MHC-I expression renders catecholaminergic neurons susceptible to T-cell-mediated degeneration

Carolina Cebrián, Fabio A Zucca, Pierluigi Mauri, Julius A Steinbeck, Lorenz Studer, Clemens R Scherzer, Ellen Kanter, Sadna Budhu, Jonathan Mandelbaum, Jean P Vonsattel, Luigi Zecca, John D Loike, David Sulzer, Carolina Cebrián, Fabio A Zucca, Pierluigi Mauri, Julius A Steinbeck, Lorenz Studer, Clemens R Scherzer, Ellen Kanter, Sadna Budhu, Jonathan Mandelbaum, Jean P Vonsattel, Luigi Zecca, John D Loike, David Sulzer

Abstract

Subsets of rodent neurons are reported to express major histocompatibility complex class I (MHC-I), but such expression has not been reported in normal adult human neurons. Here we provide evidence from immunolabel, RNA expression and mass spectrometry analysis of postmortem samples that human catecholaminergic substantia nigra and locus coeruleus neurons express MHC-I, and that this molecule is inducible in human stem cell-derived dopamine (DA) neurons. Catecholamine murine cultured neurons are more responsive to induction of MHC-I by gamma-interferon than other neuronal populations. Neuronal MHC-I is also induced by factors released from microglia activated by neuromelanin or alpha-synuclein, or high cytosolic DA and/or oxidative stress. DA neurons internalize foreign ovalbumin and display antigen derived from this protein by MHC-I, which triggers DA neuronal death in the presence of appropriate cytotoxic T cells. Thus, neuronal MHC-I can trigger antigenic response, and catecholamine neurons may be particularly susceptible to T-cell-mediated cytotoxic attack.

Conflict of interest statement

Competing financial interests

The authors declare no competing financial interests.

Figures

Figure 1. Human SN and LC express…
Figure 1. Human SN and LC express MHC-I
A) Fluorescent images showing double immunolabel for the neuronal nuclei marker, Fox-3 (green) and human MHC-I (HLA A, B and C, red) in human postmortem hippocampal/entorhinal cortex sections and striatal sections from control individuals. White circles demonstrate that there is no overlap between neurons and HLA+ structures. Scale = 60 μm. B, C) Brightfield and immunofluorescence images of SN stained for B) HLA A, B, and C (red) and TH (green) and C) β2m (red) and TH (green). NM was identified under brightfield illumination. Encircled neurons demonstrate that TH+ neurons display HLA immunolabel, overlapping in particular with NM. Scale = 100 μm. D) Confocal immunofluorescent label of TH (green) and HLA (A, B and C, red) in SN and LC control and PD samples. The first row shows a representative example of a SN DA neuron that does not express HLA. When HLA was observed in SN and LC neurons (examples in the second through fifth rows), it was often associated with NM. Scale = 50 μm. E and F) Proportion of TH+ neurons with HLA (A, B and C) immunolabel in the SN (E) and the LC (F). Cell counts were performed in three sections per brain of 8 control individuals and 8 PD patients (SN) and 8 control individuals and 9 PD patients (LC). Data are presented as the mean ± SEM. p > 0.05 (ns) in E (Mann-Whitney U test) and * p < 0.05 in F (Mann-Whitney U test). G) Percentage of TH+ and TH− neurons in the SN and LC (control and PD) labeled for HLA. The arrows in all panels point to blood vessels. Each experiment was repeated at least in triplicate. BF = brightfield; CTRL = control; ns = non significant.
Figure 2. Human SN and LC express…
Figure 2. Human SN and LC express MHC-I with local CTLs
A) V-VIP immunostain (purple) indicates HLA (A, B and C) in control and PD SN and LC samples. NM appears as brown precipitate. Arrows indicate labeled cells in which the chromogen fills and outlines cell bodies and occasional dendrites. The arrowhead indicates a NM+ neuron devoid of cytosolic immunocytochemical label. Scale = 50 μm. B) Immunoelectron microscopy images demonstrating antigenicity to HLA (A, B and C) within NM granules of control SN and LC (white arrows). Blood vessel endothelium was used as a positive control for the HLA antibody and showed immunolabel (black arrows); the erythrocyte showed very little staining (white outlined arrowhead). Lipid bodies (asterisk) within NM organelles did not exhibit HLA immunolabel. The black outlined arrowhead depicts a lysosome with HLA label in an LC neuron. Scale = 250 nm. C) MS/MS spectrum (parent ion m/z 788.2=[M+3H+]3+) corresponding to the peptide NTQTDRESLRNLRCYYNQS observed in the analysis of NM isolated from SN tissue. See Table 1 for details. D) β2m, HLA-A, and HLA-C genes are robustly expressed in laser-captured NM SN neurons of control individuals. Data are presented as the mean ± SEM. Samples from 17 control subjects were analyzed. E) Double immunofluorescence in the SN of human of a postmortem control sample. NM was observed under brightfield microscopy. An arrow indicates a CD8+ T-cell in contact with a NM+ neuron. Scale = 10 μm. F) Immunofluorescence in hES derived DA neurons, showing HLA (A, B and C) immunolabel in the cell body (upper panel) and dendrites (lower panel) of TH+ neurons exposed to human IFN-γ, but not in neurons treated with the vehicle. Asterisks indicate cell bodies and arrows point at dendrites. Scale = 10 μm. Each experiment was repeated at least in triplicate, and within each experiment, each condition was also performed at least three times. BF = brightfield; CTRL = control.
Figure 3. Induced MHC-I by murine catecholamine…
Figure 3. Induced MHC-I by murine catecholamine neurons
A) MHC-I immunolabel in postnatally-derived cultured SN DA neurons from wild type and β2m KO mice imaged by confocal microscopy. The upper row shows untreated DA (TH: green) neurons. The bottom row shows MHC-I (red) expressing DA neurons exposed to IFN-γ. The arrow indicates a MHC-I expressing astrocyte (bottom row). Scale = 30 μm. B) Dose response of MHC-I induction by IFN-γ in neurons obtained from various brain regions. Data are presented as the mean ± SEM (ns = nonsignificant; ** p < 0.01; *** p < 0.001, Two-way ANOVA test). C) IFN-γ released by microglia stimulated with LPS, NM or α-syn. Data are presented as the mean ± SEM (** p < 0.01; *** p < 0.001, One-way ANOVA with Tukey post-hoc test). D) Expression of MHC-I after exposing primary VM neurons to medium from microglia pre-stimulated with LPS, NM or α-syn (wild type, nitrated or A53T mutant). Data are presented as the mean ± SEM (* p < 0.05; ** p < 0.01; *** p < 0.001, One-way ANOVA with Tukey post-hoc test). E) Percentage of VM neurons that expressed MHC-I after exposure to microglial conditioned as in C), with or without a neutralizing antibody for IFN-γ. Data are presented as the mean ± SEM (ns = nonsignificant; * p < 0.05; ** p < 0.01, two-tailed Student’s T-test). Each experiment was repeated at least in triplicate, and within each experiment, each condition was also performed at least three times. For each condition, neurons on n = 24 fields at 20X were quantified. ab = antibody; CTRL = control; ns = nonsignificant; WT = wild type.
Figure 4. MHC-I induction in VM DA…
Figure 4. MHC-I induction in VM DA neurons is dependent on oxidative stress
A) Example of TH/MHC-I double immunolabel of primary cultures of VM neurons after treatment with L-DOPA, which induced the presence of both NM (arrow) and MHC-I. Scale = 30 μm. B) Fraction of TH+ and TH− neurons that displayed NM following L-DOPA. Data are presented as mean ± SEM (ns = nonsignificant, two-tailed Student’s T-test). C) The fraction of TH+ and TH− neurons that displayed plasma membrane MHC-I following L-DOPA. Data are presented as mean ± SEM (* p < 0.05, two-tailed Student’s T-test). Each experiment was repeated at least in triplicate and within each experiment, each condition was also performed at least three times. For each condition, neurons on n = 24 fields at 20X were quantified. BF = brightfield; CTRL = control.
Figure 5. VM DA neurons load and…
Figure 5. VM DA neurons load and display antigen
A) VM DA neurons immunolabeled for TH and SIINFEKL-MHC-I. B) The fraction of TH+ neurons that were labeled for SIINFEKL-MHC-I following vehicle, IFN-γ, OVA, IFN-γ + SIINFEKL, and IFN-γ + OVA. Data are presented as the mean ± SEM (*** p < 0.001, One-way ANOVA with Tukey post-hoc test). Each experiment was repeated at least in triplicate and within each experiment, each condition was also performed at least three times. For each condition, neurons on n = 24 fields at 20X were quantified. Scale = 30 μm. IFN = interferon gamma; SIIN = SIINFEKL; Veh = vehicle.
Figure 6. VM DA neurons that display…
Figure 6. VM DA neurons that display antigen are killed by CD8+ T cells
A) The fraction of VM DA neurons surviving in the presence of SIINFEKL, OT-1 cells pre-pulsed with SIINFEKL, and IFN-γ (100 ng/ml) in cultures obtained from wild type and β2m KO mice. Data are presented as the mean ± SEM (ns = non significant; ** p < 0.01, One-way ANOVA with Tukey post-hoc test). B) Survival of VM DA neurons incubated with IFN-γ, SIINFEKL peptide, and OT-1 cells with the pan-caspase inhibitor Z-VAD-FMK, the Fas/Fas ligand antagonist, Kp7-6, the perforin/granzyme antagonist, concanamycin A, or the combination of Kp7-6 and concanamycin A. Data are presented as the mean ± SEM (ns = non significant; * p < 0.05, *** p < 0.001, One-way ANOVA with Tukey post-hoc test). C) VM DA neuron survival with SIINFEKL, OT-1 cells pre-pulsed with SIINFEKL, and microglial medium previously exposed to LPS, α-syn or NM. Data are presented as the mean ± SEM (ns = non significant; * p < 0.05, ** p < 0.01; *** p < 0.001, One-way ANOVA with Tukey post-hoc test). Each experiment was repeated at least in triplicate and within each experiment, each condition was also performed at least three times. For each condition, neurons on n = 24 fields at 20X were quantified. ConA = concanamycin A; CTRL = control; IFN = interferon gamma; ns = non significant; SIIN = SIINFEKL; Veh = vehicle.

References

    1. Chemali M, Radtke K, Desjardins M, English L. Alternative pathways for MHC class I presentation: a new function for autophagy. Cell Mol Life Sci. 2011;68:1533–1541.
    1. Wong GH, Bartlett PF, Clark-Lewis I, Battye F, Schrader JW. Inducible expression of H-2 and Ia antigens on brain cells. Nature. 1984;310:688–91.
    1. Neumann H, Cavalié A, Jenne DE, Wekerle H. Induction of MHC class I genes in neurons. Science. 1995;28, 269(5223):549–52.
    1. Lindå H, Hammarberg H, Piehl F, Khademi M, Olsson T. Expression of MHC class I and beta2-microglobulin in rat spinal motoneurons: regulatory influences by IFN-gamma and axotomy. Exp Neurol. 1999;150:282–295.
    1. Huh SG, Boulanger LM, Du H, Riquelme PA, Brotz TM, Shatz CJ. Functional requirement for class I MHC in CNS development and plasticity. Science. 2000;290:2155–9.
    1. Letellier M, Willson ML, Gautheron V, Mariani J, Lohof AM. Normal adult climbing fiber monoinnervation of cerebellar Purkinje cells in mice lacking MHC class I molecules. Dev Neurobiol. 2008;68(8):997–1006.
    1. Needleman LA, Liu XB, El-Sabeawy F, Jones EG, McAllister AK. MHC class I molecules are present both pre-a and postsynaptically in the visula cortex during postnatal development and in adulthood. Proc Natl Acad Sci U S A. 2010;107:16999–7004.
    1. Liu J, et al. The expression pattern of classical MHC class I molecules in the development of mouse central nervous system. Neurochem Res. 2012;38(2):290–9.
    1. Goddard CA, Butts DA, Shatz CJ. Regulation of CNS synapses by neuronal MHC class I. Proc Natl Acad Sci U S A. 2007;104(16):6828–33.
    1. Corriveau RA, Huh GS, Shatz CJ. Regulation of class I MHC gene expression in the developing and mature CNS by neural activity. Neuron. 1998;21(3):505–20.
    1. Glynn MW, et al. MHCI negatively regulates synapse density during the establishment of cortical connections. Nat Neurosci. 2011;14(4):442–51.
    1. Oliveira AL, et al. A role for MHC class I molecules in synaptic plasticity and regeneration of neurons after axotomy. Proc Natl Acad Sci U S A. 2004;101:17843–8.
    1. Nelson PA, Sage JR, Wood SC, Davenport CM, Anagnostaras SG, Boulanger LM. MHC class I immune proteins are critical for hippocampus-dependent memory and gate NMDAR-dependent hippocampal long-term depression. Learn Mem. 2013;20(9):505–17.
    1. Medana IM, Gallimore A, Oxenius A, Martinic MM, Wekerle H, Neumann H. MHC class I-restricted killing of neurons by virus-specific CD8+ T lymphocytes is effected through the Fas/FasL, but not the perforin pathway. Eur J Immunol. 2000;30:3623–33.
    1. Meuth SG, et al. Cytotoxic CD8+ T cell-neuron interactions: perforin-dependent electrical silencing precedes but is not causally linked to neuronal cell death. J Neurosci. 2009;29:15397–409.
    1. Chevalier GE, et al. Neurons are MHC class I-dependent targets for CD8 T cells upon neurotropic viral infection. PLoS Pathog. 2011;7(11):e1002393.
    1. Tooyama I, Kimura H, Akiyama H, McGeer PL. Reactive microglia express class I and class II major histocompatibility complex antigens in Alzheimer’s disease. Brain Res. 1990;523:273–280.
    1. McGeer PL, Itagaki S, Boyes BE, McGeer EG. Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology. 1988;38:1285–91.
    1. Imamura K, et al. Cytokine production of activated microglia and decrease in neurotrophic factors of neurons in the hippocampus of Lewy body disease brains. Acta Neuropathol. 2005;109:141–150.
    1. Bien CG, et al. Destruction of neurons by cytotoxic T cells: a new pathogenic mechanism in Rasmussen’s encephalitis. Ann Neurol. 2002;51:311–8.
    1. Prabowo AS, Iyer AM, Anink JJ, Spliet WG, van Rijen PC, Aronica E. Differential expression of major histocompatibility complex class I in developmental glioneuronal lesions. J Neuroinflammation. 2013;24(10):12.
    1. Zhang A, et al. The expression patterns of MHC class I molecules in the developmental human visual system. Neurochem Res. 2013;38(2):273–81.
    1. Zhang A, et al. The spatiotemporal expression of MHC class I molecules during human hippocampal formation development. Brain Res. 2013;5(1529):26–38.
    1. Sulzer D, et al. Neuromelanin biosynthesis is driven by excess cytosolic catecholamines not accumulated by synaptic vesicles. Proc Natl Acad Sci U S A. 2000;97:11869–11874.
    1. Zecca L, et al. New melanic pigments in the human brain that accumulate in aging and block environmental toxic metals. Proc Natl Acad Sci U S A. 2008;105:17567–72.
    1. Zheng B, et al. Global PD Gene Expression (GPEX) Consortium, PGC-1α, a potential therapeutic target for early intervention in Parkinson’s disease. Sci Transl Med. 2008;52:52ra73.
    1. Brochard V, et al. Infiltration of CD4+ lymphocytes into the brain contributes to neurodegeneration in a mouse model of Parkinson disease. J Clin Invest. 2009;119:182–192.
    1. Kriks S, et al. Dopamine neurons derived from human ES cells efficiently engraft in animal models of Parkinson’s disease. Nature. 2011;480:547–51.
    1. van den Elsen PJ, Holling TM, Kuipers HF, van der Stoep N. Transcriptional regulation of antigen presentation. Curr Opin Immunol. 2004;16:67–75.
    1. Mogi M, Kondo T, Mizuno Y, Nagatsu T. p53 protein, interferon-gamma, and NF-kappaB levels are elevated in the parkinsonian brain. Neurosci Lett. 2007;414:94–97.
    1. Sulzer D, Trudeau LE, Rayport S. In: Parkinson’s Disease: Molecular and Therapeutic Insights from Model Systems. Nass R, editor. Academic Press; 2008. pp. 491–504.
    1. Kawanokuchi J, et al. Production of interferon-gamma by microglia. Mult Scler. 2006;12:558–64.
    1. Foix C, Nicolesco J. Suivi d’un apéndice sur l’anatomie pathologique de la maladie de Parkinson. Paris: Masson et Cie; 1925. Anatomie cérébrale. Les noyaux gris centraux et la región Mésencéphalo-sous-optique; pp. 508–538.
    1. Zhang W, et al. Microglial PHOX and Mac-1 are essential to the enhanced dopaminergic neurodegeneration elicited by A30P and A53T mutant alpha-synuclein. Glia. 2007;55:1178–1188.
    1. Zhang W, et al. Neuromelanin activates microglia and induces degeneration of dopaminergic neurons: implications for progression of Parkinson’s disease. Neurotox Res. 2011;19:63–72.
    1. Polymeropoulos MH, et al. Mutation in the alpha-synuclein gene identified in families with Parkinson’s disease. Science. 1997;276(5321):2045–7.
    1. Singleton AB, et al. alpha-Synuclein locus triplication causes Parkinson’s disease. Science. 2003;302(5646):841.
    1. Teoh CY, Davies KJ. Potential roles of protein oxidation and the immunoproteasome in MHC class I antigen presentation: the ‘PrOxI’ hypothesis. Arch Biochem Biophys. 2004;423:88–96.
    1. Fahn S, Sulzer D. Neurodegeneration and neuroprotection in Parkinson disease. NeuroRx. 2004;1(1):139–54.
    1. Conway KA, Rochet JC, Bieganski RM, Lansbury PT., Jr Kinetic stabilization of the alpha-synuclein protofibril by a dopamine-alpha-synuclein adduct. Science. 2001;294(5545):1346–9.
    1. Norris EH, et al. Reversible inhibition of alpha-synuclein fibrillization by dopaminochrome-mediated conformational alterations. J Biol Chem. 2005;280(22):21212–9.
    1. Martinez-Vicente M, et al. Dopamine-modified alpha-synuclein blocks chaperone-mediated autophagy. J Clin Invest. 2008;118:777–788.
    1. Muñoz P, Huenchuguala S, Paris I, Segura-Aguilar J. Dopamine oxidation and autophagy. Parkinsons Dis. 2012;2012:920953. doi: 10.1155/2012/920953.
    1. Mosharov EV, et al. Interplay between cytosolic dopamine, calcium, and alpha-synuclein causes selective death of substantia nigra neurons. Neuron. 2009;30:218–229.
    1. Pardo B, Mena MA, Casarejos MJ, Paíno CL, De Yébenes JG. Toxic effects of L-DOPA on mesencephalic cell cultures: protection with antioxidants. Brain Res. 1995;682(1–2):133–43.
    1. Falk K, et al. Both human and mouse cells expressing H-2Kb and ovalbumin process the same peptide, SIINFEKL. Cell Immunol. 1993;150(2):447–52.
    1. Budhu S, et al. CD8+ T cell concentration determines their efficiency in killing cognate antigen-expressing syngeneic mammalian cells in vitro and in mouse tissues. J Exp Med. 2010;207:223–235.
    1. Marsh SG, et al. Nomenclature for factors of the HLA system. Tissue Antigens. 2005;65:301–69.
    1. Chan CS, Gertler TS, Surmeier DJ. Calcium homeostasis, selective vulnerability and Parkinson’s disease. Trends Neurosci. 2009;32:249–256.
    1. Kleijmeer MJ, et al. Antigen loading of MHC class I molecules in the endocytic tract. Traffic. 2001;2:124–137.
    1. Fiegl D, et al. Amphisomal route of MHC class I cross-presentation in bacteria-infected dendritic cells. J Immunol. 2013;190(6):2791–806.
    1. Zhang W, et al. Human neuromelanin: an endogenous microglial activator for dopaminergic neuron death. Front Biosci (Elite Ed) 2013;5:1–11.
    1. Béraud D, et al. Microglial activation and antioxidant responses induced by the Parkinson’s disease protein α-synuclein. J Neuroimmune Pharmacol. 2013;8(1):94–117.
    1. Block ML, Zecca L, Hong JS. Microglia-mediated neurotoxicity: uncovering the molecular mechanisms. Nat Rev Neurosci. 2007;8(1):57–69.
    1. Lull ME, Block ML. Microglial activation and chronic neurodegeneration. Neurotherapeutics. 2010;7(4):354–65.
    1. Zhao YN, Wang F, Fan YX, Ping GF, Yang JY, Wu CF. Activated microglia are implicated in cognitive deficits, neuronal death, and successful recovery following intermittent ethanol exposure. Behav Brain Res. 2013;236(1):270–82.
    1. Double KL. Neuronal vulnerability in Parkinson’s disease. Parkinsonism Relat Disord. 2012;18 (Suppl 1):S52–4.
    1. Harris TH, et al. Generalized Lévy walks and the role of chemokines in migration of effector CD8+ T cells. Nature. 2012;486:545–8.
    1. German DC, Eagar T, Sonsalla PK. Parkinson’s Disease: A Role for the Immune System. Curr Mol Pharmacol. 2011 In press.
    1. Farkas E, De Jong GI, de Vos RA, Jansen Steur EN, Luiten PG. Pathological features of cerebral cortical capillaries are doubled in Alzheimer’s disease and Parkinson’s disease. Acta Neuropathol. 2000;100(4):395–402.
    1. Abraham S, Nagaraj AS, Basak S, Manjunath R. Japanese encephalitis virus utilizes the canonical pathway to activate NF-kappaB but it utilizes the type I interferon pathway to induce major histocompatibility complex class I expression in mouse embryonic fibroblasts. J Virol. 2010;84:5485–93.
    1. Mangano EN, et al. Interferon-γ plays a role in paraquat-induced neurodegeneration involving oxidative and proinflammatory pathways. Neurobiol Aging. 2011;33(7):1411–26.
    1. Chakrabarty P, et al. Interferon-γ induces progressive nigrostriatal degeneration and basal ganglia calcification. Nat Neurosci. 2011;14:694–696.
    1. Almeida CM, Galvão M, de L, Ferreira PL, Braga WS. Interferon-induced Parkinsonism in a patient with chronic hepatitis C. Arq Neuropsiquiatr. 2009;67:715–6.
    1. Zecca L, et al. The role of iron and copper molecules in the neuronal vulnerability of locus coeruleus and substantia nigra during aging. Proc Natl Acad Sci U S A. 2004;101(26):9843–8.
    1. Gentleman RC, et al. J. Bioconductor: open software development for computational biology and bioinformatics. Genome Biol. 2004;5:R80.
    1. Simunovic F, et al. Gene expression profiling of substantia nigra dopamine neurons: further insights into Parkinson’s disease pathology. Brain. 2009;132:1795–1809.
    1. Rothenberg BE, Voland JR. beta2 knockout mice develop parenchymal iron overload: A putative role for class I genes of the major histocompatibility complex in iron metabolism. Proc Natl Acad Sci U S A. 1996;93:1529–34.
    1. Tallóczy Z, et al. Methamphetamine inhibits antigen processing, presentation, and phagocytosis. PLoS Pathog. 2008;4:e28.
    1. Hogquist KA, et al. T cell receptor antagonist peptides induce positive selection. Cell. 1994;76:17–27.

Source: PubMed

3
購読する