Next Article in Journal
Transmembrane Pressure during Micro- and Diafiltration of Milk Affects the Release of Non-Sedimentable Caseins
Next Article in Special Issue
Food Allergies and Parasites in Children
Previous Article in Journal
Valorization Alternatives of Tropical Forest Fruits Based on the Açai (Euterpe oleracea) Processing in Small Communities
Previous Article in Special Issue
Establishment of Food Allergy Model in Dextran Sulfate Sodium Induced Colitis Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Food Allergens of Plant Origin

1
US Department of Agriculture, Agricultural Research Service, Pacific West Area, Western Regional Research Center, 800 Buchanan Street, Albany, CA 94710, USA
2
Key Laboratory of Precision Nutrition and Food Quality, Key Laboratory of Functional Dairy, Ministry of Education, College of Food Science and Nutritional Engineering, China Agricultural University, Beijing 100083, China
3
School of Food Science and Technology, Jiangnan University, Wuxi 214122, China
4
Collaborative Innovation Center of Food Safety and Quality Control, Jiangnan University, Wuxi 214122, China
5
Department of Obstetrics and Gynecology, The First Affiliated Hospital of USTC, Division of Life Sciences and Medicine, University of Science and Technology of China, Hefei 230001, China
*
Author to whom correspondence should be addressed.
Foods 2023, 12(11), 2232; https://doi.org/10.3390/foods12112232
Submission received: 15 April 2023 / Revised: 24 May 2023 / Accepted: 26 May 2023 / Published: 1 June 2023

Abstract

:
This review presents an update on the physical, chemical, and biological properties of food allergens in plant sources, focusing on the few protein families that contribute to multiple food allergens from different species and protein families recently found to contain food allergens. The structures and structural components of the food allergens in the allergen families may provide further directions for discovering new food allergens. Answers as to what makes some food proteins allergens are still elusive. Factors to be considered in mitigating food allergens include the abundance of the protein in a food, the property of short stretches of the sequence of the protein that may constitute linear IgE binding epitopes, the structural properties of the protein, its stability to heat and digestion, the food matrix the protein is in, and the antimicrobial activity to the microbial flora of the human gastrointestinal tract. Additionally, recent data suggest that widely used techniques for mapping linear IgE binding epitopes need to be improved by incorporating positive controls, and methodologies for mapping conformational IgE binding epitopes need to be developed.

Graphical Abstract

1. Introduction

Food allergies are adverse immune responses to foods. The symptoms of a food allergy range from mild hives and itching to life-threatening anaphylaxis. In the US, it was estimated that up to 26 million adults [1] and 6 million children [2] have food allergies. Depending on the methods of studies, the sub-population suffering from food allergies in Europe was estimated between 0.8% (by positive food challenge) to 19.9% (by a survey of self-reported food allergies) [3], and the overall food allergy prevalence in Asia is comparable to that in the West [4]. Food allergies are among the top causes of anaphylaxis that lead to children’s visits to emergency departments in the United States [5]. In addition, the situation is worsening as food allergy prevalence has increased in the past few decades [2,6,7,8]. Most food allergy reactions are the immediate type that happens within hours of food intake. They are reactions to proteins (except for a small number of cases, see below) mediated by immunoglobulin E (IgE) antibodies. Food allergy reactions happen because the immune system has previously, for unknown reasons, mistaken a food protein as a dangerous invader, switched the class of T helper cells that determines whether the B cells produce IgG or IgE, and developed IgE antibodies against the protein in a so-called sensitization stage. Sensitization to food can happen in an individual when that person consumes it for the first time. It can also occur in people even though they have been eating the food safely for years to decades. In the sensitization stage, abnormal immune responses promote the class switching of B-cells to produce IgE antibodies specific to a food protein and clonal expansion of naive and IgE+ memory B-cell populations [9]. IgE molecules bind to the surface of mast cells and basophils through their association with the high-affinity IgE receptor FcεRI. Subsequent consumption of the food by the patient leads to allergen cross-linking of the IgE antibodies, which, in turn, results in the initiation of allergic reactions via signaling through the high-affinity receptor for the Fc region of immunoglobulin E (IgE) or FcεRI [10,11]. Extensive research on food allergies has been conducted in recent years. Most of these efforts involved studying the genetic, environmental, and other factors that cause a sub-population to develop a food allergy [12]. In comparison, less research has been devoted to studies on the offending allergens. Nevertheless, research interest in allergic food has increased with the upsurge in the prevalence of food allergies, and more and more allergens have been identified, especially in recent years. Here, we present the result of a comprehensive review and analysis of the chemistry of well-established and recently defined food allergens and their families.

2. Food Allergens

Allergens are given names by the Allergen Nomenclature Sub-committee, which operates under the auspices of the World Health Organization (WHO) and the International Union of Immunological Societies (IUIS) [13]. The approved name contains three parts with separations by a space, three letters for the genus being the first part, one letter for the species as the second part, and an Arabic number. The letters are those at the beginning of the genus and species names and the Arabic number indicates the order of the identification of the allergens in that species [13]. The fourth letter of the genus and/or the second letter of the species is included when necessary to remove ambiguity. The WHO/IUIS Allergen Nomenclature Sub-committee also maintains a database of allergens with designated names. This database currently contains 248 food allergens of plant sources from 76 species.
In the early days of plant protein studies, they were classified based on their solubility and extractability in a series of solvent extractions and they were grouped into four groups (albumins, globulins, prolamins, and glutelins) [14]. With the advancement of knowledge of the function, biochemical, and molecular properties of proteins, plant proteins can also be classified into three groups based on their functions, structural and metabolic proteins, protective proteins, and storage proteins. Metabolic proteins can be named by their biochemical activity, while storage proteins are generally those without a known cellular activity other than the storage of nitrogen, carbon, and sulfur for the development of the next generation of the plant. Protective proteins are those with a function to defend the plant against pests, microbial pathogens, or environmental stresses. In the field of modern protein research, one of the methods of obtaining valuable information on proteins by analyzing their sequence, structure, and function is the Pfam classification of protein families based on hidden Markov model profiles [15]. At present, proteins are classified into about 19 thousand Pfam signatures [16].
The number of protein families that contain proteins from plant sources that are known to be capable of eliciting allergic responses in atopic individuals is several orders of magnitude lower compared to the total number of Pfam signatures. There are thousands of proteins in a mature plant seed [17,18], but 79% of plant food allergens belong to only 12 protein families. Table 1 listed the protein families that are known to contain more than two food allergens along with the number of known allergens in each of the families. Some of the allergens (e.g., chitinases) include two or more domains that belong to different protein families. In addition to those shown in Table 1, six protein families contain two allergens. Twenty-four allergens of plant sources belong to protein families that contain only one known allergen. Nine allergens do not belong to any of the classified Pfam families, i.e., searching the Pfam database with the allergen sequences using the online search tool at the site of the Pfam database did not find any hit. Note that minimal sequence information was available about four of the allergens in the database.
The protein family with the most food allergens from plant sources is PF00234 (the protease inhibitor/seed storage/LTP family), which has 74 known allergens. This family includes the plant nonspecific lipid transfer proteins (NsLTP), such as NsLTP1 and NsLTP2, the 2S albumin seed storage proteins, trypsin/alpha-amylase inhibitors, and other proteins. The next protein family is the Cupin-1 family, which includes 39 allergens that are in the WHO/IUIS allergen database. They are also seed storage proteins with close to half of the allergens belonging to the 11S legumins and half belonging to the 7S vicilins. Three of these allergens are now renamed as isoforms of other allergens, but their entries in the database stay due to historical reasons and literature references. With 27 food allergens, the profilin family ranked third in containing more food allergens from plant sources. Thus, the top three protein families contain more than half of the known plant food allergens. In addition, numerous other profilins are known to be pollen allergens. This indicates that the biological activities, physical–chemical properties, and conserved structures of the allergens may play a role in determining or promoting their allergenicity. The following describes the leading plant food allergen families:
Nonspecific lipid transfer proteins. Nonspecific lipid transfer proteins (NsLTPs) are found in all land plants [19]. They are small proteins with molecular masses of around ten kDa. They were demonstrated in vitro to be able to bind and transport various phospholipids to chloroplasts or mitochondria without specificity [20]. NsLTPs are plant pathogenesis-related proteins known as PR-14, and a number of them have been demonstrated to have antimicrobial activities including NsLTPs in wheat (Triticum aestivum L.) [21,22] and mung bean (Vigna radiata L. R. Wilczek) [23]. NsLTPs can be identified by an eight-cysteine residue motif (8CM). Based on the number of residues separating one cysteine from the next and the conservation of residue types at specific positions of the flanking sequences, the NsLTPs can be divided into two types. The 8CM motif for NsLTP1 is CX2VX5–7C[VLI]XY[LAV]X8–13CCXGX12DX[QKR]X2CXCX16–21PX2CX13–15C, and that for NsLTP2 is CX4LX2CX9–11P[ST]X2CCX5QX2–4C[LF]CX2[ALI]X[DN]PX10–12[KR]X4–5CX3–4PX0–2C [24], where X with a subscript number represents the number of non-conserved amino acids residues and allowed residue variation at a single position is placed in a square bracket. Thus, C1X7–10C2X12–17C3C4X8–19C5XC6X19–24C7X4–15C8 can be used to describe the 8CM of the plant NsLTPs, where the Cys residues are numbered from 1 to 8. The functions of NsLTPs are not well understood, but their expression levels are known to be high in most tissues, indicating that they may be essential for the reproduction and survival of plants. Four NsLTP2s and 38 NsLTP1s are known to be food allergens. Known NsLTP food allergens from the major allergen sources recognized by US Food and Drug Administration (FDA) include peanut (Arachis hypogaea L.) allergen Ara h 9 [25] and Ara h 17 [26], almond (Prunus dulcis (Mill.) D.A.Webb) allergen Pru du 3 [26], chestnut (Castanea sativa Mill.) allergen Cas s 8 [27], hazelnut (Corylus avellana L.) allergen Cor a 8 [28], walnut (Juglans regia L.) allergen Jug r 3 [29], and wheat allergen Tri a 14 [30]. Furthermore, the NsLTPs from many plants not used for food are known to be pollen allergens.
The crystal structure Cor a 8 [31] was the first structure reported for an NsLTP food allergen from the major allergen sources, though that of wheat allergen Tri a 14 [32] and the solution structure of Tri a 14 [33] were reported many years ago before it was identified as a food allergen. As shown in Figure 1A, the cysteines in the 8CM of Cor a 8 form four disulfide bonds. Protein structures were generated with the CCP4MG program [34]. The structures of many other NsLTP1 food and pollen allergens are also available. The conservation of these disulfide bond connectivities (between C1–C6, C2–C3, C4–C7, and C5–C8) in NsLTP1s maintains the tertiary contacts of the secondary structural elements and ensures a stable hydrophobic cavity for lipid binding [35]. Moreover, the structure of rice (Oryza sativa L.) NsLTP2 was determined by NMR, which showed an overall structure similar to that of an NsLTP1. The disulfide bond connectivities in NsLTP2 (C1–C5, C2–C3, C4–C7, and C6–C8) are different from those in NsLTP1 [36].
2S albumins. Plant proteins coagulable by heat and soluble in water were called albumins in the early 20th century for their properties that resembled hen egg albumin [14]. The 2S albumins migrated with a 2S sedimentation coefficient during sucrose gradient centrifugation [37]. The 2S albumins also contain an 8CM similar to that of the NsLTPs but with longer sequences separating C2 and C3 and C6 and C7. Seed storage proteins are believed to accumulate in developing seeds to act as a nitrogen reserve for germination [38,39]. The 2S albumins were considered to be a major group of storage proteins in many dicotyledonous plant species [40] that also play a role in providing sulfur reserve in the seed [37]. The 2S albumins have also been suggested to have antimicrobial activities [41,42,43]. Known 2S food allergens from the major allergen sources recognized by FDA include peanut allergen Ara h 2 [44], soybean (Glycine max) allergen Gly m 8 [45], Brazil nut (Bertholletia excelsa Silva Manso) allergen Ber e 1 [46], cashew (Anacardium occidentale L.) allergen Ana o 3 [47], hazelnut allergen Cor a 14 [48], pecan (Carya illinoinensis (Wangenh.) K.Koch) allergen Car i 1 [49], pistachio (Pistacia vera L.) allergen Pis v 1 [50], Stone pine (Pinus pinea L.) allergen Pin p 1 [51], sesame (Sesamum indicum L.) allergens Ses i 1 [52] and Ses i 2 [53], and Black walnut (Juglans nigra L.) allergens Jug n 1 and walnut allergen Jug r 1 [54].
The structures of many 2S albumins including food allergens in castor beans (Ricinus communis L.) (Ric c 1) [55], rapeseed (Brassica napus L.) (Bra n 1) [56], and Brazil nuts (Ber e 1) [57] have been reported. The first structure reported for a 2S albumin allergen from the major allergen sources is that of Ara h 6, which was determined by NMR using recombinantly expressed Ara h 6 with uniform 15N and 13C labeling [58]. Three peanut 2S albumins have been identified as food allergens. They are Ara h 2, Ara h 6, and Ara h 7. The structure of Ara h 2 was also reported (Figure 1B). It was determined by X-ray crystallography using recombinantly expressed Ara h 2 with a maltose-binding protein fused to the N-terminal to enhance its solubility and aid its crystallization [59]. These 2S albumins have the same disulfide bond connections as NsLTP2. However, Ara h 6 has an additional disulfide bond, which is formed by an extra cysteine between C6 and C7 (C6′) and another cysteine residue after C8 (C8′), as shown by one of the models of its structures determined by NMR (Figure 1C).
11S legumins. Both the 11S and the 7S seed storage proteins belong to the cupin superfamily, which was initially recognized based on a 50% sequence identity between the wheat protein germin and a slime mold (Physarum polycephalum) protein spherulin [60]. Germin is an unusually thermostable protein produced during the early phase of germination in wheat embryos. The sequence similarity was then extended to a group of germin-like proteins and globulin storage proteins. Globulins were classified as those soluble in dilute salt solution but insoluble in water [14]. After structural information on canavalin [61] and phaseolin [62] became available, sequence alignment revealed a much larger group of proteins in this superfamily, and the family was given the name cupin [63] (from the Latin term ‘cupa’ which means small barrel). The cupin superfamily contains monocupin, bicupins, and multicupins. It is known to be one of the most functionally diverse protein superfamilies [64] including various proteins with enzymatic functions, non-enzymatic transcription factors, and the 11S and 7S seed storage proteins.
The signature of the cupins includes two sequence motifs separated by an inter-motif sequence with variable length (from 11 amino acids to over a hundred residues). The first motif was defined as GX5HXHX3–4EX6G, and the second motif was characterized as GX5PXGX2HX3N. The two histidines and the glutamate in motif 1 and the histidine in motif 2 may act as ligands to bind metal ions. In many cupins with enzymatic activity, a metal ion is part of the active site [65]. However, the motifs are now known to tolerate variations and not all cupins have a metal ligand [64]. Nevertheless, residues other than those specified above can also provide metal ligand coordination [66].
The 11S globulins are the most widespread among seed storage protein groups. They are present in monocot and eudicot seeds, as well as in conifers and other gymnosperms. They are particularly abundant in legume seeds and are often called legumins. Typical legumins have molecular weights (MW) of about 300–450 kD and consist of six subunits of about 60 kD. These subunits are the products of a multigene family. Each subunit is post-translationally processed to give rise to an acidic (MW about 40 kD) chain and a basic (~20 kD) chain [67]. The acidic and the basic chains are linked by a single disulfide bond. The 11S globulins are rarely, if ever, glycosylated. This family of proteins accounts for many of the known major food allergens from the FDA-recognized major allergen sources including peanut allergen Ara h 3 [68], soybean allergen Gly m 6 [69], almond allergen Pru du 6 [70], Brazil nut allergen Ber e 2, cashew allergen Ana o 2 [71], hazelnut allergen Cor a 9 [72], macadamia (Macadamia integrifolia Maiden and Betche) allergen Mac i 2 [73,74], pecan allergen Car i 4 [75], pistachio allergen Pis v 2 [50], sesame allergens Ses i 6 and Ses i 7 [76], and walnut allergens Jug n 4 [77] and Jug r 4 [78].
The first structure reported for a legumin food allergen from the FDA-recognized major food sources is that of Gly m 6, which was determined before it was designated as a food allergen [13,79]. The 11S seed storage proteins in many species have more than one type of subunit and five for Gly m 6 [79]. Mature 11S proteins are hexamers that can be composed of different subunits, making it problematic for crystallization. The crystal structure of Gly m 6 was determined by purifying the protein from genetically modified soybeans with four of the subunits of the 11S protein deleted. The peanut 11S food allergen is also known to be coded by at least five different genes [80]. Nevertheless, the population of the mature protein that is composed of the translation from a single gene may be high, and the crystallization of Ara h 3 from wild-type peanuts was successful [81]. The first crystal structure of a peanut allergen was that of Ara h 3 (Figure 1D) [82]. The structure of another 11S allergen from the FDA-recognized major food sources, Pru du 6, was also determined [83]. Generally, the 11S allergens are a dimer of trimers. While the doughnut-shaped trimer was made up of three subunits by head-to-tail associations, the back-to-back binding of the trimers forms the hexameric structure of the native molecule. The dimerization of the trimers buries the N-terminus of the basic subdomain which was generated as a result of the cleavage at a conserved peptidase recognition site. The C-terminal of the acidic domain, however, moved away before the trimer–trimer interface to facilitate the packing of the mature hexamer, making it impossible to express the 11S allergen recombinantly with most of the commonly used strategies [83]. The structure of an 11S putative allergens purified from coconut was also determined recently [84,85].
7S vicilins. The 7S globulins are called vicilins, and they are also present in flowering plants and other spermatophytes. Vicilins are trimeric proteins of MW of ~150–190 kD, with a typical subunit MW of ~50 kD. No disulfide bond was found in vicilins, but proteolytic processing and glycosylation may occur [86,87]. Thus, the subunit structure of vicilins revealed by SDS-PAGE is similar in the absence or presence of reducing agents. Vicilins also account for many known major food allergens from the FDA-recognized major allergen sources including peanut allergen Ara h 1 [88], soybean allergen Gly m 5 [69], almond vicilin [89], cashew allergen Ana o 1 [90], hazelnut allergen Cor a 11 and Cor a 16 [91,92], macadamia allergen Mac i 1 [73], pecan allergen Car i 2 [93], pistachio allergen Pis v 3 [94], Korean pine (Pinus koraiensis Siebold and Zucc.) allergen Pin k 2 [95,96], walnut allergens Jug n 2 and Jug r 2 [97], and sesame allergen Ses i 3 [53].
Vicilin leader peptides. Vicilins from some species, such as pea (Pisum sativum L.) allergen Pis s 1 [98], contain just the di-cupin region and a signal peptide. However, vicilins from other species are known to have a variable region between the signal peptide, which can be predicted [99], and the C-terminal di-cupin domains [100,101,102]. This variable leader peptide (VLP) was also called vicilin leader peptide. When they are found in a food independent of the mature vicilin protein with demonstrated allergenicity, they are designated as vicilin iso allergens by the WHO/IUUS Allergen Nomenclature Subcommittee, e.g., Ara h 1.0101 (26–84). In many vicilins, this region can consist of one (as in almond vicilin allergen [89] and peanut allergen Ara h 1 [88]) or more (as in pecan allergen Car i 2 [93]) repeats of a coupled-C3C (cC3C) motif which has a quintet of cysteines arranged as a pair of CX3C linked by 8–12 amino acids. Interestingly, none of this variable region, part of it, or the entirety of the region can be found in the native vicilin purified from the seeds, depending on the plant species. The cC3C area of macadamia nut vicilin was reported to have antimicrobial activities [103]. Available data on the cC3C repeat in the variable region of vicilin food allergens are summarized in Figure 2.
The structures of the cC3C of the VLP of Ara h 1, the two individual cC3C repeats of the VLP of Ana o 1 and Pis v 3, and three cC3C repeats of the VLP of Jug r 2 have been determined by NMR [104,105]. The structures of such cC3C are typical disulfide bond stabilized helix hairpins and the best model of the structure bundle of the second cC3C repeat of Jug r 2 is shown in Figure 1E.
The VLP was also referred to as vicilin buried peptide [104,105]. However, none of the known structures of vicilins included the VLP region of proteins. The vicilins used for structure determination were either those purified from seeds without a cC3C repeat in the mature proteins or recombination proteins that lacked a cC3C region. Figure 1F shows the structure of peanut allergen Ara h 1 [106,107]. A vicilin allergen is a doughnut-shaped trimer, similar to the trimers that dimerize to form the hexameric legumin food allergens. The crystal structures of other vicilin food allergens from the FDA-recognized major food sources have also been reported. These include hazelnut allergen Cor a 11 [108], pecan allergen Car i 2 [93], and pine nut allergen Pin k 2 [95]. The structure of one of the vicilin subunits of soybean that is an isoform of Gly m 5 was also determined [109]. The structures determined earlier for the cupin family proteins were also vicilin structures [61,62].
The cC3C proteins. The cC3C region defines a separate protein family (PF04702) named vicilin_N in the Pfam classification. However, an almond protein that is made up entirely of cC3C repeats has also been identified as a food allergen (Pru du 8) [110]. Its orthologs can also be found in many other species [110,111]. One class of known antimicrobial peptides is the α-hairpinins, which are known to be produced from multimodular precursor proteins. They contain a single cC3C mofit [112] and assume a helix-loop-helix structure with two stabilizing disulfide bonds, as demonstrated by the structure of EcAMP1 from seeds of barnyard grass (Figure 1G) [113]. Whether this disulfide fortified helix-loop-helix is a building block, and the structure of the PF04702 family of allergens remains to be investigated.
Profilins. Profilin is an actin-associated protein that is present in all eukaryotic cells. It was first identified as an allergen in birch pollen [114]. Now it is known to be food allergens in nuts and beans [44,115,116], fruit [117,118,119], and vegetables [120,121]. It has also been identified as pollen allergens from numerous plant sources [122,123]. Thus, it is considered to be a panallergen [123,124] that is often responsible for cross-reactivity [122,124,125]. In the FDA-recognized major allergen sources, known profilin allergens include peanut allergen Ara h 5 [44], Soybean allergen Gly m 3 [116], almond allergen Pru du 4 [126], hazelnut allergen Cor a 2 [127], walnut allergen Jug r 7, and wheat allergen Tri a 12.
The first reported three-dimensional structure of a profilin food allergen is the crystal structure of peanut allergen Ara h 5 (Figure 1H) [128]. Also available is the crystal structure of another profilin food allergen recently determined for muskmelon (Cucumis melo L.) allergen Cuc m 2 [129]. In addition, a number of structures of profilin airway and contact allergens have been reported, including those of birch (Betula verrucosa Ehrh) allergen Bet v 2 [130], corn (Zea mays L.) allergen Zea m 12 and latex (Hevea brasiliensis (Willd. ex A.Juss.) Müll.Arg.) pollen allergen Hev b 8 [131]. Many structures of mammalian profilins have been reported, including those of human, mouse, rat, and cattle profilins.
Pathogenesis related-10 protein. The plant pathogenesis-related 10 (PR-10) proteins are homologous to one of several families of proteins that are upregulated in response to pathogens (such as viruses, bacteria, and fungi) or induced by a wound or adverse environmental stresses [132,133]. Currently, PR proteins are divided into seventeen classes [134]. PR-10 proteins are not related to other classes of PR proteins, and they are expressed constitutively in many parts of the plant [135]. Members of this class of proteins are known to bind to small hydrophobic compounds and biologically important ligands such as fatty acids, cytokinins, sterols, flavonoids, and emodin [136,137]. They are also known to have antimicrobial activities and are reported to have RNase activities [138,139].
PR-10 from 19 species is known to be food allergens including those from peanut (Ara h 8) [140], soybean (Gly m 4), walnut (Jug r 5), and almond (Pru du 1). PR-10 is also known to be pollen allergens in chestnut (Cas s 1) [141], hazelnuts (Cor a 1) [142], and numerous other species. Thus, PR-10 is also one of the cross-reactive plant allergens considered to be a panallergen [143].
The structures have been determined for numerous PR-10 allergens including many food allergens. From the FDA recognized major food allergen sources, hazelnut allergen Cor a 1 [144], peanut allergen Ara h 8 [145] (Figure 1I), and soybean allergen Gly m 4 [146] have been structurally characterized.
Oleosin. Ten known food allergens belong to the oleosin protein family (PF01277). However, they are food allergens from four species; four of them are peanuts allergens, three of them are hazelnut allergens, and two are sesame allergens. The Tartarian buckwheat (Fagopyrum tataricum (L.) Gaertn) allergen Fag t 6 is a recent addition to the oleosin allergen family [147]. Oleosins are one of the main classes of proteins associated with plant oil bodies [148]. Oil bodies are natural oil droplets that can be found in yeast, mammalian cells, and many parts of a plant including leaves, flowers, pollens, and seeds [149]. The oil bodies of plant seeds contain mainly triacylglycerols that are enclosed by a monolayer of phospholipids and associated membrane proteins [150]. Oil bodies are dynamic cellular organelles that function as storage structures, protein trafficking, and lipid catabolism [151]. Oil body-associated proteins are involved in diverse activities, including freeze tolerance and defense against fungal infection [152]. Oleosins were also reported to have acyltransferase and phospholipase activities [153]. To date, no structure of an oleosin food allergen has been determined, but it was predicted to have a central conserved hydrophobic domain that anchors to the phospholipid monolayer flanked by two none conserved amphipathic domains that reside on the surface of oil bodies [154].
Chitinases. Seven of the known food allergens contain a domain belonging to the protein family of chitin recognition protein (PF00187) which is one of the numerous subgroups of carbohydrate-binding modules. Four of the other allergens are pathogenesis-related proteins known as PR-3. They are class I and class IV chitinases that also have a domain belonging to the glycoside hydrolase family 19 (PF00182). Chitin, one of the most abundant polysaccharides on the planet, is a β-(1,4)-linked homopolymer of N-acetylglucosamine. Chitin is an indispensable element of the cell wall for keeping the cellular integrity of Fungi [155,156]. It is also an essential element of nematode eggshell and pharynx [157] and the cuticles and gut lining of insects [158]. Chitinases are a diverse group of enzymes that catalyze the degradation of chitin by cleaving the β-(1–4) linkages, and some of them are also known to have antimicrobial activities [159]. The common buckwheat (Fagopyrum esculentum Moench) allergen Feg e 4 is known to be an antimicrobial peptide [160]. The structures of many chitinases have been determined, including that of corn allergen Zea m 8 with 3 disulfide bonds (PDB: 4MCK; Figure 1J).
Thaumatin-like proteins. Thaumatin-like proteins are also pathogenesis-related proteins known as PR-5 [134]. They are produced in response to a variety of stresses in many plants. Thaumatin purified from the katamfe (Thaumatococcus daniellii) berries taste 105 times sweeter than sucrose on a molar base [161]. Known thaumatin-like food allergens include apple (Malus domestica Borkh) allergen Mal d 2, banana (Musa acuminata Colla) allergen Mus a 4, bell pepper (Capsicum annuum L.) allergen Cap a 1, cherry (Prunus avium (L.) L.) allergen Pru av 2, kiwi (Actinidia deliciosa (A.Chev.) C.F.Liang and A.R.Ferguson) allergen Act d 2, and peach (Prunus persica (L.) Batsch) allergen Pru p 2. The thaumatin-like proteins are also known as pollen allergens. Numerous structures of wild-type and mutant thaumatin-like proteins have been characterized [162], including banana allergen Mus a 4 (Figure 1K) [163].
Gibberellin-regulated proteins. Six of the known food allergens belong to the gibberellin-regulated protein (GRP) family (GASA: PF02704). They are also known as Snakin, GASA, GRP, and peamaclein-like proteins. GRPs are ubiquitously expressed in monocots andeudicots. Many different GRPs have been identified in various plant species including rice [164], apple [165], and wheat [166]. GRPs play important roles in regulating different biological processes such as seed germination, root formation, stem elongation, flowering, etc. [167]. GRP genes are located on many different chromosomes in the genome of a plant [164,166], GRPs are small proteins expressed with a signal peptide, and the mature protein is localized to different parts of the cell including cell wall, chloroplast, vacuole, extracellular membrane, and mitochondria [164]. The main characteristic of the GRPs is their C-terminal 60-amino acid GASA domain which contains 12 conserved cysteines in a pattern of C1X3C2X3C3X7–11C4X3C5X2C6C7X2C8X1–3C9X11C10X1–2C11X11–14KC12P [168]. At the N-terminal of the mature GRPs is a variable region that is reported to contain 7–31 residues rich in polar amino acids [165,169]. Whether there is a correlation between the length of the N-terminal variable region of the GRPs and their localization to different plant organs, reproductive structures, or organelles of plant cells is not well studied. Besides the GRP food allergens, there are also GRP pollen allergens in the WHO/IUIS allergen database. The full sequences of these pollen allergens are currently not available, and all the GRP food allergens have a 4-residue N-terminal variable region in the mature proteins. The Clustal Omega [170] alignment of the sequences of these allergens and those of selected GRPs from other food is shown in Figure 3. A number of GRP proteins, including potato snakin1, have been demonstrated to have antimicrobial activities [171,172].
The crystal structure of potato snakin1 was recently determined by obtaining crystals of a racemic mixture of the L- and D-proteins obtained with total chemical synthesis [173]. The structure of the potato snakin1 and the disulfide bond connections of this antimicrobial protein is shown in Figure 1L. It is worth noting that C2X3C3X7–11C4X3C5 in the GRPs can be considered as a cC3C motif. Moreover, this cC3C has overlapping C3C on both sides of the motif, C1X3C2 on one side, and C5X3C7 on the other. Thus, C1, C2, C3, C4, C5, and C7 and the sequence gaps separating these cysteines constitute an extended cC3C. Because a full turn of an α-helix is 3.6 amino acids, a three-residue separation places the cysteines in a C3C on the same side of an α-helix. A cC3C provides two disulfide bonds to stabilize the two α-helixes connected by the loop constituted by the residues between the couple pair of C3Cs. An overlapping C3C (i.e., C3C3C) places three cysteines on the same side of an α-helix. The coupling of two overlapping C3Cs contributes three disulfide bonds for stabilizing the antiparallel α-helixes, as illustrated by the GRP food allergens.

3. IgE Epitopes

In sensitizing a predisposed individual, both the properties of the linear peptide sequence of an allergen and the properties of the allergen surface may play crucial roles [174]. In general, the majority of antibodies are specific to conformational epitopes [175,176]. However, IgE epitope studies have been focused nearly exclusively on linear epitopes. Earlier identification of linear IgE-binding epitopes was carried out by chemically synthesizing overlapping peptides spanning the whole sequence of an allergen and detecting the presence of IgE antibodies that recognize any of the peptides in dot blot experiments [177]. Recently, microarray technology has enabled the printing of a large number of peptides on chips for linear IgE epitope mapping [178,179]. So far, using sera from specific patient groups, dominant linear IgE epitopes in several allergens have been identified for selected allergens including those in milk [180,181,182], peanuts [68,183,184,185,186], brown shrimp [187], English walnuts [188], cashews [71,90], oyster [189], etc. Recently, we reported a method of recombinantly expressing overlapping peptides fused with human TL1A [190]. The peptide also contained a His tag that was used as a positive control for producing uniform spots. Incorporating a positive control in synthesizing the peptide spots or in printing the microarrays by adding a short epitope tag in future studies may result in better Information about dominant epitopes. Such information can be used to generate hypoallergic peptides and hypoallergens for the development of peptide immunotherapies and food allergen immunotherapies [191].
In contrast, hardly any information is available about conformational IgE epitopes of food allergens although such information is essential for fully understanding the molecular basis of protein allergenicity. To our knowledge, the structures of βLG in complex with the antigen-binding fragment (Fab) of a recombinant IgE [192], IgE-derived Fabs complex with pollen and contact allergens [193,194,195], and the FV region of a monoclonal IgE (raised against a 2,4-dinitrophenyl hapten) in complex with a peptide displayed on the active site loop of thioredoxin [196] are the only structures available for characterizing a conformational IgE epitope experimentally. Nevertheless, many investigations attempted to infer information about IgE epitopes by studying the co-structures of allergens and monoclonal IgG antibodies. Thus, currently, methodologies for determining dominant IgE epitopes, especially for conformational epitopes studies, seem to be inadequate to meet the research requirements.
The current wisdom attributes differences among individuals with different food allergies to the antibody–allergen interface. Patients with persistent allergies are believed to have IgE antibodies that recognize mainly linear epitopes, while individuals with transient allergies have IgE antibodies that recognize a higher proportion of conformational epitopes [197]. However, to our knowledge, there is no applicable method to map conformational epitopes recognized by the IgE repertoire in the serum of a patient. We believe there is a possibility that the current opinions about the importance of conformational epitopes in persistent allergies might be incorrect because currently no method can be used systematically map conformational epitopes.

4. Thermostability

For a food protein to sensitize a person or trigger reactions in a patient, the protein cannot be fully hydrolyzed in the gastrointestinal tract. The protein or smaller peptide pieces of it must survive the attack from digestive enzymes in the gastrointestinal tract and reach the immune system with intact IgE epitopes. The stability of food proteins may be one of the contributing factors to their allergenicity, and allergens from many important plant food allergen families are known to be stable in terms of heat resistance. The peach NsLTP allergen Pru p 3 was reported to maintain its secondary structures after being heated to 95 °C for 15 min at pH 3.4 [198]. The denaturation temperature of Brazil nut allergen Ber e 1 and peanut allergen Ara h 2 (both 2S seed storage proteins) exceeded 100 °C at neutral pH [58,199], and the thermal denaturation of Ber e 1 at low pH was fully reversible [199]. Peanut allergen Ara h 3, an 11S protein, was reported to denature at 92 °C [200], and pine nut allergen Pin k 2, a 7S protein, is also known to have a denaturation temperature over 93 °C [96]. The peach GRP allergen Pru p 7 was reported to maintain its native structure at 90 °C [201]. However, plant food allergens from other dominating families do not have high stability to resist thermal denaturation. The muskmelon allergen Cuc m 2, a profilin, has a reported melting temperature of about 56 °C [129], and celery allergen Api g 1, a PR-10, has a reported denaturation temperature between 70 and 73 °C [202]. Peanut oleosin allergens Ara h 10 and Ara h 11 were reported to have a denaturation midpoint lower than 60 °C [203], and the thermal unfolding of apple allergen Mal d 2 was reported to be reversible which happened at 70 °C [204]. Thus, the thermostability of plant food allergens varies considerably. It might be logical to consider stability to heat treatment as a contributing factor to the allergenicity of a protein. However, heating can modify some amino acids of an allergen, changing its IgE binding properties and altering the relative abundance of soluble allergens within a food [205,206,207]. Thermostability alone probably cannot be used reliably to predict if a protein would be an allergen.

5. Resistance to Digest

For the allergenicity assessment of genetically modified organisms with a newly introduced protein, regulatory agencies in different countries and regional organizations require the stability of the protein when treated with simulated gastric and intestinal fluid. In the literature, however, many different methods were used for studying the digestion stability of food proteins [208]. For example, to mimic the conditions in the stomach, simulated gastric digestion with pepsin was at different pH (1.2–2.5), and there is no standard for the duration of the reaction [199,209,210]. Furthermore, the digestion stabilities of different known allergens reported in a single study can be quite different when assessed parallel with the same protocol [211], and proteins with or without known to be allergenic were demonstrated to have similar stability to pepsin digestion [209,212]. The food matrix and food processing may also affect the digestibility of food proteins [213]. Although we were not aware of a systematic study investigating whether there is a difference between the abilities of allergens and other proteins of plant sources in surviving the digestive systems of allergic and control subjects, it is known that IgE may be present in the gut in food allergy patients [214], and a portion of two ingested food allergens of animal sources (Bos d 5, Gal d 2) is absorbed and transported throughout the body in an antibody-reactive form including through the gut of control subjects [215]. Intuitively, the digestibility may be an important contributor to the allergenicity of a food protein, but there is no standard method to characterize it and available data indicated that it alone cannot be used as a reliable indicator in identifying food allergens.

6. Cross-Reactivity

As noticed above, PR-10 and profilins, actin-associated proteins present in all eukaryotic cells, have been considered panallergens that may cause cross-reactions. The sequence similarities among the orthologous allergens in some other protein families can also be high, even those in species that are not close phylogenetic relatives. Significant cross-reactivity among legumes was recognized [216], and cross-reactivities between peanuts, a legume, and tree nuts were also frequently reported [217]. Specific IgE to different allergens has been reported to be most helpful in predicting patients’ cross-reactions to different foods. For example, omega-5 gliadin-specific IgE antibodies of wheat-allergic patients were reported to be a useful predictor for barley allergy [218]. Ara h 2 could contribute to the high incidence of peanut-allergic individuals’ sensitization to tree nuts [217]. However, though the co-sensitization between legumes in patient groups allergic to different legumes could be partially attributed to the 2S albumins, and the main allergens responsible for the co-sensitization in other studies were reported to be the 11S and 7S globulins [219]. In addition, IgE cross-reactivities were also reported between peanut allergens Ara h 1, Ara h 2, and Ara h 3 [220], but a recent study suggested the reported cross-reactivities likely resulted from the complex formation of the allergens [221]. Furthermore, IgE cross-reactivity between different foods was reported as not always being clinically relevant [219]. This is not surprising as IgE antibodies specific to more than one epitope of a monomeric allergen are required to cross-link IgE receptors to trigger allergic reactions. Further investigations are needed to provide better information about allergen cross-reactivity with patients’ IgE antibodies and their usage in predicting possible allergies. Nevertheless, thermostability, resistance to peptidase digestion, and sequence similarities are among the factors to consider for new proteins’ allergenicity. Although a lot of quality data have been published and experts’ opinions and regulatory guidelines updated, a clear roadmap for assessing the allergenicity and safety of new proteins is still an unachieved objective, and the challenges in accessing their safe use remain to be overcome [222].

7. Summary and Perspectives

In summary, information obtained from stability studies suggested that humans ingest folded, native allergens even if the food is conventionally processed. It appears that the importance of conformational epitopes in food allergens is currently underestimated due to the lack of information. Knowledge of more conformational epitopes is needed to understand better the structural determinants of protein allergenicity.
Many intrinsic characteristics of a food protein may be contributing factors for the protein to be allergenic. The amino acid sequences of proteins drive their three-dimensional structures. Proteins are separated into thoughts of families based on their sequence and structural features. Known food allergens can be grouped into families based on which protein families they belong to. The number of known allergen families is much smaller than the total number of protein families. Thus, the structure of a protein, including its shape and surface properties, may contribute to its allergenicity. Taken together, if a protein’s orthologs in many foods are known to be allergens, it may be suspected to be an allergen and warrant further investigation. On the other hand, recent identification of new food allergens belonging to protein families not known to have allergens before indicated that more research on food allergens are needed to better understand the possible relevance of the properties used for protein classification with the allergenicity of food proteins. In addition, further investigation targeting those protein families with only one allergen representative and assessing the allergenicity of the orthologous proteins with patient sera may also provide important information.
Because the determinants of linear IgE epitopes can be as short as eight amino acids, the sequence property of a short stretch of a peptide in the protein may also contribute to the allergenicity of the protein. Directly linked to the molecular structure of a protein is its thermostability which is also believed to be an important factor for protein allergenicity. Food allergens may also share the characteristics of being able to resist digestion. The abundance of a protein may make it more likely for a portion of it to survive proteolysis than other proteins with similar resistance to digestion. However, one cannot confidently predict the allergenic potential of a protein solely based on one of these factors.

Author Contributions

Conceptualization, Y.Z., H.C., C.L. and T.J. Writing—review and editing, Y.Z., H.C., C.L. and T.J. Visualization, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported, in part, by the US Department of Agriculture, Agricultural Research Service.

Data Availability Statement

The data used to support the findings of this study can be made available by the corresponding author upon request.

Conflicts of Interest

The authors declare no conflict of interest associated with this publication. The mention of trade names or commercial products in this publication is solely to provide specific information and does not imply recommendation or endorsement by USDA. USDA is an equal opportunity provider and employer.

References

  1. Sicherer, S.H.; Sampson, H.A. Food allergy: A review and update on epidemiology, pathogenesis, diagnosis, prevention, and management. J. Allergy Clin. Immunol. 2018, 141, 41–58. [Google Scholar] [CrossRef]
  2. Gupta, R.S.; Springston, E.E.; Warrier, M.R.; Smith, B.; Kumar, R.; Pongracic, J.; Holl, J.L. The prevalence, severity, and distribution of childhood food allergy in the United States. Pediatrics 2011, 128, e9–e17. [Google Scholar] [CrossRef] [PubMed]
  3. Spolidoro, G.C.I.; Amera, Y.T.; Ali, M.M.; Nyassi, S.; Lisik, D.; Ioannidou, A.; Rovner, G.; Khaleva, E.; Venter, C.; van Ree, R.; et al. FREQUENCY of food allergy in Europe: An updated systematic review and meta-analysis. Allergy 2023, 78, 351–368. [Google Scholar] [CrossRef] [PubMed]
  4. Lee, A.J.; Thalayasingam, M.; Lee, B.W. Food allergy in Asia: How does it compare? Asia Pac. Allergy 2013, 3, 3–14. [Google Scholar] [CrossRef] [PubMed]
  5. Motosue, M.S.; Bellolio, M.F.; Van Houten, H.K.; Shah, N.D.; Campbell, R.L. National trends in emergency department visits and hospitalizations for food-induced anaphylaxis in US children. Pediatr. Allergy Immunol. 2018, 29, 538–544. [Google Scholar] [CrossRef]
  6. Gupta, R.S.; Warren, C.M.; Smith, B.M.; Jiang, J.L.; Blumenstock, J.A.; Davis, M.M.; Schleimer, R.P.; Nadeau, K.C. Prevalence and severity of food allergies among US adults. JAMA Netw. Open 2019, 2, e185630. [Google Scholar] [CrossRef] [PubMed]
  7. Sicherer, S.H.; Munoz-Furlong, A.; Sampson, H.A. Prevalence of peanut and tree nut allergy in the United States determined by means of a random digit dial telephone survey: A 5-year follow-up study. J. Allergy Clin. Immunol. 2003, 112, 1203–1207. [Google Scholar] [CrossRef]
  8. Kumfer, A.M.; Commins, S.P. Primary prevention of food allergy. Curr. Allergy Asthma Rep. 2019, 19, 7. [Google Scholar] [CrossRef]
  9. Bhalla, P.L.; Singh, M.B. Biotechnology-based allergy diagnosis and vaccination. Trends Biotechnol. 2008, 26, 153–161. [Google Scholar] [CrossRef]
  10. Galli, S.J.; Tsai, M.; Piliponsky, A.M. The development of allergic inflammation. Nature 2008, 454, 445–454. [Google Scholar] [CrossRef]
  11. Fattakhova, G.; Masilamani, M.; Borrego, F.; Gilfillan, A.M.; Metcalfe, D.D.; Coligan, J.E. The high-affinity immunoglobulin-E receptor (FcepsilonRI) is endocytosed by an AP-2/clathrin-independent, dynamin-dependent mechanism. Traffic 2006, 7, 673–685. [Google Scholar] [CrossRef] [PubMed]
  12. Krempski, J.W.; Dant, C.; Nadeau, K.C. The origins of allergy from a systems approach. Ann. Allergy Asthma Immunol. 2020, 125, 507–516. [Google Scholar] [CrossRef]
  13. IUIS/WHO-Allergen-Nomenclature-Subcommittee. Allergen nomenclature. Bull. World Health Organ. 1994, 72, 797–806. [Google Scholar]
  14. Osborne, T.B. The Vegetable Proteins, 2nd ed.; Green and Co.: London, UK, 1924; p. 440. [Google Scholar]
  15. El-Gebali, S.; Mistry, J.; Bateman, A.; Eddy, S.R.; Luciani, A.; Potter, S.C.; Qureshi, M.; Richardson, L.J.; Salazar, G.A.; Smart, A.; et al. The Pfam protein families database in 2019. Nucleic Acids Res. 2019, 47, D427–D432. [Google Scholar] [CrossRef]
  16. Paysan-Lafosse, T.; Blum, M.; Chuguransky, S.; Grego, T.; Pinto, B.L.; Salazar, G.A.; Bileschi, M.L.; Bork, P.; Bridge, A.; Colwell, L.; et al. InterPro in 2022. Nucleic Acids Res. 2023, 51, D418–D427. [Google Scholar] [CrossRef] [PubMed]
  17. Gallardo, K.; Job, C.; Groot, S.P.; Puype, M.; Demol, H.; Vandekerckhove, J.; Job, D. Proteomic analysis of arabidopsis seed germination and priming. Plant Physiol. 2001, 126, 835–848. [Google Scholar] [CrossRef] [PubMed]
  18. Mooney, B.P.; Krishnan, H.B.; Thelen, J.J. High-throughput peptide mass fingerprinting of soybean seed proteins: Automated workflow and utility of UniGene expressed sequence tag databases for protein identification. Phytochemistry 2004, 65, 1733–1744. [Google Scholar] [CrossRef]
  19. Salminen, T.A.; Blomqvist, K.; Edqvist, J. Lipid transfer proteins: Classification, nomenclature, structure, and function. Planta 2016, 244, 971–997. [Google Scholar] [CrossRef]
  20. Kader, J.C.; Julienne, M.; Vergnolle, C. Purification and characterization of a spinach-leaf protein capable of transferring phospholipids from liposomes to mitochondria or chloroplasts. Eur. J Biochem. 1984, 139, 411–416. [Google Scholar] [CrossRef]
  21. Sun, J.Y.; Gaudet, D.A.; Lu, Z.X.; Frick, M.; Puchalski, B.; Laroche, A. Characterization and antifungal properties of wheat nonspecific lipid transfer proteins. Mol. Plant Microbe Interact. 2008, 21, 346–360. [Google Scholar] [CrossRef]
  22. Odintsova, T.I.; Slezina, M.P.; Istomina, E.A.; Korostyleva, T.V.; Kovtun, A.S.; Kasianov, A.S.; Shcherbakova, L.A.; Kudryavtsev, A.M. Non-specific lipid transfer proteins in Triticum kiharae Dorof. et Migush.: Identification, characterization and expression profiling in response to pathogens and resistance inducers. Pathogens 2019, 8, 221. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, S.Y.; Wu, J.H.; Ng, T.B.; Ye, X.Y.; Rao, P.F. A non-specific lipid transfer protein with antifungal and antibacterial activities from the mung bean. Peptides 2004, 25, 1235–1242. [Google Scholar] [CrossRef]
  24. Wang, N.J.; Lee, C.C.; Cheng, C.S.; Lo, W.C.; Yang, Y.F.; Chen, M.N.; Lyu, P.C. Construction and analysis of a plant non-specific lipid transfer protein database (nsLTPDB). BMC Genom. 2012, 13 (Suppl. S1), S9. [Google Scholar] [CrossRef]
  25. Krause, S.; Reese, G.; Randow, S.; Zennaro, D.; Quaratino, D.; Palazzo, P.; Ciardiello, M.A.; Petersen, A.; Becker, W.M.; Mari, A. Lipid transfer protein (Ara h 9) as a new peanut allergen relevant for a Mediterranean allergic population. J. Allergy Clin. Immunol. 2009, 124, 771–778.e775. [Google Scholar] [CrossRef]
  26. WHO/IUIS-Allergen-Nomenclature-Sub-Committee. Allergen Nomenclature. Available online: http://www.allergen.org/ (accessed on 14 April 2023).
  27. Sanchez-Monge, R.; Blanco, C.; Lopez-Torrejon, G.; Cumplido, J.; Recas, M.; Figueroa, J.; Carrillo, T.; Salcedo, G. Differential allergen sensitization patterns in chestnut allergy with or without associated latex-fruit syndrome. J. Allergy Clin. Immunol. 2006, 118, 705–710. [Google Scholar] [CrossRef]
  28. Pastorello, E.A.; Vieths, S.; Pravettoni, V.; Farioli, L.; Trambaioli, C.; Fortunato, D.; Luttkopf, D.; Calamari, M.; Ansaloni, R.; Scibilia, J.; et al. Identification of hazelnut major allergens in sensitive patients with positive double-blind, placebo-controlled food challenge results. J. Allergy Clin. Immunol. 2002, 109, 563–570. [Google Scholar] [CrossRef]
  29. Pastorello, E.A.; Farioli, L.; Pravettoni, V.; Robino, A.M.; Scibilia, J.; Fortunato, D.; Conti, A.; Borgonovo, L.; Bengtsson, A.; Ortolani, C. Lipid transfer protein and vicilin are important walnut allergens in patients not allergic to pollen. J. Allergy Clin. Immunol. 2004, 114, 908–914. [Google Scholar] [CrossRef]
  30. Sander, I.; Rozynek, P.; Rihs, H.P.; van Kampen, V.; Chew, F.T.; Lee, W.S.; Kotschy-Lang, N.; Merget, R.; Bruning, T.; Raulf-Heimsoth, M. Multiple wheat flour allergens and cross-reactive carbohydrate determinants bind IgE in baker’s asthma. Allergy 2011, 66, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  31. Offermann, L.R.; Bublin, M.; Perdue, M.L.; Pfeifer, S.; Dubiela, P.; Borowski, T.; Chruszcz, M.; Hoffmann-Sommergruber, K. Structural and functional characterization of the hazelnut allergen Cor a 8. J. Agric. Food Chem. 2015, 63, 9150–9158. [Google Scholar] [CrossRef] [PubMed]
  32. Charvolin, D.; Douliez, J.P.; Marion, D.; Cohen-Addad, C.; Pebay-Peyroula, E. The crystal structure of a wheat nonspecific lipid transfer protein (ns-LTP1) complexed with two molecules of phospholipid at 2.1 A resolution. Eur. J. Biochem. 1999, 264, 562–568. [Google Scholar] [CrossRef]
  33. Tassin-Moindrot, S.; Caille, A.; Douliez, J.P.; Marion, D.; Vovelle, F. The wide binding properties of a wheat nonspecific lipid transfer protein. Solution structure of a complex with prostaglandin B2. Eur. J. Biochem. 2000, 267, 1117–1124. [Google Scholar] [CrossRef] [PubMed]
  34. McNicholas, S.; Potterton, E.; Wilson, K.S.; Noble, M.E. Presenting your structures: The CCP4mg molecular-graphics software. Acta Crystallogr. D Biol. Crystallogr. 2011, 67, 386–394. [Google Scholar] [CrossRef]
  35. Sy, D.; Le Gravier, Y.; Goodfellow, J.; Vovelle, F. Protein stability and plasticity of the hydrophobic cavity in wheat ns-LTP. J. Biomol. Struct. Dyn. 2003, 21, 15–29. [Google Scholar] [CrossRef] [PubMed]
  36. Samuel, D.; Liu, Y.J.; Cheng, C.S.; Lyu, P.C. Solution structure of plant nonspecific lipid transfer protein-2 from rice (Oryza sativa). J. Biol. Chem. 2002, 277, 35267–35273. [Google Scholar] [CrossRef]
  37. Youle, R.J.; Huang, A.H.C. Occurrence of low-molecular weight and high cysteine containing albumin storage proteins in oilseeds of diverse species. Am. J. Bot. 1981, 68, 44–48. [Google Scholar] [CrossRef]
  38. Utsumi, S. Plant food protein engineering. Adv. Food Nutr. Res. 1992, 36, 89–208. [Google Scholar]
  39. Shewry, P.R. Plant storage proteins. Biol. Rev. Camb. Philos. Soc. 1995, 70, 375–426. [Google Scholar] [CrossRef]
  40. Shewry, P.R.; Napier, J.A.; Tatham, A.S. Seed storage proteins: Structures and biosynthesis. Plant Cell 1995, 7, 945–956. [Google Scholar]
  41. Candido Ede, S.; Pinto, M.F.; Pelegrini, P.B.; Lima, T.B.; Silva, O.N.; Pogue, R.; Grossi-de-Sa, M.F.; Franco, O.L. Plant storage proteins with antimicrobial activity: Novel insights into plant defense mechanisms. FASEB J. 2011, 25, 3290–3305. [Google Scholar] [CrossRef]
  42. Odintsova, T.I.; Rogozhin, E.A.; Sklyar, I.V.; Musolyamov, A.K.; Kudryavtsev, A.M.; Pukhalsky, V.A.; Smirnov, A.N.; Grishin, E.V.; Egorov, T.A. Antifungal activity of storage 2S albumins from seeds of the invasive weed dandelion Taraxacum officinale Wigg. Protein Pept. Lett. 2010, 17, 522–529. [Google Scholar] [CrossRef] [PubMed]
  43. Pelegrini, P.B.; Noronha, E.F.; Muniz, M.A.; Vasconcelos, I.M.; Chiarello, M.D.; Oliveira, J.T.; Franco, O.L. An antifungal peptide from passion fruit (Passiflora edulis) seeds with similarities to 2S albumin proteins. Biochim. Biophys. Acta 2006, 1764, 1141–1146. [Google Scholar] [CrossRef] [PubMed]
  44. Kleber-Janke, T.; Crameri, R.; Appenzeller, U.; Schlaak, M.; Becker, W.M. Selective cloning of peanut allergens, including profilin and 2S albumins, by phage display technology. Int. Arch. Allergy Immunol. 1999, 119, 265–274. [Google Scholar] [CrossRef]
  45. Klemans, R.J.; Knol, E.F.; Michelsen-Huisman, A.; Pasmans, S.G.; de Kruijf-Broekman, W.; Bruijnzeel-Koomen, C.A.; van Hoffen, E.; Knulst, A.C. Components in soy allergy diagnostics: Gly m 2S albumin has the best diagnostic value in adults. Allergy 2013, 68, 1396–1402. [Google Scholar] [CrossRef]
  46. Pastorello, E.A.; Farioli, L.; Pravettoni, V.; Ispano, M.; Conti, A.; Ansaloni, R.; Rotondo, F.; Incorvaia, C.; Bengtsson, A.; Rivolta, F.; et al. Sensitization to the major allergen of Brazil nut is correlated with the clinical expression of allergy. J. Allergy Clin. Immunol. 1998, 102, 1021–1027. [Google Scholar] [CrossRef]
  47. Robotham, J.M.; Wang, F.; Seamon, V.; Teuber, S.S.; Sathe, S.K.; Sampson, H.A.; Beyer, K.; Seavy, M.; Roux, K.H. Ana o 3, an important cashew nut (Anacardium occidentale L.) allergen of the 2S albumin family. J. Allergy Clin. Immunol. 2005, 115, 1284–1290. [Google Scholar] [CrossRef] [PubMed]
  48. Garino, C.; Zuidmeer, L.; Marsh, J.; Lovegrove, A.; Morati, M.; Versteeg, S.; Schilte, P.; Shewry, P.; Arlorio, M.; van Ree, R. Isolation, cloning, and characterization of the 2S albumin: A new allergen from hazelnut. Mol. Nutr. Food Res. 2010, 54, 1257–1265. [Google Scholar] [CrossRef] [PubMed]
  49. Sharma, G.M.; Irsigler, A.; Dhanarajan, P.; Ayuso, R.; Bardina, L.; Sampson, H.A.; Roux, K.H.; Sathe, S.K. Cloning and characterization of 2S albumin, Car i 1, a major allergen in pecan. J. Agric. Food Chem. 2011, 59, 4130–4139. [Google Scholar] [CrossRef] [PubMed]
  50. Ahn, K.; Bardina, L.; Grishina, G.; Beyer, K.; Sampson, H.A. Identification of two pistachio allergens, Pis v 1 and Pis v 2, belonging to the 2S albumin and 11S globulin family. Clin. Exp. Allergy 2009, 39, 926–934. [Google Scholar] [CrossRef] [PubMed]
  51. Cabanillas, B.; Crespo, J.F.; Maleki, S.J.; Rodriguez, J.; Novak, N. Pin p 1 is a major allergen in pine nut and the first food allergen described in the plant group of gymnosperms. Food Chem. 2016, 210, 70–77. [Google Scholar] [CrossRef]
  52. Pastorello, E.A.; Varin, E.; Farioli, L.; Pravettoni, V.; Ortolani, C.; Trambaioli, C.; Fortunato, D.; Giuffrida, M.G.; Rivolta, F.; Robino, A.; et al. The major allergen of sesame seeds (Sesamum indicum) is a 2S albumin. J. Chromatogr. B Biomed. Sci. Appl. 2001, 756, 85–93. [Google Scholar] [CrossRef]
  53. Beyer, K.; Bardina, L.; Grishina, G.; Sampson, H.A. Identification of sesame seed allergens by 2-dimensional proteomics and Edman sequencing: Seed storage proteins as common food allergens. J. Allergy Clin. Immunol. 2002, 110, 154–159. [Google Scholar] [CrossRef]
  54. Teuber, S.S.; Dandekar, A.M.; Peterson, W.R.; Sellers, C.L. Cloning and sequencing of a gene encoding a 2S albumin seed storage protein precursor from English walnut (Juglans regia), a major food allergen. J. Allergy Clin. Immunol. 1998, 101, 807–814. [Google Scholar] [CrossRef]
  55. Pantoja-Uceda, D.; Bruix, M.; Gimenez-Gallego, G.; Rico, M.; Santoro, J. Solution structure of RicC3, a 2S albumin storage protein from Ricinus communis. Biochemistry 2003, 42, 13839–13847. [Google Scholar] [CrossRef] [PubMed]
  56. Pantoja-Uceda, D.; Palomares, O.; Bruix, M.; Villalba, M.; Rodriguez, R.; Rico, M.; Santoro, J. Solution structure and stability against digestion of rproBnIb, a recombinant 2S albumin from rapeseed: Relationship to its allergenic properties. Biochemistry 2004, 43, 16036–16045. [Google Scholar] [CrossRef] [PubMed]
  57. Rundqvist, L.; Tengel, T.; Zdunek, J.; Bjorn, E.; Schleucher, J.; Alcocer, M.J.; Larsson, G. Solution structure, copper binding and backbone dynamics of recombinant Ber e 1-the major allergen from Brazil nut. PLoS ONE 2012, 7, e46435. [Google Scholar] [CrossRef] [PubMed]
  58. Lehmann, K.; Schweimer, K.; Reese, G.; Randow, S.; Suhr, M.; Becker, W.M.; Vieths, S.; Rosch, P. Structure and stability of 2S albumin-type peanut allergens: Implications for the severity of peanut allergic reactions. Biochem. J. 2006, 395, 463–472. [Google Scholar] [CrossRef]
  59. Mueller, G.A.; Gosavi, R.A.; Pomes, A.; Wunschmann, S.; Moon, A.F.; London, R.E.; Pedersen, L.C. Ara h 2: Crystal structure and IgE binding distinguish two subpopulations of peanut allergic patients by epitope diversity. Allergy 2011, 66, 878–885. [Google Scholar] [CrossRef]
  60. Lane, B.G.; Bernier, F.; Dratewka-Kos, E.; Shafai, R.; Kennedy, T.D.; Pyne, C.; Munro, J.R.; Vaughan, T.; Walters, D.; Altomare, F. Homologies between members of the germin gene family in hexaploid wheat and similarities between these wheat germins and certain Physarum spherulins. J. Biol. Chem. 1991, 266, 10461–10469. [Google Scholar] [CrossRef]
  61. Ko, T.P.; Ng, J.D.; McPherson, A. The three-dimensional structure of canavalin from jack bean (Canavalia ensiformis). Plant Physiol. 1993, 101, 729–744. [Google Scholar] [CrossRef]
  62. Lawrence, M.C.; Izard, T.; Beuchat, M.; Blagrove, R.J.; Colman, P.M. Structure of phaseolin at 2.2 A resolution. Implications for a common vicilin/legumin structure and the genetic engineering of seed storage proteins. J. Mol. Biol. 1994, 238, 748–776. [Google Scholar] [CrossRef]
  63. Dunwell, J.M. Cupins: A new superfamily of functionally diverse proteins that include germins and plant storage proteins. Biotechnol. Genet. Eng. Rev. 1998, 15, 1–32. [Google Scholar] [CrossRef] [PubMed]
  64. Dunwell, J.M.; Purvis, A.; Khuri, S. Cupins: The most functionally diverse protein superfamily? Phytochemistry 2004, 65, 7–17. [Google Scholar] [CrossRef] [PubMed]
  65. Hansen, T.; Schlichting, B.; Felgendreher, M.; Schonheit, P. Cupin-type phosphoglucose isomerases (Cupin-PGIs) constitute a novel metal-dependent PGI family representing a convergent line of PGI evolution. J. Bacteriol. 2005, 187, 1621–1631. [Google Scholar] [CrossRef] [PubMed]
  66. Jin, T.; Wang, Y.; Chen, Y.W.; Fu, T.J.; Kothary, M.H.; McHugh, T.H.; Zhang, Y. Crystal structure of the Korean pine (Pinus koraiensis) 7S seed storage protein with copper ligands. J. Agric. Food Chem. 2014, 62, 222–228. [Google Scholar] [CrossRef]
  67. Mills, E.N.; Jenkins, J.; Marigheto, N.; Belton, P.S.; Gunning, A.P.; Morris, V.J. Allergens of the cupin superfamily. Biochem. Soc. Trans. 2002, 30, 925–929. [Google Scholar] [CrossRef]
  68. Rabjohn, P.; Helm, E.M.; Stanley, J.S.; West, C.M.; Sampson, H.A.; Burks, A.W.; Bannon, G.A. Molecular cloning and epitope analysis of the peanut allergen Ara h 3. J. Clin. Investig. 1999, 103, 535–542. [Google Scholar] [CrossRef] [PubMed]
  69. Holzhauser, T.; Wackermann, O.; Ballmer-Weber, B.K.; Bindslev-Jensen, C.; Scibilia, J.; Perono-Garoffo, L.; Utsumi, S.; Poulsen, L.K.; Vieths, S. Soybean (Glycine max) allergy in Europe: Gly m 5 (beta-conglycinin) and Gly m 6 (glycinin) are potential diagnostic markers for severe allergic reactions to soy. J. Allergy Clin. Immunol. 2009, 123, 452–458. [Google Scholar] [CrossRef]
  70. Willison, L.N.; Tripathi, P.; Sharma, G.; Teuber, S.S.; Sathe, S.K.; Roux, K.H. Cloning, expression and patient IgE reactivity of recombinant Pru du 6, an 11S globulin from almond. Int. Arch. Allergy Immunol. 2011, 156, 267–281. [Google Scholar] [CrossRef]
  71. Wang, F.; Robotham, J.M.; Teuber, S.S.; Sathe, S.K.; Roux, K.H. Ana o 2, a major cashew (Anacardium occidentale L.) nut allergen of the legumin family. Int. Arch. Allergy Immunol. 2003, 132, 27–39. [Google Scholar] [CrossRef]
  72. Beyer, K.; Grishina, G.; Bardina, L.; Grishin, A.; Sampson, H.A. Identification of an 11S globulin as a major hazelnut food allergen in hazelnut-induced systemic reactions. J. Allergy Clin. Immunol. 2002, 110, 517–523. [Google Scholar] [CrossRef]
  73. Kabasser, S.; Pratap, K.; Kamath, S.; Taki, A.C.; Dang, T.; Koplin, J.; Perrett, K.; Hummel, K.; Radauer, C.; Breiteneder, H.; et al. Identification of vicilin, legumin and antimicrobial peptide 2a as macadamia nut allergens. Food Chem. 2022, 370, 131028. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, Y.; Bhardwaj, S.R.; Lyu, S.C.; Chinthrajah, S.; Nadeau, K.C.; Li, C. Expression, purification, characterization, and patient IgE reactivity of new macadamia nut iso-allergen. Protein Expr. Purif. 2022, 203, 106211. [Google Scholar] [CrossRef] [PubMed]
  75. Sharma, G.M.; Irsigler, A.; Dhanarajan, P.; Ayuso, R.; Bardina, L.; Sampson, H.A.; Roux, K.H.; Sathe, S.K. Cloning and characterization of an 11S legumin, Car i 4, a major allergen in pecan. J. Agric. Food Chem. 2011, 59, 9542–9552. [Google Scholar] [CrossRef]
  76. Beyer, K.; Grishina, G.; Bardina, L.; Sampson, H.A. Identification of 2 new sesame seed allergens: Ses i 6 and Ses i 7. J. Allergy Clin. Immunol. 2007, 119, 1554–1556. [Google Scholar] [CrossRef]
  77. Zhang, Y.Z.; Du, W.X.; Fan, Y.; Yi, J.; Lyu, S.C.; Nadeau, K.C.; Thomas, A.L.; McHugh, T. Purification and characterization of a black walnut (Juglans nigra) allergen, Jug n 4. J. Agric. Food Chem. 2017, 65, 454–462. [Google Scholar] [CrossRef]
  78. Wallowitz, M.; Peterson, W.R.; Uratsu, S.; Comstock, S.S.; Dandekar, A.M.; Teuber, S.S. Jug r 4, a legumin group food allergen from walnut (Juglans regia Cv. Chandler). J. Agric. Food Chem. 2006, 54, 8369–8375. [Google Scholar] [CrossRef] [PubMed]
  79. Adachi, M.; Kanamori, J.; Masuda, T.; Yagasaki, K.; Kitamura, K.; Mikami, B.; Utsumi, S. Crystal structure of soybean 11S globulin: Glycinin A3B4 homohexamer. Proc. Natl. Acad. Sci. USA 2003, 100, 7395–7400. [Google Scholar] [CrossRef] [PubMed]
  80. Yan, Y.-S.; Lin, X.-D.; Zhang, Y.-S.; Wang, L.; Wu, K.; Huang, S.-Z. Isolation of peanut genes encoding arachins and conglutins by expressed sequence tags. Plant Sci. 2005, 169, 439–445. [Google Scholar] [CrossRef]
  81. Jin, T.; Howard, A.; Zhang, Y.Z. Purification, crystallization and initial crystallographic characterization of peanut major allergen Ara h 3. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2007, 63, 848–851. [Google Scholar] [CrossRef]
  82. Jin, T.; Guo, F.; Chen, Y.-W.; Howard, A.; Zhang, Y.-Z. Crystal structure of Ara h 3, a major allergen in peanut. Mol. Immunol. 2009, 46, 1796–1804. [Google Scholar] [CrossRef]
  83. Jin, T.; Albillos, S.M.; Guo, F.; Howard, A.; Fu, T.J.; Kothary, M.H.; Zhang, Y.Z. Crystal structure of prunin-1, a major component of the almond (Prunus dulcis) allergen amandin. J. Agric. Food Chem. 2009, 57, 8643–8651. [Google Scholar] [CrossRef]
  84. Vajravijayan, S.; Nandhagopal, N.; Gunasekaran, K. Crystal structure determination and analysis of 11S coconut allergen: Cocosin. Mol. Immunol. 2017, 92, 132–135. [Google Scholar] [CrossRef]
  85. Jin, T.; Wang, C.; Zhang, C.; Wang, Y.; Chen, Y.W.; Guo, F.; Howard, A.; Cao, M.J.; Fu, T.J.; McHugh, T.H.; et al. Crystal structure of cocosin, a potential food allergen from coconut (Cocos nucifera). J. Agric. Food Chem. 2017, 65, 7560–7568. [Google Scholar] [CrossRef] [PubMed]
  86. Gatehouse, J.A.; Lycett, G.W.; Croy, R.R.; Boulter, D. The post-translational proteolysis of the subunits of vicilin from pea (Pisum sativum L.). Biochem. J. 1982, 207, 629–632. [Google Scholar] [CrossRef] [PubMed]
  87. Lopez-Torrejon, G.; Salcedo, G.; Martin-Esteban, M.; Diaz-Perales, A.; Pascual, C.Y.; Sanchez-Monge, R. Len c 1, a major allergen and vicilin from lentil seeds: Protein isolation and cDNA cloning. J. Allergy Clin. Immunol. 2003, 112, 1208–1215. [Google Scholar] [CrossRef] [PubMed]
  88. Burks, A.W.; Cockrell, G.; Stanley, J.S.; Helm, R.M.; Bannon, G.A. Recombinant peanut allergen Ara h I expression and IgE binding in patients with peanut hypersensitivity. J. Clin. Investig. 1995, 96, 1715–1721. [Google Scholar] [CrossRef]
  89. Che, H.L.; Zhang, Y.Z.; Lyu, S.C.; Nadeau, K.C.; McHugh, T. Identification of almond (Prunus dulcis) vicilin as a food allergen. J. Agric. Food Chem. 2019, 67, 425–432. [Google Scholar] [CrossRef]
  90. Wang, F.; Robotham, J.M.; Teuber, S.S.; Tawde, P.; Sathe, S.K.; Roux, K.H. Ana o 1, a cashew (Anacardium occidental) allergen of the vicilin seed storage protein family. J. Allergy Clin. Immunol. 2002, 110, 160–166. [Google Scholar] [CrossRef]
  91. Lauer, I.; Foetisch, K.; Kolarich, D.; Ballmer-Weber, B.K.; Conti, A.; Altmann, F.; Vieths, S.; Scheurer, S. Hazelnut (Corylus avellana) vicilin Cor a 11: Molecular characterization of a glycoprotein and its allergenic activity. Biochem. J. 2004, 383, 327–334. [Google Scholar] [CrossRef]
  92. Mattsson, L.; Holmqvist, M.; Porsch, H.; Larsson, H.; Pontoppidan, B.; Valcour, A.; Lidholm, J. A new vicilin-like allergen in hazelnut giving rise to a spectrum of IgE-binding low-molecular-weight N-terminal fragments. Clin. Exp. Allergy 2022, 52, 1208–1212. [Google Scholar] [CrossRef]
  93. Zhang, Y.; Lee, B.; Du, W.X.; Lyu, S.C.; Nadeau, K.C.; Grauke, L.J.; Zhang, Y.; Wang, S.; Fan, Y.; Yi, J.; et al. Identification and characterization of a new pecan [Carya illinoinensis (Wangenh.) K. Koch] allergen, Car i 2. J. Agric. Food Chem. 2016, 64, 4146–4151. [Google Scholar] [CrossRef] [PubMed]
  94. Willison, L.N.; Tawde, P.; Robotham, J.M.; Penney, R.M.t.; Teuber, S.S.; Sathe, S.K.; Roux, K.H. Pistachio vicilin, Pis v 3, is immunoglobulin E-reactive and cross-reacts with the homologous cashew allergen, Ana o 1. Clin. Exp. Allergy 2008, 38, 1229–1238. [Google Scholar] [CrossRef]
  95. Zhang, Y.Z.; Du, W.X.; Fan, Y.T.; Yi, J.; Lyu, S.C.; Nadeau, K.C.; McHugh, T.H. Identification, characterization, and initial epitope mapping of pine nut allergen Pin k 2. Food Res. Int. 2016, 90, 268–274. [Google Scholar] [CrossRef] [PubMed]
  96. Jin, T.; Albillos, S.M.; Chen, Y.W.; Kothary, M.H.; Fu, T.J.; Zhang, Y.Z. Purification and characterization of the 7S vicilin from Korean pine (Pinus koraiensis). J. Agric. Food Chem. 2008, 56, 8159–8165. [Google Scholar] [CrossRef] [PubMed]
  97. Teuber, S.S.; Jarvis, K.C.; Dandekar, A.M.; Peterson, W.R.; Ansari, A.A. Identification and cloning of a complementary DNA encoding a vicilin-like proprotein, jug r 2, from english walnut kernel (Juglans regia), a major food allergen. J. Allergy Clin. Immunol. 1999, 104, 1311–1320. [Google Scholar] [CrossRef] [PubMed]
  98. Sanchez-Monge, R.; Lopez-Torrejon, G.; Pascual, C.Y.; Varela, J.; Martin-Esteban, M.; Salcedo, G. Vicilin and convicilin are potential major allergens from pea. Clin. Exp. Allergy 2004, 34, 1747–1753. [Google Scholar] [CrossRef]
  99. Teufel, F.; Almagro Armenteros, J.J.; Johansen, A.R.; Gislason, M.H.; Pihl, S.I.; Tsirigos, K.D.; Winther, O.; Brunak, S.; von Heijne, G.; Nielsen, H. SignalP 6.0 predicts all five types of signal peptides using protein language models. Nat. Biotechnol. 2022, 40, 1023–1025. [Google Scholar] [CrossRef] [PubMed]
  100. Wichers, H.J.; De Beijer, T.; Savelkoul, H.F.; Van Amerongen, A. The major peanut allergen Ara h 1 and its cleaved-off N-terminal peptide; possible implications for peanut allergen detection. J. Agric. Food Chem. 2004, 52, 4903–4907. [Google Scholar] [CrossRef]
  101. Dubiela, P.; Kabasser, S.; Smargiasso, N.; Geiselhart, S.; Bublin, M.; Hafner, C.; Mazzucchelli, G.; Hoffmann-Sommergruber, K. Jug r 6 is the allergenic vicilin present in walnut responsible for IgE cross-reactivities to other tree nuts and seeds. Sci. Rep. 2018, 8, 11366. [Google Scholar] [CrossRef]
  102. Goggin, D.E.; Mir, G.; Smith, W.B.; Stuckey, M.; Smith, P.M. Proteomic analysis of lupin seed proteins to identify conglutin Beta as an allergen, Lup an 1. J. Agric. Food Chem. 2008, 56, 6370–6377. [Google Scholar] [CrossRef]
  103. Marcus, J.P.; Green, J.L.; Goulter, K.C.; Manners, J.M. A family of antimicrobial peptides is produced by processing of a 7S globulin protein in Macadamia integrifolia kernels. Plant J. 1999, 19, 699–710. [Google Scholar] [CrossRef] [PubMed]
  104. Foo, A.C.Y.; Nesbit, J.B.; Gipson, S.A.Y.; Cheng, H.; Bushel, P.; DeRose, E.F.; Schein, C.H.; Teuber, S.S.; Hurlburt, B.K.; Maleki, S.J.; et al. Structure, immunogenicity, and IgE cross-reactivity among walnut and peanut vicilin-buried peptides. J. Agric. Food Chem. 2022, 70, 2389–2400. [Google Scholar] [CrossRef] [PubMed]
  105. Foo, A.C.Y.; Nesbit, J.B.; Gipson, S.A.Y.; DeRose, E.F.; Cheng, H.; Hurlburt, B.K.; Kulis, M.D.; Kim, E.H.; Dreskin, S.C.; Mustafa, S.; et al. Structure and IgE cross-reactivity among cashew, pistachio, walnut, and peanut vicilin-buried peptides. J. Agric. Food Chem. 2023, 71, 2990–2998. [Google Scholar] [CrossRef]
  106. Cabanos, C.; Urabe, H.; Tandang-Silvas, M.R.; Utsumi, S.; Mikami, B.; Maruyama, N. Crystal structure of the major peanut allergen Ara h 1. Mol. Immunol. 2011, 49, 115–123. [Google Scholar] [CrossRef] [PubMed]
  107. Chruszcz, M.; Maleki, S.J.; Majorek, K.A.; Demas, M.; Bublin, M.; Solberg, R.; Hurlburt, B.K.; Ruan, S.; Mattison, C.P.; Breiteneder, H.; et al. Structural and immunologic characterization of Ara h 1, a major peanut allergen. J. Biol. Chem. 2011, 286, 39318–39327. [Google Scholar] [CrossRef]
  108. Shikhi, M.; Jain, A.; Salunke, D.M. Comparative study of 7S globulin from Corylus avellana and Solanum lycopersicum revealed importance of salicylic acid and Cu-binding loop in modulating their function. Biochem. Biophys. Res. Commun. 2020, 522, 127–132. [Google Scholar] [CrossRef]
  109. Maruyama, N.; Adachi, M.; Takahashi, K.; Yagasaki, K.; Kohno, M.; Takenaka, Y.; Okuda, E.; Nakagawa, S.; Mikami, B.; Utsumi, S. Crystal structures of recombinant and native soybean beta-conglycinin beta homotrimers. Eur. J. Biochem. 2001, 268, 3595–3604. [Google Scholar] [CrossRef]
  110. Che, H.; Zhang, Y.; Jiang, S.; Jin, T.; Lyu, S.C.; Nadeau, K.C.; McHugh, T. Almond (Prunus dulcis) allergen Pru du 8, the first member of a new family of food allergens. J. Agric. Food Chem. 2019, 67, 8626–8631. [Google Scholar] [CrossRef]
  111. Zhang, Y.; Jin, T. Almond allergens: Update and perspective on identification and characterization. J. Sci. Food Agric. 2020, 100, 4657–4663. [Google Scholar] [CrossRef]
  112. Tam, J.P.; Wang, S.; Wong, K.H.; Tan, W.L. Antimicrobial peptides from plants. Pharmaceuticals 2015, 8, 711–757. [Google Scholar] [CrossRef]
  113. Nolde, S.B.; Vassilevski, A.A.; Rogozhin, E.A.; Barinov, N.A.; Balashova, T.A.; Samsonova, O.V.; Baranov, Y.V.; Feofanov, A.V.; Egorov, T.A.; Arseniev, A.S.; et al. Disulfide-stabilized helical hairpin structure and activity of a novel antifungal peptide EcAMP1 from seeds of barnyard grass (Echinochloa crus-galli). J. Biol. Chem. 2011, 286, 25145–25153. [Google Scholar] [CrossRef] [PubMed]
  114. Valenta, R.; Duchene, M.; Pettenburger, K.; Sillaber, C.; Valent, P.; Bettelheim, P.; Breitenbach, M.; Rumpold, H.; Kraft, D.; Scheiner, O. Identification of profilin as a novel pollen allergen; IgE autoreactivity in sensitized individuals. Science 1991, 253, 557–560. [Google Scholar] [CrossRef] [PubMed]
  115. Hirschwehr, R.; Valenta, R.; Ebner, C.; Ferreira, F.; Sperr, W.R.; Valent, P.; Rohac, M.; Rumpold, H.; Scheiner, O.; Kraft, D. Identification of common allergenic structures in hazel pollen and hazelnuts: A possible explanation for sensitivity to hazelnuts in patients allergic to tree pollen. J. Allergy Clin. Immunol. 1992, 90, 927–936. [Google Scholar] [CrossRef]
  116. Rihs, H.P.; Chen, Z.; Rueff, F.; Petersen, A.; Rozynek, P.; Heimann, H.; Baur, X. IgE binding of the recombinant allergen soybean profilin (rGly m 3) is mediated by conformational epitopes. J. Allergy Clin. Immunol. 1999, 104, 1293–1301. [Google Scholar] [CrossRef]
  117. Reindl, J.; Rihs, H.P.; Scheurer, S.; Wangorsch, A.; Haustein, D.; Vieths, S. IgE reactivity to profilin in pollen-sensitized subjects with adverse reactions to banana and pineapple. Int. Arch. Allergy Immunol. 2002, 128, 105–114. [Google Scholar] [CrossRef]
  118. Lopez-Torrejon, G.; Crespo, J.F.; Sanchez-Monge, R.; Sanchez-Jimenez, M.; Alvarez, J.; Rodriguez, J.; Salcedo, G. Allergenic reactivity of the melon profilin Cuc m 2 and its identification as major allergen. Clin. Exp. Allergy 2005, 35, 1065–1072. [Google Scholar] [CrossRef]
  119. Lopez-Torrejon, G.; Ibanez, M.D.; Ahrazem, O.; Sanchez-Monge, R.; Sastre, J.; Lombardero, M.; Barber, D.; Salcedo, G. Isolation, cloning and allergenic reactivity of natural profilin Cit s 2, a major orange allergen. Allergy 2005, 60, 1424–1429. [Google Scholar] [CrossRef]
  120. Ballmer-Weber, B.K.; Wuthrich, B.; Wangorsch, A.; Fotisch, K.; Altmann, F.; Vieths, S. Carrot allergy: Double-blinded, placebo-controlled food challenge and identification of allergens. J. Allergy Clin. Immunol. 2001, 108, 301–307. [Google Scholar] [CrossRef]
  121. Luttkopf, D.; Ballmer-Weber, B.K.; Wuthrich, B.; Vieths, S. Celery allergens in patients with positive double-blind placebo-controlled food challenge. J. Allergy Clin. Immunol. 2000, 106, 390–399. [Google Scholar] [CrossRef]
  122. Santos, A.; Van Ree, R. Profilins: Mimickers of allergy or relevant allergens? Int. Arch. Allergy Immunol. 2011, 155, 191–204. [Google Scholar] [CrossRef]
  123. Scala, E.; Alessandri, C.; Palazzo, P.; Pomponi, D.; Liso, M.; Bernardi, M.L.; Ferrara, R.; Zennaro, D.; Santoro, M.; Rasi, C.; et al. IgE recognition patterns of profilin, PR-10, and tropomyosin panallergens tested in 3,113 allergic patients by allergen microarray-based technology. PLoS ONE 2011, 6, e24912. [Google Scholar] [CrossRef] [PubMed]
  124. Scheurer, S.; Wangorsch, A.; Nerkamp, J.; Skov, P.S.; Ballmer-Weber, B.; Wuthrich, B.; Haustein, D.; Vieths, S. Cross-reactivity within the profilin panallergen family investigated by comparison of recombinant profilins from pear (Pyr c 4), cherry (Pru av 4) and celery (Api g 4) with birch pollen profilin Bet v 2. J. Chromatogr. B Biomed. Sci. Appl. 2001, 756, 315–325. [Google Scholar] [CrossRef]
  125. Bauer, L.; Ebner, C.; Hirschwehr, R.; Wuthrich, B.; Pichler, C.; Fritsch, R.; Scheiner, O.; Kraft, D. IgE cross-reactivity between birch pollen, mugwort pollen and celery is due to at least three distinct cross-reacting allergens: Immunoblot investigation of the birch-mugwort-celery syndrome. Clin. Exp. Allergy 1996, 26, 1161–1170. [Google Scholar] [CrossRef] [PubMed]
  126. Tawde, P.; Venkatesh, Y.P.; Wang, F.; Teuber, S.S.; Sathe, S.K.; Roux, K.H. Cloning and characterization of profilin (Pru du 4), a cross-reactive almond (Prunus dulcis) allergen. J. Allergy Clin. Immunol. 2006, 118, 915–922. [Google Scholar] [CrossRef] [PubMed]
  127. Hansen, K.S.; Ballmer-Weber, B.K.; Luttkopf, D.; Skov, P.S.; Wuthrich, B.; Bindslev-Jensen, C.; Vieths, S.; Poulsen, L.K. Roasted hazelnuts—Allergenic activity evaluated by double-blind, placebo-controlled food challenge. Allergy 2003, 58, 132–138. [Google Scholar] [CrossRef]
  128. Wang, Y.; Fu, T.J.; Howard, A.; Kothary, M.H.; McHugh, T.H.; Zhang, Y. Crystal structure of peanut (Arachis hypogaea) allergen Ara h 5. J. Agric. Food Chem. 2013, 61, 1573–1578. [Google Scholar] [CrossRef]
  129. Kapingidza, A.B.; Pye, S.E.; Hyduke, N.; Dolamore, C.; Pote, S.; Schlachter, C.R.; Commins, S.P.; Kowal, K.; Chruszcz, M. Comparative structural and thermal stability studies of Cuc m 2.0101, Art v 4.0101 and other allergenic profilins. Mol. Immunol. 2019, 114, 19–29. [Google Scholar] [CrossRef] [PubMed]
  130. Fedorov, A.A.; Ball, T.; Mahoney, N.M.; Valenta, R.; Almo, S.C. The molecular basis for allergen cross-reactivity: Crystal structure and IgE-epitope mapping of birch pollen profilin. Structure 1997, 5, 33–45. [Google Scholar] [CrossRef]
  131. Mares-Mejia, I.; Martinez-Caballero, S.; Garay-Canales, C.; Cano-Sanchez, P.; Torres-Larios, A.; Lara-Gonzalez, S.; Ortega, E.; Rodriguez-Romero, A. Structural insights into the IgE mediated responses induced by the allergens Hev b 8 and Zea m 12 in their dimeric forms. Sci. Rep. 2016, 6, 32552. [Google Scholar] [CrossRef]
  132. Breiteneder, H.; Ebner, C. Molecular and biochemical classification of plant-derived food allergens. J. Allergy Clin. Immunol. 2000, 106, 27–36. [Google Scholar] [CrossRef]
  133. Hoffmann-Sommergruber, K. Pathogenesis-related (PR)-proteins identified as allergens. Biochem. Soc. Trans. 2002, 30, 930–935. [Google Scholar] [CrossRef] [PubMed]
  134. Ali, S.; Ganai, B.A.; Kamili, A.N.; Bhat, A.A.; Mir, Z.A.; Bhat, J.A.; Tyagi, A.; Islam, S.T.; Mushtaq, M.; Yadav, P.; et al. Pathogenesis-related proteins and peptides as promising tools for engineering plants with multiple stress tolerance. Microbiol. Res. 2018, 212–213, 29–37. [Google Scholar] [CrossRef] [PubMed]
  135. Fernandes, H.; Michalska, K.; Sikorski, M.; Jaskolski, M. Structural and functional aspects of PR-10 proteins. FEBS J. 2013, 280, 1169–1199. [Google Scholar] [CrossRef] [PubMed]
  136. Koistinen, K.M.; Soininen, P.; Venalainen, T.A.; Hayrinen, J.; Laatikainen, R.; Perakyla, M.; Tervahauta, A.I.; Karenlampi, S.O. Birch PR-10c interacts with several biologically important ligands. Phytochemistry 2005, 66, 2524–2533. [Google Scholar] [CrossRef] [PubMed]
  137. Mogensen, J.E.; Wimmer, R.; Larsen, J.N.; Spangfort, M.D.; Otzen, D.E. The major birch allergen, Bet v 1, shows affinity for a broad spectrum of physiological ligands. J. Biol. Chem. 2002, 277, 23684–23692. [Google Scholar] [CrossRef]
  138. Bantignies, B.; Seguin, J.; Muzac, I.; Dedaldechamp, F.; Gulick, P.; Ibrahim, R. Direct evidence for ribonucleolytic activity of a PR-10-like protein from white lupin roots. Plant Mol. Biol. 2000, 42, 871–881. [Google Scholar] [CrossRef]
  139. Koistinen, K.M.; Kokko, H.I.; Hassinen, V.H.; Tervahauta, A.I.; Auriola, S.; Karenlampi, S.O. Stress-related RNase PR-10c is post-translationally modified by glutathione in birch. Plant Cell Environ. 2002, 25, 707–715. [Google Scholar] [CrossRef]
  140. Mittag, D.; Akkerdaas, J.; Ballmer-Weber, B.K.; Vogel, L.; Wensing, M.; Becker, W.M.; Koppelman, S.J.; Knulst, A.C.; Helbling, A.; Hefle, S.L.; et al. Ara h 8, a Bet v 1-homologous allergen from peanut, is a major allergen in patients with combined birch pollen and peanut allergy. J. Allergy Clin. Immunol. 2004, 114, 1410–1417. [Google Scholar] [CrossRef]
  141. Kos, T.; Hoffmann-Sommergruber, K.; Ferreira, F.; Hirschwehr, R.; Ahorn, H.; Horak, F.; Jager, S.; Sperr, W.; Kraft, D.; Scheiner, O. Purification, characterization and N-terminal amino acid sequence of a new major allergen from European chestnut pollen-Cas s 1. Biochem. Biophys. Res. Commun. 1993, 196, 1086–1092. [Google Scholar] [CrossRef]
  142. Florvaag, E.; Holen, E.; Vik, H.; Elsayed, S. Comparative studies on tree pollen allergens. XIV. Characterization of the birch (Betula verrucosa) and hazel (Corylus avellana) pollen extracts by horizontal 2-D SDS-PAGE combined with electrophoretic transfer and IgE immunoautoradiography. Ann. Allergy 1988, 61, 392–400. [Google Scholar]
  143. Hauser, M.; Roulias, A.; Ferreira, F.; Egger, M. Panallergens and their impact on the allergic patient. Allergy Asthma Clin. Immunol. 2010, 6, 1. [Google Scholar] [CrossRef] [PubMed]
  144. Jacob, T.; von Loetzen, C.S.; Reuter, A.; Lacher, U.; Schiller, D.; Schobert, R.; Mahler, V.; Vieths, S.; Rosch, P.; Schweimer, K.; et al. Identification of a natural ligand of the hazel allergen Cor a 1. Sci. Rep. 2019, 9, 8714. [Google Scholar] [CrossRef] [PubMed]
  145. Hurlburt, B.K.; Offermann, L.R.; McBride, J.K.; Majorek, K.A.; Maleki, S.J.; Chruszcz, M. Structure and function of the peanut panallergen Ara h 8. J. Biol. Chem. 2013, 288, 36890–36901. [Google Scholar] [CrossRef] [PubMed]
  146. Berkner, H.; Neudecker, P.; Mittag, D.; Ballmer-Weber, B.K.; Schweimer, K.; Vieths, S.; Rosch, P. Cross-reactivity of pollen and food allergens: Soybean Gly m 4 is a member of the Bet v 1 superfamily and closely resembles yellow lupine proteins. Biosci. Rep. 2009, 29, 183–192. [Google Scholar] [CrossRef] [PubMed]
  147. Chen, F.; Li, H.; Fan, X.; Li, Y.; Zhang, C.; Zhu, L.; Hu, J.; Kombe Kombe, A.J.; Xie, J.; Yin, D.; et al. Identification of a novel major allergen in buckwheat seeds: Fag t 6. J. Agric. Food Chem. 2021, 69, 13315–13322. [Google Scholar] [CrossRef]
  148. Shao, Q.; Liu, X.; Su, T.; Ma, C.; Wang, P. New insights into the role of seed oil body proteins in metabolism and plant development. Front. Plant Sci. 2019, 10, 1568. [Google Scholar] [CrossRef] [PubMed]
  149. Huang, A.H. Oleosins and oil bodies in seeds and other organs. Plant Physiol. 1996, 110, 1055–1061. [Google Scholar] [CrossRef]
  150. Nikiforidis, C.V. Structure and functions of oleosomes (oil bodies). Adv. Colloid Interface Sci. 2019, 274, 102039. [Google Scholar] [CrossRef]
  151. Jolivet, P.; Acevedo, F.; Boulard, C.; d’Andrea, S.; Faure, J.D.; Kohli, A.; Nesi, N.; Valot, B.; Chardot, T. Crop seed oil bodies: From challenges in protein identification to an emerging picture of the oil body proteome. Proteomics 2013, 13, 1836–1849. [Google Scholar] [CrossRef]
  152. Shimada, T.L.; Hayashi, M.; Hara-Nishimura, I. Membrane dynamics and multiple functions of oil bodies in seeds and leaves. Plant Physiol. 2018, 176, 199–207. [Google Scholar] [CrossRef]
  153. Parthibane, V.; Rajakumari, S.; Venkateshwari, V.; Iyappan, R.; Rajasekharan, R. Oleosin is bifunctional enzyme that has both monoacylglycerol acyltransferase and phospholipase activities. J. Biol. Chem. 2012, 287, 1946–1954. [Google Scholar] [CrossRef]
  154. Tzen, J.T.C. Integral Proteins in Plant Oil Bodies. ISRN Bot. 2012, 2012, 173954. [Google Scholar] [CrossRef]
  155. Geoghegan, I.; Steinberg, G.; Gurr, S. The role of the fungal cell wall in the infection of plants. Trends Microbiol. 2017, 25, 957–967. [Google Scholar] [CrossRef]
  156. Durkin, C.A.; Mock, T.; Armbrust, E.V. Chitin in diatoms and its association with the cell wall. Eukaryot. Cell 2009, 8, 1038–1050. [Google Scholar] [CrossRef] [PubMed]
  157. Chen, Q.; Peng, D. Nematode chitin and application. Adv. Exp. Med. Biol. 2019, 1142, 209–219. [Google Scholar]
  158. Merzendorfer, H.; Zimoch, L. Chitin metabolism in insects: Structure, function and regulation of chitin synthases and chitinases. J. Exp. Biol. 2003, 206, 4393–4412. [Google Scholar] [CrossRef]
  159. Guan, Y.; Chye, M.L. A Brassica juncea chitinase with two-chitin binding domains show anti-microbial properties against phytopathogens and Gram-negative bacteria. Plant Signal. Behav. 2008, 3, 1103–1105. [Google Scholar] [CrossRef]
  160. Fujimura, M.; Minami, Y.; Watanabe, K.; Tadera, K. Purification, characterization, and sequencing of a novel type of antimicrobial peptides, Fa-AMP1 and Fa-AMP2, from seeds of buckwheat (Fagopyrum esculentum Moench.). Biosci. Biotechnol. Biochem. 2003, 67, 1636–1642. [Google Scholar] [CrossRef]
  161. van der Wel, H.; Loeve, K. Isolation and characterization of thaumatin I and II, the sweet-tasting proteins from Thaumatococcus daniellii Benth. Eur. J. Biochem. 1972, 31, 221–225. [Google Scholar] [CrossRef]
  162. Masuda, T.; Okubo, K.; Murata, K.; Mikami, B.; Sugahara, M.; Suzuki, M.; Temussi, P.A.; Tani, F. Subatomic structure of hyper-sweet thaumatin D21N mutant reveals the importance of flexible conformations for enhanced sweetness. Biochimie 2019, 157, 57–63. [Google Scholar] [CrossRef] [PubMed]
  163. Leone, P.; Menu-Bouaouiche, L.; Peumans, W.J.; Payan, F.; Barre, A.; Roussel, A.; Van Damme, E.J.; Rouge, P. Resolution of the structure of the allergenic and antifungal banana fruit thaumatin-like protein at 1.7-A. Biochimie 2006, 88, 45–52. [Google Scholar] [CrossRef]
  164. Muhammad, I.; Li, W.Q.; Jing, X.Q.; Zhou, M.R.; Shalmani, A.; Ali, M.; Wei, X.Y.; Sharif, R.; Liu, W.T.; Chen, K.M. A systematic in silico prediction of gibberellic acid stimulated GASA family members: A novel small peptide contributes to floral architecture and transcriptomic changes induced by external stimuli in rice. J. Plant Physiol. 2019, 234–235, 117–132. [Google Scholar] [CrossRef]
  165. Fan, S.; Zhang, D.; Zhang, L.; Gao, C.; Xin, M.; Tahir, M.M.; Li, Y.; Ma, J.; Han, M. Comprehensive analysis of GASA family members in the Malus domestica genome: Identification, characterization, and their expressions in response to apple flower induction. BMC Genom. 2017, 18, 827. [Google Scholar] [CrossRef]
  166. Cheng, X.; Wang, S.; Xu, D.; Liu, X.; Li, X.; Xiao, W.; Cao, J.; Jiang, H.; Min, X.; Wang, J.; et al. Identification and analysis of the GASR Gene family in common wheat (Triticum aestivum L.) and characterization of TaGASR34, a gene associated with seed dormancy and germination. Front. Genet. 2019, 10, 980. [Google Scholar] [CrossRef] [PubMed]
  167. Zhang, S.; Wang, X. One new kind of phytohormonal signaling integrator: Up-and-coming GASA family genes. Plant Signal. Behav. 2017, 12, e1226453. [Google Scholar] [CrossRef] [PubMed]
  168. Silverstein, K.A.; Moskal, W.A., Jr.; Wu, H.C.; Underwood, B.A.; Graham, M.A.; Town, C.D.; VandenBosch, K.A. Small cysteine-rich peptides resembling antimicrobial peptides have been under-predicted in plants. Plant J. 2007, 51, 262–280. [Google Scholar] [CrossRef] [PubMed]
  169. Aubert, D.; Chevillard, M.; Dorne, A.M.; Arlaud, G.; Herzog, M. Expression patterns of GASA genes in Arabidopsis thaliana: The GASA4 gene is up-regulated by gibberellins in meristematic regions. Plant Mol. Biol. 1998, 36, 871–883. [Google Scholar] [CrossRef] [PubMed]
  170. Sievers, F.; Higgins, D.G. The Clustal Omega multiple alignment package. Methods Mol. Biol. 2021, 2231, 3–16. [Google Scholar] [CrossRef]
  171. Rodriguez-Decuadro, S.; Barraco-Vega, M.; Dans, P.D.; Pandolfi, V.; Benko-Iseppon, A.M.; Cecchetto, G. Antimicrobial and structural insights of a new snakin-like peptide isolated from Peltophorum dubium (Fabaceae). Amino Acids 2018, 50, 1245–1259. [Google Scholar] [CrossRef]
  172. Almasia, N.I.; Nahirnak, V.; Hopp, H.E.; Vazquez-Rovere, C. Potato Snakin-1: An antimicrobial player of the trade-off between host defense and development. Plant Cell Rep. 2020, 39, 839–849. [Google Scholar] [CrossRef]
  173. Yeung, H.; Squire, C.J.; Yosaatmadja, Y.; Panjikar, S.; Lopez, G.; Molina, A.; Baker, E.N.; Harris, P.W.; Brimble, M.A. Radiation damage and racemic protein crystallography reveal the unique structure of the GASA/Snakin protein superfamily. Angew. Chem. Int. Ed. Engl. 2016, 55, 7930–7933. [Google Scholar] [CrossRef] [PubMed]
  174. Pomes, A.; Chruszcz, M.; Gustchina, A.; Minor, W.; Mueller, G.A.; Pedersen, L.C.; Wlodawer, A.; Chapman, M.D. 100 Years later: Celebrating the contributions of x-ray crystallography to allergy and clinical immunology. J. Allergy Clin. Immunol. 2015, 136, 29–37.e10. [Google Scholar] [CrossRef]
  175. Barlow, D.J.; Edwards, M.S.; Thornton, J.M. Continuous and discontinuous protein antigenic determinants. Nature 1986, 322, 747–748. [Google Scholar] [CrossRef]
  176. Dall’antonia, F.; Pavkov-Keller, T.; Zangger, K.; Keller, W. Structure of allergens and structure based epitope predictions. Methods 2014, 66, 3–21. [Google Scholar] [CrossRef]
  177. Burks, A.W.; Shin, D.; Cockrell, G.; Stanley, J.S.; Helm, R.M.; Bannon, G.A. Mapping and mutational analysis of the IgE-binding epitopes on Ara h 1, a legume vicilin protein and a major allergen in peanut hypersensitivity. Eur. J. Biochem. 1997, 245, 334–339. [Google Scholar] [CrossRef] [PubMed]
  178. Shreffler, W.G.; Lencer, D.A.; Bardina, L.; Sampson, H.A. IgE and IgG4 epitope mapping by microarray immunoassay reveals the diversity of immune response to the peanut allergen, Ara h 2. J. Allergy Clin. Immunol. 2005, 116, 893–899. [Google Scholar] [CrossRef] [PubMed]
  179. Shreffler, W.G.; Beyer, K.; Chu, T.H.; Burks, A.W.; Sampson, H.A. Microarray immunoassay: Association of clinical history, in vitro IgE function, and heterogeneity of allergenic peanut epitopes. J. Allergy Clin. Immunol. 2004, 113, 776–782. [Google Scholar] [CrossRef]
  180. Matsumoto, N.; Okochi, M.; Matsushima, M.; Ogawa, A.; Takase, T.; Yoshida, Y.; Kawase, M.; Isobe, K.; Kawabe, T.; Honda, H. Development of peptide arrays for detection of IgE-binding epitopes in cow’s milk allergens. J. Biosci. Bioeng. 2009, 107, 324–330. [Google Scholar] [CrossRef] [PubMed]
  181. Jarvinen, K.M.; Chatchatee, P.; Bardina, L.; Beyer, K.; Sampson, H.A. IgE and IgG binding epitopes on alpha-lactalbumin and beta-lactoglobulin in cow’s milk allergy. Int. Arch. Allergy Immunol. 2001, 126, 111–118. [Google Scholar] [CrossRef] [PubMed]
  182. Lin, J.; Bardina, L.; Shreffler, W.G.; Andreae, D.A.; Ge, Y.; Wang, J.; Bruni, F.M.; Fu, Z.; Han, Y.; Sampson, H.A. Development of a novel peptide microarray for large-scale epitope mapping of food allergens. J. Allergy Clin. Immunol. 2009, 124, 315–322.e3. [Google Scholar] [CrossRef]
  183. Shin, D.S.; Compadre, C.M.; Maleki, S.J.; Kopper, R.A.; Sampson, H.; Huang, S.K.; Burks, A.W.; Bannon, G.A. Biochemical and structural analysis of the IgE binding sites on ara h1, an abundant and highly allergenic peanut protein. J. Biol. Chem. 1998, 273, 13753–13759. [Google Scholar] [CrossRef]
  184. Stanley, J.S.; King, N.; Burks, A.W.; Huang, S.K.; Sampson, H.; Cockrell, G.; Helm, R.M.; West, C.M.; Bannon, G.A. Identification and mutational analysis of the immunodominant IgE binding epitopes of the major peanut allergen Ara h 2. Arch. Biochem. Biophys. 1997, 342, 244–253. [Google Scholar] [CrossRef]
  185. Rouge, P.; Culerrier, R.; Sabatier, V.; Granier, C.; Rance, F.; Barre, A. Mapping and conformational analysis of IgE-binding epitopic regions on the molecular surface of the major Ara h 3 legumin allergen of peanut (Arachis hypogaea). Mol. Immunol. 2009, 46, 1067–1075. [Google Scholar] [CrossRef] [PubMed]
  186. Flinterman, A.E.; Knol, E.F.; Lencer, D.A.; Bardina, L.; den Hartog Jager, C.F.; Lin, J.; Pasmans, S.G.; Bruijnzeel-Koomen, C.A.; Sampson, H.A.; van Hoffen, E.; et al. Peanut epitopes for IgE and IgG4 in peanut-sensitized children in relation to severity of peanut allergy. J. Allergy Clin. Immunol. 2008, 121, 737–743.e710. [Google Scholar] [CrossRef] [PubMed]
  187. Reese, G.; Ayuso, R.; Leong-Kee, S.M.; Plante, M.J.; Lehrer, S.B. Characterization and identification of allergen epitopes: Recombinant peptide libraries and synthetic, overlapping peptides. J. Chromatogr. B Biomed. Sci. Appl. 2001, 756, 157–163. [Google Scholar] [CrossRef]
  188. Robotham, J.M.; Teuber, S.S.; Sathe, S.K.; Roux, K.H. Linear IgE epitope mapping of the English walnut (Juglans regia) major food allergen, Jug r 1. J. Allergy Clin. Immunol. 2002, 109, 143–149. [Google Scholar] [CrossRef]
  189. Fang, L.; Li, G.; Zhang, J.; Gu, R.; Cai, M.; Lu, J. Identification and mutational analysis of continuous, immunodominant epitopes of the major oyster allergen Crag 1. Clin. Immunol. 2019, 201, 20–29. [Google Scholar] [CrossRef] [PubMed]
  190. Zhang, Y.; Bhardwaj, S.R.; Vilches, A.; Breksa, A.; Lyu, S.C.; Chinthrajah, S.; Nadeau, K.C.; Jin, T. IgE binding epitope mapping with TL1A tagged peptides. Mol. Immunol. 2022, 153, 194–199. [Google Scholar] [CrossRef]
  191. Yang, L.; Kulis, M. Hypoallergenic proteins for the treatment of food allergy. Curr. Allergy Asthma Rep. 2019, 19, 15. [Google Scholar] [CrossRef]
  192. Niemi, M.; Jylha, S.; Laukkanen, M.L.; Soderlund, H.; Makinen-Kiljunen, S.; Kallio, J.M.; Hakulinen, N.; Haahtela, T.; Takkinen, K.; Rouvinen, J. Molecular interactions between a recombinant IgE antibody and the beta-lactoglobulin allergen. Structure 2007, 15, 1413–1421. [Google Scholar] [CrossRef]
  193. Padavattan, S.; Flicker, S.; Schirmer, T.; Madritsch, C.; Randow, S.; Reese, G.; Vieths, S.; Lupinek, C.; Ebner, C.; Valenta, R.; et al. High-affinity IgE recognition of a conformational epitope of the major respiratory allergen Phl p 2 as revealed by X-ray crystallography. J. Immunol. 2009, 182, 2141–2151. [Google Scholar] [CrossRef] [PubMed]
  194. Garcia-Ramirez, B.; Mares-Mejia, I.; Rodriguez-Hernandez, A.; Cano-Sanchez, P.; Torres-Larios, A.; Ortega, E.; Rodriguez-Romero, A. A native IgE in complex with profilin provides insights into allergen recognition and cross-reactivity. Commun. Biol. 2022, 5, 748. [Google Scholar] [CrossRef] [PubMed]
  195. Mitropoulou, A.N.; Bowen, H.; Dodev, T.S.; Davies, A.M.; Bax, H.J.; Beavil, R.L.; Beavil, A.J.; Gould, H.J.; James, L.K.; Sutton, B.J. Structure of a patient-derived antibody in complex with allergen reveals simultaneous conventional and superantigen-like recognition. Proc. Natl. Acad. Sci. USA 2018, 115, E8707–E8716. [Google Scholar] [CrossRef]
  196. James, L.C.; Roversi, P.; Tawfik, D.S. Antibody multispecificity mediated by conformational diversity. Science 2003, 299, 1362–1367. [Google Scholar] [CrossRef] [PubMed]
  197. NIH_Expert_Panel. Food Allergey: Report of the NIH Expert Panel on Food Allergy Research; NIH: Bethesda, MD, USA, 2006.
  198. Wildner, S.; Griessner, I.; Stemeseder, T.; Regl, C.; Soh, W.T.; Stock, L.G.; Volker, T.; Alessandri, C.; Mari, A.; Huber, C.G.; et al. Boiling down the cysteine-stabilized LTP fold—Loss of structural and immunological integrity of allergenic Art v 3 and Pru p 3 as a consequence of irreversible lanthionine formation. Mol. Immunol. 2019, 116, 140–150. [Google Scholar] [CrossRef]
  199. Koppelman, S.J.; Nieuwenhuizen, W.F.; Gaspari, M.; Knippels, L.M.; Penninks, A.H.; Knol, E.F.; Hefle, S.L.; de Jongh, H.H. Reversible denaturation of Brazil nut 2S albumin (Ber e1) and implication of structural destabilization on digestion by pepsin. J. Agric. Food Chem. 2005, 53, 123–131. [Google Scholar] [CrossRef] [PubMed]
  200. Van Boxtel, E.L.; van den Broek, L.A.; Koppelman, S.J.; Gruppen, H. Legumin allergens from peanuts and soybeans: Effects of denaturation and aggregation on allergenicity. Mol. Nutr. Food Res. 2008, 52, 674–682. [Google Scholar] [CrossRef]
  201. Tuppo, L.; Spadaccini, R.; Alessandri, C.; Wienk, H.; Boelens, R.; Giangrieco, I.; Tamburrini, M.; Mari, A.; Picone, D.; Ciardiello, M.A. Structure, stability, and IgE binding of the peach allergen Peamaclein (Pru p 7). Biopolymers 2014, 102, 416–425. [Google Scholar] [CrossRef]
  202. Rib-Schmidt, C.; Riedl, P.; Meisinger, V.; Schwaben, L.; Schulenborg, T.; Reuter, A.; Schiller, D.; Seutter von Loetzen, C.; Rosch, P. pH and heat resistance of the major celery allergen Api g 1. Mol. Nutr. Food Res. 2018, 62, e1700886. [Google Scholar] [CrossRef]
  203. Cabanos, C.; Katayama, H.; Tanaka, A.; Utsumi, S.; Maruyama, N. Expression and purification of peanut oleosins in insect cells. Protein J. 2011, 30, 457–463. [Google Scholar] [CrossRef]
  204. Smole, U.; Bublin, M.; Radauer, C.; Ebner, C.; Breiteneder, H. Mal d 2, the thaumatin-like allergen from apple, is highly resistant to gastrointestinal digestion and thermal processing. Int. Arch. Allergy Immunol. 2008, 147, 289–298. [Google Scholar] [CrossRef]
  205. Nesbit, J.B.; Hurlburt, B.K.; Schein, C.H.; Cheng, H.; Wei, H.; Maleki, S.J. Ara h 1 structure is retained after roasting and is important for enhanced binding to IgE. Mol. Nutr. Food Res. 2012, 56, 1739–1747. [Google Scholar] [CrossRef]
  206. Mattison, C.P.; Bren-Mattison, Y.; Vant-Hull, B.; Vargas, A.M.; Wasserman, R.L.; Grimm, C.C. Heat-induced alterations in cashew allergen solubility and IgE binding. Toxicol. Rep. 2016, 3, 244–251. [Google Scholar] [CrossRef]
  207. Mattison, C.P.; Grimm, C.C.; Li, Y.; Chial, H.J.; McCaslin, D.R.; Chung, S.Y.; Bren-Mattison, Y.; Wasserman, R.L. Identification and characterization of Ana o 3 modifications on arginine-111 residue in heated cashew nuts. J. Agric. Food Chem. 2017, 65, 411–420. [Google Scholar] [CrossRef]
  208. Verhoeckx, K.; Bogh, K.L.; Dupont, D.; Egger, L.; Gadermaier, G.; Larre, C.; Mackie, A.; Menard, O.; Adel-Patient, K.; Picariello, G.; et al. The relevance of a digestibility evaluation in the allergenicity risk assessment of novel proteins. Opinion of a joint initiative of COST action ImpARAS and COST action INFOGEST. Food Chem. Toxicol. 2019, 129, 405–423. [Google Scholar] [CrossRef] [PubMed]
  209. Fu, T.J.; Abbott, U.R.; Hatzos, C. Digestibility of food allergens and nonallergenic proteins in simulated gastric fluid and simulated intestinal fluid-a comparative study. J. Agric. Food Chem. 2002, 50, 7154–7160. [Google Scholar] [CrossRef] [PubMed]
  210. Bogh, K.L.; Barkholt, V.; Rigby, N.M.; Mills, E.N.; Madsen, C.B. Digested Ara h 1 loses sensitizing capacity when separated into fractions. J. Agric. Food Chem. 2012, 60, 2934–2942. [Google Scholar] [CrossRef] [PubMed]
  211. Scheurer, S.; Lauer, I.; Foetisch, K.; San Miguel Moncin, M.; Retzek, M.; Hartz, C.; Enrique, E.; Lidholm, J.; Cistero-Bahima, A.; Vieths, S. Strong allergenicity of Pru av 3, the lipid transfer protein from cherry, is related to high stability against thermal processing and digestion. J. Allergy Clin. Immunol. 2004, 114, 900–907. [Google Scholar] [CrossRef] [PubMed]
  212. Kenna, J.G.; Evens, R.M. Digestibility of proteins in simulated gastric fluid [Abstract]. Toxicologist 2000, 54, 141. [Google Scholar]
  213. Smith, F.; Pan, X.Y.; Bellido, V.; Toole, G.A.; Gates, F.K.; Wickham, M.S.J.; Shewry, P.R.; Bakalis, S.; Padfield, P.; Mills, E.N.C. Digestibility of gluten proteins is reduced by baking and enhanced by starch digestion. Mol. Nutr. Food Res. 2015, 59, 2034–2043. [Google Scholar] [CrossRef] [PubMed]
  214. Mayer, L. Mucosal immunity. Pediatrics 2003, 111, 1595–1600. [Google Scholar] [CrossRef] [PubMed]
  215. Husby, S.; Foged, N.; Host, A.; Svehag, S.E. Passage of dietary antigens into the blood of children with coeliac disease. Quantification and size distribution of absorbed antigens. Gut 1987, 28, 1062–1072. [Google Scholar] [CrossRef]
  216. Martinez San Ireneo, M.; Ibanez Sandin, M.D.; Fernandez-Caldas, E. Hypersensitivity to members of the botanical order Fabales (legumes). J. Investig. Allergol. Clin. Immunol. 2000, 10, 187–199. [Google Scholar] [PubMed]
  217. de Leon, M.P.; Drew, A.C.; Glaspole, I.N.; Suphioglu, C.; O’Hehir, R.E.; Rolland, J.M. IgE cross-reactivity between the major peanut allergen Ara h 2 and tree nut allergens. Mol. Immunol. 2007, 44, 463–471. [Google Scholar] [CrossRef] [PubMed]
  218. Yanagida, N.; Takei, M.; Saito, A.; Sato, S.; Ebisawa, M. Clinical cross-reactivity of wheat and barley in children with wheat allergy. Pediatr. Allergy Immunol. 2022, 33, e13878. [Google Scholar] [CrossRef]
  219. Smits, M.; Verhoeckx, K.; Knulst, A.; Welsing, P.; de Jong, A.; Gaspari, M.; Ehlers, A.; Verhoeff, P.; Houben, G.; Le, T.M. Co-sensitization between legumes is frequently seen, but variable and not always clinically relevant. Front. Allergy 2023, 4, 1115022. [Google Scholar] [CrossRef]
  220. Bublin, M.; Kostadinova, M.; Radauer, C.; Hafner, C.; Szepfalusi, Z.; Varga, E.M.; Maleki, S.J.; Hoffmann-Sommergruber, K.; Breiteneder, H. IgE cross-reactivity between the major peanut allergen Ara h 2 and the nonhomologous allergens Ara h 1 and Ara h 3. J. Allergy Clin. Immunol. 2013, 132, 118–124. [Google Scholar] [CrossRef]
  221. Warmenhoven, H.J.M.; Hulsbos, L.; Dreskin, S.C.; Akkerdaas, J.H.; Versteeg, S.A.; van Ree, R. IgE cross-inhibition between Ara h 1 and Ara h 2 is explained by complex formation of both major peanut allergens. J. Allergy Clin. Immunol. 2023. [Google Scholar] [CrossRef]
  222. EFSA Panel on Genetically Modified Organisms (GMO); Mullins, E.; Bresson, J.L.; Dalmay, T.; Dewhurst, I.C.; Epstein, M.M.; George Firbank, L.; Guerche, P.; Hejatko, J.; Naegeli, H.; et al. Scientific opinion on development needs for the allergenicity and protein safety assessment of food and feed products derived from biotechnology. EFSA J. 2022, 20, e07044. [Google Scholar] [CrossRef]
Figure 1. Structures of representative members of protein families that contain the majority of the known food allergens of plant origin. The name of the allergen or protein and the protein family/subfamily is indicated below every structure following the (AL) sequence label of the individual panels. The coordinates of the structures were downloaded from the WorldWide Protein Data Bank, and the graphics displays were generated using the CCP4MG program. The PDB codes for the structures are included in the figure labels along with the names of the allergens. Each structure is shown as a ribbon diagram with a blend-through coloring scheme displaying the N-terminal blue and the C-terminal red, except for the multimeric allergens Ara h 1 and Ara h 3 where the monomers were blended through different color ranges. Two panels of Ara h 1 are presented, with the right panel being the left panel rotated about a horizontal axis parallel to the paper pointing to the right. The side chains of cysteines that are involved in disulfide bonds are shown as ball-and-stick. Cysteines that are conserved in well-defined sequence motifs are labeled with their numbering in the motifs (see text).
Figure 1. Structures of representative members of protein families that contain the majority of the known food allergens of plant origin. The name of the allergen or protein and the protein family/subfamily is indicated below every structure following the (AL) sequence label of the individual panels. The coordinates of the structures were downloaded from the WorldWide Protein Data Bank, and the graphics displays were generated using the CCP4MG program. The PDB codes for the structures are included in the figure labels along with the names of the allergens. Each structure is shown as a ribbon diagram with a blend-through coloring scheme displaying the N-terminal blue and the C-terminal red, except for the multimeric allergens Ara h 1 and Ara h 3 where the monomers were blended through different color ranges. Two panels of Ara h 1 are presented, with the right panel being the left panel rotated about a horizontal axis parallel to the paper pointing to the right. The side chains of cysteines that are involved in disulfide bonds are shown as ball-and-stick. Cysteines that are conserved in well-defined sequence motifs are labeled with their numbering in the motifs (see text).
Foods 12 02232 g001
Figure 2. cC3C repeats of the variable leader peptide of known vicilin allergens. The domain structure of vicilin is shown at the top. Several features of the N-terminal variable region of vicilin between the signal peptide (Sig. P.) and the C-terminal cupin domains are shown below the domain structure. A question mark in the second column indicates the full sequence of the variable region of the allergen is not available. “ND” means no data available, and superscripts indicate the reference number of the cited literature. The variable leader peptides are derived by determining the signal peptides and the N-terminal peptides of the natural allergens. The sequences of the allergens were downloaded from the protein database at NCBI (https://www.ncbi.nlm.nih.gov/protein/, accessed on 1 April 2023). The signal peptides were predicted using SignalP 6.0 and the N-termini were those reported in the relevant references. The reference numbers are given as superscripts in the table cells. Almond vicilin was reported to be an allergen but does not have an Allergen Nomenclature Subcommittee designated allergen name.
Figure 2. cC3C repeats of the variable leader peptide of known vicilin allergens. The domain structure of vicilin is shown at the top. Several features of the N-terminal variable region of vicilin between the signal peptide (Sig. P.) and the C-terminal cupin domains are shown below the domain structure. A question mark in the second column indicates the full sequence of the variable region of the allergen is not available. “ND” means no data available, and superscripts indicate the reference number of the cited literature. The variable leader peptides are derived by determining the signal peptides and the N-terminal peptides of the natural allergens. The sequences of the allergens were downloaded from the protein database at NCBI (https://www.ncbi.nlm.nih.gov/protein/, accessed on 1 April 2023). The signal peptides were predicted using SignalP 6.0 and the N-termini were those reported in the relevant references. The reference numbers are given as superscripts in the table cells. Almond vicilin was reported to be an allergen but does not have an Allergen Nomenclature Subcommittee designated allergen name.
Foods 12 02232 g002
Figure 3. Sequence alignment of known GRP allergens. Cysteines are shown with a magenta background. The sequences of a large number of GRPs with a variable number of residues between the signal peptide and the first cysteine are also available (see text). Positions that tolerate amino acid changes are shown with a green background. Conserved residues in a sequence alignment included more GRPs are shown with a red background. The numbering of the cysteines and the disulfide bond connectivities are shown on top of the alignment. The sequences of the allergens were obtained following the allergen.org links and the alignment of the sequences was generated using Clustal Omega.
Figure 3. Sequence alignment of known GRP allergens. Cysteines are shown with a magenta background. The sequences of a large number of GRPs with a variable number of residues between the signal peptide and the first cysteine are also available (see text). Positions that tolerate amino acid changes are shown with a green background. Conserved residues in a sequence alignment included more GRPs are shown with a red background. The numbering of the cysteines and the disulfide bond connectivities are shown on top of the alignment. The sequences of the allergens were obtained following the allergen.org links and the alignment of the sequences was generated using Clustal Omega.
Foods 12 02232 g003
Table 1. Protein families with the most known food allergens from plant sources.
Table 1. Protein families with the most known food allergens from plant sources.
Protein Family aNo. of AllergensProteinsNo. of Allergens
PF00234: Tryp_alpha_amyl74Nsltp42
2S albumin26
Others6
PF00190: Cupin_13911S19
7S20
PF00235: Profilin27Profilin
PF00407: Bet_v_119PR-10
PF01277: Oleosin10Oleosin
PF00187: Chitin_bind_17 Chitinases
(PF00182: Glyco_hydro_19)
PF00314: Thaumatin6Thaumatin-like
PF02704: GASA6Gibberellin regulated protein
PF04702: Vicilin_N6N-terminal of vicilin cC3C antimicrobial protein
PF00112: Peptidase_C1
PF08246: Inhibitor_I293Cysteine protease
PF00197: Kunitz_legume3Protease inhibitors
PF00304: Gamma-thionin3Defensin
a The sequences of the allergen were obtained by following the links in the allergen.org database, and the protein families were determined by searching the Pfam/InterPro database (www.ebi.ac.uk/interpro, accessed on 14 April 2023).
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

Zhang, Y.; Che, H.; Li, C.; Jin, T. Food Allergens of Plant Origin. Foods 2023, 12, 2232. https://doi.org/10.3390/foods12112232

AMA Style

Zhang Y, Che H, Li C, Jin T. Food Allergens of Plant Origin. Foods. 2023; 12(11):2232. https://doi.org/10.3390/foods12112232

Chicago/Turabian Style

Zhang, Yuzhu, Huilian Che, Caiming Li, and Tengchuan Jin. 2023. "Food Allergens of Plant Origin" Foods 12, no. 11: 2232. https://doi.org/10.3390/foods12112232

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