Single domain intrabodies against WASP inhibit TCR-induced immune responses in transgenic mice T cells

Intrabody technology provides a novel approach to decipher the molecular mechanisms of protein function in cells. Single domains composed of only the variable regions (VH or VL) of antibodies are the smallest recombinant antibody fragments to be constructed thus far. In this study, we developed transgenic (Tg) mice expressing the VH or VL single domains derived from a monoclonal antibody raised against the N-terminal domain of Wiskott–Aldrich syndrome protein (WASP), which is an adaptor molecule in immune cells. In T cells from anti-WASP VH and VL single domain Tg mice, interleukin-2 production induced by T cell receptor (TCR) stimulation were impaired, and specific interaction between the WASP N-terminal domain and the Fyn SH3 domain was strongly inhibited by masking the binding sites in WASP. These results strongly suggest that the VH/VL single domain intrabodies are sufficient to knockdown the domain function of target proteins in the cytosol.

[1]  Y. Kurosawa,et al.  Intrabodies against the EVH1 domain of Wiskott–Aldrich syndrome protein inhibit T cell receptor signaling in transgenic mice T cells , 2005, The FEBS journal.

[2]  K. D. Hardman,et al.  Single-chain antigen-binding proteins. , 1988, Science.

[3]  A. Abo,et al.  A Role for Wiskott-Aldrich Syndrome Protein in T-cell Receptor-mediated Transcriptional Activation Independent of Actin Polymerization* , 2001, The Journal of Biological Chemistry.

[4]  Lucy J. Holt,et al.  Domain antibodies: proteins for therapy. , 2003, Trends in biotechnology.

[5]  A. Rao,et al.  Transcriptional regulation of the IL-2 gene. , 1995, Current opinion in immunology.

[6]  R. Kane,et al.  Nanoparticle-mediated cytoplasmic delivery of proteins to target cellular machinery. , 2010, ACS nano.

[7]  S. Chen,et al.  Design, intracellular expression, and activity of a human anti-human immunodeficiency virus type 1 gp120 single-chain antibody. , 1993, Proceedings of the National Academy of Sciences of the United States of America.

[8]  Y. Takemoto,et al.  Distinct binding patterns of HS1 to the Src SH2 and SH3 domains reflect possible mechanisms of recruitment and activation of downstream molecules. , 1996, International immunology.

[9]  Mitsuru Sato,et al.  Cytoplasmic expression and specific binding of the VH/VL single domain intrabodies in transfected NIH3T3 cells. , 2009, Experimental and molecular pathology.

[10]  T. Rabbitts,et al.  Intrabodies based on intracellular capture frameworks that bind the RAS protein with high affinity and impair oncogenic transformation , 2003, The EMBO journal.

[11]  H. Grey,et al.  Antigen recognition by H-2-restricted T cells. I. Cell-free antigen processing , 1983, The Journal of experimental medicine.

[12]  T. Rabbitts,et al.  Intracellular antibodies and challenges facing their use as therapeutic agents. , 2003, Trends in molecular medicine.

[13]  Mitsuru Sato,et al.  Identification of Fyn as the binding partner for the WASP N-terminal domain in T cells. , 2011, International immunology.

[14]  C. Terhorst,et al.  T cells of patients with the Wiskott-Aldrich syndrome have a restricted defect in proliferative responses. , 1993, Journal of immunology.

[15]  N. Tsuji,et al.  Overexpression of the Wiskott-Aldrich Syndrome Protein N-Terminal Domain in Transgenic Mice Inhibits T Cell Proliferative Responses Via TCR Signaling Without Affecting Cytoskeletal Rearrangements1 , 2001, The Journal of Immunology.

[16]  S. Kanner,et al.  Wiskott-Aldrich syndrome/X-linked thrombocytopenia: WASP gene mutations, protein expression, and phenotype. , 1997, Blood.

[17]  V. Didenko,et al.  Polyethyleneimine as a transmembrane carrier of fluorescently labeled proteins and antibodies. , 2005, Analytical biochemistry.

[18]  J. Bender,et al.  Phase I Trial , 1983 .

[19]  E. Remold-O’Donnell,et al.  Defects in Wiskott-Aldrich syndrome blood cells. , 1996, Blood.

[20]  P. Hogan,et al.  Transcription factors of the NFAT family: regulation and function. , 1997, Annual review of immunology.

[21]  B. Heng,et al.  Making cell-permeable antibodies (Transbody) through fusion of protein transduction domains (PTD) with single chain variable fragment (scFv) antibodies: potential advantages over antibodies expressed within the intracellular environment (Intrabody). , 2005, Medical hypotheses.

[22]  Philip R. Cohen,et al.  Wiskott-Aldrich syndrome protein-deficient mice reveal a role for WASP in T but not B cell activation. , 1998, Immunity.

[23]  R. Bruccoleri,et al.  Protein engineering of antibody binding sites: recovery of specific activity in an anti-digoxin single-chain Fv analogue produced in Escherichia coli. , 1988, Proceedings of the National Academy of Sciences of the United States of America.

[24]  H. Kohler,et al.  Chemical engineering of cell penetrating antibodies. , 2001, Journal of immunological methods.

[25]  L. Bracco,et al.  Restoration of transcriptional activity of p53 mutants in human tumour cells by intracellular expression of anti-p53 single chain Fv fragments , 1999, Oncogene.

[26]  K. Siminovitch,et al.  Fyn and PTP-PEST–mediated Regulation of Wiskott-Aldrich Syndrome Protein (WASp) Tyrosine Phosphorylation Is Required for Coupling T Cell Antigen Receptor Engagement to WASp Effector Function and T Cell Activation , 2004, The Journal of experimental medicine.

[27]  J. Rothman,et al.  Peptide-binding specificity of the molecular chaperone BiP , 1991, Nature.

[28]  J. Hartwig,et al.  WIP, a protein associated with wiskott-aldrich syndrome protein, induces actin polymerization and redistribution in lymphoid cells. , 1997, Proceedings of the National Academy of Sciences of the United States of America.

[29]  S. Muyldermans,et al.  Naturally occurring antibodies devoid of light chains , 1993, Nature.

[30]  S. Munro,et al.  An hsp70-like protein in the ER: Identity with the 78 kd glucose-regulated protein and immunoglobulin heavy chain binding protein , 1986, Cell.

[31]  T. Rabbitts,et al.  Functional intracellular antibody fragments do not require invariant intra-domain disulfide bonds. , 2008, Journal of molecular biology.

[32]  K. Siminovitch,et al.  Antigen Receptor–Induced Activation and Cytoskeletal Rearrangement Are Impaired in Wiskott-Aldrich Syndrome Protein–Deficient Lymphocytes , 1999, The Journal of experimental medicine.

[33]  A. Cattaneo,et al.  The selection of intracellular antibodies. , 1998, Trends in biotechnology.

[34]  H. Ochs,et al.  The Wiskott-Aldrich syndrome: studies of lymphocytes, granulocytes, and platelets. , 1980, Blood.

[35]  K. Miura,et al.  N‐WASP, a novel actin‐depolymerizing protein, regulates the cortical cytoskeletal rearrangement in a PIP2‐dependent manner downstream of tyrosine kinases. , 1996, The EMBO journal.

[36]  Ravi S Kane,et al.  Regulation of stem cell signaling by nanoparticle-mediated intracellular protein delivery. , 2011, Biomaterials.

[37]  M. Stocks Intrabodies as drug discovery tools and therapeutics. , 2005, Current opinion in chemical biology.

[38]  Martin R. Johnson,et al.  A cancer gene therapy approach utilizing an anti-erbB-2 single-chain antibody-encoding adenovirus (AD21): a phase I trial. , 2000, Clinical cancer research : an official journal of the American Association for Cancer Research.

[39]  K. Siminovitch,et al.  WIP is a chaperone for Wiskott–Aldrich syndrome protein (WASP) , 2007, Proceedings of the National Academy of Sciences.

[40]  C. Barbas,et al.  Functional deletion of the CCR5 receptor by intracellular immunization produces cells that are refractory to CCR5-dependent HIV-1 infection and cell fusion. , 2000, Proceedings of the National Academy of Sciences of the United States of America.