Original contribution
T cell receptor-stimulated generation of hydrogen peroxide inhibits MEK-ERK activation and lck serine phosphorylation

https://doi.org/10.1016/S0891-5849(03)00318-6Get rights and content

Abstract

Previous studies indicated that antigen receptor (TcR) stimulation of mature T cells induced rapid generation of reactive oxygen species (ROS). The goal of the current study was to examine the role(s) of ROS in TcR signal transduction, with a focus upon the redox-sensitive MAPK family. TcR cross-linking of primary human T blasts and Jurkat human T cells rapidly activated the ERK, JNK, p38 and Akt kinases within minutes, and was temporally associated with TcR-stimulated production of hydrogen peroxide (H2O2). TcR-induced activation of ERK was selectively augmented and sustained in the presence of pharmacologic antioxidants that can quench or inhibit H2O2 production (NAC, MnTBAP and Ebselen, but not DPI), while activation of JNK and Akt were largely unaffected. This was paralleled by concurrent changes in MEK1/2 phosphorylation, suggesting that ROS acted upstream of MEK-ERK activation. Molecular targeting of H2O2 by overexpression of peroxiredoxin II, a thioredoxin dependent peroxidase, also increased and sustained ERK and MEK activation upon TcR cross-linking. Enhancement of ERK phosphorylation by antioxidants correlated with increased and sustained serine phosphorylation of the src-family kinase lck, a known ERK substrate. Thus, the data suggest that TcR-stimulated production of hydrogen peroxide negatively feeds back to dampen antigen-stimulated ERK activation and this redox-dependent regulation may serve to modulate key steps in TcR signaling.

Introduction

Contrary to the view of reactive oxygen species (ROS) as damaging, toxic agents, it has been established that multiple cell surface receptors promote the intracellular generation of ROS that serve as key second messengers. Thus, stimulation of receptors by agonists such as angiotensin II, PDGF, or TGFβ induces production of ROS, which are required for control of protein kinase activation, gene expression, or cell proliferation. In lymphocytes, stimulation with lectins (conA or PHA) 1, 2, 3, mitogens (PMA or superantigens) 1, 4, 5, anti-CD28 [6], or the anti-CD3 4, 6, 7, have been proposed to induce oxidant generation and were suggested to regulate cell activation, proliferation and/or survival. While a regulatory role for ROS in these systems has been proposed, the mechanism(s) by which they regulate signal transduction and subsequent biological processes has not been well described.

Exposure to exogenous oxidants, sometimes at nontoxic concentrations, has been suggested to activate a number of signaling pathways, such as src-family kinases 8, 9, NF-κB dependent transcription 10, 11, Ras 12, 13, Akt [14], or the JAK-STAT pathway 15, 16. An important redox-sensitive signaling pathway is the MAPK family of protein kinases, including ERK, JNK, and p38. Addition of oxidants to cells has been shown to activate each of these kinases and they have been characterized as stress responsive elements (reviewed in [17]).

The MAPK family of protein kinases represent a group (ERK1/2, JNK, p38, ERK5) of serine/threonine protein kinases that have the unusual characteristic of requiring phosphorylation for activation on both threonine and tyrosine in a T-x-Y motif that is present on an activation loop. This phosphorylation is catalyzed by a group of relatively specific dual specificity kinases (MAPK kinases) that phosphorylate both Thr and Tyr of the respective MAPK and lead to activation of the kinase. The relative specificity of these MAPK kinases has led to a model in which modules exist for the selective activation of MAPKs. Despite this selective interaction, there is significant redundancy in the activation of the MAPK kinases (MEK, MKK) by the upstream MAPK kinase kinase (MEKK). Multiple MEKK enzymes have been identified, including Raf, MEKK1-4, Tpl-2, and ASK, and in many cases they are capable of activating different MEK or MKK. Further diversity is encountered when one considers the number of signals, both intracellular and extracellular, that can lead to activation of different MEKK. For example, the Raf family of protein kinases can be activated by inputs from proteins such as Ras, PKC family members, Src, PAK, or Akt [18]. Thus, in the case of ERK activation in T cells, signals proceed through Raf and MEK, but the upstream activation signals may be derived from multiple upstream pathways [19].

As exogenous oxidants have targeted activation of MAPK pathways, it was not completely unexpected that receptor-stimulated, endogenously generated ROS would also regulate MAPK activation. In fact, PDGF-stimulated ROS production is required for activation of ERK in VSMC or fibroblasts [20]. Also in VSMC, stimulation with angiotensin II-induced production of ROS that were necessary for JNK or p38 MAPK activation 21, 22. Similarly, serotonin (5-HT) exposure induced ROS generation in mesangial cells, fibroblasts and smooth muscle cells that, in all cases, controlled ERK activation 23, 24. Thus, many studies on receptor-stimulated production of ROS have also focused upon their ability to activate members of the MAPK family of enzymes. This family of enzymes is not the only signaling target of oxidants in the cell, however. Receptor-stimulated production of ROS has also been shown to target activation of key signaling molecules such as Ras, Src, STAT1, and Akt (reviewed in 25, 26).

Our previous data has indicated that cross-linking the T cell receptor (TcR) also leads to the rapid production of both superoxide anion and hydrogen peroxide [7]. The goal of the current study was to investigate whether TcR-stimulated production of ROS would regulate certain signaling pathways. The data extend previous results, indicating that TcR cross-linking induces production of hydrogen peroxide within 2–4 min in a MEK1-dependent manner in both Jurkat T cells and primary human T blasts. Furthermore, detailed kinetic analysis of MAPK activation indicated that the rapid generation of hydrogen peroxide served to selectively inhibit TcR-stimulated phosphorylation of MEK and ERK in both cell models. Using pharmacologic antioxidants and overexpression of the thioredoxin-dependent peroxidase, peroxiredoxin II, TcR-induced activation of MEK and ERK was enhanced and sustained with little or no effect on the activation of JNK or Akt. Further analysis of potential ERK substrates suggested that enhanced ERK activation was associated with increased serine phosphorylation of lck. Thus, the data suggest that TcR cross-linking elicits generation of hydrogen peroxide in a MEK-dependent pathway that then acts in a negative feedback loop to limit MEK-ERK activation and subsequent phosphorylation of important regulatory proteins in T cell function.

Section snippets

Reagents

Dihydroethidium (DHE) and dichlorodihydrofluorescein diacetate (DCFDA) were obtained from Molecular Probes (Eugene, OR, USA). Manganese (III) tetrakis (4-benzoic acid) porphyrin (MnTBAP) was from Alexis (San Diego, CA, USA), while ebselen, PD98059, U0126, and diphenylene iodonium (DPI) were from Calbiochem (San Diego, CA, USA). N-acetyl-l-cysteine, ionomycin, PMA, rabbit anti-hamster IgG, goat anti-mouse IgG, and all other chemicals were obtained from Sigma (St. Louis, MO, USA). All cell

TcR cross-linking induces MAPK phosphorylation

Our previous data have shown that TcR cross-linking induced rapid generation of both hydrogen peroxide and superoxide anion in mature T cells [7]. To study whether TcR signal transduction pathways are sensitive to reactive oxygen species (ROS), the kinetics of MAPK and Akt activation in Jurkat T cells was determined. The phosphorylation status of these kinases was determined by Western blot analyses using corresponding phospho-specific antibodies, which are proposed to recognize only the

Discussion

Our previous studies have established that TcR cross-linking in mature T cells also induces rapid generation of intracellular ROS [7]. However the molecular signaling mechanisms that are targeted by TcR-stimulated production of oxidants in T cells are poorly understood. The current study extends our previous observations 4, 7 to show that H2O2 production occurs as early as 2–4 min after TcR cross-linking in both cultured cell lines and primary T cells. Using detailed kinetic analysis of MAPK

Abbreviations

  • DCFDA—dichlorodihydrofluorescein diacetate

  • DHE—dihydroethidium

  • MnTBAP—Manganese (III) tetrakis (4-benzoic acid) porphyrin

  • NAC—N-acetyl-l-cysteine

  • Prx II—peroxiredoxin II

  • ROS—reactive oxygen species

  • TcR—T cell receptor

Acknowledgements

The authors would like to acknowledge Dr. Sue Goo Rhee for expression plasmids and antibodies for Prx II. We thank Cheng-Kui Qu, Greg Carey and Pierre Henkart for critical reading of the manuscript. This project was supported by a Scientist Development Grant (#0030033N) from the American Heart Association and by funds from the American Red Cross.

References (64)

  • J.P. Crow

    Dichlorodihydrofluorescein and dihydrorhodamine 123 are sensitive indicators of peroxynitrite in vitroimplications for intracellular measurement of reactive nitrogen and oxygen species

    Nitric Oxide

    (1997)
  • S.W. Kang et al.

    Mammalian peroxiredoxin isoforms can reduce hydrogen peroxide generated in response to growth factors and tumor necrosis factor-alpha

    J. Biol. Chem

    (1998)
  • M. Morey et al.

    Modulation of the Ras/MAPK signalling pathway by the redox function of selenoproteins in Drosophila melanogaster

    Dev. Biol

    (2001)
  • T.C. Meng et al.

    Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo

    Mol. Cell

    (2002)
  • H.M. Lander et al.

    Activation of the receptor for advanced glycation end products triggers a p21(ras)-dependent mitogen-activated protein kinase pathway regulated by oxidant stress

    J. Biol. Chem

    (1997)
  • S.R. Lee et al.

    Reversible inactivation of the tumor suppressor PTEN by H2O2

    J. Biol. Chem

    (2002)
  • D. Xu et al.

    Oxidants painting the cysteine chapelredox regulation of PTPs

    Dev. Cell

    (2002)
  • S.R. Lee et al.

    Reversible inactivation of protein-tyrosine phosphatase 1B in A431 cells stimulated with epidermal growth factor

    J. Biol. Chem

    (1998)
  • R.K. Vadlamudi et al.

    p62, a phosphotyrosine-independent ligand of the SH2 domain of p56lck, belongs to a new class of ubiquitin-binding proteins

    J. Biol. Chem

    (1996)
  • T. Geetha et al.

    Structure and functional properties of the ubiquitin binding protein p62

    FEBS Lett

    (2002)
  • R.N. Germain

    The T cell receptor for antigensignaling and ligand discrimination

    J. Biol. Chem

    (2001)
  • A. Kosugi et al.

    Involvement of SHP-1 tyrosine phosphatase in TCR-mediated signaling pathways in lipid rafts

    Immunity

    (2001)
  • S.M. Jones et al.

    Differential expression of glutathione transferases by native and cultured human lymphocytes

    Biochem. Pharmacol

    (1988)
  • W. Droge et al.

    Thiols and the immune systemeffect of N-acetylcysteine on T cell system in human subjects

    Methods Enzymol

    (1995)
  • C. Sekkat et al.

    Oxidative phenomena are implicated in human T-cell stimulation

    Immunology

    (1988)
  • G. Pani et al.

    Endogenous oxygen radicals modulate protein tyrosine phosphorylation and JNK-1 activation in lectin-stimulated thymocytes

    Biochem. J

    (2000)
  • M.S. Williams et al.

    The role of reactive oxygen intermediates in TcR-induced death of T cell blasts and hybridomas

    J. Immunol

    (1996)
  • M. Los et al.

    IL-2 gene expression and NF-kappa B activation through CD28 requires reactive oxygen production by 5-lipoxygenase

    EMBO J

    (1995)
  • S. Devadas et al.

    Discrete generation of superoxide and hydrogen peroxide by T cell receptor stimulationselective regulation of mitogen-activated protein kinase activation and fas ligand expression

    J. Exp. Med

    (2002)
  • D.J. Sulciner et al.

    rac1 regulates a cytokine-stimulated, redox-dependent pathway necessary for NF-kappaB activation

    Mol. Cell. Biol

    (1996)
  • R. Schreck et al.

    Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-κB transcription factor and HIV-1

    EMBO J

    (1991)
  • B. Schieffer et al.

    Role of NAD(P)H oxidase in angiotensin II-induced JAK/STAT signaling and cytokine induction

    Circ. Res

    (2000)
  • Cited by (0)

    View full text