Hostname: page-component-7bb8b95d7b-dvmhs Total loading time: 0 Render date: 2024-09-05T23:27:26.895Z Has data issue: false hasContentIssue false

Molecular mechanisms of leukocyte trafficking in T-cell-mediated skin inflammation: insights from intravital imaging

Published online by Cambridge University Press:  20 August 2009

James A. Deane
Affiliation:
Centre for Inflammatory Diseases, Monash University Department of Medicine, Monash Medical Centre, Clayton, Australia.
Michael J. Hickey*
Affiliation:
Centre for Inflammatory Diseases, Monash University Department of Medicine, Monash Medical Centre, Clayton, Australia.
*
*Corresponding author: Michael Hickey, Centre for Inflammatory Diseases, Monash University Department of Medicine, Monash Medical Centre, 246 Clayton Rd., Clayton, 3168, Australia, Tel: +61 3 9594 5519; Fax: +61 3 9594 6495; E-mail: michael.hickey@med.monash.edu.au

Abstract

Infiltration of T cells is a key step in the pathogenesis of the inflammatory skin diseases atopic dermatitis, allergic contact dermatitis and psoriasis. Understanding the mechanisms of T cell recruitment to the skin is therefore of fundamental importance for the discovery and application of novel therapies for these conditions. Studies of both clinical samples and experimental models of skin inflammation have implicated specific adhesion molecules and chemokines in lymphocyte recruitment. In particular, recent studies using advanced in vivo imaging techniques have greatly increased our understanding of the kinetics and molecular basis of this process. In this review, we summarise the current understanding of the cellular immunology of antigen-driven dermal inflammation and the roles of adhesion molecules and chemokines. We focus on results obtained using intravital microscopy to examine the dermal microvasculature and interstitium to determine the mechanisms of T cell recruitment and migration in experimental models of T-cell-mediated skin inflammation.

Type
Review Article
Copyright
Copyright © Cambridge University Press 2009

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

References

1Saint-Mezard, P. et al. (2004) Allergic contact dermatitis. European Journal of Dermatology 14, 284-295Google ScholarPubMed
2Jin, H. et al. (2009) Animal models of atopic dermatitis. Journal of Investigative Dermatology 129, 31-40CrossRefGoogle ScholarPubMed
3Hamid, Q., Boguniewicz, M. and Leung, D.Y. (1994) Differential in situ cytokine gene expression in acute versus chronic atopic dermatitis. Journal of Clinical Investigation 94, 870-876CrossRefGoogle ScholarPubMed
4Fallon, P.G. et al. (2009) A homozygous frameshift mutation in the mouse Flg gene facilitates enhanced percutaneous allergen priming. Nature Genetics 41, 602-608CrossRefGoogle ScholarPubMed
5Griffiths, C.E. and Barker, J.N. (2007) Pathogenesis and clinical features of psoriasis. Lancet 370, 263-271CrossRefGoogle ScholarPubMed
6Ellis, C.N. et al. (1986) Cyclosporine improves psoriasis in a double-blind study. Journal of the American Medical Association 256, 3110-3116CrossRefGoogle Scholar
7Martin, S.F. (2004) T lymphocyte-mediated immune responses to chemical haptens and metal ions: implications for allergic and autoimmune disease. International Archives of Allergy and Immunology 134, 186-198CrossRefGoogle ScholarPubMed
8Gocinski, B.L. and Tigelaar, R.E. (1990) Roles of CD4+ and CD8+ T cells in murine contact sensitivity revealed by in vivo monoclonal antibody depletion. Journal of Immunology 144, 4121-4128CrossRefGoogle ScholarPubMed
9Roediger, B. et al. (2008) Visualizing dendritic cell migration within the skin. Histochemistry and Cell Biology 130, 1131-1146CrossRefGoogle ScholarPubMed
10Robert, C. and Kupper, T.S. (1999) Inflammatory skin diseases, T cells, and immune surveillance. New England Journal of Medicine 341, 1817-1828CrossRefGoogle ScholarPubMed
11Szczepanik, M. et al. (2003) B-1 B cells mediate required early T cell recruitment to elicit protein-induced delayed-type hypersensitivity. Journal of Immunology 171, 6225-6235CrossRefGoogle Scholar
12Campos, R.A. et al. (2003) Cutaneous immunization rapidly activates liver invariant Valpha14 NKT cells stimulating B-1 B cells to initiate T cell recruitment for elicitation of contact sensitivity. Journal of Experimental Medicine 198, 1785-1796CrossRefGoogle ScholarPubMed
13Tsuji, R.F. et al. (2002) B cell-dependent T cell responses: IgM antibodies are required to elicit contact sensitivity. Journal of Experimental Medicine 196, 1277-1290CrossRefGoogle ScholarPubMed
14Sallusto, F. et al. (1999) Two subsets of memory T lymphocytes with distinct homing potentials and effector functions. Nature 401, 708-712CrossRefGoogle ScholarPubMed
15Askenase, P.W. et al. (2004) Extravascular T-cell recruitment requires initiation begun by Valpha14+ NKT cells and B-1 B cells. Trends in Immunology 25, 441-449CrossRefGoogle ScholarPubMed
16Ptak, W., Herzog, W.R. and Askenase, P.W. (1991) Delayed-type hypersensitivity initiation by early-acting cells that are antigen mismatched or MHC incompatible with late-acting, delayed-type hypersensitivity effector T cells. Journal of Immunology 146, 469-475CrossRefGoogle ScholarPubMed
17Tsuji, R.F. et al. (2000) Early local generation of C5a initiates the elicitation of contact sensitivity by leading to early T cell recruitment. Journal of Immunology 165, 1588-1598CrossRefGoogle ScholarPubMed
18Hwang, J.M. et al. (2004) A critical temporal window for selectin-dependent CD4+ lymphocyte homing and initiation of late-phase inflammation in contact sensitivity. Journal of Experimental Medicine 199, 1223-1234CrossRefGoogle ScholarPubMed
19Norman, M.U. et al. (2008) Multichannel fluorescence spinning disk microscopy reveals early endogenous CD4 T cell recruitment in contact sensitivity via complement. Journal of Immunology 180, 510-521CrossRefGoogle ScholarPubMed
20Doebis, C. et al. (2008) Cellular players and role of selectin ligands in leukocyte recruitment in a T-cell-initiated delayed-type hypersensitivity reaction. American Journal of Pathology 173, 1067-1076CrossRefGoogle Scholar
21Martin, S. et al. (2000) Peptide immunization indicates that CD8+ T cells are the dominant effector cells in trinitrophenyl-specific contact hypersensitivity. Journal of Investigative Dermatology 115, 260-266CrossRefGoogle ScholarPubMed
22Bour, H. et al. (1995) Major histocompatibility complex class I-restricted CD8+ T cells and class II-restricted CD4+ T cells, respectively, mediate and regulate contact sensitivity to dinitrofluorobenzene. European Journal of Immunology 25, 3006-3010CrossRefGoogle ScholarPubMed
23Bouloc, A., Cavani, A. and Katz, S.I. (1998) Contact hypersensitivity in MHC class II-deficient mice depends on CD8 T lymphocytes primed by immunostimulating Langerhans cells. Journal of Investigative Dermatology 111, 44-49CrossRefGoogle ScholarPubMed
24Yokozeki, H. et al. (2000) Signal transducer and activator of transcription 6 is essential in the induction of contact hypersensitivity. Journal of Experimental Medicine 191, 995-1004CrossRefGoogle ScholarPubMed
25Nakae, S. et al. (2002) Antigen-specific T cell sensitization is impaired in IL-17-deficient mice, causing suppression of allergic cellular and humoral responses. Immunity 17, 375-387CrossRefGoogle ScholarPubMed
26Wang, B. et al. (2000) CD4+ Th1 and CD8+ type 1 cytotoxic T cells both play a crucial role in the full development of contact hypersensitivity. Journal of Immunology 165, 6783-6790CrossRefGoogle ScholarPubMed
27Dilulio, N.A. et al. (1999) Groalpha-mediated recruitment of neutrophils is required for elicitation of contact hypersensitivity. European Journal of Immunology 29, 3485-34953.0.CO;2-B>CrossRefGoogle ScholarPubMed
28Atarashi, K. et al. (2009) Skin application of ketoprofen systemically suppresses contact hypersensitivity by inducing CD4(+) CD25(+) regulatory T cells. Journal of Dermatological Science 53, 216-221CrossRefGoogle ScholarPubMed
29Ring, S. et al. (2006) CD4+ CD25+ regulatory T cells suppress contact hypersensitivity reactions by blocking influx of effector T cells into inflamed tissue. European Journal of Immunology 36, 2981-2992CrossRefGoogle ScholarPubMed
30Xu, H., DiIulio, N.A. and Fairchild, R.L. (1996) T cell populations primed by hapten sensitization in contact sensitivity are distinguished by polarized patterns of cytokine production: interferon gamma-producing (Tc1) effector CD8+ T cells and interleukin (IL) 4/IL-10-producing (Th2) negative regulatory CD4+ T cells. Journal of Experimental Medicine 183, 1001-1012CrossRefGoogle ScholarPubMed
31Siegmund, K. et al. (2005) Migration matters: regulatory T-cell compartmentalization determines suppressive activity in vivo. Blood 106, 3097-3104CrossRefGoogle ScholarPubMed
32Mempel, T.R. et al. (2006) Regulatory T cells reversibly suppress cytotoxic T cell function independent of effector differentiation. Immunity 25, 129-141CrossRefGoogle ScholarPubMed
33Sather, B.D. et al. (2007) Altering the distribution of Foxp3(+) regulatory T cells results in tissue-specific inflammatory disease. Journal of Experimental Medicine 204, 1335-1347CrossRefGoogle ScholarPubMed
34Gerszten, R.E. et al. (1999) MCP-1 and IL-8 trigger firm adhesion of monocytes to vascular endothelium under flow conditions. Nature 398, 718-723CrossRefGoogle ScholarPubMed
35Phillipson, M. et al. (2006) Intraluminal crawling of neutrophils to emigration sites: a molecularly distinct process from adhesion in the recruitment cascade. Journal of Experimental Medicine 203, 2569-2575CrossRefGoogle Scholar
36Hickey, M.J. and Kubes, P. (2009) Intravascular immunity: the host-pathogen encounter in blood vessels. Nature Reviews Immunology 9, 364-375CrossRefGoogle ScholarPubMed
37Kanwar, S. et al. (1997) The association between α4-integrin, P-selectin, and E-selectin in an allergic model of inflammation. Journal of Experimental Medicine 185, 1077-1087CrossRefGoogle Scholar
38Kanwar, S. et al. (1999) Overlapping roles for L-selectin and P-selectin in antigen-induced immune responses in the microvasculature. Journal of Immunology 162, 2709-2716CrossRefGoogle ScholarPubMed
39Johnston, B., Issekutz, T.B. and Kubes, P. (1996) The α4-integrin supports leukocyte rolling and adhesion in chronically inflamed postcapillary venules in vivo. Journal of Experimental Medicine 183, 1995-2006CrossRefGoogle ScholarPubMed
40Hickey, M.J., Granger, D.N. and Kubes, P. (1999) Molecular mechanisms underlying IL-4-induced leukocyte recruitment in vivo: a critical role for the α4-integrin. Journal of Immunology 163, 3441-3448CrossRefGoogle Scholar
41James, W.G., Bullard, D.C. and Hickey, M.J. (2003) Critical role of the alpha 4 integrin/VCAM-1 pathway in cerebral leukocyte trafficking in lupus-prone MRL/fas(lpr) mice. Journal of Immunology 170, 520-527CrossRefGoogle ScholarPubMed
42Vachino, G. et al. (1995) P-selectin glycoprotein ligand-1 is the major counter-receptor for P-selectin on stimulated T cells and is widely distributed in non-functional form on many lymphocytic cells. Journal of Biological Chemistry 270, 21966-21974CrossRefGoogle ScholarPubMed
43Ley, K. et al. (2007) Getting to the site of inflammation: the leukocyte adhesion cascade updated. Nature Reviews Immunology 7, 678-689CrossRefGoogle Scholar
44Bromley, S.K., Mempel, T.R. and Luster, A.D. (2008) Orchestrating the orchestrators: chemokines in control of T cell traffic. Nature Immunology 9, 970-980CrossRefGoogle ScholarPubMed
45Springer, T.A. (1994) Traffic signals of lymphocyte recirculation and leukocyte emigration: The multistep paradigm. Cell 76, 301-314CrossRefGoogle ScholarPubMed
46Rao, R.M. et al. (2007) Endothelial-dependent mechanisms of leukocyte recruitment to the vascular wall. Circulation Research 101, 234-247CrossRefGoogle Scholar
47Shulman, Z. et al. (2009) Lymphocyte crawling and transendothelial migration require chemokine triggering of high-affinity LFA-1 integrin. Immunity 30, 384-396CrossRefGoogle ScholarPubMed
48Wegmann, F. et al. (2006) ESAM supports neutrophil extravasation, activation of Rho, and VEGF-induced vascular permeability. Journal of Experimental Medicine 203, 1671-1677CrossRefGoogle ScholarPubMed
49Stefanidakis, M. et al. (2008) Endothelial CD47 interaction with SIRPgamma is required for human T-cell transendothelial migration under shear flow conditions in vitro. Blood 112, 1280-1289CrossRefGoogle ScholarPubMed
50Thompson, R.D. et al. (2001) Platelet-endothelial cell adhesion molecule-1 (PECAM-1)-deficient mice demonstrate a transient and cytokine-specific role for PECAM-1 in leukocyte migration through the perivascular basement membrane. Blood 97, 1854-1860CrossRefGoogle ScholarPubMed
51Woodfin, A., Voisin, M.B. and Nourshargh, S. (2007) PECAM-1: a multi-functional molecule in inflammation and vascular biology. Arteriosclerosis Thrombosis and Vascular Biology 27, 2514-2523CrossRefGoogle ScholarPubMed
52Millan, J. et al. (2006) Lymphocyte transcellular migration occurs through recruitment of endothelial ICAM-1 to caveola- and F-actin-rich domains. Nature Cell Biology 8, 113-123CrossRefGoogle ScholarPubMed
53McHale, J.F. et al. (1999) Vascular endothelial cell expression of ICAM-1 and VCAM-1 at the onset of eliciting contact hypersensitivity in mice: evidence for a dominant role of TNF-alpha. Journal of Immunology 162, 1648-1655CrossRefGoogle ScholarPubMed
54Elices, M.J. et al. (1993) The integrin VLA-4 mediates leukocyte recruitment to skin inflammatory sites in vivo. Clinical and Experimental Rheumatology 11 Suppl 8, S77-80Google ScholarPubMed
55Issekutz, T.B. (1993) Dual inhibition of VLA-4 and LFA-1 maximally inhibits cutaneous delayed-type hypersensitivity-induced inflammation. American Journal of Pathology 143, 1286-1293Google ScholarPubMed
56Scheynius, A., Camp, R.L. and Pure, E. (1993) Reduced contact sensitivity reactions in mice treated with monoclonal antibodies to leukocyte function-associated molecule-1 and intercellular adhesion molecule-1. Journal of Immunology 150, 655-663CrossRefGoogle ScholarPubMed
57Sligh, J.E. Jr. et al. (1993) Inflammatory and immune responses are impaired in mice deficient in intercellular adhesion molecule 1. Proceedings of the National Academy of Sciences of the United States of America 90, 8529-8533CrossRefGoogle ScholarPubMed
58Xu, H. et al. (1994) Leukocytosis and resistance to septic shock in intercellular adhesion molecule 1-deficient mice. Journal of Experimental Medicine 180, 95-109CrossRefGoogle ScholarPubMed
59Tedder, T.F., Steeber, D.A. and Pizcueta, P. (1995) L-selectin-deficient mice have impaired leukocyte recruitment into inflammatory sites. Journal of Experimental Medicine 181, 2259-2264CrossRefGoogle ScholarPubMed
60Catalina, M.D. et al. (1996) The route of antigen entry determines the requirement for L-selectin during immune responses. Journal of Experimental Medicine 184, 2341-2351CrossRefGoogle ScholarPubMed
61Catalina, M.D., Estess, P. and Siegelman, M.H. (1999) Selective requirements for leukocyte adhesion molecules in models of acute and chronic cutaneous inflammation: participation of E- and P- but not L-selectin. Blood 93, 580-589CrossRefGoogle Scholar
62Harari, O.A. et al. (1999) Endothelial cell E- and P-selectin up-regulation in murine contact sensitivity is prolonged by distinct mechanisms occurring in sequence. Journal of Immunology 163, 6860-6866CrossRefGoogle ScholarPubMed
63Labow, M.A. et al. (1994) Characterization of E-selectin-deficient mice: demonstration of overlapping function of the endothelial selectins. Immunity 1, 709-720CrossRefGoogle ScholarPubMed
64Subramaniam, M. et al. (1995) Reduced recruitment of inflammatory cells in a contact hypersensitivity response in P-selectin-deficient mice. Journal of Experimental Medicine 181, 2277-2282CrossRefGoogle Scholar
65Staite, N.D. et al. (1996) Inhibition of delayed-type contact hypersensitivity in mice deficient in both E-selectin and P-selectin. Blood 88, 2973-2979CrossRefGoogle ScholarPubMed
66Hirata, T. et al. (2000) P-selectin glycoprotein ligand 1 (PSGL-1) is a physiological ligand for E-selectin in mediating T helper 1 lymphocyte migration. Journal of Experimental Medicine 192, 1669-1676CrossRefGoogle ScholarPubMed
67Fuhlbrigge, R.C. et al. (1997) Cutaneous lymphocyte antigen is a specialized form of PSGL-1 expressed on skin-homing T-cells. Nature 389, 978-981CrossRefGoogle ScholarPubMed
68Weninger, W. et al. (2000) Specialized contributions by alpha(1,3)-fucosyltransferase-IV and FucT-VII during leukocyte rolling in dermal microvessels. Immunity 12, 665-676CrossRefGoogle ScholarPubMed
69Eppihimer, M.J. et al. (1996) Heterogeneity of expression of E- and P-selectins in vivo. Circulation Research 79, 560-569CrossRefGoogle ScholarPubMed
70Bevilacqua, M.P. et al. (1987) Identification of an inducible endothelial-leukocyte adhesion molecule. Proceedings of the National Academy of Sciences of the United States of America 84, 9238-9242CrossRefGoogle ScholarPubMed
71Austrup, F. et al. (1997) P- and E-selectin mediate recruitment of T-helper-1 but not T-helper-2 cells into inflammed tissues. Nature 385, 81-83CrossRefGoogle Scholar
72Fuhlbrigge, R.C. et al. (2006) CD43 is a ligand for E-selectin on CLA+ human T cells. Blood 107, 1421-1426CrossRefGoogle ScholarPubMed
73Hidalgo, A. et al. (2007) Complete identification of E-selectin ligands on neutrophils reveals distinct functions of PSGL-1, ESL-1, and CD44. Immunity 26, 477-489CrossRefGoogle ScholarPubMed
74Alcaide, P. et al. (2007) The 130-kDa glycoform of CD43 functions as an E-selectin ligand for activated Th1 cells in vitro and in delayed-type hypersensitivity reactions in vivo. Journal of Investigative Dermatology 127, 1964-1972CrossRefGoogle ScholarPubMed
75Matsumoto, M. et al. (2007) CD43 collaborates with P-selectin glycoprotein ligand-1 to mediate E-selectin-dependent T cell migration into inflamed skin. Journal of Immunology 178, 2499-2506CrossRefGoogle ScholarPubMed
76Smithson, G. et al. (2001) Fuc-TVII is required for T helper 1 and T cytotoxic 1 lymphocyte selectin ligand expression and recruitment in inflammation, and together with Fuc-TIV regulates naive T cell trafficking to lymph nodes. Journal of Experimental Medicine 194, 601-614CrossRefGoogle Scholar
77Erdmann, I. et al. (2002) Fucosyltransferase VII-deficient mice with defective E-, P-, and L-selectin ligands show impaired CD4+ and CD8+ T cell migration into the skin, but normal extravasation into visceral organs. Journal of Immunology 168, 2139-2146CrossRefGoogle Scholar
78DeGrendele, H.C. et al. (1996) CD44 and its ligand hyaluronate mediate rolling under physiologic flow: a novel lymphocyte-endothelial cell primary adhesion pathway. Journal of Experimental Medicine 183, 1119-1130CrossRefGoogle Scholar
79Bonder, C.S. et al. (2006) Use of CD44 by CD4+ Th1 and Th2 lymphocytes to roll and adhere. Blood 107, 4798-4806CrossRefGoogle ScholarPubMed
80Camp, R.L. et al. (1993) CD44 is necessary for optimal contact allergic responses but is not required for normal leukocyte extravasation. Journal of Experimental Medicine 178, 497-507CrossRefGoogle Scholar
81Gonda, A. et al. (2005) CD44, but not L-selectin, is critically involved in leucocyte migration into the skin in a murine model of allergic dermatitis. Experimental Dermatology 14, 700-708CrossRefGoogle ScholarPubMed
82Schmits, R. et al. (1997) CD44 regulates hematopoietic progenitor distribution, granuloma formation, and tumorigenicity. Blood 90, 2217-2233CrossRefGoogle ScholarPubMed
83Werr, J. et al. (1998) Beta1 integrins are critically involved in neutrophil locomotion in extravascular tissue In vivo. Journal of Experimental Medicine 187, 2091-2096CrossRefGoogle ScholarPubMed
84Werr, J. et al. (2000) Integrin alpha(2)beta(1) (VLA-2) is a principal receptor used by neutrophils for locomotion in extravascular tissue. Blood 95, 1804-1809CrossRefGoogle Scholar
85de Fougerolles, A.R. et al. (2000) Regulation of inflammation by collagen-binding integrins alpha1beta1 and alpha2beta1 in models of hypersensitivity and arthritis. Journal of Clinical Investigation 105, 721-729CrossRefGoogle Scholar
86Cepek, K.L. et al. (1994) Adhesion between epithelial cells and T lymphocytes mediated by E-cadherin and the alpha E beta 7 integrin. Nature 372, 190-193CrossRefGoogle Scholar
87Huehn, J. et al. (2004) Developmental stage, phenotype, and migration distinguish naive- and effector/memory-like CD4+ regulatory T cells. Journal of Experimental Medicine 199, 303-313CrossRefGoogle ScholarPubMed
88Lehmann, J. et al. (2002) Expression of the integrin alpha Ebeta 7 identifies unique subsets of CD25+ as well as CD25- regulatory T cells. Proceedings of the National Academy of Sciences of the United States of America 99, 13031-13036CrossRefGoogle ScholarPubMed
89Suffia, I. et al. (2005) A role for CD103 in the retention of CD4 + CD25+ Treg and control of Leishmania major infection. Journal of Immunology 174, 5444-5455CrossRefGoogle ScholarPubMed
90Schon, M.P. et al. (2000) Cutaneous inflammatory disorder in integrin alphaE (CD103)-deficient mice. Journal of Immunology 165, 6583-6589CrossRefGoogle ScholarPubMed
91Lammermann, T. et al. (2008) Rapid leukocyte migration by integrin-independent flowing and squeezing. Nature 453, 51-55CrossRefGoogle ScholarPubMed
92Middleton, J. et al. (2002) Leukocyte extravasation: chemokine transport and presentation by the endothelium. Blood 100, 3853-3860CrossRefGoogle ScholarPubMed
93Zlotnik, A. and Yoshie, O. (2000) Chemokines: a new classification system and their role in immunity. Immunity 12, 121-127CrossRefGoogle ScholarPubMed
94Forster, R. et al. (1999) CCR7 coordinates the primary immune response by establishing functional microenvironments in secondary lymphoid organs. Cell 99, 23-33CrossRefGoogle ScholarPubMed
95Gunn, M.D. et al. (1999) Mice lacking expression of secondary lymphoid organ chemokine have defects in lymphocyte homing and dendritic cell localization. Journal of Experimental Medicine 189, 451-460CrossRefGoogle ScholarPubMed
96Kabashima, K. et al. (2007) CXCL12-CXCR4 engagement is required for migration of cutaneous dendritic cells. American Journal of Pathology 171, 1249-1257CrossRefGoogle ScholarPubMed
97Baekkevold, E.S. et al. (2001) The CCR7 ligand elc (CCL19) is transcytosed in high endothelial venules and mediates T cell recruitment. Journal of Experimental Medicine 193, 1105-1112CrossRefGoogle ScholarPubMed
98Campbell, J.J. et al. (1999) The chemokine receptor CCR4 in vascular recognition by cutaneous but not intestinal memory T cells. Nature 400, 776-780CrossRefGoogle Scholar
99Morales, J. et al. (1999) CTACK, a skin-associated chemokine that preferentially attracts skin-homing memory T cells. Proceedings of the National Academy of Sciences of the United States of America 96, 14470-14475CrossRefGoogle ScholarPubMed
100Schaerli, P. et al. (2004) A skin-selective homing mechanism for human immune surveillance T cells. Journal of Experimental Medicine 199, 1265-1275CrossRefGoogle ScholarPubMed
101Goebeler, M. et al. (2001) Differential and sequential expression of multiple chemokines during elicitation of allergic contact hypersensitivity. American Journal of Pathology 158, 431-440CrossRefGoogle ScholarPubMed
102Reiss, Y. et al. (2001) CC chemokine receptor (CCR)4 and the CCR10 ligand cutaneous T cell-attracting chemokine (CTACK) in lymphocyte trafficking to inflamed skin. Journal of Experimental Medicine 194, 1541-1547CrossRefGoogle ScholarPubMed
103Homey, B. et al. (2002) CCL27-CCR10 interactions regulate T cell-mediated skin inflammation. Nature Medicine 8, 157-165CrossRefGoogle ScholarPubMed
104Campbell, J.J., O'Connell, D.J. and Wurbel, M.A. (2007) Cutting Edge: Chemokine receptor CCR4 is necessary for antigen-driven cutaneous accumulation of CD4 T cells under physiological conditions. Journal of Immunology 178, 3358-3362CrossRefGoogle ScholarPubMed
105Lu, B. et al. (1998) Abnormalities in monocyte recruitment and cytokine expression in monocyte chemoattractant protein 1-deficient mice. Journal of Experimental Medicine 187, 601-608CrossRefGoogle ScholarPubMed
106Mayrovitz, H.N. (1992) Leukocyte rolling: a prominent feature of venules in intact skin of anaesthetized hairless mice. American Journal of Physiology 262, H157-H161Google ScholarPubMed
107Janssen, G.H. et al. (1994) Spontaneous leukocyte rolling in venules in untraumatized skin of conscious and anesthetized animals. American Journal of Physiology 267, H1199-1204Google ScholarPubMed
108Auffray, C. et al. (2007) Monitoring of blood vessels and tissues by a population of monocytes with patrolling behavior. Science 317, 666-670CrossRefGoogle ScholarPubMed
109Audoy-Remus, J. et al. (2008) Rod-shaped monocytes patrol the brain vasculature and give rise to perivascular macrophages under the influence of proinflammatory cytokines and angiopoietin-2. Journal of Neuroscience 28, 10187-10199CrossRefGoogle ScholarPubMed
110Hickey, M.J. et al. (1999) Varying roles of E-selectin and P-selectin in different microvascular beds in response to antigen. Journal of Immunology 162, 1137-1143CrossRefGoogle ScholarPubMed
111Hickey, M.J. et al. (2002) Leukocyte-endothelial cell interactions are enhanced in dermal postcapillary venules of MRL/faslpr (lupus-prone) mice: roles of P- and E-selectin. Journal of Immunology 168, 4728-4736CrossRefGoogle Scholar
112Chiang, E.Y. et al. (2007) Imaging receptor microdomains on leukocyte subsets in live mice. Nature Methods 4, 219-222CrossRefGoogle ScholarPubMed
113Miller, M.J. et al. (2002) Two-photon imaging of lymphocyte motility and antigen response in intact lymph node. Science 296, 1869-1873CrossRefGoogle ScholarPubMed
114Mempel, T.R., Henrickson, S.E. and von Andrian, U.H. (2004) T-cell priming by dendritic cells in lymph nodes occurs in three distinct phases. Nature 427, 154-159CrossRefGoogle ScholarPubMed
115Kawakami, N. et al. (2005) Live imaging of effector cell trafficking and autoantigen recognition within the unfolding autoimmune encephalomyelitis lesion. Journal of Experimental Medicine 201, 1805-1814CrossRefGoogle ScholarPubMed
116Mrass, P. et al. (2006) Random migration precedes stable target cell interactions of tumor-infiltrating T cells. Journal of Experimental Medicine 203, 2749-2761CrossRefGoogle ScholarPubMed
117Mrass, P. et al. (2008) CD44 mediates successful interstitial navigation by killer T cells and enables efficient antitumor immunity. Immunity 29, 971-985CrossRefGoogle ScholarPubMed
118Matheu, M.P. et al. (2008) Imaging of effector memory T cells during a delayed-type hypersensitivity reaction and suppression by Kv1.3 channel block. Immunity 29, 602-614CrossRefGoogle ScholarPubMed
119Ng, L.G. et al. (2008) Migratory dermal dendritic cells act as rapid sensors of protozoan parasites. PLoS Pathogens 4, e1000222CrossRefGoogle ScholarPubMed
120Pariser, D.M. et al. (2005) Clinical efficacy of efalizumab in patients with chronic plaque psoriasis: results from three randomized placebo-controlled Phase III trials: part I. Journal of Cutaneous Medicine and Surgery 9, 303-312CrossRefGoogle ScholarPubMed
121Pugashetti, R. and Koo, J. (2009) Efalizumab discontinuation: a practical strategy. Journal of Dermatological Treatment 20, 132-136CrossRefGoogle ScholarPubMed
122Clark, R.A. et al. (2006) The vast majority of CLA+ T cells are resident in normal skin. Journal of Immunology 176, 4431-4439CrossRefGoogle ScholarPubMed
123Kitagaki, H. et al. (1995) Immediate-type hypersensitivity response followed by a late reaction is induced by repeated epicutaneous application of contact sensitizing agents in mice. Journal of Investigative Dermatology 105, 749-755CrossRefGoogle ScholarPubMed
124Kitagaki, H. et al. (1997) Repeated elicitation of contact hypersensitivity induces a shift in cutaneous cytokine milieu from a T helper cell type 1 to a T helper cell type 2 profile. Journal of Immunology 159, 2484-2491CrossRefGoogle Scholar
125Man, M.Q. et al. (2008) Characterization of a hapten-induced, murine model with multiple features of atopic dermatitis: structural, immunologic, and biochemical changes following single versus multiple oxazolone challenges. Journal of Investigative Dermatology 128, 79-86CrossRefGoogle ScholarPubMed
126Fujita, T. et al. (2007) Phase-dependent roles of E-selectin during chronic contact hypersensitivity responses. American Journal of Pathology 170, 1649-1658CrossRefGoogle ScholarPubMed
127Shimada, Y. et al. (2003) L-selectin or ICAM-1 deficiency reduces an immediate-type hypersensitivity response by preventing mast cell recruitment in repeated elicitation of contact hypersensitivity. Journal of Immunology 170, 4325-4334CrossRefGoogle ScholarPubMed
128Nakano, H. et al. (1998) A novel mutant gene involved in T-lymphocyte-specific homing into peripheral lymphoid organs on mouse chromosome 4. Blood 91, 2886-2895CrossRefGoogle ScholarPubMed

Further reading, resources and contacts

The Medline Plus website is a service of the US National Library of Medicine and the National Institutes of Health, and provides information on skin conditions:

Ley, K. et al. (2007) Getting to the site of inflammation: the leukocyte adhesion cascade updated. Nature Reviews Immunology 7, 678-689CrossRefGoogle Scholar
Cahalan, M.D. and Parker, I. (2008) Choreography of Cell Motility and Interaction Dynamics Imaged by Two-Photon Microscopy in Lymphoid Organs. Annual Reviews in Immunology 26, 585-626CrossRefGoogle ScholarPubMed
Pittet, M.J. and Mempel, T.R. (2008) Regulation of T-cell migration and effector functions: insights from in vivo imaging studies. Immunology Reviews 221, 107-129CrossRefGoogle ScholarPubMed
Saint-Mezard, P. et al. (2004) Allergic contact dermatitis. European Journal of Dermatology 14, 284-295Google ScholarPubMed
Saint-Mezard, P. et al. (2004) The role of CD4+ and CD8+ T cells in contact hypersensitivity and allergic contact dermatitis. European Journal of Dermatology 14, 131-138Google ScholarPubMed
Sticherling, M. (2005) Mechanisms of psoriasis. Drug Discovery Today: Disease Mechanisms 2, 275-281CrossRefGoogle Scholar
Boehncke, W.H. (2005) Lymphocyte homing to the skin. Immunology, Immunopathology and Therapeutic Perspectives. CRC Press, Boca RatonGoogle Scholar
Ley, K. et al. (2007) Getting to the site of inflammation: the leukocyte adhesion cascade updated. Nature Reviews Immunology 7, 678-689CrossRefGoogle Scholar
Cahalan, M.D. and Parker, I. (2008) Choreography of Cell Motility and Interaction Dynamics Imaged by Two-Photon Microscopy in Lymphoid Organs. Annual Reviews in Immunology 26, 585-626CrossRefGoogle ScholarPubMed
Pittet, M.J. and Mempel, T.R. (2008) Regulation of T-cell migration and effector functions: insights from in vivo imaging studies. Immunology Reviews 221, 107-129CrossRefGoogle ScholarPubMed
Saint-Mezard, P. et al. (2004) Allergic contact dermatitis. European Journal of Dermatology 14, 284-295Google ScholarPubMed
Saint-Mezard, P. et al. (2004) The role of CD4+ and CD8+ T cells in contact hypersensitivity and allergic contact dermatitis. European Journal of Dermatology 14, 131-138Google ScholarPubMed
Sticherling, M. (2005) Mechanisms of psoriasis. Drug Discovery Today: Disease Mechanisms 2, 275-281CrossRefGoogle Scholar
Boehncke, W.H. (2005) Lymphocyte homing to the skin. Immunology, Immunopathology and Therapeutic Perspectives. CRC Press, Boca RatonGoogle Scholar