Next Article in Journal
Effect of Furnace Structure on Burden Distribution and Gas Flow in Sinter Vertical Cooling Furnace
Next Article in Special Issue
Loss of Vitamin E While Baking and Heating French Fries
Previous Article in Journal
Bending Strength Evaluation of Three Dimensional Double Rachel Geosynthetic Cementitious Composite Mat (GCCM) under Environmental Conditions
Previous Article in Special Issue
Development and Quality Attributes of Paste Sausage Supplemented with Common Squid (Todarodes pacificus) Tailored for the Elderly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oat and Oat Processed Products—Technology, Composition, Nutritional Value, and Health

by
Danuta Leszczyńska
1,*,
Anna Wirkijowska
2,
Alan Gasiński
3,
Dominika Średnicka-Tober
4,
Joanna Trafiałek
5 and
Renata Kazimierczak
4,*
1
Department of Cereal Crop Production, Institute of Soil Science and Plant Cultivation—State Research Institute, Czartoryskich 8, 24-100 Puławy, Poland
2
Division of Engineering and Cereals Technology, Department of Plant Food Technology and Gastronomy, University of Life Sciences in Lublin, Skromna 8, 20-704 Lublin, Poland
3
Department of Fermentation and Cereals Technology, Wrocław University of Environmental and Life Science, Chełmońskiego 37, 51-630 Wroclaw, Poland
4
Department of Functional and Organic Food, Institute of Human Nutrition Sciences, Nowoursynowska 159c St., 02-776 Warsaw, Poland
5
Department of Food Gastronomy and Food Hygiene, Institute of Human Nutrition Sciences, Nowoursynowska 159c St., 02-776 Warsaw, Poland
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2023, 13(20), 11267; https://doi.org/10.3390/app132011267
Submission received: 8 August 2023 / Revised: 22 September 2023 / Accepted: 11 October 2023 / Published: 13 October 2023

Abstract

:
Oat has been known in food technology and human nutrition for a very long time. Its rich chemical composition and high nutritional value make it of interest to scientists in the field of food processing technology as well as nutritionists. Low-processed, whole-grain oat products rich in biologically active substances with well-proven preventive and therapeutic effects include not only dehulled oat grains and groats but also a wide range of breakfast cereals. These products fit into the definition of functional foods and are considered excellent prebiotics. The continuous development of cereal processing technologies can improve existing cereal products and help to create new ones; however, it also increases the possibility of raw material over-processing, thus decreasing its functional properties. Therefore, monitoring technological progress and the quality of the products obtained is of great value and interest for nutritionists and consumers. The work presented here aims at systematizing existing knowledge on oat products, their impact on human health, and progress in oat processing technology. It also brings insight into various new avenues for the utilization of oat products in food technology.

1. Introduction

Analysis of nutritional, chemical, and physical properties of oat grain, with emphasis on β-glucans, has led to its appreciation in human nutrition, mainly due to its health-promoting properties, including cholesterol-lowering, blood glucose-stabilizing, anti-cancer, and anti-inflammatory effects [1,2,3,4].
The oat (Avena L.) genus is represented by 25 annual plant species, including field crops, wild species, and weeds. Avena sativa is the main cultivated oat species, comprising about 90% of the world’s oat production. Other cultivated oat species, with minor importance, include black oats (Avena strigosa) and red oats (Avena byzantina). At the same time, common wild oat (Avena fatua) is a weed of cereal crops, especially oats.
Oat appeared in cultivation several thousand years later than wheat and barley, so it was unknown to ancient agricultural cultures. It was considered a secondary crop, as it initially accompanied the cultivation of other crops, mainly emmer wheat (Triticum dicoccum Schübl.), as a segetal weed.
Nowadays, globally, oat is grown on approximately 10 million hectares, accounting for only 1.5% of the acreage occupied by cereals. Oat grain is mainly used for animal feeds and human consumption but is also gaining importance in the pharmaceutical and beauty industries.
Both the hulled (more widely grown) and the naked (unhulled) forms of oat are known. The naked form is characterized by the lack of glumellae in the grain. A high proportion of glumellae in the grain of the hulled form reduces its nutritional value and limits its usefulness for feeding monogastric animals (such as poultry). However, it is an excellent component of the feed for horses, cattle, and sheep.
Oat grain is a raw material valued in cereal processing due to its nutritional and health-promoting qualities. Products of oat grain processing can be divided into three major categories, including (I) products obtained by milling: flakes (instant, mountain, ordinary), pearl barley, bran, flour, and groats; (II) products with the addition of oat grain or substances derived from oat grain: confectionery, special breads (wheat-oat bread, fine bakery products), cereal/fruit mixes (muesli), oat preparations; and (III) products derived from oat grain (e.g., β-glucan) used in pharmacology, cosmetics, brewing, and chemical industries [5,6,7,8].
Oat grains, apart from highly valuable nutritional compounds, contain substances with anti-nutritional and toxic effects, i.e., saponins (avenacosides), which can irreversibly connect to the cell membrane, thus increasing its permeability [9]. Phytic acid present in oat grain may also have a negative health impact, as it has, among others, the ability to block the bioavailability of some micronutrients, such as iron, copper, zinc, or magnesium [10]. Additionally, due to the high fiber content of oat products, their digestion may lead to flatulence, bloating, and diarrhea [11].
The present article provides an overview of the composition of oat grains and various oat products and systematizes existing knowledge about various possibilities for utilizing health-promoting properties of oat grain ingredients or oat products.

2. Chemical Characterization of Oat Grain as a Raw Material in the Food Industry

The total protein content of oat grain is relatively high compared to other cereals, ranging from 7.4 to 24.5% d. m. (dry mass) (Table 1) [12]. The protein content of naked cultivars is about 5% higher than that of hulled cultivars [13]. Oat proteins can be divided into 4 main groups, such as globulins (50–80% d. m.), prolamins (4–15% d. m.), albumins (1–12% d. m.), and glutelins (10% d. m.) [14]. Proteins isolated from oat grain are characterized by high digestibility (90.3–94.2% d. m.) and biological value (74.5–79.6% d. m.) [15].
Oat protein has been reported to have a higher digestibility-corrected amino acid score (PDCAAS) than wheat protein, but lower than soybean or pea protein [16]. The amino acid composition of oat grain is more favorable compared to other cereals due to its higher content of essential amino acids (lysine, methionine, threonine, tyrosine, leucine, valine, and phenylalanine) (Table 2) [12,13,17,18]. Research shows that people with coeliac disease, both adults and children, can safely consume products containing oat protein [15,19]. It is documented that wheat protein, gliadin, triggers inflammation in patients with coeliac disease. However, the corresponding protein in oats, avenin, has been shown not to contain similar immunogenic sequences present in gliadin [20].
The fat content of oat grain reaches approx. 2.2–11% d.m. Oat grain has the ability to accumulate more lipids in the endosperm compared to other cereals [21]. Unsaturated fatty acids account for 80% of all fatty acids present in oat grains, among which α-linolenic (1–5%), linoleic (24–48%), oleic (29–53%), docosahexaenoic, eicosapentaenoic, and arachidonic acid are the most abundant ones [22,23,24]. Among saturated fatty acids, palmitic acid is a predominant one, accounting for approx. 21.4–22.7% of total fatty acids in oats. Naked-grain oat cultivars are characterized by a significantly higher fat content than hulled oat cultivars [24,25]. When studying new oat cultivars, Kouřimská [21] noted that naked-grain cultivars had significantly higher amounts of linoleic acid and lower amounts of palmitic acid than hulled oat varieties. All analyzed oat samples had low atherogenicity (0.17–0.19) and thrombogenicity (0.30–0.34) indices. This demonstrates that the tested cultivars can be a good source of nutritionally valuable oil, which can play an important role in the prevention of cardiovascular disease [21].
The starch content of oat grain is lower compared to other cereals (Table 1). Moreover, compared to other cereals, oat starch has a smaller granule size, higher amylose content, and high viscosity and water retention capacity. Due to these characteristics, it is widely used in food products as a thickening, gelling, and coating agent [26].
Oat grain is a good source of dietary fiber, which is important for maintaining the proper functioning of the digestive system and the body as a whole. The content of total dietary fiber in oat grain ranges from 1 to 30% [13,27] and varies depending on, i.e., agronomic conditions, the anatomical structure of the kernel, the thickness of the seed coat and aleurone layer, the thickness of the endosperm cell walls, the degree of lignification of the seed coat, and whether the oat cultivar is hulled or naked. Cereal grains subjected to dehulling and milling lose significant amounts of fiber-rich parts. The ratio between soluble and insoluble fiber fractions is also altered. The total dietary fiber (TDF) content of hulled and naked oats was previously compared in some studies [13]. It was found that naked oats had significantly higher TDF levels compared to hulled oats, with values of 17.63% and 22.97%, respectively. Sykut-Domańska et al. reported that the average soluble fiber (SDF) and insoluble fiber (IDF) content of oat grain was estimated as 4.6–6.93% and 11.19–15.9%, respectively [28]. Naked oat cultivars were characterized by a lower content of both soluble and insoluble dietary fiber compared to dehusked oats.
Oat grain is an extremely valuable source of soluble fiber, especially (1-3)(1-4)-β-D-glucans, which range from 3.08 to as much as 8% d. m. [28,29]. β-glucans belong to very important fiber components due to their multidirectional preventive and therapeutic effects in many chronic non-communicable diseases [30]. β-glucans of cereal grains are composed of long unbranched chains of β-D-glucose linked by β-1,3-glycosidic (30%) and β-1,4-glycosidic (70%) bonds. Due to their structure, these polysaccharides comprise 82% of the water-soluble fraction [31]. Oat β-glucan was found to be structurally different from other cereal plant β-glucans. Most commonly, the ratio of the (DP3) and (DP4) fractions is presented in the form of a quotient considered an indicator of β-glucan structure [32]. This quotient determines physical properties such as rheological properties in solution and gel state and solubility [32]. Another indicator that influences the aforementioned properties of β-glucan is its molecular weight. β-glucan from oats has the highest molecular weight among cereal plants, ranging from 65 to 3100 × 103 Da [32,33,34]. β-(1,3)-(1,4)-D-glucan from oat grain dissolves well in water, especially warm water, and above 50 °C it dissolves completely. During dissolution, β-glucan absorbs large amounts of water, forming gums with considerable viscosity [35,36,37].
Due to the uneven distribution of (1-3)(1-4)-β-D-glucans in the grain, their content in the individual milling products can vary considerably [28,38,39]. In the case of oat and barley grains, the (1-3)(1-4)-β-D-glucans are mainly concentrated in the cell walls of the endosperm proper. The content of (1-3)(1-4)-β-D glucans in the cell walls of the endosperm of oat grains is higher compared to barley grains, at approximately 85% and 75%, respectively. Among oat products, the β-glucan content in oat bran is 4.7–8.3% d. m., in rolled oats: 2.3–8.5%, quick oats: 2.2–7.7%, and instant oats: 1.4–5.5% [40,41].
Oat grain is an important source of many minerals essential for human health. The most important macro- and micronutrients present in oat grain include Ca, Mg, Fe, Mn, Cu, Zn, P, and K [42]. Studies show that Ca and Mn in oat grain are mainly located in the aleurone layer and the embryo disc. P, K, Fe, Cu, and Zn accumulate mainly in the aleurone layer and embryo [43]. The high phosphorus and potassium content makes oatmeal a valuable source of minerals in the diet. Oat products such as flakes, flour, and bars retain most of the minerals present in the raw grain [41]. Butt et al. have shown that oat grains are rich in B vitamins such as thiamine (vitamin B1), riboflavin (vitamin B2), niacin (vitamin B3), and pantothenic acid (vitamin B5) [44]. In addition, oatmeal and oat products also contain vitamin E, which has an antioxidant function and helps protect cells from oxidative damage [44].
Oat grain also contains bioactive substances such as phytosterols (sitosterol, campesterol, and stigmasterol) with a potential to lower blood levels of low-density lipoproteins (LDLs) and inhibit atherosclerosis [45], lignans (secoisolaricirisins and matairesinol) shown to exhibit anticancer and antioxidant properties [46], and avenanthramides—phenolic compounds that are found exclusively in oat grains. Studies show that avenanthramides exhibit anti-inflammatory, antioxidant, and neuroprotective effects [47].

3. Oat in Food Technology

Oat grains are popularly used in various branches of the food industry (Table 3) [48]. One of the most popular oat products is oat flakes, often eaten as a breakfast cereal. Oat flakes are made from cleaned oat grain, in which lipolytic enzymes have been inactivated [49]. Typically, two types of oat flakes are produced: ordinary oat flakes (flaked directly after the process of grain cleaning and inactivation of lipases) and instant oat flakes (subjected additionally to hydrothermal treatment, and thus not requiring further processing/boiling before consumption) [50]. Oat flakes are also used as one of the components of cereal/fruit/nuts mixtures (muesli) or are processed into granola (oat flakes with sugar, oil, and nuts or fruits, subjected to the baking process) [51,52]. Oat grain flour is another popular oat product. Even though it is characterized by worse technological parameters compared to wheat flour, it is often used as an additive in the production of bread due to the large amount of previously discussed health-promoting substances in its composition [53,54]. In addition, oat flour has a very low glycemic index, unlike the popular wheat or rye flours, so it can be used in the production of bread for diabetics [55]. Oat grains certified as gluten-free are also broadly used to produce bread suitable for people suffering from celiac disease [56].
Oat groats, produced from oat grains in de-hulling and polishing processes, can be a substitute for such products as rice, pasta, or potatoes [57,58]. Another application of oat grains in the food industry is the production of sprouts. The sprouts are obtained as a result of the grain germination process and are intended to be consumed as a whole—as a sprout with germinating grain [59]. The sprouts are characterized by high enzymatic activity and contain a set of nutritionally important ingredients, such as vitamins (A, B, C, and E) and minerals (i.e., fluorine, iron, zinc, copper, calcium, and iodine). In addition, sprouted grains contain substances that can improve the taste and smell of food products [60,61,62,63]. Oat malts should be mentioned as another important application of oat grains [64]. Malt is produced by controlled soaking, germination, drying, and degermination of grains, primarily barley, but some malthouses produce malt from oat grains, which gives some types of beer unusual organoleptic characteristics. Such practices became popular in the 21st century, due to the so-called ‘beer revolution’, which significantly impacted the brewing industry [65,66,67,68]. Oat flakes are also used in the brewing industry due to the fact that beers brewed with the use of oat malt possess a certain creaminess, particularly well-received in dark beers or necessary for beer styles such as ‘oatmeal stout’ [69,70].
It is important to mention that with the growing popularity of vegan and vegetarian diets, the demand for protein concentrates made from plant-derived substrates has also increased [71]. One of the most popular sources of protein for the production of these food products is soybeans; however, oat grains have also found application in this branch of industry, especially in the region of Europe where soybeans, due to the climate, are not so popularly grown [14,72,73,74]. There is no way to omit here one of the other key food products for vegans and vegetarians, popularly made from oat grains, which are plant-based milk substitutes [75]. Oat ‘milk’ is made from crushed oat flakes or flour, which are subjected to hydrothermal processes and the action of β-amylase. Due to the increased content of maltose and dextrins, a drink with a degree of sweetness similar to lactose is obtained [76]. Oat grains are suitable for making oat drinks, but can also be used to make fermented products from plant-based drinks, such as oat ‘yoghurts’ [77]. It is also popular to use oat bran (a by-product of hulled oat milling) as an addition to desserts, milk, yogurt, or other food products in order to increase their fiber content [78,79]. In addition, the oat hydrolysates, i.e., oat β-glucans with amylodextrins and small amounts of minerals, fat, and protein, are also produced from oat grains. The main oat hydrolysates from this group of food products are called Oatrim. These preparations contain various amounts of β-glucans (Oatrim 3, 5, 10 containing 3, 5, and 10% of β-glucans) [63]. These hydrolysates are used in the production of low-calorie ice creams, food concentrates, sauces, biscuits, sausages, and low-fat mayonnaise [80,81,82]. In addition, an increase in the use of β-glucans in the dairy industry can be observed, which results from their positive impact on the sensory, structural, and rheological properties of the processed products, such as yogurts or ice cream [83,84].
Table 3. Examples of possible uses of oat grain and its products in the food industry [48].
Table 3. Examples of possible uses of oat grain and its products in the food industry [48].
ProductsProcess/Application MethodsCharacteristicSource
Wholegrain productsWhole oat grainsHydrothermal treatmentDecreased phytic acid contentChen et al., 2020 [85]
Germination treatmentStrengthened antioxidant activity
Fermentation treatment
Flakes (also called oatmeal)Defatting treatmentEnhanced stabilitySibakov et al., 2014 [86]
Liu et al., 2019 [87]
Konak et al., 2014 [88]
Espinosa-Solis et al., 2019 [89]
Sobota et al., 2015 [90]
Hüttner et al., 2010 [91]
Flour, branHydrothermal treatmentUsed for production of the following:
Pasta (decreased hardness),
Cookies (modified taste and texture parameters),
As an addition to meat products (reduced hardness, chewiness and viscosity, increased capacity),
bread,
Drinks (water holding)
Fermented drinksMilk substitutesFermentation treatmentIntended for people suffering from the following:
Lactose intolerance,
Allergy to milk protein,
Irritable bowel syndrome,
Low immunity,
High cholesterol,
Diarrhea,
Colorectal cancer
Mäkinen et al., 2016 [92]
Cui et al., 2023 [48]
Selmerón et al., 2015 [93]
Staka et al., 2015 [94]
Vasudha and Mishra 2013 [95]
Angelov et al., 2018 [96]
Brückner-Gühmann et al., 2019 [97]
Probiotic microorganisms
Non-dairy yogurt
Oat concentratesβ-glucan, starch, and proteinFermentation treatment
Hydratation treatment
Direct mixing
Isolates used in the production of the following:
Bread,
Pasta (oat β-glucans significantly increased viscosity of pasta; 10–15% additive of β-glucans yielded functional pasta containing 3.3–5.5 g β-glucans/100 g with high cooking quality and sensory attributes),
Fermented skimmed milk (decrease of blood serum cholesterol),
Meat products: chicken breast meat (increase in the content of soluble proteins and gel strength), low-fat beef patties (fat replacement, moisture retention enhanced)
Ronda et al., 2015 [98]
Krawęcka et al., 2020 [99]
Brückner-Gühmann et al., 2019 [97]
Lazaridou et al., 2014 [100]
Omana et al., 2011 [101]
Piñero et al., 2008 [102]

4. Oat and Human Health

Oat grains and oat products are classified as functional foods due to their high content of β-glucans, phytosterols, antioxidant compounds (e.g., polyphenolic acids, tocols, phytic acid, avenanthramides), and polyunsaturated fatty acids [103]. The available data show that consumption of 100 g of oatmeal is able to cover the daily requirement for seven essential amino acids, with only sulfur amino acids and lysine being deficient [104].
Dietary fiber consists of two fractions—soluble (SDF) and insoluble (IDF). Both dietary fiber fractions have a positive effect on the human body, but each in a different way. The interaction of both fractions results in the best effect [44]. Oat soluble fiber is a fraction that contains mainly (1-3)(1-4)-β-D-glucans. SDF forms highly viscous gels, increases the density of the digestive contents, and prolongs intestinal transit time. It has the ability to capture toxic compounds and also prevents their absorption in the intestine. It has a detoxifying effect (due to the presence of glucuronic acid) and plays an important role in alleviating lipid metabolism disorders (reduces cholesterol concentration, binds bile acids, increases excretion of fats in the stool, and delays absorption of triacylglycerols). It also causes a decrease in the time of glucose absorption [105,106].
Insoluble dietary fibers usually have high water-holding capacity and the ability to promote softening of digesta, thus supporting regular bowel movements and contributing to increased fecal bulk [44,107]. Nevertheless, a high intake of insoluble dietary fiber could be also associated with some negative nutritional impacts—it may interact with nutritionally important minerals, resulting in their lower bioavailability and thus increasing the risk of mineral deficiencies [44].

5. Oat in the Treatment of Diseases of Affluence

The risk of many chronic non-communicable diseases such as obesity, type 2 diabetes, cardiovascular diseases, and many types of cancers can be reduced by consuming an adequate dose of β-glucans in the daily diet [102,105,106]. A recent clinical study on a group of nearly 50 patients showed the beneficial effects of high molar weight oat β-glucans in chronic gastritis in humans. Consumption of β-glucans resulted in reduced mucosal damage and positive changes in fecal SCFA levels, peripheral blood serum glutathione metabolism, and antioxidant defense parameters. The described effects were noted after 30 days of β-glucans use, shedding new light on the nutritional treatment of chronic gastritis [108]. The results of in vitro studies show that oat fiber, especially β-glucan from oats, also plays an important role in preventing infectious diseases and cancer by inducing trained immunity through metabolic reprogramming. These results indicate that dietary fiber can maintain long-term reactivity of the innate immune system [109].

5.1. Oat in the Prevention and Treatment of Metabolic and Cardiovascular Diseases

A well-documented property of β-glucan is its beneficial effect on glucose metabolism. For this reason, the use of oatmeal treatment is particularly recommended for people suffering from advanced type 2 diabetes. The hypoglycemic effect of β-glucans starts already at the stage of food digestion in the stomach. In an acidic environment, these soluble components of dietary fiber form solutions of very high viscosity, slowing down the digestion process. Consequently, there is a slower ejection of gastric contents into the duodenum and a slower further digestion of carbohydrates. Due to that fact, there is a lower rate of glucose release, which translates into lower postprandial glycemia and lower insulin requirements [110]. In a study conducted on patients with type 2 diabetes, it was found that the consumption of oat fiber reduced glycated hemoglobin (HbA1c), an indicator that controls blood glucose levels, leading to lower levels [111]. Research also shows the potential of avenanthramides from oats to lower postprandial blood glucose levels. Both avenanthramides C and B have been shown to significantly inhibit the activity of the intestinal glucose-transporting proteins GLUT2 and SGLT1. These results suggest that avenanthramides may contribute to the antihyperglycemic properties associated with oat consumption [112]. Oat products, especially flour, bran, and oatmeal, have a significantly lower glycemic index compared to barley, wheat, or maize products [105,113].
The in vivo study has shown that consumption of β-glucans from oat helped to lower total cholesterol and LDL fraction in plasma, with an additional positive effect on body weight reduction in obese individuals [5]. A study of the effects of β-glucans on cardiovascular disease risks (in vivo in animals) has shown that they reduced LDL cholesterol and very low-density lipoprotein (VLDL) by 25–31% and 0.2–2.3%, respectively, and lowered total cholesterol and triglycerides levels [114]. β-glucans have the effect of reducing the absorption of cholesterol from the small intestine by forming a gel that binds cholesterol molecules. In one study, 33 volunteers with high LDL cholesterol levels consuming 3 g of β-glucans per day for 4 weeks acquired a 5.4% reduction in LDL cholesterol levels [115]. Another study of 60 overweight or obese individuals found that consuming 6 g of β-glucans per day for 12 weeks reduced total cholesterol by 5.5% and LDL cholesterol by 7.5% [116]. The findings of these studies confirm that the consumption of oat products can be an effective way to lower blood cholesterol levels, with β-glucans being the main component responsible for this effect. Studies have shown that human intestinal microflora metabolizes dietary fiber and β-glucans and provides the body with short-chain fatty acids. Increasing the ratio of propionate to acetic acid (the main substrate for cholesterol biosynthesis) results in a decrease in cholesterol biosynthesis [117]. The study showed that propionic acid and butyric acid decreased mRNA levels of 3-hydroxy-3-methylglutaryl-CoA reductase (HMG-Co-A) (rate-limiting enzyme) of cholesterol synthesis in Caco-2/TC-7 enterocytes [118].
Regular consumption of oat grain and oat products can help prevent and treat hypertension and cardiovascular diseases. Hypertension is a major contributor to heart diseases including coronary heart disease, heart attack, and stroke [119]. The mechanism of the influence of oat products on hypertension is complex and involves many factors. Oatmeal and other oat products contain high amounts of dietary fiber, β-glucans, and phytosterols, which may have beneficial effects on the cardiovascular system. β-glucan, which is the main component of oat fiber, works in several ways to lower blood pressure. Firstly, β-glucans reduce the absorption of cholesterol in the gut, which reduces the risk of atherosclerosis and cardiovascular disease. Secondly, β-glucans increase salt excretion from the body, which helps regulate blood pressure. Thirdly, β-glucans stimulate the production of nitric oxide, which has a vasodilatory effect on the blood vessels and lowers blood pressure [30]. It has been shown that the consumption of 5–7 g of β-glucans per day by individuals suffering from hypertension was effective in reducing diastolic blood pressure values by 4 mm Hg and systolic blood pressure values by 7 mm Hg [31,103].
Phytosterols, which are natural substances found in oat grain, also have blood pressure-lowering effects. Phytosterols prevent the absorption of cholesterol in the gut, which reduces the risk of atherosclerosis and cardiovascular disease. In this way, phytosterols help to maintain healthy blood vessels and regulate blood pressure [120].

5.2. Oat in Cancer Therapy

Cancers, which involve the uncontrolled growth of cells, are now counted, along with cardiovascular disease, among the main diseases of civilization of the 21st century. The β-glucans found in oat grain have been shown to be stimulators that activate, among other things, cytokines and macrophages, which are responsible for protecting the body from various infections and keeping its tissues in good condition (in vitro studies) [121,122]. The macrophages are characterized by the ability to engulf and destroy cancer cells and various pathogens [121,123]. Products extracted from oat grain have been confirmed to possess strong antioxidant (and thus anticancer) activity due to the presence of compounds such as, i.e., hydroxycinnamic and hydroxybenzoic acids, polyphenol derivatives, and vitamin E in their composition. Polyphenol derivatives show the ability to bind copper and iron, preventing their involvement during oxidation processes [104]. It has also been shown that β-glucans can counteract colorectal cancer. Under the influence of β-glucan contained in oat grain, some carcinogenic compounds (cresols, nitrosamines, estrogens) are dispersed, improving the rheological properties of the colonic contents. This prevents the fecal mass from leading to dangerous stasis, which creates inflammatory foci, often leading to ulceration, which in turn leads to cancerous foci [124,125]. In vivo and in vitro studies have confirmed that intestinal bacteria may contribute to the development of colorectal cancer by releasing genotoxic virulence factors, as well as by producing cancer-associated metabolites [126,127]. β-glucans are fermented in the cecum and colon by microflora [128,129], stimulate the growth of the probiotic Bifidobacterium and Lactobacillus bacteria, and promote the production of short-chain fatty acids (SCFAs) while inhibiting the growth of putrefactive bacteria [130,131,132,133]. Furthermore, β-glucans were shown to have the potential to prevent stomach, lung, larynx, pharynx, esophagus, breast, ovaries, uterus, and prostate cancer [33,34,121,123,124,125,134]. The destruction of tumor cells, through inhibition of tumor growth, related to the antiproliferative and proapoptotic properties of β-glucans was previously reported. β-glucans were also presented as cytotoxin-destroying, antimutagenic substances [134]. β-glucans shield human DNA and may play a huge role in the prevention of many cancers wherever there is exposure to mutagens and carcinogenic environmental effects. Gibinski et al. added that β-glucan from oat grain introduced into the body induces increased macrophage activity [33,34,123]. It has been shown that not only macrophages but also other cells of the immune system have receptors that can be activated by β-glucan. Activated macrophages engulf dead cells, resulting in tumor shrinkage [121,123,124,125].

5.3. Oat in the Fight against Overweight and Obesity

The consumption of high-fiber oat products has a positive effect on maintaining a healthy body weight and a normal body mass index [135,136,137].
Birketvedt et al. [138], in an in vivo study (a group of 60 patients), showed that the same diet without fiber resulted in a weight loss of 5.8 kg, and enriched with fiber in a loss of 8 kg. High-fiber foods are characterized by a reduced energy density, i.e., for the same meal weight and volume, they have a lower caloric value [135]. Products with a high fiber content enforce the natural need for prolonged chewing time, which increases the amount of saliva and gastric acids produced, and this in turn stimulates a more rapid achievement of the feeling of satiety, lasting even for several hours after a meal. The ability of dietary fiber (especially insoluble fraction) to bind a significant amount of water makes it a low-calorie filler that is good in satiating the feeling of hunger in the consumer [139]. Moreover, by forming viscous gels in the lumen of the small intestine, dietary fiber slows down the absorption of nutrients from food [140]. By stabilizing blood glucose and insulin levels, it also influences the body’s hormonal balance. High insulin levels change the body’s fat metabolism, promote fat accumulation in the form of body fat, and increase the feeling of hunger [141]. Both in vivo and in vitro studies confirm that oat products high in (1-3)(1-4)-β-D-glucans play an important role in the diet. By stabilizing blood glucose levels, they contribute to lowering insulin levels and make it possible to reduce hunger and maintain a feeling of satiety long after consumption [110,142,143]. Kirwan et al. have shown that eating a cereal meal 45 min before exercise improved exercise time, maintained euglycemia for longer during exercise, and resulted in greater total carbohydrate oxidation during exercise. This meal provided a significant performance and metabolic advantage [144].

5.4. Oat in the Diet of People with Inflammatory Bowel Disease

Inflammatory bowel disease (IBD) is a chronic disease characterized by inflammation of the intestinal mucosa caused by an impaired immune response. It includes Crohn’s disease (CD) and ulcerative colitis (UC). Various methods are used to treat this disease, including pharmacotherapy, surgery, and lifestyle and dietary changes [145]. Cereals are not recommended in the diet of IBD patients with the exception of oat and rice [146]. Animal studies have shown that β-glucans from oat relieve disease symptoms, e.g., diarrhea [147]. Clinical studies confirm that oat products may have a beneficial effect on IBD patients due to the high content of soluble fiber including β-glucans, which have an anti-inflammatory effect and a positive effect on the intestinal microbiome mainly due to an increase in short-chain fatty acids and especially butyric acid in the feces [148,149].

5.5. Oat in the Diet of People with Coeliac Disease

Coeliac disease (CD) is an increasingly common genetic condition today. It is generally estimated to affect 1:200 or even 1:100 people [150,151,152,153]. To date, the only available treatment for patients with coeliac disease is a lifelong gluten-free diet. Despite strict adherence to the diet, patients have difficulty achieving full restoration of the intestinal microflora, which plays a role in the processing and absorption of nutrients [154]. The inclusion of oat products in therapeutic diets improves their texture, gives a greater feeling of satiety, and increases their nutritional value due to the presence of, among others, valuable minerals, vitamins, dietary fiber, and tocochromanols [150,151,152,153]. Studies have shown that oat grain can be tolerated without significant changes in clinical symptoms [155,156], but there may be histological, serological, and immunological signs indicating an inflammatory reaction in the intestinal mucosa without signs of disease (in vitro studies) [157,158]. In vivo study showed that people who are sensitive to oat grain may have an increased incidence of diarrhea as a result of an inflammatory reaction in the intestinal mucosa (154). It may be that the processing of oat grain (such as drying, fermentation, gluten-free cleaning) and the choice of cultivar may be important factors in determining the body’s response to its consumption [159,160].

6. Conclusions

A healthy diet and fitness are the main pillars of the lifestyle being currently strongly promoted in order to ensure a long life in good mental and physical health. For this reason, there is a growing demand for foods with high nutritional value, minimally processed, and specific health-promoting properties. Oats and oat products are one of the answers to this trend. Whole-grain oat products have clinically documented preventive and curative effects on cardiovascular disease factors such as hypertension and atherosclerosis, obesity, gastrointestinal disorders, diet-related cancers, and many other conditions.
Oat grain is a valuable yet inexpensive raw material that can be used extensively in food technology. It is worthwhile to take an interest in oat in the sphere of research, processing, and use in nutrition. Wider use of oat grain in the food industry can create great opportunities for rational nutrition and thus contribute to improving the health of the population. However, the processing of oat grain meets certain difficulties related to, for example, antinutritional compounds and high fat content in the grain. Breeding work on new varieties of oat is still in progress, so it is important to constantly monitor the possibilities of their application. In addition, the growing market for oat grain concentrates and isolates allows for their wider use in the production of functional and convenient foods. The recently developing global trend of migration of regional products from one part of the world to another coinciding with efforts to replace expensive imported raw materials with their domestic substitutes increases the potential for wider use of oat grains and their compounds—for example, oat β-glucans. Further research is also needed to support oat processing in the pursuit of obtaining products of high nutritional value with functional, health-promoting properties.

Author Contributions

Conceptualization, D.L., A.W., A.G., D.Ś.-T., J.T. and R.K.; writing—original draft, D.L., A.W. and A.G.; writing—review and editing, D.L., A.W., A.G., D.Ś.-T., J.T. and R.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The results are securely maintained by the authors and can be provided upon request to individuals or groups who express interest.

Acknowledgments

We gratefully acknowledge the technical support of Justyna Obidzińska and Bartosz Szumigaj in the preparation of the submitted version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martínez-Villaluenga, C.; Peñas, E. Health benefits of oat: Current evidence and molecular mechanisms. Curr. Opin. Food Sci. 2017, 14, 26–31. [Google Scholar] [CrossRef]
  2. Schlörmann, W.; Glei, M. Potential health benefits of β-glucan from barley and oat Processing of barley and oat in Germany. Ernahr. Umsch. 2017, 64, 145–149. [Google Scholar] [CrossRef]
  3. Gangopadhyay, N.; Hossain, M.; Rai, D.; Brunton, N. A Review of Extraction and Analysis of Bioactives in Oat and Barley and Scope for Use of Novel Food Processing Technologies. Molecules 2015, 20, 10884–10909. [Google Scholar] [CrossRef] [PubMed]
  4. Kristek, A.; Schär, M.Y.; Soycan, G.; Alsharif, S.; Kuhnle, G.G.C.; Walton, G.; Spencer, J.P.E. The gut microbiota and cardiovascular health benefits: A focus on wholegrain oats. Nutr. Bull. 2018, 43, 358–373. [Google Scholar] [CrossRef]
  5. Othman, R.A.; Moghadasian, M.H.; Jones, P.J.H. Cholesterol-lowering effects of oat β-glucan. Nutr. Rev. 2011, 69, 299–309. [Google Scholar] [CrossRef]
  6. Chauhan, D.; Kumar, K.; Kumar, S.; Kumar, H. Effect of Incorporation of Oat Flour on Nutritional and Organoleptic Characteristics of Bread and Noodles. Curr. Res. Nutr. Food Sci. J. 2018, 6, 148–156. [Google Scholar] [CrossRef]
  7. Angioloni, A.; Collar, C. Suitability of Oat, Millet and Sorghum in Breadmaking. Food Bioprocess Technol. 2013, 6, 1486–1493. [Google Scholar] [CrossRef]
  8. Šubarić, D.; Babić, J.; Lalić, A.; Ačkar, Đ.; Kopjar, M. Isolation and characterisation of starch from different barley and oat varieties. Czech J. Food Sci. 2011, 29, 354–360. [Google Scholar] [CrossRef]
  9. Kozińska, N.; Tokarska, K.; Chudy, M.; Wojciechowski, K. Cytotoxicity of Quillaja saponaria Saponins towards Lung Cells Is Higher for Cholesterol-Rich Cells. Biophysica 2021, 1, 126–136. [Google Scholar] [CrossRef]
  10. Abdulwaliyu, I.; Arekemase, S.O.; Adudu, J.A.; Batari, M.L.; Egbule, M.N.; Okoduwa, S.I.R. Investigation of the medicinal significance of phytic acid as an indispensable anti-nutrient in diseases. Clin. Nutr. Exp. 2019, 28, 42–61. [Google Scholar] [CrossRef]
  11. Fabiano, G.A.; Shinn, L.M.; Antunes, A.E. Relationship between Oat Consumption, Gut Microbiota Modulation, and Short-Chain Fatty Acid Synthesis: An Integrative Review. Nutrients 2023, 15, 3534. [Google Scholar] [CrossRef] [PubMed]
  12. Leszczyńska, D. Potential use of oats. Przegląd Zbożowo-Młynarski 2021, 64, 39–41. [Google Scholar]
  13. Sterna, V.; Zute, S.; Brunava, L. Oat Grain Composition and its Nutrition Benefice. Agric. Agric. Sci. Procedia 2016, 8, 252–256. [Google Scholar] [CrossRef]
  14. Boukid, F. Oat proteins as emerging ingredients for food formulation: Where we stand? Eur. Food Res. Technol. 2021, 247, 535–544. [Google Scholar] [CrossRef]
  15. Mirmoghtadaie, L.; Kadivar, M.; Shahedi, M. Effects of succinylation and deamidation on functional properties of oat protein isolate. Food Chem. 2009, 114, 127–131. [Google Scholar] [CrossRef]
  16. Spaen, J.; Silva, J.V.C. Oat proteins: Review of extraction methods and techno-functionality for liquid and semi-solid applications. LWT 2021, 147, 111478. [Google Scholar] [CrossRef]
  17. Kriger, O.V.; Kashirskikh, E.V.; Babich, O.O.; Noskova, S.Y. Oat Protein Concentrate Production. Foods Raw Mater. 2018, 6, 47–55. [Google Scholar] [CrossRef]
  18. Li, R.; Xiong, Y.L. Ultrasound-induced structural modification and thermal properties of oat protein. LWT 2021, 149, 111861. [Google Scholar] [CrossRef]
  19. Mohamed, A.; Biresaw, G.; Xu, J.; Hojilla-Evangelista, M.P.; Rayas-Duarte, P. Oats protein isolate: Thermal, rheological, surface and functional properties. Food Res. Int. 2009, 42, 107–114. [Google Scholar] [CrossRef]
  20. Kilmartin, C. Avenin fails to induce a Th1 response in coeliac tissue following in vitro culture. Gut 2003, 52, 47–52. [Google Scholar] [CrossRef]
  21. Kouřimská, L.; Sabolová, M.; Horčička, P.; Rys, S.; Božik, M. Lipid content, fatty acid profile, and nutritional value of new oat cultivars. J. Cereal Sci. 2018, 84, 44–48. [Google Scholar] [CrossRef]
  22. Meesapyodsuk, D.; Qiu, X. A Peroxygenase Pathway Involved in the Biosynthesis of Epoxy Fatty Acids in Oat. Plant Physiol. 2011, 157, 454–463. [Google Scholar] [CrossRef]
  23. Doehlert, D.C.; Angelikousis, S.; Vick, B. Accumulation of Oxygenated Fatty Acids in Oat Lipids During Storage. Cereal Chem. J. 2010, 87, 532–537. [Google Scholar] [CrossRef]
  24. Sterna, V.; Zute, S.; Brunava, L.; Vicupe, Z. Lipid Composition of Oat Grain Grown in Latvia. 9th Balt. Conf. Food Sci. Technol. Food Consum. Well-Being 2014, 77–80. [Google Scholar]
  25. Sykut-Domańska, E.; Rzedzicki, Z.; Nita, Z. Chemical composition variability of naked and husked oat grain (Avena sativa L.). Cereal Res. Commun. 2013, 41, 327–337. [Google Scholar] [CrossRef]
  26. Zhu, F. Structures, properties, modifications, and uses of oat starch. Food Chem. 2017, 229, 329–340. [Google Scholar] [CrossRef] [PubMed]
  27. Rzedzicki, Z.; Zarzycki, P.; Wirkijowska, A.; Sobota, A.; Sykut-domańska, E.; Bartoszek, K.; Kuzawińska, E. Zboża niechlebowe źródłem błonnika w profilaktyce i zwalczaniu chorób cywilizacyjnych. Pol. J. Agron. 2016, 25, 19–26. [Google Scholar]
  28. Sykut-Domańska, E.; Rzedzicki, Z.; Zarzycki, P.; Sobota, A.; Błaszczak, W. Distribution of (1,3)(1,4)-Beta-D-Glucans in Grains of Polish Oat Cultivars and Lines (Avena sativa L.). Pol. J. Food Nutr. Sci. 2016, 66, 51–56. [Google Scholar] [CrossRef]
  29. Yang, Z.; Xie, C.; Bao, Y.; Liu, F.; Wang, H.; Wang, Y. Oat: Current state and challenges in plant-based food applications. Trends Food Sci. Technol. 2023, 134, 56–71. [Google Scholar] [CrossRef]
  30. Ho, H.V.T.; Sievenpiper, J.L.; Zurbau, A.; Blanco Mejia, S.; Jovanovski, E.; Au-Yeung, F.; Jenkins, A.L.; Vuksan, V. The effect of oat β -glucan on LDL-cholesterol, non-HDL-cholesterol and apoB for CVD risk reduction: A systematic review and meta-analysis of randomised-controlled trials. Br. J. Nutr. 2016, 116, 1369–1382. [Google Scholar] [CrossRef]
  31. Pins, J.J.; Geleva, D.; Keenan, J.M.; Frazel, C.; O’Connor, P.J.; Cherney, L.M. Do whole-grain oat cereals reduce the need for antihypertensive medications and improve blood pressure control? J. Fam. Pract. 2002, 51, 353–359. [Google Scholar]
  32. Kaur, R.; Sharma, M.; Ji, D.; Xu, M.; Agyei, D. Structural Features, Modification, and Functionalities of Beta-Glucan. Fibers 2019, 8, 1. [Google Scholar] [CrossRef]
  33. Gibiński, M. Charakterystyka chemiczna i ż ywieniowa hydrolizatów owsianych o niskim stopniu scukrzenia. Food Sci. Technol. Qual. 2008, 6, 65–76. [Google Scholar]
  34. Gibiński, M. β–Glukany Owsa Jako Składnik Żywności Funkcjonalnej. Żywność Nauka Technol. Jakość 2008, 2, 15–29. [Google Scholar]
  35. Mohebbi, Z.; Homayouni, A.; Azizi, M.H.; Hosseini, S.J. Effects of beta-glucan and resistant starch on wheat dough and prebiotic bread properties. J. Food Sci. Technol. 2018, 55, 101–110. [Google Scholar] [CrossRef]
  36. Holtekjølen, A.K.; Olsen, H.H.R.; Færgestad, E.M.; Uhlen, A.K.; Knutsen, S.H. Variations in water absorption capacity and baking performance of barley varieties with different polysaccharide content and composition. LWT Food Sci. Technol. 2008, 41, 2085–2091. [Google Scholar] [CrossRef]
  37. Skendi, A.; Biliaderis, C.G.; Papageorgiou, M.; Izydorczyk, M.S. Effects of two barley β-glucan isolates on wheat flour dough and bread properties. Food Chem. 2010, 119, 1159–1167. [Google Scholar] [CrossRef]
  38. Wirkijowska, A.; Rzedzicki, Z.; Kasprzak, M.; Błaszczak, W. Distribution of (1-3)(1-4)-β-d-glucans in kernels of selected cultivars of naked and hulled barley. J. Cereal Sci. 2012, 56, 496–503. [Google Scholar] [CrossRef]
  39. Marconi, E.; Graziano, M.; Cubadda, R. Composition and Utilization of Barley Pearling By-Products for Making Functional Pastas Rich in Dietary Fiber and β-Glucans. Cereal Chem. J. 2000, 77, 133–139. [Google Scholar] [CrossRef]
  40. Sykut-domańska, E. Charakterystyka wybranych sortymentów zbóż śniadaniowych dostępnych na rynku polskim i brytyjskim. Bromat. Chem. 2012, 45, 72–82. [Google Scholar]
  41. Achremowicz, B.; Haber, T.; Kaszuba, J.; Puchalski, C.; Wiśniewski, R. Płatki zbożowe—ocena porównawcza. Część I Porównanie składu chemicznego i mineralnego. Postępy Tech. Przetwórstwa Spożywczego 2016, 2, 97–102. [Google Scholar]
  42. de Oliveira Maximino, J.V.; Barros, L.M.; Pereira, R.M.; de Santi, I.I.; Aranha, B.C.; Busanello, C.; Viana, V.E.; Freitag, R.A.; Batista, B.L.; Costa de Oliveira, A.; et al. Mineral and Fatty Acid Content Variation in White Oat Genotypes Grown in Brazil. Biol. Trace Elem. Res. 2021, 199, 1194–1206. [Google Scholar] [CrossRef] [PubMed]
  43. Deng, G.; Vu, M.; Korbas, M.; Bondici, V.F.; Karunakaran, C.; Christensen, D.; Bart Lardner, H.A.; Yu, P. Distribution of micronutrients in Arborg oat (Avena sativa L.) using synchrotron X-ray fluorescence imaging. Food Chem. 2023, 421, 135661. [Google Scholar] [CrossRef]
  44. Butt, M.S.; Tahir-Nadeem, M.; Khan, M.K.I.; Shabir, R.; Butt, M.S. Oat: Unique among the cereals. Eur. J. Nutr. 2008, 47, 68–79. [Google Scholar] [CrossRef]
  45. Weggemans, R.M.; Trautwein, E.A. Relation between soy-associated isoflavones and LDL and HDL cholesterol concentrations in humans: A meta-analysis. Eur. J. Clin. Nutr. 2003, 57, 940–946. [Google Scholar] [CrossRef] [PubMed]
  46. Saarinen, N.M.; Wärri, A.; Airio, M.; Smeds, A.; Mäkelä, S. Role of dietary lignans in the reduction of breast cancer risk. Mol. Nutr. Food Res. 2007, 51, 857–866. [Google Scholar] [CrossRef] [PubMed]
  47. Andersson, A.A.M.; Dimberg, L.; Åman, P.; Landberg, R. Recent findings on certain bioactive components in whole grain wheat and rye. J. Cereal Sci. 2014, 59, 294–311. [Google Scholar] [CrossRef]
  48. Cui, L.; Jia, Q.; Zhao, J.; Hou, D.; Zhou, S. A comprehensive review on oat milk: From oat nutrients and phytochemicals to its processing technologies, product features, and potential applications. Food Funct. 2023, 14, 5858–5869. [Google Scholar] [CrossRef]
  49. Zwer, P. Oats: Characteristics and quality requirements. In Cereal Grains; Wrigley, C.W., Batey, I.L.B.T.-C.G., Eds.; Elsevier: Amsterdam, The Netherlands, 2010; pp. 163–182. ISBN 978-1-84569-563-7. [Google Scholar]
  50. Strychar, R. Oats: Chemistry and technology. In World Oat Production, Trade, and Usage.; Webster, F., Wood, P., Eds.; American Association of Cereal Chemists, Inc. (AACC): Saint Paul, MN, USA, 2011; pp. 1–10. [Google Scholar]
  51. Sobota, A.; Rzedzicki, Z.; Sobieraj, M. Badania Składu Chemicznego Płatków Musli. Bromatol. Chem. Toksykol. 2012, 45, 131–137. [Google Scholar]
  52. Aigster, A.; Duncan, S.E.; Conforti, F.D.; Barbeau, W.E. Physicochemical properties and sensory attributes of resistant starch-supplemented granola bars and cereals. LWT Food Sci. Technol. 2011, 44, 2159–2165. [Google Scholar] [CrossRef]
  53. Gambu, H.; Gibi, M.; Pastuszka, D.; Mickowska, B.; Ziobro, R.; Witkowicz, R. the Application of Residual Oats Flour in Bread Production in Order To Improve. Acta Sci. Pol. Technol. Aliment 2011, 10, 313–325. [Google Scholar]
  54. Raihan, M.; Saini, C.S. Evaluation of various properties of composite flour from oats, sorghum, amaranth and wheat flour and production of cookies thereof. Int. Food Res. J. 2017, 24, 2278–2284. [Google Scholar]
  55. Kim, H.J.; White, P.J. In Vitro Digestion Rate and Estimated Glycemic Index of Oat Flours from Typical and High β-Glucan Oat Lines. J. Agric. Food Chem. 2012, 60, 5237–5242. [Google Scholar] [CrossRef]
  56. Smulders, M.J.M.; van de Wiel, C.C.M.; van den Broeck, H.C.; van der Meer, I.M.; Israel-Hoevelaken, T.P.M.; Timmer, R.D.; van Dinter, B.-J.; Braun, S.; Gilissen, L.J.W.J. Oats in healthy gluten-free and regular diets: A perspective. Food Res. Int. 2018, 110, 3–10. [Google Scholar] [CrossRef] [PubMed]
  57. Borowy, T.; Kubiak, M.S. Produkty przerobu owsa. Piekarstwo. Spec. Czas. Piekarzy 2012, 2, 47–49. [Google Scholar]
  58. Przygodzki, R.; Korbas, E.; Jóźwiak, I.; Langner, R. Właściwości Fizyczne Kasz Instant Nowej Generacji. Postępy Nauk. Technol. Przem. Rolno-Spożywczego 2012, 67, 52–63. [Google Scholar]
  59. Liska, D.J.; Dioum, E.; Chu, Y.; Mah, E. Narrative Review on the Effects of Oat and Sprouted Oat Components on Blood Pressure. Nutrients 2022, 14, 4772. [Google Scholar] [CrossRef]
  60. Brandolini, A.; Hidalgo, A. Wheat germ: Not only a by-product. Int. J. Food Sci. Nutr. 2012, 63, 71–74. [Google Scholar] [CrossRef]
  61. Oliinyk, S.; Samokhvalova, O.; Lapitska, N.; Kucheruk, Z. Studying the influence of meats from wheat and oat germs, and rose hips, on the formation of quality of rye w heat dough and bread. East. -Eur. J. Enterp. Technol. 2020, 1, 59–65. [Google Scholar] [CrossRef]
  62. Panfil, P.; Dorica, B.; Sorin, C.; Emilian, M.; Ersilia, A.; Iosif, G. Biochemical characterization of flour obtained from germinated cereals (wheat, barley and oat). Rom. Biotechnol. Lett. 2014, 19, 9772–9777. [Google Scholar]
  63. Kim, S.; Inglett, G.E.; Liu, S.X. Content and Molecular Weight Distribution of Oat β-Glucan in Oatrim, Nutrim, and C-Trim Products. Cereal Chem. J. 2008, 85, 701–705. [Google Scholar] [CrossRef]
  64. Klose, C.; Mauch, A.; Wunderlich, S.; Thiele, F.; Zarnkow, M.; Jacob, F.; Arendt, E.K. Brewing with 100% Oat Malt. J. Inst. Brew. 2011, 117, 411–421. [Google Scholar] [CrossRef]
  65. Gasiński, A.; Kawa-Rygielska, J.; Błażewicz, J.; Leszczyńska, D. Malting procedure and its impact on the composition of volatiles and antioxidative potential of naked and covered oat varieties. J. Cereal Sci. 2022, 107, 103537. [Google Scholar] [CrossRef]
  66. Briggs, D. Malts and Malting, 1st ed.; Springer: New York, NY, USA, 1998. [Google Scholar]
  67. Garavaglia, C.; Swinnen, J. The Craft Beer Revolution. Choices 2017, 32, 1–8. [Google Scholar]
  68. Garavaglia, C.; Swinnen, J. Economics of the Craft Beer Revolution: A Comparative International Perspective. In Economic Perspectives on Craft Beer; Garavaglia, C., Swinnen, J., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 3–51. ISBN 978-3-319-58235-1. [Google Scholar]
  69. Schnitzenbaumer, B.; Kaspar, J.; Titze, J.; Arendt, E.K. Implementation of commercial oat and sorghum flours in brewing. Eur. Food Res. Technol. 2014, 238, 515–525. [Google Scholar] [CrossRef]
  70. Strong, G.; England, K. Beer Judge Certification Program; University of California: Los Angeles, CA, USA, 2019. [Google Scholar]
  71. Ismail, B.P.; Senaratne-Lenagala, L.; Stube, A.; Brackenridge, A. Protein demand: Review of plant and animal proteins used in alternative protein product development and production. Anim. Front. 2020, 10, 53–63. [Google Scholar] [CrossRef]
  72. González-Pérez, S.; Arellano, J.B. Vegetable protein isolates. In Handbook of Hydrocolloids; Phillips, G.O., Williams, P.A.B.T.-H., Second, E., Eds.; Elsevier: Amsterdam, The Netherlands, 2009; pp. 383–419. ISBN 978-1-84569-414-2. [Google Scholar]
  73. Fuhrman, J.; Ferreri, D.M. Fueling the Vegetarian (Vegan) Athlete. Curr. Sports Med. Rep. 2010, 9, 233–241. [Google Scholar] [CrossRef]
  74. D’adamo, C.; Sahin, A. Soy Foods and Supplementation: A Review of Commonly Perceived Health Benefits and Risks. Altern. Ther. 2014, 20, 39–52. [Google Scholar]
  75. Aydar, E.F.; Tutuncu, S.; Ozcelik, B. Plant-based milk substitutes: Bioactive compounds, conventional and novel processes, bioavailability studies, and health effects. J. Funct. Foods 2020, 70, 103975. [Google Scholar] [CrossRef]
  76. Paul, A.A.; Kumar, S.; Kumar, V.; Sharma, R. Milk Analog: Plant based alternatives to conventional milk, production, potential and health concerns. Crit. Rev. Food Sci. Nutr. 2020, 60, 3005–3023. [Google Scholar] [CrossRef]
  77. Bernat, N.; Cháfer, M.; González-Martínez, C.; Rodríguez-García, J.; Chiralt, A. Optimisation of oat milk formulation to obtain fermented derivatives by using probiotic Lactobacillus reuteri microorganisms. Food Sci. Technol. Int. 2015, 21, 145–157. [Google Scholar] [CrossRef] [PubMed]
  78. Steinert, R.; Raederstorff, D.; Wolever, T. Effect of Consuming Oat Bran Mixed in Water before a Meal on Glycemic Responses in Healthy Humans—A Pilot Study. Nutrients 2016, 8, 524. [Google Scholar] [CrossRef] [PubMed]
  79. Jane, M.; McKay, J.; Pal, S. Effects of daily consumption of psyllium, oat bran and polyGlycopleX on obesity-related disease risk factors: A critical review. Nutrition 2019, 57, 84–91. [Google Scholar] [CrossRef] [PubMed]
  80. Lee, S.; Kim, S.; Inglett, G.E. Effect of Shortening Replacement with Oatrim on the Physical and Rheological Properties of Cakes. Cereal Chem. J. 2005, 82, 120–124. [Google Scholar] [CrossRef]
  81. Mathews, R.; Kamil, A.; Chu, Y. Global review of heart health claims for oat beta-glucan products. Nutr. Rev. 2020, 78, 78–97. [Google Scholar] [CrossRef]
  82. Dimopoulos, G.; Tsantes, M.; Taoukis, P. Effect of high pressure homogenization on the production of yeast extract via autolysis and beta-glucan recovery. Innov. Food Sci. Emerg. Technol. 2020, 62, 102340. [Google Scholar] [CrossRef]
  83. Khorshidian, N.; Yousefi, M.; Shadnoush, M.; Mortazavian, A.M. An Overview of β-Glucan Functionality in Dairy Products. Curr. Nutr. Food Sci. 2018, 14, 280–292. [Google Scholar] [CrossRef]
  84. Mykhalevych, A.; Polishchuk, G.; Nassar, K.; Osmak, T.; Buniowska-Olejnik, M. β-Glucan as a Techno-Functional Ingredient in Dairy and Milk-Based Products—A Review. Molecules 2022, 27, 6313. [Google Scholar] [CrossRef]
  85. Chen, G.; Liu, Y.; Zeng, J.; Tian, X.; Bei, Q.; Wu, Z. Enhancing three phenolic fractions of oats (Avena sativa L.) and their antioxidant activities by solid-state fermentation with Monascus anka and Bacillus subtilis. J. Cereal Sci. 2020, 93, 102940. [Google Scholar] [CrossRef]
  86. Sibakov, J.; Abecassis, J.; Barron, C.; Poutanen, K. Electrostatic separation combined with ultra-fine grinding to produce β-glucan enriched ingredients from oat bran. Innov. Food Sci. Emerg. Technol. 2014, 26, 445–455. [Google Scholar] [CrossRef]
  87. Liu, S.; Li, Y.; Obadi, M.; Jiang, Y.; Chen, Z.; Jiang, S.; Xu, B. Effect of steaming and defatting treatments of oats on the processing and eating quality of noodles with a high oat flour content. J. Cereal Sci. 2019, 89, 102794. [Google Scholar] [CrossRef]
  88. Konak, Ü.İ.; Ercili-Cura, D.; Sibakov, J.; Sontag-Strohm, T.; Certel, M.; Loponen, J. CO2-defatted oats: Solubility, emulsification and foaming properties. J. Cereal Sci. 2014, 60, 37–41. [Google Scholar] [CrossRef]
  89. Espinosa-Solis, V.; Zamudio-Flores, P.B.; Tirado-Gallegos, J.M.; Ramírez-Mancinas, S.; Olivas-Orozco, G.I.; Espino-Díaz, M.; Hernández-González, M.; García-Cano, V.G.; Sánchez-Ortíz, O.; Buenrostro-Figueroa, J.J.; et al. Evaluation of Cooking Quality, Nutritional and Texture Characteristics of Pasta Added with Oat Bran and Apple Flour. Foods 2019, 8, 299. [Google Scholar] [CrossRef] [PubMed]
  90. Sobota, A.; Rzedzicki, Z.; Zarzycki, P.; Kuzawińska, E. Application of common wheat bran for the industrial production of high-fibre pasta. Int. J. Food Sci. Technol. 2015, 50, 111–119. [Google Scholar] [CrossRef]
  91. Hüttner, E.K.; Bello, F.D.; Arendt, E.K. Rheological properties and bread making performance of commercial wholegrain oat flours. J. Cereal Sci. 2010, 52, 65–71. [Google Scholar] [CrossRef]
  92. Mäkinen, O.E.; Wanhalinna, V.; Zannini, E.; Arendt, E.K. Foods for Special Dietary Needs: Non-dairy Plant-based Milk Substitutes and Fermented Dairy-type Products. Crit. Rev. Food Sci. Nutr. 2016, 56, 339–349. [Google Scholar] [CrossRef]
  93. Salmerón, I.; Thomas, K.; Pandiella, S.S. Effect of potentially probiotic lactic acid bacteria on the physicochemical composition and acceptance of fermented cereal beverages. J. Funct. Foods 2015, 15, 106–115. [Google Scholar] [CrossRef]
  94. Staka, A.; Bodnieks, E.; Puķītis, A. Impact of Oat-Based Products on Human Gastrointestinal Tract. Proc. Latv. Acad. Sci. Sect. B Nat. Exact Appl. Sci. 2015, 69, 145–151. [Google Scholar] [CrossRef]
  95. Vasudha, S.; Mishra, H.N. Non dairy probiotic beverages. Int. Food Res. J. 2013, 20, 7–15. [Google Scholar]
  96. Angelov, A.; Yaneva-Marinova, T.; Gotcheva, V. Oats as a matrix of choice for developing fermented functional beverages. J. Food Sci. Technol. 2018, 55, 2351–2360. [Google Scholar] [CrossRef]
  97. Brückner-Gühmann, M.; Vasil’eva, E.; Culetu, A.; Duta, D.; Sozer, N.; Drusch, S. Oat protein concentrate as alternative ingredient for non-dairy yoghurt-type product. J. Sci. Food Agric. 2019, 99, 5852–5857. [Google Scholar] [CrossRef] [PubMed]
  98. Ronda, F.; Perez-Quirce, S.; Lazaridou, A.; Biliaderis, C.G. Effect of barley and oat β-glucan concentrates on gluten-free rice-based doughs and bread characteristics. Food Hydrocoll. 2015, 48, 197–207. [Google Scholar] [CrossRef]
  99. Krawęcka, A.; Sobota, A.; Sykut-Domańska, E. Physicochemical, Sensory, and Cooking Qualities of Pasta Enriched with Oat β-Glucans, Xanthan Gum, and Vital Gluten. Foods 2020, 9, 1412. [Google Scholar] [CrossRef] [PubMed]
  100. Lazaridou, A.; Serafeimidou, A.; Biliaderis, C.G.; Moschakis, T.; Tzanetakis, N. Structure development and acidification kinetics in fermented milk containing oat β-glucan, a yogurt culture and a probiotic strain. Food Hydrocoll. 2014, 39, 204–214. [Google Scholar] [CrossRef]
  101. Omana, D.; Plastow, G.; Betti, M. The use of β-glucan as a partial salt replacer in high pressure processed chicken breast meat. Food Chem. 2011, 129, 768–776. [Google Scholar] [CrossRef]
  102. Piñero, M.P.; Parra, K.; Huerta-Leidenz, N.; Arenas de Moreno, L.; Ferrer, M.; Araujo, S.; Barboza, Y. Effect of oat’s soluble fibre (β-glucan) as a fat replacer on physical, chemical, microbiological and sensory properties of low-fat beef patties. Meat Sci. 2008, 80, 675–680. [Google Scholar] [CrossRef]
  103. Lange, E. Produkty Owsiane Jako Żywność Funkcjonalna. Żywność Nauka Technol. Jakość 2010, 3, 7–24. [Google Scholar]
  104. Zygmunt Zdrojewicz, A.L.A.W. Influence of consumption the oatmeal on human body. Med. Rodz. 2017, 20, 118–123. [Google Scholar]
  105. Daou, C.; Zhang, H. Oat Beta-Glucan: Its Role in Health Promotion and Prevention of Diseases. Compr. Rev. Food Sci. Food Saf. 2012, 11, 355–365. [Google Scholar] [CrossRef]
  106. El Khoury, D.; Cuda, C.; Luhovyy, B.L.; Anderson, G.H. Beta glucan: Health benefits in obesity and metabolic syndrome. J. Nutr. Metab. 2012, 2012, 851362. [Google Scholar] [CrossRef]
  107. Gidley, M.J.; Yakubov, G.E. Functional categorisation of dietary fibre in foods: Beyond ‘soluble’ vs ‘insoluble’. Trends Food Sci. Technol. 2019, 86, 563–568. [Google Scholar] [CrossRef]
  108. Gudej, S.; Filip, R.; Harasym, J.; Wilczak, J.; Dziendzikowska, K.; Oczkowski, M.; Jałosińska, M.; Juszczak, M.; Lange, E.; Gromadzka-Ostrowska, J. Clinical Outcomes after Oat Beta-Glucans Dietary Treatment in Gastritis Patients. Nutrients 2021, 13, 2791. [Google Scholar] [CrossRef]
  109. Pan, W.; Hao, S.; Zheng, M.; Lin, D.; Jiang, P.; Zhao, J.; Shi, H.; Yang, X.; Li, X.; Yu, Y. Oat-Derived β-Glucans Induced Trained Immunity Through Metabolic Reprogramming. Inflammation 2020, 43, 1323–1336. [Google Scholar] [CrossRef] [PubMed]
  110. Brennan, C.S. Dietary fibre, glycaemic response, and diabetes. Mol. Nutr. Food Res. 2005, 49, 560–570. [Google Scholar] [CrossRef] [PubMed]
  111. Kristek, A.; Wiese, M.; Heuer, P.; Kosik, O.; Schär, M.Y.; Soycan, G.; Alsharif, S.; Kuhnle, G.G.C.; Walton, G.; Spencer, J.P.E. Oat bran, but not its isolated bioactive β -glucans or polyphenols, have a bifidogenic effect in an in vitro fermentation model of the gut microbiota. Br. J. Nutr. 2019, 121, 549–559. [Google Scholar] [CrossRef] [PubMed]
  112. Zhouyao, H.; Malunga, L.N.; Chu, Y.F.; Eck, P.; Ames, N.; Thandapilly, S.J. The inhibition of intestinal glucose absorption by oat-derived avenanthramides. J. Food Biochem. 2022, 46, e14324. [Google Scholar] [CrossRef]
  113. Jenkins, A.; Jenkins, D.; Zdravkovic, U.; Würsch, P.; Vuksan, V. Depression of the glycemic index by high levels of β-glucan fiber in two functional foods tested in type 2 diabetes. Eur. J. Clin. Nutr. 2002, 56, 622–628. [Google Scholar] [CrossRef]
  114. Bae, I.Y.; Kim, S.M.; Lee, S.; Lee, H.G. Effect of enzymatic hydrolysis on cholesterol-lowering activity of oat β-glucan. N. Biotechnol. 2010, 27, 85–88. [Google Scholar] [CrossRef]
  115. Zhang, J.; Li, L.; Song, P.; Wang, C.; Man, Q.; Meng, L.; Cai, J.; Kurilich, A. Randomized controlled trial of oatmeal consumption versus noodle consumption on blood lipids of urban Chinese adults with hypercholesterolemia. Nutr. J. 2012, 11, 54. [Google Scholar] [CrossRef]
  116. Shimizu, C.; Kihara, M.; Aoe, S.; Araki, S.; Ito, K.; Hayashi, K.; Watari, J.; Sakata, Y.; Ikegami, S. Effect of high β-glucan barley on serum cholesterol concentrations and visceral fat area in Japanese men—A randomized, double-blinded, placebo-controlled trial. Plant Foods Hum. Nutr. 2008, 63, 21–25. [Google Scholar] [CrossRef]
  117. Theuwissen, E.; Mensink, R.P. Water-soluble dietary fibers and cardiovascular disease. Physiol. Behav. 2008, 94, 285–292. [Google Scholar] [CrossRef] [PubMed]
  118. Alvaro, A.; Solà, R.; Rosales, R.; Ribalta, J.; Anguera, A.; Masana, L.; Vallvé, J.C. Gene expression analysis of a human enterocyte cell line reveals downregulation of cholesterol biosynthesis in response to short-chain fatty acids. IUBMB Life 2008, 60, 757–764. [Google Scholar] [CrossRef] [PubMed]
  119. Muntner, P.; Miles, M.A.; Jaeger, B.C.; Hannon, L.; Hardy, S.T.; Ostchega, Y.; Wozniak, G.; Schwartz, J.E. Blood Pressure Control Among US Adults, 2009 to 2012 Through 2017 to 2020. Hypertension 2022, 79, 1971–1980. [Google Scholar] [CrossRef] [PubMed]
  120. De Spirt, S.; Stahl, W.; Tronnier, H.; Sies, H.; Bejot, M.; Maurette, J.-M.; Heinrich, U. Intervention with flaxseed and borage oil supplements modulates skin condition in women. Br. J. Nutr. 2008, 101, 440–445. [Google Scholar] [CrossRef]
  121. Choromanska, A.; Kulbacka, J.; Rembialkowska, N.; Pilat, J.; Oledzki, R.; Harasym, J.; Saczko, J. Anticancer properties of low molecular weight oat beta-glucan—An in vitro study. Int. J. Biol. Macromol. 2015, 80, 23–28. [Google Scholar] [CrossRef]
  122. Ghosh, S.K.; Sanyal, T. Anti-cancer property of Lenzites betulina (L) Fr. on cervical cancer cell lines and its anti-tumor effect on HeLa-implanted mice. bioRxiv 2019, 2019, 540567. [Google Scholar]
  123. Lee, J.S.; Lee, S.H.; Jang, Y.M.; Lee, J.D.; Lee, B.H.; Jung, J.Y. Macrophage and Anticancer Activities of Feed Additives on β-Glucan from Schizophyllum commune in Breast Cancer Cells. J. Korean Soc. Food Sci. Nutr. 2011, 40, 949–955. [Google Scholar] [CrossRef]
  124. Chen, J.; Zhang, X.D.; Jiang, Z. The Application of Fungal Beta-glucans for the Treatment of Colon Cancer. Anticancer. Agents Med. Chem. 2013, 13, 725–730. [Google Scholar] [CrossRef]
  125. Baldassano, S.; Accardi, G.; Vasto, S. Beta-glucans and cancer: The influence of inflammation and gut peptide. Eur. J. Med. Chem. 2017, 142, 486–492. [Google Scholar] [CrossRef]
  126. Kostic, A.D.; Chun, E.; Robertson, L.; Glickman, J.N.; Gallini, C.A.; Michaud, M.; Clancy, T.E.; Chung, D.C.; Lochhead, P.; Hold, G.L.; et al. Fusobacterium nucleatum Potentiates Intestinal Tumorigenesis and Modulates the Tumor-Immune Microenvironment. Cell Host Microbe 2013, 14, 207–215. [Google Scholar] [CrossRef]
  127. Zackular, J.P.; Rogers, M.A.M.; Ruffin, M.T.; Schloss, P.D. The Human Gut Microbiome as a Screening Tool for Colorectal Cancer. Cancer Prev. Res. 2014, 7, 1112. [Google Scholar] [CrossRef] [PubMed]
  128. Daniel, M.; Wisker, E.; Rave, G.; Feldheim, W. Fermentation in human subjects of nonstarch polysaccharides in mixed diets, but not in a barley fiber concentrate, could be predicted by in vitro fermentation using human fecal inocula. J. Nutr. 1997, 127, 1981–1988. [Google Scholar] [CrossRef] [PubMed]
  129. Topping, D.L.; Clifton, P.M. Short-Chain Fatty Acids and Human Colonic Function: Roles of Resistant Starch and Nonstarch Polysaccharides. Physiol. Rev. 2018, 81, 1031–1064. [Google Scholar] [CrossRef] [PubMed]
  130. Pozuelo, M.J.; Agis-Torres, A.; Hervert-Hernández, D.; Elvira López-Oliva, M.; Muñoz-Martínez, E.; Rotger, R.; Goñi, I. Grape Antioxidant Dietary Fiber Stimulates Lactobacillus Growth in Rat Cecum. J. Food Sci. 2012, 77, H59–H62. [Google Scholar] [CrossRef] [PubMed]
  131. Nilsson, U.; Johansson, M.; Nilsson, Å.; Björck, I.; Nyman, M. Dietary supplementation with β-glucan enriched oat bran increases faecal concentration of carboxylic acids in healthy subjects. Eur. J. Clin. Nutr. 2008, 62, 978–984. [Google Scholar] [CrossRef]
  132. Hinnebusch, B.F.; Meng, S.; Wu, J.T.; Archer, S.Y.; Hodin, R.A. The Effects of Short-Chain Fatty Acids on Human Colon Cancer Cell Phenotype Are Associated with Histone Hyperacetylation. J. Nutr. 2002, 132, 1012–1017. [Google Scholar] [CrossRef]
  133. Thomas, L.V.; Ockhuizen, T.; Suzuki, K. Exploring the influence of the gut microbiota and probiotics on health: A symposium report. Br. J. Nutr. 2014, 112, S1–S18. [Google Scholar] [CrossRef]
  134. Chaichian, S.; Moazzami, B.; Sadoughi, F.; Haddad Kashani, H.; Zaroudi, M.; Asemi, Z. Functional activities of beta-glucans in the prevention or treatment of cervical cancer. J. Ovarian Res. 2020, 13, 24. [Google Scholar] [CrossRef]
  135. Howarth, N.C.; Sc, M.; Saltzman, E.; Roberts, S.B.; Ph, D. Dietary Fiber and Weight Regulation. Nutr. Rev. 2001, 59, 129–139. [Google Scholar] [CrossRef]
  136. Baboota, R.K.; Bishnoi, M.; Ambalam, P.; Kondepudi, K.K.; Sarma, S.M.; Boparai, R.K.; Podili, K. Functional food ingredients for the management of obesity and associated co-morbidities—A review. J. Funct. Foods 2013, 5, 997–1012. [Google Scholar] [CrossRef]
  137. Patel, S. Cereal bran fortified-functional foods for obesity and diabetes management: Triumphs, hurdles and possibilities. J. Funct. Foods 2015, 14, 255–269. [Google Scholar] [CrossRef]
  138. Birketvedt, G.S.; Aaseth, J.; Florholmen, J.R.; Ryttig, K. Long Term Effect of Fibre Supplement and Reduced Energy Intake on Body Weight and Blood Lipids in Overweight Subjects. Acta Medica 2000, 43, 129–132. [Google Scholar] [CrossRef]
  139. Ludwig, D.S. Dietary Glycemic Index and Obesity. J. Nutr. 2000, 130, 280S–283S. [Google Scholar] [CrossRef] [PubMed]
  140. Liefschitz, C.H.; Grusak, M.A.; Butte, N.F. Carbohydrate digestion in humans from a β-glucan-enriched barley is reduced. J. Nutr. 2002, 132, 2593–2596. [Google Scholar] [CrossRef] [PubMed]
  141. Ludwig, D.S.; Pereira, M.A.; Kroenke, C.H.; Hilner, J.E.; Van Horn, L.; Slattery, M.L.; Jacobs, D.R. Dietary Fiber, Weight Gain, and Cardiovascular Disease Risk Factors in Young Adults. JAMA 1999, 282, 1539. [Google Scholar] [CrossRef] [PubMed]
  142. Östman, E.; Rossi, E.; Larsson, H.; Brighenti, F.; Björck, I. Glucose and insulin responses in healthy men to barley bread with different levels of (1→3;1→4)-β-glucans; predictions using fluidity measurements of in vitro enzyme digests. J. Cereal Sci. 2006, 43, 230–235. [Google Scholar] [CrossRef]
  143. Weickert, M.; Möhlig, M.; Schöfl, C.; Arafat, A.; Otto, B.; Viehoff, H.; Koebnick, C.; Kohl, A.; Spranger, J.; Pfeiffer, A. Cereal fiber improves whole-body insulin. Diabetes Care 2006, 29, 773–780. [Google Scholar] [CrossRef]
  144. Kirwan, J.P.; Cyr-Campbell, D.; Campbell, W.W.; Scheiber, J.; Evans, W.J. Effects of moderate and high glycemic index meals on metabolism and exercise performance. Metabolism 2001, 50, 849–855. [Google Scholar] [CrossRef]
  145. Singh, R.K.; Chang, H.W.; Yan, D.; Lee, K.M.; Ucmak, D.; Wong, K.; Abrouk, M.; Farahnik, B.; Nakamura, M.; Zhu, T.H.; et al. Influence of diet on the gut microbiome and implications for human health. J. Transl. Med. 2017, 15, 73. [Google Scholar] [CrossRef]
  146. Cox, S.R.; Lindsay, J.O.; Fromentin, S.; Stagg, A.J.; McCarthy, N.E.; Galleron, N.; Ibraim, S.B.; Roume, H.; Levenez, F.; Pons, N.; et al. Effects of Low FODMAP Diet on Symptoms, Fecal Microbiome, and Markers of Inflammation in Patients With Quiescent Inflammatory Bowel Disease in a Randomized Trial. Gastroenterology 2020, 158, 176–188.e7. [Google Scholar] [CrossRef]
  147. Liu, B.; Lin, Q.; Yang, T.; Zeng, L.; Shi, L.; Chen, Y.; Luo, F. Oat β-glucan ameliorates dextran sulfate sodium (DSS)-induced ulcerative colitis in mice. Food Funct. 2015, 6, 3454–3463. [Google Scholar] [CrossRef] [PubMed]
  148. Campmans-Kuijpers, M.J.E.; Dijkstra, G. Food and Food Groups in Inflammatory Bowel Disease (IBD): The Design of the Groningen Anti-Inflammatory Diet (GrAID). Nutrients 2021, 13, 1067. [Google Scholar] [CrossRef] [PubMed]
  149. Hallert, C.; Björck, I.; Nyman, M.; Pousette, A.; Grännö, C.; Svensson, H. Increasing Fecal Butyrate in Ulcerative Colitis Patients by Diet: Controlled Pilot Study. Inflamm. Bowel Dis. 2003, 9, 116–121. [Google Scholar] [CrossRef]
  150. Comino, I. Role of oats in celiac disease. World J. Gastroenterol. 2015, 21, 11825. [Google Scholar] [CrossRef]
  151. Fric, P.; Gabrovska, D.; Nevoral, J. Celiac disease, gluten-free diet, and oats. Nutr. Rev. 2011, 69, 107–115. [Google Scholar] [CrossRef] [PubMed]
  152. Kaukinen, K.; Collin, P.; Huhtala, H.; Mäki, M. Long-Term Consumption of Oats in Adult Celiac Disease Patients. Nutrients 2013, 5, 4380–4389. [Google Scholar] [CrossRef]
  153. Silano, M.; Penas Pozo, E.; Uberti, F.; Manferdelli, S.; Del Pinto, T.; Felli, C.; Budelli, A.; Vincentini, O.; Restani, P. Diversity of oat varieties in eliciting the early inflammatory events in celiac disease. Eur. J. Nutr. 2014, 53, 1177–1186. [Google Scholar] [CrossRef]
  154. Marasco, G.; Cirota, G.G.; Rossini, B.; Lungaro, L.; Di Biase, A.R.; Colecchia, A.; Volta, U.; De Giorgio, R.; Festi, D.; Caio, G. Probiotics, Prebiotics and Other Dietary Supplements for Gut Microbiota Modulation in Celiac Disease Patients. Nutrients 2020, 12, 2674. [Google Scholar] [CrossRef]
  155. Garsed, K.; Scott, B.B. Can oats be taken in a gluten-free diet? A systematic review. Scand. J. Gastroenterol. 2007, 42, 171–178. [Google Scholar] [CrossRef]
  156. Pinto-Sánchez, M.I.; Causada-Calo, N.; Bercik, P.; Ford, A.C.; Murray, J.A.; Armstrong, D.; Semrad, C.; Kupfer, S.S.; Alaedini, A.; Moayyedi, P.; et al. Safety of Adding Oats to a Gluten-Free Diet for Patients With Celiac Disease: Systematic Review and Meta-analysis of Clinical and Observational Studies. Gastroenterology 2017, 153, 395–409.e3. [Google Scholar] [CrossRef]
  157. Lionetti, E.; Gatti, S.; Galeazzi, T.; Caporelli, N.; Francavilla, R.; Cucchiara, S.; Roggero, P.; Malamisura, B.; Iacono, G.; Tomarchio, S.; et al. Safety of Oats in Children with Celiac Disease: A Double-Blind, Randomized, Placebo-Controlled Trial. J. Pediatr. 2018, 194, 116–122.e2. [Google Scholar] [CrossRef] [PubMed]
  158. Sjöberg, V.; Hollén, E.; Pietz, G.; Magnusson, K.E.; Fälth-Magnusson, K.; Sundström, M.; Holmgren Peterson, K.; Sandström, O.; Hernell, O.; Hammarström, S.; et al. Noncontaminated dietary oats may hamper normalization of the intestinal immune status in childhood celiac disease. Clin. Transl. Gastroenterol. 2014, 5, e58. [Google Scholar] [CrossRef] [PubMed]
  159. Valido, E.; Stoyanov, J.; Bertolo, A.; Hertig-Godeschalk, A.; Zeh, R.M.; Flueck, J.L.; Minder, B.; Stojic, S.; Metzger, B.; Bussler, W.; et al. Systematic Review of the Effects of Oat Intake on Gastrointestinal Health. J. Nutr. 2021, 151, 3075–3090. [Google Scholar] [CrossRef] [PubMed]
  160. Kemppainen, T.A.; Heikkinen, M.T.; Ristikankare, M.K.; Kosma, V.-M.; Sontag-Strohm, T.S.; Brinck, O.; Salovaara, H.O.; Julkunen, R.J. Unkilned and large amounts of oats in the coeliac disease diet: A randomized, controlled study. Scand. J. Gastroenterol. 2008, 43, 1094–1101. [Google Scholar] [CrossRef]
Table 1. Chemical composition of oat and other cereal grains (d. m.) [12].
Table 1. Chemical composition of oat and other cereal grains (d. m.) [12].
Cereal Species and ProductProteinLipidsCarbohydratesDietary Fiber
Whole oat grain with husk (range)7.4–16.22.2–9.253–6620–38
Whole oat grain without husk (range)10.5–24.53.1–1562–757.8–12.2
Naked oat (range)14–19.58.3–11.469–728.6–12.1
Wheat (average)13.52.367.712.1
Rye (average)10.22.063.916.1
Whole barley grain without husk (average)12.02.465.915.4
Corn (average)10.54.371.99.4
Table 2. The content of exogenous amino acids in various cereal grains (g·kg−1 protein) [12].
Table 2. The content of exogenous amino acids in various cereal grains (g·kg−1 protein) [12].
Exogenous Amino AcidsWheatRyeTriticaleBarleyOatFAO/WHO Standard
Lysine (Lys)263834324255
Methionine (Met)171717172517
Tryptophan (Trp)131611121910
Valine (Val)465342545350
Isoleucine (Ile)343532353940
Leucine (Leu)697577727470
Threonine (Thr)263231293340
Phenylalanine (Phe)435250515326
Sum274318294302338308
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Leszczyńska, D.; Wirkijowska, A.; Gasiński, A.; Średnicka-Tober, D.; Trafiałek, J.; Kazimierczak, R. Oat and Oat Processed Products—Technology, Composition, Nutritional Value, and Health. Appl. Sci. 2023, 13, 11267. https://doi.org/10.3390/app132011267

AMA Style

Leszczyńska D, Wirkijowska A, Gasiński A, Średnicka-Tober D, Trafiałek J, Kazimierczak R. Oat and Oat Processed Products—Technology, Composition, Nutritional Value, and Health. Applied Sciences. 2023; 13(20):11267. https://doi.org/10.3390/app132011267

Chicago/Turabian Style

Leszczyńska, Danuta, Anna Wirkijowska, Alan Gasiński, Dominika Średnicka-Tober, Joanna Trafiałek, and Renata Kazimierczak. 2023. "Oat and Oat Processed Products—Technology, Composition, Nutritional Value, and Health" Applied Sciences 13, no. 20: 11267. https://doi.org/10.3390/app132011267

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop