• Keine Ergebnisse gefunden

Traumatic damage to the nervous system including spinal cord injury (SCI) causes extensive inflammation and the invasion of microglia in and around the affected regions. Previous work has shown, that activation and accumulation of microglia around the epicenter of the lesion is thought to be involved in the quite limited axonal recovery of the CNS (Kitayama et al., 2011; Popovich et al., 1999). The molecular mechanisms blocking neurite regrowth and regeneration are not completely understood but they implicate microglia-derived substances acting as axon growth inhibitors.

Results of the present thesis support evidence that activation of microglia attenuate neurite outgrowth of co-cultured neurons without direct cell-cell surface contact.

Thus, we hypothesized that microglia-derived secreted molecules must be responsible for the poor axonal elongation (Fig. 8 A). By performing experimental manipulations I found that the induction of microglial iNOS and the presence of NO were responsible for blocking neurite growth following growth cone collapse (Scheiblich and Bicker, 2015b). Intriguingly, inhibition of iNOS/NO production via

Discussion 29

induction of HO-1/CO signaling not only attenuated the inflammatory response of reactive microglia (see discussion before) but also allowed for neurite outgrowth (Scheiblich and Bicker, 2015a). In support of these observations we found, that the regulation of neurite outgrowth is caused downstream the NO pathway (Fig. 8 B) (Roloff et al., 2015).

Figure 8: Neurite outgrowth of human model neurons (magenta) is regulated by microglia (green). (A) Microglial activation result in the induction of the inducible nitric oxide synthase (iNOS) and subsequently in the release of nitric oxide (NO). The presence of NO blocks neurite outgrowth of the neurons. However, induction of heme oxygenase-1 (HO-1) and the release of carbon monoxide (CO) rescue the detrimental effects of microglia-derived NO on neurite outgrowth. (B) Sine both, NO and CO, have the potential to bind to the soluble guanylyl cyclase (sGC) and thus to regulate the conversion of guanosine triphosphate (GTP) to cyclic guanosine monophosphate (cGMP) by the protein kinase G (PKG) it is likely that considerable parts of neurite growth are regulated downstream the sGC/cGMP cascade. In support of this, treatment with the commercial pain reliever Ibuprofen (RhoA (Ras homolog gene family, member A GTPase) and Rho kinase ROCK (Rho-associated coiled coil forming protein serine/threonine kinase) inhibiting agent) greatly increased neurite length of the human model neurons while activation of RhoA/ROCK result in growth cone collapse followed by the inhibition of neurite growth.

We and others revealed a pivotal role for RhoA (Ras homolog gene family, member A GTPase) and the Rho kinase ROCK (Rho-associated coiled coil forming protein serine/threonine kinase) signaling on neurite outgrowth of neurons (Boomkamp et al., 2012; DeGeer and Lamarche-Vane, 2013; Tönges et al., 2011; Wu et al., 2009).

Treatment with RhoA/ROCK inhibiting agents increased neurite length of the human

30 Discussion

model neurons (Roloff et al., 2015). Moreover, activation of RhoA caused cytoskeletal changes eventually leading to a growth cone collapse and suppression of axonal re-extension (Gallo and Letourneau, 2004). Since PKG, which is activated downstream the NO/cGMP cascade, regulates the expression and phosphorylation of RhoA (Sauzeau et al., 2003), it is possible that reactive microglia mediate their harmful effects via NO/cGMP/PKG - RhoA/ROCK activation (Fig. 8).

There are, however, also indications that NO-initiated growth cone collapse, axon retraction and reconfiguration of axonal microtubules are not mediated via cGMP/PKG (Ernst et al., 2000; He et al., 2002; Hess et al., 1993). A key event in NO signaling to the axonal cytoskeleton is triggered through the S-nitrosylation of the microtubule-associated protein 1B (MAP1B) (Stroissnigg et al., 2007) which is highly expressed during development and regeneration (Gordon-Weeks and Fischer, 2000;

Soares et al., 2002). S-nitrosylation of MAP1B increases its microtubule binding and inhibits dynein actions leading to a reduction of neurite extension force (Stroissnigg et al., 2007)

Even though microglia-mediated neurite outgrowth inhibition is thought to be partially responsible for the limited axonal recovery, it may initially be advantageous to prevent the ingrowth of neuronal processes into the hostile environment of wounded nervous tissue. Manipulation of microglial activation might thus offer a potential pharmacological target for controlling neuronal repair under acute inflammatory/traumatic conditions, but the exact time frame for a potential treatment has to be carefully evaluated (Block and Hong, 2005; Block et al., 2007; Greter and Merad, 2013; Rock and Peterson, 2006; Scheiblich and Bicker, 2015a).

Conclusion 31

Conclusion

This thesis describes the interrelationship between NO and CO in regulating cellular characteristics of microglia under basal conditions and after activation. For reasons of reducing animal consumption in research the BV-2 cell line was introduced as model system for primary microglia in functional assays. Our experiments on microglial activation, cell migration, phagocytosis of neuronal fragment, and the impact of microglia on neurite outgrowth of human model neurons contribute to better understand the signaling pathways underlying microglial modulation in pathogenic conditions (Fig. 9). Moreover, results of the present thesis support a therapeutic potential of NO synthesis blocking agents and CO-releasing molecules (CORMs) to medicate excessive inflammation within the CNS and to control the inflammatory responsiveness of reactive microglia.

Figure 9: Summary of the key findings of the present thesis. Cellular characteristics of microglia (green), including activation, migration, phagocytosis, and the impact on neurite outgrowth of human model neurons (magenta), are dually regulated by antagonistic interacting nitric oxide (NO) and carbon monoxide (CO) signal transduction pathways.

32 References

References

Akira, S., and Takeda, K. (2004). Toll-like receptor signalling. Nat. Rev. Immunol. 4, 499–

511.

Arimoto, T., and Bing, G. (2003). Up-regulation of inducible nitric oxide synthase in the substantia nigra by lipopolysaccharide causes microglial activation and neurodegeneration. Neurobiol. Dis. 12, 35–45.

Bach, F.H. (2005). Heme oxygenase-1: a therapeutic amplification funnel. FASEB J. 19, 1216–1219.

Bal-Price, A., and Brown, G.C. (2001). Inflammatory neurodegeneration mediated by nitric oxide from activated glia-inhibiting neuronal respiration, causing glutamate release and excitotoxicity. J. Neurosci. 21, 6480–6491.

Banati, R.B., Gehrmann, J., Schubert, P., and Kreutzberg, G.W. (1993). Cytotoxicity of microglia. Glia 7, 111–118.

Bani-Hani, M.G., Greenstein, D., Mann, B.E., Green, C.J., and Motterlini, R. (2006a). A carbon monoxide-releasing molecule (CORM-3) attenuates lipopolysaccharide-and interferon-gamma-induced inflammation in microglia. Pharmacol. Rep. 58, 132-144.

Bani-Hani, M.G., Greenstein, D., Mann, B.E., Green, C.J., and Motterlini, R. (2006b).

Modulation of thrombin-induced neuroinflammation in BV-2 microglia by carbon monoxide-releasing molecule 3. J. Pharmacol. Exp. Ther. 318, 1315–1322.

Baranano, D.E., and Snyder, S.H. (2001). Neural roles for heme oxygenase: Contrasts to nitric oxide synthase. Proc. Natl. Acad. Sci. 98, 10996–11002.

Bender, A.T., and Beavo, J.A. (2006). Cyclic nucleotide phosphodiesterases: molecular regulation to clinical use. Pharmacol. Rev. 58, 488–520.

Blasi, E., Barluzzi, R., Bocchini, V., Mazzolla, R., and Bistoni, F. (1990). Immortalization of murine microglial cells by a v-raf / v-myc carrying retrovirus. J. Neuroimmunol. 27, 229–237.

Block, M.L., and Hong, J.-S. (2005). Microglia and inflammation-mediated neurodegeneration: Multiple triggers with a common mechanism. Prog. Neurobiol. 76, 77–98.

Block, M.L., Zecca, L., and Hong, J.-S. (2007). Microglia-mediated neurotoxicity: uncovering the molecular mechanisms. Nat Rev Neurosci 8, 57–69.

Boehning, D., Sedaghat, L., Sedlak, T.W., and Snyder, S.H. (2004). Heme oxygenase-2 is activated by calcium-calmodulin. J. Biol. Chem. 279, 30927–30930.

Boje, K.M., and Arora, P.K. (1992). Microglial-produced nitric oxide and reactive nitrogen oxides mediate neuronal cell death. Brain Res. 587, 250–256.

References 33

Bolaños, J.P., Almeida, A., Stewart, V., Peuchen, S., Land, J.M., Clark, J.B., and Heales, S.J. (1997). Nitric oxide‐mediated mitochondrial damage in the brain: Mechanisms and implications for neurodegenerative diseases. J. Neurochem. 68, 2227–2240.

Boomkamp, S.D., Riehle, M.O., Wood, J., Olson, M.F., and Barnett, S.C. (2012). The development of a rat in vitro model of spinal cord injury demonstrating the additive effects of Rho and ROCK inhibitors on neurite outgrowth and myelination. Glia 60, 441–456.

Brown, G. (2007). Mechanisms of inflammatory neurodegeneration: iNOS and NADPH oxidase. Biochem. Soc. Trans. 35, 1119–1121.

Brown, G.C., and Neher, J.J. (2010). Inflammatory neurodegeneration and mechanisms of microglial killing of neurons. Mol. Neurobiol. 41, 242–247.

Chao, C.C., Hu, S., Sheng, W.S., Bu, D., Bukrinsky, M.I., and Peterson, P.K. (1996).

Cytokine‐stimulated astrocytes damage human neurons via a nitric oxide mechanism.

Glia 16, 276–284.

Chen, A., Kumar, S.M., Sahley, C.L., and Muller, K.J. (2000). Nitric oxide influences injury-induced microglial migration and accumulation in the leech CNS. J. Neurosci. 20, 1036–1043.

Choi, A., and Alam, J. (1996). Heme oxygenase-1: function, regulation, and implication of a novel stress-inducible protein in oxidant-induced lung injury. Am. J. Respir. Cell Mol.

Biol. 15, 9–19.

Colasanti, M., and Suzuki, H. (2000). The dual personality of NO. Trends Pharmacol. Sci. 21, 249–252.

Corradin, S.B., Mauël, J., Donini, S.D., Quattrocchi, E., and Ricciardi‐Castagnoli, P. (1993).

Inducible nitric oxide synthase activity of cloned murine microglial cells. Glia 7, 255–

262.

Cross, A.K., and Woodroofe, M.N. (1999). Chemokines induce migration and changes in actin polymerization in adult rat brain microglia and a human fetal microglial cell line in vitro. J. Neurosci. Res. 55, 17–23.

DeGeer, J., and Lamarche-Vane, N. (2013). Rho GTPases in neurodegeneration diseases.

Exp. Cell Res. 319, 2384–2394.

Dibaj, P., Nadrigny, F., Steffens, H., Scheller, A., Hirrlinger, J., Schomburg, E.D., Neusch, C., and Kirchhoff, F. (2010). NO mediates microglial response to acute spinal cord injury under ATP control in vivo. Glia 58, 1133–1144.

Duan, Y., Sahley, C.L., and Muller, K.J. (2009). ATP and NO dually control migration of microglia to nerve lesions. Dev. Neurobiol. 69, 60–72.

Elliott, M.R., Chekeni, F.B., Trampont, P.C., Lazarowski, E.R., Kadl, A., Walk, S.F., Park, D., Woodson, R.I., Ostankovich, M., and Sharma, P. (2009). Nucleotides released by apoptotic cells act as a find-me signal to promote phagocytic clearance. Nature 461, 282–286.

34 References

Ernst, A.F., Gallo, G., Letourneau, P.C., and McLoon, S.C. (2000). Stabilization of growing retinal axons by the combined signaling of nitric oxide and brain-derived neurotrophic factor. J. Neurosci. 20, 1458–1469.

Ferris, C.D., Jaffrey, S.R., Sawa, A., Takahashi, M., Brady, S.D., Barrow, R.K., Tysoe, S.A., Wolosker, H., Barañano, D.E., and Doré, S. (1999). Haem oxygenase-1 prevents cell death by regulating cellular iron. Nat. Cell Biol. 1, 152–157.

Gallo, G., and Letourneau, P.C. (2004). Regulation of growth cone actin filaments by guidance cues. J. Neurobiol. 58, 92–102.

Garthwaite, J. (2008). Concepts of neural nitric oxide-mediated transmission. Eur. J.

Neurosci. 27, 2783–2802.

Glass, C.K., Saijo, K., Winner, B., Marchetto, M.C., and Gage, F.H. (2010). Mechanisms underlying inflammation in neurodegeneration. Cell 140, 918–934.

Gordon-Weeks, P.R., and Fischer, I. (2000). MAP1B expression and microtubule stability in growing and regenerating axons. Microsc. Res. Tech. 48, 63–74.

Greter, M., and Merad, M. (2013). Regulation of microglia development and homeostasis.

Glia 61, 121–127.

Hanisch, U.-K., and Kettenmann, H. (2007). Microglia: active sensor and versatile effector cells in the normal and pathologic brain. Nat. Neurosci. 10, 1387–1394.

Hayashi, S., Omata, Y., Sakamoto, H., Higashimoto, Y., Hara, T., Sagara, Y., and Noguchi, M. (2004). Characterization of rat heme oxygenase-3 gene. Implication of processed pseudogenes derived from heme oxygenase-2 gene. Gene 336, 241–250.

Haynes, S.E., Hollopeter, G., Yang, G., Kurpius, D., Dailey, M.E., Gan, W.-B., and Julius, D.

(2006). The P2Y12 receptor regulates microglial activation by extracellular nucleotides. Nat. Neurosci. 9, 1512–1519.

He, Y., Yu, W., and Baas, P.W. (2002). Microtubule reconfiguration during axonal retraction induced by nitric oxide. J. Neurosci. 22, 5982–5991.

Henn, A., Lund, S., Hedtjärn, M., Schrattenholz, A., Pörzgen, P., and Leist, M. (2009). The suitability of BV2 cells as alternative model system for primary microglia cultures or for animal experiments examining brain inflammation. ALTEX 26, 83–94.

Hess, D.T., Patterson, S.I., Smith, D.S., and Skene, J.P. (1993). Neuronal growth cone collapse and inhibition of protein fatty acylation by nitric oxide. Nature 366, 562–565.

Hofmann, F., Feil, R., Kleppisch, T., and Schlossmann, J. (2006). Function of cGMP-dependent protein kinases as revealed by gene deletion. Physiol. Rev. 86, 1–23.

Horvath, R.J., Nutile-McMenemy, N., Alkaitis, M.S., and DeLeo, J.A. (2008). Differential migration, LPS-induced cytokine, chemokine, and NO expression in immortalized BV-2 and HAPI cell lines and primary microglial cultures. J. Neurochem. 107, 557–569.

References 35

Hunot, S., Boissiere, F., Faucheux, B., Brugg, B., Mouatt-Prigent, A., Agid, Y., and Hirsch, E.

(1996). Nitric oxide synthase and neuronal vulnerability in Parkinson’s disease.

Neuroscience 72, 355–363.

Ignarro, L.J. (1990). Nitric oxide. A novel signal transduction mechanism for transcellular communication. Hypertension 16, 477–483.

Itoh, K., Chiba, T., Takahashi, S., Ishii, T., Igarashi, K., Katoh, Y., Oyake, T., Hayashi, N., Satoh, K., and Hatayama, I. (1997). An Nrf2/small Maf heterodimer mediates the induction of phase II detoxifying enzyme genes through antioxidant response elements. Biochem. Biophys. Res. Commun. 236, 313–322.

Kettenmann, H., and Verkhratsky, A. (2011). [Neuroglia--living nerve glue]. Fortschritte Neurol. Psychiatr. 588–597.

Keyse, S.M., and Tyrrell, R.M. (1987). Both near ultraviolet radiation and the oxidizing agent hydrogen peroxide induce a 32-kDa stress protein in normal human skin fibroblasts.

J. Biol. Chem. 262, 14821–14825.

Keyse, S.M., and Tyrrell, R.M. (1989). Heme oxygenase is the major 32-kDa stress protein induced in human skin fibroblasts by UVA radiation, hydrogen peroxide, and sodium arsenite. Proc. Natl. Acad. Sci. 86, 99–103.

Kharitonov, V.G., Sharma, V.S., Pilz, R.B., Magde, D., and Koesling, D. (1995). Basis of Axonal Growth through RGMa. PLoS ONE 6, e25234.

Knipp, S., and Bicker, G. (2009). Regulation of enteric neuron migration by the gaseous messenger molecules CO and NO. Development 136, 85–93.

Koglin, M., and Behrends, S. (2002). Biliverdin IX is an endogenous inhibitor of soluble guanylyl cyclase. Biochem. Pharmacol. 64, 109–116.

Kreutzberg, G.W. (1996). Microglia: a sensor for pathological events in the CNS. Trends Neurosci. 19, 312–318.

Kumar, A., Chen, S.-H., Kadiiska, M.B., Hong, J.-S., Zielonka, J., Kalyanaraman, B., and Mason, R.P. (2014). Inducible nitric oxide synthase is key to peroxynitrite-mediated, LPS-induced protein radical formation in murine microglial BV2 cells. Free Radic.

Biol. Med. 73, 51-59.

Lawson, L., Perry, V., Dri, P., and Gordon, S. (1990). Heterogeneity in the distribution and morphology of microglia in the normal adult mouse brain. Neuroscience 39, 151–170.

Lee, S., and Suk, K. (2007). Heme oxygenase-1 mediates cytoprotective effects of immunostimulation in microglia. Biochem. Pharmacol. 74, 723–729.

36 References

Lee, J.Y., Jhun, B.S., Oh, Y.T., Lee, J.H., Choe, W., Baik, H.H., Ha, J., Yoon, K.-S., Kim, S.S., and Kang, I. (2006). Activation of adenosine A3 receptor suppresses lipopolysaccharide-induced TNF-α production through inhibition of PI 3-kinase/Akt and NF-κB activation in murine BV2 microglial cells. Neurosci. Lett. 396, 1–6.

Lee, S.M., Yune, T.Y., Kim, S.J., Kim, Y.C., Oh, Y.J., Markelonis, G.J., and Oh, T.H. (2004).

Minocycline inhibits apoptotic cell death via attenuation of TNF-α expression following iNOS/NO induction by lipopolysaccharide in neuron/glia co-cultures. J. Neurochem.

91, 568–578.

Liao, H. (2004). Tenascin-R plays a role in neuroprotection via its distinct domains that coordinate to modulate the microglia function. J. Biol. Chem. 280, 8316–8323.

Liu, B., and Hong, J.-S. (2003). Role of microglia in inflammation-mediated neurodegenerative diseases: mechanisms and strategies for therapeutic intervention.

J. Pharmacol. Exp. Ther. 304, 1–7.

Liu, Y., Zhu, B., Wang, X., Luo, L., Li, P., Paty, D.W., and Cynader, M.S. (2003). Bilirubin as a potent antioxidant suppresses experimental autoimmune encephalomyelitis:

implications for the role of oxidative stress in the development of multiple sclerosis. J.

Neuroimmunol. 139, 27–35.

Loihl, A.K., and Murphy, S. (1998). Expression of nitric oxide synthase-2 in glia associated with CNS pathology. Prog. Brain Res. 118, 253–267.

Lu, D.-Y., Tsao, Y.-Y., Leung, Y.-M., and Su, K.-P. (2010). Docosahexaenoic acid suppresses neuroinflammatory responses and induces heme oxygenase-1 expression in BV-2 microglia: implications of antidepressant effects for Omega-3 fatty acids. Neuropsychopharmacology 35, 2238–2248.

Lucin, K.M., and Wyss-Coray, T. (2009). Immune activation in brain aging and neurodegeneration: too much or too little? Neuron 64, 110–122.

Madhusoodanan, K., and Murad, F. (2007). NO-cGMP signaling and regenerative medicine involving stem cells. Neurochem. Res. 32, 681–694.

Maines, M.D. (1997). The heme oxygenase system: a regulator of second messenger gases.

Annu. Rev. Pharmacol. Toxicol. 37, 517–554.

Mander, P., and Brown, G.C. (2005). Activation of microglial NADPH oxidase is synergistic with glial iNOS expression in inducing neuronal death: a dual-key mechanism of inflammatory neurodegeneration. J. Neuroinflammation 2, 20.

Mariño, G., and Kroemer, G. (2013). Mechanisms of apoptotic phosphatidylserine exposure.

Cell Res. 23, 1247-1248.

McNaught, K.S.P., and Brown, G.C. (1998). Nitric oxide causes glutamate release from brain synaptosomes. J. Neurochem. 70, 1541–1546.

Min, K.-J., Yang, M., Kim, S.-U., Jou, I., and Joe, E. (2006). Astrocytes induce hemeoxygenase-1 expression in microglia: a feasible mechanism for preventing excessive brain inflammation. J. Neurosci. 26, 1880–1887.

References 37

Minghetti, L., and Levi, G. (1998). Microglia as effector cells in brain damage and repair:

focus on prostanoids and nitric oxide. Prog. Neurobiol. 54, 99–125.

Moncada, S., Palmer, R., and Higgs, E. (1989). The biological significance of nitric oxide formation from L-arginine. Biochem Soc Trans 17, 642–644.

Morgan, S.C., Taylor, D.L., and Pocock, J.M. (2004). Microglia release activators of neuronal proliferation mediated by activation of mitogen-activated protein kinase, phosphatidylinositol-3-kinase/Akt and delta-Notch signalling cascades: Microglia enhance neuronal survival and proliferation. J. Neurochem. 90, 89–101.

Moss, D.W., and Bates, T.E. (2001). Activation of murine microglial cell lines by lipopolysaccharide and interferon-gamma causes NO-mediated decreases in mitochondrial and cellular function. Eur. J. Neurosci. 13, 529–538.

Motterlini, R., and Otterbein, L.E. (2010). The therapeutic potential of carbon monoxide. Nat.

Rev. Drug Discov. 9, 728–743.

Nath, K.A., Balla, G., Vercellotti, G.M., Balla, J., Jacob, H.S., Levitt, M., and Rosenberg, M.E.

(1992). Induction of heme oxygenase is a rapid, protective response in rhabdomyolysis in the rat. J. Clin. Invest. 90, 267-270.

Nathan, C. (1992). Nitric oxide as a secretory product of mammalian cells. FASEB J. 6, 3051–3064.

Neher, J.J., Neniskyte, U., and Brown, G.C. (2012). Primary phagocytosis of neurons by inflamed microglia: potential roles in neurodegeneration. Front. Pharmacol. 3:27.

Neumann, H., Kotter, M.R., and Franklin, R.J.M. (2009). Debris clearance by microglia: an essential link between degeneration and regeneration. Brain 132, 288–295.

Nimmerjahn, A., Kirchhoff, F., and Helmchen, F. (2005). Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 308, 1314–1318.

Ohsawa, K., and Kohsaka, S. (2011). Dynamic motility of microglia: Purinergic modulation of microglial movement in the normal and pathological brain. Glia 59, 1793–1799.

Ohsawa, K., Irino, Y., Nakamura, Y., Akazawa, C., Inoue, K., and Kohsaka, S. (2007).

Involvement of P2X4 and P2Y12 receptors in ATP-induced microglial chemotaxis.

Glia 55, 604–616.

Otterbein, L.E., and Choi, A.M. (2000). Heme oxygenase: colors of defense against cellular stress. Am. J. Physiol.-Lung Cell. Mol. Physiol. 279, L1029–L1037.

Otterbein, L.E., Soares, M.P., Yamashita, K., and Bach, F.H. (2003). Heme oxygenase-1:

unleashing the protective properties of heme. Trends Immunol. 24, 449–455.

Polazzi, E., Gianni, T., and Contestabile, A. (2001). Microglial cells protect cerebellar granule neurons from apoptosis: Evidence for reciprocal signaling. Glia 36, 271–280.

Popovich, P.G., Guan, Z., Wei, P., Huitinga, I., van Rooijen, N., and Stokes, B.T. (1999).

Depletion of hematogenous macrophages promotes partial hindlimb recovery and

38 References

neuroanatomical repair after experimental spinal cord injury. Exp. Neurol. 158, 351–

365.

Rock, R.B., and Peterson, P.K. (2006). Microglia as a pharmacological target in infectious and inflammatory diseases of the brain. J. Neuroimmune Pharmacol. 1, 117–126.

Roloff, F., Scheiblich, H., Dewitz, C., Dempewolf, S., Stern, M., and Bicker, G. (2015).

Enhanced neurite outgrowth of human model (NT2) neurons by small-molecule inhibitors of Rho/ROCK signaling. PLoS ONE 10, e0118536.

Ryter, S.W., Otterbein, L.E., Morse, D., and Choi, A.M. (2002). Heme oxygenase/carbon monoxide signaling path-ways: Regulation and functional significance. In Oxygen/Nitrogen Radicals: Cell Injury and Disease, (Springer), 249–263.

Sauzeau, V., Rolli-Derkinderen, M., Marionneau, C., Loirand, G., and Pacaud, P. (2003).

RhoA expression is controlled by nitric oxide through cGMP-dependent protein kinase activation. J. Biol. Chem. 278, 9472–9480.

Savill, J., Dransfield, I., Gregory, C., and Haslett, C. (2002). A blast from the past: clearance of apoptotic cells regulates immune responses. Nat. Rev. Immunol. 2, 965–975.

Sawle, P., Foresti, R., Mann, B.E., Johnson, T.R., Green, C.J., and Motterlini, R. (2005).

Carbon monoxide-releasing molecules (CO-RMs) attenuate the inflammatory response elicited by lipopolysaccharide in RAW264.7 murine macrophages. Br. J.

Pharmacol. 145, 800–810.

Scheiblich, H., and Bicker, G. (2015a). Regulation of microglial migration, phagocytosis, and neurite outgrowth by HO-1/CO signaling. Dev. Neurobiol. 75, 854–876.

Scheiblich, H., and Bicker, G. (2015b). Nitric oxide regulates antagonistically phagocytic and neurite outgrowth inhibiting capacities of microglia. Dev. Neurobiol. in press.

Scheiblich, H., Roloff, F., Singh, V., Stangel, M., Stern, M., and Bicker, G. (2014). Nitric oxide/cyclic GMP signaling regulates motility of a microglial cell line and primary microglia in vitro. Brain Res. 1564, 9–21.

Schilling, T., Stock, C., Schwab, A., and Eder, C. (2004). Functional importance of Ca2+‐activated K+ channels for lysophosphatidic acid‐induced microglial migration.

Eur. J. Neurosci. 19, 1469–1474.

Sierra, A., Abiega, O., Shahraz, A., and Neumann, H. (2013). Janus-faced microglia:

beneficial and detrimental consequences of microglial phagocytosis. Front. Cell.

Neurosci. 7:6.

Soares, M.P., and Bach, F.H. (2009). Heme oxygenase-1: from biology to therapeutic potential. Trends Mol. Med. 15, 50–58.

Soares, S., Boxberg, Y. Von, Lombard, M., Ravaille‐Veron, M., Fischer, I., Eyer, J., and Nothias, F. (2002). Phosphorylated MAP1B is induced in central sprouting of primary afferents in response to peripheral injury but not in response to rhizotomy. Eur. J.

Neurosci. 16, 593–606.

References 39

Srisook, K., Han, S.-S., Choi, H.-S., Li, M.-H., Ueda, H., Kim, C., and Cha, Y.-N. (2006). CO from enhanced HO activity or from CORM-2 inhibits both O2− and NO production and downregulates HO-1 expression in LPS-stimulated macrophages. Biochem.

Pharmacol. 71, 307–318.

Stansley, B., Post, J., and Hensley, K. (2012). A comparative review of cell culture systems for the study of microglial biology in Alzheimer’s disease. J. Neuroinflammation 9, 115.

Stocker, R., Yamamoto, Y., McDonagh, A.F., Glazer, A.N., and Ames, B.N. (1987). Bilirubin is an antioxidant of possible physiological importance. Science 235, 1043–1046.

Streit, W.J., Walter, S.A., and Pennell, N.A. (1999). Reactive microgliosis. Prog. Neurobiol.

57, 563–581.

Stroissnigg, H., Trančíková, A., Descovich, L., Fuhrmann, J., Kutschera, W., Kostan, J., Meixner, A., Nothias, F., and Propst, F. (2007). S-nitrosylation of microtubule-associated protein 1B mediates nitric-oxide-induced axon retraction. Nat. Cell Biol. 9, 1035–1045.

Suzuki, J., Denning, D.P., Imanishi, E., Horvitz, H.R., and Nagata, S. (2013). Xk-related protein 8 and CED-8 promote phosphatidylserine exposure in apoptotic cells. Science 341, 403–406.

Syapin, P. (2008). Regulation of haeme oxygenase‐1 for treatment of neuroinflammation and brain disorders. Br. J. Pharmacol. 155, 623–640.

Terazawa, R., Akimoto, N., Kato, T., Itoh, T., Fujita, Y., Hamada, N., Deguchi, T., Iinuma, M., Noda, M., Nozawa, Y., et al. (2013). A kavalactone derivative inhibits lipopolysaccharide-stimulated iNOS induction and NO production through activation of Nrf2 signaling in BV2 microglial cells. Pharmacol. Res. 71, 34–43.

Tönges, L., Koch, J.-C., Bähr, M., and Lingor, P. (2011). ROCKing regeneration: Rho kinase inhibition as molecular target for neurorestoration. Front. Mol. Neurosci. 4:39.

Turcanu, V., Dhouib, M., and Poindron, P. (1998a). Heme oxygenase inhibits nitric oxide synthase by degrading heme: a negative feedback regulation mechanism for nitric oxide production. Transplant. Proc. 30, 4184–4185.

Turcanu, V., Dhouib, M., and Poindron, P. (1998b). Nitric oxide synthase inhibition by haem oxygenase decreases macrophage nitric-oxide-dependent cytotoxicity: a negative feedback mechanism for the regulation of nitric oxide production. Res. Immunol. 149, 741–744.

Vicente, A.M., Guillén, M.I., and Alcaraz, M.J. (2001). Modulation of haem oxygenase-1 expression by nitric oxide and leukotrienes in zymosan-activated macrophages. Br. J.

Pharmacol. 133, 920–926.

Vincent, S.R. (1994). Nitric oxide: a radical neurotransmitter in the central nervous system.

Prog. Neurobiol. 42, 129–160.

40 References

Walter, L., Franklin, A., Witting, A., Wade, C., Xie, Y., Kunos, G., Mackie, K., and Stella, N.

(2003). Nonpsychotropic cannabinoid receptors regulate microglial cell migration. J.

Neurosci. 23, 1398–1405.

Wang, M.-J., Lin, W.-W., Chen, H.-L., Chang, Y.-H., Ou, H.-C., Kuo, J.-S., Hong, J.-S., and Jeng, K.-C.G. (2002). Silymarin protects dopaminergic neurons against lipopolysaccharide-induced neurotoxicity by inhibiting microglia activation. Eur. J.

Neurosci. 16, 2103–2112.

Wang, W.W., Smith, D.L., and Zucker, S.D. (2004). Bilirubin inhibits iNOS expression and NO production in response to endotoxin in rats. Hepatology 40, 424–433.

Wu, D., Yang, P., Zhang, X., Luo, J., Haque, M.E., Yeh, J., Richardson, P.M., Zhang, Y., and Bo, X. (2009). Targeting a dominant negative rho kinase to neurons promotes axonal outgrowth and partial functional recovery after rat rubrospinal tract lesion. Mol. Ther.

Wu, D., Yang, P., Zhang, X., Luo, J., Haque, M.E., Yeh, J., Richardson, P.M., Zhang, Y., and Bo, X. (2009). Targeting a dominant negative rho kinase to neurons promotes axonal outgrowth and partial functional recovery after rat rubrospinal tract lesion. Mol. Ther.