Next Article in Journal
RG-I Domain Matters to the In Vitro Fermentation Characteristics of Pectic Polysaccharides Recycled from Citrus Canning Processing Water
Previous Article in Journal
Effect of Mixing Time on Properties of Whole Wheat Flour-Based Cookie Doughs and Cookies
Previous Article in Special Issue
Sweet Chestnut (Castanea sativa Mill.) Nutritional and Phenolic Composition Interactions with Chestnut Flavor Physiology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Composition of Nuts and Their Potential Health Benefits—An Overview

1
CITAB, Centre for the Research and Technology of Agro-Environmental and Biological Sciences, Inov4Agro, Institute for Innovation, Capacity Building and Sustainability of Agri-Food Production, University of Trás-of-Montes and Alto Douro, Quinta de Prados, 5000-801 Vila Real, Portugal
2
CQ-VR, Chemistry Research Centre—Vila Real, University of Trás-os-Montes and Alto Douro, Quinta de Prados, 5000-801 Vila Real, Portugal
*
Author to whom correspondence should be addressed.
Foods 2023, 12(5), 942; https://doi.org/10.3390/foods12050942
Submission received: 17 November 2022 / Revised: 15 February 2023 / Accepted: 16 February 2023 / Published: 23 February 2023

Abstract

:
The possibility that nut intake may defend human health is an interesting point of view and has been investigated worldwide. Consequently, nuts are commonly promoted as healthy. In recent decades, the number of investigations proposing a correlation between nut consumption and a decrease in the risk of key chronic diseases has continued to increase. Nuts are a source of intake of fiber, and dietary fiber is associated with a reduced occurrence of obesity and cardiovascular diseases. Nuts likewise provide minerals and vitamins to the diet and supply phytochemicals that function as antioxidant, anti-inflammatory, and phytoestrogens agents and other protective mechanisms. Therefore, the main goal of this overview is to summarize current information and to describe the utmost new investigation concerning the health benefits of certain nuts.

1. Introduction

Currently, consumers are concerned about making a diversified and well-balanced diet. Therefore, the inclusion of nuts in the diet has undergone significant increases due to a growing recognition of their unique nutritional value, distinctive taste, flavor, nutraceutical properties, and healthy bioactive compounds, including high-quality proteins, fibers, minerals, tocopherols, phytosterols, and phenolic compounds [1]. Nuts are usually described as dry fruits with an edible seed and a hard shell, with cashews (Anacardium occidentale), walnuts (Juglans regia), almonds (Prunus dulcis), chestnuts (Castanea sativa), pistachios (Pistacia vera), and hazelnuts (Corylus avellana) as the ones with higher production worldwide [2]. There is the recognition that nuts are a good source of many nutrients, including monounsaturated and polyunsaturated fatty acid profile, vitamins E and K, selected minerals such as magnesium, copper, potassium, and selenium, dietary fibers, carotenoids, and phytosterols with potential antioxidant action [3]. In addition, the ease of transport due to their size makes them even more recommended to be consumed in all situations. In addition, the ingesting of nuts is often related to reducing risk factors for chronic diseases, due to the fatty acid profiles, squalene, fibers, vegetable proteins, minerals, vitamins, carotenoids, and phytosterols with potential antioxidant action [4]. Curiously, in all nuts, most of the antioxidants are located in the pellicle, as shown for almonds [5,6] and peanuts [7], and they are lost when the skin is removed [8]. In addition, in pistachios, most of the antioxidants are destroyed when the hard shells are cracked [9]. This review article aims to synthesize the current state of knowledge on the nutritional composition and health outcomes of some selected nuts.

2. Nuts

2.1. Proteins

Nuts are a rich source of proteins and essential amino acids as indicated by the USDA National Nutrient Database for Standard Reference [10] and as presented in Table 1. The major sources of proteins are peanuts, almonds, and pistachios, while chestnuts are the poorest in proteins. Chung et al. [11] reported higher protein content for some of these nuts, which can be ascribed to different geographic regions. The protein content also varied within the same nut species, denoting a significant effect of cultivar [12,13,14,15]. Other factors, such as the harvest year, post-harvest storage, and even the processing method [16] can affect the content of proteins in nuts. For example, Dodevbka et al. [17] reported differences between raw, boiled, and roasted nut samples from Serbia. The seed storage proteins are the main type of proteins present in nuts and are responsible for nut allergies [18]. Except for chestnuts, the other nine nuts referred to in Table 1 are the most common nuts capable of triggering adverse allergic reactions in some people. The proteins involved in nut allergy belong to different families, especially 2S albumins, globulins (legumins and vicilins), non-specific lipid transfer proteins (nsLTP), plant pathogenesis-related proteins (PR-10), profilins and oleosins [18]. Amandin, a legumin-type protein, is the most abundant protein in almonds, while the PR-10 Cor a 1 is the main allergen in hazelnuts, and the legumin Ju r 4 is the prevalent allergen in walnuts. The 2S albumins Car I 1 and Pis v 1 are dominant in pecan nuts and pistachios, respectively. The vicilins Ana o 1, Ara h 1, and Mac Ii1 are the most common allergens in cashew nuts, peanuts, and macadamia nuts, respectively [18].
Regarding the amino acid profile of each nut (Table 1), there is a considerable variation in the content of essential and non-essential amino acids. The nut protein composition is dominated by hydrophobic amino acids, followed by acidic, basic, and hydrophilic amino acids [19]. Among the non-essential amino acids, glutamic acid is the most important, ranging from 0.02 g/100 g in chestnuts to 6.21 g/100 g in almonds. The second major non-essential amino acid is arginine ranging from 0.12 g/100 g in chestnuts to 3.08 g/100 g in peanuts, followed by aspartic acid that ranges between 0.03 g/100 g in chestnuts and 3.15 g/100 g in peanuts. Leucine is the most essential amino acid, followed by phenylalanine and valine. Chestnuts present the lowest values of these essential amino acids (0.10, 0.07, and 0.09 g/100 g for leucine, phenylalanine and valine, respectively), while peanuts are the richest source of leucine and phenylalanine, and pistachios are the richest source of valine. Although the amino acid profile can differ significantly with variety and location, studies with 23 hazelnuts in northeast China [20] revealed the dominance of the same non-essential and essential amino acids (glutamic acid, arginine, aspartic acid, and leucine) described in Table 1 for hazelnut. The composition and dominance of essential and non-essential amino acids can influence several attributes of nuts, including the taste, aroma, or color being used, for example, for the characterization of almond cultivars [21].
The essential amino acid contents and their digestibility determine the nutritional value of a food protein. Although nut proteins are often recognized as incomplete proteins (i.e., do not contain all essential amino acids) when compared to animal proteins, their consumption is strongly associated with cardiovascular health [22]. Moreover, the presence of large quantities of arginine in all tree nuts has positive effects on immune response and inflammation, and cardiovascular function, including its key role in reducing the risk of cardiovascular disease and reproductive performance [23]. The health benefits of nut consumption can be enhanced by combining different protein sources to provide adequate levels of all essential amino acids.

2.2. Vitamins

Vitamins are essential for a balanced and healthy diet. Nuts contain fat-soluble vitamins (ascorbic acid, B1, B2, B3, B6) and antioxidants such as α-tocopherol (vitamin E), promoting better health, playing an important role against the aging process, improving brain function, and helping consumers to have healthy skin [24,25]. According to studies carried out by several researchers, the existence of vitamin C (ascorbic acid) is an important antioxidant for human colon cells [26,27]. The nut’s nutritional value depends on its chemical composition, and this is the result of the interaction of the cultivar (genotype), meteorological factors such as temperature and radiation, and production practices [28,29,30]. As Table 2 shows, walnuts, almonds, pine nuts, and hazelnuts are especially rich in vitamin E. Almonds, cashews, pistachios, walnuts, and peanuts are abundant sources of B vitamins. The concentration of folic acid was higher in pistachios and chestnuts. It is also the chestnuts that reveal the highest amount of vitamin C.

2.3. Minerals

Nuts are also rich sources of minerals such as magnesium and potassium (Table 3). In recent years, increased consumption of nuts has been considered good for human health to increase the intake of certain minerals, and they are considered a heart-healthy snacks when eaten in moderation [34]. Nuts are an important food source of minerals such as copper and magnesium. These two minerals may be protective against coronary heart disease. Nuts are also fairly high in potassium, particularly pistachio and cashew nuts (Table 1). Most nuts have a decent amount of zinc and iron, but pine nuts, cashews, and almonds stand above the rest. In contrast, nuts do not have a high content of calcium, still, some nuts such as almonds are better in terms of calcium content.

2.4. Fiber

Fiber is a health-promoting nut ingredient. The intake of dietary fiber is inversely related to obesity, type two diabetes, cancer, and cardiovascular disease according to epidemiological and clinical studies [38]. Among nuts, almonds present the highest content of fiber (Table 4), with a clear effect of genotype influencing its amount recorded [39]. Some works have highlighted the influence of genotype on the fiber content found in almonds [40], ranging from 6.88% to 9.74% in blanched almonds [41], showing that almond skin is also responsible for the fiber content of this nut, as it is composed of around 60% of fiber [42]. However, not all available data follow the same trend, with similar values of fiber recorded for different cultivars [43,44,45]. Cashews have the lowest fiber content among the referred nuts, with recent works pointing at values always around 3% to 4% [46], with no apparent significant effect of the cultivar on its content, although comprehensive studies are lacking for this specific nut. Chestnuts are considered to be a good source of dietary fiber [47], with similar values to those of cashews. The content of fiber in chestnuts has been the subject of studies that cannot find a trend on the factors behind their variation. Some authors point out the clear effect of cultivar on fiber content [43,44,48,49] or area of production [50] or year [43,44]. However, other works clearly state the similar content of fiber, regardless of cultivar [51,52,53]. The hazelnut fiber content is usually referred to as ranging from 6.5 g/100 g to 9.7 g/100 g [54]. Researchers have found higher amounts of fiber in some cultivars, such as the Turkish tombul hazelnut (12.9 g/100 g) [54], or other cultivars, with fiber values ranging from 9.8 g/100 g to 13.2 g/100 g (dry weight basis), being lowest in Tonda di Giffoni and highest in Campanica [55]. This also shows the variation of fiber content among cultivars, also recorded in the comparison of sixteen hazelnut cultivars [56]. For pistachio, the available works dealing with fiber content are few. However, early data indicate a content of 1.1–2.0% [57,58] although more recent works show considerably higher values. Dreher [59], Bulló et al. [60], and Terzo et al. [61] refer to values of fiber as around 10%, with Rabadán et al. [39] suggesting that the major factor between variations is the crop year and related weather conditions. Finally, walnut presents an intermediate amount of fiber when compared to other nuts. Although the majority of available works indicate values ranging from 4% to 6% [59,62,63,64,65], some authors have found considerably different amounts of fiber, namely Özcan [66], which indicates 1.8%, and Özcan et al. [67] that reports values between 3.8% and 3.9%. Again, the major factor affecting the fiber content of walnut is the genotype, with a slight effect also found to be caused by the crop year and related weather conditions [39].

2.5. Lipids and Fatty Acids

Nuts are rich in several nutrients, although with great differences between them and minor but sometimes still significant variations within cultivars. Lipid content and fatty acid profile are two of the parameters that can change considerably when discussing nut composition (Table 4 and Table 5). Besides these great variations between nut species, changes in lipid content and profile can also occur due to several other factors, with genotype as one of the most important that influences nut composition. Recent works show that genotype and the environment are key factors behind changes in several compositional parameters of some nuts, namely fat content [39]. There are some very good examples in the available literature, and to illustrate this fact, we will refer only to some for each nut.
For almonds, Summo et al. [69], working with samples from a germplasm collection under the same growing condition, recorded variations of lipid content, depending on the cultivar, from 42.4% to 56.2% (fresh weight). Barreca et al. [70] also reported a significant cultivar effect on the content of lipids in almonds. Almonds are also known for their interesting fatty acid profile, which is mainly composed of monounsaturated (MUFA) (60%) and polyunsaturated (PUFA) (30%) fatty acids, with a predominance of oleic, linoleic, palmitic, or stearic acids [43,44,71,72]. The work of Summo et al. [69] also shows the effect of the genotype on the fatty acid profile. Although major fatty acids are the same across the studied cultivars, changes can be observed in the individual amount of each fatty acid, as well as for the sum of unsaturated (mono- or polyunsaturated) and saturated (SFA) fractions. For cashew nuts, recent studies show great variability in fat content and associated fatty acid profiles when comparing different production regions. The work of Rico et al. [46], analyzing 11 cashew origins, shows that fat content can vary from 45.05 g/100 g in Vietnamese samples to 50.40 g/100 g in samples from Kenya. In the fatty acid profiles, oleic, linoleic, and palmitic acids are the three major ones. Although monounsaturated fatty acids represent the major fraction in all samples, followed by saturated fatty acids, at least in one sample, the second most important fraction is polyunsaturated fatty acids.
Chestnuts are featured with low-fat content and compared to other nuts, such as hazelnut, macadamia, pecan, or almond, chestnuts, exhibit the lowest fat content [73]. However, in this minor chestnut fraction, fat-soluble bioactive compounds, such as tocols and phytosterols, are present in higher quantities when compared to fat-rich nuts. They contain a high quantity of essential fatty acids (those that must be provided by food intake, as they are not synthesized in the body but are necessary for health) [74,75,76], either saturated or unsaturated, linked to several processes involved in health and chronic diseases [77]. Among them, the most important unsaturated fatty acids are linoleic and linolenic acids [75,78]. Fat content and fatty acid profiles can, as for other nuts, change significantly among cultivars. A thorough study of 17 chestnut cultivars produced in Portugal shows significant variations ranging in fat content from 1.67% to 3.50% [76]. Chestnut fat is primarily composed of three fatty acids, namely linoleic, oleic, and palmitic acids, with a predominance of polyunsaturated fatty acids. However, when comparing samples, significant variations of these fractions can be seen, with some presenting almost the same amount of mono- and polyunsaturated fatty acids. Similarly, the amount of saturated fatty acid also recorded significant variations across cultivars.
Among nuts, hazelnut presents one of the highest contents of fat, above 60% (Table 5), with some authors indicating the amount of fat above 70%, depending on the cultivar [79] or even on the canopy position of the fruits [80]. The fat present in hazelnuts is mainly composed of MUFA, representing around 80% of the total fatty acid content, and oleic acid is the major individual monounsaturated fatty acid [81,82,83]. Polyunsaturated fatty acids represent the second major fraction in hazelnut fat, almost exclusively due to the content of linoleic acid [84,85]. However, some works have found that SFA can represent the second major group of fatty acids [66,69], influenced by the higher content of palmitic acid.
Like most other nuts, pistachio is rich in fat, the available works indicating values around 50% [86,87,88,89], although some cultivars can have increased fat content, reaching values as high as 74.15% [90]. Following the trend of other nuts, pistachio fat is rich in unsaturated fatty acids, namely MUFA. This fraction is mainly composed of oleic acid, with a contribution from palmitoleic acid, while the second most important fraction, PUFA, is mainly composed of linoleic acid [83,89,91]. Regarding SFA, the minor fatty acid fraction is made almost entirely of palmitic acid [82,86].
The fat content of walnut is very high, with average values that can be surpassed only by hazelnuts [82]. Although the fat content is in the 60% range, considerable variations have been observed when comparing cultivars. Values varied between 49% [92] and 82% [93]. However, as referred before, most of the works show values of fat around 60%, with some variations associated with the studied cultivar [94,95]. Walnut fat is mostly composed of unsaturated fatty acid, namely PUFA, while MUFA is the second most important type of fatty acid [96,97,98]. Linoleic and linolenic acids are the ones responsible for the high amount of PUFA, with oleic as the major MUFA. Regarding SFA content, palmitic and stearic are the ones present in higher amounts [92,94,95] (Table 5).

2.6. Phenolic Compounds

Like in numerous other crops, phenolics are present in nuts. Many studies are reporting the beneficial effects of nut consumption on human health, including cardioprotective, neuroprotective, antidiabetic, anti-inflammatory, and antioxidant properties [99,100,101,102,103]. Studies have shown that the consumption of nuts improves flood lipoprotein profile [104,105] and gut microbiota [106]. These health effects are mainly due to the presence of several type of compounds, including phenolics, as reported by Lamuel-Raventos and Onge [102]. Each nut species presents its typical phenolic profile and content. For example, Liu et al. [107] found a high content of phenolics, such as vanillic acid, catechin, naringin, quercetin, and ellagic acid, in chestnuts, while Smeriglio et al. [108], in almonds, found a high content of phenolics, such as quercetin, kaempferol, and isorhamnetin. Instead, Taş and Gökmen [109] reported high levels of procyanidins A and B, trimers and tetramers, and prodelphinidin in peanuts. Table 6 shows several examples of phenolics found in the most common species of edible nuts. Similar to other crops, the variation in both profile and content of phenolic of nuts is highly related to genotype, cultural practices, climate conditions, fruit ripeness stage, storage, and post-harvest settings [110,111,112,113,114,115]. In addition, differences in the methods used to extract and quantify phenolic compounds (e.g., microwave-assisted extraction—MAE; supercritical CO2 extraction—SC-CO2; enzyme-assisted extraction—EAE; pressurized liquid extraction—PLE) by researchers may interfere with the number of phenolic compounds identified. However, based on the literature, it is possible to find a more or less common pattern.
The most abundant phenolics in almonds are catechin, epicatechin, protocatechuic acid, ferulic acid, kaempferol, and isorhamnetin [108,117]; in chestnuts are gallic acid, vanillic acid, syringic acid, catechin, and ellagic acid [119]; while in hazelnuts, the preponderance is for the catechin, epicatechin gallate, and gallic acid [121,122]; in peanuts, p-hydroxybenzoic acid, p-coumaric acid, ferulic acid, and epicatechin dominate [124]; in pistachios, gallic acid, syringic acid, catechin, and epicatechin [126]; while pecans and walnuts have in common high contents of chlorogenic, caffeic, p-coumaric, ferulic, ellagic and syringic acids [118]. In general, all nuts have in common the presence of high amounts of phenolic acids and flavonoids. The anthocyanins are present in vestigial amounts and are therefore not considered.
All these compounds are highly important because they have been associated with important beneficial effects on human health, as reported in the review of Lamuel-Raventos and Onge [102] and De Souza et al. [103]. Consumer perception of their beneficial effects has increased the intake of nuts. Different important findings from researchers have also contributed to the increment of such products in the human diet. For example, Brown et al. [130] found that higher nut consumption was associated with a reduced prevalence of high cholesterol and blood pressure, diabetes, and gallstones, due to the richness of phenolic compounds. In addition, Musarra-Pizzo et al. [131] tested a mix of phenolics present in natural almond skin and found that epicatechin and catechin were able to stop the growth of Staphylococcus aureus, suggesting that extracts from almond skins can be used to develop novel products for topical use. Neuroprotective effects against Alzheimer’s disease were found in almonds, hazelnuts, and walnuts due to their richness in tocopherols and phenolics [132].

2.7. Aroma and Flavor Compounds

The aroma compound profile of nuts is dependent on geographical origin and thermal processing and the presence of microorganisms. In almonds, several studies indicate aldehydes as the major volatiles, namely benzaldehyde [133,134,135] with a characteristic bitter-almond flavor, although this compound might not be found in several cultivars [136,137]. Besides terpenoids and substances derived from amino acids, volatiles are usually present as a result of the oxidation of fatty acids [138].
Processing causes several modifications, either in the number of compounds, but also in the chemical classes present [136,137]. In the work of Elmore et al. [139], they verified that walnuts from China and Ukraine contained high levels of lipid-derived volatiles from the linoleic acid breakdown (hexanal, pentanal, 1-hexanol, and 1-pentanol) and α-linolenic acid breakdown (1-penten-3-ol), whereas Chilean walnuts contained high levels of alkylbenzenes. Pyrazines are the major group of aromatic compounds in peanuts. They are formed by the thermally induced the Maillard reaction. The same applies to other nuts, such as pistachio and hazelnut. It is the roasting process that makes the fruit commercially viable and valuable, improving the nut’s sales and sensory characteristics [86]. Two pyrazines represent peanut flavor: 2,5-dimethyl pyrazine (with a characteristic nutty aroma) and 2-methoxy-5-methyl pyrazine (roasted nutty aroma) (Figure 1).
In hazelnuts, the results from Kiefl and Schieberle [140] showed that the aroma-active compounds 2-acetyl-1-pyrroline, 2-propionyl-1-pyrroline, 5-methyl-(E)-2-hepten-4-one (fibertone), 2,3-diethyl-5-methyl pyrazine, 3,5-dimethyl-2-ethyl pyrazine, and 2-furfurylthiol are appropriate odorant indicators to distinguish the several nut aromas. Specifically, the roasted or nutty aroma of roasted hazelnuts was developed if both 5-methyl-(E)-2-hepten-4-one and 3-methyl-4-heptanone were higher than 450 μg/kg, whereas the sum of the two 2-acyl-1-pyrrolines and two pyrazines should not exceed 400 μg/kg to avoid an over-roasted odor. A favored aroma can be obtained for each cultivar if specific temperatures, roasting techniques, and roasting times can be applied.
One major quality concern related to nuts is the development of off-flavors due to the formation of oxidative degradation products [141,142]. Various volatiles are involved in off-flavor; 1-Pentanol, 1-hexanol, and hexanal are the most important volatiles involved in off-flavor, and their presence at the highest levels is a synonym of nut degradation.

3. Impact of Nuts Processing on Nutrients and Phytochemicals

The phytochemicals in tree nuts have been linked to various health benefits, but processing steps can affect their bioavailability. Nuts can be processed in various ways to create different final products. For example, nuts that are consumed are often dehulled, peeled, blanched, and roasted [143,144,145]. Roasting is a common processing method used to preserve the quality and storability of nuts. It improves the flavor, aroma, color, texture, and appearance of the nuts through non-enzymatic reactions, such as Maillard browning. Roasting also inactivates enzymes that accelerate nutrient deterioration, remove microorganisms and food contaminants, and reduce degradative reactions such as lipid oxidation and rancidity, which are major factors that limit the shelf life of nuts. Additionally, the roasting process alters the microstructure and chemical composition of nuts, resulting in changes such as moisture reduction, modifications to lipids, changes in color, and the development of unique roasted flavors through the Maillard reaction [146,147]. Thus, the roasting process improves the nuts’ sensory characteristics such as flavor, color, taste, texture, appearance, and crispiness [145]. This improves the overall sensory characteristics of the nuts, making them more appealing to consume. The antioxidant activity, nutritional content, and total phenolic compounds in nuts may decrease after blanching and peeling, but roasting can improve these factors by releasing bound phenolic compounds and forming Maillard reaction products such as melanins [109,147]. However, the research on the effect of roasting on the phenolic compounds in nuts is limited. Based on available studies, the impact of roasting on the phenolic compounds in nuts can vary depending on the roasting temperature and duration. Some studies indicate that lower temperatures or shorter heating times may increase phenolic compounds, but higher temperatures or longer heating periods may decrease phenolic compounds [148]. For example, in hazelnuts, the content of flavan-3-ols (catechin and epicatechin) decreases significantly when roasted, with significant differences observed between raw nuts with skin and roasted nuts without skin. Thermal treatment also negatively impacts the content of procyanidin dimers and trimers in hazelnuts. Studies have shown that polyphenols in hazelnuts are mostly present in the skin and that roasting reduces the levels of phenolic compounds in most nuts, not only because of the removal of the skin but also due to the chemical degradation of many phenolic compounds [149]. These compounds are highly unstable and may be lost during processing, particularly when heat treatment is involved. Roasting can also alter the levels of antioxidants in the nuts, as the level of individual phenolics is higher in whole unroasted nuts [150] and alters the protein profile and allergenic properties [151]. Previous studies suggest that roasting enhances the allergenicity of roasted peanuts compared to raw peanuts [152], but the same was not observed in almonds [153]. The antioxidant activity of raw and roasted nuts depends on the type of nut and the roasting conditions. According to Schlörmann et al. [154], roasting can lead to a decrease in antioxidant activity in some nuts (hazelnut and walnut), but in others (almond and pistachio), the activity remains stable or is slightly enhanced. This decrease in activity is due to the loss of polyphenols due to thermal treatment, but the formation of antioxidant-active compounds due to Maillard reactions can counter this effect. The impact of roasting on bioavailability is still uncertain and requires further research. It is also important to evaluate the necessity of thermal processing by proving that nutritional and other properties are of great value, with antinutrients considerably decreased [155]. Thermal processing significantly reduces the protein, ash, and fiber content. The decrease in protein content may be caused by high-temperature denaturation and/or solubilization [156]. Additionally, the precipitation of mineral components leads to a decrease in ash content. In terms of carbohydrate content, roasting can increase it from 4.17% to 5.5%. This may be due to the hydrolysis of carbohydrates and to the reduction of other compounds in nuts due to thermal processing, making them easier to capture [109].
Concerning the effect of hot water blanching on protein composition, the results also depend on species and conditions of thermal processing [18]. Tian et al. [157] demonstrated that subjecting peanuts to 100 °C for 20 min reduced their allergenicity, due to the denaturation of allergenic proteins and to the transition of low molecular weight to the boiling water [158]. In turn, boiling almonds for 10 min [153] or cashews and pistachios for 60 min did not affect their properties [159].

4. Nut Consumer Perceptions of Health Benefits

Plant science research has been primarily focused on increasing production, with health benefits as a minor concern. The food industry is currently adapting its market trends to accommodate sustainability values, especially those related to health benefits, as they are increasingly researched by consumers [160], based, on nuts, on the phytonutrients present in these foods [161]. The current use of phytonutrients by food producers and the knowledge of their effect on the prevention of chronic disease points out the need for a careful look at crop production strategies (fertilization, season, soil fertility, and irrigation) affecting the quantitative and qualitative profiles of these compounds, but also to post-harvest techniques (processing or packaging) that can modify phytonutrients [161]. There is mounting evidence of the potential health benefits of a nut-rich diet. The ingestion of phytochemicals from nuts and their positive influence on several diseases (cancer, heart disease, stroke, hypertension, birth defects, cataracts, diabetes, diverticulosis, and obesity) are established [68,162,163,164]. There are many phytochemicals present in nuts that can be responsible for their health-promoting activities. Of those, one must refer to the vitamins, carotenoids, phenolic acids, or flavonoids, and their role in the prevention of certain cancers and cardiovascular diseases, but also to phytoestrogens, organosulfur compounds, fiber, or isothiocyanates (reviewed by several authors [68,165,166]). Nuts have been traditionally looked at as a high-fat and high-calorie food that should be consumed in moderation, which may be part of the reason why their intake is still below the recommended amount [167,168,169]. Although the link between weight gain and nut intake has been disproven [170], the usual high cost of nuts is another barrier to the increase in daily intake by consumers. The intake of nuts has been linked to several benefits to health, including favorable plasma lipid profiles, reduced risk of coronary heart disease, certain types of cancer, stroke, atherosclerosis, type-2 diabetes, inflammation, and several other chronic diseases [68,171,172]. However, it appears that consumers are not fully aware of the potential benefits of the intake of nuts. Recent works have shown that consumers link nuts to the high content of fat and proteins and that they are healthy. Nevertheless, a large percentage of consumers are still not aware of the link between nuts and the effects on blood cholesterol, cardiovascular disease risk, obesity, cancer, or diabetes [173,174,175,176,177,178].
Recently, there has been a huge effort to emphasize the beneficial action for the health by changing consumers’ eating habits, leading them to increase the consumption of certain foods such as nuts. There is no doubt that an informed consumer makes better decisions when choosing certain foods. In the long term, a higher intake of nuts will lead to clear benefits in the health sector, but other sectors will also benefit, such as producers and sellers.
Major concerns of the food industry related to the production and commercialization of nuts are the effects of processing and storage on the quality of nuts. Both temperature and humidity after harvesting can influence the appearance, moisture content, texture, and sensory characteristics of nuts [179]. Specifically, higher post-harvest temperature conditions can reduce crispness, increase moisture content and change oiliness and sweetness, resulting in the development of rancidity [180]. According to Mexis et al. [181], the alteration of sensorial characteristics leads to the formation of unpleasant flavors in pistachios, almonds, peanuts, and walnuts, as a result of alterations in the oxidation rate caused by high storage temperatures. It is mentioned in this study that storage temperatures of 30, 36, and 40 °C showed that nuts are more rancid compared to those stored at 8, 10, 20, or 25 °C.
Another possible alternative to increase the shelf life of nuts is the use of suitable packaging to reduce the problems mentioned above. Food can be packaged properly using modified atmosphere packaging or vacuum packaging to control the oxidation reaction [182]. The packaging material will be an important aspect to take into account as it will influence the shelf life of the nuts, will affect the respiration and transpiration rates of the fruit, as well as, the development of microorganisms. Fernandes et al. [182], in their comparative studies of chestnut conservation packages, concluded that chestnut conservation through the use of a specific packaging can have a substantial impact on preserving the color and texture of the fruit, preventing loss of weight, microbial growth, and in maintaining the water content of the fruit.
Consumer demand for eco-friendly and sustainable product packaging has proven to be remarkably stable and robust in recent years, including willingness to pay more for eco-friendly packaging. Consumers also recognize the value of reuse. The refillable packaging is proving to be a versatile and valuable solution for consumer products. Therefore, a holistic view of these issues is a growing requirement for everyone involved, from production, conservation, and marketing of this type of food product. Food choice is one of the most frequent human decisions and is determined by a complex set of factors and interrelated determinants [183]. Although several models attempting to explain that process have been proposed, one of the most accepted is the Total Food Quality Model [184]. This model can be divided into three parameters: ‘search’, ‘experience’, and ‘credence’ attributes. The first two (search attributes, such as appearance or price, and experience characteristics, such as flavor or taste) are those more easily observed by consumers and can be straightforwardly experienced by them. For credence properties, such as health and nutritional benefits, the consumer cannot validate those claims [185].
This is even more important in the current society, where the available fast food supply is large and more easily responds to the fast-paced life of consumers, with nutritionally poor foods taking place of a healthier diet.

5. Conclusions and Final Remarks

Nuts are a good source of many bioactive compounds with recognized health benefits, such as tocopherols, vitamins, and phenolic compounds. However, acquiring knowledge about the variation of bioactive compounds during fruit development and the ripening stage is crucial. How global environmental change and innovative crop production technology affect tree physiology and thus yield and fruit quality is at the moment mostly unknown. The development of species-specific strategies that improve both fruit quality and nutritional properties without significantly affecting yield should be aimed at by future research studies. The selection of high-yielding nut species and cultivars well-adapted to the different growing regions and future climatic conditions, with improved fruit traits, are needed to produce fruits with excellent quality and high consumer acceptability.

Author Contributions

All authors have contributed equally to this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Funds FCT Portuguese Foundation for Science and Technology—Portugal and COMPETE under the projects UIDB/00616/2020 and UIDP/00616/2020 (CQ-VR) and UIDB/04033/2020 (CITAB).

Acknowledgments

The authors would like to thank the CITAB/Inov4Agro Center for the Research and Technology of Agro-Environmental and Biological Sciences/Institute for Innovation, Capacity Building, and Sustainability of Agri-Food Production and Chemistry Research Center—Vila Real (CQ-VR) for their financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alasalvar, C.; Salvadó, J.S.; Ros, E. Bioactives and health benefits of nuts and dried fruits. Food Chem. 2020, 314, 126192. [Google Scholar] [CrossRef]
  2. INC 2022. Annual Report 2020/2021. Available online: https://www.nutfruit.org/files/transparency/1621876452_ANNUAL_REPORT_2021_final.pdf (accessed on 18 September 2022).
  3. de Souza, R.G.M.; Schincaglia, R.M.; Pimentel, G.D.; Mota, J.F. Nuts and human health outcomes: A systematic review. Nutrients 2017, 9, 1311. [Google Scholar] [CrossRef] [Green Version]
  4. Maguire, L.S.; O’Sullivan, S.M.; Galvin, K.; O’Connor, T.P.; O’Brien, N.M. Fatty acid profile, tocopherol, squalene and phytosterol content of walnuts, almonds, peanuts, hazelnuts and the macadamia nut. Int. J. Food Sci. Nutr. 2004, 55, 171–178. [Google Scholar] [CrossRef]
  5. Chen, C.-Y.; Milbury, P.E.; Lapsley, K.; Blumberg, J.B. Flavonoids from almond skins are bioavailable and act synergistically with vitamins C and E to enhance hamster and human LDL resistance to oxidation. J. Nutr. 2005, 135, 1366–1373. [Google Scholar] [CrossRef] [Green Version]
  6. Milbury, P.E.; Chen, C.Y.; Dolnikowski, G.G.; Blumberg, J.B. Determination of flavonoids and phenolics and their distribution in almonds. J. Agric. Food Chem. 2006, 54, 5027–5033. [Google Scholar] [CrossRef]
  7. Lou, H.; Yuan, H.; Ma, B.; Ren, D.; Ji, M.; Oka, S. Polyphenols from peanut skins and their free radical-scavenging effects. Phytochemistry 2004, 65, 2391–2399. [Google Scholar] [CrossRef]
  8. Blomhoff, R.; Carlsen, M.H.; Frost Andersen, L.; Jacobs, D.R., Jr. Health benefits of nuts, potential role of antioxidants. Br. J. Nutr. 2006, 96, S52–S60. [Google Scholar] [CrossRef] [Green Version]
  9. Seeram, N.P.; Zhang, Y.; Henning, S.M.; Lee, R.; Niu, Y.; Lin, G.; Heber, D. Pistachio skin phenolics are destroyed by bleaching resulting in reduced antioxidative capacities. J. Agric. Food Chem. 2006, 54, 7036–7040. [Google Scholar] [CrossRef]
  10. USDA, Agricultural Research Service, Nutrient Data Laboratory. USDA National Nutrient Database for Standard Reference, Legacy. Version Current: April 2018. Available online: http://www.ars.usda.gov/nutrientdata (accessed on 18 September 2022).
  11. Chung, K.H.; Shin, K.O.; Hwang, H.J.; Choi, K.S. Chemical composition of nuts and seeds sold in Korea. Nutr. Res. Pract. 2013, 7, 82–88. [Google Scholar] [CrossRef] [Green Version]
  12. Kafkas, E.; Attar, S.H.; Gundesli, M.A.; Ozcan, A.; Ergun, M. Phenolic and fatty acid profile, and protein content of different walnut cultivars and genotypes (Juglans regia L.) grown in the USA. Int. J. Fruit Sci. 2020, 20, S1711–S1720. [Google Scholar] [CrossRef]
  13. Müller, A.K.; Helms, U.; Rohrer, C.; Möhler, M.; Hellwig, F.; Glei, M.; Schwerdtle, T.; Lorkowski, S.; Dawczynski, C. Nutrient composition of different hazelnut cultivars grown in Germany. Foods 2020, 9, 1596. [Google Scholar] [CrossRef]
  14. Nankya, R.; Mulumba, J.W.; Lwandasa, H.; Matovu, M.; Isabirye, B.; De Santis, P.; Jarvis, D.I. Diversity in nutrient content and consumer preferences of sensory attributes of peanut (Arachis hypogaea L.) varieties in Ugandan agroecosystems. Sustainability 2021, 13, 2658. [Google Scholar] [CrossRef]
  15. Ferrari, V.; Gil, G.; Heinzen, H.; Zoppolo, R.; Ibanez, F. Influence of cultivar on nutritional composition and nutraceutical potential of pecan growing in Uruguay. Front. Nut. 2022, 9, 868054. [Google Scholar] [CrossRef]
  16. Bolling, B.; Chen, C.; McKay, D.; Blumberg, J. Tree nut phytochemicals: Composition, antioxidant capacity, bioactivity, impact factors. A systematic review of almonds, Brazils, cashews, hazelnuts, macadamias, pecans, pine nuts, pistachios and walnuts. Nutr. Res. Rev. 2011, 24, 244–275. [Google Scholar] [CrossRef] [Green Version]
  17. Dodevska, M.; Kukic, M.J.; Sofrenic, I.; Tesevic, V.; Jankovic, M.; Djordjevic, B.; Ivanovic, N.D. Similarities and differences in the nutritional composition of nuts and seeds in Serbia. Front. Nutr. 2022, 9, 1003125. [Google Scholar] [CrossRef]
  18. Luparelli, A.; Losito, I.; De Angelis, E.; Pilolli, R.; Lambertini, F.; Monaci, L. Tree nuts and peanuts as a source of beneficial compounds and a threat for allergic consumers: Overview on methods for their detection in complex food products. Foods 2022, 11, 728. [Google Scholar] [CrossRef]
  19. Venkatachalam, M.; Sathe, S.K. Chemical composition of selected edible nut seeds. J. Agric. Food Chem. 2006, 54, 4705–4714. [Google Scholar] [CrossRef]
  20. Fan, L.; Ren, J.; Yang, Y.; Zhang, L. Comparative analysis on essential nutrient compositions of 23 wild Hazelnuts (Corylus heterophylla) grown in Northeast China. J. Food Qual. 2020, 2020, 9475961. [Google Scholar] [CrossRef] [Green Version]
  21. Serón, L.A.; Garrigós, E.S.; Berenguer, V.; Grané-Teruel, N. Characterisation of 19 almond cultivars on the basis of their free amino acids composition. Food Chem. 1998, 61, 455–459. [Google Scholar] [CrossRef]
  22. Hertzler, S.R.; Lieblein-Boff, J.C.; Weiler, M.; Allgeier, C. Plant proteins: Assessing their nutritional quality and effects on health and physical function. Nutrients 2020, 12, 3704. [Google Scholar] [CrossRef]
  23. Arya, S.S.; Salve, A.R.; Chauhan, S. Peanuts as functional food: A review. J Food Sci Technol. 2016, 53, 31–41. [Google Scholar] [CrossRef] [Green Version]
  24. Suna, S.; Avşar, B.; Koçer, S.; Çopur, Ö.U. Effects of different pretreatments on the physicochemical characteristics and quality criteria of chestnut (Castanea sativa Mill.) pickle: A new value-added product. J. Food Process. Preserv. 2021, 45, e15669. [Google Scholar] [CrossRef]
  25. Pinto, D.; Rodrigues, F.; Braga, N.; Santos, J.; Pimentel, F.B.; Palmeira-De-Oliveira, A.; Oliveira, M.B.P.P. The Castanea sativa bur as a new potential ingredient for nutraceutical and cosmetic outcomes: Preliminary studies. Food Funct. 2016, 8, 201–208. [Google Scholar] [CrossRef]
  26. Beton, K.; Brozek-Pluska, B. Vitamin C-Protective role in oxidative stress conditions induced in human normal colon cells by Label-Free Raman Spectroscopy and Imaging. Int. J. Mol. Sci. 2021, 22, 6928. [Google Scholar] [CrossRef]
  27. Choupina, A.B. Nutritional and health potential of European chestnut. Rev. Cienc. Agrar. 2019, 42, 801–807. [Google Scholar]
  28. Cevriye, M.E.R.T.; Ertürk, Ü. Chemical compositions and sugar profiles of consumed chestnut cultivars in the Marmara Region, Turkey. Not. Bot. Horti Agrobot. 2017, 45, 203–207. [Google Scholar] [CrossRef] [Green Version]
  29. Mota, M.; Pinto, T.; Vilela, A.; Marques, T.; Borges, A.; Caço, J.; Ferreira-Cardoso, J.; Raimundo, F.; Gomes-Laranjo, J. Irrigation positively affects the chestnut’s quality: The chemical composition, fruit size and sensory attributes. Sci. Hortic. 2018, 238, 177–186. [Google Scholar] [CrossRef]
  30. Santos, M.J.; Pinto, T.; Vilela, A. Sweet chestnut (Castanea sativa Mill.) nutritional and phenolic composition interactions with chestnut flavor physiology. Foods 2022, 11, 4052. [Google Scholar] [CrossRef]
  31. Açkurt, F.; Özdemir, M.; Biringen, G.; Löker, M. Effects of geographical origin and variety on vitamin and mineral composition of hazelnut (Corylus avellana L.) varieties cultivated in Turkey. Food Chem. 1999, 65, 309–313. [Google Scholar] [CrossRef]
  32. Sathe, S.; Monaghan, E.; Kshiesagar, H.; Venkatachalam, M. Chemical composition of edible nut seeds and its implications in human health. In Tree Nuts Composition, Phytochemicals and Health Effects; Alsalvar, C., Shahidi, F., Eds.; Taylor & Francis Group: Boca Raton, FL, USA, 2008; pp. 12–29. [Google Scholar] [CrossRef]
  33. Câmara, C.R.S.; Schlegel, V. A Review on the potential human health benefits of the black walnut: A comparison with the english walnuts and other tree nuts. Int. J. Food Prop. 2016, 19, 2175–2189. [Google Scholar] [CrossRef]
  34. Pradhan, C.; Peter, N.; Dileep, N. Nuts as dietary source of fatty acids and micro nutrients in human health. In Nuts and Nut Products in Human Health and Nutrition; Rao, V., Rao, L., Ahiduzzaman, M., Aminul, A.K.M.A., Eds.; IntechOpen: London, UK, 2021; ISBN 978-1-78985-511-1. [Google Scholar] [CrossRef]
  35. Alhassan, K.; Agbenorhevi, J.K.; Asibuo, J.Y.; Sampson, G.O. Proximate composition and functional properties of some new groundnut accessions. J. Food Secur. 2017, 5, 9–12. [Google Scholar] [CrossRef] [Green Version]
  36. Caglarirmak, N.; Batkan, A.C. Nutrients and biochemistry of nuts in different consumption types in Turkey. J. Food Proc. Preserv. 2005, 29, 407–423. [Google Scholar] [CrossRef]
  37. Ros, E. Health benefits of nut consumption. Nutrients 2010, 2, 652–682. [Google Scholar] [CrossRef] [Green Version]
  38. Lattimer, J.M.; Haub, M.D. Effects of dietary fibre and its components on metabolic health. Nutrients 2010, 2, 1266–1289. [Google Scholar] [CrossRef] [Green Version]
  39. Rabadán, A.; Álvarez-Ortí, M.; Pardo, J.E. A comparison of the effect of genotype and weather conditions on the nutritional composition of most important commercial nuts. Sci. Hortic. 2019, 244, 218–224. [Google Scholar] [CrossRef]
  40. Ruggeri, S.; Cappelloni, M.; Gambelli, L.; Nicoli, S.; Carnovale, E. Chemical composition and nutritive value of nuts grown in Italy. Ital. J. Food Sci. 1998, 3, 243–251. [Google Scholar]
  41. Romero, A.; Vargas, F.J.; Tous, J.; Ninot, A.; Miarnau, X. New almond varieties from IRTA’s breeding programme:(1) chemical composition. In Proceedings of the V International Symposium on Pistachios and Almonds, Sanliurfa, Turkey, 6–10 October 2009; Volume 912, pp. 477–484. [Google Scholar]
  42. Mandalari, G.; Tomaino, A.; Arcoraci, T.; Martorana, M.; Turco, V.L.; Cacciola, F.; Rich, G.; Bisignano, C.; Saija, A.; Dugo, P.; et al. Characterization of polyphenols, lipids and dietary fibre from almond skins (Amygdalus communis L.). J. Food Compos. Anal. 2010, 23, 166–174. [Google Scholar] [CrossRef]
  43. Barreira, J.C.; Casal, S.; Ferreira, I.C.; Peres, A.M.; Pereira, J.A.; Oliveira, M.B.P. Supervised chemical pattern recognition in almond (Prunus dulcis) portuguese PDO cultivars: PCA-and LDA-based triennial study. J. Agric. Food Chem. 2012, 60, 9697–9704. [Google Scholar] [CrossRef]
  44. Barreira, J.C.; Casal, S.; Ferreira, I.C.; Peres, A.M.; Pereira, J.A.; Oliveira, M.B.P. Chemical characterization of chestnut cultivars from three consecutive years: Chemometrics and contribution for authentication. Food Chem. Toxicol. 2012, 50, 2311–2317. [Google Scholar] [CrossRef] [Green Version]
  45. Yada, S.; Huang, G.; Lapsley, K. Natural variability in the nutrient composition of California-grown almonds. J. Food Comp. Anal. 2013, 30, 80–85. [Google Scholar] [CrossRef] [Green Version]
  46. Rico, R.; Bulló, M.; Salas-Salvadó, J. Nutritional composition of raw fresh cashew (Anacardium occidentale L.) kernels from different origin. Food Sci. Nutr. 2016, 4, 329–338. [Google Scholar] [CrossRef] [PubMed]
  47. Gonçalves, B.; Borges, O.; Costa, H.S.; Benett, R.; Santos, M.; Silva, A.P. Metabolite composition of chestnut (Castanea sativa Mill.) upon cooking: Proximate analysis, fibre, organic acids and phenolics. Food Chem. 2010, 122, 154–160. [Google Scholar] [CrossRef]
  48. Borges, O.; Gonçalves, B.; Carvalho, J.L.S.; Correia, P.; Silva, A.P. Nutritional quality of chestnut (Castanea sativa Mill.) cultivars from Portugal. Food Chem. 2008, 106, 976–984. [Google Scholar] [CrossRef]
  49. Barreira, J.C.; Casal, S.; Ferreira, I.C.; Oliveira, M.B.P.; Pereira, J.A. Nutritional, fatty acid and triacylglycerol profiles of Castanea sativa Mill. cultivars: A compositional and chemometric approach. J. Agric. Food Chem. 2009, 57, 2836–2842. [Google Scholar] [CrossRef]
  50. Pena-Méndez, E.M.; Hernández-Suárez, M.; Díaz-Romero, C.; Rodríguez-Rodríguez, E. Characterization of various chestnut cultivars by means of chemometrics approach. Food Chem. 2008, 107, 537–544. [Google Scholar] [CrossRef]
  51. Pereira-Lorenzo, S.; Ramos-Cabrer, A.M.; Díaz-Hernández, M.B.; Ciordia-Ara, M.; Ríos-Mesa, D. Chemical composition of chestnut cultivars from Spain. Sci. Hortic. 2006, 107, 306–314. [Google Scholar] [CrossRef] [Green Version]
  52. De Vasconcelos, M.C.; Bennett, R.N.; Rosa, E.A.S.; Ferreira-Cardoso, J.V. Metabolites composition of fresh kernels from three cultivars of Portuguese chestnut. In Proceedings of the II Iberian Congress on Chestnut, Vila Real, Portugal, 20–22 June 2007; Volume 784, pp. 43–48. [Google Scholar]
  53. Dimitri, G.; Mastrocola, D.; Sacchetti, G.; Neri, L. Chemical composition and functional properties of three sweet chestnut (Castanea sativa Mill.) ecotypes from Italy. In Proceedings of the IV International Chestnut Symposium, Beijing, China, 25–28 September 2008; Volume 844, pp. 41–46. [Google Scholar]
  54. Alasalvar, C.; Shahidi, F.; Liyanapathirana, C.M.; Ohshima, T. Turkish tombul hazelnut (Corylus avellana L.). 1. Compositional characteristics. J. Agric. Food Chem. 2003, 51, 3790–3796. [Google Scholar] [CrossRef]
  55. Savage, G.P.; McNeil, D.L. Chemical composition of hazelnuts (Corylus avellana L.) grown in New Zealand. Int. J. Food Sci. Nutr. 1998, 49, 199–203. [Google Scholar] [CrossRef]
  56. Solar, A.; Stampar, F. Characterisation of selected hazelnut cultivars: Phenology, growing and yielding capacity, market quality and nutraceutical value. J. Sci. Food Agric. 2011, 91, 1205–1212. [Google Scholar] [CrossRef]
  57. Kamangar, T.; Farsam, H. Composition of pistachio kernels of various Iranian origins. J. Food Sci. 1977, 42, 1135–1136. [Google Scholar] [CrossRef]
  58. Maskan, M.; Karataş, Ş. Storage stability of whole-split pistachio nuts (Pistachia vera L.) at various conditions. Food Chem. 1999, 66, 227–233. [Google Scholar] [CrossRef]
  59. Dreher, M.L. Pistachio nuts: Composition and potential health benefits. Nutr. Rev. 2012, 70, 234–240. [Google Scholar] [CrossRef]
  60. Bulló, M.; Juanola-Falgarona, M.; Hernández-Alonso, P.; Salas-Salvadó, J. Nutrition attributes and health effects of pistachio nuts. Br. J. Nutr. 2015, 113, S79–S93. [Google Scholar] [CrossRef] [Green Version]
  61. Terzo, S.; Baldassano, S.; Caldara, G.F.; Ferrantelli, V.; Lo Dico, G.; Mulè, F.; Amato, A. Health benefits of pistachios consumption. Nat. Prod. Res. 2019, 33, 715–726. [Google Scholar] [CrossRef]
  62. Ozkan, G.; Koyuncu, M.A. Physical and chemical composition of some walnut (Juglans regia L.) genotypes grown in Turkey. Grasas y Aceites 2005, 56, 141–146. [Google Scholar] [CrossRef]
  63. Savage, G.P. Chemical composition of walnuts (Juglans regia L.) grown in New Zealand. Plant Foods Hum. Nutr. 2001, 56, 75–82. [Google Scholar] [CrossRef]
  64. Gharibzahedi, S.M.T.; Mousavi, S.M.; Ghahderijani, M.; Dadashpour, A. Assessment of physical and chemical aspects of new Persian walnut cultivars to optimize process conditions. Int. J. Food Eng. 2012, 8, 1. [Google Scholar] [CrossRef]
  65. Gharibzahedi, S.M.T.; Mousavi, S.M.; Hamedi, M.; Khodaiyan, F. Determination and characterization of kernel biochemical composition and functional compounds of Persian walnut oil. J. Food Sci. Technol. 2014, 51, 34–42. [Google Scholar] [CrossRef] [Green Version]
  66. Özcan, M.M. Some nutritional characteristics of fruit and oil of walnut (Juglans regia L.) growing in Turkey. Iran. J. Chem. Chem. Eng. 2009, 28, 57–62. [Google Scholar] [CrossRef]
  67. Özcan, M.M.; İman, C.; Arslan, D. Physicochemical properties, fatty acid and mineral content of some walnuts (Juglans regia L.) types. Agric. Sci. 2010, 1, 62. [Google Scholar] [CrossRef] [Green Version]
  68. Amarowicz, R.; Gong, Y.; Pegg, R.B. Recent advances in our knowledge of the biological properties of nuts. In Wild Plants, Mushrooms and Nuts: Functional Food Properties and Applications; Ferreira, I.C.F.R., Morales, P., Barros, L., Eds.; John Wiley & Sons, Ltd.: London, UK, 2017; pp. 377–409. ISBN 9781118944653. [Google Scholar] [CrossRef]
  69. Summo, C.; Palasciano, M.; De Angelis, D.; Paradiso, V.M.; Caponio, F.; Pasqualone, A. Evaluation of the chemical and nutritional characteristics of almonds (Prunus dulcis (Mill). DA Webb) as influenced by harvest time and cultivar. J. Sci. Food Agric. 2018, 98, 5647–5655. [Google Scholar] [CrossRef] [Green Version]
  70. Barreca, D.; Nabavi, S.M.; Sureda, A.; Rasekhian, M.; Raciti, R.; Silva, A.S.; Annunziata, G.; Arnone, A.; Tenore, G.C.; Süntar, İ.; et al. Almonds (Prunus dulcis Mill. D. A. Webb): A source of nutrients and health-promoting compounds. Nutrients 2020, 12, 672. [Google Scholar] [CrossRef] [Green Version]
  71. Oliveira, I.; Meyer, A.S.; Afonso, S.; Aires, A.; Goufo, P.; Trindade, H.; Gonçalves, B. Phenolic and fatty acid profiles, α-tocopherol and sucrose contents, and antioxidant capacities of understudied Portuguese almond cultivars. J. Food Biochem. 2019, 43, e12887. [Google Scholar] [CrossRef]
  72. Oliveira, I.; Meyer, A.S.; Afonso, S.; Sequeira, A.; Vilela, A.; Goufo, P.; Trindade, H.; Gonçalves, B. Effects of different processing treatments on almond (Prunus dulcis) bioactive compounds, antioxidant activities, fatty acids, and sensorial characteristics. Plants 2020, 9, 1627. [Google Scholar] [CrossRef]
  73. Alasalvar, C.; Bolling, B.W. Review of nut phytochemicals, fat-soluble bioactives, antioxidant components and health effects. Br. J. Nutr. 2015, 113, S68–S78. [Google Scholar] [CrossRef]
  74. Kaur, N.; Chugh, V.; Gupta, A. Essential fatty acids as functional components of foods—A review. J. Food Sci. Technol. 2014, 51, 2289–2303. [Google Scholar] [CrossRef] [Green Version]
  75. Barreira, J.C.M.; Ferreira, I.C.F.R.; Oliveira, M.B.P.P. Bioactive compounds of Chestnut (Castanea sativa Mill.) BT. In Bioactive Compounds in Underutilized Fruits and Nuts; Murthy, H.N., Bapat, V.A., Eds.; Springer International Publishing: Cham, Switzerland, 2020; pp. 303–313. ISBN 978-3-030-30182-8. [Google Scholar]
  76. Borges, O.P.; Carvalho, J.S.; Correia, P.R.; Silva, A.P. Lipid and fatty acid profiles of Castanea sativa Mill. Chestnuts of 17 native Portuguese cultivars. J. Food Comp. Anal. 2007, 20, 80–89. [Google Scholar] [CrossRef]
  77. Benatti, P.; Peluso, G.; Nicolai, R.; Calvani, M. Polyunsaturated fatty acids: Biochemical, nutritional and epigenetic properties. J. Am. Coll. Nutr. 2004, 23, 281–302. [Google Scholar] [CrossRef]
  78. España, M.S.A.; Galdón, B.R.; Romero, C.D.; Rodríguez, E.R. Fatty acid profile in varieties of chestnut fruits from Tenerife (Spain). CyTA-J. Food 2011, 9, 77–81. [Google Scholar] [CrossRef] [Green Version]
  79. Lucchetti, S.; Ambra, R.; Pastore, G. Effects of peeling and/or toasting on the presence of tocopherols and phenolic compounds in four Italian hazelnut cultivars. Eur. Food Res. Technol. 2018, 244, 1057–1064. [Google Scholar] [CrossRef]
  80. Pannico, A.; Cirillo, C.; Giaccone, M.; Scognamiglio, P.; Romano, R.; Caporaso, N.; Sacchi, R.; Basile, B. Fruit position within the canopy affects kernel lipid composition of hazelnuts. J. Sci. Food Agric. 2017, 97, 4790–4799. [Google Scholar] [CrossRef] [Green Version]
  81. Oliveira, I.; Sousa, A.; Morais, J.S.; Ferreira, I.C.; Bento, A.; Estevinho, L.; Pereira, J.A. Chemical composition, and antioxidant and antimicrobial activities of three hazelnut (Corylus avellana L.) cultivars. Food Chem. Toxicol. 2008, 46, 1801–1807. [Google Scholar] [CrossRef]
  82. Kalogeropoulos, N.; Chiou, A.; Ioannou, M.S.; Karathanos, V.T. Nutritional evaluation and health promoting activities of nuts and seeds cultivated in Greece. Int. J. Food Sci. Nutr. 2013, 64, 757–767. [Google Scholar] [CrossRef]
  83. Freitas, J.B.; Naves, M.M.V. Chemical composition of nuts and edible seeds and their relation to nutrition and health. Rev. Nutr. 2010, 23, 269–279. [Google Scholar] [CrossRef]
  84. Köksal, A.İ.; Artik, N.; Şimşek, A.; Güneş, N. Nutrient composition of hazelnut (Corylus avellana L.) varieties cultivated in Turkey. Food Chem. 2006, 99, 509–515. [Google Scholar] [CrossRef]
  85. Turan, A. Effect of drying methods on fatty acid profile and oil oxidation of hazelnut oil during storage. Eur. Food Res. Technol. 2018, 244, 2181–2190. [Google Scholar] [CrossRef]
  86. Rodríguez-Bencomo, J.J.; Kelebek, H.; Sonmezdag, A.S.; Rodríguez-Alcalá, L.M.; Fontecha, J.; Selli, S. Characterization of the aroma-active, phenolic, and lipid profiles of the pistachio (Pistacia vera l.) nut as affected by the single and double roasting process. J. Agric. Food Chem. 2015, 63, 7830–7839. [Google Scholar] [CrossRef] [Green Version]
  87. Roozban, M.R.; Mohamadi, N.; Vahdati, K. Fat content and fatty acid composition of four Iranian Pistachio (Pistacia vera L.) varieties grown in Iran. In Proceedings of the IV International Symposium on Pistachios and Almonds, Tehran, Iran, 22–25 May 2005; Volume 726, pp. 573–578. [Google Scholar]
  88. Harmankaya, M.; Özcan, M.M.; Juhaimi, F.A. Mineral contents and proximate composition of Pistacia vera kernels. Environ. Monit. Assess. 2014, 186, 4217–4221. [Google Scholar] [CrossRef]
  89. Rabadán, A.; Álvarez-Ortí, M.; Gómez, R.; de Miguel, C.; Pardo, J.E. Influence of genotype and crop year in the chemometrics of almond and pistachio oils. J. Sci. Food Agric. 2018, 98, 2402–2410. [Google Scholar] [CrossRef]
  90. Zribi, F.; Ghrab, M.; Mnafki, N.; Ayadi, M.; Ben Mimoun, M. Biochemical characterization of pistachio germplasm grown in Tunisia. In Proceedings of the V International Symposium on Pistachios and Almonds, Sanliurfa, Turkey, 6–10 October 2009; Volume 912, pp. 469–476. [Google Scholar]
  91. Catalán, L.; Alvarez-Ortí, M.; Pardo-Giménez, A.; Gomez, R.; Rabadan, A.; Pardo, J.E. Pistachio oil: A review on its chemical composition, extraction systems, and uses. Eur. J. Lipid Sci. 2017, 119, 1600126. [Google Scholar] [CrossRef]
  92. Muradoglu, F.; Oguz, H.I.; Yildiz, K. Some chemical composition of walnut (Juglans regia L.) selections from Eastern Turkey. Afr. J. Agric. Res. 2010, 5, 2379–2385. [Google Scholar]
  93. Beyhan, O.; Ozcan, A.; Ozcan, H.; Kafkas, E.; Kafkas, S.; Sutyemez, M.; Ercisli, S. Fat, fatty acids and tocopherol content of several walnut genotypes. Not. Bot. Horti Agrobot. Cluj Napoca 2017, 45, 437–441. [Google Scholar] [CrossRef] [Green Version]
  94. Pereira, J.A.; Oliveira, I.; Sousa, A.; Ferreira, I.C.; Bento, A.; Estevinho, L. Bioactive properties and chemical composition of six walnut (Juglans regia L.) cultivars. Food Chem. Toxicol. 2008, 46, 2103–2111. [Google Scholar] [CrossRef]
  95. Amaral, J.S.; Casal, S.; Pereira, J.A.; Seabra, R.M.; Oliveira, B.P. Determination of sterol and fatty acid compositions, oxidative stability, and nutritional value of six walnut (Juglans regia L.) cultivars grown in Portugal. J. Agric. Food Chem. 2003, 51, 7698–7702. [Google Scholar] [CrossRef] [Green Version]
  96. Ertürk, U.; Şisman, T.; Yerlikaya, C.; Ertürk, O.; Karadeniz, T. Chemical Composition and Nutritive Value of Selected Walnuts (Juglans regia L.) from Turkey. In Proceedings of the VII International Walnut Symposium, Taiyuan, China, 20–23 July 2013; Volume 1050, pp. 231–234. [Google Scholar]
  97. Zhai, M.Z.; Wang, D.; Tao, X.D.; Wang, Z.Y. Fatty acid compositions and tocopherol concentrations in the oils of 11 varieties of walnut (Juglans regia L.) grown at Xinjiang, China. J. Hortic. Sci. Biotechnol. 2015, 90, 715–718. [Google Scholar] [CrossRef]
  98. Hayes, D.; Angove, M.J.; Tucci, J.; Dennis, C. Walnuts (Juglans regia) chemical composition and research in human health. Crit Rev. Food Sci. Nutr. 2016, 56, 1231–1241. [Google Scholar] [CrossRef]
  99. Vinson, J.A.; Cai, Y. Nuts, especially walnuts, have both antioxidant quantity and efficacy and exhibit significant potential health benefits. Food Funct. 2012, 3, 134–140. [Google Scholar] [CrossRef] [Green Version]
  100. Pribis, P.; Bailey, R.N.; Russell, A.A.; Kilsby, M.A.; Hernandez, M.; Craig, W.J.; Grajales, T.; Shavlik, D.J.; Sabate, J. Effects of walnut consumption on cognitive performance in young adults. Br. J. Nutr. 2012, 107, 1393–1401. [Google Scholar] [CrossRef] [Green Version]
  101. Poulose, S.M.; Miller, M.G.; Shukitt-Hale, B. Role of walnuts in maintaining brain health with age. J. Nutr. 2014, 144, 561S–566S. [Google Scholar] [CrossRef] [Green Version]
  102. Lamuel-Raventos, R.M.; St. Onge, M.-P. Prebiotic nut compounds and human microbiota. Crit. Rev. Food Sci. Nutr. 2017, 57, 3154–3163. [Google Scholar] [CrossRef]
  103. Gervasi, T.; Barreca, D.; Laganà, G.; Mandalari, G. Health benefits related to tree nut consumption and their bioactive compounds. Int. J. Mol. Sci. 2021, 22, 5960. [Google Scholar] [CrossRef] [PubMed]
  104. Sabate, J.; Wien, M. Nuts, blood lipids and cardiovascular disease. Asia Pac. J. Clin. Nutr. 2010, 19, 131–136. [Google Scholar]
  105. Del Gobbo, L.C.; Falk, M.C.; Feldman, R.; Lewis, K.; Mozaffarian, D. Effects of tree nuts on blood lipids, apolipoproteins, and blood pressure: Systematic review, meta-analysis, and dose-response of 61 controlled intervention trials. Am. J. Clin. Nutr. 2015, 102, 1347–1356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Cardona, F.; Andrés-Lacueva, C.; Tulipani, S.; Tinahones, F.J.; Queipo-Ortuno, M.I. Benefits of polyphenols on gut microbiota and implications on human health. J. Nutr. Biochem. 2013, 24, 1415–1422. [Google Scholar] [CrossRef] [Green Version]
  107. Liu, F.; Chang, X.; Hu, X.; Brennan, C.S.; Guo, X. Effect of thermal processing on phenolic profiles and antioxidant activities in Castanea mollissima. Int. J. Food Sci. Technol. 2017, 52, 439–447. [Google Scholar] [CrossRef]
  108. Smeriglio, A.; Mandalari, G.; Bisigano, C.; Filocarmo, A.; Barreca, D.; Belloco, E.; Trombeta, D. Polyphenolic content and biological properties of Avola almond (Prunus dulcis Mill. D.A. Webb) skin and its industrial by products. Ind. Crops Prod. 2015, 83, 283–293. [Google Scholar] [CrossRef]
  109. Taş, G.N.; Gökmen, V. Phenolic compounds in natural and roasted nuts and their skins: A brief review. Curr. Opin. Food Sci. 2017, 14, 103–109. [Google Scholar] [CrossRef]
  110. Wojdyło, A.; Turkiewicz, I.P.; Tkacz, K.; Nowicka, P.; Bobak, Ł. Nuts as functional foods: Variation of nutritional and phytochemical profile and their in vitro bioactive properties. Food Chem. X 2022, 15, 100418. [Google Scholar] [CrossRef]
  111. Chandrasekara, N.; Shahidi, F. Effect of roasting on phenolic content and antioxidant activities of whole cashew nuts, kernels, and testa. J. Agric. Food Chem. 2011, 59, 5006–5014. [Google Scholar] [CrossRef]
  112. Dolatabadi, K.S.M.; Dehghan, G.; Hosseini, S.; Jahanban Esfahlan, A. Effect of five years storage on total phenolic content and antioxidant capacity of almond (Amygdalus communis L.) hull and shell from different genotypes. Avicenna J. Phytomed. 2015, 5, 26–33. [Google Scholar]
  113. Pycia, K.; Kapusta, I.; Jaworska, G. Changes in antioxidant activity, profile, and content of polyphenols and tocopherols in common hazel Seed (Corylus avellana L.) depending on variety and harvest date. Molecules 2019, 25, 43. [Google Scholar] [CrossRef] [Green Version]
  114. Granata, M.U.; Bracco, F.; Catoni, R.; Cavalloro, V.; Martino, E. Secondary metabolites profile and physiological leaf traits in wild and cultivated Corylus avellana under different nutritional status. Nat. Prod. Res. 2019, 30, 3100–3107. [Google Scholar] [CrossRef]
  115. Caltagirone, C.; Peano, C.; Sottile, F. Post-harvest industrial processes of almond (Prunus dulcis L. Mill) in Sicily influence the nutraceutical properties of by-products at harvest and during storage. Front. Nutr. 2021, 8, 659378. [Google Scholar] [CrossRef]
  116. Bhagwat, S.; Haytowitz, D.B.; Holden, J.M. USDA Database for the Flavonoid Content of Selected Foods, Release 3.1. Available online: https://www.ars.usda.gov/arsuserfiles/80400525/data/flav/flav_r03-1.pdf (accessed on 18 October 2018).
  117. Gültekin-Özgünven, M.; Davarci, F.; Pasli, A.A.; Demir, N.; Özcelic, B. Determination of phenolic compounds by ultra-high liquid chromatography-tandem mass spectrometry: Applications in nuts. LWT-Food Sci. Technol. 2015, 64, 42–49. [Google Scholar] [CrossRef]
  118. Abe, L.T.; Lajolo, F.M.; Genovese, M.I. Comparison of phenol content and antioxidant capacity of nuts. Food Sci. Technol. 2010, 1 (Suppl. 30), 254–259. [Google Scholar] [CrossRef] [Green Version]
  119. Otles, S.; Seleck, I. Phenolics compounds and antioxidant activities of chestnut (Castanea sativa Mill.) fruits. Qual. Assur. Saf. Crops Foods 2012, 4, 199–205. [Google Scholar] [CrossRef]
  120. Özdemir, K.S.; Yılmaz, C.; Durmaz, G.; Gökmen, V. Hazelnut skin powder: A new brown colored functional ingredient. Food Res. Int. Part B 2014, 65, 291–297. [Google Scholar] [CrossRef]
  121. Bertolino, M.; Belviso, S.; Bello, B.D.; Ghirardello, D.; Giordano, M.; Rolle, L.; Gerbi, V.; Zeppa, G. Influence of the addition of different hazelnut skins on the physicochemical, antioxidant, polyphenol and sensory properties of yogurt. LWT-Food Sci. Technol. 2015, 63, 1145–1154. [Google Scholar] [CrossRef] [Green Version]
  122. Yuan, B.; Lu, M.; Eskridge, K.M.; Isom, L.D.; Hanna, M.A. Extraction, identification, and quantification of antioxidant phenolics from hazelnut (Corylus avellana L.) shells. Food Chem. 2018, 244, 7–15. [Google Scholar] [CrossRef]
  123. Cabo, S.; Aires, A.; Carvalho, R.; Vilela, A.; Pascual-Seva, N.; Silva, A.P.; Gonçalves, B. Kaolin, Ascophyllum nodosum and salicylic acid mitigate effects of summer stress improving hazelnut quality. J. Sci. Food Agric. 2021, 101, 459–475. [Google Scholar] [CrossRef]
  124. Sales, J.M.; Resurreccion, A.V.A. Phenolic profile, antioxidants, and sensory acceptance of bioactive-enhanced peanuts using ultrasound and UV. Food Chem. 2010, 122, 795–803. [Google Scholar] [CrossRef]
  125. De Camargo, A.C.; Regitano-d’Arce, M.A.B.; Gallo, C.R.; Shahidi, F. Gamma-irradiation induced changes in microbiological status, phenolic profile and antioxidant activity of peanut skin. J. Funct. Foods 2015, 12, 129–143. [Google Scholar] [CrossRef]
  126. Attree, R.; Du, B.; Xu, B. Distribution of phenolic compounds in seed coat and cotyledon, and their contribution to antioxidant capacities of red and black seed coat peanuts (Arachis hypogaea L.). Ind. Crops Prod. 2017, 67, 448–456. [Google Scholar] [CrossRef]
  127. Grace, M.H.; Esposito, D.; Timmers, M.A.; Xiong, J.; Yousef, G.; Komarnytskya, S.; Lila, M.A. In vitro lipolytic, antioxidant and anti-inflammatory activities of roasted pistachio kernel and skin constituents. Food Funct. 2016, 7, 4285–4298. [Google Scholar] [CrossRef]
  128. Jakopic, J.; Verberic, R.; Stampar, F. Extraction of phenolic compounds from green walnut fruits in different solvents. Acta Agric. Slov. 2009, 93, 11–15. [Google Scholar] [CrossRef] [Green Version]
  129. Bujdsó, G.; Végvári, G.; Hajnal, V.; Ficzek, G.; Tóth, M. Phenolic profile of the kernel of selected Persian walnut (Juglans regia L.) cultivars. Not. Bot. Horti Agrobot. 2014, 42, 24–29. [Google Scholar] [CrossRef] [Green Version]
  130. Brown, R.C.; Gray, A.R.; Tey, S.L.; Chisholm, A.; Burley, V.; Greenwood, D.C.; Cade, J. Associations between nut consumption and health vary between omnivores, vegetarians, and vegans. Nutrients 2017, 9, 1219. [Google Scholar] [CrossRef] [Green Version]
  131. Musarra-Pizzo, M.; Ginestra, G.; Smeriglio, A.; Pennisi, R.; Sciortino, M.T.; Mandalari, G. The antimicrobial and antiviral activity of polyphenols from almond (Prunus dulcis L.) skin. Nutrients 2019, 11, 2355. [Google Scholar] [CrossRef] [Green Version]
  132. Gorji, N.; Moeini, R.; Memariani, Z. Almond, hazelnut and walnut, three nuts for neuroprotection in Alzheimer’s disease: A neuropharmacological review of their bioactive constituents. Pharmacol. Res. 2018, 129, 115–127. [Google Scholar] [CrossRef]
  133. Mexis, S.; Badeka, A.; Kontominas, M. Quality evaluation of raw ground almond kernels (Prunus dulcis): Effect of active and modified atmosphere packaging, container oxygen barrier and storage conditions. Innov. Food Sci. Emerg. Technol. 2009, 10, 580–589. [Google Scholar] [CrossRef]
  134. Lee, J.; Xiao, L.; Zhang, G.; Ebeler, S.; Mitchell, A. Influence of storage on volatile profiles in roasted almonds (Prunus dulcis). J. Agric. Food. Chem. 2014, 62, 11236–11245. [Google Scholar] [CrossRef] [Green Version]
  135. Valdés, A.; Beltran, A.; Karabagias, I.; Badeka, A.; Kontominas, M.; Garrigos, M. Monitoring the oxidative stability and volatiles in blanched, roasted and fried almonds under normal and accelerated storage conditions by DSC, thermogravimetric analysis and ATR–FTIR. Eur. J. Lipid Sci. Technol. 2015, 117, 1199–1213. [Google Scholar] [CrossRef]
  136. Erten, E.; Cadwallader, K. Identification of predominant aroma components of raw, dry roasted and oil roasted almonds. Food Chem. 2017, 217, 244–253. [Google Scholar] [CrossRef]
  137. Oliveira, I.; Malheiro, R.; Meyer, A.S.; Pereira, J.A.; Gonçalves, B. Application of chemometric tools for the comparison of volatile profile from raw and roasted regional and foreign almond cultivars (Prunus dulcis). J. Food Sci. Technol. 2019, 56, 3764–3776. [Google Scholar] [CrossRef] [Green Version]
  138. García, A.V.; Romero, R.S.; Polo, A.J.; Moya, S.P.; Pérez, S.E.M.; Sanahuja, A.B. Volatile profile of nuts, key odorants and analytical methods for quantification. Foods 2021, 10, 1611. [Google Scholar] [CrossRef]
  139. Elmore, J.S.; Nisyrios, I.; Mottram, D.S. Analysis of the headspace aroma compounds of walnuts (Juglans regia L.). Flavour Fragr. J. 2005, 20, 501–506. [Google Scholar] [CrossRef]
  140. Kiefl, J.; Schieberle, P. Evaluation of process parameters governing the aroma generation in three hazelnut cultivars (Corylus avellana L.) by correlating quantitative key odorant profiling with sensory evaluation. J. Agric. Food Chem. 2013, 61, 5236–5244. [Google Scholar] [CrossRef]
  141. Mexis, S.F.; Kontominas, M.G. Effect of g-irradiation on the physicochemical and sensory properties of hazelnuts (Corylus avellana L.). Radiat. Phys. Chem. 2009, 78, 407–413. [Google Scholar] [CrossRef]
  142. Wilson-Kakashita, G.; Gerdes, D.L.; Hall, W.R. The effect of gamma irradiation on the quality of English walnuts (Juglans regia). LWT-Food Sci. Technol. 1995, 28, 17–20. [Google Scholar] [CrossRef]
  143. Duduzile Buthelezi, N.M.; Samukelo Magwaza, L.; Zeray Tesfay, S. Postharvest pre-storage processing improves antioxidants, nutritional and sensory quality of macadamia nuts. Sci. Hortic. 2019, 251, 197–208. [Google Scholar] [CrossRef]
  144. Chang, S.K.; Alasalvar, C.; Bolling, B.W.; Shahidi, F. Nuts and their co-products: The impact of processing (roasting) on phenolics, bioavailability, and health benefits—A comprehensive review. J. Funct. Foods 2016, 26, 88–122. [Google Scholar] [CrossRef]
  145. Hojjati, M.; Calín-Sánchez, Á.; Razavi, S.H.; Carbonell-Barrachina, Á.A. Effect of roasting on colour and volatile composition of pistachios (Pistacia vera L.). Int. J. Food Sci. 2013, 48, 437–443. [Google Scholar] [CrossRef]
  146. Saklar, S.; Katnas, S.; Ungan, S. Determination of optimum hazelnut roasting conditions. Int. J. Food Sci. 2001, 36, 271–281. [Google Scholar] [CrossRef]
  147. Taş, N.G.; Gökmen, V. Maillard reaction and caramelization during hazelnut roasting: A multiresponse kinetic study. Food Chem. 2017, 221, 1911–1922. [Google Scholar]
  148. Lin, J.T.; Liu, S.C.; Hu, C.C.; Shyu, C.Y.; Hsu, Y.D.J. Effects of roasting temperature and duration on fatty acid composition, phenolic composition, Maillard reaction degree and antioxidant attribute of almond (Prunus dulcis) kernel. Food Chem. 2016, 190, 520–528. [Google Scholar] [CrossRef]
  149. Schmitzer, V.; Slatnar, A.; Veberic, R.; Stampar, F.; Solar, A. Roasting affects phenolic composition and antioxidative activity of hazelnuts (Corylus avellana L.). J. Food Sci. 2011, 76, S14–S19. [Google Scholar] [CrossRef]
  150. Alamprese, C.; Ratti, S.; Rossi, M. Effects of roasting conditions on hazelnut characteristics in a two-step process. J. Food Eng. 2009, 95, 272–279. [Google Scholar] [CrossRef]
  151. Cuadrado, C.; Sanchiz, Á.; Linacero, R. Nut Allergenicity: Effect of food processing. Allergies 2021, 1, 150–162. [Google Scholar] [CrossRef]
  152. Maleki, S.J.; Chung, S.Y.; Champagne, E.T.; Raufman, J.P. The effects of roasting on the allergenic properties of peanut proteins. J. Allergy Clin. Immunol. 2000, 106, 763–768. [Google Scholar] [CrossRef] [Green Version]
  153. Venkatachalam, M.; Teuber, S.S.; Roux, K.H.; Sathe, S.K. Effects of roasting, blanching, autoclaving, and microwave heating on antigenicity of almond (Prunus dulcis L.) proteins. J. Agric. Food Chem. 2002, 50, 3544–3548. [Google Scholar] [CrossRef]
  154. Schlörmann, W.; Birringer, M.; Böhm, V.; Löber, K.; Jahreis, G.; Lorkowski, S.; Muller, A.K.; Schone, F.; Glei, M. Influence of roasting conditions on health-related compounds in different nuts. Food Chem. 2015, 180, 77–85. [Google Scholar] [CrossRef]
  155. Arinola, S.O.; Adesina, K. Effect of thermal processing on the nutritional, antinutritional, and antioxidant properties of Tetracarpidium conophorum (African walnut). J. Food Proc. 2014, 2014, 418380. [Google Scholar] [CrossRef] [Green Version]
  156. Kita, A.; Figiel, A. Effect of roasting on properties of walnuts. Pol. J. Food Nutr. Sci. 2007, 57, 89–94. [Google Scholar]
  157. Tian, Y.; Rao, H.; Zhang, K.; Tao, S.; Xue, W.T. Effects of different thermal processing methods on the structure and allergenicity of peanut allergen Ara h 1. Food Sci. Nutr. 2018, 6, 1706–1714. [Google Scholar] [CrossRef] [Green Version]
  158. Verhoeckx, K.C.M.; Vissers, Y.M.; Baumert, J.L.; Faludi, R.; Feys, M.; Flanagan, S.; Herouet-Guicheney, C.; Holzhauser, T.; Shimojo, R.; van der Bolt, N.; et al. Food processing and allergenicity. Food Chem. Toxicol. 2015, 80, 223–240. [Google Scholar] [CrossRef]
  159. Cuadrado, C.; Cheng, H.; Sanchiz, A.; Ballesteros, I.; Easson, M.; Grimm, C.C.; Dieguez, M.C.; Linacero, R.; Burbano, C.; Maleki, S.J. Influence of enzymatic hydrolysis on the allergenic reactivity of processed cashew and pistachio. Food Chem. 2018, 241, 372–379. [Google Scholar] [CrossRef]
  160. Samoggia, A.; Nicolodi, S. Consumer’s perception of fruit innovation. J. Int. Food Agribus. Mark. 2017, 9, 92–108. [Google Scholar] [CrossRef]
  161. Patil, B.S.; Uckoo, R.M.; Jayaprakasha, G.K.; Palma, M.A. Consumers’ changing perceptions of quality: Revisiting the science of fruit and vegetable cultivation for improved health benefits. Acta Hortic. 2016, 1120, 459–468. [Google Scholar] [CrossRef]
  162. Van Duyn, M.A.S. Year 2000 Dietary Guidelines: The Case for Fruits and Vegetables First; Produce for Better Health Foundation: Wilmington, DE, USA, 1999. [Google Scholar]
  163. Kader, A. Importance of fruits, nuts and vegetables in human nutrition and health. Perish. Handl. Q. 2001, 106, 6. [Google Scholar]
  164. Ströhle, A. Vegetables and fruits in prevention: The German Nutrition Society (DGE) opinion confirms: High consumption of vegetables and fruits reduces risk of contracting diseases. Dtsch. Apoth. Ztg. 2012, 152, 75–77. [Google Scholar] [CrossRef] [Green Version]
  165. Scheerens, J.C. Phytochemicals and the consumer: Factors affecting fruit and vegetable consumption and the potential for increasing small fruit in the diet. Hort Technol. 2001, 11, 547–556. [Google Scholar] [CrossRef] [Green Version]
  166. Kris-Etherton, P.M.; Hecker, K.D.; Bonanome, A.; Coval, S.M.; Binkoski, A.E.; Hilpert, K.F.; Etherton, T.D. Bioactive compounds in foods: Their role in the prevention of cardiovascular disease and cancer. Am. J. Med. 2002, 113, 71–88. [Google Scholar] [CrossRef]
  167. Jenab, M.; Sabate, J.; Slimani, N.; Ferrari, P.; Mazuir, M.; Casagrande, C.; Deharveng, G.; Tjønneland, A.; Olsen, A.; Overvad, K.; et al. Consumption and portion sizes of tree nuts, peanuts and seeds in the European Prospective Investigation into Cancer and Nutrition (EPIC) cohorts from 10 European countries. Br. J. Nutr. 2006, 96, S12–S23. [Google Scholar] [CrossRef] [Green Version]
  168. O’Neil, C.E.; Keast, D.R.; Nicklas, T.A.; Fulgoni, V.L. Out-of-hand nut consumption is associated with improved nutrient intake and health risk markers in US children and adults: National Health and Nutrition Examination Survey 1999–2004. Nutr. Res. 2012, 32, 185–194. [Google Scholar] [CrossRef]
  169. O’Neil, C.E.; Nicklas, T.A.; Fulgoni, V.L. Tree nut consumption is associated with better nutrient adequacy and diet quality in adults: National Health and Nutrition Examination Survey 2005–2010. Nutrients 2015, 7, 595–607. [Google Scholar] [CrossRef] [Green Version]
  170. Flores-Mateo, G.; Rojas-Rueda, D.; Basora, J.; Ros, E.; Salas-Salvadó, J. Nut intake and adiposity: Meta-analysis of clinical trials. Am. J. Clin. Nutr. 2013, 97, 346–1355. [Google Scholar] [CrossRef] [Green Version]
  171. Alasalvar, C.; Shahidi, F. Tree Nuts: Composition, Phytochemicals, and Health Effects; CRC Press: Boca Raton, FL, USA, 2008; 340p, ISBN 9780849337352. [Google Scholar]
  172. Wang, J.; Lee Bravatti, M.A.; Johnson, E.J.; Raman, G. Daily almond consumption in cardiovascular disease prevention via LDL-C change in the U.S. population: A cost-effectiveness analysis. BMC Public Health 2020, 20, 558. [Google Scholar] [CrossRef]
  173. Pawlak, R.; Colby, S.; Herring, J. Beliefs, benefits, barriers, attitude, intake and knowledge about peanuts and tree nuts among WIC participants in eastern North Carolina. Nutr. Res. Pract. 2009, 3, 220–225. [Google Scholar] [CrossRef] [Green Version]
  174. Pawlak, R.; London, H.A.; Colby, S.; Wall-Bassett, E.; Sira, N. Perception of nut intake among individuals with or at risk for heart disease and/or diabetes. J. Behav. Health 2012, 1, 185–188. [Google Scholar] [CrossRef] [Green Version]
  175. Souza, R.G.; Gomes, A.C.; Naves, M.M.; Mota, J.F. Nuts and legume seeds for cardiovascular risk reduction: Scientific evidence and mechanisms of action. Nutr. Rev. 2015, 73, 335–347. [Google Scholar] [CrossRef]
  176. Yong, L.C.; Gray, A.R.; Chisholm, A.; Leong, S.L.; Tey, S.L.; Brown, R.C. Barriers to and facilitators and perceptions of nut consumption among the general population in New Zealand. Public Health Nutr. 2017, 20, 3166–3182. [Google Scholar] [CrossRef] [Green Version]
  177. Li, H.; Li, X.; Yuan, S.; Jin, Y.; Lu, J. Nut consumption and risk of metabolic syndrome and overweight/obesity: A meta-analysis of prospective cohort studies and randomized trials. Nutr. Metab. 2018, 15, 46. [Google Scholar] [CrossRef] [Green Version]
  178. Viale, P.H. The benefits of nuts for cancer prevention. J. Adv. Pract. Oncol. 2019, 10, 102–103. [Google Scholar] [CrossRef]
  179. Shirmohammadi, M.; Chandrasekaran, I.; Singh, C. Effect of post-harvest processes and storage conditions on aging and quality of fruit nuts. In Proceedings of the CSBE/SCGAB 2018 Annual Conference, School of Engineering, University of Guelph, Guelph, ON, Canada, 22–25 July 2018; Available online: https://www.researchgate.net/publication/335160450 (accessed on 17 October 2022).
  180. Kader, A.A. Impact of nut postharvest handling, deshelling, drying and storage on quality. In Improving the Safety and Quality of Nuts; Harris, L.J., Ed.; Woodhead Publishing Ltd.: Cambridge, UK, 2013; Volume 1, pp. 22–34. [Google Scholar]
  181. Mexis, S.F.; Badeka, A.V.; Riganakos, K.A.; Karakostas, K.X.; Kontominas, M.G. Effect of packaging and storage conditions on quality of shelled walnuts. Food Control 2009, 20, 743–751. [Google Scholar] [CrossRef]
  182. Fernandes, L.; Pereira, E.L.; Fidalgo, M.C.; Gomes, A.; Ramalhosa, E. Effect of modified atmosphere, vacuum and polyethylene packaging on physicochemical and microbial quality of chestnuts (Castanea sativa) during storage. Int. J. Fruit Sci. 2020, 20, S785–S801. [Google Scholar] [CrossRef]
  183. Köster, E.P. Diversity in the determinants of food choice: A psychological perspective. Food Qual. Prefer. 2009, 20, 70–82. [Google Scholar] [CrossRef]
  184. Grunert, K.G.; Larsen, H.H.; Madsen, T.K.; Baadsgaard, A. Market Orientation in Food and Agriculture; Kluwer Academic: Dordrecht, The Netherlands, 1996. [Google Scholar]
  185. Brunsø, K.; Birch, D.; Memery, J.; Temesi, Á.; Lakner, Z.; Lang, M.; Dean, D.; Grunert, K.G. Core dimensions of food-related lifestyle: A new instrument for measuring food involvement, innovativeness and responsibility. Food Qual. Prefer. 2021, 91, 104192. [Google Scholar] [CrossRef]
Figure 1. Some representative flavor compounds in almonds, peanuts, and hazelnuts.
Figure 1. Some representative flavor compounds in almonds, peanuts, and hazelnuts.
Foods 12 00942 g001
Table 1. Protein and amino acid contents of selected nuts (compiled from [10]).
Table 1. Protein and amino acid contents of selected nuts (compiled from [10]).
NutProtein (g/100 g)Amino Acids (g/100 g of Portion)
TrpThrIleLeuLysMetCysPheTyrValArgHisAlaAspGluGlyProSer
Almond16.8–25.40.2110.6010.7511.470.5680.1570.2151.130.450.8552.460.5390.9992.646.211.430.9690.912
Cashew nut17.5–19.00.2870.6880.7891.470.9280.3620.3930.9510.5081.092.120.4560.8371.804.510.9370.8121.08
Chestnut1.630.0180.0580.0640.0960.0960.0380.0520.0690.0450.0910.1160.0450.1090.0280.0210,0840,0860.081
Hazelnut14.5–15.20.1930.4970.5451.060.420.2210.2770.6630.3620.7012.210.4320.731.683.710.7240.5610.735
Macadamia nut7.55–8.580.0670.370.3140.6020.0180.0230.0060.6650.5110.3631.400.1950.3881.102.270.4540.4680.419
Peanut25.80.250.8830.9071.670.9260.3170.3311.381.051.083.080.6521.023.155.391.551.141.27
Pecan nut9.0–9.30.0930.3060.3360.5980.2870.1830.1520.4260.2150.4111.180.2620.3970.9291.830.4530.3630.474
Pine nut13.70.1070.370.5420.9910.540.2590.2890.5240.5090.6872.410.3410.6841.32.930.6910.6730.835
Pistachio19.4–22.10.2510.6840.9171.601.140.360.2921.090.5091.252.130.5120.9731.884.31.010.9381.28
Walnut14.4–16.00.170.5960.6251.170.4240.2360.2080.7110.4060.7532.280.3910.6961.832.820.8160.7060.934
Trp—Tryptophan, Thr—Threonine, Ile—Isoleucine, Leu—Leucine, Lys—Lysine, Met—Methionine, Cys—Cystine, Phe—Phenylalanine, Tyr—Tyrosine, Val—Valine, Arg—Arginine, His—Histidine, Ala—Alanine, Asp—Aspartic Acid, Glu—Glutamic Acid, Gly—Glycine, Pro—Proline, Ser—Serine.
Table 2. Vitamin contents (mg/100 g) of selected nuts (source: [10,11,19,31,32,33]).
Table 2. Vitamin contents (mg/100 g) of selected nuts (source: [10,11,19,31,32,33]).
NutAscorbic Acid
(C)
Vit A (IU)Niacin
(B3)
Thiamine (B1)Riboflavin
(B2)
Pyridoxine
(B6)
Folic Acid
(B9)
Pantothenic Acid
(B5)
α-Tocopherol
(E)
Almond3.62–3.900.063.62–3.900.210.80–1.140.10.040.32.4–25.9
Cashew nut1.06–1.10-1.06–1.100.420.06–0.100.40.250.90.0–0.9
Hazelnut1.81201.810.300.100.2–0.6ND0.93.5–15.0
Peanut5.75–12.10-5.75–12.100.600.04–0.100.1–0.30.240.60.4
Pine nut4.40294.40 0.200.1ND0.32.5–9.3
Pistachio1.304151.300.870.16–0.201.751.000.50.3–2.3
Walnut0.47–1.13200.47–1.130.340.15–0.200.5–0.60.980.60.1–13.0
Chestnut40.2261.10.140.02ND580.48-
Not detected—ND.
Table 3. Mineral contents (mg/100 g) of selected nuts (source: [10,11,19,31,32,33,35,36,37]).
Table 3. Mineral contents (mg/100 g) of selected nuts (source: [10,11,19,31,32,33,35,36,37]).
NutNaMgKCaCuZnFe
Almond1.002757282480.90–1.031.91–3.123.71–6.21
Cashew nut12.00292660370.560.96–5.783.82–6.68
Chestnut2.0030.00484.0019.000.4180.490.94
Hazelnut0.70–0.98140–163514–68084–1140.65–0.991.95–2.960.56–4.70
Peanut1.30–18.00168–173558–70567–920.75–0.830.44–3.270.58–4.58
Pine nut2.00251–265597161.32–1.603.08–6.455.53–6.64
Pistachio1.00–9.36117 -121642–1025107–1710.75–1.702.77–6.720.41–8.86
Walnut2.00158–201441–52361–982.541.52–3.372.91–5.74
Table 4. Fiber and lipid contents (%) of selected nuts (adapted from Amarowicz et al. [68]).
Table 4. Fiber and lipid contents (%) of selected nuts (adapted from Amarowicz et al. [68]).
NutFiber (%)Lipid (%)
Almond11.8–13.043.3–50.6
Cashew nut1.4–3.342.8–43.9
Chestnut2.3–3.71.6–7.4
Hazelnut3.4–9.759.8–61.5
Pistachio10.344.4–45.4
Walnut6.764.5–65.2
Table 5. Fatty acid composition of selected nuts (g/100 g nut) (Source: [10]).
Table 5. Fatty acid composition of selected nuts (g/100 g nut) (Source: [10]).
NutSFAMUFAPUFA
TotalPalmitic 16:0Stearic 18:0TotalOleic 18:1Palmitoleic 16:1TotalLinoleic 18:2Linolenic 18:3
Almond3.8023.0830.70431.55131.2940.22712.32912.3240.003
Cashew7.7833.9163.22323.79723.5230.1367.8457.7820.062
Chestnut0.4250.3840.0210.7800.7490.0210.8940.7980.095
Hazelnut4.4643.0971.26545.65245.4050.1167.9207.8330.087
Pistachio5.9075.2650.47823.25722.6740.49514.38014.0910.289
Walnut6.1264.4041.6598.9338.7990.134 (C20:1)47.17438.0939.080
Table 6. Main phenolic compounds found in the most common nuts (skins + kernels).
Table 6. Main phenolic compounds found in the most common nuts (skins + kernels).
NutPhenolic CompoundReference
AlmondCatechin, epicatechin, naringenin, eriodictyol, gallic acid, caffeic acid, chlorogenic acid, o-coumaric acid, p-coumaric acid ferulic acid, hydroxybenzoic acid, protocatechuic, vanillic acid, quercetin, kaempferol, isorhamnetin[71,108,116,117]
ChestnutGallic acid, syringic acid, chlorogenic acid, ferulic acid, vanillic acid, catechin, naringin, quercetin, ellagic acid[107,118,119]
HazelnutGallic acid, protocatechuic acid, caffeic acid, o-coumaric acid, p-coumaric acid, ferullic acid, catechin, epicatechin, epicatechin gallate, rutin[116,117,120,121,122,123]
PeanutCatechin, epicatechin, quercetin, isorhamnetin, gallic acid, protocatechuic, caffeic acid, p-coumaric acid, procyanidins A and B, trimers and tetramers, prodelphinidin[109,117,124,125,126]
Pecan nutEllagic acid, catechin, gallic acid, hydroxybenzoic acid, trans-cinnamic acid, syringic acid, caffeic acid, p-coumaric acid, ferulic acid, naringenin, apigenin, quercetin, rutin, kaempferol, isorhamnetin, resveratrol[118]
PistachioCyanidin, gallic acid, protocatechuic, eriodictyol, catechin, epicatechin, epicatechin gallate, luteolin, quercetin, myricetin, procyanidin B1, trimers, and tetramers[109,117,127]
WalnutVanillic acid, catechin, pyrocatechin, protocatechuic acid, epicatechin, syringic acid, gallic acid, juglone and cinnamic acid, ellagic acid, rutin[118,128,129]
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

Gonçalves, B.; Pinto, T.; Aires, A.; Morais, M.C.; Bacelar, E.; Anjos, R.; Ferreira-Cardoso, J.; Oliveira, I.; Vilela, A.; Cosme, F. Composition of Nuts and Their Potential Health Benefits—An Overview. Foods 2023, 12, 942. https://doi.org/10.3390/foods12050942

AMA Style

Gonçalves B, Pinto T, Aires A, Morais MC, Bacelar E, Anjos R, Ferreira-Cardoso J, Oliveira I, Vilela A, Cosme F. Composition of Nuts and Their Potential Health Benefits—An Overview. Foods. 2023; 12(5):942. https://doi.org/10.3390/foods12050942

Chicago/Turabian Style

Gonçalves, Berta, Teresa Pinto, Alfredo Aires, Maria Cristina Morais, Eunice Bacelar, Rosário Anjos, Jorge Ferreira-Cardoso, Ivo Oliveira, Alice Vilela, and Fernanda Cosme. 2023. "Composition of Nuts and Their Potential Health Benefits—An Overview" Foods 12, no. 5: 942. https://doi.org/10.3390/foods12050942

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