of 20 Canadian Lentil Cultivars and Their Contribution to Antioxidant Activity and Inhibitory Effects on α-Glucosidase and Pancreatic Lipase. Food Chem. 2015, 172, 862. DOI: 10.1016/j.foodchem.2014.09.144.
  • Baptista, A.; Pinho, O.; Pinto, E.; Casal, S.; Mota, C.; Ferreira, I. M. P. L. V. O. Characterization of Protein and Fat Composition of Seeds from Common Beans (Phaseolus Vulgaris L.), Cowpea (Vigna Unguiculata L. Walp) and Bambara Groundnuts (Vigna Subterranea L. Verdc) from Mozambique. J. Food Meas. Charact. 2017, 11(2), 442. DOI: 10.1007/s11694-016-9412-2.
  • de Almeida Costa, G. E.; da Silva Queiroz-Monici, K.; Pissini Machado Reis, S. M.; de Oliveira, A. C. Chemical Composition, Dietary Fibre and Resistant Starch Contents of Raw and Cooked Pea, Common Bean, Chickpea and Lentil Legumes. Food Chem. 2006, 94(3), 327. DOI: 10.1016/j.foodchem.2004.11.020.
  • Lu, Z. X.; He, J. F.; Zhang, Y. C.; Bing, D. J. Composition, Physicochemical Properties of Pea Protein and Its Application in Functional Foods. Crit. Rev. Food Sci. Nutr. 2020, 60(15), 2593. DOI: 10.1080/10408398.2019.1651248.
  • Guillon, F.; Champ, M. M. J. Carbohydrate Fractions of Legumes: Uses in Human Nutrition and Potential for Health. Br. J. Nutr. 2002, 88(S3), 293. DOI: 10.1079/BJN2002720.
  • Millar, K. A.; Gallagher, E.; Burke, R.; McCarthy, S.; Barry-Ryan, C. Proximate Composition and Anti-Nutritional Factors of Fava-Bean (Vicia Faba), Green-Pea and Yellow-Pea (Pisum Sativum) Flour. J. Food Compost. Anal. 2019, 82, 103233. DOI: 10.1016/j.jfca.2019.103233.
  • Owusu-ansah, Y. J.; McCurdy, S. M. Pea Proteins: A Review of Chemistry, Technology of Production, and Utilization. Food Rev. Int. 1991, 7(1), 103. DOI: 10.1080/87559129109540903.
  • Alajaji, S. A.; El-Adawy, T. A. Nutritional Composition of Chickpea (Cicer Arietinum L.) as Affected by Microwave Cooking and Other Traditional Cooking Methods. J. Food Compost. Anal. 2006, 19(8), 806. DOI: 10.1016/j.jfca.2006.03.015.
  • Wallace, T. C.; Murray, R.; Zelman, K. M. The Nutritional Value and Health Benefits of Chickpeas and Hummus. Nutrients. 2016, 8(12), 766. DOI: 10.3390/nu8120766.
  • Rivas-Vega, M.; Goytortúa-Bores, E.; Brauer, J.; Salazar-García, M.; Cruz-Suarez, L.; Nolasco, H.; Civera-Cerecedo, R. Nutritional Value of Cowpea (Vigna Unguiculata L. Walp) Meals as Ingredients in Diets for Pacific White Shrimp (Litopenaeus Vannamei Boone). Food Chem. 2006, 97(1), 41. DOI: 10.1016/j.foodchem.2005.03.021.
  • Khalil, A. H.; Mansour, E. H. The Effect of Cooking, Autoclaving and Germination on the Nutritional Quality of Faba Beans. Food Chem. 1995, 54(2), 177. DOI: 10.1016/0308-8146(95)00024-D.
  • Joshi, M.; Timilsena, Y.; Adhikari, B. Global Production, Processing and Utilization of Lentil: A Review. J. Integr. Agric. 2017, 16(12), 2898. DOI: 10.1016/S2095-3119(17)61793-3.
  • Grela, E. R.; Kiczorowska, B.; Samolińska, W.; Matras, J.; Kiczorowski, P.; Rybiński, W.; Hanczakowska, E. Chemical Composition of Leguminous Seeds: Part I—Content of Basic Nutrients, Amino Acids, Phytochemical Compounds, and Antioxidant Activity. Eur. Food Res. Technol. 2017, 243(8), 1385. DOI: 10.1007/s00217-017-2849-7.
  • Arab, E. A.; Helmy, I.; Bareh, G. Nutritional Evaluation and Functional Properties of Chickpea (Cicer Arietinum L.) Flour and the Improvement of Spaghetti Produced from Its. J. Am. Sci. 2010, 6(10), 1055.
  • Gonçalves, A.; Goufo, P.; Barros, A.; Domínguez-Perles, R.; Trindade, H.; Rosa, E. A. S.; Ferreira, L.; Rodrigues, M. Cowpea (Vigna Unguiculata L. Walp), a Renewed Multipurpose Crop for a More Sustainable Agri-Food System: Nutritional Advantages and Constraints. J. Sci. Food Agric. 2016, 96(9), 2941. DOI: 10.1002/jsfa.7644.
  • Jezierny, D.; Mosenthin, R.; Bauer, E. The Use of Grain Legumes as a Protein Source in Pig Nutrition: A Review. Anim. Feed Sci. Technol. 2010, 157(3–4), 111. DOI: 10.1016/j.anifeedsci.2010.03.001.
  • Aguirre Santos, E. A.; Gómez Aldapa, C. A. EVALUACIÓN DE LAS CARACTERÍSTICAS FISICOQUÍMICAS EN LA ESPECIE DE FRIJOL PHASEOLUS VULGARIS DE LAS VARIEDADES; PINTO SALTILLO, BAYO VICTORIA Y NEGRO SAN LUIS. 2010. https://www.uaeh.edu.mx/investigacion/icbi/LI_FisicAlim/Carlos_Aldapa/4.pdf
  • Mejri, F.; Ben Khoud, H.; Njim, L.; Baati, T.; Selmi, S.; Martins, A.; Serralheiro, M. L. M.; Rauter, A. P.; Hosni, K. In vitro and in vivo Biological Properties of Pea Pods (Pisum Sativum L.). Food Biosci. 2019, 32, 100482. DOI: 10.1016/j.fbio.2019.100482.
  • Hanan, E.; Rudra, S. G.; Sagar, V. R.; Sharma, V. Utilization of Pea Pod Powder for Formulation of Instant Pea Soup Powder. J. Food Process. Preserv. 2020, 44(11), e14888. DOI: 10.1111/jfpp.14888.
  • Brigide, P.; Canniatt-Brazaca, S.; Silva, M. Nutritional Characteristics of Biofortified Common Beans. Food Sci. Technol (Campinas). 2014, 34(3), 493. DOI: 10.1590/1678-457x.6245.
  • Tosh, S. M.; Yada, S. Dietary Fibres in Pulse Seeds and Fractions: Characterization, Functional Attributes, and Applications. Food Res. Int. 2010, 43(2), 450.
  • Champ, M. M. J. Non-Nutrient Bioactive Substances of Pulses. Br. J. Nutr. 2002, 88(S3), 307. DOI: 10.1079/BJN2002721.
  • Dalgetty, D. D.; Baik, B. -K. Isolation and Characterization of Cotyledon Fibers from Peas, Lentils, and Chickpeas. Cereal Chem. J. 2003, 80(3), 310. DOI: 10.1094/CCHEM.2003.80.3.310.
  • Gawalko, E.; Garrett, R. G.; Warkentin, T.; Wang, N.; Richter, A. Trace Elements in Canadian Field Peas: A Grain Safety Assurance Perspective. Food Addit. Contam., Part A. 2009, 26(7), 1002. DOI: 10.1080/02652030902894389.
  • Maphosa, Y.; Jideani, V. A. Dietary Fiber Extraction for Human Nutrition—A Review. Food Rev. Int. 2016, 32(1), 98. DOI: 10.1080/87559129.2015.1057840.
  • Maheri-Sis, N.; Chamani, M.; Ali-Asghar, S.; Mirza-Aghazadeh, A.; Aghajanzadeh-Golshani, A. Nutritional Evaluation of Kabuli and Desi Type Chickpeas (Cicer Arietinum L.) for Ruminants Using in vitro Gas Production Technique. Afr. J. Biotechnol. 2008, 7(16), 2946–2951. DOI:10.5897/AJB08.501.
  • Campos-Vega, R.; Reynoso-Camacho, R.; Pedraza-Aboytes, G.; Acosta-Gallegos, J. A.; Guzman-Maldonado, S. H.; Paredes-Lopez, O.; Oomah, B. D.; Loarca-Piña, G. Chemical Composition and in vitro Polysaccharide Fermentation of Different Beans (Phaseolus Vulgaris L.). J. Food Sci. 2009, 74(7), T59. DOI: 10.1111/j.1750-3841.2009.01292.x.
  • Wang, N.; Hatcher, D. W.; Tyler, R. T.; Toews, R.; Gawalko, E. J. Effect of Cooking on the Composition of Beans (Phaseolus Vulgaris L.) and Chickpeas (Cicer Arietinum L.). Food Res. Int. 2010, 43(2), 589. DOI: 10.1016/j.foodres.2009.07.012.
  • Ovando-Martínez, M.; Osorio-Díaz, P.; Whitney, K.; Bello-Pérez, L. A.; Simsek, S. Effect of the Cooking on Physicochemical and Starch Digestibility Properties of Two Varieties of Common Bean (Phaseolus Vulgaris L.) Grown Under Different Water Regimes. Food Chem. 2011, 129(2), 358. DOI: 10.1016/j.foodchem.2011.04.084.
  • Mateos-Aparicio, I.; Redondo-Cuenca, A.; Villanueva-Suárez, M. -J.; Zapata-Revilla, M. -A.; Tenorio-Sanz, M. -D. Pea Pod, Broad Bean Pod and Okara, Potential Sources of Functional Compounds. LWT - Food Sci. Technol. 2010, 43(9), 1467. DOI: 10.1016/j.lwt.2010.05.008.
  • Antova, G. A.; Stoilova, T. D.; Ivanova, M. M. Proximate and Lipid Composition of Cowpea (Vigna Unguiculata L.) Cultivated in Bulgaria. J. Food Compost. Anal. 2014, 33(2), 146. DOI: 10.1016/j.jfca.2013.12.005.
  • Crépon, K.; Marget, P.; Peyronnet, C.; Carrouée, B.; Arese, P.; Duc, G. Nutritional Value of Faba Bean (Vicia Faba L.) Seeds for Feed and Food. Field Crops Res. 2010, 115(3), 329. DOI: 10.1016/j.fcr.2009.09.016.
  • Savage, G.; Deo, S. The Nutritional Value of Peas (Pisum Sativum). A Literature Review. Nutr. Abstr. Rev. 1989, 59, 66.
  • Sánchez-Mata, M. C.; Peñuela-Teruel, M. J.; Cámara-Hurtado, M.; Díez-Marqués, C.; Torija-Isasa, M. E. Determination of Mono-, Di-, and Oligosaccharides in Legumes by High-Performance Liquid Chromatography Using an Amino-Bonded Silica Column. J. Agric. Food Chem. 1998, 46(9), 3648. DOI: 10.1021/jf980127w.
  • Kumari, T.; Deka, S. C. Potential Health Benefits of Garden Pea Seeds and Pods: A Review. Legume Sci. 2021, 3(2), e82. DOI: 10.1002/leg3.82.
  • McCrory, M. A.; Hamaker, B. R.; Lovejoy, J. C.; Eichelsdoerfer, P. E. Pulse Consumption, Satiety, and Weight Management. Adv. Nutr. (Bethesda, Md). 2010, 1(1), 17. DOI: 10.3945/an.110.1006.
  • Martín-Cabrejas, M. A. Legumes: An Overview. In Legumes: Nutritional Quality, Processing and Potential Health Benefits, Cambridge, UK: Royal Society of Chemistry, 2019; pp 1–18. DOI:10.1039/9781788015721-00001.
  • Caprioli, G.; Giusti, F.; Ballini, R.; Sagratini, G.; Vila-Donat, P.; Vittori, S.; Fiorini, D. Lipid Nutritional Value of Legumes: Evaluation of Different Extraction Methods and Determination of Fatty Acid Composition. Food Chem. 2016, 192, 965. DOI: 10.1016/j.foodchem.2015.07.102.
  • Winham, D.; Webb, D.; Barr, A. Beans and Good Health. Nutr. Today. 2008, 43(5), 201. DOI: 10.1097/01.NT.0000303354.21347.45.
  • Mudryj Adriana, N.; Yu, N.; Aukema Harold, M. Nutritional and Health Benefits of Pulses. Appl. Physiol. Nutr. Metab. 2014, 39(11), 1197. DOI: 10.1139/apnm-2013-0557.
  • Neilson, A. P.; Goodrich, K. M.; Ferruzzi, M. G. Chapter 15 - Bioavailability and Metabolism of Bioactive Compounds from Foods. In Nutrition in the Prevention and Treatment of Disease, Fourth ed.; Coulston, A.M., Boushey, C.J., Ferruzzi, M.G. and Delahanty, L.M., Eds.; Cambridge, Massachusetts, United States: Academic Press, 2017; p. 301.
  • Espinosa-Alonso, L. G.; Lygin, A.; Widholm, J. M.; Valverde, M. E.; Paredes-Lopez, O. Polyphenols in Wild and Weedy Mexican Common Beans (Phaseolus Vulgaris L.). J. Agric. Food. Chem. 2006, 54(12), 4436. DOI: 10.1021/jf060185e.
  • Oomah, B. D.; Kotzeva, L.; Allen, M.; Bassinello, P. Z. Microwave and Micronization Treatments Affect Dehulling Characteristics and Bioactive Contents of Dry Beans (Phaseolus Vulgaris L.). J. Sci. Food Agric. 2014, 94(7), 1349. DOI: 10.1002/jsfa.6418.
  • Arribas, C.; Cabellos, B.; Cuadrado, C.; Guillamón, E.; Pedrosa, M. M. Bioactive Compounds, Antioxidant Activity, and Sensory Analysis of Rice-Based Extruded Snacks-Like Fortified with Bean and Carob Fruit Flours. Foods. 2019, 8(9), 381. DOI: 10.3390/foods8090381.
  • Pedrosa, M. M.; Varela, A.; Domínguez-Timón, F.; Tovar, C. A.; Moreno, H. M.; Borderías, A. J.; Díaz, M. T. Comparison of Bioactive Compounds Content and Techno-Functional Properties of Pea and Bean Flours and Their Protein Isolates. Plant Foods Human Nutr. 2020, 75(4), 642. DOI: 10.1007/s11130-020-00866-4.
  • Han, H.; Baik, B. -K. Antioxidant Activity and Phenolic Content of Lentils (Lens Culinaris), Chickpeas (Cicer Arietinum L.), Peas (Pisum Sativum L.) and Soybeans (Glycine Max), and Their Quantitative Changes During Processing. Int. J. Food Sci. Technol. 2008, 43(11), 1971. DOI: 10.1111/j.1365-2621.2008.01800.x.
  • Kirse, A.; KrkljFa, D. Integrated Evaluation of Cowpea (Vigna Unguiculata (L.) Walp.) and Maple Pea (Pisum Sativum Var. Arvense L.) Spreads. Agron. Res. 2015, 13, 956.
  • Cavalcante, R.; Araújo, M.; Rocha, M.; Moreira-Araújo, R. Effect of Thermal Processing on Chemical Compositions, Bioactive Compounds, and Antioxidant Activities of Cowpea Cultivars. Rev. Caatinga. 2017, 30(4), 1050. DOI: 10.1590/1983-21252017v30n426rc.
  • Ali, A. K.; Toliba, A. O.; Rady, A. H.; El-Sahy, K. M. EFFECT of GAMMA RADIATION on PHYTOCHEMICAL COMPOUNDS in FABA BEAN (Vicia Faba L.). Zagazig J. Agri. Res. 2019, 46(3), 757–767.
  • Duan, S. -C.; Kwon, S. -J.; Eom, S. -H. Effect of Thermal Processing on Color, Phenolic Compounds, and Antioxidant Activity of Faba Bean (Vicia Faba L.) Leaves and Seeds. Antioxidants. 2021, 10(8), 1207. DOI: 10.3390/antiox10081207.
  • Johnson, J. B.; Skylas, D. J.; Mani, J. S.; Xiang, J.; Walsh, K. B.; Naiker, M. Phenolic Profiles of Ten Australian Faba Bean Varieties. Molecules. 2021, 26(15), 4642. DOI: 10.3390/molecules26154642.
  • Irakli, M.; Kargiotidou, A.; Tigka, E.; Beslemes, D.; Fournomiti, M.; Pankou, C.; Stavroula, K.; Tsivelika, N.; Vlachostergios, D. N. Genotypic and Environmental Effect on the Concentration of Phytochemical Contents of Lentil (Lens Culinaris L.). Agronomy. 2021, 11(6), 1154. DOI: 10.3390/agronomy11061154.
  • Nwosu, J. N.; Ojukwu, M.; Ogueke, C. C.; Ahaotu, I.; Owuamanam, C. I. The Antinutritional Properties and Ease of Dehulling on theProximate Composition of Pigeon Pea (Cajanus Cajan) as Affected by Malting. International Journal of Life Sciences. 2013, 2(2), 60–67.
  • Sahu, M.; Verma, D.; Haris, K. Phytochemical Analysis of the Leaf, Stem and Seed Extracts of Cajanus Cajan L (Dicotyledoneae: Fabaceae). World J. Pharm. Pharm. Sci. 2014, 3, 694.
  • Sharma, S.; Singh, A.; Singh, B. Characterization of in vitro Antioxidant Activity, Bioactive Components, and Nutrient Digestibility in Pigeon Pea (Cajanus Cajan) as Influenced by Germination Time and Temperature. J. Food Biochem. 2019, 43(2), e12706. DOI: 10.1111/jfbc.12706.
  • Xu, B.; Chang, S. K. Phytochemical Profiles and Health-Promoting Effects of Cool-Season Food Legumes as Influenced by Thermal Processing. J. Agric. Food. Chem. 2009, 57(22), 10718. DOI: 10.1021/jf902594m.
  • Fares, C.; Suriano, S.; Codianni, P.; Marciello, U.; Russo, M.; Menga, V. Phytochemical Profile of Chickpea Cultivars Grown in Conventional and Organic Farms in Southern, Italy. Org. Agri. 2021, 11(4), 589. DOI: 10.1007/s13165-021-00365-z.
  • Kalogeropoulos, N.; Chiou, A.; Ioannou, M.; Karathanos, V. T.; Hassapidou, M.; Andrikopoulos, N. K. Nutritional Evaluation and Bioactive Microconstituents (Phytosterols, Tocopherols, Polyphenols, Triterpenic Acids) in Cooked Dry Legumes Usually Consumed in the Mediterranean Countries. Food Chem. 2010, 121(3), 682. DOI: 10.1016/j.foodchem.2010.01.005.
  • Aguilera, Y.; Estrella, I.; Benitez, V.; Esteban, R. M.; Martín-Cabrejas, M. A. Bioactive Phenolic Compounds and Functional Properties of Dehydrated Bean Flours. Food Res. Int. 2011, 44(3), 774. DOI: 10.1016/j.foodres.2011.01.004.
  • Aguilera, Y.; Dueñas, M.; Estrella, I.; Hernández, T.; Benitez, V.; Esteban, R. M.; Martín-Cabrejas, M. A. Phenolic Profile and Antioxidant Capacity of Chickpeas (Cicer Arietinum L.) as Affected by a Dehydration Process. Plant Foods Hum. Nutr. 2011, 66(2), 187. DOI: 10.1007/s11130-011-0230-8.
  • Aparicio, X.; Juárez-López, B. A.Polyphenolics Concentration and Antiradical Capacity of Common Bean Varieties (Phaseolus vulgaris L.) after Thermal Treatment. In Food Science and Food Biotechnology Essentials: A Contemporary Perspective, 1st ed.; Nevárez-Moorillón, G. V., Ortega-Rivas, E., Eds., Durango, Mexico: Asociación Mexicana de Ciencias de loa Alimentos,2012; pp 25.
  • Fouad, A. A.; Rehab, F. M. Effect of Germination Time on Proximate Analysis, Bioactive Compounds and Antioxidant Activity of Lentil (Lens Culinaris Medik.) Sprouts. Acta Sci. Pol. Technol. Aliment. 2015, 14(3), 233. DOI: 10.17306/J.AFS.2015.3.25.
  • de Mejía, E. G.; Guzmán-Maldonado, S. H.; Acosta-Gallegos, J. A.; Reynoso-Camacho, R.; Ramírez-Rodríguez, E.; Pons-Hernández, J. L.; González-Chavira, M. M.; Castellanos, J. Z.; Kelly, J. D. Effect of Cultivar and Growing Location on the Trypsin Inhibitors, Tannins, and Lectins of Common Beans (Phaseolus Vulgaris L.) Grown in the Semiarid Highlands of Mexico. J. Agric. Food. Chem. 2003, 51(20), 5962. DOI: 10.1021/jf030046m.
  • Kwon, S. -J.; Kim, D. -G.; Kim, J.; Kang, K. -Y.; Lee, M. -K.; Hong, M.; Kim, J. -B.; Eom, S.; Kang, S. -Y.; Ha, B. -K., et al. Phytochemical Compounds and Antioxidant Activity in the Grain of Selected Faba Bean (Vicia Faba) Genotypes. Plant Breed. Biotechnol. 2018, 6(1), 65. DOI: 10.9787/PBB.2018.6.1.65.
  • Iniestra-González, J. J.; Ibarra-Pérez, F. J.; Gallegos-Infante, J. A.; Rocha-Guzmán, N. E.; Laredo, R. F. G.; Ibarra-Pérez, F. Factores antinutricios y actividad antioxidante en variedades mejoradas de frijol común (Phaseolus vulgaris). Agrociencia. 2005, 39(1), 603. DOI: 10.1016/j.lwt.2004.11.002.
  • Ryan, E.; Galvin, K.; O’Connor, T. P.; Maguire, A. R.; O’Brien, N. M. Phytosterol, Squalene, Tocopherol Content and Fatty Acid Profile of Selected Seeds, Grains, and Legumes. Plant Foods Hum. Nutr. 2007, 62(3), 85. DOI: 10.1007/s11130-007-0046-8.
  • Ferreira, C. D.; Bubolz, V. K.; da Silva, J.; Dittgen, C. L.; Ziegler, V.; de Oliveira Raphaelli, C.; de Oliveirade Oliveira, M. Changes in the Chemical Composition and Bioactive Compounds of Chickpea (Cicer Arietinum L.) Fortified by Germination. LWT. 2019, 111, 363. DOI: 10.1016/j.lwt.2019.05.049.
  • Eashwarage, I. S.; Herath, H. M. T.; Gunathilake, K. G. T. Dietary Fibre, Resistant Starch and Elevencommonly Consumed Legumes (Mung Bean, Cowpea, Soybean and Horse Gram) in Sri Lanka. Research Journal of Chemical Sciences. 2017, 7(2), 1–7.
  • Adamidou, S.; Nengas, I.; Grigorakis, K.; Nikolopoulou, D.; Jauncey, K. Chemical Composition and Antinutritional Factors of Field Peas (Pisum Sativum), Chickpeas (Cicer Arietinum), and Faba Beans (Vicia Faba) as Affected by Extrusion Preconditioning and Drying Temperatures. Cereal Chem. J. 2011, 88(1), 80. DOI: 10.1094/CCHEM-05-10-0077.
  • Sharma, S.; Singh, A.; Sharma, U.; Kumar, R.; Yadav, N. Effect of Thermal Processing on Anti Nutritional Factors and in vitro Bioavailability of Minerals in Desi and Kabuli Cultivars of Chick Pea Grown in North India. Legume RES. Int. J. 2017, 41. DOI: 10.18805/LR-3708.
  • Suneja, Y.; Kaur, S.; Gupta, A. K.; Kaur, N. Levels of Nutritional Constituents and Antinutritional Factors in Black Gram (Vigna Mungo L. Hepper). Food Res. Int. 2011, 44(2), 621. DOI: 10.1016/j.foodres.2010.12.020.
  • Margier, M.; Georgé, S.; Hafnaoui, N.; Remond, D.; Nowicki, M.; Du Chaffaut, L.; Amiot, M. J.; Reboul, E. Nutritional Composition and Bioactive Content of Legumes: Characterization of Pulses Frequently Consumed in France and Effect of the Cooking Method. Nutrients. 2018, 10(11), 1668. DOI: 10.3390/nu10111668.
  • Shi, L.; Mu, K.; Arntfield, S. D.; Nickerson, M. T. Changes in Levels of Enzyme Inhibitors During Soaking and Cooking for Pulses Available in Canada. J. Food Sci. Technol. 2017, 54(4), 1014. DOI: 10.1007/s13197-017-2519-6.
  • Barbana, C.; Boye, J. I. In vitro Protein Digestibility and Physico-Chemical Properties of Flours and Protein Concentrates from Two Varieties of Lentil (Lens Culinaris). Food Funct. 2013, 4(2), 310. DOI: 10.1039/C2FO30204G.
  • Sharma, A.; Sehgal, S. Effect of Processing and Cooking on the Antinutritional Factors of Faba Bean (Vicia Faba). Food Chem. 1992, 43(5), 383. DOI: 10.1016/0308-8146(92)90311-O.
  • Duhan, A.; Khetarpaul, N.; Bishnoi, S. Saponin Content and Trypsin Inhibitor Activity in Processed and Cooked Pigeon Pea Cultivars. Int. J. Food Sci. Nutr. 2001, 52(1), 53. DOI: 10.1080/09637480020027200.
  • Guajardo-Flores, D.; García-Patiño, M.; Serna-Guerrero, D.; Gutiérrez-Uribe, J. A.; Serna-Saldívar, S. O. Characterization and Quantification of Saponins and Flavonoids in Sprouts, Seed Coats and Cotyledons of Germinated Black Beans. Food Chem. 2012, 134(3), 1312. DOI: 10.1016/j.foodchem.2012.03.020.
  • Lafarga, T.; Villaró, S.; Bobo, G.; Simó, J.; Aguiló-Aguayo, I. Bioaccessibility and Antioxidant Activity of Phenolic Compounds in Cooked Pulses. Int. J. Food Sci. Technol. 2019, 54(5), 1816. DOI: 10.1111/ijfs.14082.
  • Padhi, E. M.; Liu, R.; Hernández, M.; Tsao, R.; Ramdath, D. D. Total Polyphenol Content, Carotenoid, Tocopherol and Fatty Acid Composition of Commonly Consumed Canadian Pulses and Their Contribution to Antioxidant Activity. J. Funct. Foods. 2017, 38, 602. DOI: 10.1016/j.jff.2016.11.006.
  • Kalogeropoulos, N.; Chiou, A.; Ioannou, M.; Karathanos, V.; Hassapidou, M.; Andrikopoulos, N. Nutritional Evaluation and Bioactive Microconstituents (Phytosterols, Tocopherols, Polyphenols, Triterpenic Acids) in Cooked Dry Legumes Usually Consumed in the Mediterranean Countries. Food Chem. 2010, 121(3), 682. DOI: 10.1016/j.foodchem.2010.01.005.
  • Giusti, F.; Caprioli, G.; Ricciutelli, M.; Vittori, S.; Sagratini, G. Determination of Fourteen Polyphenols in Pulses by High Performance Liquid Chromatography-Diode Array Detection (HPLC-DAD) and Correlation Study with Antioxidant Activity and Colour. Food Chem. 2017, 221, 689. DOI: 10.1016/j.foodchem.2016.11.118.
  • Nachbar, M. S.; Oppenheim, J. D. Lectins in the United States Diet: A Survey of Lectins in Commonly Consumed Foods and a Review of the Literature. Am. J. Clin. Nutr. 1980, 33(11), 2338. DOI: 10.1093/ajcn/33.11.2338.
  • Singh, B.; Singh, J. P.; Shevkani, K.; Singh, N.; Kaur, A. Bioactive Constituents in Pulses and Their Health Benefits. J. Food Sci. Technol. 2017, 54(4), 858. DOI: 10.1007/s13197-016-2391-9.
  • Rochfort, S.; Panozzo, J. Phytochemicals for Health, the Role of Pulses. J. Agric. Food Chem. 2007, 55(20), 7981. DOI: 10.1021/jf071704w.
  • Singh, J. P.; Kaur, A.; Shevkani, K.; Singh, N. Influence of Jambolan (Syzygium Cumini) and Xanthan Gum Incorporation on the Physicochemical, Antioxidant and Sensory Properties of Gluten-Free Eggless Rice Muffins. Int. J. Food Sci. Technol. 2015, 50(5), 1190. DOI: 10.1111/ijfs.12764.
  • Singh, B.; Singh, J. P.; Kaur, A.; Singh, N. Bioactive Compounds in Banana and Their Associated Health Benefits–A Review. Food Chem. 2016, 206, 1. DOI: 10.1016/j.foodchem.2016.03.033.
  • Ganesan, K.; Xu, B. A Critical Review on Phytochemical Profile and Health Promoting Effects of Mung Bean (Vigna Radiata). Food Sci. Hum. Wellness. 2018, 7(1), 11. DOI: 10.1016/j.fshw.2017.11.002.
  • Smulikowska, S.; Pastuszewska, B.; Święch, E.; Ochtabińska, A.; Mieczkowska, A.; Nguyen, V. C.; Buraczewska, L. Tannin Content Affects Negatively Nutritive Value of Pea for Monogastrics. J. Anim. Feed Sci. 2001, 10(3), 511. DOI: 10.22358/jafs/68004/2001.
  • Hou, D.; Feng, Q.; Tang, J.; Shen, Q.; Zhou, S. An Update on Nutritional Profile, Phytochemical Compounds, Health Benefits, and Potential Applications in the Food Industry of Pulses Seed Coats: A Comprehensive Review. Crit. Rev. Food Sci. Nutr. 2022, 1. DOI: 10.1080/10408398.2022.2095973.
  • Dueñas, M.; Sun, B.; Hernández, T.; Estrella, I.; Spranger, M. I. Proanthocyanidin Composition in the Seed Coat of Lentils (Lens Culinaris L.). J. Agric. Food Chem. 2003, 51(27), 7999. DOI: 10.1021/jf0303215.
  • López, A.; El-Naggar, T.; Dueñas, M.; Ortega, T.; Estrella, I.; Hernández, T.; Gómez-Serranillos, M. P.; Palomino, O. M.; Carretero, M. E. Effect of Cooking and Germination on Phenolic Composition and Biological Properties of Dark Beans (Phaseolus Vulgaris L.). Food Chem. 2013, 138(1), 547. DOI: 10.1016/j.foodchem.2012.10.107.
  • Xu, B. J.; Yuan, S. H.; Chang, S. K. Comparative Analyses of Phenolic Composition, Antioxidant Capacity, and Color of Cool Season Legumes and Other Selected Food Legumes. J. Food Sci. 2007, 72(2), S167. DOI: 10.1111/j.1750-3841.2006.00261.x.
  • Ruiz-López, M. A.; Barrientos-Ramírez, L.; García-López, P. M.; Valdés-Miramontes, E. H.; Zamora-Natera, J. F.; Rodríguez-Macias, R.; Salcedo-Pérez, E.; Bañuelos-Pineda, J.; Vargas-Radillo, J. J. Nutritional and Bioactive Compounds in Mexican Lupin Beans Species: A Mini-Review. Nutrients. 2019, 11(8), 1785. DOI: 10.3390/nu11081785.
  • Wiedemann, M.; Gurrola-Díaz, C. M.; Vargas-Guerrero, B.; Wink, M.; García-López, P. M.; Düfer, M. Lupanine Improves Glucose Homeostasis by Influencing KATP Channels and Insulin Gene Expression. Molecules. 2015, 20(10), 19085. DOI: 10.3390/molecules201019085.
  • Khan, M. K.; Karnpanit, W.; Nasar-Abbas, S. M.; Huma, Z. -E.; Jayasena, V. Phytochemical Composition and Bioactivities of Lupin: A Review. Int. J. Food Sci. Technol. 2015, 50(9), 2004. DOI: 10.1111/ijfs.12796.
  • Benveniste, P. Sterol Biosynthesis. Annu. Rev. Plant Biol. 1986, 37(1), 275. DOI: 10.1146/annurev.pp.37.060186.001423.
  • Muzquiz, M.; Varela, A.; Burbano, C.; Cuadrado, C.; Guillamón, E.; Pedrosa, M. M. Bioactive Compounds in Legumes: Pronutritive and Antinutritive Actions. Implications for Nutrition and Health. Phytochem. Rev. 2012, 11(2–3), 227. DOI: 10.1007/s11101-012-9233-9.
  • Han, I. H.; Baik, B. -K. Oligosaccharide Content and Composition of Legumes and Their Reduction by Soaking, Cooking, Ultrasound, and High Hydrostatic Pressure. Cereal Chem. J. 2006, 83(4), 428. DOI: 10.1094/CC-83-0428.
  • Gobbetti, M.; DeAngelis, M.; Di Cagno, R.; Polo, A.; Rizzello, C. G. The Sourdough Fermentation is the Powerful Process to Exploit the Potential of Legumes, Pseudo-Cereals and Milling By-Products in Baking Industry. Crit. Rev. Food Sci. Nutr. 2020, 60(13), 2158. DOI: 10.1080/10408398.2019.1631753.
  • Singh, J.; Basu, P. S. Non-Nutritive Bioactive Compounds in Pulses and Their Impact on Human Health: An Overview. Food Nutr. Sci. 2012, 03(12), 1664–1672. DOI: 10.4236/fns.2012.312218.
  • Kamran, F. Enzymatic Hydrolysis of Lupin (Lupinus Angustifolius) Protein: Isolation and Characterization of Bioactive Peptides; Sydney, Australia: Western Sydney University, 2017.
  • Adamcová, A.; Laursen, K. H.; Ballin, N. Z. Lectin Activity in Commonly Consumed Plant-Based Foods: Calling for Method Harmonization and Risk Assessment. Foods. 2021, 10(11), 2796. DOI: 10.3390/foods10112796.
  • Cuadrado, C.; Hajos, G.; Burbano, C.; Pedrosa, M. M.; Ayet, G.; Muzquiz, M.; Pusztai, A.; Gelencser, E. Effect of Natural Fermentation on the Lectin of Lentils Measured by Immunological Methods. Food Agric. Immunol. 2002, 14(1), 41. DOI: 10.1080/09540100220137655.
  • Kumar, S.; Verma, A. K.; Das, M.; Jain, S. K.; Dwivedi, P. D. Clinical Complications of Kidney Bean (Phaseolus Vulgaris L.) Consumption. Nutrition. 2013, 29(6), 821. DOI: 10.1016/j.nut.2012.11.010.
  • Oomah, B. D.; Blanchard, C.; Balasubramanian, P. Phytic Acid, Phytase, Minerals, and Antioxidant Activity in Canadian Dry Bean (Phaseolus Vulgaris L.) Cultivars. J. Agric. Food Chem. 2008, 56(23), 11312. DOI: 10.1021/jf801661j.
  • West, L.; Greger, J.; White, A.; Nonnamaker, B. J. In vitro Study on Saponin-Mineral Complexation. J. Food Sci. 2006, 43(4), 1342. DOI: 10.1111/j.1365-2621.1978.tb15309.x.
  • Coda, R.; Katina, K.; Rizzello, C. G. Bran Bioprocessing for Enhanced Functional Properties. Curr. Opin. Food Sci. 2015, 1, 50. DOI: 10.1016/j.cofs.2014.11.007.
  • Luo, Y.; Gu, Z.; Han, Y.; Chen, Z. The Impact of Processing on Phytic Acid, in vitro Soluble Iron and Phy/Fe Molar Ratio of Faba Bean (Vicia Faba L.). J. Sci. Food Agric. 2009, 89(5), 861. DOI: 10.1002/jsfa.3525.
  • Alonso, R.; Aguirre, A.; Marzo, F. Effects of Extrusion and Traditional Processing Methods on Antinutrients and in vitro Digestibility of Protein and Starch in Faba and Kidney Beans. Food Chem. 2000, 68(2), 159. DOI: 10.1016/S0308-8146(99)00169-7.
  • Khattab, R. Y.; Arntfield, S. D. Nutritional Quality of Legume Seeds as Affected by Some Physical Treatments 2. Antinutritional Factors. LWT - Food Sci. Technol. 2009, 42(6), 1113. DOI: 10.1016/j.lwt.2009.02.004.
  • Shimelis, E. A.; Rakshit, S. K. Effect of Processing on Antinutrients and in vitro Protein Digestibility of Kidney Bean (Phaseolus Vulgaris L.) Varieties Grown in East Africa. Food Chem. 2007, 103(1), 161. DOI: 10.1016/j.foodchem.2006.08.005.
  • Luthria, D.; Pastor-Corrales, M. Phenolic Acid Content of Fifteen Dry Edible Bean (Phaseolus Vulgaris L) Varieties. J. Food Compost. Anal. 2006, 19(2–3), 205. DOI: 10.1016/j.jfca.2005.09.003.
  • Bishnoi, S.; Khetarpaul, N.; Yadav, R. K. Effect of Domestic Processing and Cooking Methods on Phytic Acid and Polyphenol Contents of Pea Cultivars (Pisum Sativum). Plant Foods Hum. Nutr. 1994, 45(4), 381. DOI: 10.1007/BF01088088.
  • Xu, B.; Chang, S. K. C. Effect of Soaking, Boiling, and Steaming on Total Phenolic Contentand Antioxidant Activities of Cool Season Food Legumes. Food Chem. 2008, 110(1), 1. DOI: 10.1016/j.foodchem.2008.01.045.
  • Aguilera, Y.; Dueñas, M.; Estrella, I.; Hernández, T.; Benitez, V.; Esteban, R. M.; Martín-Cabrejas, M. A. Evaluation of Phenolic Profile and Antioxidant Properties of Pardina Lentil as Affected by Industrial Dehydration. J. Agric. Food. Chem. 2010, 58(18), 10101. DOI: 10.1021/jf102222t.
  • Khandelwal, S.; Udipi, S. A.; Ghugre, P. Polyphenols and Tannins in Indian Pulses: Effect of Soaking, Germination and Pressure Cooking. Food Res. Int. 2010, 43(2), 526. DOI: 10.1016/j.foodres.2009.09.036.
  • Sharma, A.; Sehgal, S. Effect of Domestic Processing, Cooking and Germination on the Trypsin Inhibitor Activity and Tannin Content of Faba Bean (Vicia Faba). Plant Foods Human Nutr. 1992, 42(2), 127. DOI: 10.1007/BF02196465.
  • Yasmin, A.; Zeb, A.; Khalil, A. W.; Paracha, G. M. D.; Khattak, A. B. Effect of Processing on Anti-Nutritional Factors of Red Kidney Bean (Phaseolus Vulgaris) Grains. Food Bioprocess. Technol. 2008, 1(4), 415. DOI: 10.1007/s11947-008-0125-3.
  • Abd El-Hady, E. A.; Habiba, R. A. Effect of Soaking and Extrusion Conditions on Antinutrients and Protein Digestibility of Legume Seeds. LWT - Food Sci. Technol. 2003, 36(3), 285. DOI: 10.1016/S0023-6438(02)00217-7.
  • Morales, P.; Cebadera-Miranda, L.; Cámara, R. M.; Reis, F. S.; Barros, L.; Berrios, J. D. J.; Ferreira, I. C.; Cámara, M. Lentil Flour Formulations to Develop New Snack-Type Products by Extrusion Processing: Phytochemicals and Antioxidant Capacity. J. Funct. Foods. 2015, 19, 537. DOI: 10.1016/j.jff.2015.09.044.
  • Ciudad-Mulero, M.; Fernández-Ruiz, V.; Cuadrado, C.; Arribas, C.; Pedrosa, M. M.; Berrios, J. D. J.; Pan, J.; Morales, P. Novel Gluten-Free Formulations from Lentil Flours and Nutritional Yeast: Evaluation of Extrusion Effect on Phytochemicals and Non-Nutritional Factors. Food Chem. 2020, 315, 126175. DOI: 10.1016/j.foodchem.2020.126175.
  • Shohag, M.; Wei, Y.; Yang, X. Changes of Folate and Other Potential Health-Promoting Phytochemicals in Legume Seeds as Affected by Germination. J. Agric. Food Chem. 2012, 60(36), 9137. DOI: 10.1021/jf302403t.
  • López-Amorós, M.; Hernández, T.; Estrella, I. Effect of Germination on Legume Phenolic Compounds and Their Antioxidant Activity. J. Food Compost. Anal. 2006, 19(4), 277. DOI: 10.1016/j.jfca.2004.06.012.
  • Aguilera, Y.; Díaz, M. F.; Jiménez, T.; Benítez, V.; Herrera, T.; Cuadrado, C.; Martín-Pedrosa, M.; Martín-Cabrejas, M. A. Changes in Nonnutritional Factors and Antioxidant Activity During Germination of Nonconventional Legumes. J. Agric. Food Chem. 2013, 61(34), 8120. DOI: 10.1021/jf4022652.
  • Guo, X.; Li, T.; Tang, K.; Liu, R. H. Effect of Germination on Phytochemical Profiles and Antioxidant Activity of Mung Bean Sprouts (Vigna Radiata). J. Agric. Food Chem. 2012, 60(44), 11050. DOI: 10.1021/jf304443u.
  • Moktan, K.; Ojha, P. Quality Evaluation of Physical Properties, Antinutritional Factors, and Antioxidant Activity of Bread Fortified with Germinated Horse Gram (Dolichus Uniflorus) Flour. Food Sci. Nutr. 2016, 4(5), 766. DOI: 10.1002/fsn3.342.
  • Khalil, M. M. Effect of Soaking, Germination, Autoclaving and Cooking on Chemical and Biological Value of Guar Compared with Faba Bean. Food/Nahrung. 2001, 45(4), 246. DOI: 10.1002/1521-3803(20010801)45:4<246:AID-FOOD246>3.0.CO;2-F.
  • Alonso, R.; Orúe, E.; Marzo, F. Effects of Extrusion and Conventional Processing Methods on Protein and Antinutritional Factor Contents in Pea Seeds. Food Chem. 1998, 63(4), 505. DOI: 10.1016/S0308-8146(98)00037-5.
  • Wilson, K. A.; Papastoitsis, G.; Hartl, P.; Tan-Wilson, A. L. Survey of the Proteolytic Activities Degrading the Kunitz Trypsin Inhibitor and Glycinin in Germinating Soybeans (Glycine Max) 1. Plant Physiol. 1988, 88(2), 355. DOI: 10.1104/pp.88.2.355.
  • Guajardo-Flores, D.; Serna-Saldívar, S. O.; Gutiérrez-Uribe, J. A. Evaluation of the Antioxidant and Antiproliferative Activities of Extracted Saponins and Flavonols from Germinated Black Beans (Phaseolus Vulgaris L.). Food Chem. 2013, 141(2), 1497. DOI: 10.1016/j.foodchem.2013.04.010.
  • Boateng, J.; Verghese, M.; Walker, L.; Ogutu, S. Effect of Processing on Antioxidant Contents in Selected Dry Beans (Phaseolus Spp. L.). LWT Food Sci. Technol. 2008, 41(9), 1541. DOI: 10.1016/j.lwt.2007.11.025.
  • Habtie, T.; Admassu, S.; Asres, K. Effects of Processing Methods on Some Phytochemicals Present in the Seeds of Lupinus Albus L. Grown in Ethiopia. Ethiopian Pharm. J. 2009, 27(2). DOI: 10.4314/epj.v27i2.58269.
  • Siah, S.; Konczak, I.; Wood, J. A.; Agboola, S.; Blanchard, C. L. Effects of Roasting on Phenolic Composition and in vitro Antioxidant Capacity of Australian Grown Faba Beans (Vicia Faba L.). Plant Foods Human Nutr. 2014, 69(1), 85. DOI: 10.1007/s11130-013-0400-y.
  • Seena, S.; Sridhar, K. R.; Arun, A. B.; Young, C. -C. Effect of Roasting and Pressure-Cooking on Nutritional and Protein Quality of Seeds of Mangrove Legume Canavalia Cathartica from Southwest Coast of India. J. Food Compost. Anal. 2006, 19(4), 284. DOI: 10.1016/j.jfca.2005.05.004.
  • Aremu, M.; Ibrahim, H.; Ekanem, B. Effect of Processing on In–Vitro Protein Digestibility and Anti–Nutritional Properties of Three Underutilized Legumes Grown in Nigeria. British Biotechnol. J. 2016, 14(1), 1–10. DOI: 10.9734/BBJ/2016/22581.
  • Bolek, S. Effects of Roasting on Bioavailability and Bioactivities of Vigna Angularis and Potential of Coffee-Like Beverage. J. Food Sci. 2022, 87(3), 911. DOI: 10.1111/1750-3841.16048.
  • Kamalasundari, S.; Babu, R.; Umamaheswari, T. Effect of Domestic Processing Methods on Anti-Nutritional Factors and Its Impact on the Bio-Availability Proteins and Starch in Commonly Consumed Whole Legumes. Asian J. Dairy Food Res. 2019, 38(1). DOI: 10.18805/ajdfr.DR-1410.
  • Jiratanan, T.; Liu, R. H. Antioxidant Activity of Processed Table Beets (Beta Vulgaris Var, Conditiva) and Green Beans (Phaseolus Vulgaris L.). J. Agric. Food Chem. 2004, 52(9), 2659. DOI: 10.1021/jf034861d.
  • Preti, R.; Rapa, M.; Vinci, G. Effect of Steaming and Boiling on the Antioxidant Properties and Biogenic Amines Content in Green Bean (Phaseolus Vulgaris) Varieties of Different Colours. J. Food Qual. 2017, 2017, 2017. DOI: 10.1155/2017/5329070.
  • Ojo, M. A.; Ade-Omowaye, B.; Ngoddy, P. Influence of Soaking and Hydrothermal Techniques on Antinutritional Components and in vitro Multienzymes Protein Digestibility of Vigna Racemosa–An Underutilised Hard-To-Cook Legume. Ann Food Sci Technol. 2017, 18, 385.
  • Bolade, M. K. Individualistic Impact of Unit Operations of Production, at Household Level, on Some Antinutritional Factors in Selected Cowpea-Based Food Products. Food Sci. Nutr. 2016, 4(3), 441. DOI: 10.1002/fsn3.306.
  • Qin, G.; Ter Elst, E. R.; Bosch, M. W.; van der Poel, A. F. B. Thermal Processing of Whole Soya Beans: Studies on the Inactivation of Antinutritional Factors and Effects on Ileal Digestibility in Piglets. Anim. Feed Sci. Technol. 1996, 57(4), 313. DOI: 10.1016/0377-8401(95)00863-2.
  • Batista, K. A.; Pereira, W. J.; Moreira, B. R.; Silva, C.; Fernandes, K. F. Effect of Autoclaving on the Nutritional Quality of Hard-To-Cook Common Beans (Phaseolus Vulgaris). Int. J. Environ. Agric. Biotechnol. 2020, 5(1), 22. DOI: 10.22161/ijeab.51.4.
  • Ertaş, N.; Türker, S. Bulgur Processes Increase Nutrition Value: Possible Role in in-Vitro Protein Digestability, Phytic Acid, Trypsin Inhibitor Activity and Mineral Bioavailability. J. Food Sci. Technol. 2014, 51(7), 1401. DOI: 10.1007/s13197-012-0638-7.
  • Rehman, Z. -U.; Shah, W. Thermal Heat Processing Effects on Antinutrients, Protein and Starch Digestibility of Food Legumes. Food Chem. 2005, 91(2), 327. DOI: 10.1016/j.foodchem.2004.06.019.
  • Piecyk, M.; Wołosiak, R.; Drużynska, B.; Worobiej, E. Chemical Composition and Starch Digestibility in Flours from Polish Processed Legume Seeds. Food Chem. 2012, 135(3), 1057. DOI: 10.1016/j.foodchem.2012.05.051.
  • Siah, S.; Wood, J. A.; Agboola, S.; Konczak, I.; Blanchard, C. L. Effects of Soaking, Boiling and Autoclaving on the Phenolic Contents and Antioxidant Activities of Faba Beans (Vicia Faba L.) Differing in Seed Coat Colours. Food Chem. 2014, 142, 461. DOI: 10.1016/j.foodchem.2013.07.068.
  • Jood, S.; Chauhan, B. M.; Kapoor, A. C. Polyphenols of Chickpea and Blackgram as Affected by Domestic Processing and Cooking Methods. J. Sci. Food Agric. 1987, 39(2), 145. DOI: 10.1002/jsfa.2740390207.
  • Kalpanadevi, V.; Mohan, V. R. Effect of Processing on Antinutrients and in vitro Protein Digestibility of the Underutilized Legume, Vigna Unguiculata (L.) Walp Subsp. Unguiculata. LWT - Food Sci. Technol. 2013, 51(2), 455. DOI: 10.1016/j.lwt.2012.09.030.
  • Tungmunnithum, D.; Drouet, S.; Lorenzo, J. M.; Hano, C. Effect of Traditional Cooking and in vitro Gastrointestinal Digestion of the Ten Most Consumed Beans from the Fabaceae Family in Thailand on Their Phytochemicals, Antioxidant and Anti-Diabetic Potentials. Plants. 2021, 11(1), 67. DOI: 10.3390/plants11010067.
  • Nyau, V.; Prakash, S.; Rodrigues, J.; Farrant, J. Domestic Cooking Effects of Bambara Groundnuts and Common Beans in the Antioxidant Properties and Polyphenol Profiles. J Food Res. 2017, 6(2), 10.5539. DOI: 10.5539/jfr.v6n2p24.
  • Palermo, M.; Pellegrini, N.; Fogliano, V. The Effect of Cooking on the Phytochemical Content of Vegetables. J. Sci. Food Agric. 2014, 94(6), 1057. DOI: 10.1002/jsfa.6478.
  • Carlini, C. R.; Udedibie, A. B. Comparative Effects of Processing Methods on Hemagglutinating and Antitryptic Activities of Canavalia Ensiformis and Canavalia Braziliensis Seeds. J. Agric. Food Chem. 1997, 45(11), 4372.
  • Trugo, L. C.; Ramos, L. A.; Trugo, N. M. F.; Souza, M. C. P. Oligosaccharide Composition and Trypsin Inhibitor Activity of P. Vulgaris and the Effect of Germination on the α-Galactoside Composition and Fermentation in the Human Colon. Food Chem. 1990, 36(1), 53. DOI: 10.1016/0308-8146(90)90007-Q.
  • Egbe, I. A.; Akinyele, I. O. Effect of Cooking on the Antinutritional Factors of Lima Beans (Phaseolus Lunatus). Food Chem. 1990, 35(2), 81. DOI: 10.1016/0308-8146(90)90022-V.
  • Bressani, R.; Elías, L. G., Nutritional Role of Polyphenols in Beans. In Polyphenols in cereals and legumes: proceedings of a symposium, IDRC, Ottawa, ON, CA: 1979.
  • Nicolás-García, M.; Perucini-Avendaño, M.; Jiménez-Martínez, C.; Perea-Flores, M. D. J.; Gómez-Patiño, M. B.; Arrieta-Báez, D.; Dávila-Ortiz, G. Bean Phenolic Compound Changes During Processing: Chemical Interactions and Identification. J. Food Sci. 2021, 86(3), 643. DOI: 10.1111/1750-3841.15632.
  • Avanza, M.; Acevedo, B.; Chaves, M.; Añón, M. Nutritional and Anti-Nutritional Components of Four Cowpea Varieties Under Thermal Treatments: Principal Component Analysis. LWT - Food Sci. Technol. 2013, 51(1), 148. DOI: 10.1016/j.lwt.2012.09.010.
  • Aguilera, Y.; Dueñas, M.; Estrella, I.; Hernández, T.; Benitez, V.; Esteban, R. M.; Martín-Cabrejas, M. A. Phenolic Profile and Antioxidant Capacity of Chickpeas (Cicer Arietinum L.) as Affected by a Dehydration Process. Plant Foods for Human Nutrition. Plant Foods Human Nutr. 2011, 66(2), 187. DOI: 10.1007/s11130-011-0230-8.
  • Luthria, D. L.; Pastor-Corrales, M. A. Phenolic Acids Content of Fifteen Dry Edible Bean (Phaseolus Vulgaris L.) Varieties. J. Food Compost. Anal. 2006, 19(2–3), 205. DOI: 10.1016/j.jfca.2005.09.003.
  • Fredrikson, M.; Alminger, M. L.; Carlsson, N. -G.; Sandberg, A. -S. Phytate Content and Phytate Degradation by Endogenous Phytase in Pea (Pisum Sativum). J. Sci. Food Agric. 2001, 81(12), 1139. DOI: 10.1002/jsfa.918.
  • Francis, G.; Makkar, H. P. S.; Becker, K. Antinutritional Factors Present in Plant-Derived Alternate Fish Feed Ingredients and Their Effects in Fish. Aquaculture. 2001, 199(3–4), 197. DOI: 10.1016/S0044-8486(01)00526-9.
  • Kumar, Y.; Basu, S.; Goswami, D.; Devi, M.; Shivhare, U. S.; Vishwakarma, R. K. Anti-Nutritional Compounds in Pulses: Implications and Alleviation Methods. Legume Sci. 2022, 4(2), e111. DOI: 10.1002/leg3.111.
  • Egounlety, M.; Aworh, O. C. Effect of Soaking, Dehulling, Cooking and Fermentation with Rhizopus Oligosporus on the Oligosaccharides, Trypsin Inhibitor, Phytic Acid and Tannins of Soybean (Glycine Max Merr.), Cowpea (Vigna Unguiculata L. Walp) and Groundbean (Macrotyloma Geocarpa Harms). J. Food Eng. 2003, 56(2), 249.
  • Berrios, J. D. J. Extrusion Processing of Main Commercial Legume Pulses. Adv. Food Extrusion Technol . 2016, 1, 209.
  • Tas, A. A.; Shah, A. U. The Replacement of Cereals by Legumes in Extruded Snack Foods: Science, Technology and Challenges. Trends Food Sci. Technol. 2021, 116, 701. DOI: 10.1016/j.tifs.2021.08.016.
  • Patil, S. S.; Brennan, M. A.; Mason, S. L.; Brennan, C. S. The Effects of Fortification of Legumes and Extrusion on the Protein Digestibility of Wheat Based Snack. Foods. 2016, 5(4), 26. DOI: 10.3390/foods5020026.
  • Alam, M. S.; Kaur, J.; Khaira, H.; Gupta, K. Extrusion and Extruded Products: Changes in Quality Attributes as Affected by Extrusion Process Parameters: A Review. Crit. Rev. Food Sci. Nutr. 2016, 56(3), 445. DOI: 10.1080/10408398.2013.779568.
  • Rathod, R. P.; Annapure, U. S. Effect of Extrusion Process on Antinutritional Factors and Protein and Starch Digestibility of Lentil Splits. LWT - Food Sci. Technol. 2016, 66, 114. DOI: 10.1016/j.lwt.2015.10.028.
  • Pasqualone, A.; Costantini, M.; Coldea, T. E.; Summo, C. Use of Legumes in Extrusion Cooking: A Review. Foods. 2020, 9(7), 958. DOI: 10.3390/foods9070958.
  • Kamau, E. H.; Nkhata, S. G.; Ayua, E. O. Extrusion and Nixtamalization Conditions Influence the Magnitude of Change in the Nutrients and Bioactive Components of Cereals and Legumes. Food Sci. Nutr. 2020, 8(4), 1753. DOI: 10.1002/fsn3.1473.
  • Marzo, F.; Alonso, R.; Urdaneta, E.; Arricibita, F. J.; Ibáñez, F. Nutritional Quality of Extruded Kidney Bean (Phaseolus Vulgaris L. Var. Pinto) and Its Effects on Growth and Skeletal Muscle Nitrogen Fractions in Rats. J. Anim. Sci. 2002, 80(4), 875. DOI: 10.2527/2002.804875x.
  • Karaaslan, M.; Yilmaz, F. M.; Cesur, Ö.; Vardin, H.; Ikinci, A.; Dalgiç, A. C. Drying Kinetics and Thermal Degradation of Phenolic Compounds and Anthocyanins in Pomegranate Arils Dried Under Vacuum Conditions. Int. J. Food Sci. Technol. 2014, 49(2), 595. DOI: 10.1111/ijfs.12342.
  • Arribas, C.; Cabellos, B.; Cuadrado, C.; Guillamón, E.; Pedrosa, M. M. The Effect of Extrusion on the Bioactive Compounds and Antioxidant Capacity of Novel Gluten-Free Expanded Products Based on Carob Fruit, Pea and Rice Blends. Innovative Food Sci. Emerg. Technol. 2019, 52, 100. DOI: 10.1016/j.ifset.2018.12.003.
  • Espinoza-Moreno, R. J.; Reyes-Moreno, C.; Milán-Carrillo, J.; López-Valenzuela, J. A.; Paredes-López, O.; Gutiérrez-Dorado, R. Healthy Ready-To-Eat Expanded Snack with High Nutritional and Antioxidant Value Produced from Whole Amarantin Transgenic Maize and Black Common Bean. Plant Foods Human Nutr. 2016, 71(2), 218. DOI: 10.1007/s11130-016-0551-8.
  • Zieliński, H.; Michalska, A.; Piskuła, M. K.; Kozłowska, H. Antioxidants in Thermally Treated Buckwheat Groats. Mol. Nutr. Food Res. 2006, 50(9), 824. DOI: 10.1002/mnfr.200500258.
  • Pedrosa, M. M.; Cuadrado, C.; Burbano, C.; Allaf, K.; Haddad, J.; Gelencsér, E.; Takács, K.; Guillamón, E.; Muzquiz, M. Effect of Instant Controlled Pressure Drop on the Oligosaccharides, Inositol Phosphates, Trypsin Inhibitors and Lectins Contents of Different Legumes. Food Chem. 2012, 131(3), 862. DOI: 10.1016/j.foodchem.2011.09.061.
  • Lorenz, K.; Jansen, G.; Harper, J. Nutrient Stability of Full-Fat Soy Flour and Corn-Soy Blends Produced by Low-Cost Extrusion. Cereal Foods World. 1980, 25(4), 161–162.
  • Nikmaram, N.; Leong, S. Y.; Koubaa, M.; Zhu, Z.; Barba, F. J.; Greiner, R.; Oey, I.; Roohinejad, S. Effect of Extrusion on the Anti-Nutritional Factors of Food Products: An Overview. Food Control. 2017, 79, 62. DOI: 10.1016/j.foodcont.2017.03.027.
  • Laxmi, G.; Chaturvedi, N.; Richa, S. The Impact of Malting on Nutritional Composition of Foxtail Millet, Wheat and Chickpea. J. Nutr. Food Sci. 2015, 5(5), 1. DOI: 10.4172/2155-9600.1000407.
  • Oghbaei, M.; Prakash, J.; Yildiz, F. Effect of Primary Processing of Cereals and Legumes on Its Nutritional Quality: A Comprehensive Review. Cogent Food Agric. 2016, 2(1), 1136015. DOI: 10.1080/23311932.2015.1136015.
  • Samtiya, M.; Aluko, R. E.; Dhewa, T. Plant Food Anti-Nutritional Factors and Their Reduction Strategies: An Overview. Food Prod. Process. Nutr. 2020, 2(1), 6. DOI: 10.1186/s43014-020-0020-5.
  • Rehman, Z. -U.; Shah, W. H. Thermal Heat Processing Effects on Antinutrients, Protein and Starch Digestibility of Food Legumes. Food Chem. 2005, 91(2), 327. DOI: 10.1016/j.foodchem.2004.06.019.
  • López-Martínez, L. X.; Leyva-López, N.; Gutiérrez-Grijalva, E. P.; Heredia, J. B. Effect of Cooking and Germination on Bioactive Compounds in Pulses and Their Health Benefits. J. Funct. Foods. 2017, 38, 624. DOI: 10.1016/j.jff.2017.03.002.
  • Mamilla, R. K.; Mishra, V. K. Effect of Germination on Antioxidant and ACE Inhibitory Activities of Legumes. LWT. 2017, 75, 51. DOI: 10.1016/j.lwt.2016.08.036.
  • Patterson, C. A.; Curran, J.; Der, T. Effect of Processing on Antinutrient Compounds in Pulses. Cereal Chem. J. 2017, 94(1), 2. DOI: 10.1094/CCHEM-05-16-0144-FI.
  • Mubarak, A. E. Nutritional Composition and Antinutritional Factors of Mung Bean Seeds (Phaseolus Aureus) as Affected by Some Home Traditional Processes. Food Chem. 2005, 89(4), 489. DOI: 10.1016/j.foodchem.2004.01.007.
  • Bueno-Borges, L. B.; Sartim, M. A.; Gil, C. C.; Sampaio, S. V.; Rodrigues, P. H. V.; Regitano-d’Arce, M. A. B. Sacha Inchi Seeds from Sub-Tropical Cultivation: Effects of Roasting on Antinutrients, Antioxidant Capacity and Oxidative Stability. J. Food Sci. Technol. 2018, 55(10), 4159. DOI: 10.1007/s13197-018-3345-1.
  • Oluseyi, E. O.; Temitayo, O. M. Chemical and Functional Properties of Fermented, Roasted and Germinated Tamarind (Tamarindus Indica) Seed Flours. Nutr. Food Sci. 2015, 45(1), 97–111. DOI: 10.1108/NFS-11-2013-0131.
  • Udensi, E.; Ekwu, F.; Isinguzo, J. Antinutrient Factors of Vegetable Cowpea (Sesquipedalis) Seeds During Thermal Processing. Pak. J. Nutr. 2007, 6(2), 194. DOI: 10.3923/pjn.2007.194.197.
  • Cardador-Martínez, A.; Maya-Ocaña, K.; Ortiz-Moreno, A.; Herrera-Cabrera, B. E.; Dávila-Ortiz, G.; Múzquiz, M.; Martín-Pedrosa, M.; Burbano, C.; Cuadrado, C.; Jiménez-Martínez, C. Effect of Roasting and Boiling on the Content of Vicine, Convicine and L-3,4-Dihydroxyphenylalanine in Vicia Faba L. J. Food Qual. 2012, 35(6), 419. DOI: 10.1111/jfq.12006.
  • Pedrosa, M. M.; Guillamón, E.; Arribas, C. Autoclaved and Extruded Legumes as a Source of Bioactive Phytochemicals: A Review. Foods. 2021, 10(2), 379. DOI: 10.3390/foods10020379.
  • Ertop, M. H.; Bektaş, M. Enhancement of Bioavailable Micronutrients and Reduction of Antinutrients in Foods with Some Processes. Food And Health. 2018, 4(3), 159. DOI: 10.3153/FH18016.
  • Abd Allah, M. A.; Foda, Y. H.; Abu Salem, F. M.; Abd Allah, Z. S. Treatments for Reducing Total Vicine in Egyptian Faba Bean (Giza 2 Variety). Plant Foods Human Nutr. 1988, 38(3), 201. DOI: 10.1007/BF01092859.
  • Zhang, B.; Deng, Z.; Tang, Y.; Chen, P. X.; Liu, R.; Ramdath, D. D.; Liu, Q.; Hernandez, M.; Tsao, R. Effect of Domestic Cooking on Carotenoids, Tocopherols, Fatty Acids, Phenolics, and Antioxidant Activities of Lentils (Lens Culinaris). J. Agric. Food Chem. 2014, 62(52), 12585. DOI: 10.1021/jf504181r.
  • Granito, M.; Paolini, M.; Pérez, S. Polyphenols and Antioxidant Capacity of Phaseolus Vulgaris Stored Under Extreme Conditions and Processed. LWT - Food Sci. Technol. 2008, 41(6), 994. DOI: 10.1016/j.lwt.2007.07.014.
  • Nakitto, A. M.; Muyonga, J. H.; Nakimbugwe, D. Effects of Combined Traditional Processing Methods on the Nutritional Quality of Beans. Food Sci. Nutr. 2015, 3(3), 233. DOI: 10.1002/fsn3.209.
  • Amarowicz, R.; Pegg, R. B. Legumes as a Source of Natural Antioxidants. Eur. J. Lipid Sci. Technol. 2008, 110(10), 865. DOI: 10.1002/ejlt.200800114.
  • Bento, J. A. C.; Lanna, A. C.; Bassinello, P. Z.; Oomah, B. D.; Pimenta, M. E. B.; Carvalho, R. N.; Moreira, A. S. Aging Indicators for Stored Carioca Beans. Food Res. Int. 2020, 134, 109249. DOI: 10.1016/j.foodres.2020.109249.
  • Cilla, A.; Bosch, L.; Barberá, R.; Alegría, A. Effect of Processing on the Bioaccessibility of Bioactive Compounds – a Review Focusing on Carotenoids, Minerals, Ascorbic Acid, Tocopherols and Polyphenols. J. Food Compost. Anal. 2018, 68, 3. DOI: 10.1016/j.jfca.2017.01.009.
  • Lafarga, T.; Villaró, S.; Bobo, G.; Simó, J.; Aguiló-aguayo, I. Bioaccessibility and Antioxidant Activity of Phenolic Compounds in Cooked Pulses. Int. J. Food Sci. Technol. 2019, 54(5), 1816. DOI: 10.1111/ijfs.14082.
  • Parada, J.; Aguilera, J. M. Food Microstructure Affects the Bioavailability of Several Nutrients. J. Food Sci. 2007, 72(2), R21. DOI: 10.1111/j.1750-3841.2007.00274.x.
  • Scalbert, A.; Williamson, G. Dietary Intake and Bioavailability of Polyphenols. J. Nutr. 2000, 130(8), 2073S. DOI: 10.1093/jn/130.8.2073S.
  • Wang, T.; He, F.; Chen, G. Improving Bioaccessibility and Bioavailability of Phenolic Compounds in Cereal Grains Through Processing Technologies: A Concise Review. J. Funct. Foods. 2014, 7, 101. DOI: 10.1016/j.jff.2014.01.033.
  • Zhang, B.; Deng, Z.; Tang, Y.; Chen, P. X.; Liu, R.; Ramdath, D. D.; Liu, Q.; Hernandez, M.; Tsao, R. Bioaccessibility, in vitro Antioxidant and Anti-Inflammatory Activities of Phenolics in Cooked Green Lentil (Lens Culinaris). J. Funct. Foods. 2017, 32, 248. DOI: 10.1016/j.jff.2017.03.004.
  • Sancho, R. A. S.; Pavan, V.; Pastore, G. M. Effect of in vitro Digestion on Bioactive Compounds and Antioxidant Activity of Common Bean Seed Coats. Food Res. Int. 2015, 76, 74. DOI: 10.1016/j.foodres.2014.11.042.
  • Chen, P. X.; Dupuis, J. H.; Marcone, M. F.; Pauls, P. K.; Liu, R.; Liu, Q.; Tang, Y.; Zhang, B.; Tsao, R. Physicochemical Properties and in vitro Digestibility of Cooked Regular and Nondarkening Cranberry Beans (Phaseolus Vulgaris L.) and Their Effects on Bioaccessibility, Phenolic Composition, and Antioxidant Activity. J. Agric. Food Chem. 2015, 63(48), 10448. DOI: 10.1021/acs.jafc.5b04005.
  • Hithamani, G.; Srinivasan, K. Bioaccessibility of Polyphenols from Wheat (Triticum Aestivum), Sorghum (Sorghum Bicolor), Green Gram (Vigna Radiata), and Chickpea (Cicer Arietinum) as Influenced by Domestic Food Processing. J. Agric. Food Chem. 2014, 62(46), 11170. DOI: 10.1021/jf503450u.
  • Akillioglu, H. G.; Karakaya, S. Changes in Total Phenols, Total Flavonoids, and Antioxidant Activities of Common Beans and Pinto Beans After Soaking, Cooking, and in vitro Digestion Process. Food Sci. Biotechnol. 2010, 19(3), 633. DOI: 10.1007/s10068-010-0089-8.
  • McInerney, J. K.; Seccafien, C. A.; Stewart, C. M.; Bird, A. R. Effects of High Pressure Processing on Antioxidant Activity, and Total Carotenoid Content and Availability, in Vegetables. Innovative Food Sci. Emerg. Technol. 2007, 8(4), 543. DOI: 10.1016/j.ifset.2007.04.005.
  • Oghbaei, M.; Prakash, J. Effects of Processing and Digestive Enzymes on Retention, Bioaccessibility and Antioxidant Activity of Bioactive Components in Food Mixes Based on Legumes and Green Leaves. Food Biosci. 2013, 4, 21. DOI: 10.1016/j.fbio.2013.08.001.
  • Ofuya, Z.; Akhidue, V. The Role of Pulses in Human Nutrition: A Review. J. Appl. Sci. Environ. Manage. 2005, 9(3), 99. DOI: 10.4314/jasem.v9i3.17361.
  • Kamboj, R.; Nanda, V. Proximate Composition, Nutritional Profile and Health Benefits of Legumes-A Review. Legume Res. 2018, 41, 3. DOI: 10.18805/LR-3748.
  • Yeap, S. K.; Mohd Ali, N.; Mohd Yusof, H.; Alitheen, N. B.; Beh, B. K.; Ho, W. Y.; Koh, S. P.; Long, K. Antihyperglycemic Effects of Fermented and Nonfermented Mung Bean Extracts on Alloxan-Induced-Diabetic Mice. J. Biomed. Biotechnol. 2012, 2012, 2012. DOI: 10.1155/2012/285430.
  • Moreno-Valdespino, C. A.; Luna-Vital, D.; Camacho-Ruiz, R. M.; Mojica, L. Bioactive Proteins and Phytochemicals from Legumes: Mechanisms of Action Preventing Obesity and Type-2 Diabetes. Food Res. Int. 2020, 130, 108905. DOI: 10.1016/j.foodres.2019.108905.
  • Bai, Y.; Chang, J.; Xu, Y.; Cheng, D.; Liu, H.; Zhao, Y.; Yu, Z. Antioxidant and Myocardial Preservation Activities of Natural Phytochemicals from Mung Bean (Vigna Radiata L.) Seeds. J. Agric. Food Chem. 2016, 64(22), 4648. DOI: 10.1021/acs.jafc.6b01538.
  • Duh, P. -D.; Du, P. -C.; Yen, G. -C. Action of Methanolic Extract of Mung Bean Hulls as Inhibitors of Lipid Peroxidation and Non-Lipid Oxidative Damage. Food Chem. Toxicol. 1999, 37(11), 1055. DOI: 10.1016/S0278-6915(99)00096-4.
  • Pingfan, W. Research Progress on the Biological Activities and Functions of Mung Beans. J. Chin. Inst. Food Sci. Technol. 2004, 9(13), 3286–3291.
  • Thembo, K.; Vismer, H.; Nyazema, N.; Gelderblom, W.; Katerere, D. Antifungal Activity of Four Weedy Plant Extracts Against Selected Mycotoxigenic Fungi. J. Appl. Microbiol. 2010, 109(4), 1479. DOI: 10.1111/j.1365-2672.2010.04776.x.
  • Manikandaselvi, S.; David, R.; Aravind, S.; Ravikumar, R.; Thinagarbabu, R.; Nandhini, S. Anti-Anemic Activity of Sprouts of Vigna Radiata L. in Male Albino Rats. Int. J. Pharm. Pharm. Sci. 2015, 7(11), 263–267.
  • Venkateshwarlu, E.; Reddy, K. P.; Dilip, D. Potential of Vigna Radiata (L.) Sprouts in the Management of Inflammation and Arthritis in Rats: Possible Biochemical Alterations. Indian J. Exp. Biol. 2016, 54(1), 37–43.
  • Luo, J.; Cai, W.; Wu, T.; Xu, B. Phytochemical Distribution in Hull and Cotyledon of Adzuki Bean (Vigna Angularis L.) and Mung Bean (Vigna Radiate L.), and Their Contribution to Antioxidant, Anti-Inflammatory and Anti-Diabetic Activities. Food Chem. 2016, 201, 350. DOI: 10.1016/j.foodchem.2016.01.101.
  • Xu, B.; Chang, S. K. C. Phenolic Substance Characterization and Chemical and Cell-Based Antioxidant Activities of 11 Lentils Grown in the Northern United States. J. Agric. Food Chem. 2010, 58(3), 1509. DOI: 10.1021/jf903532y.
  • Singh, B.; Singh, J. P.; Kaur, A.; Singh, N. Phenolic Composition and Antioxidant Potential of Grain Legume Seeds: A Review. Food Res. Int. 2017, 101, 1. DOI: 10.1016/j.foodres.2017.09.026.
  • Kumar, V.; Sinha, A. K.; Makkar, H. P. S.; Becker, K. Dietary Roles of Phytate and Phytase in Human Nutrition: A Review. Food Chem. 2010, 120(4), 945. DOI: 10.1016/j.foodchem.2009.11.052.