• Non ci sono risultati.

30 Mast Cells in the Skin

N/A
N/A
Protected

Academic year: 2022

Condividi "30 Mast Cells in the Skin"

Copied!
10
0
0

Testo completo

(1)

30 Mast Cells in the Skin

M.K. Church

Mast cells are a heterogeneous group of tissue-dwelling cells with roles in conditions as diverse as allergy, para- site infestation, inflammation, angiogenesis, and tissue remodeling. The cells were named Mastzellen in 1876 by Paul Ehrlich because they looked stuffed (German

„gemästet“, „Mast“) and he believed that the intracellu- lar granules, which appeared purple in color when stained with aniline blue dyes, contained phagocytosed materials or nutrients [1]. This change in color, or metachromasia, we now know to represent the interac- tion of the dyes with the highly acidic heparin con- tained within mast cell granules.

The life of the human mast cells begins in the bone marrow as a pluripotent stem cell which enters the bloodstream early on in its development (Fig. 30.1).

Studies of culturing mast cells from cord blood suggest that the precursors are a CD34+/CD38+/HLA-DRpop- ulation of cells [2]. In vivo in mastocytosis, immature mast cells have been recognized as mononuclear cells

Fig. 30.1. Development of mast cells and granulocytes. The right-hand side of the diagram represents a mucosal surface.

Note the close association of MCTCwith nerves and blood ves- sels and MCTwith mucosal epithelium

that both express mRNA for SCF and have SCF recep- tors (SCFR, CD 117) on their cell membranes [3]. From the blood the precursors migrate into the tissues where, under the influence of local microenvironmen- tal factors, they undergo their final phases of differenti- ation and maturation into recognizable mast cells com- plete with cytoplasmic granules and receptors for IgE.

Again, studies of culturing mast cells from cord blood suggest that stem cell factor (SCF) and IL-6 are impor- tant for mast cell maturation [2, 4]. It is pertinent at this stage to distinguish mast cells from basophils, which were originally thought to be circulating mast cells, but are actually related more closely to eosinophils, devel- oping in the bone marrow from granulocyte precur- sors and entering the circulation only when fully mature [5].

Mast cells are distinguished immunocytochemically by their neutral protease content, the MCTphenotype containing only tryptase and the MCTCphenotype con- taining both tryptase and chymase [6]. Initially, these respective subtypes were suggested to be the equiva- lents of the „mucosal“ and „connective tissue“ previ- ously described in experimental animals. However, it is now realized that variable amounts of both mast cell subtypes are present within any given tissue, their rela- tive abundance changing with disease. For example, in allergy MCT, which appear to be „immune system- related“ mast cells with a primarily role in host defense, increase in numbers at mucosal surfaces and allergic foci. Conversely, increased numbers of MCTC, which appear to be „nonimmune system-related“ mast cells with functions in angiogenesis and tissue remod- eling rather than immunological protection, are asso- ciated with fibrosis. However, it should be remembered that both phenotypes express Fc5 RI and may, there- fore, participate fully in IgE-dependent allergic reac- tions.

30

(2)

Mast cells are relatively abundant in human skin, being found in the greatest density in the papillary der- mis and the superficial dermal zone immediately below the dermal-epidermal junction [7]. They are concen- trated particularly around dermal nerve endings and blood vessels [8, 9] and are, therefore, ideally situated to influence the function of both. Normal skin contains around 7,000 mast cells per mm3[10, 11] which equates to a histamine content of 12 – 20 ng/mg tissue [12, 13].

Skin mast cell numbers increase dramatically in sev- eral diseases. For example, the histamine content of the skin in Beh¸cet’s disease is reported to be twice that of normal skin [13] while mast cell numbers are 10-fold higher in urticarial lesional skin [14] and are even higher in urticaria pigmentosa [15]. In a study using antibodies to tryptase and chymase, the number of mast cells in the superficial dermis of mastocytosis lesions was 40,985 ± 21,772 /mm3(mean ± SD) com- pared with 7347 ± 2973 /mm3in normal skin. Further- more, the cells in skin lesions of mastocytosis were exclusively MCTC [10]. Mast cell hyperplasia is also associated with skin tumors such as basal cell carcino- ma [16] and melanoma [17, 18].

Although histamine has been found in significant amounts in the epidermis [19, 20], mast cells are rarely observed in this layer in normal skin. Whether this indicates histamine synthesis by keratinocytes, as indi- cated by murine studies [21], or the ability of keratino- cytes to take up histamine is not clear.

30.1

Mast Cell Activation

Mast cells may be activated by both immunological and nonimmunological mechanisms (Fig. 30.2). To facili- tate immunological activation, human mast cells have

Fig. 30.2. Activation sites of the human skin mast cell

104to 105high affinity (Ka ~ 109/M) receptors (Fc5 RI) for immunoglobulin E (IgE) on the plasma-membrane [22, 23]. Mast cell Fc5 RI is composed of four subunits, an IgE-binding [ -chain, a q -chain, and two * -chains [24]. The presence of the signal-amplifying q -chain [25] in the heterotetrameric [ q * * mast cell and baso- phil receptor complex distinguishes it from the [ * * heterotrimeric receptor of dendritic cells and mono- cytes [23]. Cross linkage of these receptors by multiva- lent allergen stimulates phosphorylation of immunore- ceptor tyrosine activation motifs (ITAMs) [23] and ini- tiation of the biochemical cascade which leads to the release of both preformed and newly generated media- tors.

Nonimmunological activation, which appears to be unique among human mast cells, may be initiated in two ways, by complement fractions and by basic secre- tagogues. Human skin mast cells alone express on the plasma membrane CD88, the receptor for the anaphy- latoxin C5a, allowing them to be activated in comple- ment-mediated disease [26, 27]. Also, skin mast cells alone respond to a variety of basic nonimmunological secretagogues, including neuropeptides, compound 48/80, and drugs such as morphine, codeine, and mus- cle relaxants [28, 29]. These agents stimulate a common activation site on the mast cell membrane which is associated with a pertussis toxin-sensitive G protein [30, 31]. The ability of human skin mast cells, but not those from other tissues, to respond to anaphylatoxins and basic nonimmunological secretagogues explains the flushing reactions observed in sensitive individuals in the absence of overt rhinorrhea or bronchoconstric- tion. Such responses may also be involved in physical urticarias.

In vitro studies of human isolated skin mast cells have shown distinct differences between immunologi- cal and nonimmunological mast cell activation. IgE- dependent activation is relatively slow, taking around 5 – 6 min to reach completion, and requires the pres- ence of extracellular calcium. It is a „complete“ stimu- lus in that it causes release of preformed mediators and initiates the synthesis of the eicosanoids prostaglandin D2and leukotriene C4. In contrast, stimulation of mast cells with basic secretagogues and C5a causes a much more rapid release of histamine, being complete within 30 sec. This release mechanism in which G protein acti- vation leads to subsequent activation of phospholipase C to increase in intracellular inositol triphosphate lev- els, proceeds in the absence of extracellular calcium,

(3)

Histidine

Histamine Histidine decarboxylase

N-Methylhistamine Imidazole acetic acid Diamine oxidase (histaminase) Histamine N-

methyltransferase

(30%) (70%)

calcium mobilization from the endoplasmic reticulum being sufficient to support degranulation. Also, this is an „incomplete“ stimulus in that histamine release is accompanied by negligible eicosanoid generation [30].

Despite these biochemical and temporal differences, degranulation induced by both secretagogues is indis- tinguishable under the electron microscope, proceed- ing by compound exocytosis [32]. From these data it seems likely that IgE-dependent and neuropeptide stimulation of human skin mast cells activate distinct biochemical pathways which eventually merge to stim- ulate exocytosis of their preformed granule-associated mediators.

30.2

Mast Cell Mediators

30.2.1

Early Phase Mediators

The secretory granule of the human mast cell contains a crystalline complex of preformed inflammatory mediators within a matrix of heparin proteoglycan.

The granule-associated mediator most readily associ- ated with the mast cell is the simple diamine histamine.

Histamine is synthesized in the Golgi apparatus of mast cells by decarboxylation of the amino acid, histi- dine, under the influence of histidine decarboxylase (Fig. 30.3). Following synthesis, histamine becomes ionically bound to acidic residues of the glycosamino- glycan (GAG) side chains of heparin proteoglycan [33].

Histamine is present within the granules at ~100 mM, equivalent to about ~4 pg/cell in skin mast cells [34]. In

Fig. 30.3. Synthesis and catabolism of histamine

the extracellular environment, the effects of histamine are normally of relatively short duration as it is rapidly metabolized, usually within 1 – 2 min, by histamine-N- methyltransferase (~70 %), and by diamine oxidase (histaminase) (~30 %) (Fig. 30.3). Interestingly, re- duced diamine oxidase activity has been associated with recurrent urticaria [35].

Histamine exerts many effects pertinent to the immediate phase of allergic responses, including initi- ation of the wheal and flare response initially described by Thomas Lewis in 1927 [36] (Fig. 30.4). The wheal reaction is a direct effect of histamine acting on H1- receptors firstly on local vascular smooth muscle [37]

to cause vasodilatation and then on endothelial cells [38] of postcapillary venules to allow the exudation of plasma proteins [39]. The action of histamine on sen- sory C-fibers also initiates the flare and itch responses [40 – 42].

The other „early phase“ mediators released from the skin mast cell are PGD2and LTC4[43]. While there is little evidence for a role of the former in skin inflam- mation, the success, particularly in individuals with a variant LTC4synthase allele [44], of leukotriene recep- tor antagonists in atopic dermatitis [45] and urticaria [46] suggests that this eicosanoid may be more impor- tant in skin disease than considered previously. It should be emphasized at this point that eosinophils are probably the major producers of LTC4in asthma. The findings that there are large deposits of extracellular eosinophil granule proteins in the skin in both urticar- ia [47] and atopic eczema [48, 49] and raised circulat- ing levels of eosinophil major basic protein in atopic eczema [50] suggest that eosinophils may also make

Fig. 30.4. Wheal and flare in human skin

(4)

Mast Cell

MCT

Tryptase Tryptase

MCTC

Chymase Carboxypeptidase

Cathepsin G a significant contribution to allergic inflammation,

and probably to LTC4generation, in the skin.

30.2.2 Proteoglycans

Proteoglycans are macromolecules which comprise a protein core to which glycosaminoglycan (GAG) side chains are covalently bound (Fig. 30.5). The diverse biological roles of proteoglycans range from simple mechanical support functions to effects on cell differ- entiation and proliferation, cell adhesion and motility, and tissue morphogenesis [51]. In the human mast cell the dominant proteoglycans is a relatively low molecu- lar mass species of heparin (~ 60 kDa) with small amounts of chondroitin E [52, 53]. The highly anionic nature of the GAG side chains is responsible for many of the functions of proteoglycans, such as the noncova- lent binding of the mast cell proteases, tryptase, chy- mase, and carboxypeptidase A, tryptase being in a complex distinct from that of the other two enzymes [54].

A wide variety of biological functions have been attributed to heparin as an extracellular mast cell mediator. Probably the most widely recognized is its capacity to interfere with blood coagulation by enhancing the ability of antithrombin 3 (AT-III) to inhibit the serine proteases involved in the coagulation cascade. Heparin neutralizes the cytotoxic actions of the eosinophil-derived basic proteins [55] and has anticomplementary activity [56]. Heparin also modu- lates the function of the mast cell proteases. It stabilizes

Fig. 30.5. Heparin and chondroitin sulfate proteoglycans

tryptase in its biologically active tetrameric form at neutral pH [57]. As there appear to be no endogenous inhibitors of tryptase, it is likely that it is restricted to having only very local effects because, as it diffuses away from heparin, it rapidly dissociates into inactive monomers. Heparin also enhances the activity of chy- mase [58].

30.2.3

Neutral Proteases

Neutral proteases comprise the majority of the protein of the secretory granules of human mast cells and rep- resent the major mediators of this cell type on a weight basis, MCTCcontaining up to 60 pg proteases per cell (Fig. 30.6). Until recently, the study of these abundant mast cell products has been neglected, but they are now attracting attention as important mediators of allergic disease and as potential targets for therapeutic inter- vention.

The major mast cell protease, tryptase, originally identified by Schwartz in 1981 [59], is a ~130-kDa tet- rameric serine protease which is stored in a fully active form complexed with heparin in the granule [60, 61].

Tryptase is secreted in proteoglycan complexes with molecular weights of 200 – 250 and 400 – 560 kDa [54].

The large size of these active complexes will severely limit diffusion away from sites of mast cell activation and helps to explain why increases in circulating levels of tryptase seem to occur only following the massive mast cell activation of anaphylactic shock. In the extra- cellular environment, the neutral pH allows tryptase to

Fig. 30.6.

(5)

become enzymatically functional. The properties of tryptase pertinent to the skin include: cleavage of the vasodilator neuropeptide, calcitonin gene-related pep- tide (CGRP) [62]; a kallikrein-like activity [63]; and, cleavage of matrix components, including 75-kDa gela- tinase/type IV collagen, fibronectin and type VI colla- gen [64, 65], which would be in keeping with participa- tion in processes of tissue destruction and remodeling.

Tryptase also has mitogenic activity for fibroblasts [66].

Chymase is a 30-kDa monomeric protease stored in the same secretory granules as tryptase in the MCTC subset of mast cells [67]. Like tryptase, chymase is pre- sent in a catalytically active form in the granule [68].

Chymase is inhibited by the circulating inhibitors -an- tichymotrypsin and 1-antitrypsin [69], an absolute necessity when it is realized that this protease can cleave angiotensin I to angiotensin II more effectively than angiotensin-converting enzyme [70]. Chymase degrades the neuropeptide neurotensin [71], but not substance P or VIP [70]. A role for chymase in tissue destruction and remodeling is also suggested from its ability to activate stromelysin and interstitial collage- nase, to convert procollagen I to collagen-sized frag- ments [72], and to degrade type IV collagen [73]. Fur- thermore, incubation of fresh human skin with human chymase can result in extensive separation of the epi- dermal-dermal layers [74].

Two other proteinases, carboxypeptidase and cathepsin G, have been associated with the MCTCsub- set of human mast cells [75, 76]. Carboxypeptidase is a unique 34.5-kDa metalloproteinase which removes the carboxyl terminal residues from a range of peptides, including angiotensin, leu5-enkephalin, kinetensin, neuromedin N, and neurotensin. Cathepsin G is a chy- motryptic enzyme with a structure seemingly identical to neutrophil cathepsin G. When mast cells are activat- ed, chymase, carboxypeptidase, and cathepsin G are released together in a 400 to 500-kDa complex with proteoglycan and are likely to act in concert with the other enzymes to degrade proteins.

Although few formal studies have been performed on the part played by mast cell proteases in inflamma- tory diseases of the skin, it is clear that they have a potential role. As early as 1978, Wintroub and col- leagues [77] reported evidence of mast cell degranula- tion in lesions of bullous pemphigoid and suggested a causal relationship. This suggestion was reinforced by the studies of Briggman and colleagues [74], who

showed that the epidermal-dermal junction is highly susceptible to neutral serine proteases located in mast cells and neutrophils. Finally, the finding of IgE auto- antibodies in bullous pemphigoid suggested a way in which mast cells may be activated in the disease [78].

Another area of research which may involve mast cell proteases is the recently made association between Netherton’s syndrome, a congenital skin disorder whose symptoms include severe ichthyosis, and poly- morphisms in SPINK5, a gene encoding for the serine protease inhibitor LEKTI [79, 80]. In vitro, LEKTI has been shown to inhibit the serine proteases plasmin, subtilisin A, cathepsin G, human neutrophil elastase, and trypsin, but not chymotrypsin [81]. Whether or not LEKTI inhibits mast cell tryptase is not yet known.

However, the findings that polymorphisms of SPINK5 are also associated with atopic dermatitis, asthma, and high IgE levels [82 – 84] suggests an axial role for prote- ases in allergic disease.

30.2.4 Cytokines

The concept that the mast cell, by the generation of proinflammatory cytokines, plays a pivotal role in the stimulation and maintenance of allergic inflammation is now well established. In addition, human purified lung mast cells incubated with SCF and anti-IgE express mRNA for IL-3, IL-4, IL-5, IL-6, IL-10, IL-13, GM-CSF, and TNF[ but not IL-2 or IFN * [85].

The first association of TNF[ with mast cells was made in 1980 [86] while examining the antitumor effects of murine mast cells. Initial studies relating to human mature mast cells showed that skin explants in vitro stimulated with mast cell secretagogues produced a factor that stimulated endothelial-leukocytic adhe- sion molecule-1 (ELAM-1) expression on endothelial cells. Inhibition of this response by specific antibodies strongly suggested that it was TNF[ [87]. TNF [ is stored preformed within the granules of human skin mast cells and is released rapidly after initiation of an allergic response [88]. In contrast, macrophages and lymphocytes, which also produce large amounts of TNF[ , have little or no capacity to store it and thus only generate it slowly after transcription. Functionally, TNF[ is a key cytokine in allergic inflammation, being required to activate NF-κB, a transcription factor that upregulates the expression of mRNA for TNF[ , GM- CSF, IL-8, IL-2, IL-6, E selectin, ICAM-1, and VCAM-1

(6)

in a variety of cells including endothelial, epithelial, and mast cells [89, 90].

Of the other NF-κB-associated both IL-8 and GM- CSF have also been demonstrated to be synthesized and stored by mast cells [91, 92].

The other group of cytokines are the so-called TH2 cytokines, including IL-4, IL-5, and IL-13 which are responsible for promoting and maintaining many aspects of the allergic response [93, 94].

IL-4, the cytokine largely responsible for stimulat- ing and maintaining Th2 cell proliferation and switch- ing the B cell to IgE synthesis [95], is seen in approxi- mately 80 % of mast cells, both MCTand MCTC,of the bronchial mucosa [96, 97]. Furthermore, immunogold electron microscopy applied to ultrathin sections of human purified dispersed mast cells and biopsies has shown that IL-4 is localized to the secretory granules [98, 99]. The ability of mast cells to store IL-4 and release it upon stimulation has been suggested to be an important initiating event in allergic inflammation by stimulating the expansion of the repertoire of TH2 cytokine-producing cells in the local microenviron- ment [100].

IL-13, a cytokine with many properties in common with IL-4, has been demonstrated to be associated with mast cells in conjunctival biopsies from patients with seasonal allergic conjunctivitis [101] and be synthe- sized by mast cells in vitro [102, 103].

IL-5, which is crucial for the maturation, activation, and survival of eosinophils, has also been localized to human mast cell [93, 104]. Unlike IL-4, IL-5 is present only in 10 % of mast cells found to be IL-5 positive and is restricted to MCT[105], the subset of mast cells that are under T cell control and that increase in numbers in parasitic infestation and at sites of chronic allergic inflammation. Furthermore, in vitro studies have shown that IgE-dependent stimulation of human lung mast cells induces IL-5 production [106, 107] which may be suppressed by dexamethasone [108].

30.3 Conclusions

Mast cells possess the armory to participate actively in many forms of inflammation in the skin. Perhaps the most obvious associations are with urticaria, which, in terms of appearance, is the clinical counterpart of the wheal and flare response induced by the intradermal

injection of allergen, codeine, or histamine [40]. Per- haps the most direct evidence for mast cell involve- ment in urticaria is the study in cold-induced urticaria performed by Anderson and colleagues [109]. They inserted microdialysis fibers in the upper dermis below a site of challenge with an ice cube. During the warm- ing-up period following challenge, the development of the wheal was paralleled by an up to 80-fold increase of the levels of histamine in the dialysis effluent. Also, in chronic idiopathic urticaria, circulating autoanti- bodies against Fc5 RI, IgE, or both, occur in approxi- mately one third of patients (Fig. 30.7). Injection of autologous serum containing these antibodies causes a wheal and flare response, suggesting degranulation of cutaneous mast cells to be the cause of the urticarial condition [110, 111]. The picture in atopic eczema is far less clear. Clearly, IgE-bearing cells are axial in the dis- ease mechanism, but in the skin these include cells of the Langerhans’ cell/dendritic cell lineage as well as mast cells [112]. Thus, while the dermal mast cell plays a role in many cutaneous conditions, it is not necessari- ly a major player in them all.

Fig. 30.7. Auto antibodies in chronic urticaria

(7)

References

1. Ehrlich P (1876) Beiträge zur Kenntnis der Anilinfärbun- gen und ihrer Verwendung in der mikroskopischen Tech- nik. Arch Mikr Anat 13:263 – 277

2. Nakahata T, Toru H (2002) Cytokines regulate develop- ment of human mast cells from hematopoietic progeni- tors. Int J Hematol 75:350 – 356

3. Castells MC, Friend DS, Burckart GJ (1996) The presence of membrane bound stem cell factor on highly immature nonmetachromatic mast cells in the peripheral blood of a patient with aggressive systemic mastocytosis. J Allergy Clin Immunol 98:831 – 840

4. Saito H, Ebisawa M, Tachimoto H, Shichijo M, Fukagawa K, Matsumoto K, Iikura Y, Awaji T, Tsujimoto G, Yanagida M, Uzumaki H, Takahashi G, Tsuji K, Nakahata T (1996) Selective growth of human mast cells induced by Steel fac- tor, IL- 6, and prostaglandin E2 from cord blood mononu- clear cells. J Immunol 157:343 – 350

5. Galli SJ (1990) New insight into „the riddle of the mast cells”: microenvironmental regulation of mast cell devel- opment and phenotypic heterogeneity. Lab Invest 62:

5 – 33

6. Irani AA, Schechter NM, Craig SS, Debois G, Schwartz LB (1986) Two types of mast cells that have distinct neutral protease compositions. Proc Natl Acad Sci 82:1214 – 1218 7. Cowen T, Trigg P, Eady RA (1979) Distribution of mast

cells in human dermis: development of a mapping tech- nique. Br J Dermatol 100:635 – 640

8. Eady RAJ (1976) The mast cell: distribution and morphol- ogy. Clin Exp Dermatol 1:313 – 329

9. Wiesner-Menzel L, Schulz B, Vakilzadeh F, Czarnetzki BM (1981) Electron microscopical evidence for a direct con- tact between nerve fibres and mast cells. Acta Derm Vene- reol (Stockh) 61:465 – 469

10. Irani AA, Garriga MM, Metcalfe DD, Schwartz LB (1990) Mast cells in cutaneous mastocytosis: accumulation of the MCTCtype. Clin Exp Allergy 20:53 – 58

11. Mikhail GR, Miller-Milinska A (1964) Mast cell popula- tions in human skin. J Invest Dermatol 43:249 – 254 12. Eady RAJ, Cowen T, Marshall TF, Plummer V, Greaves MW

(1979) Mast cell population density, blood vessel density and histamine content of normal human skin. Br J Derma- tol 100:623 – 633

13. Lichtig C, Haim S, Gilhar A, Hammel I, Ludatscher R (1981) Mast cells in Behcet’s disease: ultrastructural and histamine content studies. Dermatologica 162:167 – 174 14. Natbony SF, Phillips ME, Elias JM, Godfrey HP, Kaplan AP

(1983) Histologic studies of chronic idiopathic urticaria. J Allergy Clin Immunol 71:177 – 183

15. Garriga MM, Friedman MM, Metcalfe DD (1988) A survey of the number and distribution of mast cells in the skin of patients with mast cell disorders. J Allergy Clin Immunol 82:425 – 432

16. Cawley EP, Hoch-ligeti C (1961) Association of tissue mast cells and skin tumours. Arch Dermatol 83:92 – 96 17. Carr NJ, Warren AY (1993) Mast cell numbers in melano-

cytic naevi and cutaneous neurofibromas. J Clin Pathol 46:

86 – 87

18. Duncan LM, Richards LA, Mihm Jr MC (1998) Increased

mast cell density in invasive melanoma. J Cutan Pathol 25:11 – 15

19. Malaviya R, Morrison AR, Pentland AP (1996) Histamine in human epidermal cells is induced by ultraviolet light injury. J Invest Dermatol 106:785 – 789

20. Søndergaard J, Zachariae H (1968) Epidermal histamine.

Arch Klin Exp Dermatol 233:323 – 328

21. Fitzsimons C, Engel N, Duran H, Policastro L, Cricco G, Martin G, Molinari B, Rivera E (2001) Histamine produc- tion in mouse epidermal keratinocytes is regulated during cellular differentiation. Inflamm Res 50(2):S100 – S101 22. Coleman JW, Godfrey RC (1981) The number and affinity

of IgE receptors on dispersed human lung mast cells.

Immunology 44:859 – 863

23. Kinet JP (1999) The high-affinity IgE receptor (Fc5 RI):

from physiology to pathology. Annu Rev Immunol 17:

931 – 972

24. Letourneur O, Sechi S, Willette-Brown J, Robertson MW, Kinet JP (1995) Glycosylation of human truncated Fc epsi- lon RI alpha chain is necessary for efficient folding in the endoplasmic reticulum. J Biol Chem 270:8249 – 8256 25. Lin S, Cicala C, Scharenberg AM, Kinet JP (1996) The

Fc5 RI q subunit functions as an amplifier of Fc 5 RI * -medi- ated cell activation signals. Cell 85:985 – 995

26. El Lati SG, Dahinden CA, Church MK (1994) Complement peptides C3a- and C5a-induced mediator release from dis- sociated human skin mast cells. J Invest Dermatol 102:

803 – 806

27. Valent P, Schernthaner GH, Sperr WR, Fritsch G, Agis H, Willheim M, Buhring HJ, Orfao A, Escribano L (2001) Var- iable expression of activation-linked surface antigens on human mast cells in health and disease. Immunol Rev 179:74 – 81

28. Lowman MA, Benyon RC, Church MK (1988) Character- ization of neuropeptide-induced histamine released from human dispersed skin mast cells. Br J Pharmacol 95:121 – 130

29. Stellato C, De Paulis A, Cirillo R, Mastronardi P, Mazzarella B, Marone G (1991) Heterogeneity of human mast cells and basophils in response to muscle relaxants. Anesthesiology 74:1078 – 1086

30. Church MK, El Lati SG, Caulfield JP (1991) Neuropeptide- induced secretion from human skin mast-cells. Int Arch Allergy Appl Immunol 94:310 – 318

31. Mousli M, Hugli TE, Landry Y, Bronner C (1994) Peptider- gic pathway in human skin and rat peritoneal mast cell activation. Immunopharmacology 27:1 – 11

32. Caulfield JP, el Lati S, Thomas G, Church MK (1990) Disso- ciated human foreskin mast cells degranulate in response to anti- IgE and substance P. Lab Invest 63:502 – 510 33. Uvnas B, Aborg C-H, Bergendorff A (1970) Storage of his-

tamine in mast cells. Evidence for ionic binding of hista- mine to protein carboxyls in the granule heparin protein complex. Acta Physiol Scand 336(Suppl):1 – 26

34. Benyon RC, Lowman MA, Church MK (1987) Human skin mast cells:their dispersion, purification and secretory characteristics. J Immunol 138:861 – 867

35. Lessof MH, Gant V, Hinuma K, Murphy GM, Dowling RH (1990) Recurrent urticaria and reduced diamine oxidase activity. Clin Exp Allergy 20:373 – 376

(8)

36. Lewis T (1927) The blood vessels of human skin and their responses. Shaw & Son, London

37. Schoeffter P, Godfraind T (1989) Histamine receptors in the smooth muscle of human internal mammary artery and saphenous vein. Pharmacol Toxicol 64:64 – 71 38. Jow F, Numann R (2000) Histamine increases [Ca(2+)](in)

and activates Ca-K and nonselective cation currents in cul- tured human capillary endothelial cells. J Membr Biol 173:107 – 116

39. Raud J (1989) Intravital microscopic studies on acute mast cell-dependent inflammation. Acta Physiol Scand 578:1 – 58

40. Petersen LJ, Church MK, Skov PS (1997) Histamine is released in the wheal but not the flare following challenge of human skin in vivo: A microdialysis study. Clin Exp Allergy 27:284 – 296

41. Schmelz M, Michael K, Weidner C, Schmidt R, Torebjork HE, Handwerker HO (2000) Which nerve fibers mediate the axon reflex flare in human skin? Neuroreport 11:645 – 648 42. Schmelz M (2001) A neural pathway for itch. Nat Neurosci

4:9 – 10

43. Robinson C, Benyon C, Holgate ST, Church MK (1989) The IgE- and calcium-dependent release of eicosanoids and histamine from human purified cutaneous mast cells. J Invest Dermatol 93:397 – 404

44. Sampson AP, Siddiqui S, Buchanan D, Howarth PH, Holga- te ST, Holloway JW, Sayers I (2000) Variant LTC4synthase allele modifies cysteinyl leukotriene synthesis in eosino- phils and predicts clinical response to zafirlukast.Thorax 55 2:S28 – S31

45. Capella GL, Grigerio E, Altomare G (2001) A randomized trial of leukotriene receptor antagonist montelukast in moderate-to-severe atopic dermatitis of adults. Eur J Der- matol 11:209 – 213

46. Erbagci Z (2002) The leukotriene receptor antagonist montelukast in the treatment of chronic idiopathic urti- caria: a single-blind, placebo-controlled, crossover clinical study. J Allergy Clin Immunol 110:484 – 488

47. Haas N, Motel K, Czarnetzki BM (1995) Comparative immunoreactivity of the eosinophil constituents MBP and ECP in different types of urticaria. Arch Dermatol Res 287:180 – 185

48. Kiehl P, Falkenberg K, Vogelbruch M, Kapp A (2001) Tis- sue eosinophilia in acute and chronic atopic dermatitis: a morphometric approach using quantitative image analysis of immunostaining. Br J Dermatol 145:720 – 729

49. Leiferman KM, Fujisawa T, Gray BH, Gleich GJ (1990) Extracellular deposition of eosinophil and neutrophil granule proteins in the IgE-mediated cutaneous late phase reaction. Lab Invest 62:579 – 589

50. Morita H, Yamamoto K, Kitano Y (1995) Elevation of serum major basic protein in patients with atopic dermati- tis. J Dermatol Sci 9:165 – 168

51. Kjellen L, Lindahl U (1991) Proteoglycans: structures and interactions. Annu Rev Biochem 60:443 – 475

52. Metcalfe DD, Lewis RA, Silbert JE, Rosenberg RD, Wasser- man SI, Austen KF (1979) Isolation and characterisation of heparin from human lung. J Clin Invest 4:1537 – 1543 53. Stevens RL, Fox CC, Lichtenstein LM, Austen KF (1988)

Identification of chondroitin sulfate E proteoglycans and

heparin proteoglycans in the secretory granules of human lung mast cells. Proc Natl Acad Sci 85:2284 – 2287 54. Goldstein SM, Leong J, Schwartz LB, Cooke D (1992) Prote-

ase composition of exocytosed human skin mast cell pro- tease- proteoglycan complexes: Tryptase resides in a com- plex distinct from chymase and carboxypeptidase. J Immunol 148:2475 – 2482

55. Fredens K, Dahl R, Venge P (1991) In vitro studies of the interaction between heparin and eosinophil cationic pro- tein. Allergy 46:27 – 29

56. Kazatchkine MD, Fearon DT, Silbert JE, Austen KF (1979) Surface-associated heparin inhibits zymosan-induced activation of the alternative complement pathway by aug- menting the regulatory action of the control proteins on particle bound C3b. J Exp Med 150:1202 – 1215

57. Schechter NM, Eng GY, Selwood T, McCaslin DR (1995) Structural changes associated with the spontaneous inac- tivation of the serine proteinase human tryptase. Bio- chemistry 34:10628 – 10638

58. McEuen AR, Sharma B, Walls AF (1995) Regulation of the activity of human chymase during storage and release from mast cells: the contributions of inorganic cations, pH, heparin and histamine. Biochim Biophys Acta 1267:115 – 121

59. Schwartz LB, Lewis RA, Austen KF (1981) Tryptase from human pulmonary mast cells: purification and character- ization. J Biol Chem 256:11939 – 11943

60. Church MK, Holgate ST, Shute JK, Walls AF, Sampson AP (1998) Mast cell derived mediators. In Middleton E, Reed CE, Ellis EF, Adkinson NF, Yuginger J, Busse WW (eds) Allergy: principles and practice. Mosby, St Louis, pp 146 – 161

61. Schwartz LB (1990) Tryptase from human mast cells: bio- chemistry, biology and clinical utility. Monogr Allergy 27:90 – 113

62. Walls AF, Brain SD, Desai A, Jose PJ, Hawkings E, Church MK, Williams TJ (1992) Human mast cell tryptase attenu- ates the vasodilator activity of calcitonin gene-related pep- tide. Biochem Pharmacol 43:1243 – 1248

63. Imamura T, Dubin A, Moore W, Tanaka R, Travis J (1996) Induction of vascular permeability enhancement by human tryptase: dependence on activation of prekalli- krein and direct release of bradykinin from kininogens.

Lab Invest 74:861 – 870

64. Kielty CM, Lees M, Shuttleworth CA, Woolley D (1993) Catabolism of intact type VI collagen microfibrils: suscep- tibility to degradation by serine proteinases. Biochem Bio- phys Res Commun 191:1230 – 1236

65. Lohi J, Harvima I, Keski-Oja J (1992) Pericellular sub- strates of human mast cell tryptase: 72, 000 Dalton gelati- nase and fibronectin. J Cell Biochem 50:337 – 349 66. Levi-Schaffer F, Piliponsky AM (2003) Tryptase, a novel

link between allergic inflammation and fibrosis. Trends Immunol 24:158 – 161

67. Schechter NM, Choi JK, Slavin DA, Deresienski DT, Saya- ma S, Dong G, Lavker RM, Proud D, Lazarus GS (1986) Identification of a chymotrypsin-like proteinase in human mast cells. J Immunol 137:962 – 970

68. Huntley JF, Newlands GFJ, Gibson S, Ferguson A, Miller HRP (1985) Histochemical demonstration of chymotryp-

(9)

sin like serine esterases in mucosal mast cells in four spe- cies including man. J Clin Pathol 38:375 – 384

69. Schechter NM, Sprows JL, Schoenberger OL, Lazarus GS, Cooperman BS, Rubin H (1989) Reaction of human skin chymotrypsin-like proteinase chymase with plasma pro- teinase inhibitors. J Biol Chem 264:21308 – 21315 70. Urata H, Kinoshita A, Misono KS, Bumpus FM, Husain A

(1990) Identification of a highly specific chymase as the major angiotensin II forming enzyme in the human heart.

J Biol Chem 265:22348 – 22357

71. Goldstein SM, Leong J, Bunnett NW (1991) Human Mast Cell proteases hydrolyze neurotensin, kinetensin and Leu5-enkephalin. Peptides 12:995 – 1000

72. Buckley MG, Gallagher PJ, Walls AF (1996) Altered mast cell phenotype in synovial tissues of patients with osteoar- thritis. J Allergy Clin Immunol 97:267

73. Sage H, Woodbury RG, Bornstein P (1979) Structural stud- ies on human type IV collagen. J Exp Med 254(19):9893 – 9900

74. Briggman RA, Schechter NM, Fraki JE, Lazarus GS (1984) Degradation of the epidermal-dermal junction by a pro- teolytic enzyme from human skin and human polymor- phonuclear leukocytes. J Exp Med 160:1027 – 1042 75. Goldstein SM, Kaempfer CE, Kealey JT, Wintroub BU

(1989) Human mast cell carboxypeptidase. Purification and characterization. J Clin Invest 83:1630 – 1636 76. Irani AM, Goldstein SM, Wintroub BU, Bradford T, Schwartz

LB (1991) Human mast cell carboxypeptidase. Selective localization to MCTC cells. J Immunol 147:247 – 253 77. Wintroub BU, Mihm MC, Jr., Goetzl EJ, Soter NA, Austen

KF (1978) Morphologic and functional evidence for release of mast-cell products in bullous pemphigoid. N Engl J Med 298:417 – 421

78. Dimson OG, Giudice GJ, Fu CL, Van den BF, Warren SJ, Janson MM, Fairley JA (2003) Identification of a potential effector function for IgE autoantibodies in the organ-spe- cific autoimmune disease bullous pemphigoid. J Invest Dermatol 120:784 – 788

79. Chavanas S, Bodemer C, Rochat A, Hamel-Teillac D, Ali M, Irvine AD, Bonafe JL, Wilkinson J, Taieb A, Barrandon Y, Harper JI, de Prost Y, Hovnanian A (2000) Mutations in SPINK5, encoding a serine protease inhibitor, cause Netherton syndrome. Nat Genet 25:141 – 142

80. Sprecher E, Chavanas S, DiGiovanna JJ, Amin S, Nielsen K, Prendiville JS, Silverman R, Esterly NB, Spraker MK, Gue- lig E, de Luna ML, Williams ML, Buehler B, Siegfried EC, Van Maldergem L, Pfendner E, Bale SJ, Uitto J, Hovnanian A, Richard G (2001) The spectrum of pathogenic muta- tions in SPINK5 in 19 families with Netherton syndrome:

implications for mutation detection and first case of pre- natal diagnosis. J Invest Dermatol 117:179 – 187

81. Mitsudo K, Jayakumar A, Henderson Y, Frederick MJ, Kang Y, Wang M, El Naggar AK, Clayman GL (2003) Inhi- bition of serine proteinases plasmin, trypsin, subtilisin A, cathepsin G, and elastase by LEKTI: a kinetic analysis. Bio- chemistry 42:3874 – 3881

82. Kabesch M, Peters W, Carr D, Weiland S, von Mutius E (2002) A polymorphism in the gene SPINK5 is associated with asthma in a large German population sample. Am J Resp Crit Care Med 165:A808

83. Kato A, Fukai K, Oiso N, Hosomi N, Murakami T, Ishii M (2003) Association of SPINK5 gene polymorphisms with atopic dermatitis in the Japanese population. Br J Derma- tol 148:665 – 669

84. Walley AJ, Chavanas S, Moffatt MF, Esnouf RM, Ubhi B, Lawrence R, Wong K, Abecasis GR, Jones EY, Harper JI, Hovnanian A, Cookson WO (2001) Gene polymorphism in Netherton and common atopic disease. Nat Genet 29:175 – 178

85. Okayama Y, Semper A, Holgate ST, Church MK (1995b) Multiple cytokine messenger-rna expression in human mast-cells stimulated via fc-epsilon-ri. Int Arch Allergy Immunol 107:158 – 159

86. Farram E, Nelson DS (1980) Mouse mast cells as anti- tumor effector cells. Cell Immunol 55:294 – 301

87. Klein LM, Lavker RM, Matis WL, Murphy GF (1989) Degranulation of human mast cells induces an endothelial antigen central to leukocyte adhesion. Proc Natl Acad Sci 86:8972 – 8976

88. Walsh LJ, Trinchieri G, Waldorf HA, Whitaker D, Murphy GF (1991) Human dermal mast cells contain and release tumor necrosis factor alpha, which induces endothelial leukocyte adhesion molecule 1. Proc Natl Acad Sci 88:

4220 – 4224

89. Coward WR, Okayama Y, Sagara H, Wilson SJ, Holgate ST, Church MK (2002) NF-kappaB and TNF-alpha: a positive autocrine loop in human lung mast cells? J Immunol 169:5287 – 5293

90. Ghosh S, May MJ, Kopp EB (1998) NF-κB and rel proteins:

Evolutionarily conserved mediators of immune responses.

Annu Rev Immunol 16:225 – 260

91. Möller A, Lippert U, Lessmann D, Kolde G, Hamann K, Welker P, Schadendorf D, Rosenbach T, Luger T, Czarnetz- ki BM (1993) Human mast cells produce IL-8. J Immunol 151:3261 – 3266

92. Okayama Y, Kobayashi H, Ashman LK, Dobashi K, Naka- zawa T, Holgate ST, Church MK, Mori M (1998) Human lung mast cells are enriched in the capacity to produce granulocyte-macrophage colony-stmulating factor in response to IgE-dependent stimulation. Eur J Immunol 28:708 – 715

93. Bradding P, Roberts JA, Britten KM, Montefort S, Djuka- novic R, Mueller R, Heusser CH, Howarth PH, Holgate ST (1994) Interleukin-4, -5, and -6 and tumor necrosis factor- alpha in normal and asthmatic airways: evidence for the human mast cell as a source of these cytokines. Am J Respir Cell Mol Biol 10:471 – 480

94. Kobayashi H, Okayama Y, Ishizuka T, Pawankar R, Ra C, Mori M (1998) Production of IL-13 by human lung mast cells in response to Fcepsilon receptor cross-linkage. Clin Exp Allergy 28:1219 – 1227

95. Anderson GP, Coyle AJ (1994) TH2 and ’TH2-like’ cells in allergy and asthma: pharmacological perspectives. Trends Pharmacol Sci 15:324 – 332

96. Bradding P, Feather IH, Howarth PH, Mueller R, Roberts JA, Britten K, Bews JPA, Hunt TC, Okayama Y, Heusser CH, Bullock GR, Church MK, Holgate ST (1992) Interleukin 4 is localized to and released by human mast cells. J Exp Med 176:1381 – 1386

97. Bradding P, Okayama Y, Howarth PH, Church MK, Holgate

(10)

ST (1995a) Heterogeneity of human mast cells based on cytokine content. J Immunol 155:297 – 307

98. Wilson SJ, Bradding P, Heusser C, Holgate ST, Howarth PH (1994) The subcellular localisation of interleukin 4 in the respiratory mucosa using immunoelectron microsco- py. Clin Exp Allergy 24:980

99. Wilson SJ, Shute JK, Holgate ST, Howarth PH, Bradding P (2000) Localization of interleukin IL-4 but not IL-5 to human mast cell secretory granules by immunoelectron microscopy. Clin Exp Allergy 30:493 – 500

100. Wang M, Saxon A, Diaz-Sanchez D (1999) Early IL-4 pro- duction driving Th2 differentiation in a human in vivo allergic model is mast cell derived. Clin Immunol 90:

47 – 54

101. Anderson DF, Zhang S, Bradding P, McGill JI, Holgate ST, Roche WR (2001) The relative contribution of mast cell subsets to conjunctival TH2-like cytokines. Invest Oph- thalmol Vis Sci 42:995 – 1001

102. Burd PR, Thompson WC, Max EE, Mills FC (1995) Acti- vated mast-cells produce interleukin-13. J Exp Med 181:

1373 – 1380

103. Kanbe N, Kurosawa M, Yamashita T, Kurimoto F, Yanagi- hara Y, Miyachi Y (1999) Cord-blood-derived human cul- tured mast cells produce interleukin 13 in the presence of stem cell factor. Int Arch Allergy Immunol 119:138 – 142 104. Bradding P, Feather IH, Wilson S, Bardin PG, Heusser CH,

Holgate ST, Howarth PH (1993) Immunolocalization of cytokines in the nasal mucosa of normal and perennial rhinitic subjects: the mast cell as a source of IL-4, IL-5 and IL-6 in human allergic mucosal inflammation. J Immunol 151:3853 – 3865

105. Bradding P, Okayama Y, Howarth PH, Church MK, Holga- te ST (1995b) Heterogeneity of human mast-cells based on cytokine content. J Immunol 155:297 – 307

106. Jaffe JS, Glaum MC, Raible DG, Post TJ, Dimitry E, Govin- darao D, Wang Y, Schulman ES (1995) Human lung mast cell IL-5 gene and protein expression: temporal analysis of upregulation following IgE-mediated activation. Am J Respir Cell Mol Biol 13:665 – 675

107. Okayama Y, Petit-Fr`ere C, Kassel O, Semper A, Tunon de Lara JM, Holgate ST, Church MK (1995a) The IgE-depen- dent expression of mRNA for IL-4 and IL-5 in human lung mast cells. J Immunol 155:1796 – 1808

108. Glaum MC, Jaffe JS, Gillespie DH, Raible DG, Post TJ, Wang Y, Dimitry E, Schulman ES (1995) IgE-dependent expression of interleukin-5 mRNA and protein in human lung: modulation by dexamethasone. Clin Immunol Immunopathol 75:171 – 178

109. Andersson T, Wardell K, Anderson C (1995) Human in vivo cutaneous microdialysis: estimation of histamine release in cold urticaria. Acta Derm Venereol (Stockh ) 75:343 – 347

110. Grattan CE, Francis DM, Hide M, Greaves MW (1991) Detection of circulating histamine releasing autoanti- bodies with functional properties of anti-IgE in chronic urticaria. Clin Exp Allergy 21:695 – 704

111. Sabroe RA, Fiebiger E, Francis DM, Maurer D, Seed PT, Grattan CE, Black AK, Stingl G, Greaves MW, Barr RM (2002) Classification of anti-Fc5 RI and anti-IgE autoanti- bodies in chronic idiopathic urticaria and correlation with disease severity. J Allergy Clin Immunol 110:492 – 499

112. Klubal R, Osterhoff B, Wang B, Kinet JP, Maurer D, Stingl G (1997) The high-affinity receptor for IgE is the predom- inant IgE-binding structure in lesional skin of atopic der- matitis patients. J Invest Dermatol 108:336 – 342

Riferimenti

Documenti correlati

In recent years, some emerging evidence has indicated that p53 is not the only factor controlling the relationship between ribosome biogenesis and cell proliferation, thus revealing

a) Anderluh,M.; Carbohydrates-Comprensive studies on Glycobiology and Glycotechnology 2012; c.. Ren,T.; Liu,D.;

The mixer has a minimum conversion loss of 7.8 dB and has a third order intermodulation intercept point of 30.3 dBm.. The mixer is designed using a harmonic-balance simula-

This study investigates how certain organisational, managerial, and technological factors enable the adoption of software as a service (SaaS) model for customer

Ci sono state poi attività con i volontari dell’associazione AGESCI (Associazione Guide e Scout Cattolici Italiani) e l’organizzazione di cene con il vicinato

It is also possible to further reduce the up-sweep angle in the case of the reference Prandtlplane configuration by reducing the horizontal distance between its center of gravity

This involves reading and understanding a patent application, searching the prior art (including prior patent applications and patents, scientific literature databases, etc.)

Ad un gruppo composto in prevalenza da scialpinisti (più qualche ciaspolatore) è stato proposto un questionario on-line che sottopone loro alcune