Open access

Allergenicity of Milk Proteins

Written By

Simonetta Caira, Rosa Pizzano, Gianluca Picariello, Gabriella Pinto, Marina Cuollo, Lina Chianese and Francesco Addeo

Published: 12 September 2012

DOI: 10.5772/52086

From the Edited Volume

Milk Protein

Edited by Walter L. Hurley

Chapter metrics overview

3,696 Chapter Downloads

View Full Metrics

1. Introduction

Adverse reactions to food are currently classified into toxic and non-toxic reactions. There is a normal range of concentrations of naturally occurring toxic compounds, which can easily increase during food processing. For example, thermal processing can cause the unintended and undesirable formation of toxic compounds, such as acrylamide in fried potato chips and furan in sterilized canned vegetables, together with losses of certain nutrients. The incidence of non-toxic reactions depends on individual susceptibility to a specific food or food ingredient, although these reactions are often dose-dependent. The non-toxic types may be divided further into immune- and non immune-mediated reactions. The term “hypersensitivity” is used for immune-mediated reactions, and the term 'intolerance' is used for non immune-mediated reactions (Figure 1). Immune-mediated reactions may be IgE-mediated (i.e., allergy or type I hypersensitivity) or non-IgE-mediated, whereas food intolerance may be enzymatic, pharmacologic or undefined. The incidence of immune-mediated adverse reactions to foods has increased in recent decades. In healthy subjects, orally ingested dietary proteins induce antigen-specific systemic hyporesponsiveness, termed oral tolerance. This phenomenon is well described in animal models, although the mechanisms remain unknown. Abrogation of oral tolerance or failure to induce oral tolerance may result in the development of food hypersensitivity. Immune reactions that cause tissue damage may be mediated by four reaction types that were defined by Coombs and Gell [1] (Figure 2). Type I, or anaphylactic hypersensitivity, is mediated by the reaction of an antigen with specific IgE antibodies that are strongly bound through their Fc receptor (CD23, IgεR) to the surface of the mast cell. Crosslinking of Igε receptors by divalent hapten leads immediately to the release of mediators, for example, some cytokines [primarily interleukin 4 (IL4)] and histamine, which are both activities of eosinophil chemotactic factor (ECF) and neutrophil chemotactic factor (NCF) (Figure 2A). This type of hypersensitivity occurs within minutes of antigen exposure. Some reaction mediators produce local skin, gastrointestinal and respiratory tract manifestations, and a systemic allergic reaction to an allergen that is associated with a dangerously low blood pressure. Moreover, some of the mediators exhibit chemoattractant activity and induce the infiltration of neutrophils, eosinophils, macrophages, lymphocytes and basophils within 6-12 h after challenge. The localized late-phase inflammatory response may also be mediated partly by cytokines that are released from mast cells. In type II or antibody-dependent cytotoxic hypersensitivity, antibodies recognize antigens on the surface of specific cells or tissues. Once activated, the complement system can initiate a variety of responses that can lyse and destroy cells. Phagocytic and cytotoxic K cells, which have receptors for the Fc-part of IgG or an activation component of complement, i.e., C3b, may also destroy cells (Figure 2B). In type III, or immune complex-mediated hypersensitivity, the soluble antigen can activate the complement and deposited phagocytes. Leucocytes may release tissue-damaging mediators and activate phagocytes, culminating in tissue damage. The complex can also induce thrombin-mediated platelet aggregation and release a vasoactive amine (Figure 2C). Type IV reaction, or delayed-type hypersensitivity, arises more than 24 h after an encounter with the antigen. Type IV reactions are mediated by antigen-sensitized CD4+ T cells (T helper cells) that release inflammatory mediators [e.g., IL2 and interferon-y (lFN-y)], attract phagocytes to site of infection, activate an inflammation response and lyse invading cells (Figure 2D). Because lytic enzymes are secreted from the phagocytic cells into the surrounding tissue, localized tissue destruction can progress.

Figure 1.

Figure 2.

Types of Allergy Mechanisms according to Coombs and Gell [1].

The mechanisms underlying of allergic sensitization to food include genetic susceptibility, aberrant barrier functions of the skin epithelium and gut mucosa and dysregulation of immune functions. Despite a wide range of clinical manifestations, there are at least two common prerequisites for the development of a general food allergy (FA). First, intraluminal antigens must penetrate the mucosal barrier of the intestine. Second, the absorbed antigens must cause harmful immune responses [2].

Advertisement

2. Adverse reaction to milk

Non-toxic adverse reactions to milk are primarily caused by either lactose intolerance or milk allergy.

2.1. Lactose intolerance

Milk intolerance is due to the inherited lack of the specific enzyme, β-galactosidase that is required to hydrolyze lactose. For lactose malabsorption, the most common therapeutic approach excludes lactose-containing milk from the diet. To make yogurt edible, exogenous β-galactosidases that hydrolyze yogurt lactose or probiotics for their bacterial lactase activity are added. However, further studies are required to clarify the role probiotics play in lactose intolerance therapy, which includes considering their well-known beneficial effects on intestinal function, gas metabolism and motility [3]. A prolonged contact time between β-galactosidase and lactose delays the gastrointestinal transit time and chronic lactose ingestion to improve colonic adaptation. It is known that high concentrations of β-galactosidase are physiologically present in neonates, but a genetically programmed and irreversible decline of the activity occurs after weaning [4], which results in primary lactose malabsorption. The secondary hypolactasia because of intestinal mucosa brush border damage that increases the gastrointestinal transit time is a transient and reversible condition [5]. Bloating, flatulence, abdominal pain, the passage of loose and watery stools, excessive flatus and diarrhea are gastrointestinal symptoms of lactose intolerance [6]. However, lactose occurring in the colonic lumen does not necessarily produce gastrointestinal symptoms because of the variable amount of residual intestinal mucosal lactase activity that possibly digests lactose. The availability of recombinant β-galactosidase as an exogenous lactase has resolved problems concerning bacteria that release β-galactosidase during gastric passage. By this means, lactose is partially or fully degraded to glucose and galactose and is therefore easily eliminated by simple dietary adjustments that mediate the effects of lactose intolerance [7]. An accurate diagnosis of lactose intolerance can significantly reduce patient anxiety and avoid unnecessary examinations and treatments [8].

2.2. Milk allergy

Cow's milk allergy (CMA) is a complex disorder that implies an immunologically mediated hypersensitivity reaction with varying mechanisms and clinical presentations. The type I reactions appear to be the most common immune reaction to milk. However, the dominance of IgE reactions (Type I) may be an artifact as the reaction is easy to diagnose because of an immediate IgE measurement, whereas other reaction types are more difficult to diagnose. Non-IgE-mediated hypersensitivity has been increasingly diagnosed, and it is likely that several mechanisms operate in an individual patient. In clinical work, hypersensitivities are classified as IgE- and non-IgE-mediated or as immediate and late reactions based on the appearance of the first types of symptoms [9]. Cow’s milk is a member of the “Big-8” food allergens that include egg, soy, wheat, peanuts, tree nuts, fish and shellfish in terms of prevalence [9]. The incidence of CMA varies with age. CMA is prevalent in early childhood with reported incidences between 2 and 6% [10-12] and decreases into adulthood to an incidence of 0.1–0.5% [13-14]. It has been suggested that infants have milk allergies because milk is usually the first source of foreign antigens that they ingest in large quantities, and the infant intestinal system is insufficiently developed to digest and immunologically react to milk proteins. When milk is eliminated, the inflammation response is controlled. After several years, oral tolerance is developed, and milk can again be tolerated [15]. Most milk-allergic children are considered symptom-free by 3 years of age, but several studies have indicated that older children may also have immune reactions to milk. Children whose CMA has been diagnosed at an age older than 3 years do not tend to outgrow the problem. In adults, CMA is less common than lactose intolerance [16-17], even though it has been reported that approximately 1% of the adult population has milk-specific IgE antibodies. However, studies on CMA in adults are scarce. Little is known about the clinical symptoms, eliciting doses, and allergens involved. It has established that CMA in adults is rare but serious [18]. In a study by Stöger et al. [19], the main target organs in adult CMA were the skin and the respiratory tract. Gastrointestinal (mild to moderate) and cardiovascular (severe) symptoms were observed less often in adults compared with children. Milk allergies and hypersensitivity may be more common now than they were several decades ago. Further, the prevalence of atopic diseases has also increased in recent decades. Because genetic diversity has not changed over a short period, environmental factors are believed to have influenced the phenotype. Such factors may include increased air pollution, such as passive smoking, and dietary factors, such as the duration of breastfeeding, amount of antioxidants and the type of dietary fats (favoring saturated fat and n-6 fatty acids). Another approach is the hygiene hypothesis, which states that early exposure to microbial antigens may reduce the risk of having allergies [20]. For as long as milk allergy and hypersensitivity have been experienced, and still may be, these conditions may be misdiagnosed as a disease other than an allergy/hypersensitivity, particularly if symptoms are delayed. Classical IgE-mediated milk allergies with objectively recorded skin reactions may have been diagnosed easily, whereas hypersensitivity with subjective gastrointestinal reactions may have been diagnosed as lactose intolerance, irritable bowel syndrome or some other intestinal disorder. In adults, viral infection, antibiotic treatment or stress may alter intestinal integrity so that its balance is disturbed and the number of protecting agents, such as microflora and mucosal immunity, are altered [21]. This process may result in milk hypersensitivity. Over one-third of women with IgE-mediated reactions to milk proteins have been reported to exhibit their first symptoms of hypersensitivity during or shortly after pregnancy, and one-quarter reported the first symptoms during a period of severe emotional stress [22].

Advertisement

3. Milk allergens

No single major allergen has been identified in cow’s milk according to either challenge tests or laboratory procedures; [23]. Indeed, clinical challenge tests demonstrate that most CMA patients react to several protein fractions of cow’s milk and each allergenic protein may have several epitopes, which are widely spread along the molecules. The cow milk proteins prevalently implicated in allergic responses in children are the whey proteins α-Lactalbumin (α-La)(Bos d 4) and β-Lactoglobulin (β-Lg) (Bos d 5), in addition to the casein (CN) fraction (Bos d 8) [24-26]. In adults, the predominant allergen is CN, whereas sensitization to whey proteins is rare. Biochemical characteristic of allergenic cow milk proteins are reported in Table 1. Currently, milk allergen analyses are generally based on immunoassay methods, such as enzyme-linked immunosorbent assay (ELISA) [27]. Commercial test kits are available for the determination of CN, β-Lg, or total milk proteins (CN and β-Lg) [28]. Interlaboratory studies were performed to evaluate the reliability and reproducibility of these kits [29]. Antibody cross-reactivity has been reported for some milk proteins [30]. Confirmatory tests are required to corroborate ELISA detection results and improve the detection specificity of undeclared milk allergens. For the last several decades, mass spectrometry has become the dominant technology for the identification of peptides and proteins. The primary current approaches used for protein identification are top-down [31-33] and bottom-up [34] sequencing. Top-down sequencing involves introducing the intact protein into the gas phase. The protein is identified by measuring either the protein molecular weight or its fragmentation pattern using various techniques [35]. The bottom-up approach is more common. The sample is usually digested with an enzyme, such as trypsin, followed by accurate sequence analysis by tandem mass spectrometry (MS/MS) of the proteolytic fragments. For protein identification, an algorithm is used for database searching based on amino acid sequence assignments.

Table 1.

Chemical characteristics of cow’s milk proteins and their inclusion in the official list of allergens

3.1. Casein

The CN fraction is composed of four proteins αs1-, αs2-, β-, and κ-CN, in approximate proportions of 40%, 10%, 40%, and 10%, respectively, and αs1-CN is a major allergen according to IgE and T cell recognition data [36-39]. The specificity of the IgE response to the different purified CNs has been analyzed on 58 sera from patients allergic to whole-CN [40]. Multi-sensitization was observed, which most likely corresponded to a co-sensitization to the different CN components after the disruption of the CN micelles. The CN fractions cross-link to aggregates termed nanoclusters, which combine into micelles. In the new state of aggregation, there is a hydrophobic central part and peripheral hydrophilic one that include phosphorylation sites. Although the primary structure of CNs is known, the micelle structure remains relatively unknown. αs2-, αs1-, and β-CN bind calcium to form a core that is covered with a κ-CN layer. The κ-CN latter protects CN micelles from precipitation in the presence of milk constitutive calcium ions. During the last five decades, different models for the structure of the bovine CN micelle have been proposed. Walstra has summarized the common structural elements into a “ball-shaped model” [41]. According to Walstra, a well-accepted model considers the CN micelle as follows: (i) it is roughly spherical; although it does not have a smooth surface; (ii) it is built of smaller units, termed sub-micelles, which mainly contain CN and have a mixed composition; (iii) sub-micelles vary in composition and consist of two main types-one primarily consists of αs and β-CNs and the other primarily consists of αs- and κ-CNs; (iv) the sub-micelles can be linked together by small calcium phosphate clusters bridging them; (v) the sub-micelles aggregate until they have formed a micelle in which those with κ-CN are outside; and (vi) consequently, molecular chains of the C-terminal end of κ-CN protrude from the micelle surface to form a “hairy” layer that prevents any further aggregation of sub-micelles by steric and electrostatic repulsion. The hairy layer is also held responsible for the stability of the micelles against flocculation. Destabilization of CN micelles can be made by treating milk with milk clotting enzymes, which limitedly degrade κ-CN affording to CN micelle fusion and formation of a para-κ-CN aggregate. Finally, a variable proportion, up to 5% total, consists of a heterogeneous group of CNs termed γ-CNs (γ-CN) that result from the limited proteolysis of β-CN by plasmin, the native milk protease. Plasmin disperses in low amounts from blood to milk during secretion to generate CN fragments whose structure are shown in Figure 7 with peptides labeled according to the current nomenclature [42]. CNs are highly sensitive to proteolysis and do not maintain a unique folded conformation [43], which has led them to being termed rheomorphic. Because the folded structure is limited, heating does not generally change the structure and hence its IgE binding [44]. αs1-CN, αs2-CN, and β-CN can chelate Ca2+, Zn2+, and Fe3+, respectively. The four CNs share little sequential homology, but, despite this lack of homology, simultaneous sensitization is often observed.

3.1.1. αs1-CN

αs1-CN consists of major and minor polypeptides that include the same amino acid sequence but have different degrees of phosphorylation [45]. Allergenic epitopes were identified by Spuergin et al. [46] in αs1-CN regions 19–30, 93–98 and 141–150 as immunodominant epitopes. Some sequential IgE-binding regions were recognized at AA 17–36, 39–48, 69–78, 93–102, 109–120, 123–132, 139–154,159–174, and 173–194 using sera from nine older children (> 9 years old), and the epitopes AA 69–78 and 175–192 were recognized by 60% and 80% of sera, respectively, from older children [47]. A later study by Elsayed et al. [48] has demonstrated that the N- and C-terminal peptides, AA 16–35 and 136–155, respectively, have the highest human IgE-binding affinity, and AA 1–18 and 181–199 exhibited high binding to rabbit IgG. cDNA coding for αs1-CN from a bovine mammary gland cDNA library with allergic patients’ IgE Abs has been isolated. IgE epitopes of αs1-CN were determined with recombinant fragments and synthetic peptides that spanned the αs1-CN sequence using microarray components and sera from 66 cow’s milk-sensitized patients. The allergenic activity of recombinant αs1-CN and the αs1-CN-derived peptides exhibited IgE reactivity, but mainly the intact recombinant αs1-CN induced strong basophil degranulation. The results indicate that αs1-CN contains several sequential IgE epitopes, but the isolated peptides were less potent than the complete allergen in initiating effector cell degranulation. These results suggest that primarily intact αs1-CN or larger IgE-reactive portions thereof are responsible for IgE-mediated symptoms of FA [49].

3.1.2. αs2-CN

αs2-CN is the most hydrophilic of the four CNs because of an anionic group cluster. αs2-CN consists of two major and several minor components that exhibit varying levels of phosphorylation. αs2-CN contains two cysteines and forms disulfide-linked dimers. Using 99 synthetic decapeptides, 10 regions binding IgE from the sera were identified as allergenic, i.e., 31–44, 43–56, 83–100, 93–108, 105–114, 117–128, 143–158, 157–172, 165–188, and 191–200 [50]. Studies on the presence of αs2-CN epitopes 87–96 and 159–168 with weak binding to 145–154 and 171–180 with originate from individuals with persistent cow’s milk allergy. Patients with transient allergies exhibited only weak binding to αs2-CN peptides [51].

3.1.3. -CN

β-CN is the most hydrophobic component of total CN. Sequence variants are known because of both partial proteolysis and variant genes [42]. β-CN is less phosphorylated than αs1-CN and αs2-CN, with five potential phosphorylated sites located in the N-terminal region. Plasmin cleaves β-CN into γ1-, γ2-, and γ3-CN. Synthetic decapeptides were used to estimate the β-CN region that binds IgE of patients, and peptides 1–16, 45–54, 55–70, 83–92, 107–120, 135–144, 149–164, 167–184, and 185–208 were described as typical for patients with persistent CMA. Sera from eight young patients exhibited a simpler IgE pattern because the sera were bound to peptides 1–16, 45–54, 83–92, 107–120, and 135–144 with weak binding to residues 57-66 and the C-terminal region [52].

3.1.4. κ-CN

κ-CN consists of a major carbonate-free component and a mini-bonded polymer that ranges from dimers to octamers. κ-CN plays an important role in the stability and coagulation properties of milk [89]. κ-CN is most likely more structured than αs- and β-CN and contains specific disulfide bonds [53-54]. The β-CN can rearrange on heating [55]. κ-CN is sensitive to proteolysis and is hydrolyzed by chymosin to produce para-κ-CN and caseinomacropeptide in the cheese-making process. Allergenic potential sequences remain in cooked cheeses. κ-CN is essential to the stability of CN micelles [55]. Diagnosing patients with persistent cow’s milk allergies with the use of 80 overlapping synthetic decapeptides helped to identify some regions that bind to IgE from sera, specifically sequences 15–24, 37–46, 55–80, 83–92, and 105–116 [37].

3.2. Whey proteins

Whey proteins contain two major allergens, β-Lg and α-La, and minor constituents, e.g., lactoferrin, bovine serum albumin (BSA), and immunoglobulins. In this fraction, proteolytic fragments of CN and fat globule membrane proteins can occur.

3.2.1. β-Lg

β-Lg (Bos d 5) is the major whey protein in ruminant milks, comprising 50% of the total whey protein. β-Lg is found in the milk of other mammals but is missing from the milk from rodents, lagomorphs and humans. Notably, 13-76% of patients are found to react with β-Lg. β-Lg has a molecular weight of approximately 18 kDa [37] and belongs to the lipocalin superfamily. The β-Lg (Bos d 5) allergen is capable of binding lipids, including retinol, β-carotene, saturated and unsaturated fatty acids, and aliphatic hydrocarbons, and transporting hydrophobic molecules, which is an important function [56-57]. Under physiological conditions, β-Lg is an equilibrium mixture of monomeric and dimeric forms. The proportion of monomers increases after heating to 70°C. β-Lg contains five cysteine residues, of which four are engaged in intra-chain disulfide bridges. Because of the single unpaired cysteine, β-Lg predominantly exists as a stable dimer that tends to dissociate into monomers at a pH between 2 and 3 [58]. β-Lg is present in several variants, i.e., A, B, and C, that are found in the Jersey breed. β-Lg is sensitive to thermal processes. Ehn et al. [59] reported that heating β-Lg to 74°C and 90°C reduced IgE binding significantly. Heating to 90 °C reduced IgE binding more extensively. Chen et al. [60] reported that nearly 90% loss and denaturation of β-Lg are observed in processed milk and that high heat is the major cause of protein aggregation. Circular dichroism demonstrated no significant conformational changes at temperatures below 70°C for as long as 480 s. The rapid changes of β-Lg occurred between 80°C and 95°C. Fifty percent of the maximal changes could be reached within 15 s. Guyomarc’h et al. [61] reported that large micellar aggregates, 4 × 106 Da, are formed upon heating milk that contained 3:1 ratios of β-Lg and α-La together with κ-CN and αs2-CN. Proteolysis and use of monoclonal antibodies proved that β-Lg possesses many allergenic epitopes spread over the β-Lg structure [62]. Major human IgE epitopes for β-Lg amino acid fragments are composed of residues 41–60, 102–124, and 149–162; intermediate 1–8, 25–40, and 92–100; and minor 9–14, 84–91, 125–135, and 78–83 [63]. Similar IgE epitope regions (21–40, 40–60, 107–117, and 148–168) were reported for a rat model of β-Lg allergy [64,65].

3.2.2. α-La

α-La (Bos d 4) is a homologue of C-type lysozymes. It is a member of glycohydrolyase family 22 and Pfam family and weighs 14,186 Da in the mature form and between 15,840 to 16,690 Da for the glycosylated forms. α-La is stabilized by binding to calcium. Polverino de Laureto et al. [66] reported that α-La is cleaved by pepsin at pH 2 in the region of residues 34–57, which produces large fragments. Veprintsev et al. [67] reported that differential scanning calorimetry of α-La at pH 8.1 exhibited transitions at 20°C–30°C with calcium chelator ethylene glycol tetra-acetic acid and near 70°C with the addition of calcium. McGuffey et al. [68] investigated the heating effects of purified α-La and demonstrated that the extent of irreversible aggregation varies at temperatures between 67°C and 95°C. When milk is heated to 95°C, α-La denatures more slowly than β-Lg. The folded α-La structure is destabilized at low pHs with the formation of a molten globule[69]. The stability to denaturation is also strongly lowered by the reduction of disulfides [70]. Disulfide exchange can occur during thermal denaturation, which leads to the formation of aggregates [71]. Although evolved from a lysozyme [72], the function of α-La is to form a complex with galactosyltransferase, which alters the substrate specificity and increase the lactose synthase rate in milk. The galactosyltransferase and α-La is termed lactose synthase [73]. α-La alone does not have any catalytic activity as a lysozyme or a synthase. Several other properties of α-La and possible additional functions have been described [74], which include binding of several ligands and antimicrobial activity, both as a complete molecule [75] or as peptides [76]. The cytotoxic effects against mammalian cells have also been investigated [77]. The major component of α-La is unglycosylated. However, a mass spectrum of α-La contains at least 15 distinct peaks [78], and a minor glycosylated form (approximately 10%) results from asparagine 45 glycosylation. A study concerning the allergenic properties of α-La demonstrated that in 60% of the study patients, allergic sera were specific for intact α-La with only 40% binding to peptides obtained after tryptic hydrolysis. Residue 17–58 was the most frequently recognized in the sequence 59–93, 99–108, and 109–123 [79]. The linear epitopes were identified by using sera of patients suffering from persistent allergies and IgE to cow’s milk levels > 100 kU(A)/L. Serum IgE bound most strongly to peptides 1–16, 13–26, 47–58, and 93–102 [80].

3.3.3. Minor allergens

Bovine serum albumin (BSA) (Bos d 6), which is a heat-labile protein, is a major allergen in beef but a minor allergen in milk [81-84]. Accordingly, beef allergic individuals are at risk of being allergic to cow's milk and vice versa. BSA allergies account for 0-88% of sensitization events, whereas clinical symptoms occur in up to 20% of patients. BSA is one of the proteins most frequently involved in binding with circulating IgE [85-86]. Bovine immunoglobulins (Bos d 7) may be also responsible for clinical symptoms in CMA.

Advertisement

4. Post-translational modifications

In evaluating the allergenic potential of a protein, post-translational modifications of amino acid residues should be considered in addition to sequential and conformational IgE binding domains. Notably, such modifications may either generate additional IgE epitopes or induce changes in protein folding that affect IgE-protein interactions. Accordingly, recombinant allergens do not generally have the IgE-binding capacity of their natural counterparts, most likely because of a deficiency in the post-translational events [87]. Regarding milk proteins, selective phosphorylation of serine residues in all of the four CNs, O-glycosylation of threonine residues in κ-CN and N-glycosylation of asparagine residues in α-La have been long described.

4.1. Phosphorylation

The removal of phosphate groups from CNs significantly reduces the CN-binding capacity of IgE from patients who suffer from milk allergies, which indicates that at least part of anti-CN IgE is directed against CN domains that comprise a major phosphorylation site [88, 89]. It has been suggested that currently observed co- and cross-sensitization to the different CNs that are encoded by different genes and display few amino acid sequence homologies can be caused by the occurrence of common highly conserved major sites of phosphorylation, i.e., the Ca2+ binding CN sequence SerP-SerP-SerP-Glu-Glu that corresponds to αs1-CN 66-70, β-CN 17-21 and αs2-CN 8-12 and 56-60 [87]. Most likely, sensitization to milk is caused by a large release of phosphopeptides that are resistant to further degradation by digestive enzymes [90] during intestinal proteolysis of milk proteins. However, serine phosphorylation poorly affects the overall antigenic potential of individual CNs. Notably, antisera raised against native β- and αs2-CNs can recognize their targets after dephosphorylation or deletion of a major phosphorylation site [87]. Furthermore, polyclonal antisera that are produced in rabbits using a bovine β-CN 1-28 phosphorylated peptide as an antigen have been utilized to detect all of the tryptic phosphopeptides that originate exclusively from the 1-28 region of β-CN, regardless of the content of the phosphorylated Ser residues, and none of those generated by the other bovine CN fractions [91]. β-CN from human milk contains the phosphopeptide cluster 5Glu-Ser-Leu-SerP-SerP-SerP-Glu-Glu12, also found in the bovine β-CN sequence 14-21; however, a lower level of phosphorylation has been generally observed. For example, according to the phosphopeptide analysis of human milk that is reported in Table 2 and Figure 3, the 2092.8 Da component, which corresponds to β-CN(f2-18)2P, caused the third peak in intensity order in combination with the fully phosphorylated components. This lower phosphorylation level is lacking in its bovine counterpart. The overall higher degree of phosphorylation of bovine CNs can play a role in sensitizing humans to bovine milk.

4.2. Glycosylation

The effect of glycosylation on the allergenic potential of milk proteins has been long disregarded despite efforts to identify the domains responsible for the allergenicity of milk proteins, mostly based on an epitope mapping approach. Notably, the role of carbohydrate epitopes in initiating an allergic reaction is still unclear [92]. Potential glycosylation sites have been identified in major milk proteins, i.e., N45 and N71 of mature α-La [93] and T131, T133, T135, T136, T142, and S141 of mature κ-CN [94]. Approximately 10% of α-La has been found glycosylated at N45, giving rise to at least 14 distinct peaks by electrospray-ionization mass spectrometry analysis [95]. However, these glycosylated forms were not included among the IgE epitopes in a study because they were not detected by matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) analysis of α-La; notwithstanding in the same study, IgE reactivity of sera from patients allergic to α-La were proven to be sensitive to periodic acid treatment [93]. As reported in Table 3, the genetic variant A of water buffalo α-La that carries an N45→D45 substitution cannot be glycosylated. To assess the effect of glycosylation on the allergenicity of α-La, it might be used as substrate for IgE reactivity testing of sera from patients sensitized to α-La.

Table 2.

Identification of human milk soluble TCA 12% peptide fractions enriched on hydroxyapatite after MALDI-TOF analysis by FindPept (http://www.expasy.org/tools/findpept.html) software.

Despite some indications that the allergenic character has been identified in the glycosidic moiety of native κ-CN [96], at present this issue remains to be settled. Glycated forms of κ-CN account for approximately 40% of the κ-CN that normally occurs in bovine milk, but glycans are not randomly distributed among potential glycosylation sites. The hierarchy of glycan addition proceeds according to the order T131, T142, T133, whereas the other sites remain latent until these sites are occupied [97]. κ-CN is cleaved by chymosin during the primary stage of cheese making at the peptide bond F105-M106. The C-terminal 106-169 fragment, known as glycopeptide because all of the glucides originally present in κ-CN are retained, is released and lost in the whey. Therefore, cheese is devoid of any glycosylated major component. Potential allergenicity of κ-CN glycoforms has been suggested by analyzing the IgE binding capacity of an individual human serum from an adult atopic patient who had outgrown a cow milk allergy in early childhood. Bovine κ-CN has been selectively recognized by IgE immunostaining of an electrophoretic profile of milk proteins. No additional IgE-reactive proteins other than bovine κ-CN have been found in either bovine cheese, regardless of the cheese making technology and time ripening, or in ewe, goat and water buffalo milk. Moreover, chemical removal of glucide chains from bovine κ-CN has not impaired IgE binding, thus proving a primary involvement of the glycoside moiety of the protein in IgE recognition. According to the specificity displayed by IgE, N-acetylneuraminic acid as a terminal unit of a tetrasaccharide chain has been argued to be an IgE epitope [98].

Figure 3.

Mass spectrum of human milk soluble TCA 12% fraction enriched on hydroxyapatite by MALDI-TOF.

Table 3.

Position and amino acid differences among the genetic variant of α-La from four animal species.

Advertisement

5. “Allergenomics”

The application of proteomic methodologies for the analysis of food allergens has been termed “allergenomics” [99]. For “type I” FA, IgE-binding indicates that the target carries the risk to be an allergen. MS-based proteomic methods have identified many proteins and allergens. MALDI [100] and electrospray ionization (ESI) [101] and MS/MS sequencing are the techniques most widely used to produce high-quality spectra of post-translationally modified peptides [102-104] or intact proteins (see [105]). The characterization of glycosylated allergens has been partly overcome by specific enrichment using lectin or hydrophilic resins (HILIC) prior to MS analysis [106,107]. Native and de-glycosylated peptides are analyzed by MALDI or ESI-MS. Because of the difficulty of profiling oligosaccharides released by glycoprotein [108-110], glycan profiles are obtained after permethylation of the oligosaccharide chains according to the procedure of Das et al. [111]. By this means, glycosylated site(s) are identified together with the peptide backbone. Although widespread, several studies are dedicated to milk protein analysis for the detection of allergenic proteins or peptides in dairy products. Thus is determined by the concentration of allergenic compounds in food products that are often secondarily masked by dominant non-allergenic proteins. Among the various methods currently used to detect allergens in food products, immunochemical techniques that rely on antibody-binding properties have been developed. Commercially based kits are used for rapid screening, and enzyme-linked immunosorbent assays (ELISA) provide evaluations. Limits of detection (LOD) attained by ELISA tests are in the range of 1–5 ppm. Because the epitopes to be detected and their possible cross-reactivity with matrix components are unknown, detection reliability strongly depends on various factors that include the thermal changes of whey proteins, which are of primary importance. Furthermore, in several foods, linear epitopes can be released by parent protein hydrolysis, whereas retain their allergenic potential. As cited above, MS measurements can be finalized to evaluate the molecular mass of proteins and derived-peptides (MS1), determine the amino acid sequence and identify post-translational modifications (MS/MS or MSn). Two-dimensional electrophoresis (2DE) separates proteins according to the pI (i.e., first dimension, isoelectric focusing, IEF) and subsequently molecular weight (SDS-PAGE) in an orthogonal dimension. By this means, the separation of thousands of proteins has been achieved using highly specific stains that visualize specific protein classes, i.e., phosphoproteins, or nonspecific stains that simultaneously target total proteins without particular functional groups. Protein spots are localized, excised from the gel and subjected to an in-gel tryptic digestion. Mass spectrometry analysis either by MALDI reflectron TOF or microcapillary liquid chromatography MS-MS detects the proteins based on the expected masses of peptides available in databases and other plant proteins in pollen diffusates. More directly, tandem mass spectrometry is used to identify the peptide sequence and search for allergens in databases. More elegantly, allergens are localized on the gels after a one- or two-dimensional electrophoretic separation followed by a nitrocellulose transfer of the proteins (i.e., western blotting), which is stained with sera from allergic patients as a source of specific IgE. Combined with the analysis by mass spectrometry of electrophoretically separated allergens, immunoblotting is useful for the rapid determination of allergen identities. Allergen–IgE complexes are also detected using conjugated anti-human IgE as a secondary antibody. Once localized in a 2DE map, the allergen can be monitored using allergen specific antibodies [112,113]. Immune-reactive allergenic protein(s) are identified along the immunoblots by comparison with a reference electrophoretic map. All of the major milk proteins are allergen candidates because sera of allergic patients contain various percentages of immune-reactive proteins that are recognized by IgE [114]. The order of milk protein allergenicity is as follows: αs2-CN>αs1-CN>β-CN>κ-CN>β-Lg> serum albumin > IgG-heavy chain>Lf. This list contains allergenic proteins that have been identified by experiments on several MS platforms [115]. Thus far, it is quite difficult to find a separation method that can accommodate the diversity of proteins equally. Therefore, modern separation techniques have been performed off-line or by online ion-exchange/reversed phase liquid chromatography prior to MS analysis. ESI-MS is currently the interface most frequently used to perform an LC separation of intact proteins. The protein identification is most commonly achieved after a proteolytic digestion and molecular weight determination of the LC-separated peptides. “Shotgun” proteomics is the most effective LC/MS-based strategy because a trypsin-digested protein sample generates thousands of peptides that are subsequently separated by LC prior to MS/MS sequencing [103,116]. Proteins with at least one matching peptide are candidates to occur in the sample. However, with CNs, it is difficult to determine which fractions are present in the sample if they share the same set of phosphopeptides or have only one constituent peptide detected. LC-Q/TOF MS/MS has been used to detect wine CN as a fining agent. Two peptides were identified from αs1-CN and four peptides from the tryptic digestion of α- and β-CNs [117]. A similar strategy could be applied to monitoring allergens in processed milk products. Signature peptides could be identified as CN or whey protein allergens by submitting protein concentrates to trypsinolysis. In this manner, information on the molecular weight of the intact allergen is lost, but cross-reactive immunogenic peptides can be discriminated. Because of the higher sensitivity of MS in the detection of peptides, MS expands the dynamic range of the protein species detected. MS is a method for discovering “hidden” or traces of allergens. Proteomics has become pivotal to the development of modern structural immunology and to the understanding of interacting systems that are involved in immune responses, regardless of FA status.

5.1. Allergen quantification methods

Difference gel electrophoresis (DIGE) is utilized to compare multiple proteins in samples migrating in parallel in the same chamber. The proteins are labeled with three distinct fluorescent dyes on the same 2D gel and differentially visualized via fluorescence at different wavelengths. This methodology enables the detection of a differential presence of proteins and small differences in protein abundance. Allergens can be quantified by LC–MS. An accurate evaluation of the protein/peptide requires a suitable standard. In the direct quantification of intact proteins, the intensity of multi-charged analyte ions is compared with that of an internal or external standard. For example, quantification of cow’s milk allergens in fruit juice samples [118] and whey drink [119] was performed by simultaneously monitoring several multiple-charged ions of whey protein components. With a similar approach, internal standard β-Lg was used to quantify non-bovine β-Lg in different milk-derived products [120]. The use of “bottom-up” methods, such as SILAC, ICAT, and iTRAQ, for quantitative analysis in proteomics has progressively increased [121,122]. Quantification of allergens in complex samples requires simple and precise methods of analysis, such as selected reaction monitoring (SRM) [123]. SRM is presently considered the gold standard for absolute quantification, whereas multiple reaction monitoring (MRM) can monitor the masses of selected signature peptides. For an allergen evaluation, internal reference peptides are required for food product monitoring by LC–SRM MS [124]. Hydrolyzed protein samples are spiked with known amounts of synthetic peptides and monitored by LC–MS in the SRM mode. Absolute amounts of peptide(s) are determined by the ratio of the ion intensities of natural and synthetic peptides (Figure 4).

Figure 4.

Mass Spectrometry procedure analysis for protein absolute quantification. Proteotypic peptides are selected with a preliminary fullscan. For quantitative analysis whole protein extracts are trypsinized and the peptide mixture is spiked with external standard peptide. Proteotypic peptides (blue colored) selected as analytical probes of the target protein(s), are quantified by comparing the ionic intensities.

This strategy has demonstrated its validity for using signature peptides as analytical surrogates to measure allergens in crude protein extracts. One advantage of the SRM procedure is the possibility of one-step monitoring of a variety of allergens. Recently LC–triple quadrupole MS operating in an MRM mode has been effectively demonstrated to simultaneously detect allergens from seven different potentially allergenic matrices, such as milk, eggs, soy, hazelnuts, peanuts, walnuts and almonds. The detection limits were in the 10-1000 µg/g range. However, prior knowledge of allergens was required to monitor the most suitable allergenic peptides [125]. Based on the above-specified considerations, allergen evaluation requires the following: (i) allergen extraction from the food; (ii) enzyme proteolysis, usually trypsinolysis; and (iii) identification of signature peptides that are characteristic of food proteins or food ingredients. The signature peptides should be determined experimentally by prior LC/MS analysis of food-derived digested protein extracts.

5.2. Standardization of allergen preparations

Although the search for clinically relevant allergens has progressed, the characterization of allergens still requires studies on milk proteins as starting material. Pure native and recombinant allergens are needed as reference materials to calibrate methods among different laboratories. Recently, a panel of 46 food plant and animal allergens [126,127,128] has been made available within an EU-funded research project. In a recent 2DE application, calibration has been utilized for microbial complex protein systems using data obtained by MS [129]. Developing more allergen standards could be realized in the near future. Moreover, the search for allergenic sequence stretches would comprise only those immuno-dominant produced during digestion that can to translocate the gut barrier and reach the mucosal immune system. Among the digestion/adsorption models of food protein stability, pepsin digestion has been included in the Food and Agriculture Organization/World Health Organization to assess food safety [130]. A model study has established milk-derived peptide candidate-mediated resistance to proteases to display allergic effects. The survival of milk protein epitopes [131,114] requires structure determination. To this end, a simulated digestion of bovine milk proteins in vitro that includes the sequential use of pepsin, pancreatic proteases, and extracts of human intestinal brush border membranes, has allowed the identification of produced peptides by MS. The presence of characteristic β-Lg resistant peptides could implicate β-Lg in the case of a cow’s milk allergy [132]. The identification by MS of peptides arising from simulated digestion is complicated by a lack of enzyme specificity. Currently, there is no treatment to fully resolve or provide long-term remission from FA allergies. The research for therapy is mainly focused on the introduction of anti-IgE antibodies and specific oral tolerance induction. Immunotherapy appears to be an attractive approach; however, the risk of anaphylaxis should be considered. To this purpose, engineered proteins have been designed, i.e., anaphylaxis-initiating epitopes have been removed within these proteins, while preserving the tolerance-inducing epitopes [133-135]. It appears clear that to successfully pursue similar strategies, the precise identification of epitopes is necessary. It is expected that such approaches will be extended to an increasing number of food sources, whereby MS will play a key role for characterizing novel protein entities. The accurate characterization of the offending sequences could also be the starting point for developing less allergenic food products through the use of enzymatic, microbiological and technological processes to effectively remove allergens [136–139].

Advertisement

6. Dairy research versus CMA

6.1. Milk and dairy products from mammals different from cow

According to the current clinical approach to FA and intolerance based on an elimination diet, the treatment of choice is complete avoidance of cow milk. Although of moderate importance in an adult diet, cow milk elimination has a significant nutritional significance in the infant diet, especially during early childhood. Milk from other mammals has been suggested as a possible alternative to cow milk. At first, goat milk had been proposed as a hypoallergenic infant food or cow milk substitute in human diet, but much of this thesi has no credible scientific evidence. Despite the immunological cross-reactivity between cow and goat milk proteins, due to the close biochemical similarity associated with the same phylogenetic origin [140], it has been estimated that from 40 to 100% of patients allergic to cow milk proteins can tolerate goat milk intake [141]. However, clinical and immunochemical studies aimed at evaluating goat milk safety for cow milk allergic subjects have demonstrated that goat milk cannot be a substitute for cow milk without risk of anaphylactic reactions [142,143]. It has been suggested that evidence for goat milk tolerance in clinical trials can be due, at least in part, to a higher number of genetic polymorphisms in goat CNs, especially for αs1-CN [144]. Null or reduced expression of αs1-CN in individual goats; consequently, the overall αs1-CN content in goat bulk milk is lower than that found in cow bulk milk. According to this general finding, and taking into account that little β-Lg persists in cheese, fresh cheese produced from raw milk has been suggested to be a promising hypoallergenic protein source [145]. Unexpectedly, water buffalo milk yogurt has successfully been employed as an alternative food for children with cow milk allergies [146] despite homologous proteins from cow and water buffalo milk [147]. In contrast, several studies have reported the existence of allergies to goat and sheep milk [148-151] and cheese [152,153] in patients with tolerance to cow milk proteins. Overall, this type of allergy is less common and occurs later than that initiated by cow milk proteins, which is likely because goat and sheep dairy products are not usually included in an infant diet. Moreover, IgE epitopes have been widely recognized in the CN components of goat and sheep milk. Differences in the degree of CN phosphorylation, on average lower in goat and sheep milk than in cow milk, rather than differences in IgE epitope sequences, may be involved in initiating selective allergies to goat and sheep milk, as observed in recent cases. In addition to the four ruminant species of dairy interest (i.e., cow, water buffalo, sheep and goat), other monogastric mammals produce milk for human consumption, such as mares and donkeys. Mare milk, which is more similar to human than cow milk, has been proven to be an acceptable substitute of cow milk for children with severe IgE-mediated cow milk allergy; although the evidence of its tolerability by a supervised oral challenge test is recommended [154]. However, mare milk availability is limited, and its collection is difficult. Donkey milk provides nutritional adequacy and excellent palatability similar to that of mare milk but is more readily available. The composition of donkey milk is more similar to human milk than cow milk because of the higher lactose content (6.5 vs. 5 g/100 mL), lower protein content (1.2 vs. 3.2 g/100 mL), lower CN/whey protein ratio (approximately 1 vs. 4) and a higher non-protein nitrogen fraction level (0.29 vs. 0.18%) [155]. These features have prompted clinicians to propose donkey milk as a valuable breast milk substitute. Additionally, donkey milk intake has demonstrated positive effects in the diet therapy of patients allergic to cow milk proteins [156]. Although the mechanism of this tolerance is unclear, the reduced allergenic properties of donkey milk can be related to the structural differences compared with bovine milk. Because of scientific and clinical interest in donkey milk, characterization of the whey protein fraction [157], caseome [158], and the minor protein components [159] of donkey milk have been recently provided. Presently, milk from mammals with a geographically restricted distribution area, such as reindeer living in Northern Europe, has been utilized to overcome immunological cross-reactivity among proteins from mammals other than cows. In particular, β-Lg from reindeer milk, although belonging to the lipocalin family and similar to its homologous bovine protein, lacks several IgE epitopes of bovine β-Lg that are involved in CMA [160]. Recently, camel milk, mainly available in the Gulf area and Mauritania, is of growing interest to both nutritionists and pediatricians because of its high nutritive value and unique electrophoretic protein patterns, which strongly suggest a different immunological reactivity of camel milk proteins with respect to the bovine counterparts [161].

6.2. Gut microflora

It has been suggested that gut flora may be involved in the etiology of atopic diseases. It has been demonstrated that the gut microflora differs in children with high or low rates of allergy. Commensal gut flora play a role in inducing an oral tolerance, and the importance of the intestinal microbiota in developing food allergies is essential at early ages when the mucosal barrier and immune system are still immature. Probiotics interact with the mucosal immune system by the same pathways as commensal bacteria. A recent study has demonstrated that probiotic bacteria induced in vivo increased plasma levels of IL-10 and total IgA in children with allergic predisposition. Many clinical studies have reported significant benefits by probiotics supplementation in FA prevention and management. However, not everyone agrees on the effectiveness of probiotics supplementation. The differences are most likely related to the selected populations and probiotic strains used. The hygiene hypothesis proposes that disturbances in the gastrointestinal microbiota are associated with increased prevalence of allergic and autoimmune diseases [162]. Changes in the establishment of gut microbiota have been observed in Western infants [163,164]. This is most likely because of improved hygiene and cleanliness in Western countries and excessive use of antibiotics, which causes a reduced bacterial stimulus. Several clinical studies have reported differences in the composition of bacterial communities in the feces of children with and without allergic diseases. Many of those studies have highlighted the involvement of Bifidobacterium and Bacteroides in the protection against the development of atopy [165-168], but this observation remains a matter of debate [162]. Moreover, the mechanisms underlying such protective effects remain elusive. There is increasing evidence that T-regulatory cells derived from the thymus or induced in the periphery including the gut mucosa [169,170] are key players of immune regulation [171-173]. Using a single strain mouse model and defined bacterial communities and conventional mice, it has been recently demonstrated that the gut microbiota plays a protective role against allergen sensitization and allergic response in a mouse model of FA [174]. The difference between healthy and allergic children may be in their microflora. At 3 weeks of age, infants in whom atopy developed then had more Clostridia and fewer bifidobacteria in their feces compared with infants who remained healthy. Moreover, fecal bifidobacteria microflorae were different between healthy and allergic children; the healthy infants' microflora was mainly Bifidobacterium bifidum, whereas the microflora was mainly Bifidobacteria adolescentis in the allergic infants. It can be hypothesized that individual species, rather than an entire genus, can affect the manifestation of allergy. In a recent study, the microflorae of milk-hypersensitive and control adults before and after a 4-week supplementation with probiotic bacteria (Lactobacillus rhamnosus GG, ATCC 53103) have been studied. The anaerobic microflora before supplementation was comparable between the healthy and hypersensitive subjects, whereas the response after supplementation was different. The number of bifidobacteria in the healthy subjects increased significantly after supplementation. However, this did not occur with the supplementation in milk-hypersensitive subjects; this may be because of altered intestinal integrity. However, other studies have suggested a beneficial effect of probiotic bacteria in milk-hypersensitive subjects. In one study, symptoms of hypersensitivity abated along with an elimination diet in 28% of the patients in 4 years. It can be hypothesized that with milk elimination and long-lasting probiotic treatment, the intestinal severity of IgE-mediated hypersensitivity reactions may increase the intestinal microflora or even eliminate them.

6.3. Reduction of allergenicity of milk proteins by hydrolysis

The main objective of the milk industry is to supply products while preserving both the nutritive value and safety against developing allergies. Nutritional value is preserved by exposing liquid or powdered milk to low heat treatments to reduce heat susceptible amino acid side chain modifications and preserve the integrity of triacylglycerols, native vitamins and other milk components. As noted above, infants can develop milk allergies because of increased gut permeability to large molecules, in addition to other causes [175]. This result is supported by measurements of unmodified proteins or partially modified proteins in the sera of infants and adults [176]. Milk proteins have a molecular mass between 14 and 80 kDa. To reduce allergenicity, milk proteins can be submitted to different hydrolysis procedures. Attempts to classify products by protein hydrolysis include “extensive” or “high degree” hydrolysis and “partial” or “low degree” hydrolysis. The rationale of such a classification is the spectrum of peptide molecular weights or the ratio of α amino acids to total nitrogen. For quality assurance, in vitro product characterization requires size measurements of the peptides that are generated by protein hydrolysis and then an in vivo allergenicity determination. The in vivo step would include evaluating immunogenic or allergenic effects in a recipient infant. The European Union regulates that infant formulas contain immunoreactive proteins in quantities lower than 1% of nitrogen compounds [177] to reduce allergenicity knowing that only pure amino acids are strictly non-allergenic. This criterion could be encountered by milk proteins that have undergone extensive hydrolysis partially to cleave amino acids [178-180]. In contrast, formulas with moderately reduced allergenicity (partially hydrolyzed) are not recommended for the treatment of allergies because of the high amounts of residual allergens [181]. The low quantity of native proteins

or residual high molecular mass peptides may produce adverse effects in highly sensitive patients. Therefore, a milk hydrolysate can be considered safe and non-allergenic if the nitrogen fraction does not contain unmodified milk proteins or high molecular mass peptides [178-180]. In the latter case, the product could be classified in the “low degree” protein hydrolysate category. However, the antigenic properties of protein hydrolysates may not be dependent on the molecular size of the peptide components alone [182]. By comparing protein structures with known allergens and allergen epitopes, protein allergenicity has been predicted [183]. Although this is true for crystallized proteins, such as α-La, β-Lg and Lf, this procedure cannot be applied to uncrystallized CNs. Because infants who are diagnosed for milk protein allergies must ingest foods that exclude the causal protein, including those ingested by the mother and filtered in breast milk, extensively hydrolyzed milk formulas are used for the development of appropriate dietary and management strategies. Preclinical testing of infant formulas is necessary to characterize the molecular properties and residual antigenicity of proteins [184-186]. Stringent criteria specify that extensively hydrolyzed CN with a molecular weight below 5000 Da should be reduced by at least 99.99%. There is a need for accurate diagnostic methods to confirm the amount of extensively hydrolyzed CN. Milk for allergic infants would consist of extensively hydrolyzed CN and whey proteins of which at least 99.99% of the hydrolysis products have molecular weights below 5000 Da. The crucial criterion is for the level of allergens to be sufficiently low as to cause no significant reaction, even in infants who are highly allergic to cow's milk. There have been no reports of adverse reactions because of whey [187, 188] and CN hydrolysates [189-191]. Therefore, caution must be maintained that milk formulas destined to infants with milk allergies contain correctly hydrolyzed proteins. This generic indication requires that molecular properties and residual antigenicity of proteins would be characterized [184-186]. In vitro incubation of milk proteins with pepsin, trypsin, and chymotrypsin causes the cleavage of numerous peptides of various sizes. Bacterial, fungal and plant proteases may also act as hydrolyzing agents. Various enzyme combinations, such as alcalase, pancreatin and enzymes from fungal sources, have been utilized to produce protein hydrolysates. Commercial hypoallergenic products are currently characterized by an average degree of hydrolysis (DH) of the protein components. The DH 19 milk protein value is calculated from the increase of the number of primary amino groups compared to that of native proteins. In practice, the DH value could vary from 1 to 100% in the case of total hydrolysis of the proteins. In the case of partly hydrolyzed proteins, intact and partly hydrolyzed proteins are visible bands along an electrophoretic pattern of the products. Two commercial formula preparations with DH values of 6.3 and 1.3% contained some intact β-Lg and peptides with an Mr between 6000 and 8000 Da [192]. Using gel permeation chromatography, quantitative results on peptides with an Mr larger than 10 kDa were obtained [192]. Regardless of the technique used, descriptive information was obtained on either the molecular mass or the origin of the peptides. As a result, consumption of infant formulas by allergic patients cannot be attributed to one specific protein or high molecular mass peptide. To suppress or reduce the antigenicity of peptides, natural enzyme cleaving

of many or most of the peptide bonds is required. In this manner, epitopes that determine the antigenicity of the protein molecules are destroyed. This result proves that evaluation of the adequacy of infant formula composition in preventing or delaying antigenicity is not based solely on DH or Mr determination. In highly sensitized infants with IgE-mediated cow's milk allergies, life-threatening anaphylactic shock usually develops shortly after the consumption of claimed hypoallergenic milk products in which a number of epitopes would have survived in the highly proteolytic environment. The possibility of using well-characterized monoclonal antibodies in ELISA tests can be used for assessing the origin of immunoreactive bovine milk proteins. Because the clinical significance of residual antigenicity requires prior molecular approaches, hypoallergenic products may first be screened for peptide identification in hydrolyzed milk products. Although protein hydrolysates can provide a positive effect, they can contain undefined peptide components, which are undesirable for pharmaceutical production purposes. In many cases, hydrolysates are produced by methods that are not well-controlled. Other complications arise from the raw starting material and differences in processing that lead to lot-to-lot hydrolysate composition variability. For these reasons, constant chemically defined products are needed. The data presented here represent the initial steps that have been taken to identify peptides treated with pepsin (P) and trypsin (T) and were used in succession to hydrolyze commercial milk protein powder (PT hydrolysate). To mimic commercial milk hydrolysates, the protein powder was treated with enzymes after a thermal shock deactivation treatment. Subsequently, the peptides are identified. RP-HPLC fractionation was used to aid with the peptide separation. In Figure 5, the hydrolysate was analyzed to demonstrate a correlation between proteolytic enzymes and the presence of peptides. Commercial milk protein hydrolysates may contain trace amounts of allergenic proteins whose molecular weights were determined by MALDI-TOF analysis. Among the number of peptides present in the hydrolyzed sample, the proteins/peptides exhibited molecular masses less than or equal to 2431 Da (Figure 6). This means that the CNs and whey proteins were digested by pepsin and trypsin into peptides with masses less than 3000 Da. This type of hydrolysate is not expected to elicit allergic reactions in already sensitized allergic patients (neither anaphylactic shock nor positive passive cutaneous anaphylaxis), as verified in experimental animals [193]. LC-ESI-MS/MS analysis was performed on the hydrolysate to identify peptides occurring herein. No sequence peptides with 3 or 4 residues were detected because the MALDI signals were acquired at a mass gate of m/z 400. Because some short peptides were in the hydrolysate, milk proteins were hydrolyzed by P and T into oligopeptides with different biological activities. In Figure 7, the amino acid sequence of the four bovine CN fractions and β-Lg are reported with a subscript that indicates the number of amino acid residues in MS/MS-identified fragment.

The proteins in the milk powder sample, which contained modified amino acid residues that may indicate the quality of the protein in milk powder, were not examined within the present work.

Figure 5.

MALDI spectra of a sample of milk powder before and after sequential hydrolysis with pepsin and trypsin. A search for residual intact proteins and high molecular derived peptides (a) and measurements of molecular mass value of oligopeptides in the mixture (b). No peptide at a molecular mass higher than 3430.97 Da was observed in the MALDI spectrum.

Figure 6.

MALDI analysis of a milk sample after sequential hydrolysis with pepsin and trypsin. A molecular mass value corresponds to that of oligopeptides in mixture. No peptide at a molecular mass higher than 2207.1 Da was observed in the MALDI spectrum.

Advertisement

7. Oral immunotherapy as future perspective in CMA management

The primary treatment for managing food allergies is eliminating the offending food from the diet. In the case of milk, it is extremely difficult to achieve complete elimination because milk can be masked in any number of foods, which may lead to unwanted severe reactions. The natural course of a cow’s milk protein allergy is the acquisition of tolerance spontaneously through an elimination diet, and 85% of patients overcome CMA by the time they are 4-5 years old [195-198]. In recent years, a number of studies have been published regarding desensitization or oral tolerance to food antigens, particularly to cow’s milk. Major advances in understanding the immunological processes involved in the development of CMA have revealed a considerable number of allergenic epitopes and the heterogeneity of allergic responses. Importantly, an elimination diet of dairy foods has negative consequences in terms of inadequate calcium and vitamin intake. In the literature, several conflicting studies have reported on possible desensitizing therapies in the treatment of FA allergies. The possibility to obtain an oral desensitization is now gaining acceptance widely, even if the mechanism is still unclear. Oral tolerance or desensitization is the active non-response of the immune system to an antigen through sublingual, oral administration. Tolerance or the long-term loss of allergic reactivity follows a desensitization treatment. In the literature, several studies have reported on possible physiopathogenetic mechanisms of oral desensitization, but the exact mechanism is still unknown. That tolerance may be involved in the mechanism of desensitization is still uncertain [199-201]. A growing understanding of the molecular and cellular mechanisms of oral tolerance is reinforcing advances in potential therapies for food allergies and is pivotal to eventually curing allergies in sensitized individuals. Oral desensitization should be taken into consideration in the management of food-allergic patients even if the physiopathogenetic mechanisms are still unexplained. Moreover, this treatment should be considered, particularly for children, because elimination from the diet of some foods (e.g., milk and eggs) for these patients could cause psychological and/or nutritional problems. Oral immunotherapy (oral desensitization) may be a promising treatment strategy for cow’s milk allergy in children and valuable for other foods, such as eggs or peanuts. Although the mechanisms of IgE-mediated allergies are fairly well understood, the immunology and variety of non-IgE-mediated reactions remains largely unknown. A better understanding of these allergy mechanisms is a prerequisite to the development of improved diagnostics, which in turn will facilitate an improved understanding of the epidemiology of CMA, particularly for non-IgE-mediated reactions.

Advertisement

8. Conclusions

A better understanding will also aid the development of hypoallergenic dairy products, especially for adults with CMA for whom there is currently a dearth of suitable low-allergenic dairy products. Some of the risk factors for developing CMA have been identified; a familiar history of atopy is one of the main determinants. However, the mechanisms of allergic sensitization and the precise interactions between genetics and various environmental factors that lead to CMA remain unclear. The first few months of life, during which the immune system is still maturing, appear to be a critical risk period for allergic sensitization. For at-risk infants with at least one atopic parent, breastfeeding during this period is currently the best identified preventative strategy; the use of hydrolyzed formulas is recommended for babies who cannot be breastfed. The use of immunomodulatory dietary adjuvants, such as probiotics, is an emerging approach with considerable promise for primary prevention. For CMA sufferers, the avoidance of dietary milk proteins remains the only effective management strategy but carries with it nutritional implications, particularly for adequate vitamin and calcium intake as well as protein and energy where unorthodox alternative diets are implemented. Increasing knowledge of the molecular and cellular mechanisms of oral tolerance reinforces the advances in potential FA therapies and is pivotal to eventually curing allergies in sensitized individuals. Unraveling the links between innate and adaptive immunity and characterizing the roles of dendritic cells and T cells in directing immune responses and homeostasis to environmental antigens are likely to remain a focus of fundamental FA research in the coming years.

Acknowledgement

The Authors gratefully acknowledge the American Journal Experts Association for the text revision (http://www.journalexperts.com/). This work was partly supported by the financial aid to C.L. from MIUR, Program PRIN-2008 HNHAT7-004.

References

  1. 1. GellP. G. H. R. R. A.CoombsThe classification of allergic reactions underlying disease. In: Coombs, R.R.A., Gells, P.G.H. (Eds.) Clinical Aspects of Immunology. Blackwell, Oxford 1963
  2. 2. JohanssonS. G. O.BieberT.DahlR.FriedmannP. S.LanierB. Q.LockeyR. F.MotalaC.OrtegaMartell. J. A. T. A. E.Platts-MillsRing. J.ThienF.Van CauwenbergeP.WilliamsH. C.Revised nomenclature for allergy for global use: Report of the Nomenclature Review Committee of the World Allergy Organization, October 2003The Journal of Allergy and Clinical Immunology 2004;1135832836
  3. 3. MontaltoM.CuriglianoV.SantoroL.VastolaM.CammarotaG.MannaR.GasbarriniA.GasbarriniG.Management and treatment of lactose malabsorption. World Journal of Gastroenterology 2006122187191
  4. 4. WangY.HarveyC. B.HolloxE. J.PhillipsA. D.PoulterM.ClayP.Walker-SmithJ. A.SwallowD. M.The genetically programmed down-regulation of lactase in children. Gastroenterology 1998114612301236
  5. 5. LabayenI.ForgaL.GonzálezA.Lenoir-WijnkoopI.NutrR.MartínezJ. A.Relationship between lactose digestion, gastrointestinal transit time and symptoms in lactose malabsorbers after dairy consumption. Alimentary Pharmacology & Therapeutics 2001154543549
  6. 6. Shaw AD, Davies GJ. Lactose intolerance: problems in diagnosis and treatment.Journal of Clinical Gastroenterology 1999283208216
  7. 7. RosadoJ. L.SolomonsN. W.LiskerR.BourgesH.Enzyme replacement therapy for primary adult lactase deficiency. Effective reduction of lactose malabsorption and milk intolerance by direct addition of beta-galactosidase to milk at mealtime. Gastroenterology 198487510721082
  8. 8. Swagerty DL Jr, Walling AD, Klein RM. Lactose intolerance.American Family Physician 200265918451850
  9. 9. Crittenden RG, Bennett LE. Cow’s milk allergy: a complex disorder.Journal of the American College of Nutrition 2005Suppl) 582S-91S.
  10. 10. HillD. J.HoskingC. S.ZhieC. Y.LeungR.BaratwidjajaK.IikuraY.IyngkaranN.Gonzalez-AndayaA.WahL. B.HsiehK. H.Thefrequency.offood.allergyin.AustraliaAsiaEnvironmental Toxicology and Pharmacology 1997
  11. 11. BMExlFritsché. R.Cow’s milk protein allergy and possible means for its prevention. Nutrition 2001
  12. 12. Garcia-AraM. C.Boyano-MartinezM. T.Diaz-PenaJ. M.Martin-MunozM. F.Martin-EstebanM.Cow’s milk-specific immuno-globulin E levels as predictors of clinical reactivity in the follow-up of the cow’s milk allergy infants. Clinical & Experimental Allergy 2004346866870
  13. 13. WoodsR. K.ThienF.RavenJ.WaltersE. H.MAAbramsonPrevalence of food allergies in young adults and their relationship to asthma, nasal allergies, and eczema. Annals of Allergy, Asthma & Immunology 2002882183189
  14. 14. BindelsJ. G.HoijerM.Allergenslatest.developmentsnewest.techniquesBulletin of the International Dairy Federation 200035131
  15. 15. HøstA.Dietary products used in infants for treatment and prevention of food allergy. Joint Statement of the European Society for Paediatric Allergology and Clinical Immunology (ESPACI) Committee on HypoallergenicFormulas and the European Society for Paediatric Gastroenterology, Hepatology and Nutrition (ESPGHAN) Committee on Nutrition. Archives of Disease in Childhood 19998118084
  16. 16. Distler JW. Food allergy Children.An update on diagnosis and treatment. ADVANCE for NPs & Pas 2010122630
  17. 17. JansenJ. J.KardinaalA. F.HuijbersG.Vlieg-BoerstraB. J.MartensB. P.OckhuizenT.Prevalence of food allergy and intolerance in the adult Dutch population. Journal of Allergy and Clinical Immunology 1994932446456
  18. 18. Lam-YH.Van HoffenE.MichelsenA.GuikersK.Van Der TasC. H. W.Bruijnzeel-KoomenC. A. F. M.KnulstA. C.Cow’s milk allergy in adults is rare but severe: both casein and whey proteins are involved. Clinical & Experimental Allergy 20083869951002
  19. 19. StögerP.WuthrichB.TypeI.allergyto.cowmilk.proteinsin.adultsA retrospective study of 34 adult milk and cheese allergic patients. International Archives of Allergy and Immunology 19931024399407
  20. 20. RautavaS.RuuskanenO.OuwehandA.SalminenS.IsolauriE.The hygiene hypothesis of atopic disease-an extended version. Journal of Pediatric Gastroenterology and Nutrition 200438437888
  21. 21. StrannegårdO.StrannegårdI. L.The causes of the increasing prevalence of allergy: is atopy a microbial deprivation disorder? Allergy 200156291102
  22. 22. BitoT.KandaE.TanakaM.FukunagaA.HorikawaT.NishigoriC.Cows milk-dependent exercise-induced anaphylaxis under the condition of a premenstrual or ovulatory phase following skin sensitization. Allergology International 2008574437439
  23. 23. SavilahtiE.KuitunenM.Allergenicity of cow milk proteins. Journal of Pediatrics 1992Pt 2) S1220
  24. 24. International Union of Immunological SocietiesAllergen Nomenclature Sub-Committee. Allergen Nomenclature. Retrieved from http://www.allergen.org/Allergen.aspx.Accessed 2009
  25. 25. Wal J-M. Cow’s milk proteins/allergens. Annals of Allergy, Asthma & Immunology2002Suppl 9) 3-10.
  26. 26. RestaniP.BallabioC.Di LorenzoC.TripodiS.FiocchiA.Molecular aspects of milk allergens and their role in clinical events. Analytical and Bioanalytical Chemistry 200939514756
  27. 27. PlebaniA.RestaniP.NaselliA.GalliC. L.MeiniA.CavagniG.UgazioA. G.PoiesiC.Monoclonal and polyclonal antibodies against casein components of cow milk for evaluation of residual antigenic activity in ‘hypoallergenic’ infant formulas. Clinical & Experimental Allergy 1997278949956
  28. 28. PomsR. E.KleinC. L.AnklamE.Methods for allergen analysis in food: a review. Food Additives and Contaminants 2004211131
  29. 29. RozenfeldP.DocenaG. H.AnonM. C.CAFossatiDetection and identification of a soy protein component that cross-reacts with caseins from cow’s milk. Clinical & Experimental Immunology 200213014958
  30. 30. Loo JA, Edmonds CG, Smith RD.Primary sequence information from intact proteins by electrospray ionization tandem mass spectrometry. Science 19902484952201204
  31. 31. Senko MW, Speir JP, McLafferty FW.Collisional activation of large multiply charged ions using Fourier transform mass spectrometry. Analytical Chemistry 1994661828012808
  32. 32. Little DP, Speir JP, Senko MW, O’Connor PB, McLafferty FW.Infrared multiphoton dissociation of large multiply charged ions for biomolecule sequencing. Analytical Chemistry 1994661828092815
  33. 33. OhH.BreukerK.SzeS. K.GeY.CarpenterB. K.Mc LaffertyF. W.Secondary and tertiary structures of gaseous protein ions characterized by electron capture dissociation mass spectrometry and photofragment spectroscopy. Proceedings of the National Academy of Sciences 200299251586315868
  34. 34. YatesJ. R. I. I. I.Database searching using mass spectrometry data. Electrophoresis 1998196893900
  35. 35. WysockiV. H.ResingK. A.ZhangQ.ChengG.Mass spectrometry of peptides and proteins. Methods 2005353211222
  36. 36. SpuerginP.MuellerH.WalterM.SchiltzE.ForsterJ.Allergenic epitopes of bovine αs1-casein recognized by human IgE ad IgG. Allergy 1996515306312
  37. 37. ChatchateeP.JarvinenK. M.BardinaL.BeyerK.SampsonH. A.Identificationof.IgE.IgG-binding.epitopeson.αs1-caseindifferences.inpatients.withpersistent.transientcow’s.milkallergy.Journal of Allergy and Clinical Immunology 20011072379383
  38. 38. RuiterB.TregoatV.M’RabetL.GarssenJ.CABruijnzeel-KoomenKnol. E. F.HoffenE.Characterizationof. T.cellepitopes.inαs1-casein.incow’s.milkallergic.atopicnon-atopicchildren.Clinical & Experimental Allergy 2006363303310
  39. 39. RuiterB.RozemullerE. H.van DijkA. J.GarssenJ.CABruijnzeel-KoomenTilanus. M. G.KnolE. F.van HoffenE.Roleof.humanleucocyte.antigenD. Q.inthe.presentationof. T.cellepitopes.inthe.majorcow’s.milkallergen.αs1-caseinInternational Archives of Allergy and Immunology 20071432119126
  40. 40. BernardH.CreminonC.YvonM.WalJ. M.Specificity of the human IgE response to the different purified caseins in allergy to cow’s milk proteins. International Archives of Allergy and Immunology 19981151235244
  41. 41. WalstraP.Caseinsub-micelles.dotheyexist?.InternationalDairy.Journal1999
  42. 42. FarrellH. M.Jr Jimenez-FloresR.BleckG. T.BrownE. M.ButlerJ. E.CreamerL. K.HicksC. L.HollarC. M.Ng-Kwai-HangK. F.SwaisgoodH. E.Nomenclature of the proteins of cows’ milk-sixth revision. Journal of Dairy Science 200487616411674
  43. 43. Horne, D.S.Casein structure, self-assembly and gelation.Current Opinion in Colloid & Interface Science 2002
  44. 44. KohnoY.HonmaK.SaitoK.ShimojoN.TsunooH.KaminogawaS.NiimiH.Preferential recognition of primary protein structures of alpha-casein by IgG and IgE antibodies of patients with milk allergy. Annales of Allergy 1994735419422
  45. 45. MonaciL.TregoatV,van.HengelA. J.AnklamE.Milk allergens, their characteristics and their detection in food: A review. European Food Research And Technology 20062232149179
  46. 46. SpuerginP.WalterM.SchiltzE.DeichmannK.ForsterJ.MuellerH.Allergenicity of alpha-caseins from cow, sheep, and goat. Allergy 1997523293298
  47. 47. VilaL.BeyerK.JärvinenK. M.ChatchateeP.BardinaL.SampsonH. A.Role of conformational and linear epitopes in the achievement of tolerance in cow’s milk allergy. Clinical & Experimental Allergy 2001311015991606
  48. 48. ElsayedS.HillD. J.DoT. V.Evaluation of the allergenicity and antigenicity of bovine-milk alphas1-casein using extensively purified synthetic peptides. Scandinavian Journal of Immunology 200460548693
  49. 49. SchulmeisterU.HochwallnerH.SwobodaI.Focke-TejklM.GellerB.NystrandM.HärlinA.ThalhamerJ.ScheiblhoferS.KellerW.NiggemannB.QuirceS.EbnerC.MariA.PauliG.HerzU.ValentaR.SpitzauerS.Cloningexpression.mappingof.allergenicdeterminants.ofalpha.S1-caseina.majorcow’s.milkallergen.The Journal of Immunology 20091821170197029
  50. 50. BusseP. J.JärvinenK. M.VilaL.BeyerK.SampsonH. A.Identification of sequential IgE-binding epitopes on bovine alpha(s2)-casein in cow’s milk allergic patients. International Archives of Allergy and Immunology 200212919396
  51. 51. JärvinenK. M.BeyerK.VilaL.ChatchateeP.BusseP. J.SampsonH. A.B-cell epitopes as a screening instrument for persistent cow’s milk allergy. Journal of Allergy and Clinical Immunology. 20021102293297
  52. 52. ChatchateeP.JärvinenK. M.BardinaL.VilaL.BeyerK.SampsonH. A.Identificationof.IgE.IgG.bindingepitopes.onbeta.kappa-caseinin.cow’smilk.allergicpatients.Clinical & Experimental Allergy 200131812561262
  53. 53. RasmussenL. K.HøjrupP.PetersenT. E.Themultimeric.structuredisulfide-bondingpattern.ofbovine.kappa-caseinEuropean Journal of Biochemistry 19922071215222
  54. 54. RasmussenL. K.HøjrupP.PetersenT. E.Disulphide arrangement in bovine caseins: localization of intrachain disulphide bridges in monomers of kappa- and alpha s2-casein from bovine milk. Journal of Dairy Research 1994614485493
  55. 55. Creamer LK, Plowman JE, Liddell MJ, Smith MH, Hill JP. Micelle stability: kappa-casein structure and function.Journal of Dairy Science 1998811130043012
  56. 56. Godovac-ZimmermannJ.ContiA.LiberatoriJ.BraunitzerG.Theamino-acid.sequenceof.beta-lactoglobulinI. I.fromhorse.colostrum.Equuscaballus.Perissodactylabeta-lactoglobulinsare.retinol-bindingproteins.Biological chemistry Hoppe-Seyler 19853666601608
  57. 57. KontopidisG.HoltC.SawyerL.Invitedreview.beta-lactoglobulinbinding.propertiesstructure.functionJournal of Dairy Science 2004874785796
  58. 58. McKenzie HA, Sawyer WH, Smith MB.Optical rotatory dispersion and sedimentation in the study of association-dissociation: bovine beta-lactoglobulins near pH 5. Biochimica and Biophysica Acta 196714717392
  59. 59. BMEhnEkstrand. B.BengtssonU.AhlstedtS.Modificationof.IgE.bindingduring.heatprocessing.ofthe.cow’smilk.allergenbeta-lactoglobulin.Journal of Agricultural and Food Chemistry 200452513981403
  60. 60. Chen, WL, Hwang MT, Liau CY, Ho JC, Hong KC, and Mao SJ.Beta-lactoglobulin is a thermal marker in processed milk as studied by electrophoresis and circular dichroic spectra. Journal of Dairy Science 200588516181630
  61. 61. Guyomarc’hF.LawA. J.DalgleishD. G.Formation of soluble and micelle-bound protein aggregates in heated milk. Journal of Agricultural Food Chemistry 2003511646524660
  62. 62. ClementG.BoquetD.FrobertY.BernardH.NegroniL.Chatel-MJ.Adel-PatientK.CreminonC.Wal-MJ.GrassiJ.Epitopic characterization of native bovine b-lactoglobulin. Journal of Immunological Methods 2002
  63. 63. SéloI.ClémentG.BernardH.ChatelJ.CréminonC.PeltreG.WalJ.Allergy to bovine beta-lactoglobulin: specificity of human IgE to tryptic peptides. Clinical & Experimental Allergy 199929810551063
  64. 64. FritschéR.Adel-PatientK.BernardH.Martin-PaschoudC.SchwarzC.Ah-LeungS.WalJ. M.IgIgE-mediated rat mast cell triggering with tryptic and synthetic peptides of bovine beta-lactoglobulin. International Archives of Allergy and Immunology 20051384291297
  65. 65. MillerK.MeredithC.SeloI.WalJ. M.Allergy to bovine beta-lactoglobulin: specificity of immunoglobulin E generated in the Brown Norway rat to tryptic and synthetic peptides. Clinical & Experimental Allergy 1999291216961704
  66. 66. Polverino deLaureto. P.FrareE.GottardoR.Van DaelH.FontanaA.Partly folded states of members of the lysozyme/lactalbumin superfamily: a comparative study by circular dichroism spectroscopy and limited proteolysis. Protein Science 2002111229322946
  67. 67. Veprintsev DB, Permyakov SE, Permyakov EA, Rogov VV, Cawthern KM, Berliner LJ.Cooperative thermal transitions of bovine and human apo-alpha-lactalbumins: evidence for a new intermediate state. FEBS Letters 19974123625628
  68. 68. McGuffey MK, Epting KL, Kelly RM, Foegeding EA.Denaturation and aggregation of three alpha-lactalbumin preparations at neutral pH. Journal of Agricultural and Food Chemistry 200553831823190
  69. 69. Redfield C.Using nuclear magnetic resonance spectroscopy to study molten globule states of proteins.Methods 200434112132
  70. 70. Chang JY.Evidence for the underlying cause of diversity of the disulfide folding pathway. Biochemistry 2004431545224529
  71. 71. LivneyY. D.VerespejE.DalgleishD. G.Steric effects governing disulfide bond interchange during thermal aggregation in solutions of beta-lactoglobulin B and alpha-lactalbumin. Journal of Agricultural and Food Chemistry 2003512780988106
  72. 72. McKenzie HA. alpha-Lactalbumins and lysozymes. EXS199675365
  73. 73. RamakrishnanB.QasbaP. K.Crystal structure of lactose synthase reveals a large conformational change in its catalytic component, the beta1,4-galactosyltransferase-I. Journal of Molecular Biology 20013101205218
  74. 74. Permyakov EA and Berliner LJ. alpha-Lactalbumin: Structure and function. FEBS Letters20004733269274
  75. 75. HakanssonA.SvenssonM.MossbergA. K.SabharwalH.LinseS.LazouI.LonnerdalB.SvanborgC. A.foldingvariant.ofalpha-lactalbumin.withbactericidal.activityagainst.Streptococcuspneumoniae.Molecular Microbiololgy 2000353589600
  76. 76. PellegriniA.ThomasU.BramazN.HunzikerP.vonFellenberg. R.Isolation and identification of three bactericidal domains in the bovine alpha-lactalbumin molecule. Biochimica and Biophysica Acta 199914263439448
  77. 77. Permyakov SE, Pershikova IV, Khokhlova TI, Uversky VN, Permyakov EA.No need to be HAMLET or BAMLET to interact with histones: binding of monomeric alpha-lactalbumin to histones and basic poly-amino acids. Biochemistry 2004431955755582
  78. 78. SlangenC. J.VisserS.Use of mass spectrometry To rapidly characterize the heterogeneity of bovine alpha-lactalbumin. Journal of Agricultural and Food Chemistry 1999471145494556
  79. 79. MaynardF.JostR.WalJ. M.HumanIg. E.bindingcapacity.oftryptic.peptidesfrom.bovine-Lactalbuminalpha.International Archives of Allergy and Immunology 19971134478488
  80. 80. WróblewskaB.JędrychowskiL.Milkallergens.In Lucjan Jędrychowski edtr Chemical and Biological Properties of Food Allergens Lucjan Jędrychowski, Harry Wichers Editors CRC Press, 2010ISBN 142005855X, 9781420058550
  81. 81. SzepfalusiZ.EbnerC.UrbanekR.EbnerH.ScheinerO.Boltz-NitulescuG.KraftD.Detectionof.IgE.antibodiesspecific.forallergens.incow.milkcowdander.International Archives of Allergy and Immunology 19931023288294
  82. 82. HanG. D.MatsunoM.ItoG.IkeuchtY.SuzukiA.Meatallergy.investigationof.potentialallergenic.proteinsin.beefBioscience, Biotechnology, and Biochemistry 200064918871895
  83. 83. FiocchiA.RestaniP.RivaE.Beef allergy in children. Nutrition 2000166454457
  84. 84. FiocchiA.RestaniP.RivaE.MirriG. P.SantiniI.BernardoL.GalliC. L.Heat treatment modifies the allergenicity of beef and bovine serum albumin. Allergy 1998538798802
  85. 85. RestaniP.FiocchiA.BerettaB.VelonaT.GiovanniniM.GalliC. L.Meatallergy. I. I. I.Proteinsinvolved.cross-reactivitybetween.differentanimal.speciesJournal of the American College of Nutrition 1997164383389
  86. 86. WerfelS.CookeS. K.SampsonH. A.Clinical reactivity to beef in children allergic to cow’s milk. Journal of Allergy and Clinical Immunology 1997993293300
  87. 87. van ReeR.van LeeuwenW. A.AalberseR. C.How far can we simplify in vitro diagnostics for grass pollen allergy?: A study with 17 whole pollen extracts and purified natural and recombinant major allergens. Journal of Allergy and Clinical Immunology 19981022184190
  88. 88. BernardH.MeiselH.CreminonC.WalJ. M.Post-translational phosphorylation affects the IgE binding capacity of caseins. FEBS Lett. 2000
  89. 89. MolinaA. C. T.AlliI.BoyeJ. I.Effects of dephosphorylation and proteolysis on immunogenicity of caseins. Milchwissenschaft-Milk Science International 2007624371375
  90. 90. SchlimmeE.MeiselH.Bioactive peptides derived from milk proteins. Structural, physiological and analytical aspects. Nahrung/Food 1995391120
  91. 91. PizzanoR.MANicolaiPadovano. P.FerrantiP.BaroneF.AddeoF.Immunochemical evaluation of bovine beta-casein and its 1-28 phosphopeptide in cheese during ripening. Journal of Agricultural and Food Chemistry 2000481045554560
  92. 92. MariA.IgE.to-ReactiveCross.CarbohydrateDeterminants.analysisof.thedistribution.appraisalof.thein.vivoinvitro.reactivityInternational Archives of Allergy and Immunology 20021294286295
  93. 93. HochwallnerH.SchulmeisterU.SwobodaI.Focke-TejklM.CivajV.BalicN.NystrandM.HärlinA.ThalhamerJ.ScheiblhoferS.KellerW.PavkovT.ZafredD.NiggemannB.QuirceS.MariA.PauliG.EbnerC.PapadopoulosN. G.HerzU.van TolE. A.ValentaR.SpitzauerS.Visualization of clustered IgE epitopes on alpha-lactalbumin Journal of Allergy and Clinical Immunology 2010125612791285e9.
  94. 94. TranN. T.DaaliY.CherkaouiS.TavernaM.NeeserJ. R.VeutheyJ. L.Routine o-glycan characterization in nutritional supplements--a comparison of analytical methods for the monitoring of the bovine kappa-casein macropeptide glycosylation. Journal of Chromatography A 2001
  95. 95. SlangenC. J.VisserS.Use of mass spectrometry To rapidly characterize the heterogeneity of bovine alpha-lactalbumin. Journal of Agricultural and Food Chemistry 1999471145494556
  96. 96. HeineW.WutzkeK. D.RadkeM.Decreasing the immunogenicity of milk proteins by desialinization. Infusionstherapie 1989166264266
  97. 97. Holland JW, Deeth HC, Alewood PF.Proteomic analysis of kappa-casein micro-heterogeneity. Proteomics 200443743752
  98. 98. PizzanoR.MANicolaiManzo. C.GiannattasioM.AddeoF.HumanIg. E.bindingto.theglycosidic.moietyof.bovinekappa-casein.Journal of Agricultural and Food Chemistry 2005532079717975
  99. 99. AkagawaM.HandoyoT.IshiiT.KumazawaS.MoritaN.SuyamaK.Proteomic analysis of wheat flour allergens. Journal of Agricultural and Food Chemistry 2007551768636870
  100. 100. KarasM.HillenkampF.Laser desorption ionization of proteins with molecular masses exceeding 10,000 daltons. Chemistry 1988602022992301
  101. 101. FennJ. B.MannM.MengC. K.WongS. F.WhitehouseC. M.Electrospray ionization for mass spectrometry of large biomolecules. Science 198924649266471
  102. 102. AebersoldR.MannM.Massspectrometry-based.proteomicsNature.20034226928198207
  103. 103. Yates JR 3rd.Mass spectral analysis in proteomics. Annual Review of Biophysics and Biomolecular Structure 200433297
  104. 104. DomonB.AebersoldR.Mass spectrometry and protein analysis. Science 20063125771212217
  105. 105. HanX.AslanianA.YatesJ. R.3rd Mass spectrometry for proteomics. Current Opinion in Chemical Biology 2008125483490
  106. 106. HirabayashiJ.Lectin-basedstructural.glycomicsglycoproteomics.glycanprofiling.Glycoconjugate Journal 2004
  107. 107. WuhrerM.de BoerA. R.DeelderA. M.Structural glycomics using hydrophilic interaction chromatography (HILIC) with mass spectrometry. Mass Spectrometry Review 2009282192206
  108. 108. Harvey DJ.Analysis of carbohydrates and glycoconjugates by matrix-assisted laser desorption/ionization mass spectrometry: an update for the period 2005-2006. Mass Spectrometry Review 20113011100
  109. 109. An HJ, Peavy TR, Hedrick JL, Lebrilla CB.Determination of N-glycosylation sites and site heterogeneity in glycoproteins. Analytical Chemistry 2003752056285637
  110. 110. ZaiaJ.Mass spectrometry of oligosaccharides. Mass Spectrometry Review 2004233161227
  111. 111. DasB. C.GeroS. D.LedererE.N-methylation of N-acyl oligopeptides. Biochemical and Biophysical Research Communications 1967292211215
  112. 112. De AngelisM.Di CagnoR.MinerviniF.RizzelloC. G.GobbettiM.Two-dimensional electrophoresis and IgE-mediated food allergy. Electrophoresis 2010311321262136
  113. 113. PetersenA.Two-dimensional electrophoresis replica blotting: a valuable technique for the immunological and biochemical characterization of single components of complex extracts. Proteomics 20033712061214
  114. 114. NataleM.BissonC.MontiG.PeltranA.GaroffoL. P.ValentiniS.FabrisC.BertinoE.CosciaA.ContiA.Cow’s milk allergens identification by two-dimensional immunoblotting and mass spectrometry. Molecular Nutrition & Food Research 2004485363369
  115. 115. MonaciL.LositoI.PalmisanoF.ViscontiA.Reliable detection of milk allergens in food using a high-resolution, stand-alone mass spectrometer. Journal of AOAC International 201194410341042
  116. 116. LuB.XuT.ParkS. K.Mc ClatchyD. B.LiaoL.YatesJ. R.3rd Shotgun protein identification and quantification by mass spectrometry in neuroproteomics. Methods in Molecular Biology 2009566229
  117. 117. WeberD.RaymondP.Ben-RejebS.LauB.Development of a liquid chromatography-tandem mass spectrometry method using capillary liquid chromatography and nanoelectrospray ionization-quadrupole time-of-flight hybrid mass spectrometer for the detection of milk allergens. Journal of Agricultural and Food Chemistry 200654516041610
  118. 118. KuppannanK.AlbersD. R.SchaferB. W.DielmanD.YoungS. A.Quantification and characterization of maize lipid transfer protein, a food allergen, by liquid chromatography with ultraviolet and mass spectrometric detection. Analytical Chemistry 2011832516524
  119. 119. HuberC. G.PremstallerA.Evaluation of volatile eluents and electrolytes for high-performance liquid chromatography-electrospray ionization mass spectrometry and capillary electrophoresis-electrospray ionization mass spectrometry of proteins. I. Liquid chromatography. Journal of Chromatography A 1999849116173
  120. 120. CzerwenkaC.MaierI.PotocnikN.PittnerF.LindnerW.Absolute quantitation of beta-lactoglobulin by protein liquid chromatography-mass spectrometry and its application to different milk products. Analytical Chemistry 2007791451655172
  121. 121. WuW. W.WangG.BaekS. J.ShenR. F.Comparativestudy.ofthree.proteomicquantitative.methodsD. I. G. E. c. I. C. A. T.usingi. T. R. A. Q.gel-or. D.-ML. C.A. L. D. I. T. O. F. T. O. F.Journal of Proteome Research 200653651658
  122. 122. GingrasA. C.GstaigerM.RaughtB.AebersoldR.Analysis of protein complexes using mass spectrometry. Nature Reviews Molecular Cell Biology 200788645654
  123. 123. PicottiP.RinnerO.StallmachR.DautelF.FarrahT.DomonB.WenschuhH.AebersoldR.High-throughput generation of selected reaction-monitoring assays for proteins and proteomes. Nature Methods. 2010714346
  124. 124. CareriM.CostaA.ElviriL.LagosJ. B.MangiaA.TerenghiM.CeretiA.GaroffoL. P.Use of specific peptide biomarkers for quantitative confirmation of hidden allergenic peanut proteins Ara h 2 and Ara h 3/4 for food control by liquid chromatography-tandem mass spectrometry. Analytical and Bioanalytical Chemistry 2007389619011907
  125. 125. HeickJ.FischerM.KerbachS.TammU.PoppingB.Application of a liquid chromatography tandem mass spectrometry method for the simultaneous detection of seven allergenic foods in flour and bread and comparison of the method with commercially available ELISA test kits. Journal of AOAC International 201194410601068
  126. 126. MannM.HendricksonR. C.PandeyA.Analysis of proteins and proteomes by mass spectrometry. Annual Review of Biochemistry 200170437
  127. 127. ViethsS.Hoffmann-SommergruberK.EuroPrevall.foodallergen.libraryMolecular Nutrition & Food Research 2008Suppl 2:S1578
  128. 128. SanchoA. I.Hoffmann-SommergruberK.AlessandriS.ContiA.GiuffridaM. G.ShewryP.BMJensenSkov. P.ViethsS.Authentication of food allergen quality by physicochemical and immunological methods. Clinical & Experimental Allergy 2010
  129. 129. MaassS.SieversS.ZühlkeD.KuzinskiJ.SappaP. K.MuntelJ.HesslingB.BernhardtJ.SietmannR.VölkerU.HeckerM.BecherD.Efficientglobal-scale.quantificationof.absoluteprotein.amountsby.integrationof.targetedmass.spectrometrytwo-dimensionalgel-based.proteomicsAnalytical Chemistry 2011
  130. 130. FAO/WHO, Evaluation of Allergenicity of Genetically Modified Foods, FAO, Rome,20011
  131. 131. Williams SC, Badley RA, Davis PJ, Puijk WC, Meloen RH.Identification of epitopes within beta lactoglobulin recognised by polyclonal antibodies using phage display and PEPSCAN. Journal of Immunological Methods 19982131117
  132. 132. PicarielloG.FerrantiP.FierroO.MamoneG.CairaS.Di LucciaA.MonicaS.AddeoF.Peptides surviving the simulatedgastrointestinal digestion of milk proteins: biological and toxicological implications. Journal of Chromatography B: AnalyticalTechnologies in the Biomedical and Life Sciences 2010
  133. 133. BannonG. A.CockrellG.ConnaughtonC.WestC. M.HelmR.StanleyJ. S.KingN.RabjohnP.SampsonH. A.BurksA. W.Engineeringcharacterization.invitro.efficacyof.themajor.peanutallergens.foruse.inimmunotherapy.International Archives of Allergy and Immunology 2001
  134. 134. BurksA. W.KingN.BannonG. A.Modification of a major peanut allergen leads to loss of IgE binding.International Archives of Allergy and Immunology 1999
  135. 135. SwobodaI.Bugajska-SchretterA.LinhartB.VerdinoP.KellerW.SchulmeisterU.SperrW. R.ValentP.PeltreG.QuirceS.DouladirisN.PapadopoulosN. G.ValentaR.SpitzauerS. A.recombinanthypoallergenic.parvalbuminmutant.forimmunotherapy.ofIg.E-mediatedfish.allergyThe Journal of Immunology 20071781062906296
  136. 136. PrioultG.PecquetS.FlissI.Stimulationof.interleukin-productionby.acidicbeta-lactoglobulin-derived.peptideshydrolyzed.withLactobacillus.paracaseiN. C.C24peptidasesClinical and Diagnostic Laboratory Immunology 2004112266271
  137. 137. PrioultG.PecquetS.FlissI.Allergenicity of acidic peptides from bovine beta-lactoglobulin is reduced by hydrolysis with Bifidobacterium lactis NCC362 enzymes International Dairy Journal 2005155439448
  138. 138. Di CagnoR.RizzelloC. G.De AngelisM.CassoneA.GiulianiG.BenedusiA.LimitoneA.SuricoR. F.GobbettiM.Use of selected sourdough strains of Lactobacillus for removing gluten and enhancing the nutritional properties of gluten-free bread. Journal Of Food Protection 200871714911495
  139. 139. GianfraniC.SicilianoR. A.FacchianoA. M.CamarcaA.MazzeoM. F.CostantiniS.SalvatiV. M.MauranoF.MazzarellaG.IaquintoG.BergamoP.RossiM.Transamidation of wheat flour inhibits the response to gliadin of intestinal T cells in celiac disease. Gastroenterology 20071333780789
  140. 140. RestaniP.BerettaB.FiocchiA.BallabioC.GalliC. L.Cross-reactivity between mammalian proteins. Annals of Allergy, Asthma & Immunology 20028961115
  141. 141. Park YW.Hypo-allergenic and therapeutic significance of goat milk. Small Ruminant Research 1994142151159
  142. 142. Bellioni-BusincoB.PaganelliR.LucentiP.GiampietroP. G.PerbornH.BusincoL.Allergenicity of goat’s milk in children with cow’s milk allergy. Journal of Allergy and Clinical Immunology 1999103611911194
  143. 143. PesslerF.NejatM.Anaphylactic reaction to goat’s milk in a cow’s milk-allergic infant. Pediatric Allergy and Immunology 2004152183185
  144. 144. BallabioC.ChessaS.RignaneseD.GigliottiC.PagnaccoG.TerraccianoL.FiocchiA.RestaniP.CaroliA. M.Goat milk allergenicity as a function of αs₁-casein genetic polymorphism. Journal of Dairy Science 20119429981004
  145. 145. TomotakeH.KatagiriM.FujitaM.YamatoM.Preparation of fresh cheese from caprine milk as a model for the reduction of allergenicity. Journal of Nutritional Science and Vitaminology (Tokyo) 2009553296300
  146. 146. SheehanW. J.PhipatanakulW.Tolerance to water buffalo milk in a child with cow milk allergy. Annals of Allergy, Asthma & Immunology 2009
  147. 147. D’AuriaE.AgostoniC.GiovanniniM.RivaE.ZetterströmR.FortinR.GreppiG. F.BonizziL.RoncadaP.Proteomic evaluation of milk from different mammalian species as a substitute for breast milk. Acta Paediatrica 2005941217081713
  148. 148. LamblinC.BourrierT.OrlandoJ. P.SauvageC.WallaertB.Allergie aux laits de chèvre et de brebis sans allergie associée au lait de vache. Revue Française d’Allergologie et d’Immunologie Clinique 2001412165168
  149. 149. BidatE.RancéF.BaranèsT.GoulamhoussenS.Goat’s milk and sheep’s milk allergies in children in the absence of cow’s milk allergy. Revue Française d’Allergologie et d’Immunologie Clinique 2003434273277
  150. 150. MuñozMartín. T.de la HozCaballer. B.MarañónLizana. F.GonzálezMendiola. R.PrietoMontaño. P.SánchezCano. M.Selective allergy to sheep’s and goat’s milk proteins. Allergologia et Immunopathologia (Madr) 20043213942
  151. 151. Ah-LeungS.BernardH.BidatE.PatyE.RancéF.ScheinmannP.WalJ. M.Allergy to goat and sheep milk without allergy to cow’s milk. Allergy. 2006611113581365
  152. 152. WüthrichB.JohanssonS. G.Allergy to cheese produced from sheep’s and goat’s milk but not to cheese produced from cow’s milk. Journal of Allergy and Clinical Immunology 1995962270273
  153. 153. UmpiérrezA.QuirceS.MarañónF.CuestaJ.García-VillamuzaY.LahozC.SastreJ.Allergy to goat and sheep cheese with good tolerance to cow cheese. Clinical & Experimental Allergy 199929810641068
  154. 154. BusincoL.GiampietroP. G.LucentiP.LucaroniF.PiniC.Di FeliceG.IacovacciP.CuradiC.OrlandiM.Allergenicity of mare’s milk in children with cow’s milk allergy. Journal of Allergy and Clinical Immunology. 2000105510311034
  155. 155. GuoH. Y.PangK.ZhangX. Y.ZhaoL.ChenS. W.DongM. L.RenF. Z.Compositionphysiochemical.propertiesnitrogen.fractiondistribution.aminoacid.profileof.donkeymilk.Journal of Dairy Science. 200790416351643
  156. 156. MontiG.BertinoE.MuratoreM. C.CosciaA.CresiF.SilvestroL.FabrisC.FortunatoD.GiuffridaM. G.ContiA.Efficacy of donkey’s milk in treating highly problematic cow’s milk allergic children: an in vivo and in vitro study. Pediatric Allergy and Immunology 2007183258264
  157. 157. CunsoloV.SalettiR.MuccilliV.FotiS.Characterization of the protein profile of donkey’s milk whey fraction. Journal of Mass Spectrometry 200742911621174
  158. 158. ChianeseL.CalabreseM. G.FerrantiP.MaurielloR.GarroG.De SimoneC.QuartoM.AddeoF.CosenzaG.RamunnoL.Proteomiccharacterization.ofdonkey.milk"caseome.Journal of Chromatography A 201012172948344840
  159. 159. CunsoloV.MuccilliV.FasoliE.SalettiR.RighettiP. G.FotiS.Poppea’sbath.liquorthe.secretproteome.ofthe-donkey’s.milkJournal of Proteomics 20117410208399
  160. 160. SuutariT. J.ValkonenK. H.KarttunenT. J.BMEhnEkstrand. B.BengtssonU.VirtanenV.NieminenM.KokkonenJ.IgE.crossreactivity.betweenreindeer.bovinemilk.beta-lactoglobulinsin.cow’smilk.allergicpatients.Journal of Investigational Allergology and Clinical Immunology 2006165296302
  161. 161. El -AgamyE. I.NawarM.ShamsiaS. M.AwadS.HaenleinG. F. W.Are camel milk proteins convenient to the nutrition of cow milk allergic children? Small Ruminant Research 200982116
  162. 162. OkadaH.KuhnC.Feillet The ‘hygiene hypothesis’ for autoimmune and allergic diseases: an updatClinical and Experimental Immunology 2010; 160(1) 1-9.
  163. 163. CampeottoF.Waligora-DuprietA. J.Doucet-PopulaireF.KalachN.DupontC.MJButelEstablishment of the intestinal microflora in neonates. Gastroenterologie Clinique et Biologique 2007315533542
  164. 164. AdlerberthI.WoldA. E.Establishment of the gut microbiota in Western infants. Acta Paediatrica 2009982229238
  165. 165. StsepetovaJ.SeppE.JulgeK.VaughanE.MikelsaarM.de VosW. M.Molecularly assessed shifts of Bifidobacterium ssp. and less diverse microbial communities are characteristic of 5-year-old allergic children. FEMS Immunology & Medical Microbiology. 2007512260269
  166. 166. VaelC.NelenV.VerhulstS. L.GoossensH.DesagerK. N.Early intestinal Bacteroides fragilis colonisation and development of asthma. BMC Pulmonary Medicine. 2008268
  167. 167. SjögrenY. M.TomicicS.LundbergA.BöttcherM. F.BjörksténB.Sverremark-EkströmE.JenmalmM. C.Influence of early gut microbiota on the maturation of childhood mucosal and systemic immune responses. Clinical & Experimental Allergy. 2009391218421851
  168. 168. SjögrenY. M.JenmalmM. C.BöttcherM. F.BjörksténB.Sverremark-EkströmE.Altered early infant gut microbiota in children developing allergy up to 5 years of age. Clinical & Experimental Allergy 200939451826
  169. 169. ChenW.JinW.HardegenN.LeiK. J.LiL.MarinosN.Mc GradyG.WahlS. M.Conversion of peripheral CD4+CD25-naive T cells to CD4+CD25+regulatory T cells by TGF-beta induction of transcription factor Foxp3. The Journal of Experimental Medicine. 200319818751886
  170. 170. KarimM.KingsleyC. I.BushellA. R.BSSawitzkiWood. K. J.Alloantigen-inducedC.D2C. D.regulatoryT.cellscan.developin.vivofrom. C.D2C. D.precursorsin. a.thymus-independentprocess.The Journal of Immunology 20041722923928
  171. 171. O’MahonyC.ScullyP.O’MahonyD.MurphyS.O’BrienF.LyonsA.SherlockG.MacSharry. J.KielyB.ShanahanF.O’MahonyL.Commensal-inducedregulatory. T.cellsmediate.protectionagainst.pathogen-stimulatedN.F-kappaB.activationP.LoPLoS Pathogens 2008e1000112.
  172. 172. LyonsA.O’MahonyD.O’BrienF.MacSharry. J.SheilB.CeddiaM.RussellW. M.ForsytheP.BienenstockJ.KielyB.ShanahanF.O’MahonyL.Bacterial strain-specific induction of Foxp3+ T regulatory cells is protective in murine allergy models. Clinical & Experimental Allergy 2010405811819
  173. 173. Round JL, Mazmanian SK. Inducible Foxp3+ regulatory T-cell development by a commensal bacterium of the intestinal microbiota.Proceedings of the National Academy of Sciences 2010107271220412209
  174. 174. RodriguezB.PrioultG.BibiloniR.NicolisI.MercenierA.MJButel-DuprietWaligora.A. J.Germ-free status and altered caecal subdominant microbiota are associated with a high susceptibility to cow’s milk allergy in mice. FEMS Microbiology Ecology 2011761133144
  175. 175. HøstA.KoletzkoB.DreborgS.MuraroA.WahnU.AggettP.BressonJ. L.HernellO.LafeberH.MichaelsenK. F.MicheliJ. L.RigoJ.WeaverL.HeymansH.StrobelS.VandenplasY.Dietary products used in infants for treatment and prevention of food allergy. Joint Statement of the European Society for Paediatric Allergology and Clinical Immunology (ESPACI) Committee on Hypoallergenic Formulas and the European Society for Paediatric Gastroenterology, Hepatology and Nutrition (ESPGHAN) Committee on Nutrition. Archives of Disease in Childhood 19998118084
  176. 176. HøstA.Cow’s milk protein allergy and intolerance in infancy. Some clinical, epidemiological and immunological aspects. Pediatric Allergy and Immunology 1994Suppl) 136
  177. 177. Commission of the European Communities.1996Official Journal of the European Commission, 39, 12.
  178. 178. de BoissieuD.MatarazzoP.DupontC.Allergy to extensively hydrolyzed cow milk proteins in infants: identification and treatment with an amino acid-based formula. The Journal of Pediatrics 19971315744747
  179. 179. HalkenS.HøstA.How hypoallergenic are hypoallergenic cow’s milk-based formulas? Allergy. 1997521211751183
  180. 180. VanderhoofJ. A.MurrayN. D.KaufmanS. S.MackD. R.AntonsonD. L.CorkinsM. R.PerryD.KrugerR.Intolerance to protein hydrolysate infant formulas: an underrecognized cause of gastrointestinal symptoms in infants. The Journal of Pediatrics 19971315741744
  181. 181. RagnoV.GiampietroP. G.BrunoG.BusincoL.Allergenicity of milk protein hydrolysate formulae in children with cow’s milk allergy. European Journal of Pediatrics 19931529760762
  182. 182. TakaseM.FukuwatariY.KawaseK.KiyosawaI.OgasaK.SuzukiS.KuroumeT.Antigenicity of casein enzymatichydrolysate. Journal of Dairy Science 1979621015701576
  183. 183. SharmaS.KumarP.BetzelC.SinghT. P.Structure and function of proteins involved in milk allergies. Journal of Chromatography B: Biomedical Sciences and Applications 2001
  184. 184. RichterW. O.JacobB.SchwandtP.Molecular weight determination of peptides by high-performance gel permeation chromatography. Analytical Biochemistry 19831332288291
  185. 185. Hames BD (ed): Get Electrophoresis of Proteins: a Practical Approach, IRL Pr.Mclean, Va. 1981
  186. 186. Coombs RRA, Devey ME, Anderson KJ.Refractoriness to anaphylactic shock after continuous feeding of cow’s milk to guinea pigs. Clinical and Experimental Immunology 1978322263271
  187. 187. Ellis MH, Short JA, Heiner DC.Anaphylaxis after ingestion of a recently hydrolyzed whey protein formula. The Journal of Pediatrics 199111874
  188. 188. BusincoL.CantaniA.MaLonghiGiampietro.PgAnaphylactic reactions to a cow’s milk whey protein hydrolyzate (Alfa-Ré, Nestlé) in infants with cow’s milk allergy. Annals of Allergy 1989624333335
  189. 189. Saylor Jd, Bahna Sl.Anaphylaxis to casein hydrolysate formula. Journal Of Pediatrics 199111817174
  190. 190. Bock SA.Probable allergic reaction to casein hydrolysate formula. Journal Of Allergy And Clinical Immunology 1989842272272
  191. 191. Sampson HA.Immunological mechanisms in adverse reactions to foods. Immunology and Allergy Clinics of North America 1991114701716
  192. 192. Van Beresteijn ECH, Meijer RJ, Schmidt DG.Residual antigenicity of hypoallergenic infant formulas and the occurrence of milk-specific IgE antibodies in patients with clinical allergy. Journal of Allergy and Clinical Immunology 1995963365374
  193. 193. Van Beresteijn, ECH, Peeters RA, Kaper JM, Ron JGM, Robben AJPM, Schmidt DG. Molecular Mass Distribution, Immunological Properties and Nutritive Value of Whey Protein Hydrolysates.Journal of Food Protection 1994577619625
  194. 194. Pelissier-PJ.Mercier-CJ.RibadeauDumas. B.Etude de la protéolyse des caséines αs1 et β bovines par la présure spécificité d’action. peptides amers libérés. Annales des Biologie Animales Biochimie Biophysic 1974142343362
  195. 195. HøstA.Clinical course of cow’s milk protein allergy and intolerance. Pediatric Allergy and Immunology 19989114852
  196. 196. HostA.HalkenS.JacobsenH. P.EstmannA.MortensenS.MygilS.The natural course of cow’s milk protein allergy/intolerance. Journal of Allergy and Clinical Immunology 1997S490
  197. 197. Bishop JM, Hill DJ, Hosking CS Natural history of cow milk allergy: Clinical outcome.The Journal of Pediatrics. 19901166862867
  198. 198. AlvaradoM. I.AlonsoE. G.ÁlvarezM.MDIbáñezLaso. M. T.Persistencia de sensibilización a proteínas de leche de vaca: estudio clínico. Allergologia et Immunopathologia 2000
  199. 199. Lowney ED.Immunological unresponsiveness to a contact sensitizer in man. Journal of Investigative Dermatology 1968516411417
  200. 200. Bellanti JA.Prevention of food allergies. Annals of Allergy 1984Pt 2) 683-688.
  201. 201. TomasiT. B.BarrW. G.ChallacombeS. J.CurranG.Oral tolerance and accessory-cell function of Peyer’s patches. Annals of the New York Academy of Sciences 1983409145

Written By

Simonetta Caira, Rosa Pizzano, Gianluca Picariello, Gabriella Pinto, Marina Cuollo, Lina Chianese and Francesco Addeo

Published: 12 September 2012