Glia-Neuron Interactions in The Mammalian Retina Review - Vecino 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

Progress in Retinal and Eye Research 51 (2016) 1e40

Contents lists available at ScienceDirect

Progress in Retinal and Eye Research


journal homepage: www.elsevier.com/locate/prer

Review

Gliaeneuron interactions in the mammalian retina


Elena Vecino a, *, 1, 2, F.David Rodriguez b, 1, Noelia Ruzafa a, 1, 2, Xandra Pereiro a, 1, 2,
Sansar C. Sharma c, d, 1
a
Department of Cell Biology and Histology, University of the Basque Country UPV/EHU, Leioa 48940, Vizcaya, Spain
b
Department of Biochemistry and Molecular Biology, E-37007, University of Salamanca, Salamanca, Spain
c
Department of Ophthalmology, Cell Biology and Anatomy, New York Medical College, Valhalla, NY 10595, USA
d
IKERBASQUE, Basque Foundation for Science at Dept. Cell Biology and Histology, UPV/EHU, Spain

a r t i c l e i n f o a b s t r a c t

Article history: The mammalian retina provides an excellent opportunity to study gliaeneuron interactions and the
Received 30 January 2015 interactions of glia with blood vessels. Three main types of glial cells are found in the mammalian retina
Received in revised form that serve to maintain retinal homeostasis: astrocytes, Müller cells and resident microglia. Müller cells,
18 May 2015 astrocytes and microglia not only provide structural support but they are also involved in metabolism,
Accepted 2 June 2015
the phagocytosis of neuronal debris, the release of certain transmitters and trophic factors and Kþ up-
Available online 23 June 2015
take. Astrocytes are mostly located in the nerve fibre layer and they accompany the blood vessels in the
inner nuclear layer. Indeed, like Müller cells, astrocytic processes cover the blood vessels forming the
Keywords: retinal blood barrier and they fulfil a significant role in ion homeostasis. Among other activities, microglia
Retina
can be stimulated to fulfil a macrophage function, as well as to interact with other glial cells and neurons
Neurons
Glial cells
by secreting growth factors. This review summarizes the main functional relationships between retinal

List of abbreviations: AAV, adeno-associated virus; Ach, acetylcholine; AGA, amadori-glycated albumin; AGE, advanced glycation end-products; AhR, aryl hydrocarbon
receptor; AKT, protein kinase B; AMD, age-related macular degeneration; AP-1, activating protein-1; APC, antigen-presenting cells; Apo, apolipoprotein; AQP, aquaporin;
ASCL1, achaete-scute family bHLH transcription factor 1; Atoh7, atonal homolog 7; ATP, adenosine triphosphate; BBB, bloodebrain barrier; Bcl-2, B-cell lymphoma 2; BDNF,
brain-derived neurotrophic factor; bFGF, basic fibroblast growth factor; BMP, bone morphogenetic protein; BV, blood vessel; CaMKII, Ca2þ/calmodulin-dependent protein
kinase; Car2, carbonic anhydrase II; CCL2, monocyte Chemotactic Protein-1; CCR2, CeC chemokine receptor type 2; CD45, cluster of differentiation 45; Cdc14A, cell division
cycle 14A; Cdk10, cyclin-dependent kinase 10; Ch, choroid; ChAT, choline acetyltransferase; CHX10, Ceh-10 homeodomain-containing homolog; CNS, central nervous system;
CNTF, ciliary neurotrophic factor; CNV, choroidal neovascularisation; CRALBP, cellular retinaldehyde-binding protein C; CREB, cAMP response element-binding protein;
CSF1R, colony stimulating factor 1 receptor; CSPG, chondroitin sulphate proteoglycans; CTGF, connective tissue growth factor; CX3CL1, chemokine (C-X3-C Motif) Ligand 1;
CX3CR1, CX3C chemokine receptor 1; DAPI, 40 ,6-diamidino-2-phenylindole; DAPT, g-Secretase inhibitor; DHA, docosahexaenoic acid; DKK3, Dickkopf-related protein 3; DR,
diabetic retinopathy; EAAT, excitatory amino acid transporter; ECM, extracellular matrix; ED-1, monoclonal antibody against glycoprotein CD68; EGF, epidermal growth
factor; EMV, extracellular membrane microvesicles; ESMV, embryonic stem cells microvesicles; FG, fluorogold; FGF, fibroblast growth factor; FKN, fraktalkine; G6P, Glucose-
6-Phosphate; GABA, g-aminobutyric acid; gadd45b, growth Arrest And DNA-Damage-Inducible Beta; GCL, ganglion cell layer; GDNF, glia-derived neurotrophic factor; GFAP,
glial fibrillary acidic protein; GFP, green fluorescent protein; GK, Goto Kakizaki; GLAST, glutamate-aspartate transporter; GLT1, glial Glutamate Transporter 1; Gpc, glypican;
GS, glutamine synthetase; hMGSC, human Müller glia stem cells; HSPG, heparin sulphate proteoglycan; IBA1, ionized calcium-binding adapter molecule 1; IF, intermediate
filaments; IFN, interferon; IGF, insulin-like growth factor; IL, interleukin; ILM, inner limiting membrane; INL, inner nuclear layer; IOP, intraocular pressure; IP3R2, IP3 receptor
type 2; IPL, inner plexiform layer; JNK, C-Jun N-terminal kinase; Kir, inwardly rectifying potassium channels; LCA, Leber congenital amaurosis; LDL, low-density lipoproteins;
LGN, lateral geniculate nucleus; LIF, leukaemia inhibitory factor; MAPK, mitogen-activated protein kinases; Mash1, mammalian achaete-scute homolog-1; MCP1, monocyte
chemoattractant protein-1; Megf10, multiple epidermal growth factor-like domains protein 10; Mertk, tyrosine-protein kinase Mer; MHC, major histocompatibility complex;
MMP, matrix metalloproteinases; mTOR, mechanistic target of rapamycin; NAD, nicotine adenine dinucleotide; NADPH, nicotinamide adenine dinucleotide phosphate; NF,
neurofilaments; NFL, nerve fibre layer; NFk&Bgr, uclear factor kappa-light-chain-enhancer of activated B cells; NGF, nerve growth factor; NMDA, N-Methyl-D-aspartate; NO,
nitric oxide; NPD1, neuroprotectin D1; NPPB, 5-nitro-2-[(3-phenylpropyl)amino]benzoic acid; NV, neovascularisation; OIR, oxygen-induced retinopathy; OLM, outer limiting
membrane; ON, optic nerve; ONL, outer nuclear layer; OPC, oligodendrocyte precursor cells; OPL, outer plexiform layer; OPN, osteopontin; OS, outer segment layer; P2Y,
metabotropic purinergic; Pax6, paired box 6; PCAG, primary closed-angle glaucoma; PCG, primary congenital glaucoma; PE, pigment epithelium; PEDF, pigment epithelium-
derived factor; PI3K, phosphatidylinositol-4,5-bisphosphate 3-kinase; PKB, protein kinase B; PLC, phospholipase C; POAG, primary open angle glaucoma; PSD95, postsynaptic
density protein 95; rd10, retinal degeneration 10; RFP, red fluorescent protein; RGC, retinal ganglion cells; RGCL, retinal ganglion cells layer; RNA, ribonucleic acid; ROS,
reactive oxygen species; RP, retinitis pigmentosa; RPE, retinal pigment epithelium; RT-PCR, reverse transcription polymerase chain reaction; RVD, regulatory Volume
Decrease; SELDI-TOF, surface-enhanced laser desorption ionization-time of flight; SNAP25, synaptosomal-associated protein 25; SP-A, surfactant protein A; Spbc25, spindle
pole body component 25; STAT, signal transducers and activators of transcription transient; STZ, streptozotozin; TAK1, transforming growth factor b-activated kinase 1; TCA,
tricarboxylic acid cycle; TGF&Bgr, transforming growth factor beta; Thy-1, thymocyte antigen 1; TN-C, tenascin C; TNF, tumour necrosis factor; TRK&Bgr, tropomyosin re-
ceptor kinase B; uPA, urinary-type plasminogen activator; VEGF, vascular endothelial growth factor; Vim, vimentin; WNT, Wingless-Type MMTV integration site family.
* Corresponding author.
1
Each co-author contributed equally (20%) to this article.
2
www.ehu.es/GOBE.

http://dx.doi.org/10.1016/j.preteyeres.2015.06.003
1350-9462/© 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
2 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Astrocytes glial cells and neurons, presenting a general picture of the retina recently modified based on experi-
Müller glia mental observations. The preferential involvement of the distinct glia cells in terms of the activity in the
Microglia retina is discussed, for example, while Müller cells may serve as progenitors of retinal neurons, astrocytes
Macrophages
and microglia are responsible for synaptic pruning. Since different types of glia participate together in
Retinal ganglion cells
certain activities in the retina, it is imperative to explore the order of redundancy and to explore the
Photoreceptors
Glaucoma heterogeneity among these cells. Recent studies revealed the association of glia cell heterogeneity with
Retinitis pigmentosa specific functions. Finally, the neuroprotective effects of glia on photoreceptors and ganglion cells under
Neuroprotection normal and adverse conditions will also be explored.
Neurotrophins © 2015 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND
Plasticity license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Integrins
Extracellular matrix

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2. Origin and morphology of the retinal glial cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Müller cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.1. Origin and morphology of Müller cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.2. Müller cell markers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2. Astrocytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.1. Origin and morphology of astrocytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2.2. Astrocyte markers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.3. Astrocytes main functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3. Microglial cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3.1. Origin, morphology and distribution of microglia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3.2. Plasticity and functional microglia phenotypes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.3. Representative molecular traits of microglia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3.4. Identification of microglia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3. Common functions of different glial cells in the retina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1. Formation of boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2. Glia as a structural support for neurons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2.1. Viscoelastic properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2.2. Architecture-mechanical forces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.3. Volume regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.4. Metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.4.1. Glucose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.4.2. Lipid metabolism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.5. Extracellular matrix generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.5.1. Laminin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.5.2. Collagen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.5.3. Vitronectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.5.4. Fibronectin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.5.5. Tenascin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.5.6. Proteoglycans . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.5.7. Intermediate Filaments (IFs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.5.8. N-Cadherin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.5.9. Integrins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.6. Regulation of neuronal activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.6.1. Gamma-Aminobutyric acid (GABA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.6.2. Glutamate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.6.3. ATP and adenosine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.7. Synaptic pruning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.8. Immune responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.9. Gliosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.10. Neuroprotection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.10.1. Neurotrophic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.10.2. Other trophic factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.11. Phagocytosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.12. Repair and regeneration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4. Specific functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.1. Light guidance of Müller cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2. Progenitor cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
5. Heterogeneity of retinal glial cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.1. Astrocyte heterogeneity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
5.2. Müller cell heterogeneity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
6. Glia and retinal pathologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
6.1. Diabetic retinopathy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
6.2. Glaucoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 3

6.3. Age-related macular degeneration (AMD) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


6.4. Retinitis pigmentosa (RP) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
7. Conclusions, remarks and future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

1. Introduction to restore homeostasis, when they malfunction, these cells become


what can be considered a primary pathogenic element. Little is
The concept of neuroglia was first proposed by Rudolf Virchow known about the regulation of gliaeglia and gliaeneuronal in-
in 1858, in reference to the cellular elements that make up the teractions within the retina. However, since the retina is a well
connective tissue of the brain “nervekitt”. The Müller cell was structured organ within the Central Nervous System (CNS), studies
described as a radial fibre in the retina by Heinrich Müller in 1851, of the different roles of glial cells in health and disease have allowed
later becoming known as Müller glial cells. In 1893, the term researchers to understand some of their interactions with neurons.
astrocyte first appeared, used by Michael von Lenhossek in refer- The interactions between glia and neurons contribute to retinal
ence to the stellate morphology (from the Greek for star, astro, and homeostasis, and while some such interactions have been defined,
cell, cyte) of the cells that Camillo Golgi had previously seen in the we are still far from fully understanding how homeostasis is
nervous system, describing them as the “glue” of the brain. These maintained in the retina. Thus, in this review we will establish
cells were soon subdivided into fibrous and protoplasmic astro- some of the common features of glial cells in the retina and we will
cytes by Kolliker and Andriezen, yet their true diversity was not describe their possible interactions with neurons.
fully appreciated until Ramon y Cajal's drawings revealed their
extraordinary pleomorphism. Based on his histological studies,
several roles for this diverse class of cells were postulated, 2. Origin and morphology of the retinal glial cells
including a role in the maintenance of brain architecture, homeo-
stasis and nutrition. In 1919, oligodendrocytes (from the Greek for 2.1. Müller cells
few, oligo, branches, dendro, and cell cyte) and microglia were
recognized as separated cell types and categorised as neuroglia by 2.1.1. Origin and morphology of Müller cells
Pio del Rio-Hortega, one of the first scientists to propose that The Müller cell is the predominant glial element in the retina,
microglia are of mesodermal origin, and that they can migrate and representing 90% of the retinal glia. These are radially oriented cells
act as phagocytes (Kettenmann and Verkhratsky, 2008). that traverse the retina from its inner (vitreal) border to the distal
The vertebrate retina contains four types of glia, each of which end of the outer nuclear layer (ONL: Fig. 2). Although Müller cells
exhibits distinct morphological, developmental and antigenic share a lineage with retinal neurons, their time of birth appears to
characteristics. Müller cells are the most predominant glial be different. Recent studies found the surprising origin of Müller
element, followed by astrocytes and microglia. The fourth glial cell cells in the neural crest cell (Liu et al., 2014a). Birth dating of retinal
type, the oligodendrocyte, is seen only occasionally in the retina, neurons and Müller cells was determined by 3H-thymidine-label-
associated with myelinated ganglion cell axons in a few species like ling (Sidman, 1961) and all retinal cell types are generated in an
the rabbit. In this review, we will focus on general aspects of the orderly fashion, the sequence of which is largely conserved among
mammalian retina and thus, we will not mention oligodendrocytes, vertebrates (Vecino and Acera, 2015). Accordingly, ganglion cells
even though their activity and interactions with other astrocytes in are the first cells to be born, followed closely by horizontal cells,
the optic nerve, and with retinal ganglion cell (RGC) axons, might cones and displaced amacrine cells. At later stages, amacrine cells,
affect retinal function in some species (Fig. 1). bipolar cells, rod cells and Müller cells are generated, and indeed,
Despite studying the glia for more than a century the full influ- the general consensus is that Müller cells are the last cells to
ence of glial cells on neuronal centres is still not fully clear. While become post-mitotic (Cepko, 1993).
Cajal and others postulated several roles for glial cells (including The idea of a functional unit in which local interactions occur
maintaining brain architecture, homoeostasis, nutrition and regu- between a group of retinal neurons and their supportive glial cell
lating neuronal communication), we now know that glia are involved gave rise to the proposal of co-operativity among cells. This idea
in many other roles within the nervous system, for example influ- was based on developmental studies, where Müller cells were
encing synaptic connectivity by controlling the genesis and mainte- thought to perform many of the metabolic, ionic and extracellular
nance of synapses, and by modifying synaptic strength and plasticity. buffering requirements of neurons, with which they share a com-
Although Müller, astrocytes and microglial cells each have a different mon progenitor (Turner and Cepko, 1987). Later, an interesting
developmental origin, they share many functions within the retina. concept arose suggesting that co-operativity exists among cells that
Some of these activities are performed simultaneously, while others arise from a common stem cell to form a columnar array
are carried out by one or other of the glial cell types. More recently (Reichenbach et al., 1993). The idea is that the retina is constituted
glial cells have been established as a source of progenitors in the by many functional units in which local interactions occur between
brain, as not only can radial glia act as pluripotent neural precursors the group of retinal neurons and their supportive Müller glial cell,
but also, some astrocytes can re-enter the cell cycle and produce all limiting the sphere of influence of the latter. Thus, each Müller cell
types of brain cells, from neurons to macroglial cells. may only have to meet the requirements of its immediate neigh-
It is known that glial cells have intrinsic signalling systems that bours for the extracellular environment to remain stable in the face
can spread in the form of Ca2þ waves and that are correlated with of intense neural activity.
glutamate release. Among other functions, this signalling plays an In retinal sections, the processes of Müller cells are seen to
important role in immunity, angiogenesis and neuroprotection. ensheath the cell bodies and processes of retinal neurons, espe-
However, while neuroglia responds to injury as part of the defence cially after retinal damage (Hernandez et al., 2010). Thus, it seems
likely that the intimate association between these two cell types
4 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Fig. 1. Schematic drawing of the cellular components of the retina: glia and neurons. The different cell types are situated in a standard large mammalian retina. Note the in-
teractions between the cells and blood vessels (BV). Amacrine cells (A), astrocytes in green (AS), bipolar cells (B), cones (C), ganglion cells (G), horizontal cells (H), Müller cells in blue
(M), microglia in red (Mi), rods (R), cones (C). Note the location of the different layers of the retina (from the most internal to the outer layers): optic nerve (ON), nerve fibre layer
(NFL), ganglion cell layer (GCL), inner plexiform layer (IPL), inner nuclear layer (INL), outer plexiform layer (OPL), outer nuclear layer (ONL), outer segment layer (OS), pigment
epithelium (PE), choroid (Ch).

enables glial cells to promote the formation of synapses, and it cells also possess a cilium (Ennis and Kunz, 1986). In the opposite
would help maintain neuronal function by providing the nerve direction, the inner process approaches the vitreal surface of the
terminals with energy substrates and neurotransmitter precursors neuroretina, forming a so-called endfoot that abuts the basal lam-
(Pfrieger and Barres, 1996). Ensheathing by Müller cells is only ina between the vitreous body and the neuroretina (the “inner
possible to observe when membrane markers, like low affinity re- limiting membrane”, ILM). In the nuclear layers, the lamellar pro-
ceptor p75, are used to stain the Müller cells, as it is not evident cesses of Müller cells form structures that envelop the cell bodies of
with cytoskeletal markers (Fig. 3). As there are no specific astrocyte neuronal cells (Dreher et al., 1992; Hama et al., 1978; Reichenbach
membrane markers, and GFAP (glial fibrillary acidic protein) and et al., 1988, 1989). In addition, Müller cells are involved in the
p75 co-localize in both Müller cells and astrocytes, it is possible that structural organization of the blooderetina barrier (Reichenbach
at least some of the ensheathing membranes around RGCs could et al., 1995). Blood capillaries are ensheathed by Müller cell pro-
belong to astrocytes. Moreover, in some species like dogs, the RGC cesses that act as a communication system for metabolic exchange
axons are packaged and intercalated between the end feet of between the vasculature and neurons, in much the same manner as
groups of Müller cells on their way to the optic nerve, forming a postulated for brain astrocytes (Fig. 2D). However, the close phys-
nest like structure. By contrast, in species like the pig or minipig, ical association between Müller cell processes and retinal vessels
the RGC axons are intercalated between the Müller cells and such still leaves open the possibility that the Müller cell per se is an
axonal packages are not found (Fig. 2). essential component of the blooderetina barrier, and as such, it
Although Müller cells from different species vary considerably may be capable of influencing its own permeability (Sarthy and
in shape, they do have some fairly universal features. At the level of Ripps, 2001).
the outer limiting membrane (OLM), two processes extend in In summary, the Müller cell population accumulates in a dense,
opposite directions and Müller cells extend apical microvilli into regular pattern, although each cells can be considered as the core of a
the sub-retinal space between the inner segments of photoreceptor columnar retinal “microunit” of neurons (Reichenbach and Robinson,
cells. The length and number of microvilli may vary between spe- 1995). This “strategic position” enables Müller cells to constitute an
cies, probably in an inverse relationship to the degree of retinal anatomical and functional link between retinal neurons and the
vascularisation (Uga and Smelser, 1973), and in some species Müller compartments with which these need to exchange molecules, i.e.:
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 5

Fig. 2. Müller cells. A, C and E are from dog retinas, B and D are from pig retinas, and F is from rat. (A) Müller cells are labelled in the retinal pigment epithelium (RPE) with
antibodies against vimentin (red) and RPE65 (green), and the nuclei were stained with DAPI (blue). (B) Retina section stained with antibodies against glutamine synthetase (GS). (C)
High magnification of a section at the level of the retinal ganglion cell layer (RGCL), labelled with antibodies against vimentin (red), showing the Müller endfeet and RGC axons
labelled with antibodies against neurofilaments (NF) in green. (D) High magnification of the Müller endfeet stained with antibodies against GS surrounding a blood vessel (BV) at
the nerve fiber layer level. (E) Müller cells and RPE of a dog retina section stained with antibodies against CRALB. (F) Rat Müller cells in culture marked with green fluorescent
protein (GFP). Scale bar for all pictures ¼ 50 mm.

retinal blood vessels, the vitreous chamber and the sub-retinal space. selectively penetrate Müller cells and have a marked affinity for
Accordingly, Müller glia fulfil many crucial roles, supporting neuronal nucleic acids (Jeon and Masland, 1993; Laties, 1983). In culture,
development, survival and information processing (Reichenbach and Müller cell nuclei can be differentiated, as they are the largest of the
Bringmann, 2010; Sarthy and Ripps, 2001). different retinal cell types.
In terms of immunohistochemical markers, the expression of
2.1.2. Müller cell markers various proteins is restricted to astrocytes and Müller cells, like
Müller cells can be easily identified by their characteristic vimentin and glial fibrillary acidic protein (GFAP), or selectively to
morphology, containing a well-developed endoplasmic reticulum Müller cells, such as the cellular retinaldehyde-binding protein
with varying amounts of glycogen granules (Hogan et al., 1971). A (CRALBP), the glutamate-aspartate transporter (GLAST), glutamine
prominent number of mitochondria are found at the endfeet synthetase (GS), the Kir4.1 subtype of inwardly rectifying potas-
(Fig. 3). Müller cells can be labelled in living retinal preparations by sium channels, aquaporin-4 (AQ4), pyruvate carboxylase, a-crys-
loading for a few minutes with vital dyes, such as Mitotracker or- tallin, the small GTP-binding protein RhoB, glutamate
ange or CellTracker green. Their nuclei can be stained with carboxypeptidase II, carbonic anhydrase C and microtubule-
ethidium bromide, diamidino yellow or chromomycin A3, which associated protein 4 (Berger et al., 1999; Bunt-Milam and Saari,
6 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Fig. 3. Interactions between Müller cells and Retinal Ganglion cells (RGCs). (A) Retina section where RGCs are labelled with antibodies against ßIII Tubulin (green) and Müller cells
are labelled with antibodies against GFAP (red). (B) Electron microscopy detail of mitochondria in the interior of the Müller cell endfeet. (C) Retina section where the RGC are
labelled with antibodies against BDNF (green) and Müller cell membranes with antibodies against p75 (red). Note that RGCs were enveloped by p75-labelled Müller processes. (D)
Electron microscopy detail of RGC (G) wrapped in Müller cell processes (M). Scale bar A), C), D) ¼ 10 mm, B) ¼ 1 mm.

1983; Linser and Moscona, 1981; Moscona et al., 1985; Parysek distributed diffusely in those that are diffusely vascularized and
et al., 1985; Santos-Bredariol et al., 2006). restricted to the vascularised regions of other retinas. This striking
correlation between vascularisation and the presence of astrocytes
2.2. Astrocytes led to the proposal that retinal astrocytes are immigrants that enter
the retina with its vasculature. While astrocytes may reside in
2.2.1. Origin and morphology of astrocytes almost all regions of the retina, regional differences in their density
Named on the basis of their stellate shape, astrocytes are almost arise. Astrocytes are absent in retinal regions with extremely thin
exclusively confined to the innermost retinal layers. The astrocytes nerve fibre layers, such as areas near the ora serrata, and they are
in the retina do not originate from the retinal embryonic epithe- absent from the fovea centralis and the peri-foveolar retina of
lium and thus, it is generally believed that these astrocytes migrate primates. Moreover, the periphery of the retina has a thin nerve
from the optic nerve and probably enter the retina along with the fibre layer and it contains only a few dispersed astrocytes, the
blood vessels (Stone and Dreher, 1987; Turner and Cepko, 1987; number of which increases as the nerve fibre layer thickens. As-
Watanabe and Raff, 1988). In the optic nerve, as in other white trocytes send processes to both blood vessels and axons. Two
matter tracts, most astrocytes and oligodendrocytes are derived groups of astrocytes can be distinguished in the pig and minipig
from local mitoses (Skoff, 1990). The ventricular cavity disappears retinas, those that are directly connected with blood vessels and
in the evagination that forms the retina and optic nerve, and it is those that seem to be connected with RGC axons. We found elon-
believed that the mitotic precursors of the oligodendrocytes and gated astrocytes and star-shaped astrocytes similar to those first
astrocytes migrate into the optic nerve from the sub-ependymal described in primates by Ogden (1978). By contrast, in rat and mice
germinal layer of the brain. According to the extent of myelina- we only found star-shaped astrocytes as described in other species
tion, there would appear to be species differences in the migration (Fig. 5).
of oligodendrocyte precursors. Thus, myelination stops at the The numbers and distribution of astrocytes depends on the
lamina cribosa in the human and rat, while in rabbit and dog density of the nerve fibres that they ensheath, as confirmed by
myelination extends beyond the lamina cribosa to reach the optic confocal microscopy and computer-assisted image reconstruction
nerve head. In some fish, where the lamina cribosa is absent, of astrocytes in the vascularised retina of pigs, rats and mice
myelination extends into the retina within the optic fibre layer (Rungger-Brandle et al., 1993), and in the disposition of astrocytes
(Wolburg, 1981). between species (Fig. 5). For example, in the fully vascularised
The presence and distribution of retinal astrocytes is correlated retina of cat, astrocytes are distributed throughout the inner retina
with the presence and distribution of retina blood vessels (Fig. 4). in a pattern that closely mirrors that of the axon bundles (Karschin
Thus, while avascular retinas contain no astrocytes, they are et al., 1986). The reconstruction of images from retinas stained for
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 7

express GFAP but not connexin-43. Type I astrocytes can be further


divided into type Ia and type Ib, the former present in the optic
nerve head while type Ib astrocytes are present in the ILM of the
retina (Hernandez et al., 2000). Interestingly, both types are
quiescent astrocytes and they only become reactive in response to
certain stimuli, including elevated intraocular pressure, excitotox-
icity and retinal ischemia (Hernandez et al., 2008). When astrocytes
become reactive they hypertrophy and enlarge their soma, they
develop thicker astrocytic processes and their GFAP-
immunoreactivity is enhanced.

2.2.3. Astrocytes main functions


Despite our increased understanding of the function of inter-
mediate filaments (IFs) in astroglial cells (for a review, see Pekny,
2001), the role of astroglial IFs in maintaining the mechanical
integrity of these cells and that of the CNS tissue remains unclear.
As the main producers of VEGF (vascular endothelial growth factor)
during both normal (Stone et al., 1995) and pathological (Ozaki
et al., 2000) vessel formation, astroglial cells are strongly impli-
cated in retinal vascularisation. However, the vascular system de-
velops normally in GFAPe/e;Vime/e mice, and VEGF RNA is
expressed at normal levels in both ischemic and control GFAPe/
e
;Vime/e retinas, indicating that the absence of IFs in astroglial cells
does not affect their role in retinal vascularisation, either in healthy
or pathological conditions. The absence of IFs in astrocytes does not
impede the production of chondroitin sulphate proteoglycans, yet
it might delay the onset of astrocyte maturation and prolong the
window for optic nerve regeneration in GFAPe/e; Vime/e mice.
These data support the hypothesis that mature astrocytes represent
a key barrier to CNS axon regeneration in vivo (Cho et al., 2005) and
thus, astrocytes may play dual roles: one beneficial and another
degenerative.
Some of the beneficial roles fulfilled by astrocytes include
Fig. 4. Astrocytes. (A) Minipig retinal section. RGCs are labelled with antibodies
neurotrophic support, enhanced mechanical support for degener-
against ßIII Tubulin (green), Müller cells and astrocytes are labelled with antibodies
against GFAP (red) and the nuclei are stained with DAPI. Note that the astrocytes follow ating axons, and the maintenance of the blooderetina barrier. In
the blood vessels from the RGC layer to the outer nuclear layer (ONL). (B) Confocal response to injury or disease, the same astrocytes express a number
image of a rat whole-mount rat retina where astrocytes were labelled with antibodies of proteins that compromise the integrity of blooderetina barrier,
against GFAP (green) and aquaporin 4 (AQ4, red) (C) Rat glia primary culture. Astro-
up-regulating the expression of various genes encoding cytokines,
cytes are indicated with an arrowhead, while the other cells are Müller cells with a
larger nuclei: vimentin labelling (green) and nuclei are stained with DAPI. Scale
chemokines and elements of the complement cascade, and pro-
bar ¼ 50 mm. moting retinal degeneration (Anderson et al., 1967; Kim et al.,
2006). Reactive astrocytes in the retina express increased levels
of MMP-9 and uPA, and accordingly, they degrade the extracellular
GFAP (astrocytes) allowed us to visualize the ensheathing of blood matrix associated with the ILM and they promote RGC death via
vessels (Figs. 4B and 5A), whereby the vitreal and lateral sides of detachment-induced apoptosis (Zhang et al., 2004). In the different
most blood vessels are in close apposition to the vitreal surface of sections that follow, we will refer to many other functions of as-
the retina. In addition, individual astrocytes can be seen to extend trocytes in more detail.
their processes into the vessel wall in such confocal images, espe-
cially in pig retinas, and while in some instances astrocytic pro- 2.3. Microglial cells
cesses extend to the vitreo-retinal surface, in other cases astrocytes
remain parallel to RGC axons. It is noteworthy that astrocytes Experimental evidence has confirmed the so called immune
communicate by gap junctions, and that both vitreous and blood privilege of the CNS on the basis of the existence of the bloodebrain
vessels can serve as a “sink” for Kþ (Newman et al., 1984). barrier (BBB), the absence of lymph or the lack of locally competent
macrophages. However, more recent research supports the notion
2.2.2. Astrocyte markers that the CNS establishes robust interactions with the immune
Astrocytic process envelops blood vessels as seen in electron system and that the microglia fulfils a key role in controlling im-
microscopy, with apparently no perivascular space between the mune responses (Carson et al., 2006). The CNS reacts through im-
adventitia of the vessel wall and the glial cells (Hogan and Feeney, mune responses that are both innate and adaptive, and that are
1963; Hogan et al., 1971). GFAP is the major constituent of the mediated by distinct populations of immune competent cells.
astrocytic intermediate filament (Bignami et al., 1972), and it rep- These cells may or may not share their embryonic origin but they
resents a reliable marker to identify and localise astrocytes in the communicate through direct contact and paracrine interactions.
mammalian retina. Under normal circumstances macroglia and Microglia, astrocytes, infiltrating T-cells and macrophages form a
mammalian Müller cells contain significant amounts of GFAP and complex and pleiotropic defence network that secures and protects
thus, astrocytes must be classified as one of two distinct types in the integrity of the CNS (Colton, 2009; Corraliza, 2014; Prinz and
the retina by attending to other molecular markers: type I astro- Priller, 2014; Ransohoff and Brown, 2012).
cytes express GFAP and connexin-43. Whereas type II astrocytes Macrophages are a morphologically and functionally diverse
8 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Fig. 5. Whole mount retinal explants. A, and C are control retinas. B and D are explants. A and B are pig retinas. C and D are mouse retinas. The astrocytes are labelled with
antibodies against GFAP (red in pig, A and B, and green in mouse, C and D). Note the changes in the astrocyte morphology between control and explant retinas. Scale bar ¼ 50 mm.

group of cells as a consequence of their embryonic origin, cellular microstructured supports in vitro, these cells demonstrate a high
differentiation, tissue localization, and exposure to both internal or degree of plasticity in distinct extracellular milieus (Amadio et al.,
external stimuli (Gordon, 2003; Gordon and Martinez, 2010; Prinz 2013).
and Priller, 2014). The macrophage family includes meningeal and
perivascular macrophages, blood circulating macrophages, den- 2.3.1. Origin, morphology and distribution of microglia
dritic cells derived from the bone marrow, and tissue resident Several cell pools of myeloid origin and with common molecular
macrophages (Wieghofer et al., 2014). The microglia form a popu- features populate the CNS, although they also exhibit important
lation of resident macrophages within the CNS that not only has differences relating to their origin and activity. Experiments in mice
immune activity but also, that participates in the development and confirmed that microglia originate in the yolk sac at an early stage
maintenance of neural networks, as well as in tissue homeostasis of embryonic development, between 8.5 and 9.5 days post-
(Fernandes et al., 2014; Gertig and Hanisch, 2014; Ransohoff and fertilisation (dpf) (Ginhoux et al., 2010; Schulz et al., 2012). Spe-
Brown, 2012). Although the origin of these cells is controversial, cifically, they derive from erythromyeloid precursors (Kierdorf
it is believed that primitive myeloid progenitors of the egg sac are et al., 2013). These stem cells undergo a sequential development
the precursors of these cells and not the bone marrow (Ginhoux from stage A1 when they express the CD45 antigen to stage A2,
et al., 2010; Schulz et al., 2012), and during adulthood a popula- when they synthesize myeloid markers that include CX3CR1
tion of local progenitors ensures the continuity of the microglia in (chemokine receptor 1) and CSF1R (colony-stimulating factor 1
the nervous system (Ajami et al., 2007). Microglia have a quite receptor). Stage A2 cells migrate to the brain parenchyma with the
specific morphology and physiology, very different from their aid of specific matrix metalloproteinases (MMPs), where they
precursors, and they adapt to their neural environment to fulfil establish their final destination (Kierdorf et al., 2013; Prinz and
their designated roles (Kettenmann et al., 2013). Priller, 2014; Prinz et al., 2011). By contrast, other macrophage
As occurs with many advances in science (over time and populations that inhabit the CNS have different myeloid origins,
through persistence), the morphological and functional definition which include the fetal liver, the aortaegonademesonephros re-
of microglia has gradually been sculpted (Ginhoux et al., 2013). The gion and the bone marrow (Kierdorf et al., 2013; Prinz and Priller,
term “mesoglia” was proposed by Robertson (Ginhoux et al., 2013; 2014).
Kettenmann et al., 2013). Del Rio-Hortega in a conference delivered Microglia colonise CNS structures during embryogenesis and in
in 1938 at the University of Oxford (UK), defined microglia as the the early postnatal period until they reach their positions in the
non-neuronal and non-astrocytic component of the “third mature brain (Harry, 2013). Colonization occurs along different
element”, of mesodermal origin, that can change its phenotype, migratory pathways, following glial structures, white matter net-
migrate and proliferate. He also separated microglia from the works or blood vessels (Rezaie and Male, 1999). In the CNS,
oligodendroglia (Del Rio-Hortega, 1939), a characterization that can microglia are distributed widely in the gray and white matter, and
still be applied today. Their plastic nature makes defining the they present local singularities related to their function (Harry,
functional phenotypes of microglia difficult. However, it seems 2013). The diverse morphologies of the microglia are related to
clear that microglia can exist in different states of activation and their activation states or functional phenotypes (Kettenmann et al.,
“quiescence”. Indeed, there seems to be a spectrum of intermediate 2011). Briefly, in the so called “resting” state they have a ramified
states or functional phenotypes that may explain their complex appearance with a small cell body and rather large nucleus. In the
physiology (Cherry et al., 2014). For instance, when using artificial “activated state” the cell morphology is more ameboid and they
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 9

exhibit pseudopodia (Davis et al., 1994). In healthy conditions his early studies, Del Rio Hortega observed these two different
microglia are found at variable densities in the different retinal morphologies and he related them to the primitive invasion of
layers, including the NFL, ganglion cell layer (GCL), inner plexiform microglia (amoeboid) and the pool of settled cells in the mature
layer (IPL), inner nuclear layer (INL) and OPL (Chen et al., 2014; brain (branched and ramified: (Del Rio-Hortega, 1939). However,
Cuenca et al., 2014; Garcia-Valenzuela et al., 2005; Noailles et al., although these two states serve as a reference point, the adaptable
2014; Santiago et al., 2014; Sobrado-Calvo et al., 2007). The cells and fine-tuned responses of microglia reinforce the existence of
have a small ovoid-shaped cell body and multiple ramifications in several intermediate and fully functional phenotypes (Cherry et al.,
the GCL of the normal retina, and they are organized into a single 2014). One basic response of microglia to local insults depends on
sheet that is distributed homogeneously (Garcia-Valenzuela et al., the presence of extracellular ATP and it involves the specific
2005). In the optic nerve, microglia ramifications appear to be extension of microglial processes to the injured site (Davalos et al.,
orientated along the fibres. This orientation is disorganized in an 2005). This morphologicalephysiological tour has been called the
experimental model of diabetes (Chen et al., 2014), and the distri- “spider effect” (the spider in the middle of the web responds to
bution and organization of microglia within the retinal layers is vibration caused by the “intruder” and moves to the site where the
altered in a variety of pathological states, when changes in cell prey is trapped, thereafter returning to its “vigilant” site in the
shape can be observed (Cuenca et al., 2014; Kezic et al., 2013). middle of the web: (Jonas et al., 2012). The concept of “resting” state
Moreover, when RGCs are labelled with a retrograde vital fluores- implies an “inactive” condition and it may lead to confusion, since it
cent dye in an experimental axotomy, microglia containing RGC suggests a passive behaviour that would only be converted to an
debris can be detected months after the injection of marker, indi- “active”, functional state under certain circumstances (for example,
cating both the phagocytotic capacity of microglia and their relative infection, malignancy, trauma, neurodegeneration, etc…). Experi-
longevity (Thanos et al., 1994). mental studies support the idea that even in healthy and homeo-
static conditions, microglia are responsive to changes in the
environment, and that they modulate numerous events by
2.3.2. Plasticity and functional microglia phenotypes secreting cytokines and other factors, and by interacting with other
Microglia fulfil specialized immune tasks, responding to specific glial cells and neurons (Kettenmann et al., 2011; Wieghofer et al.,
stimuli by secreting cytokines and neurotrophic factors. In addition, 2014). Therefore, applying the term “surveying” instead of
they establish a close relationship with neurons and other cells so “resting” to vigilant microglia seems to be more appropriate
that they contribute to the remodelling of neuronal circuits, as well (Santiago et al., 2014) (Fig. 6).
as to the elimination of cell debris and deranged synapses. Micro-
glia also play a role in pathological events that affect the CNS
(Kettenmann et al., 2013; Prinz et al., 2011; Wieghofer et al., 2014). 2.3.3. Representative molecular traits of microglia
Microglia exist in two separate morphological/physiological states, Microglia produce many substances and they also have receptor
namely “resting” and “activated” microglia (see for instance Colton, systems that process chemical signals from neighbouring or distant
2009; Kettenmann et al., 2011, 2013; Ransohoff and Perry, 2009). In cells. The complex physiology of microglia relies on their secretory

Fig. 6. Microglia. (A) Pig retina section where the microglia were labelled with antibodies against Iba-1 (red), astrocytes and Müller cells (green) were labelled with antibodies
against GFAP and the nuclei were stained with DAPI. (B) Amplification of A, where the filled arrowhead indicates a microglia cell around an astrocyte or Müller cell. (C) Amplification
of A where the empty arrowhead indicates the interaction between a microglial cell and a blood vessel. (D) Rat microglia primary culture where the mitochondria were labelled
with mitotrack (red) and the nuclei are stained with DAPI. Scale bar ¼ 50 mm.
10 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

capacity, and on the existence of multiple receptor systems for: molecule 1- and ED-1 -monoclonal antibody against glycoprotein
neurotransmitters and neuromodulators (e.g., dopamine, GABA, CD68) can be used simultaneously to differentiate microglia from
glutamate, epinephrine/norepinephrine, cannabinoids, histamine, other cell-related populations, including migrating macrophages
angiotensin II, opioids, bradykinin, neurotrophins, acetylcholine, (Ellis-Behnke et al., 2013). Since all types of macrophages and
etc…); cytokines and chemokines, including TNF-a (tumour ne- microglia express the F4/80 protein (a surface receptor of the EGF
crosis factor) or interleukins; complement; epidermal growth fac- family), this marker alone is not suitable to target microglia. His-
tor (EGF); Notch; macrophage colony-stimulating factor; peptides; tochemical detection also offers advantages in some cases. For
leukotrienes…etc. Microglia also express different transporter instance, one histochemical method uses a tomato lectin with af-
systems for solutes and ions (Kettenmann et al., 2011). finity for sugar moieties present on microglial and endothelial cells,
Microglia communicate with other glial cells and neurons by allowing the relationship of microglia with blood vessels to be
secreting factors that establish specific secretory signatures char- analysed (Villacampa et al., 2013). Microglia in different activation
acteristic of distinct functional or pathological states. The mole- states can be characterized using combinations of techniques
cules secreted by microglia and other glial cells (both proteinaceous aimed at analyzing the secretory, proteomic and transcriptional
and non-proteinaceous) are defined as their secretome, and signatures (Glanzer et al., 2007): microarrays, real-time RT-PCR
establishing its composition may be essential to discover bio- (reverse transcription polymerase chain reaction), SELDI-TOF (sur-
markers useful for diagnosis and prognosis, as well as in the search face-enhanced laser desorption ionization-time of flight), two-
for new therapies to alleviate different CNS pathologies. Here we dimensional gel electrophoresis (2D-PAGE), mass spectrometry,
name just a few of the molecules that microglia are known to quantitative western blotting…etc.
secrete, either constitutively or in response to inflammation, injury, Genetic manipulation serves to target microglia and has pro-
CNS development and degeneration or homeostasis: cytokines and vided relevant tools to distinguish these cells from other myeloid
chemokines, ROS (reactive oxygen species), nitric oxide (NO), cells. Different approaches are available and include labelling
phospholipase A2 IIA type, BDNF (brain-derived neurotrophic fac- methods (where fluorescent reporter genes like GFP -green fluo-
tor), GDNF (glial cell line-derived neurotrophic factor), IGF (insulin- rescent protein- or LacZ -beta-galactosidase- may help to monitor
like growth factor), FGF-2 (fibroblast growth factor 2) or LC1 (lip- the spatio-temporal expression of a given protein), as well as the
ocortin 1) (Jha et al., 2013). Cell contacts and soluble factors are key manipulation (activation and deletion) of genes that produce stable
elements to maintain the functional dialogue established between transgenic animal lines (see Wieghofer et al., 2014 for a review). On
the different cell types that populate the retina and they are char- occasions it is difficult to separate macrophage populations and to
acteristic of a given activation state of microglia (Langmann, 2007). assign the specific involvement of each class to particular functions.
A representative example of a neuronemicroglia interaction is the In a knock-in mouse model, where CNS microglia were labelled
FKN/CX3CR1 system (fractalkine or CX3CL1 and its G protein- with a GFP-tagged CX3CR1 and blood originated monocytes were
coupled receptor, CX3CR1). Fractalkine is a chemokine that may labelled with a RFP (red fluorescent protein)-tagged CCR2, the ge-
exist as a membrane-bound protein or in a soluble form, and it is netic markers allowed “red” infiltrating monocytes and “green”
present in neurons that bind to CX3CR1 expressed on microglia. resident macrophages to be visualized (Saederup et al., 2010). This
The CX3CL1 ligand modulates the activation of microglia in distinct method has also been applied to localize microglial populations
manners during homeostasis and in disease states (Sheridan and during development and at their final sites (Mizutani et al., 2012).
Murphy, 2013). In the retina, the action of CX3CL1 depends on Based on the self-renewal capacity and the long lifespan of
the conversion of the membrane-bound and soluble isoforms, and microglia, a genetic system has been developed where the
it participates in both neurogenesis and the response to cues that expression of Cre recombinase (an enzyme that catalyzes site-
trigger neurodegeneration (Zieger et al., 2014). Microglia also ex- specific DNA recombination events) is controlled by the promoter
press adenosine receptors that regulate their activities both in of a gene that encodes Cx3cr1 (Goldmann et al., 2013). This
healthy and disease states (Santiago et al., 2014). Accordingly, experimental system allowed an in vivo animal model to be elab-
microglial receptor proteins appear as important targets for the orated (Cx3cr1Cre mouse) in which conditional depletion of a
development of drugs to prevent or manage retinal neuro- mitogen-associated protein kinase kinase kinase (MAPK3), TAK1
inflammation. The secretory capacity of microglia depends on non- (transforming growth factor b-activated kinase 1), demonstrated
vesicular (transporters or ion channels) and vesicular mechanisms, the direct participation of microglia in autoimmune inflammation.
such as exocytosis and the release of extracellular membrane This model opens new avenues to ascertain the role of different
microvesicles (EMVs) loaded with bioactive molecules (Prada et al., genes expressed in microglia at specific time points or under spe-
2013). cific circumstances (physiological or pathological). Since many
diverse techniques are available, it is important to be aware of the
2.3.4. Identification of microglia potential and the limitations of each of these when designing ex-
Due to the important role of microglia, both in homeostasis and periments to study the presence, morphology and function of
disease, it is of particular interest to develop tools to characterize microglia.
and identify these cells, as well as the molecular dialogue they
undertake in their multiple modulatory activities. This task is
difficult, particularly since this group of cells share molecular
identities with related cells of hematopoietic origin (Wieghofer
et al., 2014). The identification of cells is based on their localiza- 3. Common functions of different glial cells in the retina
tion and morphology, as well as on the specific expression of sur-
face proteins and the functions associated to them. The The three glial cell subtypes present in the retina share many
methodological approaches may involve the use of specific anti- properties that while possibly redundant they have a preferential
bodies to localize protein epitopes (immunocytochemistry, western hierarchy in which they are performed. There are certain functions
blotting…) or the manipulation of gene expression. The identity of of retina glial cells that are more predominant in one cell type or
microglia can be assessed by using different protein markers (for another. Thus, in the following sections we will review the most
instance, IBA-1, ED-1, HLA-DR, CD-45…etc). In most cases, the recent advances related to this important issue and the role of these
antibody markers (such as IBA-1 -ionized calcium binding adaptor glial cells in maintaining retinal homeostasis.
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 11

3.1. Formation of boundaries Müller glial cells does not vary remarkably in the different
mammalian species studied so far (Lu et al., 2013), and their
Müller cells share many properties with astrocytes, including viscoelastic properties appear to be similar in man, monkey and rat.
the ability to ensheath axons, and to form glial boundaries between In general, their elastic behaviour dominates over their viscous
the retina and the blood vessels (Hollander et al., 1991). However, behaviour.
while Müller cells are formed in the retina, astrocytes migrate Müller cells act as mechanosensors, as they span vertically from
independently of vessels and they are therefore in a position to the inner to the outer retina. The Müller cells form stiff endfeet that
respond to environmental signals in avascular areas of the retina. are rich in mechanosensitive ion channels, indicating that the
Their restriction to the inner layer of retina enables them to biomechanical responsive element of the retina is located in this
respond to hypoxia in this region by expressing VEGF, thereby region. In addition to the different types of cells, the mechanical
inducing the formation of a superficial layer of blood vessels. By properties of soft tissues like the retina mainly depend on extra-
contrast, Müller cells extend across all retina layers and cannot cellular matrix components like collagen, elastin and pro-
respond in a layer-specific manner. The migration of astrocytes teoglycans. One of the main functions of proteoglycans is to retain
across the surface of the retina therefore provides a population of water, which constitutes >70% of the retina. The presence of water
macroglial cells that is strategically located to sense hypoxia in the in tissues causes collagen fibres to separate, swell and become
inner layers of retina. disorganized, which significantly affects their relative stiffness and
In addition to the glia limitans of the retina and vessels, astro- tensile strength (Chen et al., 2009a).
cytes and Müller cells both contribute to the glial sheaths around The biomechanical scaffolding of the retina, as well as its
neurons. In particular, both astrocytes and Müller cells wrap gan- biochemical homeostasis, is maintained by Müller cells and astro-
glion cell axons into bundles, contributing to the glial convergence cytes. These cells can alter their elasticity and stiffness by up- and
on the initial segments and node-like axon structures, and they also downregulating their IF proteins, the most common of which are
ensheath the somas of neurons in the RGC layer. Furthermore, GFAP and vimentin, thereby altering both the tissue-wide and
adherens junctions form between astrocytes, between Müller cells, cellular biomechanics of the retina (Bringmann et al., 2009a;
and between astrocytes and Müller cells, although not between Lindqvist et al., 2010; Lu et al., 2011). In mice lacking GFAP and
these cells and neurons or among neurons. These similarities point vimentin, mechanical challenge revealed an exaggerated fragility of
to the possibility that astrocytes and Müller cells might act inter- Müller cells (Lundkvis et al., 2004). Müller cells are mechanically
changeably in some processes, and we propose that they be heterogeneous in the lateral direction, although they seem to
regarded as variants of macroglia. Quantitative differences between display a mechanical homogeneity in the vertical direction. Due to
astrocytes and Müller cells were noted in their ensheathment of the distinct cytoskeletal organization in local regions of a Müller
neurons. Specifically, the glial sheaths around the somas of RGCs cell, the regions closer to the endfeet, soma and processes can be
are predominantly formed by Müller cells (Fig. 3A,C), while the glial discerned by their mechanical properties (Lu et al., 2006). This local
processes attached to node-like specialisations of their axons are variation in the mechanical compliance of a single Müller cell is
mainly formed by astrocytes (Fig. 5A: Stone et al., 1995). In dog likely to determine the mechanical diversity in the retina (Park and
retinas in particular we found that axons form packages that Lee, 2013). Mechanosensory Ca2þ ion channels, such as TRPV4, are
interdigitate between the endfeet of the Müller cells in way similar known to be present on Müller cells, particularly on the Müller cell
to the nests building (Fig. 2C). endfeet at the inner retinal border. Thus, these channels may also
One qualitative difference has been that in two glial cell classes provide the retina with a sensor of biomechanical changes
the gap junctions formed between astrocytes do not form between (Ryskamp et al., 2011).
Müller cells or between these two different types of cells. Thus, the In normal situations, in addition to morphological alterations
roles of the two types of glial cells, in terms of the information and due to glutamatergic neurotransmission, there are traction forces
communication related to the cells that are ensheathed could vary. on Müller cells in the retina derived from the vitreous body. In the
Moreover, reactive astrocytes form a border between the lesion and human retina, the vitreous body adheres to the retinal tissue at the
the surrounding tissues, and as resting microglia are located in peripheral retina through the major superficial retinal vessels, optic
certain areas, close to where the retina is vascularised, they form an disc, perifovea and macula. At sites of vitreo-retinal attachments,
imaginary fence that limits the penetration of any invader into the the basement membrane of the ILM can become thinner, and vit-
retina. reous fibres can adhere directly to Müller cells (Schubert, 1989).
However, there is limited information regarding the biomechanical
3.2. Glia as a structural support for neurons properties of these cells (and of the retinal tissue). Experimental
approaches are limited as it is unclear whether retinal cells and the
3.2.1. Viscoelastic properties tissue of laboratory animals suitably reflect human conditions (Lu
The biomechanical changes in the retina during its development et al., 2013).
and in disease are very important, and they could be the underlying Astrocytes are abundant in the CNS, and their role as passive
cause of several conditions. During development and after the optic nurturing cells has been broadened to a more dynamic function in
vesicle has closed, the intraocular pressure and proliferation of terms of their relationship with other glial cells and neurons.
neuroblasts causes the primitive retina to stretch. After birth, the Because of the specialized nature of astrocytes, their morphology
retina continues to expand tangentially, stretching out from the and their location in the retina, one might envision that astrocytes
centre to periphery until the eye reaches a maximum size. During in the retina may contain mechanical sensors. Accordingly, astro-
this growth period the components of the retina must be elastic. cytes might sense eye movements and other perturbations of the
The retinal neurons are significantly stiffer than the glial cells, and retina, for example an elevation of the intraocular pressure. Little is
elasticity decreases from the inner to the outer retina. Moreover, known about the indicators of the mechanical state, however it has
the two limiting membranes of the retina are stiffer that the rest of been proposed that astrocytes at the optic nerve head could induce
the layers. While the Müller cell endfeet, nerve fibre bundles and astrocytic signalling by releasing ATP (Barres and Dowling, 2010). If
GC somata are stiff structures, the least elastic part of the retina is this was to occur, ATP release could potentially stimulate microglial
the ONL as well as the inner and outer photoreceptor segments cells to migrate to the nerve head, activating astrocytes to become
(Franze et al., 2011; Lu et al., 2006). The elastic stiffness of retinal reactive and influence neurons or blood flow (Barres and Dowling,
12 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Fig. 7. Changes in the distribution of BDNF and p75 receptors in control rat retinas and following ischemia. (A, C, E, G) Retinal sections where RGCs were labelled with antibodies
against BDNF. (B, D, F, H) Müller cells labelled with antibodies against p75 receptor. (A, B, E, F) are control retinas and (C, D, G, H) are the contralateral retinas where the ischemic
insult was maintained for 1 h, followed by 2 (C and D) or 24 h of reperfusion (G and H). Note the increase in immunofluorescence staining within the ganglion cell layer of C
compared to the control (white arrows) A, and the decrease in labelling in G as compared to E. Note the loss of the layering of Müller cells stained with antibodies against p75 in H
compared to F. Scale bar ¼ 50 mm.

2010). Astrocytes respond to mechanical forces by triggering transmitters from neurons (Hossmann, 1993). The plastic changes
reactive gliosis, which might perturb their vasodilatory role by of the Müller cells take place over a relative short timescale (Fig. 7).
disrupting their vessel contacts and interfering with their gap The three distinct layers of immunoreactivity originating from the
junction coupling. Alternatively, this response could perturb the extensions of the Müller cells, that are present in the IPL of control
ionic homeostasis, since astrocytes are also vulnerable to neuronal retinas, was no longer evident, rather, it appeared as a single thick
death (Cahoy et al., 2008; Foo et al., 2011). layer after ischemia (Fig. 7F, H).
The communication between Müller cells and RGCs remains
unknown, although we did observe changes to Müller cells in ret-
3.2.2. Architecture-mechanical forces inas from animals with experimental glaucoma or axotomy
The axons within the retina of all species are organized radially (Hernandez et al., 2009).
in order to converge at the optic nerve head, yet there are impor-
tant differences in the way that astrocytes are distributed within
the retina. While in the pig, minipig and human, astrocytes are 3.3. Volume regulation
organized in two levels one more vitreal and another more internal.
The more vitreal astrocytes have a parallel disposition that can be The regulation of cell volume is of great importance to avoid
observed in flat mount retinas (Fig. 5A), while in rat and mice, the changes in neuronal excitability resulting from a decrease in the
distribution of astrocytes is more stellate appearance and their volume of the extracellular space. Cell swelling and subsequent
extensions do not run parallel to the optic nerve (Figs. 4B and 5C). Regulatory Volume Decrease (RVD) is accompanied by trans-
Astrocytes can modify their distribution in response to local membrane potential (Vm) depolarization followed by repolariza-
changes and thus, after axotomy and explant culture the pattern of tion. However, this RVD response depends closely on the
astrocytes observed in flat mount retinas is lost within a few days. It composition of the extracellular media. Although Kþ and Cl
is possible that astrocytes sense the damage to the axons and thus, channels play an important role in the RVD response of Müller cells,
they tend to lose their contacts and redistribute within the retina. their contribution is evident only if a significant driving force for
In mice and rats, astrocytes became hypertrophic and assume a KCl efflux is present (Ferna ndez et al., 2013). The changes in volume
stellate shape, as observed in glaucoma (Fig. 5). of glial cells are also associated to the viscoelastic characteristics of
Müller cells also sense the changes that take place in the inner the retina.
and outer retina (see below, “glaucoma and retinitis”). Müller cells It has been suggested that Müller cells possess endogenous cell
modulate neuronal activity by regulating the concentration of volume-regulatory mechanisms that are involved in the activity-
neuroactive substances and potassium in the extracellular space dependent regulation of the volume of the extracellular space
through a variety of transport processes (Newman and (Vogler et al., 2013). In addition to Kir (inwardly rectifying potas-
Reichenbach, 1996). This activity of Müller cells is particularly sium) channel mediated cell volume regulation, Müller cells
important in ischemia when there is a massive release of possess receptor-mediated mechanisms that prevent hypoosmotic
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 13

swelling (Uckermann et al., 2006; Wurm et al., 2008a). Thus, rates of proliferation (for instance, embryonic stem cells and
pharmacological blockade of Kir channels causes swelling in brain tumour cells: Ng et al., 2014). Glucose, obtained from external cir-
astroglial somata and processes, yet it has no effect on Müller cells. culation, from internal cellular glycogen stores or from lactate is the
The blockade of Kir channels in hypoosmotic conditions induces primary metabolite used by retinal cells to produce energy, both in
immediate swelling of Müller cell somata but no additional aerobic and anaerobic conditions (Reichenbach and Bringmann,
swelling in somata of brain astrocytes (Hirrlinger et al., 2008). 2010).
Water accumulation in retinal neurons and glial cells is a patho- Retinal cells also display specialised metabolic outcomes and
genic factor involved in retinal degeneration under ischemic- they have adapted to better serve cell house-keeping needs and
hypoxic and inflammatory conditions. Nerve growth factor (NGF) specific demands linked to processing visual inputs (photo-
prevents the osmotic swelling of Müller cells by inducing auto- transduction and neurotransmission). Within the retina, glial cells
crine/paracrine FGF signalling, as well as indirectly controlling the and neurons orchestrate complex metabolic relationships
swelling of bipolar cells by inducing a release of cytokines from (providing nutrients and metabolites, responding to signals from
Müller cells (Garcia et al., 2014). neighbouring or distant cells, or adapting the transcriptional ma-
Müller cells of the rodent retina possess a volume-regulatory chinery to supply the required metabolites) that fine tune their
glutamatergicepurinergic signal transduction cascade. Glutamate metabolic activity to their needs. For example, Müller cells resist
induces the swelling of inner retinal neurons and shrinkage of the hypoxia and low glucose environments by activating anaerobic
extracellular space. Retinal glial cells rapidly and significantly glycolysis and by oxidizing alternative substrates, such as lactate,
change their shape when glutamate is applied to retinal whole glutamate or glutamine in order to obtain energy in the form of ATP
mounts, and this effect is induced by glutamate-evoked neuronal (Kitano et al., 1996; Reichenbach and Bringmann, 2010; Stone et al.,
cell swelling (Uckermann et al., 2004; Wurm et al., 2008b). Müller 1999; Winkler et al., 2000). Compared with other tissues, the retina
cells adapt their morphology to the increased size of activated consumes more ATP (Hurley et al., 2014) and it undergoes an
retinal neurons by extending and thinning the inner stem process. elevated proportion of aerobic glycolysis, the so-called Warburg
This response is due to the opening of potassium and chloride effect (glucose is preferentially converted into lactate, instead of
channels, whereby the ion efflux compensates for the osmotic carbon dioxide and water by oxidative phosphorylation, despite the
gradient across the Müller cell membrane and prevents cellular presence of oxygen). The Warburg effect favours rapid ATP pro-
swelling (Uckermann et al., 2006). The glial water channel duction when glucose is converted into pyruvate and then fer-
aquaporin-4 (AQP4) is implicated in the control of ion and osmo- mented into lactate to regenerate oxidized NAD (nicotine adenine
homeostasis in the retina. The water flux through AQP4 is dinucleotide), and it may procure the energy needed for continuous
involved in the rapid volume regulation of retinal Müller cells in rhodopsin production or for other processes (Casson et al., 2013; Ng
response to osmotic stress and deletion of AQP4 results in an in- et al., 2014).
flammatory response of the retinal tissue (Pannicke et al., 2010). A few examples confirm that retinal cells configure a metabolic
Müller cells can maintain the integrity of the blooderetina dialogue based on the expression and distribution of enzymes and
barrier and clear metabolic waste by regulating the homeostasis of metabolic carriers. Retinal cells also exhibit different energy
extracellular pH and Kþ ions. To avoid high Kþ levels that might metabolic profiles (for instance, glia, RGCs, photoreceptors) and a
induce depolarization of retinal neurons, Müller cells take up single cell may have different energy demands in different com-
excess Kþ from the extracellular space (especially in the synaptic partments (as is the case for RGCs) while still maintaining the in-
layers of the retina), and release it into the blood and vitreous ternal homeostasis (Yu et al., 2013). Glucose and lactate are not the
humour (Reichenbach and Bringmann, 2013; Reichenbach et al., only primary metabolites catabolised to attend to the energy and
2007). Many of the homeostatic functions of Müller cells, vegetative requirements of the retina. Retinal pigment epithelium
including spatial potassium buffering and neurotransmitter uptake, (RPE) cells metabolise fatty acids from phagocytised photoreceptor
depend on the negative membrane potential (around 80 mV) outer segment membranes by beoxidation. Also, RPE cells produce
established by the ample expression of Kir4.1 in these cells, which b-hydroxybutyrate from fatty acids (a ketone body) that is further
can regulate and compensate the osmotic conditions (Bringmann incorporated into the TCA (tricarboxylic acid cycle) by retinal cells
et al., 2006; Kofuji et al., 2000). Kir 2.1 and Kir 4.1 channels were to obtain ATP, amino acids or other metabolites (due to the
recently studied under hypoxic conditions, which produced no amphibolic properties of TCA). Müller and microglial cells also
significant changes in their mRNA expression (Kang et al., 2013). establish important metabolic relationships, both in functional and
Adenosine might upregulate the expression of Kir 4.1 channels diseased retinas. In physiological conditions, Müller cells release
while it only weakly affects the expression of Kir 2.1 channels ATP to the extracellular space that is used by microglia to fuel their
in vitro. Such upregulation of Kir 4.1 channels should accelerate motility (Wang and Wong, 2014). Accordingly, an effective oxida-
retinal Kþ clearance to prevent neuronal hyperexcitation and tive and biosynthetic dialogue is established between retinal cell
excessive release of glutamate (Yu et al., 2012). components (Adijanto et al., 2014).
AQP4 are present in the plasma membranes of astrocytes that Among the well-defined house-keeping activities of astrocytes,
are in contact with micro vessels (Fig. 4B), allowing them to regu- they provide neurons with essential metabolites (e.g., glucose), and
late the extracellular space volume, waste clearance, potassium they take up Kþ and neurotransmitters from the extracellular space.
buffering, cell migration and calcium signalling (Nagelhus and Although it was suggested that astrocytes can convert glycogen to
Ottersen, 2013), playing an important role in water homeostasis glucose, it is not known under what conditions this occurs in vivo.
in the retina, moreover Müller cells play also a role in water Although glucose production via the G6Pase complex in the
regulation. endoplasmic reticulum is theoretically possible, other fates of G6P
in the cell include lactate-energy production via glycolysis and
3.4. Metabolism entry into the pentose phosphate pathway. Also, there is contro-
versial evidence that confers a glucose export capacity on astro-
3.4.1. Glucose cytes per se. The metabolic relationship between different retinal
The retina exhibits metabolic peculiarities that are related to its cell populations includes many molecules and reactions (Fig. 8),
activity. Some specific characteristics of retinal oxidative and and their detailed description is beyond the scope of this work. The
anabolic metabolism are also found in tissues that manifest high reader is referred to some recent works that deal with this matter in
14 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Fig. 8. Diagram summarizing the main interactions of glial cells with neurons. Scheme showing glucose metabolism, Kþ homeostasis, H2O, glutamate (Glu) to glutamine (Gln)
metabolism, the secretion of trophic factors and interleukins: A (astrocytes, in green), B (bipolar cells), G (ganglion cells), M (Müller cells, in blue), Mi (microglia, in red), Ph
(photoreceptors).

more detail (Hosoya and Tachikawa, 2012; Hurley et al., 2014; Ng regular exposure to photooxidative damage. As a precursor for the
et al., 2014; Reichenbach and Bringmann, 2010; Sanderson et al., neuroprotectin, NPD1, DHA helps to upregulate anti-apoptotic
2014; Yu et al., 2013). genes of the Bcl-2 family, while downregulating pro-apoptotic
and pro-inflammatory genes (Bazan, 2007). Müller cells take up
DHA (Gordon and Bazan, 1990), incorporate it into glial phospho-
3.4.2. Lipid metabolism
lipids and channel it towards the photoreceptors (Politi et al., 2001).
Anabolic activities in photoreceptors are primarily directed to-
Müller cells also express fatty acid-binding and -transfer proteins
ward the synthesis of outer segment discs, a highly membranous
for the transport of DHA and other fatty acids (Deguchi et al., 1992;
compartment of the photoreceptor and the site of the photo-
Kingma et al., 1998). A potential mechanism for the anti-
transduction cascade. The retina has the capacity to both synthesize
inflammatory effects of DHA in diabetic retinopathy was recently
cholesterol de novo and to take up lipids from the blood (Fliesler
suggested, whereby microglia recruit receptors into lipid rafts
and Keller, 1997). Circulating low-density lipoproteins (LDL) enter
(Wang et al., 2015).
the retina via LDL receptors on RPE and Müller cells (Tserentsoodol
et al., 2006). There is a lipid shuttle from Müller cells to neurons to
meet the need of neurons for lipids, especially to maintain the long
RGC axons and the photoreceptor outer segments, as well as for 3.5. Extracellular matrix generators
synapse formation (Mauch et al., 2001). Müller cells and astrocytes
synthesize ApoE and ApoJ which are assembled into cholesterol- The extracellular matrix (ECM) provided by macroglia is an
rich lipoprotein particles that are thought to be secreted into the important source of supporting and signalling factors in the retina
vitreous humour. From there the particles may be taken up by RGCs and it contributes to the ILM, which is the boundary between the
and transported centrally within the optic nerve. Although the retinal layers and the vitreous humour. ECM components are key
exact mechanism involved in this uptake is still unclear mediators of glial activation, and they have the capacity to evoke
(Amaratunga et al., 1996), the implication of Müller glia in the both regenerative and degenerative effects on RGCs. The ECM
intraretinal transport of lipids dependent on high density produced by glia is altered during reactive gliosis, with many
lipoprotein-like particles has been proposed (Tserentsoodol et al., studies reporting a dramatic increase in the production of ECM
2006). components and the re-expression of extracellular glycoproteins
Docosahexaenoic acid (DHA) is a trophic factor required by that are down regulated after development. The following ECM
photoreceptors for their disk membranes, and it drives several ac- components have been identified in the healthy adult mammalian
tivities in the retina that can help protect photoreceptor cells from retina, unless otherwise specified.
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 15

3.5.1. Laminin are unsupportive of RGC neurite outgrowth are reportedly tenascin
Laminins are a family of glycoproteins that are a major con- negative (although Tenascin C -TN-C- was not assessed), whereas
stituent of the ILM (Kohno et al., 1987), and the specific laminin neonatal astrocytes that do support RGC outgrowth are found to
chains expressed in the retina are now starting to be elucidated. express tenascin (Bahr et al., 1995). A lack of tenascin expression
Laminin acts to anchor cells to the ECM but it also has important has ben proposed to contribute to the failure of adult CNS regen-
signalling functions transduced via transmembrane receptors like eration (Bartsch et al., 1992). More recently, TN-C upregulation was
integrins and dystroglycan. In the adult human retina, laminin g1 seen in optic nerve head astrocytes after experimental injury in rats
has been detected at the ILM of retinal astrocytes, and laminins a4, (Johnson et al., 2007) and in glaucoma patients (Pena et al., 1999).
b2 and g3 have been associated with Müller cells in the GC layer. This suggests that tenascin is downregulated in cultured adult as-
Other laminin chains present in the ILM that may be produced by trocytes that perhaps do not reflect events in vivo.
retinal macroglia include laminins a1, a5 and b1. The specific het- Interestingly, TN-C dampens the reactivity of cultured human
erotrimeric complexes of these laminins in the retinal ECM have astrocytes, transforming them to a quiescent phenotype, and glia
not been fully determined (Libby et al., 2000). RGCs express laminin appear to produce their own TN-C to modulate their own reactive
chains a2 and b1 in the adult mouse retina, which suggests that phenotype in an autocrine manner (Holley et al., 2005). Perhaps,
they contribute to their own survival (Morissette and Carbonetto, the expression of TN-C during regeneration serves to maintain a
1995; Sarthy and Fu, 1990). manageable level of reactivity in order to promote a less inhibitory
environment for RGC regeneration. TN-C expressed by astrocytes
3.5.2. Collagen also acts at the retinal optic nerve junction, where it serves as a
Collagen is another glycoprotein produced by glia and a con- crucial molecular barrier preventing myelinating oligodendrocyte
stituent of the retinal ECM Collagen types I and IV was identified in precursor cells (OPCs) from entering the retina (Bartsch et al., 1994;
the ILM of the human adult retina (Jerdan and Glaser, 1986). Morcos and Chan-Ling, 2000).
Collagen type II was not found in the ILM but found in small Cultured Müller glia from P3 mice promote RGC axonal
“packages” in the vitreous humour, closely associating with Müller outgrowth and express the large TN-C isoform (Siddiqui et al.,
cell endfeet perhaps as a result of ECM remodelling (Ponsioen et al., 2009). Indeed, some isoforms of TN-C inhibit neurite extension
2005). Furthermore, immortalised human Müller cells express whereas others vigorously promote neurite outgrowth, depending
collagens Types I-VII, IX and XI (Ponsioen et al., 2008). Collagen- on their FnIII domains. The FNIII Domain D of TN-C is necessary to
XVIII was also detected at the ILM of the adult human retina. stimulate retinal outgrowth on Müller glia cells (Siddiqui et al.,
Collagen XVIII is a basement membrane heparin sulphate proteo- 2008). An increase in Müller cell proliferation has been observed
glycan (HSPG) that can be cleaved to form endostatin. Endostatin is on stiff substrates and an upregulation of connective tissue growth
a broad spectrum inhibitor of angiogenesis and it may interfere factor (CTGF) and TN-C on softer substrates (Davis et al., 2012).
with the pro-angiogenic action of growth factors like basic FGF Recently, Müller cells were shown to have the potential to re-enter
(bFGF/FGF-2) and VEGF. This suggests that the expression of Type the cell cycle and differentiate into multipotent neuronal pro-
XVIII collagen is related to retinal vascularisation in which astro- genitors that can regenerate cells in damaged areas. This can be
cytes play an important role (Keenan et al., 2012). induced in vitro by exposure to FGF2. TN-C is important in this
process as it enhances Müller cell sensitivity to FGF2. Moreover, de-
3.5.3. Vitronectin differentiation was found to be impaired in TN-C knock-out mice
Vitronectin, an extracellular glycoprotein that can bind to many (Besser et al., 2012).
components of the ECM and to integrins (specifically integrin
avb5). Vitronectin is expressed in the RGC layer of adult human and 3.5.6. Proteoglycans
mouse retinas, and it is up-regulated, along with its integrin re- Chondroitin Sulphate Proteoglycans (CSPGs) are produced by
ceptor, following retinal damage (Anderson et al., 1999; Wang et al., retinal glia and have both positive and negative effects on RGCs.
2006). While vitronectin is expressed by RGCs, it is not expressed CSPGs are crucial mediators of axonal pathfinding during devel-
by astrocytes or Müller cells (Anderson et al., 1999). opment, yet they are known to be inhibitory to axonal regeneration
(Carulli et al., 2005). The CSPG, neurocan, is upregulated by Müller
3.5.4. Fibronectin cells in adult rat retina during reactive gliosis, suggesting that these
Fibronectin is expressed by retinal astrocytes in neonatal rats cells may not support RGC regeneration (Inatani et al., 2000). A
and it plays an important role in angiogenesis (Jiang et al., 1994). comparative study of the ECM produced by different glial cells
Fibronectin was also shown to be produced by cultured adult found that the ECM of Müller cells from neonatal mice express low
mouse retinal astrocytes (Scheef et al., 2005), although there are levels of CSPGs and high levels of growth promoting molecules like
inconsistent reports regarding fibronectin expression in the adult laminin, fibronectin and TN-C. These Müller cells are a favourable
retina and it remains unclear whether or not it is found in the ILM substrate for RGC axon outgrowth and when the production of
of adult retina of different mammalian species (Chen et al., 2009b; CSPGs is inhibited, axonal growth is enhanced. This suggests that
Deeg et al., 2011; Hiscott et al., 1992; Kohno et al., 1987; Ljubimov the ECM of neonatal Müller cells contains more stimulating than
et al., 1996). Fibronectin has an important developmental role in inhibitory components and that it is a favourable substrate for RGC
vascularisation (Jiang et al., 1994), yet its function in the adult retina axons (Siddiqui et al., 2009). These results highlight the need to
requires further study. further analyse the ECM produced by adult retinal glia in order to
assess which growth promoting and growth inhibiting factors are
3.5.5. Tenascin altered with respect to neonatal cells.
Tenascins are extracellular matrix glycoproteins with adhesive Heparan Sulphate Proteoglycans The ILM of the adult human
and counter-adhesive properties that can bind to a multitude of retina contains diverse heparan sulphate proteoglycans (HSPGs)
associated ECM proteins and cell surface receptors, including including perlecan, agrin and Collagen Type XVIII (Clark et al., 2011;
members of the integrin family (specifically a8b1 a9b1 avb3 and Keenan et al., 2012). HSPGs are crucial for axonal guidance during
avb6: Morrison, 2006). Tenascins can be cleaved by matrix metal- retinal development (Chai and Morris, 1999), although their role in
loproteinases (MMPs) and serine proteinases, which alter their the adult retina is not well understood. In mature tissue, sulphated
binding capacities. Astrocytes from the adult rat optic nerve that proteoglycans are crucial for the correct function of many signal
16 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

transduction pathways and for the binding of a multitude of mol- embryonic chick RGCs are cultured on rat cortical astrocytes
ecules, including growth factors (e.g., VEGFs and FGFs), growth (Neugebauer et al., 1988). Similarly, N-Cadherin is required for
factor receptors, chemokines, matrix molecules, proteases and neurite outgrowth when in contact with Müller cells in vitro and
enzyme inhibitors. For this reason, it would be interesting to in vivo (Riehl et al., 1996). Retinal axons from embryonic mouse
explore the roles of HSPGs in relation to mature RGCs since there explants express N-cadherin when cultured on astrocytes whilst
may be multiple signalling pathways that are dependent upon this those from adult explants do not.
type of regulation. N-Cadherin fulfils crucial developmental functions in the em-
bryonic retina and it is downregulated in the adult retina (Bates
3.5.7. Intermediate Filaments (IFs) et al., 1999). However, N-Cadherin is upregulated during regener-
Following CNS injury, astrocytes and Müller cells upregulate the IF ation in the adult zebrafish retina (Liu et al., 2002). Thus, the
proteins GFAP, vimentin and nestin, which alters their physical inability of the mammalian retina to re-express N-Cadherin
properties and makes the cells more rigid. This allows them to following injury may be one of the reasons that RGCs don't
establish a glial scar that is a physical and biochemical barrier to regenerate, making N-Cadherin a potential therapeutic target.
infection, as well as a barrier to regeneration. The upregulation of IFs
also increases contact between cytoskeletal components and the ECM, 3.5.9. Integrins
enabling the cells to initiate rapid cytoskeletal alterations. Astrocytes Integrins are membrane-spanning adhesion molecules that link
and Müller cells have a polarised morphology (Enger et al., 2012) and the intracellular cytoskeleton to the ECM. Integrins are composed
the loss of this polarity may be one factor that contributes to their of a and b subunits, of which there are multiple types, and the
reactivity as they lose the crucial connection with the ECM. resulting heterodimers allow cells to respond to adhesion through a
In the retina, astrocytes and Müller cells express GFAP and number of different ECM ligands. Integrin-mediated cell-adhesion
vimentin at detectable levels in normal adult tissue and both are allows cells to communicate with their environment and the sur-
dramatically upregulated following damage to the retina (Lewis rounding ECM, favouring alterations to the extracellular milieu to
and Fisher, 2003). It is unclear if GFAP is expressed by unreactive be relayed back to the cell, which then alters its morphology
Müller cells since while there are reports that it is only expressed at accordingly. Loss of the connection between the cell and the ECM,
times of stress (Kim et al., 1998; Xu et al., 2009) it has also been for example by MMP induced proteolysis, is one way in which
detected in Müller cell processes of healthy retinas (Erickson et al., retinal degeneration can progress. Based on current evidence,
1987). Increased expression of GFAP and vimentin increases the integrin signalling in RGC-glia interactions is crucial for cell sur-
stiffness of the cells, which does not encourage axonal regeneration vival, process extension and the maintenance of correct retinal
because growing neurites prefer soft substrates (Bringmann and function, although the promiscuity of integrin subunit hetero-
Wiedemann, 2012). There is also evidence that retinal astrocytes dimers combined with their temporal expression in the retina
in zebrafish express GFAP weakly and rather, they rely on cyto- make it difficult to draw definitive conclusions regarding the
keratin IFs to maintain their structural stability, which perhaps functions of specific integrins. What is clear is that throughout
produces a less rigid scar than in mammals (Koke et al., 2010). retinal development and during degeneration there is a strong
There may also be alterations to intrinsic cell receptors in adult requirement for b1 integrin signalling to promote cell survival and
RGCs that make them unresponsive to extracellular guidance cues outgrowth. The importance of b1 class integrins in mediating RGC
since they are less capable of extending processes through 3- neurite outgrowth in the developing retina has been well charac-
dimensional cultures of astrocytes than embryonic RGCs, suggest- terised in the chick and mouse embryonic retina and the a-subunits
ing that their cellecell adhesion properties are significantly altered were found to be expressed in RGCs (Bosco et al., 1994; Bossy et al.,
(Fawcett et al., 1989). 1991; Bradshaw et al., 1995; Cann et al., 1996; Cohen et al., 1987; de
Nestin is expressed very weakly in healthy tissues but it is Curtis, 1993; Duband et al., 1992; Hall et al., 1987; Sheppard et al.,
upregulated after retinal damage, for example in induced glaucoma 1994).
(Xue et al., 2006a), optic nerve transection (Xue et al., 2006b) or Adult rat RGCs preserve the capacity to survive and regenerate
after laser injury (Kohno et al., 2006). Nestin was recently shown to in the presence of different ECM components, such as laminin,
be upregulated in a subpopulation of Müller/astrocytes in the RGC fibronectin, collagen I and collagen IV, demonstrating the tremen-
layer in a rat model of Retinitis pigmentosa. Nestin is better known dous plasticity of adult RGCs to reprogram their ability to express
as a stem cell marker and it is associated with proliferating cells, different integrins depending of the substratum in which they grow
such as stem cells and progenitor cells, and it was proposed that the (Vecino et al., 2015b). It is clear that RGCs are in contact with
re-expression of nestin signifies the dedifferentiation of Müller different ECM components produced by different glial cells within
cells (Valamanesh et al., 2013). The upregulation of nestin by Müller the retina. As the axons exit the optic nerve they come into contact
cells and astrocytes has also been reported following experimental with the ECM components of the optic nerve head, lamina cribosa,
retinal detachment in the adult rat and where differential expres- optic nerve axons and oligodendrocytes. Thus, RGCs must express
sion of nestin within the Müller cell population was observed (Luna different sets of integrins in order to survive when in contact with
et al., 2010). Synemin is another IF protein that is upregulated in different substrata and glial cell types, preserving their capacity to
both Müller glia and astrocytes and like nestin, it is similarly regenerate. We have described the beneficial effect of Müller cells
downregulated after development and its re-expression in Müller on the survival and the capacity of RGCs to regenerate neurites
cells following injury may be an indicator of de-differentiation (Fig. 9). Müller cells secrete beneficial factors but they are even
(Luna et al., 2010). more beneficial when RGCs are grown on a confluent Müller cell
In summary, the ECM substrate provided by confluent Müller substratum (Garcia et al., 2002; Higginson et al., 2013). The bene-
cells in vitro favours the growth and regeneration of RGCs (Garcia ficial effects of Müller cells therefore appear to be due to secreted
et al., 2002). Moreover, the best conditions for RGCs involve a factors and adhesion molecules present on these cells.
combination of ECM plus secreted factors from Müller cells (Vecino Adult human Müller cells express a1, a2, a3 and b1 integrin
et al., 2015b). subunits in culture (Guidry et al., 2003) and cultured rat Müller
cells from P3 rats strongly express the b1 integrin subunit (Siddiqui
3.5.8. N-Cadherin et al., 2009). Multiple integrin heterodimers are expressed in as-
N-Cadherin is essential for neurite outgrowth, as evident when trocytes and the expression of a2, a3, a6, b1 and b4 integrin
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 17

Fig. 9. Effect of Müller cells on RGCs in vitro. (A) Rat RGCs cultured on confluent pig retinal Müller cells monolayer. (B) RGCs incubated directly on the laminin-poly-lysine substrate.
RGCs were labelled with antibodies against-bIII Tubulin (green) and the nuclei were stained with DAPI (blue). Note that there are more RGCs when grown on Müller cells, and with
longer neurites. Scale bar ¼ 50 mm.

subunits by astrocytes of the adult human optic nerve head has development of the visual system depends on the integrity of the
been reported (Morrison, 2006). Thus, based on the resultant het- complement cascade, given that genetic deletion of important
erodimers, we can predict that laminin and collagen are important components of this protein family block the capacity of microglia to
ECM components in this structure. In terms of RGC-astrocyte in- properly remove specific synapses (Schafer et al., 2012). Indeed,
teractions, whilst there could be bi-directional integrin communi- astrocytes induce the expression of complement elements in the
cation in both cells, the primary role of astrocytes in this context is CNS and they act in conjunction with microglia that eliminate
likely to be the provision of binding partners for the integrin sub- complement tagged synapses during synapse pruning (Stephan
units expressed by RGCs in order to relay survival signals. et al., 2012). Astrocytes also engulf and eliminate synapses
through a mechanism independent of complement activation, both
3.6. Regulation of neuronal activity during development and in adults (Chung et al., 2013). Other mo-
lecular mechanisms that trigger the phagocytotic intervention of
Glial cells are classically viewed as passive participants in syn- microglia involve the major histocompatibility complex (MHC class
aptic function. However, there is much evidence that there is a 1), glutamate receptors and ATP signalling, etc (Miyamoto et al.,
dynamic two-way communication between glia and neurons at the 2013).
synapse (Barres, 1991). Neurotransmitters released from pre-
synaptic neurons evoke increases in Ca2þ in adjacent glia. Acti- 3.6.1. Gamma-Aminobutyric acid (GABA)
vated glia then release transmitters, including glutamate and ATP, In addition to GABA and Glutamate, Müller cells express, take up
and these glio-transmitters feedback onto the pre-synaptic termi- and exchange various neurotransmitters (Bringmann et al., 2013).
nal to either enhance or supress further neurotransmitter release. In the outer retina, Müller cells are responsible for all GABA
Transmitters released from glia can also directly stimulate post- removal, whereas they share that task with amacrine cells in the
synaptic neurons, producing either excitatory or inhibitory re- inner retina (Marc, 1992). Astrocytes express a wide variety of
sponses. Müller cells envelop synapses in the retina, like Bergmann neurotransmitter receptors (Porter and McCarthy, 1997) and glia
glia in the cerebellum or Schwann cells at the pre-synaptic responses to neurotransmitters were initially characterized using
neuromuscular junction. Thus, the precise shaping of synaptic ac- cultured astrocytes as a model system. The release of several
tivity depends upon the kinetics of both the pre-synaptic neuro- different neurotransmitters from active neurons can stimulate glia
transmitter release and transmitter re-uptake. In the retina, in intact tissue preparations, evoking increases in glia calcium.
photoreceptor cells, neurons and macroglial cells express high- These responses are mediated by several glutamate receptors,
affinity transporters for neurotransmitters, all contributing to GABAB receptors and muscarinic ACh receptors (Kang et al., 1998;
synaptic function in the retina. Pasti et al., 1997). Moreover, the capacity of glia to respond to
As previously indicated in this review the role of microglia is neuronal activity suggests that they might regulate local blood flow
mainly related to their macrophage action during damage and to in response to changes in neuronal activity (Paulson and Newman,
their capacity of secreting chemical signals that contribute to 1987), which would be crucial for retinal damage.
inflammation, cell death and cell survival. However, microglia also Transmitter release from neurons can stimulate glia and pro-
have essential activities that control synaptic connections and voke the release of glutamate, ATP and other neuroactive sub-
synaptic activity in the healthy brain, both during development and stances from glia. These gliotransmitters that are released by glia
in adulthood (Schafer et al., 2013). Microglia contact synaptic ele- can directly stimulate post-synaptic neurons, evoking either
ments, dendritic spines and peri-synaptic astrocytes, through excitatory or inhibitory responses in these cells (Newman, 2003).
processes that extend and retract dynamically (Tremblay et al., However, whether such responses also take place in vivo remains
2010). The release of microglial factors influences both neuro- unclear. Another interesting question that must be addressed is if
transmission and synaptic plasticity (Bilimoria and Stevens, 2014; the heterogeneity of glial cells and their actions influence the sur-
Schafer et al., 2013). Microglia modify the number of synapses in vival of different neurons, or if different neurons are responsible for
different circumstances and they directly eliminate specific syn- modulating resistance to damage.
aptic contacts or those in close collaboration with astrocytes
(Miyamoto et al., 2013). One key candidate that seems to play an 3.6.2. Glutamate
important role in synapse elimination/remodelling is the classical Glutamate is the main excitatory neurotransmitter in the retina,
complement cascade (Mastellos, 2014; Stephan et al., 2012; Stevens which is used in the transmission of visual signals by photorecep-
et al., 2007). For instance, the remodelling of synapses during the tors, bipolar and ganglion cells. In the outer retina, glutamate is
18 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

continuously released by photoreceptor terminals in the dark and development but that also occurs in the mature brain. Astrocytes
this release is inhibited by light. In the IPL, ON-bipolar cells release and microglia fulfil a key function in synaptic pruning through
glutamate during light exposure whereas OFF-bipolar cells release different mechanisms that ultimately combine to conform a robust
glutamate in the dark (Bringmann et al., 2013). Müller cells are the neural network. Microglia employ their phagocytotic and secretory
main elements responsible for the removal of excess glutamate activities to control the neural circuits in healthy brains (Bilimoria
released by the RGCs. The Müller cells are interposed between the and Stevens, 2014) In mouse, for example, microglia engulf syn-
vessels and RGCs, creating a barrier through which the Müller cells apses during synaptic maturation given that proteins like PSD95 (a
provide nutrients and release essential substances to the RGCs marker of excitatory postsynaptic density) and SNAP25 (a pre-
(Bringmann et al., 2009b). The major glutamate uptake carrier of synaptically localized protein) can be found inside microglial cells
Müller cells is the electrogenic glutamate-aspartate transporter (Paolicelli et al., 2011). Synaptic pruning is delayed in mice with
(GLAST, excitatory amino acid transporter-1, EAAT1: Derouiche and microglia lacking the chemokine receptor Cx3cr1, indicating that
Rauen, 1995). their neuronal networks were immature (Paolicelli et al., 2011).
After EAAT-mediated glutamate uptake into Müller cells, gluta- Astrocytes in both the adult and developing brain mediate the
mate is converted to glutamine by glutamine synthetase (GS), and elimination of synapses and therefore, they participate in the
subsequently, glutamine is transported back to the neurons for reshaping of neural circuits. Astrocytes participate in the formation
glutamate re-synthesis (Danbolt, 2001). Glucose deprivation results of functional synapses within the CNS and they can control synapse
in an increased glutamate uptake as well as the upregulation of maintenance in vitro (Ullian et al., 2001). Glypican (HSPGs) 4 (Gpc4)
EAAT1 and EAAT2 expression, indicative of a regulatory mechanism and 6 (Gpc6) are secreted by astrocytes, and they induce functional
that prevents excitotoxicity in neighbouring RGCs. Indeed, glutamate synapses between purified RGCs (Allen et al., 2012) Indeed, the
uptake by Müller cells would appear to be essential to RGC survival application of Gpc4 to neurons increases glutamatergic neuro-
(Toft-Kehler et al., 2014). The transport process itself triggers a signal transmission. Astrocytes can also eliminate synapses through two
transduction cascade that involves the activation of p60src, PI3K and phagocytotic receptors, namely Megf10 and Mertk, and their ca-
PKB (Akt) linked to extracellular Ca2þ entry, which not only promotes pacity to engulf synapses depends on neuronal activity (Chung
translation by Ser-2448 phosphorylated mTOR but also, transcrip- et al., 2013).
tional activity mediated by Ca2þ-dependent binding of inducible The complement system is part of humoural immunity,
transcription factors, such as AP-1. These results further support the involving a group of proteins that participate in the elimination of
pivotal role of glia cells in glutamatergic transmission in the retina alien cells and debris. However, new functions of this tagging sys-
(Lopez-Colome et al., 2012). Significantly, various receptor agonists, tem are observed in relation to neuronal differentiation and the
including VEGF, erythropoietin and neuropeptide Y can induce the establishment of neuronal networks in the CNS (Mastellos, 2014).
release of glutamate from Müller cells, which induces ATP release Synaptic components tagged with proteins of the complement
upon activation of mGluRs (Bruckner et al., 2012). system are eliminated by microglia and astrocytes, and they
collaborate in this process by secreting signals that induce the
3.6.3. ATP and adenosine synthesis of complement proteins by neurons (Stephan et al., 2012).
Glutamate and ATP are candidates for neuron-to-glia signalling Selective elimination of synapses during post-natal development
associated with volume (Uckermann et al., 2006). Indeed, ‘patho- requires the expression of C1q in cultured RGCs, the initiating
logical’ osmotic swelling of retinal (Müller) glia in rat retinal slices protein in the complement cascade (Stevens, 2008; Stevens et al.,
(evoked by the potassium channel blocker, barium chloride, by 2007). Moreover, the absence of the C1q or C3 protein in mice
inflammatory lipid mediators, or under pathological conditions like (downstream elements in the complement cascade) leads to
diabetes and ischemia) can be prevented by pharmacological acti- excessive retinal innervation of the LGN (lateral geniculate nucleus)
vation of purinergic P2Y and adenosine receptors (Uckermann as a consequence of defective synapse pruning, and neural activity
et al., 2006; Wurm et al., 2008a). Activation of purinergic re- regulates the engulfing capacity of microglia (Schafer et al., 2012).
ceptors was recently concluded to be fundamentally involved in the Other phagocytotic mechanisms may be implicated in the process
regulation of glial cell volume and thus, in the homeostasis of the of of synaptic pruning and thus, further studies in other regions of the
the extracellular space mediated by these cells (Wurm et al., 2011). CNS and in different experimental models will be needed to
Moreover, the inhibitory effects of VEGF and glutamate on the ascertain the precise role of microglia and astrocytes in the
swelling of mouse Müller cells is mediated by transactivation of remodelling of synapses during development and in maturity.
P2Y1R, as well as by intracellular calcium signalling through IP3R2 Indeed, many questions are still unanswered and the role of glial
activation (Grosche et al., 2013). cells in CNS remodelling remains ill-defined (Schafer et al., 2013).
Adenosine is a reactive metabolite involved in cellular A close relationship exists between Müller cells and RGCs,
communication in some pathological states and in the eye, retinal whereby glial cell processes often ensheath ganglion cell somata, as
ischemia and/or hypoxia provokes an increase of adenosine. Four well as some of the axonal, dendritic and synaptic regions of these
adenosine receptor subtypes have been identified (A1, A2A, A2B retinal neurons, an ensheathing pattern also common to astrocytic
and A3: Kang et al., 2013). In the retina, adenosine suppresses processes (Hollander et al., 1991). This intimate association be-
excitatory neurotransmission through various mechanisms, tween glia and neurons is thought to allow glial cells to promote
including the inhibition of pre-synaptic voltage-dependent calcium synapse formation and it helps maintain neuronal function by
channels (Housley et al., 2009). Since adenosine is a downstream providing nerve terminals with energy substrates and neuro-
mediator of the swelling-inhibitory actions of glutamate and ATP, it transmitter precursors (Pfrieger and Barres, 1996).
is likely that the inhibitory effect of adenosine induces chloride
channel opening (Bruckner et al., 2012). 3.8. Immune responses

3.7. Synaptic pruning The retina is considered an immune-privileged tissue and


therefore, a multitude of mechanisms operate in the eye to limit the
Synaptic pruning refers to the sculpting that reduces and re- harmful inflammation that foreign substances provoke. At the same
models excess synaptic neuronal contacts to obtain functional time, different mechanisms defend the eye from invading patho-
neuronal circuits, a process that takes place mainly during gens, with ocular cell surface and soluble factors maintaining the
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 19

eye's immunosuppressive nature. RGCs are thought to play a major fibroblasts, endothelial cells and macrophages to produce chemo-
role in the recognition and subsequent clearance of pathogens kines, promoting the recruitment of neutrophils and macrophages
during infection, and they are believed to represent the first line of to the retina, thereby resulting in further tissue damage and chronic
defence in the retina. While microglial cells are mainly responsible inflammation. It is now known how IL-6 expression is regulated in
for these effects, astrocytes and Müller cells also collaborate in this Müller cells. These cells produce high levels of IL-6 in response to
activity. Microglial cells are capable of modifying their positions in IL-1b stimulation and activation of the p38MAPK/NF-kB signalling
order to maintain homeostasis, however, both astrocytes and pathway plays a predominant role in regulating IL-6 expression (Liu
Müller cells can detect damage since both expand their projections et al., 2014c).
to the blood vessels while essentially maintaining their positions. Some activities are common to the immune system and CNS: the
Different immune cells are situated at different locations within secretion of immunoregulatory cytokines; the response to cyto-
the retina. Dendritic cells (MHC-IIþ33D1þ) are preferentially kines; and antigen presentation. These properties reflect the
located around the optic disc and the peripheral margin of the physical contact between the two systems (microglia, astrocytes
retina (Xu et al., 2007). Moreover, their role in optic nerve and and Müller cells presenting antigen to T cells) and their commu-
retinal inflammation remains to be elucidated. Perivascular mac- nication via soluble factors like cytokines. Indeed, cytokines
rophages play a critical role in retinal vascular inflammation and mediate a complex circuit of interactions, especially in the event of
they may also participate in neuronal retinal defence. Microglial lymphoid-mononuclear cell infiltration into the retina. The secre-
cells are the main innate immune cells in the retina and their tion of interferon g (IFN-g) by infiltrating activated T cells could
activation is tightly regulated by a number of inhibitory pathways, initially induce astrocytes, Müller cells and microglia to express
including CD200/CD200R (Copland et al., 2010) or CX3CL1/CX3CR1 class I and class II MHC antigens, as well as priming these cells for
(Chen et al., 2012). Complement proteins are expressed in the subsequent cytokine production that would contribute to either the
retina under normal physiological conditions (Luo et al., 2011) and initiation and/or propagation of immune responses. Müller cell-
their expression augments in pathological conditions like inflam- mediated inhibition of T cell proliferation can be overcome by
mation (Chen et al., 2010) and ageing (Chen et al., 2007). Comple- treating Müller cells with glucocorticoids, presumably preventing
ment activation is related to various retinal inflammatory and the expression of the Müller cell inhibitory factor (Roberge et al.,
degenerative diseases, and modulating these pathways may have 1991), supporting a role for Müller cells as antigen-presenting
therapeutic applications. cells (APCs) in the retina. A number of inflammatory mediators in
Recent studies have shown that Müller glia actively participate the CNS act by ultimately suppressing immune responses, inhibit-
in retinal innate immunity, specifically in infectious conditions like ing class I and II MHC expression, and dampening cytokine pro-
bacterial endophthalmitis, an activity that involves their Toll-like duction by glial cells, such as that of prostaglandins, IFN-a and IFN-
receptors, phagocytosis, and the secretion of cytokines and che- b, TGFb and neurotransmitter like norepinephrine. The induction
mokines. Müller cells also influence the adaptive arm of immunity and ultimate downregulation of immune responses and cytokine
by inhibiting the proliferation of T-cells via membrane-bound production within the retina depends on: a dynamic interaction
molecules. As a consequence of Müller cell activation, pro- between a variety of peripheral immune cells and retinal cells; the
inflammatory cytokines and other potent mediators of inflamma- activation of these cells; the presence of cytokines with pleiotropic
tion are synthesised and released (Kumar et al., 2013). effects; the concentration and location of these cytokines in the
Müller glia cells increase complement protein expression retina; and the temporal sequence in which a particular cell re-
(Kuehn et al., 2006). The complement protein C1q, an initiator of sponds to these cytokines. Thus, the ultimate outcome of immu-
the classical complement cascade, is upregulated in neurodegen- nological and inflammatory events in the retina will be determined
erative diseases like Alzheimer's and amyotrophic lateral sclerosis in part by the interplay of these parameters (Benveniste, 1993).
(ALS), as well as in rodent and human glaucoma (Stasi et al., 2006;
Stevens et al., 2007). The complement cascade has been implicated 3.9. Gliosis
in the normal elimination of synapses by RGCs during visual system
development in the DBA/2J mouse model of glaucoma (Howell In the retina, reactive gliosis alters the morphology of the retina
et al., 2007). Thus, it might be possible that complement protein through hypertrophy, Müller cells and astrocytes provoking the
deposition occurs prior to significant loss of RGCs in the IPL. These thickening and enlargement of processes. This response occurs in a
findings strongly suggest that a normal mechanism of develop- multitude of disorders like trauma, ischemic damage, infection,
mental synapse elimination is reactivated in the mature visual neuroinflammation or neurodegeneration, and in addition to the
system by the glaucomatous process, and that this process of syn- morphological changes, dramatic alterations in astrocyte gene
apse elimination may well entirely drive glaucomatous neuro- expression occurs (Hernandez et al., 2000; Pekny and Nilsson,
degeneration. Most of the secreted C1q is released by microglia, 2005). The concept of reactive astrogliosis, especially its molecu-
which express C1q mRNA strongly. However, what communication lar and cellular characteristics, are still ill-defined, and we are only
occurs between astrocytes and microglia, and the chronology of the starting to understand its multifaceted contribution to disease
events that take place in the early stages of glaucoma remain un- pathogenesis and recovery from injuries (Barres, 2008; Burda and
clear, although it is known that microglial activation is an early Sofroniew, 2014; Zamanian et al., 2012). In reactive gliosis the
event in this process (Bosco et al., 2008; Steele et al., 2006). expression of many genes is altered, and these cells exhibit distinct
Müller cells produce different cytokines in response to different morphological and functional features. Given the molecular
stimuli. For example, IL-1b (interleukin-1b) and TNF-a are potent complexity of the response of reactive astrocytes, it is unlikely that
inducers of IL-8 expression in Müller cells (Liu et al., 2014b), and IL- a single molecular target or pathway will be implicated, indicating
1b and lipopolysaccharide (LPS) also provoke IL-6 production that gliosis is a very complex reaction in which many elements are
(Yoshida et al., 2001). IL-6 is a key cytokine in the host's defence implicated (Pekny and Pekna, 2014).
against environmental stress, such as infection, inflammation and In response to infection and in certain inflammatory scenarios,
injury (Mesquida et al., 2014). Indeed, the dysregulated and Müller cells undergo reactive gliosis. Müller gliosis is a generalized,
persistent production of IL-6 has been implicated in the develop- more unspecific, yet a very complex response of Müller glia towards
ment of various autoimmune and chronic inflammatory diseases, injury or infection. Müller cells become activated by nearly every
as well as ocular disorders. These cytokines can stimulate pathological challenge in the retina (Bringmann et al., 2009a).
20 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Reactive gliosis is thought to be the result of signals received from angiogenesis in early development, as well as modifying regener-
the injured or diseased tissue that sets off a molecular cascade ative neurite growth and neovascularisation (NV) under patho-
within the glial cells to provoke a change of state (Fischer and logical conditions (Lange et al., 2012).
Bongini, 2010). Typical features of Müller cell gliosis involve Surfactant protein A (SP-A) upregulates cytokine expression in
cellular hypertrophy, up-regulation of IF expression (e.g., GFAP and lung disease associated to prematurity. SP-A expression and local-
vimentin), increased rates of proliferation, loss of function, ization has now been characterized in the mouse retina, as has its
inflammation, downregulation of GS expression, and tissue and impact on NV in the mouse. Retinal and Müller cell SP-A is upre-
vasculature remodelling (Belecky-Adams et al., 2013). gulated via the NFkB pathway and it is up regulated during the
A mild to severe continuum of the states of reactive gliosis may hypoxia phase of oxygen-induced retinopathy (OIR). A lack of SP-A
exist. In the mild to moderate forms of gliosis, the cells may un- attenuates NV in the OIR model and thus, SP-A may be a marker of
dergo hypertrophy and show some changes to their functionality, retinal inflammation during NV (Bhatti et al., 2014). Müller cells
yet if the trigger is removed, the cells might revert back to their also produce PEDF (Pigment epithelium-derived factor), TGF-b and
former condition without altering the tissue. In the more severe thrombospondin-1 as anti-angiogenic proteins that are important
forms of reactive gliosis, cells undergo hypertrophy, lose their in adjusting the ‘‘angiogenic balance’’ (Eichler et al., 2004). Recent
functionality, form glial scars that are inhibitory to axonal regen- studies suggest that in addition to PEDF (Yafai et al., 2007), TGF-b2
eration and neuronal survival, and they may also proliferate The is one of the Müller cell derived factors that limit intra-retinal NV
more severe state is marked by the persistence of these charac- (Yafai et al., 2014).
teristics. Proliferative gliosis is generally associated with chronic Within the mammalian retina, both the Müller glia and retinal
gliosis and it is usually detrimental but it can also have a beneficial astrocytes react to injury and disease. Although the mechanisms
effect (Sofroniew, 2009). In such a state, the integrity of the and triggers involved have not been fully characterized. A variety of
blooderetina barrier is lost and as a result, serum components leak growth factor signalling mechanisms have been found to be
into the perivascular space, causing the re-entry of Müller cells into important in reactive gliosis, such as those of EGF, FGF, TNF-a,
the cell cycle and their proliferation (Bringmann et al., 2009a; ciliary neurotrophic factor (CNTF), insulin and WNTs (Nakazawa
Coorey et al., 2012). et al., 2006; Sofroniew, 2009). Bone morphogenetic protein 7
The release of VEGF by Müller cells under conditions stimulating (BMP7) also induces changes in the mRNA and protein of markers
gliosis is especially interesting, as it acts in two directions. VEGF is typically associated with reactive gliosis in retinal astrocytes and
neuroprotective at low concentrations but at high concentrations it Müller glial cells, including GFAP, GS, a subset of CSPGs, MMPs, and
causes all sorts of vascular pathologies. VEGF has neurotrophic ef- other molecules (Dharmarajan et al., 2013).
fects (Zheng et al., 2012) and can prolong the survival time of cells
by inducing angiogenesis, which in turn attenuates brain edema 3.10. Neuroprotection
after cerebral ischemia/reperfusion and neuron injury. The neuro-
protective effect of VEGF was proposed to be unrelated to its 3.10.1. Neurotrophic factors
angiogenic effect but more closely related with its direct stimula- Neurotrophins and growth factors are involved in the survival,
tion of neuronal and endothelial activity (Oosthuyse et al., 2001). differentiation and development of neurons. Administration of
The regeneration of neurons together with angiogenesis could be different combinations of neurotrophins in the eye has been re-
improved by VEGF, especially after ischemia (Ma et al., 2011). ported to increase RGC survival in different experimental para-
Conversely, the excessive and prolonged expression of VEGF has digms. Glial cells are important to maintain the survival of neurons,
devastating effects on the retinal vasculature and/or Müller cell although the factors involved in the development and survival of
proliferation, which in turn might lead to retinal neuro- glial cells themselves are less well defined. Glial like Müller cells,
degeneration (Tolentino et al., 2002). Moreover, calcium contrib- astrocytes and microglia are known to synthesise neurotrophic
utes to the high-glucose-induced expression of the angiogenic factors and these can directly or indirectly mediate neurotrophic
factors HIF-1a and VEGF by retinal Müller cells. The activation of actions (Garcia et al., 2003; Ruiz-Ederra et al., 2003; Vecino et al.,
the CaMKII-CREB pathway by high-glucose may be a possible 1998a, 1998b). In order to define the cell-to-cell communication
mechanism underlying the pathogenesis of diabetic retinopathy (Li between Müller glia and neurons, it is essential to first isolate the
et al., 2012). respective cell types. When adult retinas were maintained in vitro
Müller cells are not alone in influencing the effects of VEGF as for 6 days, the expression of neurotrophins and their receptors was
astrocytes are also considered a target for ischemic damage. Apart preserved in glial cells and in RGCs, irrespective of the substrate on
from their neuroprotective effect on neurons and their detoxifica- which they were cultured (Fig. 10: Garcia et al., 2002). Moreover,
tion of various ROS at the acute stage of ischemic stroke, reactive the factors secreted by Müller cells offered protection to RGCs
astrocytes profoundly affect later post-ischemic stages. This influ- (Garcia et al., 2003).
ence may be mediated by the secretion of VEGF, which stimulates The distribution of neurotrophins and their receptors in the
the formation of new blood vessels and synaptogenesis, or by the mammalian retina has been studied in detail, even in pathological
secretion of synaptogenic thrombospondin (Christopherson et al., conditions like ischemia or following NMDA induced exitotoxicity
2005). The positive stimulatory effect of VEGF on vessel and syn- (Vecino et al., 1998b, 1999). BDNF expression by rat RGCs is
apse formation in ischemic stroke, contrasts with its induction of completely blocked after intravitreal injection of NMDA, as is
blood-retinal barrier breakdown and lymphocyte infiltration in choline acetyltransferase (ChAT) immunoreactivity associated with
autoimmune CNS inflammation (Argaw et al., 2009). It is also a subset of amacrine cells. However, immediately after NMDA in-
noteworthy that treatment with GDNF enhances neuronal survival jection, BDNF mRNA expression by RGCs and its immunoreactivity
in ischemic brain tissue and that GDNF is up regulated in reactive is enhanced, while the amacrine cell ChAT immunoreactivity is
astrocytes in the peri-infarct region (Kokaia et al., 1999). Netrin-4, is clearly reduced, and expression of the mRNA transcripts encoding
a positive regulator of retinal angiogenesis and Müller cells rhodopsin and Thy-1 does not change. However, a few hours later
generate Netrin-4 in a hypoxia- and VEGF-dependent manner, an increase in BDNF expression is no longer apparent, suggesting
another example of a multifunctional protein produced by Müller that the synthesis of BDNF increases in RGCs immediately after an
cells that may influence their complex relationship with retinal NMDA insult a natural protective mechanism in retinal cells
neurons. Netrin-4 can support both axon guidance and (Vecino et al., 1999). A similar transitory increase of BDNF and its
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 21

Fig. 10. Preservation of the neurotrophins and their receptors in (red) from adult pig retinal glia (in vivo: A, C and E) and primary cultured cells (in vitro: B, and F). (A and B) high-
affinity neurotrophin receptor, TrkA (red: C and D), low-affinity neurotrophin receptor, p75 (E and F) and neurotrophin-3 (NT3) (E). In D and F, retinal ganglion cells were stained
with anti-neurofilament antibodies (green). Scale bar ¼ 50 mm.

high-affinity TrkB receptor is also found in photoreceptors and 2hr of reperfusion (Fig. 7). Müller cells modulate neuronal activity
RGCs after ischemia-reperfusion. The mechanisms underlying by regulating the concentration of neuroactive substances and
BDNF induced neuroprotection are unknown, although it has been potassium in the extracellular space by means of a variety of
suggested that the BDNF derived upregulation of GLAST and GS transport processes (Newman and Reichenbach, 1996). Hence,
may contribute to this effect. Indeed, BDNF has a protective effect Müller cells are particularly important in ischemia when there is a
on GLAST and GS expression by Müller cells under hypoxic condi- massive release of transmitters from neurons (Hossmann, 1993).
tions, which could induce neuroprotection (Dai et al., 2012b). Thus, the changes seen in the p75 immunoreactivity associated
The Müller cell extensions that form three laminae in the IPL of with the branches of the Müller cells in the inner plexiform layer
the rat retina contain p75. After ischemia, there is a progressive after ischemia may therefore indicate changes in plasticity and be
disorganization of the laminae that is worse after 24hr than after of functional significance (Vecino et al., 1998b). Disorganization of
22 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Müller cell processes has also been observed in RP65 mutant dogs, Müller cells following retinal hypoxia or injury. PEDF functions as
in which the primary cells affected are photoreceptors, while in an anti-oxidative and anti-inflammatory agent, as well as exerting
glaucomatous rat retinas the cells affected are RGCs. These changes neuroprotective and neurotrophic effects that are mediated by
within the Müller cells again demonstrate that Müller glia have a activation of the NF-kB (nuclear factor kappa B) and ERK1 and 2
radial function in the retina that senses the changes in the outer as pathways. When secreted by Müller cells, this factor acts directly on
well as the inner retina, reacting by modulating the structure and PEDF receptors at the surface of RGCs. The upregulation of PEDF in
possibly the function of retinal neurons. Müller cells occurs early in gliosis and it may reflect an attempt to
rescue retinal neurons from further damage (Unterlauft et al.,
3.10.2. Other trophic factors 2012). Indeed, PEDF reduces RGC loss in DBA/2J mice, a mouse
Ciliary neurotrophic factor (CNTF) CNTF exerts numerous ef- model of inherited glaucoma (Zhou et al., 2009).
fects on the development, differentiation and survival of retinal The expression of PEDF, Kir4.1 and GLAST by retinal Müller cells
neurons (Rhee and Yang, 2010). The molecular mechanisms that is unbalanced at high glucose concentrations, with abnormal
explain the neuroprotective role of CNTF in the retina include: a expression closely related to oxidative stress. Decreased expression
direct action on photoreceptors to prevent apoptosis; the stimula- of PEDF induces a down-regulation of Kir4.1 in Müller cells, causing
tion of Müller cells to produce photoreceptor survival factors (Zack, membrane depolarization of Müller cells and GLAST dysfunction.
2000); enhanced synthesis or distribution of glutamate trans- Thus, retinal PEDF must balance the anti-oxidative state against the
porters, thereby improving glutamate handling and resulting in need to protect retinal neuronal and glial cells (Xie et al., 2012).
less excitotoxic damage to retinal neurons (Escartin et al., 2006); Further evidence for the neuroprotective actions of Müller cell
and an induction of metabolic plasticity and increased resistance to derived PEDF in the retina involves the intracellular signalling
metabolic damage (Escartin et al., 2007). Defining the global tran- molecules that may operate in the RGC under conditions in which
scriptional response of Müller cells to CNTF in vivo will help PEDF of glial origin exerts its pro-survival effects on RGCs. Thus,
elucidate the molecular basis of its biological actions in the retina. PEDF-induced NF-kB activation enables RGCs to protect themselves
CNTF induces rapid and extensive transcriptional changes in Müller from hypoxia- or growth-factor deprivation-induced apoptosis.
cells and several genes induced by CNTF in normal Müller cells are Although not directly demonstrated, exposing RGCs to PEDF may
also up regulated in gliotic Müller cells involved in inherited and increase the phosphorylation of the inhibitor NF-kB, I-kB, thereby
experimentally-induced retinal degeneration. Hence, CNTF would unmasking the NF-kB/NLS domain at the C terminus of NF-kB,
appear to be an inducer of gliosis in the retina. Interestingly, the stimulating its nuclear translocation and increased its DNA bind-
transcript profiles of Müller cells and astrocytes are similar ing activity (Unterlauft et al., 2014).
although they are derived from distinct cell lineages (Xue et al., Insulin-like growth factor-1 (IGF-1) is a polypeptide growth
2011). Exogenous CNTF initially targets Müller glia and they factor similar in structure and function to insulin. IGF-1 has mito-
trigger a reaction that induces the production of cytokines that act genic, metabolic and growth-stimulating activities in many cells
through gp130 in photoreceptors to promote neuronal survival and tissues, including the retina, and it promotes the survival of
(Rhee et al., 2013). Indeed, CNTF not only reduces the death of various cells following serum deprivation. This protective effect of
photoreceptors but it also rescues retinal function (Heiduschka IGF-1 most probably occurs via PI3K/Akt, although the complete
et al., 2013). In addition, lens injury upregulates CNTF in retinal details of the upstream and downstream targets have not yet been
astrocytes and Müller cells. elucidated (Guvakova, 2007).
bFGF is a pleiotropic cytokine that, in addition to its pro- Glial Derived Neurotrophic Factor (GDNF) in the retina has
angiogenic actions, may elicit further effects on retinal cells. bFGF been reported to rescue photoreceptor activity in the rd1 mouse
is found in astrocytes, Müller cells, RGCs and RPE cells. Although model of Retinitis Pigmentosa (Frasson et al., 1999) and its effects
bFGF is considered to be neuroprotective in the retina, it may also are transmitted indirectly via activation of Müller glia (Harada et al.,
have detrimental effects by stimulating aberrant vessel growth, or 2003; Hauck et al., 2006). Transcriptome profiling of the mouse
inducing the proliferation and de-differentiation of Müller cells retina in response to GDNF identified osteopontin (OPN) as a novel
(Bringmann et al., 2009a). Together, bFGF and VEGF act synergis- GDNF effector that could in part mediate in the GDNF induced
tically on microvascular endothelial cells (Yan et al., 2001), whereby neuroprotective activity of Müller cells on photoreceptors (Del Rio
the effects of bFGF are in part mediated by stimulating VEGF release et al., 2011). OPN is secreted by Müller glia and it supports photo-
from Müller and vascular endothelial cells (Hollborn et al., 2004). receptor survival in explant cultures of rd1 retinas, potentially
Müller cells represent a major source for bFGF in the ischemic exerting a direct pro-survival effect on photoreceptors. OPN has a
retina and in diabetic fibrovascular tissue. The secretion of bFGF by significant, concentration-dependent survival effect on primary
Müller cells is stimulated by various pro-angiogenic growth factors porcine photoreceptor cells, and it activates the PI3K/Akt pro-
and cytokines. Indeed, while the soluble factors released by Müller survival pathway (Del Rio et al., 2011). OPN also induces the
cells in response to transient hypoxia produce an anti-angiogenic release of VEGF, glutamate, ATP and adenosine from Müller cells.
environment for retinal vascular endothelial cells, persistent hyp- The effect of OPN was prevented by blockers of voltage-gated so-
oxia increases bFGF release from Müller cells, which may become dium channels (tetrodotoxin), T-type voltage-gated calcium chan-
increasingly important to shift the balance between stimulators nels (kurtoxin), potassium channels (clofilium) and chloride
and inhibitors of angiogenesis in the retina. In such conditions channels (5-nitro-2-(3-phenylpropylamino)-benzoic acid) (NPPB).
Müller cells are likely to provide a stimulatory environment for the The swelling-inhibitory effect of OPN was dependent on intracel-
proliferation of retinal vascular endothelial cells and bFGF may play lular calcium signalling, the activation of PLC and protein kinase C,
a key role in the stimulation of abnormal vessel growth in the retina and vesicular exocytosis of glutamate. In retinal slices, Müller glial
(Yafai et al., 2013). In diabetes, the increased formation of advanced cells are immunoreactive for OPN and Müller cell-derived OPN has
glycation end-products (AGEs) in the vitreous humour may also be autocrine effects. The neuroprotective effects of OPN may in part be
involved in the initiation and progress of intraocular NV through mediated by preventing the cytotoxic Müller cell swelling, and the
the production of bFGF by Müller cells (Ai et al., 2013). release of VEGF and adenosine by Müller cells (Wahl et al., 2013).
PEDF (Pigment epithelium derived factor) Müller glial cells VEGF (see section on Gliosis above).
are PEDF producers, a positive trophic factor expressed by activated
Müller cells. PEDF is a 50 kDa glycoprotein and it is upregulated by
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 23

3.11. Phagocytosis Indeed, reactive glia secrete trophic factors distinct to those pro-
duced by quiescent glia and the degree of gliosis also appears to be
The phagocytotic activity of Müller cells was described more crucial to determine the success of RGC regeneration.
than eight decades ago. There is no evidence that these cells can Can retinal glia be primed to have positive or negative effects on
phagocytise pathogens, however the phagocytotic properties of RGC survival and regeneration? Müller cells secrete factors to the
Müller cells were recently revaluated following Staphylococcus media that favour the survival, regeneration and branching of RGC
aureus infection. As a result, Müller cells can apparently not only axons in vitro (Garcia and Vecino, 2003; Higginson et al., 2013).
phagocytise bacteria but they can also kill them intracellularly Activation of rat retinal glia (astrocytes and Müller cells), either by
(Singh et al., 2014). lens lesion, intravitreal peripheral nerve grafting, zymosan injec-
Astrocytes play a role in the life cycle of axonal mitochondria at tion or experimental glaucoma (Lorber et al., 2009, 2008, 2002,
the head of the optic nerve. Optic nerve fibres discard their mito- 2012), enhances the regenerative potential of RGCs in vivo and
chondria, which are then engulfed by neighbouring astrocytes in vitro. It is assumed that retinal glia are activated by factors pro-
(transmitophagy: Davis et al., 2014). In the myelination transition duced after lens injury or peripheral nerve graft, or by macro-
zone, astrocytes have a constitutive phagocytotic role and they phages, which may be proteinaceous or non-proteinaceous, and
express the phagocytotic marker Mac-2, which augments in glau- that they are then stimulated to secrete neurotrophic factors that
coma, and it may play an important role in axonal degeneration enhance the growth of RGC axons. A trophic factor produced by
(Nguyen et al., 2011). activated astrocytes and Müller cells that enhances RGC axonal
The death of RGCs after optic nerve injury involves the partici- regeneration is CNTF. Its production is mediated by apolipoprotein
pation of microglia/macrophages and astroglia in a neuro- E (ApoE), which is upregulated in retinal astrocytes and Müller cells
inflammatory scenario. Simvastatin, a statin used to lower blood after optic nerve crush and intravitreal Zymosan injections (Lorber
cholesterol, may reduce NF-kB activation by TNF-a in astrocytes et al., 2009; Muller et al., 2007). Other trophic factors produced by
maintained in culture. When administered systemically, the drug reactive retinal astrocytes and Müller cells that may be involved in
suppresses RGC death possibly by ameliorating the astroglial re- RGC regeneration include BDNF, GDNF and bFGF (Bonnet et al.,
action to injury (Morishita et al., 2014). TNF-a can activate signal- 2004; Garcia et al., 2002, 2003; Garcia and Vecino, 2003; Harada
ling reactions involving NF-kB or C-Jun N-terminal kinase (JNK) et al., 2000, 2002Kinkl et al., 2003).
pathways. These pathways influence neurodegeneration and serve Although mammalian Müller glia can respond to injury, prolif-
as neuroprotectors. The detrimental and neuroprotective role of erate and express genes that are associated with retinal stem cells
TNF-a signalling was recently assessed in cultured RGCs and as- (Roesch et al., 2008), they do not function as retinal progenitors
trocytes (Dvoriantchikova and Ivanov, 2014), apparently activating in vivo. Nonetheless, these characteristics suggest that under the
NF-kB and JNK in astrocytes but not in RGCs, indicating a contrast in right circumstances, Müller glia might act as retinal progenitors
both cell populations. However, the significance of this finding that can be used for repair. The regenerative potential of Müller
in vivo remains unclear. TNF-a also induces the secretion of MMP-2 cells makes them ideal candidates for cell replacement therapy to
by optic nerve astrocytes and this secretion is attenuated by a se- treat retinal degenerative diseases, such as age-related macular
lective delta opioid agonist, SNC121 (Akhter et al., 2013). degeneration, retinitis pigmentosa or glaucoma, where loss of
Extracellular ATP acts as a ligand to purinergic receptors and vision correspond to retinal neuron loss. A regenerative strategy
P2X4 purinergic receptors (ionotropic receptors) are distributed in that mobilizes the patients’ own Müller progenitors to participate
different cell types within the retina (bipolar cell terminals, hori- in retinal repair would have several advantages over cell trans-
zontal cell somata and processes, calretinin-positive amacrine plantation therapy (Ruilin et al., 2014). Indeed, in rodent and hu-
cells…etc), microglia, Müller cells and astrocytes. These receptors man cell cultures, Müller glia can generate both neurons and glia
influence glia function and local tissue homeostasis in both physi- (Das et al., 2006). Importantly, primary human Müller glial cell
ological and pathological conditions (Vessey et al., 2014). cultures can generate photoreceptors and RGCs that have some
reparative potential when transplanted into a damaged rodent
3.12. Repair and regeneration retina. These studies suggest that human Müller glia are capable of
generating neurons under appropriate conditions and that they
Retinal astrocytes and Müller cells both produce multiple factors may be able to participate in repair. Indeed, human Müller glia stem
that are trophic to RGCs, perhaps at different levels and under cells (hMGSCs) can differentiate into functional RGC precursors
different conditions. RGCs express a number of receptors that in vitro, as confirmed by their expression of RGC markers and the
respond to trophic factors including but not restricted to leukaemia increased cytosolic calcium in response to nicotinic stimulation
inhibitory factor (LIF), CNTF, NGF, BDNF, IGF, neurotrophins, GDNF (Singhal et al., 2012).
and FGF2 (Bonnet et al., 2004; Kinkl et al., 2003; Vecino et al., Viral expression of ASCL1 is sufficient to activate a neurogenic
1998b, 2002). However, promotion of RGC survival is only tran- program in mammalian Müller glial, both in dissociated cultures
sient and does not necessarily correspond with the restoration of and in the intact retina. ASCL1 remodels the chromatin of retinal
axonal outgrowth. In fact, whilst some of these trophic factors are progenitor genes, activating their expression and downregulating
very efficient at enhancing cell survival, some of them promote glial genes. The reprogrammed Müller glial differentiate into cells
axonal growth (Vecino et al., 2002), indicating that a delicate bal- that resemble neurons morphologically, both in terms of gene
ance must exist to promote optimal RGC regeneration. expression and in their responses to neurotransmitters. These re-
Reactive glia clearly display inhibitory properties, yet an inter- sults suggest that stimulating neurogenesis in Müller glial with
mediate “activated” glial phenotype may also exist that favours ASCL1 could provide an alternative strategy for retinal repair after
axonal regeneration by upregulating factors that enhance their disease or injury (Pollak et al., 2013). Similarly, manipulating the
growth and repair, whilst keeping inhibitory responses below the Ephrin-A2/A3 pathway may represent a viable approach to pro-
threshold for damage. Acute activation of glia is neuroprotective, mote the progenitor cell fate of Müller cells, since Ephrin-A2 and A3
while sustained activation is detrimental, for example due to the inhibit the regenerative potential of Müller cells in the retina of
release of inflammatory cytokines. Enhanced reactivity or “activa- adult mice (Ruilin et al., 2014).
tion” of retinal glia has positive effects on RGCs due to the pro- Among the possible molecular and cellular mechanisms con-
duction of trophic factors, both in vivo after injury and also in vitro. trolling neuronal regeneration, p53-mediated cell cycle arrest may
24 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

play an important role in preventing Müller glia proliferation in the differentiate and re-enter the mitotic cycle. Under pathological
adult mouse retina (Ueki et al., 2012). The loss of p53 activity could conditions, hypertrophy and the proliferation of glial cells leads to
significantly enhance the proliferation potential of cells and facili- the formation of glial scars, which fill the spaces left by dying
tate the reprogramming of Müller glia into retinal progenitor cells neurons and disconnected synapses (Reichenbach and Bringmann,
(Zhao et al., 2014a). The tumour suppressor p53 plays a funda- 2010). Müller glia in mammalian retinas can respond to injury,
mental role in astrocytes, when environmental conditions are proliferate and express genes that are associated with retinal stem
deleterious, as in heightened conditions of oxidative stress. In cells (Roesch et al., 2008), displaying a gene expression profile
“super p53” mice that carry two copies of the p53 gene, the density resembling that of progenitor cells (Blackshaw et al., 2004). A
of astrocytes is greater than in the wild type and the area they subset of Müller cells express progenitor genes like Car2, Dkk3,
occupy is also augmented (Salazar et al., 2013). Chx10 and Pax6 (Roesch et al., 2008; Rowan and Cepko, 2004), and
Atoh7 is an essential transcription factor that determines the Müller cells in the adult mouse retina express cell cycle genes like
competence state of RGC progenitors. Ectopic expression of Atoh7 cyclinD3, Cdc14A, Cdk10 and Spbc25, as well as genes in the Notch
significantly increases the number of RGCs that differentiate from pathway that might regulate the re-entry of Müller glia into the cell
cultured retinal stem cells in vitro (Yao et al., 2007). Moreover, cycle under pathological conditions (Roesch et al., 2008).
Atoh7 promotes the differentiation of Müller cell-derived retinal Müller glia might adopt characteristics of retinal progenitors
stem cells into RGCs by inhibiting Notch signalling, opening up a that can be used to repair injured retina. In chicks (Fischer and Reh,
new avenue for gene therapy and optic nerve regeneration in 2001) and rats (Das et al., 2006; Ooto et al., 2004) Müller cells may
glaucoma (Song et al., 2013). even differentiate into cone and rod photoreceptors, raising the
Glutamate is the main neurotransmitter of the vertical signal- possibility that they may play a role as retinal stem cells. Indeed,
ling pathway in the retina and it induces the expression of pro- primary human Müller glial cell cultures can generate photore-
genitor cell markers in adult murine Müller glia, both in vivo and ceptors and RGCs that have some reparative potential when
in vitro (Takeda et al., 2008), as well as promoting the proliferation transplanted into a damaged rodent retina (Jayaram et al., 2014).
of Müller derived progenitor cells (Ramirez and Lamas, 2009). Thus, Müller glia might be the ideal candidate for cell replacement
Glutamate can trigger de-differentiation in primary cultures of rat therapy to treat retinal degenerative diseases where vision loss is
Müller glial cells through a process that involves the activation of caused by retinal neuron loss (Ruilin et al., 2014).
NMDA and group II metabotropic glutamate receptors, and to some Müller glia with stem cell characteristics have been identified in
extent the participation of global DNA demethylation via gadd45b the adult human eye and although there is no evidence that they
(Reyes-Aguirre et al., 2013). regenerate the retina in vivo, they can be isolated and grown in
culture and can be induced to express retinal neuron markers
4. Specific functions in vitro (Lawrence et al., 2007). Both hMGSCs and RGC precursors
differentiate, as shown by the expression of RGC markers and the
4.1. Light guidance of Müller cells increase in cytosolic Ca2þ in response to nicotinic stimulation
(Singhal et al., 2012), the former generating functional newly born
In the mammalian retina, light has to pass the entire depth RGCs in response to BMP, FGF2 and the secretase inhibitor (DAPT).
before it arrives at the photoreceptors suggesting that specific This provides important evidence that hMGSCs can be differenti-
retinal tissue optics overcome light scattering, even in the typical ated in vitro into functional RGCs, providing a potential application
non-foveate vertebrate retina. It has been proposed that Müller in cell-based therapies to treat conditions in which RGCs are
radial glial cells might act as light-guiding elements (Labin and compromised (Becker et al., 2013). In addition, an immortalized
Ribak, 2010) and that to minimize light scattering, non-foveal human Müller glial cell line (MIO-M1) has been established that
Müller cells act as living optical fibres that guide light through can maintain its characteristics over 50 sub-culture passages (Limb
the inner retinal layers towards the photoreceptors, similar to glass et al., 2002).
fibre optics (Franze et al., 2007). The funnel-shaped endfoot of Cell-derived micro vesicles (MVs: recognized as important
Müller cells could serve as a light collector at the vitreal surface of components of cellecell communication) contain mRNAs, miRNAs,
the retina and the stem processes as light-guiding fibres. Between proteins and lipids, and they can transfer their bioactive contents
the vitreous and the retina there is a ‘soft coupling’ of the light path from parent cells to cells of other origins. Embryonic stem cell
that reduces light reflection at the inner retinal surface (Franze microvesicles (ESMVs) induce de-differentiation and the pluripo-
et al., 2007). Müller cells in the living retina minimize intraretinal tency of Müller cells after prolonged exposure, opening an avenue
light scattering, conserving the diameter of the beam that hits a to employ ESMVs in therapies to induce and support the retina's
single Müller cell endfoot. Thus, high intensity light reaches the endogenous regenerative capacity (Katsman et al., 2012).
individual photoreceptors. By minimizing scattering and enhancing The viral expression of ASCL1 is sufficient to activate a neuro-
the signal/noise ratio, Müller cells conserve the spatial distribution genic programme in mammalian Müller glia, both in dissociated
of light patterns in the propagating image. The ratio between the cultures and in the intact retina. The reprogrammed Müller glia
numbers of cones and Müller cells is roughly 1:1, indicating an differentiate into cells that morphologically resemble neurons, in
optimal coupling between the light-guiding units and the func- both their gene expression and their responses to neurotransmit-
tional light pattern-sensing units (the cones: Agte et al., 2011). The ters. Hence, stimulating neurogenesis in Müller glia with ASCL1
molecular basis of the light-guiding capacity of Müller cells remains could provide an alternative strategy for retinal repair after disease
to be determined, although bundles of IFs are arranged along the or injury (Pollak et al., 2013). Manipulating the ephrin-A2/A3
light path. In tissue edema, the neuroretina loses its transparency pathway may also represent an alternative means to promote the
and the milky opacity of the edematous tissue makes it likely that progenitor cell fate of Müller cells (see above: Ruilin et al., 2014).
the light guidance through Müller cells is deteriorated, resulting in Among the possible molecular and cellular mechanisms that
increased scattering (Reichenbach and Bringmann, 2013). control neuronal regeneration, p53-mediated cell cycle arrest may
play an important inhibitory role in Müller glia behaviour in the
4.2. Progenitor cells adult mouse retina (Ueki et al., 2012). Loss of p53 activity can
significantly enhance the proliferation potential of cells and facili-
Unlike neurons, glia have the lifelong capability to de- tate the reprogramming of Müller glia into retinal progenitor cells
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 25

(Zhao et al., 2014a). The tumour suppressor p53 may also strongly markers, and in their shape. For example, AQP4 is present in large
influence astrocytes (see above). The possible influence of Atoh7 amounts in the astrocyte plasma membranes when they are in
and glutamate in promoting the generation and proliferation of contact with microvessels. AQP4 influences many process (the
Müller glia derived progenitor cells has also been addressed above. volume of the extracellular space, waste clearance, potassium
Little is known about the role of BMPs in Müller glial prolifer- buffering, cell migration and calcium signalling: Nagelhus and
ation although a novel role for BMP in the control of Müller glial Ottersen, 2013) and significantly, its expression is altered in dia-
proliferation was recently described. Treating the mouse retina betic retinopathy (Qin et al., 2012). Astrocytes may play a role in the
with EGF stimulates the BMP signalling pathway in Müller glia, transport of exosomes with different cargos between neurons
causing them to release BMP7. This autocrine BMP release activates (Edelstein and Smythies, 2014; Kawikova and Askenase, 2014). In a
the BMP receptors on Müller glia, promoting Smad1/5/8 phos- laser-induced choroidal neovascularisation (CNV) model, retinal
phorylation and an increase in Id1. Hence, BMP can influence the astrocytes release exosomes containing anti-angiogenic com-
mitogenic pathway regulating Müller glial proliferation, which has pounds that protect the eye from the deleterious effects of neo-
implications for both vitreoretinopathy and regenerative ap- vascularisation (Hajrasouliha et al., 2013).
proaches to address retinal disease (Ueki and Reh, 2013). Differential expression of TrkA, the NGF receptor, led us to
The capacity of some astrocytes to enter a re-programmed state identify two sub-populations of astrocytes in the adult human and
of pluripotency and subsequently, to serve as progenitors capable of pig retina. While most astrocytes in the nerve fibre layer (NFL)
generating neurons suggests a reparative function that merits express TrkA, a small population in the INL do not express this
attention (Buffo et al., 2008; Wang et al., 2008; Wohl et al., 2012) In receptor (Ruiz-Ederra et al., 2003). A separate study of adult human
fact, the generation of neurons from astrocytes has been reported retinal astrocytes identified 3 different sub-populations character-
in vitro and it depends on specific transduction mechanisms ized by the expression of GFAP and Vimentin; GFAPþ/Vimentinþ,
mediated by key transcription factors like Pax 6 (Heins et al., 2002) GFAPþ/Vimentin and GFAP/Vimentinþ (Perez-Alvarez et al.,
or Mash 1 (Ding et al., 2014). 2008). The physiological significance of this, if any, is not yet
known but these data provide further strong evidence for the
5. Heterogeneity of retinal glial cells heterogeneity of retinal astrocytes.
In addition to the two sub-populations of astrocytes showing
5.1. Astrocyte heterogeneity distinct TrkA expression in the INL, we found another possible sub-
population of retinal astrocytes based on their function. Thus, only
The phenotypic heterogeneity among glial cell types is an a subpopulation of astrocytes were able to take up Fluorogold
emerging area of research in glial biology. Whilst most of the injected systemically in the blood vessels of the optic nerve of pig
attention to date has focused on astrocyte subpopulations in the eyes when incubated over night in oxygenated cultured medium.
brain white matter (Chaboub and Deneen, 2012), there is some Given their morphology, we propose that these specialized cells are
evidence to support the existence of glial subpopulations within the functional “perivascular astrocytes” described previously in the
the adult mammalian retina. At least nine types of astrocytes with human eye (Wolter, 1956). These astrocytes are preferentially
different morphological characteristics have been described in the connected to blood vessels while the rest of the astrocytes in the
normal brain. Moreover, morphologically undistinguishable astro- NFL of the retina connect mainly with RGC axons (Fig. 11).
cytes can exhibit different physiological behaviour (Matyash and Interspecies differences have been observed in the distribution
Kettenmann, 2010). The existence of different astrocyte pop- of astrocytes, which is very similar in the pig and human where two
ulations is reinforced by the finding that astrocyte reactivity takes types of astrocytes are found: stellate astrocytes that form cell
place in selected areas of the glaucomatous retina, these “micro- clusters; and elongated astrocytes located in the NFL (Ruzafa et al.,
domains” possibly serving as biomarkers and predictors of RGC 2014). By contrast, only stellate astrocytes are found in rodents.
health (Formichella et al., 2014). Interspecies differences in brain astrocytes have also been found
Astrocytes form a family of diverse populations adapted to their and differences exist between rodent and human astrocytes in
specific localization. Astrocytes in different brain regions have similar areas of the cortex, including differences in size, branching
distinct physiological properties, and they express various sets of area and calcium waves (Oberheim et al., 2009).
receptors and transporters, probably in response to their local
environment. Within the retina the situation is similar and when 5.2. Müller cell heterogeneity
primary retinal astrocytes are cultured, different types are evident
that vary in their expression of neurotransmitters and molecular Little is known about the heterogeneity of Müller cells, although

Fig. 11. Astrocyte heterogeneity. (A) Pig retinal astrocytes labelled with antibodies against GFAP. (B) Astrocytes labelled with fluoro-gold FG injected into the optic nerve blood
vessels of a pig eye ball that was preserved in vitro for 12 h in oxygenated medium. Note that in A there is at least one other type of astrocyte not associated with the blood vessels
and located in parallel to the RGC axons.
26 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

only 30% of the Müller cells in vitro express class II MHC antigen,
suggesting that these might be involved in retinal immune re-
actions (Roberge et al., 1985). In an attempt to define the tran-
scriptome of Müller glia, certain genes were seen to be expressed
heterogeneously among the Müller glial cells including common
housekeeping genes like b-actin, b-2-microglobulin, heat shock
protein 1 b and TATA box binding protein (Roesch et al., 2008). It is
clear that molecular differences exist in Müller cells because the
homeodomain containing transcription factor Chx10 is expressed
only in a subset of Müller glia (Rowan and Cepko, 2004). Hetero-
geneous gene expression among Müller glial cells was observed
over a large cluster, yet further characterization of the glial genes
identified will be required to clarify the diversity among Müller
glial cells and to shed light on their potential functional diversity
(Roesch et al., 2008). A comparison of the Müller glia transcriptome
with that of other glial cell types in the CNS failed to reveal common
clusters of gene expression (Bachoo et al., 2004). This might be due
to technical differences or it might reflect a distinct genetic
constitution. However, the expression of some transcripts was
common to Müller glia and other glial cell types, and some tran-
scripts enriched in Müller glia are also enriched in astrocytes
include peroxiredoxin 6, Basigin (also called CD147), caveolin, CD9
antigen, carboxypeptidase (Roesch et al., 2008).
Isolated cells in vitro can respond to certain conditions by
rapidly changing their protein expression and transforming to a
proliferative phenotype. While vimentin remains stably expressed
in these cells, GFAP expression, which is probably induced during
their isolation from the intact retina, is rapidly lost in vitro (Fischer
et al., 2004). This is consistent with reports from the chick retina
Fig. 12. Müller cell heterogeneity in vitro. Confluent pig Müller cells monolayer
where only the sub-population of GFAP negative retinal Müller glia
cultured for 10 days and labelled with (A) antibodies against-GFAP (green) and (B)
become proliferative following NMDA-induced retinal damage, glutamine synthetase (GS, red). C) Co-localisation of both antibodies plus nuclei
possibly indicating that the primary retinal Müller glia represent stained with DAPI (blue). Note that even when there are many Müller cells in a field,
only a glial subpopulation (Hauck et al., 2014). there are only two cells with strong GS-immunofluorescence: one with strong GFAP
Other evidence for phenotypic diversity among Müller glia in- fluorescence and another with an intermediate strength of GS fluorescence but
negative for GFAP (top right). Scale bar ¼ 50 mm.
cludes the variation in nestin expression in Müller cells after retinal
detachment from adult rats (Luna et al., 2010). A different post-
injury response may be elicited in individual Müller cell pop- which all cells in a tissue participate. Likewise, when things go
ulations and since nestin is a marker of immature Müller cells, wrong due to trauma or disease, the community of retinal cells
perhaps sub-populations exist with enhanced regenerative prop- responds and collaborates to restore a healthy state. However, this
erties. Furthermore, a gene expression analysis of Müller cells in a cannot always be achieved and the damage may seriously and
mouse model of retinitis pigmentosa revealed significant inherent irreversibly affect retinal function. Indeed, many pathological
heterogeneity amongst individual wild type cells, with further di- conditions may alter the integrity of the retina. In the case of
versity amongst individual cells post-injury (Roesch et al., 2012). neurodegenerative diseases the role of microglia is variable, and
Heterogeneity in GS or GFAP expression was observed in pri- both their pathogenic engagement and neuroprotective role are
mary Müller cell cultures from adult rat or pig, and cells expressing entangled (i.e.: capacities and limits). Thus, both these roles must
high amounts of GS or GFAP lie close to others that do not express be fully understood to adequately design external interventions
these markers (Fig. 12). This was evident in both pure glial cell aimed at preventing, alleviating or eliminating the damage (Prinz
cultures as well as in Müller-RGC co-cultures. In addition, RGCs et al., 2011).
prefer to grow on Müller cells that overexpress GS, their regener- Retinal damage results from the action of different etiological
ating axons following a very precise path over the surface of Müller agents, such as elevated intraocular pressure, altered vascularisa-
cells that contain GS and avoiding the surrounding cells without GS tion, ageing, genetic defects. In different retinal neurodegenerative
(Fig. 13). This heterogeneity among Müller cells would perhaps also diseases the cell signals and changes, like the inflammatory re-
reflect different glialeneuronal interactions that may have neuro- sponses, exhibit both similar and singular characteristics, and
protective effects on retinal neurons. therefore, therapeutic interventions must be based on knowledge
Reactive Müller cells upregulate GS, possibly as a neuro- of the common and distinct patterns of the disease (Cuenca et al.,
protective mechanism to limit glutamate neurotoxicity (Gorovits 2014). The effect of a glial response contralateral to the damaged
et al., 1997). Whilst this is not a direct interaction between glia eye has been demonstrated and it is very possible that inflamma-
and RGCs, it does represent an important effect of glia on RGC tion caused by the way to increase intraocular pressure with laser
survival and retinal homeostasis. could be the cause of such contralateral reactions (Gallego et al.,
2012; Ramirez et al., 2010; Rojas et al., 2014). It is thus very
6. Glia and retinal pathologies important to pay attention to the inflammation caused in the
different experimental models of eye damage since in addition to
Microglia perform valuable activities in the retina under phys- elevation of intraocular pressure, it is possible that other retina
iological conditions by interacting with other cell populations in damage can be caused to the retina.
the retinal layers. Maintenance of homeostasis is a complex task in
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 27

retinal cell populations, including neurons, glial cells and reduced


myeloid populations (for a review see (Grigsby et al., 2014).
As a consequence of consistent high blood glucose levels in DR,
AGE appear (molecules that are non-enzymatically glycated in the
presence of aldoses), dyslipemia occurs and there is an increase in
ROS that activates retinal microglia. As a result, the microvascula-
ture, as well as other glia and neurons suffer damage (Fletcher et al.,
2005; Grigsby et al., 2014; Wong et al., 2011). Various experimental
animal models of DR have provided useful information about the
molecular events involved in the disease. Alloxan or STZ (strepto-
zotozin) are toxic agents that specifically destroy pancreatic islet
beta cells and induce diabetes (Larger et al., 1995; Mansford and
Opie, 1968), while Goto Kakizaki (GK) rats are a non-obese model
for spontaneous type 2 diabetes developed by select breeding
(Kitahara et al., 1978). The retina of these experimental animals can
be used to analyse the role of microglia, in vivo and in vitro: their
migratory, secretory and growth activities, as well as their neuro-
protective capacity. Also, human subjects can be used to study the
disease, although with evident limitations.
Microglial activation in DR has been observed at early stages of
this progressive disease and an increase in microglial density oc-
curs in the GCL and NFL, with a decrease in the IPL. The alignment of
microglial processes along the optic nerve is also disrupted (Chen
et al., 2014). It is noteworthy that the precise mechanism(s) of
glial activation in DR and other retinopathies are not fully under-
stood. A few examples illustrate the specific and partial relevance of
several actors. The accumulation of AGA (Amadori-glycated albu-
min), an AGE product in early DR induces TNF-a secretion by
microglia (Ibrahim et al., 2011). Moreover, in an experimental rat
model of diabetes, activated microglia more intensely express TNF-
a and IL-1b mRNA, and interestingly, minocycline (a tetracycline
antibiotic with anti-inflammatory activity) produces an approxi-
mate 50% decrease in the transcripts encoding these two cytokines
(Krady et al., 2005). Another agent analyzed in DR is aldose
reductase (an enzyme that catalyzes the synthesis of polyols),
which is activated in DR (Kinoshita, 1986). Aldose reductase par-
ticipates in the activation of retinal microglia and its inhibition or
Fig. 13. Preference of RGCs grown in vitro on Müller cells that express GS strongly. (A) its genetic deficiency significantly reduces the inflammatory
Rat RGCs were cultured on confluent pig retinal Müller cells monolayer where the
nuclei are labelled with DAPI (blue), the glutamine synthetase is labelled in Müller
changes caused by microglia activation (Chang et al., 2014),
cells (GS, green) and the RGCs are stained with antibodies against-ßIII Tubulin (red). ameliorating the degeneration of capillaries and the generation of
(B) RGCs have a preference for those Müller cells that express GS strongly. Scale superoxide anions found in DR (Tang et al., 2013). The participation
bar ¼ 50 mm. of signalling pathways related to the microglial adenosine receptor
A2A has also been documented (see Santiago et al., 2014).
In cases of disease or trauma, astrocytes react and become
6.1. Diabetic retinopathy activated (Pekny et al., 2014; Yang et al., 2014), undergoing a series
of changes (proliferation/migration, hypertrophy, GFAP expression,
Diabetic retinopathy (DR) is a long term complication of type 1 secretion of pro-inflammatory signals that include IL-6, MCP-1 and
and type 2 diabetes. In both the proliferative (angiogenesis MIP-2alp) aimed at tackling the insult in cooperation with other
observed) and non-proliferative (disrupted blood vessels detected) glial cells (Nahirnyj et al., 2013). As a consequence of astrocyte
forms of DR, activation of a plethora of signalling pathways ulti- activation, RGCs receive both positive and negative signals, and the
mately results in retinal damage and visual loss (Grigsby et al., disturbance of the homeostatic scenario triggers the death of
2014). Dysregulated metabolites (including glucose, amino acids, resident retinal cells. Astrocytes also respond to aggression and in
lipids, amines, vitamins, minerals, etc) disturb the homeostasis of diabetes for example, high glucose levels alter astrocyte inflam-
growth factors, neurotrophic factors, cytokines and adhesion mol- matory cytokine expression, activating NF-kB and increasing
ecules, consequently affecting capillary permeability and cell oxidative stress (Shin et al., 2014). In glaucoma and other degen-
turnover. Therefore, many related and unrelated biochemical erative retinal diseases, oxidative stress and death by apoptosis are
pathways are involved in the complex pathophysiology of DR (Ola known features, as are astrocytic TNF-a/TNFR signalling, nuclear
et al., 2012), which explains why a precise picture of the mecha- factor (NF-kB) activation, and p53 and caspase 3 upregulation
nisms related to the onset and progression of the disease is not yet (Rogers et al., 2012; Tezel et al., 2012). In neurodegeneration glial
clear. Moreover, individual genetic susceptibility and habits influ- activity also deteriorates.
ence the development and morbidity of the disease, which may Elevation of histone acetylation in Müller cells plays an impor-
eventually lead to partial or total loss of vision as a consequence of tant regulatory role in the inflammatory response during diabetes
retinal destruction and/or detachment (Grigsby et al., 2014). (Wang et al., 2012). It has also been proposed that in early phases of
Clearly, DR has an important inflammatory component, where diabetic retinopathy, Müller glial activation does not contribute to
microglia participate as both prey and enforcer, together with other neurodegeneration but that it may provoke neuroprotection
28 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

against high glucose induced neurotoxicity through a mechanism mechanisms are at play that involve both structural and diffusible
involving phosphorylation of extracellular signal-regulated kinase elements. Microglial fractalkine may be a key element of the
1/2, pERK1/2: (Matteucci et al., 2014). Since in vitro high glucose neurotoxicity exerted by microglia in an experimental model of
stimulation of Müller cells increases VEGF secretion, the specific induced glaucoma since suppression of its CX3CR1 receptor
ERK1/2 inhibitor, U0126, blocked ERK1/2 phosphorylation and induced more extensive RGC death when IOP was elevated (Wang
thus, impaired the Müller cells secretion of VEGF that is implicated et al., 2014). A relationship between the release of cytokines (IL-6,
in the development of diabetic retinopathy (Ye et al., 2010). Thus, IL-1, TNF-a, etc) and the initiation of apoptosis has been established
targeting ERK1/2 signalling in Müller cells, through the upstream (Gonzalez-Scarano and Baltuch, 1999). The intervention of neuro-
signalling pathway or nuclear transcription factors of VEGF secre- toxic agents, alone or in collaboration, has been studied in vivo and
tion, could represent a promising future treatment for diabetic in vitro, in different animal models or in humans. Thus, while there
retinopathy. is abundant information, the immune system is complex, making it
difficult to define the main causes and the accompanying phe-
6.2. Glaucoma nomena. RGC death in glaucoma probably depends on the
concurrence of several stressors (for instance ROS, glutamate,
Glaucoma encompasses several diseases that destroy RGCs and mitochondrial impairment…), as well as ill-tuned or adapted
their axons in the optic nerve through apoptosis (Garcia-Valenzuela cellular responses (Rieck, 2013). Thus, full resolution of this “puz-
et al., 1995), provoking progressive loss of the visual field and zle” will require more effort.
blindness (Mi et al., 2014; Quigley, 2011; Rieck, 2013). The aetiology Gene expression dependent on TGF-b2 activation in astrocytes
and pathophysiology of glaucoma is not well known, although cultured from the optic nerve head is reduced after antioxidant
several risk factors and mechanisms that influence the disease have treatment, indicating that antioxidants may have a place in the
been found. Among these factors, the elevation of intraocular treatment of neurodegenerative diseases like glaucoma (Yu and
pressure (IOP) is particularly important, while other factors have Welge-Lussen, 2013). Astrocytes undergo morphological changes
also all been associated with the disease, like: low diastolic ocular related to the damage produced in RGCs (Lye-Barthel et al., 2013).
perfusion pressure (calculated as 2/3 mean arterial pressure-IOP), Hypertrophia of astrocytes was observed in the optic nerve in a rat
other ocular hemodynamic parameters, systemic diseases such as in vivo model of transient intraocular hypertension, as was early
sustained high blood pressure or diabetes, ethnic group, age, up-regulation of phosphorylated STAT3 (signal transducer and
myopia, migraine, the nutritional state of RGCs (Hernandez et al., activator of transcription protein 3), a key transcription factor that
2009). regulates the expression of many genes and that may play a crucial
Human glaucoma can be classified into three main categories: role in the activation of astrocytes after an increase in IOP (Zhang
primary open angle glaucoma (POAG), primary closed-angle glau- et al., 2013). Indeed, it is proposed that the damage observed in
coma (PCAG) and primary congenital glaucoma (PCG: Ahonen et al., optic nerve axons after an increase in IOP is secondary to the
2014; Quigley, 2011). Each of these has specific and common damage induced in astrocytes, which suggests they are the primary
characteristics, related to alterations of the aqueous humour site of damage in early stages of glaucoma (Dai et al., 2012a).
outflow and abnormalities in the anterior chamber angle (Ahonen In a mouse model of hereditary glaucoma (DBA/2J), an increase
et al., 2014; Liu et al., 2013). The structural changes in the outflow in BDNF in hypertrophic astrocytes was concomitant with a deficit
system involve a thickening of the trabecular beam, cell death in in anterograde axonal transport prior to RGC synapse loss (Crish
the cribriform layer, accumulation of extracellular material and et al., 2013). Hence, astrocytes may respond to the diminished
plaque formation (Fuchshofer et al., 2006; Rohen et al., 1981). axonal transport by taking up BDNF from target neurons to mini-
However, a clear relationship between elevated IOP, structural al- mize the damage. It is very possible that the same mechanism will
terations to the Schlemm's canal cells that respond to local me- work at the onset of glaucoma and that after the first days of high
chanical changes, and the preponderant intervention of the IOP, the homeostatic mechanisms in the retina will be altered and
mechanisms related to the onset and progression of the disease are glial cells will then orchestrate the events that will provoke the
not yet fully known (Galdos and Vecino, 2011; Overby et al., 2014; death of certain RGCs. It is unclear why not all RGCs die at the same
Vecino et al., 2015a) Interestingly, therapies that lower IOP slow the time in Glaucoma, however, the glial environment of the vulnerable
progress of different types of glaucoma, yet other treatments aimed RGCs may be different and thus, these RGCs might respond to the
at avoiding cell death are required. malevolent signals in a more aggressive manner.
Due to the limitations faced when studying human individuals, In rats where glaucoma was induced by cauterization of epis-
animal models of induced or spontaneous glaucoma are efficient cleral veins, in addition to RGC death changes in amacrine and bi-
tools to discover and define the aetiological factors, pathological polar cells were found, as well as astrocyte activation in the
mechanisms, and the extent and consequences of RGC death. damaged but not the contralateral eye (Hernandez et al., 2010). The
Moreover, reliable animal models are needed to assess the thera- function of astrocyte and Müller cell activation is unknown as yet,
peutic value of drugs in preclinical trials (Smedowski et al., 2014). although one week after the IOP elevation we observe more
However, since differences in retinal anatomy and physiology be- astrocyte gliosis that decreased thereafter (Hernandez et al., 2010).
tween animal species exist, and the procedures used to obtain such It is possible that the astrocytes detect the increase in pressure
experimental models may vary, it is necessary to ascertain the directly, or through microglial or Müller cells, and they react by
advantages and limitations of a particular model, and carefully liberating neurotrophic signals that could help RGCs to resist the
choose the pertinent one in function of the damage caused to the insult. It is also possible that glial cells may provide mechanical
optic nerve and the reproducibility (Smedowski et al., 2014; Vecino, support. Another possible function of reactive astrocytes might be
2008). related to repairing the damage to the blood retina barrier (Bush
Neuroinflammation is a common aspect of many neurodegen- et al., 1999; Voskuhl et al., 2009). However, the detrimental side
erative diseases, including glaucoma, and microglia participate in of astrogliosis is that reactive astrocytes may potentially exacerbate
their pathology, although the precise extent to which they influ- the glaucomatous process by overexpressing toxins, producing a
ence each disease is not yet clear. Microglial activation has been direct toxic effect on RGCs. Müller cells overexpress GS, especially
observed both in eyes affected by ocular hypertension and in the in their inner and outer regions, while the vimentin cytoskeleton is
healthy contralateral eyes (Rojas et al., 2014), indicating complex also partially disorganized (Fig. 14). AQP9 expression is reduced in
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 29

Fig. 14. Glial cells in glaucomatous retinas. Comparative distribution of glial cells between control (A, C, E) and glaucomatous (B, D, F) rat retinas. A and B are whole-mount retinas
where the astrocytes were labelled with antibodies against-GFAP (green). Note the change in astrocyte morphology in the glaucoma retina (in B) compared to the control (A).
Vimentin (C, D) and glutamine synthetase (E, F) immunoreactivity (red) is evident in Müller cells. Nuclei were DAPI labelled (blue). Note the change in vimentin expression that is
more disorganized (D) than in the control eye (C) and GS expression is increased in F with respect to the control eye in E. Scale bar ¼ 50 mm.

RGCs of glaucomatous eyes, while glaucoma induces AQP7 involve the mild release of plasma components into the affected
expression in the Müller cell endfeet. Alternatively, the expression tissue and if the condition is sufficiently severe (e.g., infection or
of both AQP7 and AQP9 is reduced in the ciliary body of glaucom- injury), an acute inflammatory response ensues. This is charac-
atous eyes. Therefore, AQP expression seems to play some role in terized by the recruitment of neutrophils and specialized subsets
glaucoma (Tran et al., 2014). of monocytes from the circulation that help to protect the host
Under normal conditions, resident macrophages maintain tis- from infection, promoting tissue repair and restoration of ho-
sue homeostasis by removing dead cells and other debris, and by meostasis (Medzhitov, 2008, 2010). This could be the case of
producing growth factors. However, under noxious conditions the experimental damage to the eye where strong inflammation could
cellular stress response is activated that results in cell- activate resident cells and the immune system. Since astrocytes
autonomous adaption. However, communication between and microglial cells have receptors for inflammation, this could
stressed cells and other cells in a tissue may also occur, including explain the contralateral response to damage in experimental
resident macrophages. If the nature or the extent of stress is such models of hypertension when laser damage causes intense
that it affects not only individual cells but also the entire tissue inflammation of the eye. It could also be relevant to situations in
(e.g. hypoxia or hyperthermia), then a tissue-level stress response humans where inflammation of different sources (uveitis, trauma,
or parainflammation is elicited by the resident macrophages. etc…) could activate the immune system and affect the contra-
Depending on the strength of the noxious conditions, this lateral eye. We must pay close attention to the different experi-
response can involve the recruitment of different types of circu- mental models to induce hypertension and glaucoma (Ruiz-
lating inflammatory monocytes. Parainflammation may also Ederra et al., 2005; Vecino and Sharma, 2011).
30 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

6.3. Age-related macular degeneration (AMD) RPE cells appears to be responsible for the initiation and progres-
sion of AMD.
Advanced age is a risk factor for numerous pathologies, collab-
orating with other intrinsic and extrinsic cues that provoke the 6.4. Retinitis pigmentosa (RP)
initiation of a given disease. The brain and the retina suffer age-
related deterioration/degeneration (Harry, 2013), and to discover RP is a group of heterogeneous inherited diseases in which the
the importance of age in this process, which cannot be avoided or primary defect affects the retina, alone or in association to other
changed, it is paramount to unravel the influence of the other main anomalies affecting different organs. RP is characterized by
aetiological elements (for example life-style and environmental photoreceptor degeneration and progressive blindness, yet the
factors) on the initiation and progression of neurodegeneration. molecular and cellular mechanisms involved in photoreceptor
Microglia populate the retina early in the embryonic period and death are not completely understood. Only when the mechanisms
their persistence during adulthood depends on self-renewal underlying this damage are known can effective treatments follow.
through multiple divisions (Ginhoux et al., 2010). Microglial sen- As in other retinal pathologies, inflammation plays a fundamental
esce has been observed in healthy aging rats in conjunction with role in RP and it has been associated with mutations that affect
telomere shortening and reduced telomerase activity (Flanary et al., more than 70 genes related to different retinal functions (photo-
2007). Other observations indicate that old microglia suffer age- transduction, cell cycle proteins, ciliary proteins, structural pro-
related alterations in their dynamic capacities (motility, branch- teins, transcription factors, see RetNet at https://sph.uth.edu/
ing, length of processes, etc.) and that they are less responsive to retnet/disease.htm). However, in some cases the genetic defect is
external stimuli like ATP (Damani et al., 2011). Hence, the homeo- yet to be defined. Recently, a dominant mutation in the gene that
static role of microglia can be seriously altered with age (Karlstetter encodes HK1 (hexokinase 1, an enzyme that catalyzes the phos-
and Langmann, 2014). phorylation of glucose in the glucolytic pathway) was found to
AMD is a neurodegenerative disorder in which age plays a cause RP (Sullivan et al., 2014).
principal role. The most common form of this disease is the atro- Microglia sense the environment and they rapidly respond to
phic or “dry” form that is characterized by RPE degeneration and disturbances by trying to protect neurons against insults (neuro-
central retina photoreceptor death. In the exudative or “wet” form, protection). However, the persistence of damage may activate the
choroidal neo-vascularization is observed (Ma et al., 2009; Zhao neurotoxic phenotype of microglia, in an attempt to limit the
et al., 2014b). In healthy retinas, microglia are located in the inner damage (Langmann, 2007). The early intervention of microglia in
retina and they are scarce or non-existent in the outer retina and degeneration suggests that they play an active role in pathological
retino-choroidal zone. However, in AMD the distribution of mechanisms. In vitro and in vivo studies have characterized
microglia changes dramatically. Activated microglia migrate to sub- microglial behaviour in RP and other degenerative retinal diseases
retinal locations and in some cases, they come into close contact by studying their temporal transcription profiles, secretomes,
with drusen deposits (Ma et al., 2009, 2012). In an in vitro model phenotypic changes and migration (Karlstetter et al., 2010;
where microglial cells were co-cultivated with primary RPE cells Langmann, 2007). Using genetic mouse models and in vitro ap-
and in an in vivo mouse model where microglia were transplanted proaches, the complex and multiple effects of microglia as key el-
into the sub-retinal space, microglia appear to be responsible for ements in RP and other examples of inherited retinal degeneration
early alterations to RPE cells, and the initiation of a pro- can be illustrated. An analysis of microglia transcripts and
inflammatory and pro-angiogenic environment (Ma et al., 2009). morphology in a retinoschisin-deficient mouse that suffers a severe
Therefore, the translocation of microglia to the outer retina may be photoreceptor dystrophy indicated that microglia act in early stages
an essential event that triggers chronic inflammation related to the of the diseases (Gehrig et al., 2007; Karlstetter et al., 2010). In a
progress of AMD. mouse RP model, rd (carrying a spontaneous defect in the b subunit
Genetic models provide valuable information to analyse the of rod cGMP-phosphodiesterase), early activation of microglia was
pathology of age-related neurodegeneration and to study the again seen by analysing the expression of microglial chemokine
participation of microglia in such processes (Karlstetter et al., 2010). transcripts and the secretion of TNF-a, which appeared prior to or
For example, knock-out mice develop some characteristics of AMD concomitant with photoreceptor death (Zeng et al., 2005).
at old ages, such as Ccl2 (CCL2 is a chemokine) and Cx3cr1 deficient Microglia-mediated neurotoxicity in photoreceptors is also sup-
animals, and the Ccl2/Cx3cr1 double knock-out developed evident ported by the activation of microglial NADPH oxidase, which gen-
lesions consistent with AMD earlier: RPE atrophy, photoreceptor erates ROS (Zeng et al., 2014). When the rd10 mouse model of RP is
death, neovascularisation, and migration of microglia to the outer crossed with a Cx3cr1GFP/GFP mouse, microglial activation was
nuclear layer and sub-retinal spaces (Tuo et al., 2007). In a knock- evident as an early phenomenon in RP and Cx3cl1/Cx3cr1 signalling
out mice deficient in Nrf2 (nuclear erythroid 2-related factor e a significantly ameliorated microglia neurotoxicity (Peng et al., 2014).
transcription factor that regulates the expression of protective Moreover, administration of minocycline for several days inhibited
proteins against oxidative stress), the abnormal presence of microglia activation, reducing photoreceptor death and improved
microglia, regulatory T cells and IL-17 (interleukin 17) producing T vision in rd10 mice (Peng et al., 2014). Chemical destruction of
cells in the retina and at sub-retinal locations has been documented microglia in rd10 mouse retinal explants with clodronate-
(Zhao et al., 2014b). Therefore, the destruction of the RPE in early containing liposomes slowed photoreceptor death, yet the neuro-
AMD favours the infiltration of circulating blood cells that protective effect of IGF-I was attenuated when microglia were
contribute significantly to the pathology and progression of AMD. depleted. These findings suggest a misbalance in the microglial
The AhR (a nuclear aryl hydrocarbon receptor) signalling pathway neuroprotective/neurotoxic intervention in this retinal dystrophy
is a defence mechanism that detoxifies metabolic waste molecules (Arroba et al., 2011, 2014). In a rat model of RP bearing a mutation in
and xenobiotics, and its activity in human RPE cells diminishes with the rhodopsin-encoding gene (P23H), microglial cells were more
age (Hu et al., 2013). A knock-out mouse model deficient in AhR abundant in all retinal layers and they were also present in the sub-
(Ahr/) exhibits phenotypic characteristics of AMD, indicating that retinal space. Treatment of these rats with taurodeoxycholic acid
this signalling pathway may hinder the appearance of AMD (Hu prevented microglial activation and delayed photoreceptor death
et al., 2013; Kim et al., 2014). Accordingly, a complex collabora- (Noailles et al., 2014).
tion of mechanisms dependent on immune competent cells and Not only are microglial cells activated in RP but Müller cells are
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 31

more activated in the early stages of the disease (young animals) in receptor was expressed in Müller cells and it increased in the
a natural model of RP in dogs, the RP65 mutant. Müller cells also mutant retina, with expression in radial Müller fibres and sur-
respond to the ongoing disease process, although this response was rounding RGCs. Both vimentin and GFAP expression also increased,
not uniform against all the proteins examined. The low-affinity p75 although more prominently in younger animals than in adults. By

Fig. 15. Retinitis pigmentosa (RP). Comparison of the glia in controls (A, D, G, J) and adult dog models of RP, RPE65/ (B, E, H, K) and adult dog RPE65/ treated with adeno-
associated virus (AAV) vector containing a functional copy of the RPE65 gene (AAV-RPE65: C, F, I, L). (AeC) RGCs are labelled with antibodies against-BDNF (green) and the
Müller cells are stained with antibodies against p75 (red). The BDNF expression in RGCs decreased in affected retinas and increased after treatment. (DeF) GFAP expression (red)
increased in affected dogs and RPE65 expression (green) is lost in affected animals but it is recovered in treated dogs. (GeI) CRALBP was expressed in Müller cells (green) and it
appears disorganized in the affected and in the treated dogs. Increased expression of CRALBP was found in RPE in affected dogs (H) and it returned to control levels in gene-therapy
treated dogs (I). (JeL) Vimentin is expressed in Müller cells (red, DeF), and it was enhanced in the affected (K) and treated (L) dogs. Nuclei were stained with DAPI (blue). Scale
bar ¼ 50 mm.
32 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

contrast, CRALBP did not show an initial increase and it was 7. Conclusions, remarks and future directions
expressed below control levels in adult mutant dogs. It is clear that
Müller cells respond to outer retinal disease. The response was The retina contains three major classes of glia, Müller cells, as-
mild, transient and more intense in the younger than in the older trocytes and a resident subpopulation of microglial cells. All are
mutant retinas, possibly indicating a lack of progressive retinal integrated with the neurons that form the retina and allow a correct
degenerative changes in the period analyzed. However, only a functioning of the retina. Müller cells, astrocytes and microglia are
limited number of antibodies were analysed, two of which were all capable of phagocytosis, responding to immunological damage
directed at IFs. It is possible that other proteins (e.g., GC and car- and they interact with neurons, thereby modulating the synapses.
bonic anhydrase C) would respond differently as they are known to Moreover, the three cell types play a role in the metabolism of
decrease in experimental retinal detachment in adult cats (Lewis neurons, which is affected by their distribution and physiological
et al., 1994). state. Since the neurons in the retina are highly specialized, their
Therapeutic approaches in RP and other hereditary diseases metabolic demands are also very specific, and glial cells sur-
involving retinal degeneration might aim to restore the gene defect rounding these neurons provide different nutrients. Therefore, it is
by gene-replacement (Beltran et al., 2014) or stem-cell trans- crucial that when exploring the changes within the retina, an in-
plantation (He et al., 2014). Retinal degeneration might also be tegrated approach should be adopted that considers glialeneuron
palliated or slowed by controlling the neuroinflammatory frame- interactions, in order to fully understand the retina in health and
work. This can be achieved with either immune-suppressant or disease.
immunomodulatory drugs (Karlstetter et al., 2010). Another strat- The source and age of glia is an important factor when deter-
egy includes the administration of protective molecules, such as mining the support that astrocytes or Müller glia can offer to RGCs.
rod-derived cone viability factors, to preserve cone-dependent Much of the research concerning the molecular mechanisms of RGC
central vision after rod death (Leveillard et al., 2014). In addition, processing in association with retinal glia has been carried out
the application of sub-retinal photovoltaic arrays that restore using embryonic tissue, the rational being that axon outgrowth
photoreceptor function can be applied when the inner retinal cir- during development could provide insight into the deficits under-
cuits are preserved (Light et al., 2014). Indeed different therapeutic lying unsuccessful adult regeneration. Whilst much information
approaches should target the specific features of each type of RP can be gleaned from these studies, there are limitations to the
and hereditary retinal degeneration. comparisons that can be made, since there is a wealth of evidence
In diseases where the outer part of the retina is affected, showing that adult RGCs and glia have a highly altered molecular
Müller cells sense the damage in the outer retina and transmit the repertoire when compared to their embryonic and postnatal
information to the inner retina. This event can be studied in dogs counterparts. Therefore, it is imperative that attention is paid to
carrying a mutation similar to Leber Congenital Amaurosis (LCA), study adult RGCs and adult retinal glia.
a specific type of RP. This disease comprises a group of childhood- The other limitation to studies in early periods is the use of
onset, autosomal recessive retinal diseases that result in severe astrocytes from non-retinal sources. We are now only just begin-
visual impairment or blindness (den Hollander et al., 2008). LCA ning to understand the complexity of glial heterogeneity and the
is caused by mutations in the RPE65 gene, which encodes the roles of different subpopulations within a single tissue, and further
65 kDa RPE-specific isomerase involved in visual pigment research on the adult retina is required to explore the possibility
regeneration (Gu et al., 1997). The RPE65 protein is essential to that some populations of retinal glia are more supportive of RGCs
maintain retinal function and photoreceptor viability, and for than others. The molecular changes that take place in the adult
retinoid cycling between the RPE and photoreceptors (Saari, retina after damage occur rapidly, within minutes or hours. How-
2000). Mutations in RPE65 occur naturally in dogs (Aguirre ever, the return to a normal situation can take years, although some
et al., 1998), and gene therapy studies in these dogs (Acland restoration of function can be achieved in a shorter period.
et al., 2001) has led to the first gene therapy undertaken in
humans (Hauswirth et al., 2008). We studied the molecular Acknowledgements
changes that take place in the retina of the mutant dogs as they
mature (Hernandez et al., 2010) and found activation of astrocytes We acknowledge the support of The Glaucoma Foundation
in the GC layer, even when the damage to the retina originated in (TGF), Fundacio n ONCE (name of the price: International price
the RPE. Hence, Müller cells appear to sense the changes in the ONCE IþD in biomedicine and new technologies for blind people
outer part of the retina and communicate these to cells located in 2004), Fundaluce and the Grupos Consolidados del Gobierno Vasco
the inner part of the retina. Moreover, astrocytes located in the (IT437-10) to E.V., and that of the IKERBASKE, Basque Foundation of
neighbourhood of RGCs may also sense that synaptic trans- Science, Bilbao, Spain to S.C.S.
mission was disturbed and react to this malfunction. Interest-
ingly, after gene therapy with an adeno-associated virus to References
restore the mutated RP65 gene (AAV_RPE65), and even when
functional recovery was achieved (Acland et al., 2001), the glial Acland, G.M., Aguirre, G.D., Ray, J., Zhang, Q., Aleman, T.S., Cideciyan, A.V., Pearce-
Kelling, S.E., Anand, V., Zeng, Y., Maguire, A.M., Jacobson, S.G., Hauswirth, W.W.,
cells remained activated 6e24 months after treatment (Fig. 15). Bennett, J., 2001. Gene therapy restores vision in a canine model of childhood
Only the CRALBP immunoreactivity in the RPE returned to normal blindness. Nat. Genet. 28, 92e95.
levels after gene therapy, while Müller cells remained activated, Adijanto, J., Du, J., Moffat, C., Seifert, E.L., Hurle, J.B., Philp, N.J., 2014. The retinal
pigment epithelium utilizes fatty acids for ketogenesis. J. Biol. Chem. 289,
as evident by examining vimentin labelling. Müller cells can 20570e20582.
change the morphology of their membranes, adapting to the Agte, S., Junek, S., Matthias, S., Ulbricht, E., Erdmann, I., Wurm, A., Schild, D., Kas, J.A.,
changes in the damaged cells surrounding them. After this Reichenbach, A., 2011. Muller glial cell-provided cellular light guidance through
the vital guinea-pig retina. Biophys. J. 101, 2611e2619.
damage and retinal function is restored, as is this case with this
Aguirre, G.D., Baldwin, V., Pearce-Kelling, S., Narfstrom, K., Ray, K., Acland, G.M.,
gene therapy, Müller cells should return to the original state in a 1998. Congenital stationary night blindness in the dog: common mutation in
plastic-like process, although has yet to be confirmed (after the the RPE65 gene indicates founder effect. Mol. Vis. 4, 23.
submission of this review and during the process of revision a Ahonen, S.J., Kaukonen, M., Nussdorfer, F.D., Harman, C.D., Komaromy, A.M., Lohi, H.,
2014. A novel Missense mutation in ADAMTS10 in Norwegian Elkhound pri-
book on retinal glia appeared where details of each cell type are mary glaucoma. PLoS One 9, e111941.
addressed: Reichenbach and Bringmann, 2015). Ai, J., Liu, Y., Sun, J.H., 2013. Advanced glycation end-products stimulate basic
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 33

fibroblast growth factor expression in cultured Muller cells. Mol. Med. Rep. 7, insights from animal models. Brain Res.
16e20. Blackshaw, S., Harpavat, S., Trimarchi, J., Cai, L., Huang, H., Kuo, W.P., Weber, G.,
Ajami, B., Bennett, J.L., Krieger, C., Tetzlaff, W., Rossi, F.M., 2007. Local self-renewal Lee, K., Fraioli, R.E., Cho, S.H., Yung, R., Asch, E., Ohno-Machado, L., Wong, W.H.,
can sustain CNS microglia maintenance and function throughout adult life. Cepko, C.L., 2004. Genomic analysis of mouse retinal development. PLoS Biol. 2,
Nat. Neurosci. 10, 1538e1543. E247.
Akhter, N., Nix, M., Abdul, Y., Singh, S., Husain, S., 2013. Delta-opioid receptors Bonnet, D., Garcia, M., Vecino, E., Lorentz, J.G., Sahel, J., Hicks, D., 2004. Brain-
attenuate TNF-alpha-induced MMP-2 secretion from human ONH astrocytes. derived neurotrophic factor signalling in adult pig retinal ganglion cell neurite
Invest. Ophthalmol. Vis. Sci. 54, 6605e6611. regeneration in vitro. Brain Res. 1007, 142e151.
Allen, N.J., Bennett, M.L., Foo, L.C., Wang, G.X., Chakraborty, C., Smith, S.J., Bosco, A., Inman, D.M., Steele, M.R., Wu, G., Soto, I., Marsh-Armstrong, N.,
Barres, B.A., 2012. Astrocyte glypicans 4 and 6 promote formation of excitatory Hubbard, W.C., Calkins, D.J., Horner, P.J., Vetter, M.L., 2008. Reduced retina
synapses via GluA1 AMPA receptors. Nature 486, 410e414. microglial activation and improved optic nerve integrity with minocycline
Amadio, S., De Ninno, A., Montilli, C., Businaro, L., Gerardino, A., Volonte, C., 2013. treatment in the DBA/2J mouse model of glaucoma. Invest. Ophthalmol. Vis. Sci.
Plasticity of primary microglia on micropatterned geometries and spontaneous 49, 1437e1446.
long-distance migration in microfluidic channels. BMC Neurosci. 14, 121e2202, Bosco, A., Linden, R., Carri, N.G., 1994. Expression of alpha-1 integrin subunit in the
2214-2121. mammalian retina. Cell Biol. Int. 18, 211e213.
Amaratunga, A., Abraham, C.R., Edwards, R.B., Sandell, J.H., Schreiber, B.M., Fine, R.E., Bossy, B., Bossy-Wetzel, E., Reichardt, L.F., 1991. Characterization of the integrin
1996. Apolipoprotein E is synthesized in the retina by Muller glial cells, secreted alpha 8 subunit: a new integrin beta 1-associated subunit, which is promi-
into the vitreous, and rapidly transported into the optic nerve by retinal gan- nently expressed on axons and on cells in contact with basal laminae in chick
glion cells. J. Biol. Chem. 271, 5628e5632. embryos. EMBO J. 10, 2375e2385.
Anderson, D.H., Hageman, G.S., Mullins, R.F., Neitz, M., Neitz, J., Ozaki, S., Bradshaw, A.D., McNagny, K.M., Gervin, D.B., Cann, G.M., Graf, T., Clegg, D.O., 1995.
Preissner, K.T., Johnson, L.V., 1999. Vitronectin gene expression in the adult Integrin alpha 2 beta 1 mediates interactions between developing embryonic
human retina. Invest. Ophthalmol. Vis. Sci. 40, 3305e3315. retinal cells and collagen. Development 121, 3593e3602.
Anderson, D.R., Hoyt, W.F., Hogan, M.J., 1967. The fine structure of the astroglia in Bringmann, A., Grosche, A., Pannicke, T., Reichenbach, A., 2013. GABA and glutamate
the human optic nerve and optic nerve head. Trans. Am. Ophthalmol. Soc. 65, uptake and metabolism in retinal glial (Muller) cells. Front. Endocrinol. (Lau-
275e305. sanne) 4, 48.
Argaw, A.T., Gurfein, B.T., Zhang, Y., Zameer, A., John, G.R., 2009. VEGF-mediated Bringmann, A., Iandiev, I., Pannicke, T., Wurm, A., Hollborn, M., Wiedemann, P.,
disruption of endothelial CLN-5 promotes blood-brain barrier breakdown. Proc. Osborne, N.N., Reichenbach, A., 2009a. Cellular signaling and factors involved in
Natl. Acad. Sci. U. S. A. 106, 1977e1982. Müller cell gliosis: neuroprotective and detrimental effects. Prog. Retin. Eye Res.
Arroba, A.I., Alvarez-Lindo, N., van Rooijen, N., de la Rosa, E.J., 2011. Microglia- 28, 423e451.
mediated IGF-I neuroprotection in the rd10 mouse model of retinitis pigmen- Bringmann, A., Pannicke, T., Biedermann, B., Francke, M., Iandiev, I., Grosche, J.,
tosa. Invest. Ophthalmol. Vis. Sci. 52, 9124e9130. Wiedemann, P., Albrecht, J., Reichenbach, A., 2009b. Role of retinal glial cells in
Arroba, A.I., Alvarez-Lindo, N., van Rooijen, N., de la Rosa, E.J., 2014. Microglia- neurotransmitter uptake and metabolism. Neurochem. Int. 54, 143e160.
Muller glia crosstalk in the rd10 mouse model of retinitis pigmentosa. Adv. Exp. Bringmann, A., Pannicke, T., Grosche, J., Francke, M., Wiedemann, P., Skatchkov, S.N.,
Med. Biol. 801, 373e379. Osborne, N.N., Reichenbach, A., 2006. Müller cells in the healthy and diseased
Bachoo, R.M., Kim, R.S., Ligon, K.L., Maher, E.A., Brennan, C., Billings, N., Chan, S., retina. Prog. Retin. Eye Res. 25, 397e424.
Li, C., Rowitch, D.H., Wong, W.H., DePinho, R.A., 2004. Molecular diversity of Bringmann, A., Wiedemann, P., 2012. Muller glial cells in retinal disease. Oph-
astrocytes with implications for neurological disorders. Proc. Natl. Acad. Sci. U. thalmologica 227, 1e19.
S. A. 101, 8384e8389. Bruckner, E., Grosche, A., Pannicke, T., Wiedemann, P., Reichenbach, A.,
Bahr, M., Przyrembel, C., Bastmeyer, M., 1995. Astrocytes from adult rat optic nerves Bringmann, A., 2012. Mechanisms of VEGF- and glutamate-induced inhibition
are nonpermissive for regenerating retinal ganglion cell axons. Exp. Neurol. 131, of osmotic swelling of murine retinal glial (Muller) cells: indications for the
211e220. involvement of vesicular glutamate release and connexin-mediated ATP
Barres, B., Dowling, J., 2010. Roles of astrocytes and other glial cells in glaucoma and release. Neurochem. Res. 37, 268e278.
other retinal diseases. Astrocytes Glaucomatous Neurodegener. Buffo, A., Rite, I., Tripathi, P., Lepier, A., Colak, D., Horn, A.P., Mori, T., Gotz, M., 2008.
Barres, B.A., 1991. New roles for glia. J. Neurosci. 11, 3685e3694. Origin and progeny of reactive gliosis: a source of multipotent cells in the
Barres, B.A., 2008. The mystery and magic of glia: a perspective on their roles in injured brain. Proc. Natl. Acad. Sci. U. S. A. 105, 3581e3586.
health and disease. Neuron 60, 430e440. Bunt-Milam, A.H., Saari, J.C., 1983. Immunocytochemical localization of two
Bartsch, U., Bartsch, S., Dorries, U., Schachner, M., 1992. Immunohistological local- retinoid-binding proteins in vertebrate retina. J. Cell Biol. 97, 703e712.
ization of tenascin in the developing and Lesioned adult mouse optic nerve. Eur. Burda, J.E., Sofroniew, M.V., 2014. Reactive gliosis and the multicellular response to
J. Neurosci. 4, 338e352. CNS damage and disease. Neuron 81, 229e248.
Bartsch, U., Faissner, A., Trotter, J., Dorries, U., Bartsch, S., Mohajeri, H., Bush, T.G., Puvanachandra, N., Horner, C.H., Polito, A., Ostenfeld, T., Svendsen, C.N.,
Schachner, M., 1994. Tenascin demarcates the boundary between the myelin- Mucke, L., Johnson, M.H., Sofroniew, M.V., 1999. Leukocyte infiltration, neuronal
ated and nonmyelinated part of retinal ganglion cell axons in the developing degeneration, and neurite outgrowth after ablation of scar-forming, reactive
and adult mouse. J. Neurosci. 14, 4756e4768. astrocytes in adult transgenic mice. Neuron 23, 297e308.
Bates, C.A., Becker, C.G., Miotke, J.A., Meyer, R.L., 1999. Expression of polysialylated Cahoy, J.D., Emery, B., Kaushal, A., Foo, L.C., Zamanian, J.L., Christopherson, K.S.,
NCAM but not L1 or N-cadherin by regenerating adult mouse optic fibers Xing, Y., Lubischer, J.L., Krieg, P.A., Krupenko, S.A., Thompson, W.J., Barres, B.A.,
in vitro. Exp. Neurol. 155, 128e139. 2008. A transcriptome database for astrocytes, neurons, and oligodendrocytes:
Bazan, N.G., 2007. Omega-3 fatty acids, pro-inflammatory signaling and neuro- a new resource for understanding brain development and function. J. Neurosci.
protection. Curr. Opin. Clin. Nutr. Metab. Care 10, 136e141. 28, 264e278.
Becker, S., Singhal, S., Jones, M.F., Eastlake, K., Cottrill, P.B., Jayaram, H., Limb, A., Cann, G.M., Bradshaw, A.D., Gervin, D.B., Hunter, A.W., Clegg, D.O., 1996. Widespread
2013. Acquisition of RGC phenotype in human Müller glia with stem cell expression of beta1 integrins in the developing chick retina: evidence for a role
characteristics is accompanied by upregulation of functional nicotinic acetyl- in migration of retinal ganglion cells. Dev. Biol. 180, 82e96.
choline receptors. Mol. Vis. 19, 1925e1936. Carson, M.J., Doose, J.M., Melchior, B., Schmid, C.D., Ploix, C.C., 2006. CNS immune
Belecky-Adams, T.L., Chernoff, E.C., Wilson, J.M., Dharmarajan, S., 2013. Reactive privilege: hiding in plain sight. Immunol. Rev. 213, 48e65.
muller glia as potential retinal progenitors. In: Bonfanti, L. (Ed.), Neural Stem Carulli, D., Laabs, T., Geller, H.M., Fawcett, J.W., 2005. Chondroitin sulfate pro-
Cells e New Perspectives. InTech. teoglycans in neural development and regeneration. Curr. Opin. Neurobiol. 15,
Beltran, W.A., Cideciyan, A.V., Lewin, A.S., Hauswirth, W.W., Jacobson, S.G., 116e120.
Aguirre, G.D., 2014. Gene Augmentation for X-linked retinitis pigmentosa Casson, R.J., Chidlow, G., Han, G., Wood, J.P., 2013. An explanation for the Warburg
caused by mutations in RPGR. Cold Spring Harb. Perspect. Med. effect in the adult mammalian retina. Clin. Exp. Ophthalmol. 41, 517.
Benveniste, E.N., 1993. Astrocyte-microgía interactions. In: Murphy, S. (Ed.), Astro- Cepko, C., 1993. Lineage versus environment in the embryonic retina. Trends
cytes. Pharmacology and Function. Academic Press, San Diego, pp. 355e382. Neurosci. 16, 96e97. Author reply 98.
Berger, U.V., Luthi-Carter, R., Passani, L.A., Elkabes, S., Black, I., Konradi, C., Coyle, J.T., Chaboub, L.S., Deneen, B., 2012. Developmental origins of astrocyte heterogeneity:
1999. Glutamate carboxypeptidase II is expressed by astrocytes in the adult rat the final frontier of CNS development. Dev. Neurosci. 34, 379e388.
nervous system. J. Comp. Neurol. 415, 52e64. Chai, L., Morris, J.E., 1999. Heparan sulfate in the inner limiting membrane of em-
Besser, M., Jagatheaswaran, M., Reinhard, J., Schaffelke, P., Faissner, A., 2012. bryonic chicken retina binds basic fibroblast growth factor to promote axonal
Tenascin C regulates proliferation and differentiation processes during embry- outgrowth. Exp. Neurol. 160, 175e185.
onic retinogenesis and modulates the de-differentiation capacity of Muller glia Chang, K.C., Ponder, J., Labarbera, D.V., Petrash, J.M., 2014. Aldose reductase inhi-
by influencing growth factor responsiveness and the extracellular matrix bition prevents endotoxin-induced inflammatory responses in retinal microglia.
compartment. Dev. Biol. 369, 163e176. Invest. Ophthalmol. Vis. Sci. 55, 2853e2861.
Bhatti, F., Ball, G., Hobbs, R., Linens, A., Munzar, S., Akram, R., Barber, A.J., Chen, K., Zhang, Q., Wang, J., Liu, F., Mi, M., Xu, H., Chen, F., Zeng, K., 2009a. Taurine
Anderson, M.P., Elliott, M.H., Edwards, M., 2014. Pulmonary surfactant protein A protects transformed rat retinal ganglion cells from hypoxia-induced apoptosis
is expressed in the mouse retina by muller cells and impacts neovascularization by preventing mitochondrial dysfunction. Brain Res. 1279, 131e138.
in oxygen induced retinopathy. Invest. Ophthalmol. Vis. Sci. Chen, M., Forrester, J.V., Xu, H., 2007. Synthesis of complement factor H by retinal
Bignami, A., Eng, L.F., Dahl, D., Uyeda, C.T., 1972. Localization of the glial fibrillary pigment epithelial cells is down-regulated by oxidized photoreceptor outer
acidic protein in astrocytes by immunofluorescence. Brain Res. 43, 429e435. segments. Exp. Eye Res. 84, 635e645.
Bilimoria, P.M., Stevens, B., 2014. Microglia function during brain development: new Chen, M., Muckersie, E., Luo, C., Forrester, J.V., Xu, H., 2010. Inhibition of the
34 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

alternative pathway of complement activation reduces inflammation in Beckers, J., Hauck, S.M., Ueffing, M., 2011. GDNF-induced osteopontin from
experimental autoimmune uveoretinitis. Eur. J. Immunol. 40, 2870e2881. Muller glial cells promotes photoreceptor survival in the Pde6brd1 mouse
Chen, M., Zhao, J., Luo, C., Pandi, S.P., Penalva, R.G., Fitzgerald, D.C., Xu, H., 2012. model of retinal degeneration. Glia 59, 821e832.
Para-inflammation-mediated retinal recruitment of bone marrow-derived Del Rio-Hortega, P., 1939. The microglia. Lancet 233, 1023e1926.
myeloid cells following whole-body irradiation is CCL2 dependent. Glia 60, den Hollander, A.I., Roepman, R., Koenekoop, R.K., Cremers, F.P., 2008. Leber
833e842. congenital amaurosis: genes, proteins and disease mechanisms. Prog. Retin. Eye
Chen, W., Mo, W., Sun, K., Huang, X., Zhang, Y.L., Song, H.Y., 2009b. Microplasmin Res. 27, 391e419.
degrades fibronectin and laminin at vitreoretinal interface and outer retina Derouiche, A., Rauen, T., 1995. Coincidence of L-glutamate/L-aspartate transporter
during enzymatic vitrectomy. Curr. Eye Res. 34, 1057e1064. (GLAST) and glutamine synthetase (GS) immunoreactions in retinal glia: evi-
Chen, X., Zhou, H., Gong, Y., Wei, S., Zhang, M., 2014. Early spatiotemporal charac- dence for coupling of GLAST and GS in transmitter clearance. J. Neurosci. Res.
terization of microglial activation in the retinas of rats with streptozotocin- 42, 131e143.
induced diabetes. Graefes Arch. Clin. Exp. Ophthalmol. Dharmarajan, S., Gurel, Z., Wang, S., Sorenson, C.M., Sheibani, N., Belecky-
Cherry, J.D., Olschowka, J.A., O'Banion, M.K., 2014. Neuroinflammation and M2 Adams, T.L., 2013. Bone morphogenetic protein 7 regulates reactive gliosis in
microglia: the good, the bad, and the inflamed. J. Neuroinflammation 11, retinal astrocytes and Müller glia. Mol. Vis. 20, 1085e1108.
98e2094, 2011-2098. Ding, D., Xu, L., Xu, H., Li, X., Liang, Q., Zhao, Y., Wang, Y., 2014. Mash1 efficiently
Cho, K.S., Yang, L., Lu, B., Feng Ma, H., Huang, X., Pekny, M., Chen, D.F., 2005. Re- reprograms rat astrocytes into neurons. Neural Regen. Res. 9, 25e32.
establishing the regenerative potential of central nervous system axons in Dreher, Z., Robinson, S.R., Distler, C., 1992. Muller cells in vascular and avascular
postnatal mice. J. Cell Sci. 118, 863e872. retinae: a survey of seven mammals. J. Comp. Neurol. 323, 59e80.
Christopherson, K.S., Ullian, E.M., Stokes, C.C., Mullowney, C.E., Hell, J.W., Agah, A., Duband, J.L., Belkin, A.M., Syfrig, J., Thiery, J.P., Koteliansky, V.E., 1992. Expression of
Lawler, J., Mosher, D.F., Bornstein, P., Barres, B.A., 2005. Thrombospondins are alpha 1 integrin, a laminin-collagen receptor, during myogenesis and neuro-
astrocyte-secreted proteins that promote CNS synaptogenesis. Cell 120, genesis in the avian embryo. Development 116, 585e600.
421e433. Dvoriantchikova, G., Ivanov, D., 2014. Tumor necrosis factor-alpha mediates acti-
Chung, W.S., Clarke, L.E., Wang, G.X., Stafford, B.K., Sher, A., Chakraborty, C., Joung, J., vation of NF-kappaB and JNK signaling cascades in retinal ganglion cells and
Foo, L.C., Thompson, A., Chen, C., Smith, S.J., Barres, B.A., 2013. Astrocytes astrocytes in opposite ways. Eur. J. Neurosci. 40, 3171e3178.
mediate synapse elimination through MEGF10 and MERTK pathways. Nature Edelstein, L., Smythies, J., 2014. The role of epigenetic-related codes in neuro-
504, 394e400. computation: dynamic hardware in the brain. Philos. Trans. R. Soc. Lond. B Biol.
Clark, S.J., Keenan, T.D., Fielder, H.L., Collinson, L.J., Holley, R.J., Merry, C.L., van Sci. 369.
Kuppevelt, T.H., Day, A.J., Bishop, P.N., 2011. Mapping the differential distribu- Eichler, W., Yafai, Y., Wiedemann, P., Reichenbach, A., 2004. Angiogenesis related
tion of glycosaminoglycans in the adult human retina, choroid, and sclera. factors derived from retinal glial (Müller) cells in hypoxia. Neuroreport 15,
Invest. Ophthalmol. Vis. Sci. 52, 6511e6521. 1633e1637.
Cohen, J., Burne, J.F., McKinlay, C., Winter, J., 1987. The role of laminin and the Ellis-Behnke, R.G., Jonas, R.A., Jonas, J.B., 2013. The microglial system in the eye and
laminin/fibronectin receptor complex in the outgrowth of retinal ganglion cell brain in response to stimuli in vivo. J. Glaucoma 22 (Suppl. 5), S32eS35.
axons. Dev. Biol. 122, 407e418. Enger, R., Gundersen, G.A., Haj-Yasein, N.N., Eilert-Olsen, M., Thoren, A.E.,
Colton, C.A., 2009. Heterogeneity of microglial activation in the innate immune Vindedal, G.F., Petersen, P.H., Skare, O., Nedergaard, M., Ottersen, O.P.,
response in the brain. J. Neuroimmune Pharmacol. 4, 399e418. Nagelhus, E.A., 2012. Molecular scaffolds underpinning macroglial polarization:
Coorey, N.J., Shen, W., Chung, S.H., Zhu, L., Gillies, M.C., 2012. The role of glia in an analysis of retinal Muller cells and brain astrocytes in mouse. Glia 60,
retinal vascular disease. Clin. Exp. Optom. 95, 266e281. 2018e2026.
Copland, D.A., Hussain, K., Baalasubramanian, S., Hughes, T.R., Morgan, B.P., Xu, H., Ennis, S., Kunz, Y.W., 1986. Differentiated retinal Muller glia are cil-
Dick, A.D., Nicholson, L.B., 2010. Systemic and local anti-C5 therapy reduces the iatedeultrastructural evidence in the teleost Poecilia reticulata P. Cell Biol. Int.
disease severity in experimental autoimmune uveoretinitis. Clin. Exp. Immunol. Rep. 10, 611e622.
159, 303e314. Erickson, P.A., Fisher, S.K., Guerin, C.J., Anderson, D.H., Kaska, D.D., 1987. Glial
Corraliza, I., 2014. Recruiting specialized macrophages across the borders to restore fibrillary acidic protein increases in Muller cells after retinal detachment. Exp.
brain functions. Front. Cell. Neurosci. 8, 262. Eye Res. 44, 37e48.
Crish, S.D., Dapper, J.D., MacNamee, S.E., Balaram, P., Sidorova, T.N., Lambert, W.S., Escartin, C., Brouillet, E., Gubbellini, P., Trioulier, Y., Jacquard, C., 2006. Ciliary
Calkins, D.J., 2013. Failure of axonal transport induces a spatially coincident neurotrophic factor activates astrocytes, redistributes their glutamate trans-
increase in astrocyte BDNF prior to synapse loss in a central target. Neurosci- porters GLAST and GLT-1 to raft microdomains, and improves glutamate
ence 229, 55e70. handling in vivo. J. Neurosci. 26, 5978e5989.
Cuenca, N., Fernandez-Sanchez, L., Campello, L., Maneu, V., De la Villa, P., Lax, P., Escartin, C., Pierre, K., Colin, A., Brouillet, E., Delzescaux, T., 2007. Activation of as-
Pinilla, I., 2014. Cellular responses following retinal injuries and therapeutic trocytes by CNTF induces metabolic plasticity and increases resistance to
approaches for neurodegenerative diseases. Prog. Retin. Eye Res. 43C, 17e75. metabolic insults. J. Neurosci. 27, 7094e7104.
Dai, C., Khaw, P.T., Yin, Z.Q., Li, D., Raisman, G., Li, Y., 2012a. Structural basis of Fawcett, J.W., Housden, E., Smith-Thomas, L., Meyer, R.L., 1989. The growth of axons
glaucoma: the fortified astrocytes of the optic nerve head are the target of in three-dimensional astrocyte cultures. Dev. Biol. 135, 449e458.
raised intraocular pressure. Glia 60, 13e28. Fernandes, A., Miller-Fleming, L., Pais, T.F., 2014. Microglia and inflammation:
Dai, M., Xia, X.B., Xiong, S.Q., 2012b. BDNF regulates GLAST and glutamine syn- conspiracy, controversy or control? Cell. Mol. Life Sci. 71, 3969e3985.
thetase in mouse retinal Müller cells. J. Cell. Physiol. 227, 596e603. Fern andez, J.M., Di Giusto, G., Kalstein, M., Melamud, L., Rivarola, V., Ford, P.,
Damani, M.R., Zhao, L., Fontainhas, A.M., Amaral, J., Fariss, R.N., Wong, W.T., 2011. Capurro, C., 2013. Cell volume regulation in cultured human retinal Müller cells
Age-related alterations in the dynamic behavior of microglia. Aging Cell. 10, is associated with changes in transmembrane potential. PLoS One 8, e57268.
263e276. Fischer, A.J., Bongini, R., 2010. Turning Muller glia into neural progenitors in the
Danbolt, N.C., 2001. Glutamate uptake. Prog. Neurobiol. 65, 1e105. retina. Mol. Neurobiol. 42, 199e209.
Das, A.V., Mallya, K.B., Zhao, X., Ahmad, F., Bhattacharya, S., Thoreson, W.B., Fischer, A.J., Reh, T.A., 2001. Muller glia are a potential source of neural regeneration
Hegde, G.V., Ahmad, I., 2006. Neural stem cell properties of Muller glia in the in the postnatal chicken retina. Nat. Neurosci. 4, 247e252.
mammalian retina: regulation by Notch and Wnt signaling. Dev. Biol. 299, Fischer, A.J., Schmidt, M., Omar, G., Reh, T.A., 2004. BMP4 and CNTF are neuro-
283e302. protective and suppress damage-induced proliferation of Muller glia in the
Davalos, D., Grutzendler, J., Yang, G., Kim, J.V., Zuo, Y., Jung, S., Littman, D.R., retina. Mol. Cell. Neurosci. 27, 531e542.
Dustin, M.L., Gan, W.B., 2005. ATP mediates rapid microglial response to local Flanary, B.E., Sammons, N.W., Nguyen, C., Walker, D., Streit, W.J., 2007. Evidence that
brain injury in vivo. Nat. Neurosci. 8, 752e758. aging and amyloid promote microglial cell senescence. Rejuvenation Res. 10,
Davis, C.H., Kim, K.Y., Bushong, E.A., Mills, E.A., Boassa, D., Shih, T., Kinebuchi, M., 61e74.
Phan, S., Zhou, Y., Bihlmeyer, N.A., Nguyen, J.V., Jin, Y., Ellisman, M.H., Marsh- Fletcher, E.L., Phipps, J.A., Wilkinson-Berka, J.L., 2005. Dysfunction of retinal neu-
Armstrong, N., 2014. Transcellular degradation of axonal mitochondria. Proc. rons and glia during diabetes. Clin. Exp. Optom. 88, 132e145.
Natl. Acad. Sci. U. S. A. 111, 9633e9638. Fliesler, S.J., Keller, R.K., 1997. Isoprenoid metabolism in the vertebrate retina. Int. J.
Davis, E.J., Foster, T.D., Thomas, W.E., 1994. Cellular forms and functions of brain Biochem. Cell. Biol. 29, 877e894.
microglia. Brain Res. Bull. 34, 73e78. Foo, L.C., Allen, N.J., Bushong, E.A., Ventura, P.B., Chung, W.S., Zhou, L., Cahoy, J.D.,
Davis, J.T., Wen, Q., Janmey, P.A., Otteson, D.C., Foster, W.J., 2012. Muller cell Daneman, R., Zong, H., Ellisman, M.H., Barres, B.A., 2011. Development of a
expression of genes implicated in proliferative vitreoretinopathy is influenced method for the purification and culture of rodent astrocytes. Neuron 71,
by substrate elastic modulus. Invest. Ophthalmol. Vis. Sci. 53, 3014e3019. 799e811.
de Curtis, I., 1993. The alpha 6 beta 1 integrin is a laminin receptor for developing Formichella, C.R., Abella, S.K., Sims, S.M., Cathcart, H.M., Sappington, R.M., 2014.
retinal neurons. Cytotechnology 11 (Suppl. 1), S41eS43. Astrocyte reactivity: a biomarker for retinal ganglion cell health in retinal
Deeg, C.A., Eberhardt, C., Hofmaier, F., Amann, B., Hauck, S.M., 2011. Osteopontin and neurodegeneration. J. Clin. Cell. Immunol. 5.
fibronectin levels are decreased in vitreous of autoimmune uveitis and retinal Franze, K., Francke, M., Guenter, K., Christ, A.F., Koerber, N., Reichenbach, A., Guck, J.,
expression of both proteins indicates ECM re-modeling. PLoS One 6, e27674. 2011. Spatial mapping of the mechanical properties of the living retina using
Deguchi, J., Yamamoto, A., Fujiki, Y., Uyama, M., Tsukahara, I., Tashiro, Y., 1992. scanning force microscopy. Soft Matter 7, 3147e3154.
Localization of nonspecific lipid transfer protein (nsLTP ¼ sterol carrier protein Franze, K., Grosche, J., Skatchkov, S.N., Schinkinger, S., Foja, C., Schild, D.,
2) and acyl-CoA oxidase in peroxisomes of pigment epithelial cells of rat retina. Uckermann, O., Travis, K., Reichenbach, A., Guck, J., 2007. Müller cells are living
J. Histochem. Cytochem. 40, 403e410. optical fibers in the vertebrate retina. Proc. Natl. Acad. Sci. U. S. A. 104,
Del Rio, P., Irmler, M., Arango-Gonzalez, B., Favor, J., Bobe, C., Bartsch, U., Vecino, E., 8287e8292.
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 35

Frasson, M., Sahel, J.A., Fabre, M., Simonutti, M., Dreyfus, H., Picaud, S., 1999. Reti- response to extracellular matrix proteins: developmental changes and effects of
nitis pigmentosa: rod photoreceptor rescue by a calcium-channel blocker in the the cell substratum attachment antibody (CSAT). J. Cell. Biol. 104, 623e634.
rd mouse. Nat. Med. 5, 1183e1187. Hama, K., Mizukawa, A., Kosaka, T., 1978. Fine structure of the Muller cell revealed
Fuchshofer, R., Welge-Lussen, U., Lutjen-Drecoll, E., Birke, M., 2006. Biochemical and by high-voltage electron microscopy. Sens. Process. 2, 296e299.
morphological analysis of basement membrane component expression in cor- Harada, C., Harada, T., Quah, H.M., Maekawa, F., Yoshida, K., Ohno, S., Wada, K.,
neoscleral and cribriform human trabecular meshwork cells. Invest. Oph- Parada, L.F., Tanaka, K., 2003. Potential role of glial cell line-derived neuro-
thalmol. Vis. Sci. 47, 794e801. trophic factor receptors in Muller glial cells during light-induced retinal
Galdos, M., Vecino, E., 2011. Ultrastructural changes in the trabecular meshwork and degeneration. Neuroscience 122, 229e235.
increased IOP. Which came first, the chicken or the egg? Arch. Soc. Esp. Oftal- Harada, C., Harada, T., Slusher, B.S., Yoshida, K., Matsuda, H., Wada, K., 2000. N-
mol. 86, 241e242. acetylated-alpha-linked-acidic dipeptidase inhibitor has a neuroprotective ef-
Gallego, B.I., Salazar, J.J., de Hoz, R., Rojas, B., Ramirez, A.I., Salinas-Navarro, M., fect on mouse retinal ganglion cells after pressure-induced ischemia. Neurosci.
Ortin-Martinez, A., Valiente-Soriano, F.J., Aviles-Trigueros, M., Villegas- Lett. 292, 134e136.
Perez, M.P., Vidal-Sanz, M., Trivino, A., Ramirez, J.M., 2012. IOP induces upre- Harada, T., Harada, C., Kohsaka, S., Wada, E., Yoshida, K., Ohno, S., Mamada, H.,
gulation of GFAP and MHC-II and microglia reactivity in mice retina contra- Tanaka, K., Parada, L.F., Wada, K., 2002. Microglia-Müller glia cell interactions
lateral to experimental glaucoma. J. Neuroinflammation 9, 92. control neurotrophic factor production during light-induced retinal degenera-
Garcia, M., Forster, V., Hicks, D., Vecino, E., 2002. Effects of muller glia on cell sur- tion. J. Neurosci. 22, 9228e9236.
vival and neuritogenesis in adult porcine retina in vitro. Invest. Ophthalmol. Vis. Harry, G.J., 2013. Microglia during development and aging. Pharmacol. Ther. 139,
Sci. 43, 3735e3743. 313e326.
Garcia, M., Forster, V., Hicks, D., Vecino, E., 2003. In vivo expression of neuro- Hauck, S.M., Kinkl, N., Deeg, C.A., Swiatek-de Lange, M., Schoffmann, S., Ueffing, M.,
trophins and neurotrophin receptors is conserved in adult porcine retina 2006. GDNF family ligands trigger indirect neuroprotective signaling in retinal
in vitro. Invest. Ophthalmol. Vis. Sci. 44, 4532e4541. glial cells. Mol. Cell. Biol. 26, 2746e2757.
Garcia, M., Vecino, E., 2003. Role of Muller glia in neuroprotection and regeneration Hauck, S.M., von Toerne, C., Ueffing, M., 2014. The neuroprotective potential of
in the retina. Histol. Histopathol. 18, 1205e1218. retinal Müller glial cells. Adv. Exp. Med. Biol. 801, 381e387.
Garcia, T.B., Pannicke, T., Vogler, S., Berk, B.A., Grosche, A., Wiedemann, P., Seeger, J., Hauswirth, W.W., Aleman, T.S., Kaushal, S., Cideciyan, A.V., Schwartz, S.B., Wang, L.,
Reichenbach, A., Herculano, A.M., Bringmann, A., 2014. Nerve growth factor Conlon, T.J., Boye, S.L., Flotte, T.R., Byrne, B.J., Jacobson, S.G., 2008. Treatment of
inhibits osmotic swelling of rat retinal glial (Muller) and bipolar cells by leber congenital amaurosis due to RPE65 mutations by ocular subretinal in-
inducing glial cytokine release. J. Neurochem. 131, 303e313. jection of adeno-associated virus gene vector: short-term results of a phase I
Garcia-Valenzuela, E., Shareef, S., Walsh, J., Sharma, S.C., 1995. Programmed cell trial. Hum. Gene Ther. 19, 979e990.
death of retinal ganglion cells during experimental glaucoma. Exp. Eye Res. 61, He, Y., Zhang, Y., Liu, X., Ghazaryan, E., Li, Y., Xie, J., Su, G., 2014. Recent advances of
33e44. stem cell therapy for retinitis pigmentosa. Int. J. Mol. Sci. 15, 14456e14474.
Garcia-Valenzuela, E., Sharma, S.C., Pina, A.L., 2005. Multilayered retinal microglial Heiduschka, P., Renninger, D., Fischer, D., Muller, A., Hofmeister, S., Schraermeyer, U.,
response to optic nerve transection in rats. Mol. Vis. 11, 225e231. 2013. Lens injury has a protective effect on photoreceptors in the RCS rat. ISRN
Gehrig, A., Langmann, T., Horling, F., Janssen, A., Bonin, M., Walter, M., Poths, S., Ophthalmol. 2013, 814814.
Weber, B.H., 2007. Genome-wide expression profiling of the retinoschisin- Heins, N., Malatesta, P., Cecconi, F., Nakafuku, M., Tucker, K.L., Hack, M.A.,
deficient retina in early postnatal mouse development. Invest. Ophthalmol. Chapouton, P., Barde, Y.A., Gotz, M., 2002. Glial cells generate neurons: the role
Vis. Sci. 48, 891e900. of the transcription factor Pax6. Nat. Neurosci. 5, 308e315.
Gertig, U., Hanisch, U.K., 2014. Microglial diversity by responses and responders. Hernandez, M., Pearce-Kelling, S.E., Rodriguez, F.D., Aguirre, G.D., Vecino, E., 2010.
Front. Cell. Neurosci. 8, 101. Altered expression of retinal molecular markers in the canine RPE65 model of
Ginhoux, F., Greter, M., Leboeuf, M., Nandi, S., See, P., Gokhan, S., Mehler, M.F., Leber congenital amaurosis. Invest. Ophthalmol. Vis. Sci. 51, 6793e6802.
Conway, S.J., Ng, L.G., Stanley, E.R., Samokhvalov, I.M., Merad, M., 2010. Fate Hernandez, M., Rodriguez, F.D., Sharma, S.C., Vecino, E., 2009. Immunohistochem-
mapping analysis reveals that adult microglia derive from primitive macro- ical changes in rat retinas at various time periods of elevated intraocular
phages. Science 330, 841e845. pressure. Mol. Vis. 15, 2696e2709.
Ginhoux, F., Lim, S., Hoeffel, G., Low, D., Huber, T., 2013. Origin and differentiation of Hernandez, M., Urcola, J.H., Vecino, E., 2008. Retinal ganglion cell neuroprotection
microglia. Front. Cell. Neurosci. 7, 45. in a rat model of glaucoma following brimonidine, latanoprost or combined
Glanzer, J.G., Enose, Y., Wang, T., Kadiu, I., Gong, N., Rozek, W., Liu, J., treatments. Exp. Eye Res. 86, 798e806.
Schlautman, J.D., Ciborowski, P.S., Thomas, M.P., Gendelman, H.E., 2007. Hernandez, M.R., Pena, J.D., Selvidge, J.A., Salvador-Silva, M., Yang, P., 2000. Hy-
Genomic and proteomic microglial profiling: pathways for neuroprotective drostatic pressure stimulates synthesis of elastin in cultured optic nerve head
inflammatory responses following nerve fragment clearance and activation. astrocytes. Glia 32, 122e136.
J. Neurochem. 102, 627e645. Higginson, J.R., Ruzafa, N., Rodríguez-Ferna ndez, D., Vecino, E., 2013. Neuro-
Goldmann, T., Wieghofer, P., Muller, P.F., Wolf, Y., Varol, D., Yona, S., Brendecke, S.M., protection of retinal ganglion cells by Müller glia and astrocytes. Poster. In: XI
Kierdorf, K., Staszewski, O., Datta, M., Luedde, T., Heikenwalder, M., Jung, S., European Meeting on Glial Cell Function in Health and Disease.
Prinz, M., 2013. A new type of microglia gene targeting shows TAK1 to be Hirrlinger, P.G., Wurm, A., Hirrlinger, J., Bringmann, A., Reichenbach, A., 2008. Os-
pivotal in CNS autoimmune inflammation. Nat. Neurosci. 16, 1618e1626. motic swelling characteristics of glial cells in the murine hippocampus, cere-
Gonzalez-Scarano, F., Baltuch, G., 1999. Microglia as mediators of inflammatory and bellum, and retina in situ. J. Neurochem. 105, 1405e1417.
degenerative diseases. Annu. Rev. Neurosci. 22, 219e240. Hiscott, P., Waller, H.A., Grierson, I., Butler, M.G., Scott, D., 1992. Local production of
Gordon, S., 2003. Alternative activation of macrophages. Nat. Rev. Immunol. 3, fibronectin by ectopic human retinal cells. Cell Tissue Res. 267, 185e192.
23e35. Hogan, M.J., Feeney, L., 1963. The ultrastructure of the retinal vessels. Iii. Vascular-
Gordon, S., Martinez, F.O., 2010. Alternative activation of macrophages: mechanism Glial relationships. J. Ultrastruct. Res. 49, 47e64.
and functions. Immunity 32, 593e604. Hogan, M.J., Moschini, G.B., Zardi, O., 1971. Effects of Toxoplasma gondii toxin on the
Gordon, W.C., Bazan, N.G., 1990. Docosahexaenoic acid utilization during rod rabbit eye. Am. J. Ophthalmol. 72, 733e742.
photoreceptor cell renewal. J. Neurosci. 10, 2190e2202. Hollander, H., Makarov, F., Dreher, Z., van Driel, D., Chan-Ling, T.L., Stone, J., 1991.
Gorovits, R., Avidan, N., Avisar, N., Shaked, I., Vardimon, L., 1997. Glutamine syn- Structure of the macroglia of the retina: sharing and division of labour between
thetase protects against neuronal degeneration in injured retinal tissue. Proc. astrocytes and Muller cells. J. Comp. Neurol. 313, 587e603.
Natl. Acad. Sci. U. S. A. 94, 7024e7029. Hollborn, M., Jahn, K., Limb, G.A., Kohen, L., Wiedemann, P., Bringmann, A., 2004.
Grigsby, J.G., Cardona, S.M., Pouw, C.E., Muniz, A., Mendiola, A.S., Tsin, A.T., Characterization of the basic fibroblast growth factor-evoked proliferation of
Allen, D.M., Cardona, A.E., 2014. The role of microglia in diabetic retinopathy. the human Muller cell line, MIO-M1. Graefes Arch. Clin. Exp. Ophthalmol. 242,
J. Ophthalmol. 2014, 705783. 414e422.
Grosche, A., Pannicke, T., Chen, J., Wiedemann, P., Reichenbach, A., Bringmann, A., Holley, J.E., Gveric, D., Whatmore, J.L., Gutowski, N.J., 2005. Tenascin C induces a
2013. Disruption of endogenous purinergic signaling inhibits vascular endo- quiescent phenotype in cultured adult human astrocytes. Glia 52, 53e58.
thelial growth factor- and glutamate-induced osmotic volume regulation of Hosoya, K., Tachikawa, M., 2012. The inner blood-retinal barrier: molecular struc-
Müller glial cells in knockout mice. Ophthalmic. Res. 50, 209e214. ture and transport biology. Adv. Exp. Med. Biol. 763, 85e104.
Gu, S.M., Thompson, D.A., Srikumari, C.R., Lorenz, B., Finckh, U., Nicoletti, A., Hossmann, K.A., 1993. Ischemia-mediated neuronal injury. Resuscitation 26,
Murthy, K.R., Rathmann, M., Kumaramanickavel, G., Denton, M.J., Gal, A., 1997. 225e235.
Mutations in RPE65 cause autosomal recessive childhood-onset severe retinal Housley, G.D., Bringmann, A., Reichenbach, A., 2009. Purinergic signaling in special
dystrophy. Nat. Genet. 17, 194e197. senses. Trends Neurosci. 32, 128e141.
Guidry, C., Bradley, K.M., King, J.L., 2003. Tractional force generation by human Howell, G.R., Libby, R.T., Marchant, J.K., Wilson, L.A., Cosma, I.M., Smith, R.S.,
muller cells: growth factor responsiveness and integrin receptor involvement. Anderson, M.G., John, S.W., 2007. Absence of glaucoma in DBA/2J mice homo-
Invest. Ophthalmol. Vis. Sci. 44, 1355e1363. zygous for wild-type versions of Gpnmb and Tyrp1. BMC Genet. 8, 45.
Guvakova, M.A., 2007. Insulin-like growth factors control cell migration in health Hu, P., Herrmann, R., Bednar, A., Saloupis, P., Dwyer, M.A., Yang, P., Qi, X.,
and disease. Int. J. Biochem. Cell. Biol. 39, 890e909. Thomas, R.S., Jaffe, G.J., Boulton, M.E., McDonnell, D.P., Malek, G., 2013. Aryl
Hajrasouliha, A.R., Jiang, G., Lu, Q., Lu, H., Kaplan, H.J., Zhang, H.G., Shao, H., 2013. hydrocarbon receptor deficiency causes dysregulated cellular matrix meta-
Exosomes from retinal astrocytes contain antiangiogenic components that bolism and age-related macular degeneration-like pathology. Proc. Natl. Acad.
inhibit laser-induced choroidal neovascularization. J. Biol. Chem. 288, Sci. U. S. A. 110, E4069eE4078.
28058e28067. Hurley, J.B., Chertov, A.O., Lindsay, K., Giamarco, M., Cleghorn, W., Du, J.,
Hall, D.E., Neugebauer, K.M., Reichardt, L.F., 1987. Embryonic neural retinal cell Brockerhoff, S., 2014. Vertebrate photoreceptors: functional molecular bases.
36 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

Energy Metabol. Vertebr. Retina. rat retinal ganglion cells. Exp. Eye Res. 63, 105e112.
Ibrahim, A.S., El-Remessy, A.B., Matragoon, S., Zhang, W., Patel, Y., Khan, S., Al- Kofuji, P., Ceelen, P., Zahs, K.R., Surbeck, L.W., Lester, H.A., Newman, E.A., 2000.
Gayyar, M.M., El-Shishtawy, M.M., Liou, G.I., 2011. Retinal microglial activation Genetic inactivation of an inwardly rectifying potassium channel (Kir4.1 sub-
and inflammation induced by amadori-glycated albumin in a rat model of unit) in mice: phenotypic impact in retina. J. Neurosci. 20, 5733e5740.
diabetes. Diabetes 60, 1122e1133. Kohno, H., Sakai, T., Kitahara, K., 2006. Induction of nestin, Ki-67, and cyclin D1
Inatani, M., Tanihara, H., Oohira, A., Honjo, M., Kido, N., Honda, Y., 2000. Upregu- expression in Muller cells after laser injury in adult rat retina. Graefes Arch.
lated expression of neurocan, a nervous tissue specific proteoglycan, in tran- Clin. Exp. Ophthalmol. 244, 90e95.
sient retinal ischemia. Invest. Ophthalmol. Vis. Sci. 41, 2748e2754. Kohno, T., Sorgente, N., Ishibashi, T., Goodnight, R., Ryan, S.J., 1987. Immunofluo-
Jayaram, H., Jones, M.F., Eastlake, K., Cottrill, P.B., Becker, S., Wiseman, J., Khaw, P.T., rescent studies of fibronectin and laminin in the human eye. Invest. Oph-
Limb, G.A., 2014. Transplantation of photoreceptors derived from human Muller thalmol. Vis. Sci. 28, 506e514.
glia restore rod function in the P23H rat. Stem Cells Transl. Med. 3, 323e333. Kokaia, Z., Airaksinen, M.S., Nanobashvili, A., Larsson, E., Kujamaki, E., Lindvall, O.,
Jeon, C.J., Masland, R.H., 1993. Selective accumulation of diamidino yellow and Saarma, M., 1999. GDNF family ligands and receptors are differentially regulated
chromomycin A3 by retinal glial cells. J. Histochem. Cytochem. 41, 1651e1658. after brain insults in the rat. Eur. J. Neurosci. 11, 1202e1216.
Jerdan, J.A., Glaser, B.M., 1986. Retinal microvessel extracellular matrix: an immu- Koke, J.R., Mosier, A.L., Garcia, D.M., 2010. Intermediate filaments of zebrafish retinal
nofluorescent study. Invest. Ophthalmol. Vis. Sci. 27, 194e203. and optic nerve astrocytes and Muller glia: differential distribution of cyto-
Jha, M.K., Seo, M., Kim, J.H., Kim, B.G., Cho, J.Y., Suk, K., 2013. The secretome keratin and GFAP. BMC Res. Notes 3, 50.
signature of reactive glial cells and its pathological implications. Biochim. Bio- Krady, J.K., Basu, A., Allen, C.M., Xu, Y., LaNoue, K.F., Gardner, T.W., Levison, S.W.,
phys. Acta 1834, 2418e2428. 2005. Minocycline reduces proinflammatory cytokine expression, microglial
Jiang, B., Liou, G.I., Behzadian, M.A., Caldwell, R.B., 1994. Astrocytes modulate retinal activation, and caspase-3 activation in a rodent model of diabetic retinopathy.
vasculogenesis: effects on fibronectin expression. J. Cell Sci. 107 (Pt 9), Diabetes 54, 1559e1565.
2499e2508. Kuehn, M.H., Kim, C.Y., Ostojic, J., Bellin, M., Alward, W.L., Stone, E.M.,
Johnson, E.C., Jia, L., Cepurna, W.O., Doser, T.A., Morrison, J.C., 2007. Global changes Sakaguchi, D.S., Grozdanic, S.D., Kwon, Y.H., 2006. Retinal synthesis and depo-
in optic nerve head gene expression after exposure to elevated intraocular sition of complement components induced by ocular hypertension. Exp. Eye.
pressure in a rat glaucoma model. Invest. Ophthalmol. Vis. Sci. 48, 3161e3177. Res. 83, 620e628.
Jonas, R.A., Yuan, T.F., Liang, Y.X., Jonas, J.B., Tay, D.K., Ellis-Behnke, R.G., 2012. The Kumar, A., Pandey, R.K., Miller, L.J., Singh, P.K., Kanwar, M., 2013. Muller glia in
spider effect: morphological and orienting classification of microglia in retinal innate immunity: a perspective on their roles in endophthalmitis. Crit.
response to stimuli in vivo. PLoS One 7, e30763. Rev. Immunol. 33, 119e135.
Kang, J., Jiang, L., Goldman, S.A., Nedergaard, M., 1998. Astrocyte-mediated poten- Labin, A.M., Ribak, E.N., 2010. Retinal glial cells enhance human vision acuity. Phys.
tiation of inhibitory synaptic transmission. Nat. Neurosci. 1, 683e692. Rev. Lett. 104, 158102.
Kang, X., Yu, J., Wei, Y., Zhao, P., 2013. The effect of A2A receptor sntagonist on the Lange, J., Yafai, Y., Noack, A., Yang, X.M., Munk, A.B., Krohn, S., Iandiev, I.,
mRNA expression of Kir 2.1 and Kir 4.1 channels in rat retinal Müller cells under Wiedemann, P., Reichenbach, A., Eichler, W., 2012. The axon guidance molecule
hypoxic conditions in vitro. Adv. Clin. Exp. Med. 22, 825e829. Netrin-4 is expressed by Muller cells and contributes to angiogenesis in the
Karlstetter, M., Ebert, S., Langmann, T., 2010. Microglia in the healthy and degen- retina. Glia 60, 1567e1578.
erating retina: insights from novel mouse models. Immunobiology 215, Langmann, T., 2007. Microglia activation in retinal degeneration. J. Leukoc. Biol. 81,
685e691. 1345e1351.
Karlstetter, M., Langmann, T., 2014. Microglia in the aging retina. Adv. Exp. Med. Larger, E., Becourt, C., Bach, J.F., Boitard, C., 1995. Pancreatic islet beta cells drive T
Biol. 801, 207e212. cell-immune responses in the nonobese diabetic mouse model. J. Exp. Med. 181,
Karschin, A., Wassle, H., Schnitzer, J., 1986. Shape and distribution of astrocytes in 1635e1642.
the cat retina. Invest. Ophthalmol. Vis. Sci. 27, 828e831. Laties, A.M., 1983. New light on the Muller cell. Trans. Ophthalmol. Soc. U. K. 103 (Pt
Katsman, D., Stackpole, E.J., Domin, D.R., Farber, D.B., 2012. Embryonic stem cell- 4), 380e384.
derived microvesicles induce gene expression changes in Muller cells of the Lawrence, J.M., Singhal, S., Bhatia, B., Keegan, D.J., Reh, T.A., Luthert, P.J., Khaw, P.T.,
retina. PLoS One 7, e50417. Limb, G.A., 2007. MIO-M1 cells and similar muller glial cell lines derived from
Kawikova, I., Askenase, P.W., 2014. Diagnostic and therapeutic potentials of exo- adult human retina exhibit neural stem cell characteristics. Stem Cells 25,
somes in CNS diseases. Brain Res. 2033e2043.
Keenan, T.D., Clark, S.J., Unwin, R.D., Ridge, L.A., Day, A.J., Bishop, P.N., 2012. Mapping Leveillard, T., Fridlich, R., Clerin, E., Ait-Ali, N., Millet-Puel, G., Jaillard, C., Yang, Y.,
the differential distribution of proteoglycan core proteins in the adult human Zack, D., van-Dorsselaer, A., Sahel, J.A., 2014. Therapeutic strategy for handling
retina, choroid, and sclera. Invest. Ophthalmol. Vis. Sci. 53, 7528e7538. inherited retinal degenerations in a gene-independent manner using rod-
Kettenmann, H., Hanisch, U.K., Noda, M., Verkhratsky, A., 2011. Physiology of derived cone viability factors. C. R. Biol. 337, 207e213.
microglia. Physiol. Rev. 91, 461e553. Lewis, G.P., Fisher, S.K., 2003. Up-regulation of glial fibrillary acidic protein in
Kettenmann, H., Kirchhoff, F., Verkhratsky, A., 2013. Microglia: new roles for the response to retinal injury: its potential role in glial remodeling and a com-
synaptic stripper. Neuron 77, 10e18. parison to vimentin expression. Int. Rev. Cytol. 230, 263e290.
Kettenmann, H., Verkhratsky, A., 2008. Neuroglia: the 150 years after. Trends Lewis, G.P., Guerin, C.J., Anderson, D.H., Matsumoto, B., Fisher, S.K., 1994. Rapid
Neurosci. 31, 653e659. changes in the expression of glial cell proteins caused by experimental retinal
Kezic, J.M., Chen, X., Rakoczy, E.P., McMenamin, P.G., 2013. The effects of age and detachment. Am. J. Ophthalmol. 118, 368e376.
Cx3cr1 deficiency on retinal microglia in the Ins2(Akita) diabetic mouse. Invest. Li, J., Zhao, S.Z., Wang, P.P., Yu, S.P., Zheng, Z., Xu, X., 2012. Calcium mediates high
Ophthalmol. Vis. Sci. 54, 854e863. glucose-induced HIF-1a and VEGF expression in cultured rat retinal Müller cells
Kierdorf, K., Erny, D., Goldmann, T., Sander, V., Schulz, C., Perdiguero, E.G., through CaMKII-CREB pathway. Acta Pharmacol. Sin. 33, 1030e1036.
Wieghofer, P., Heinrich, A., Riemke, P., Holscher, C., Muller, D.N., Luckow, B., Libby, R.T., Champliaud, M.F., Claudepierre, T., Xu, Y., Gibbons, E.P., Koch, M.,
Brocker, T., Debowski, K., Fritz, G., Opdenakker, G., Diefenbach, A., Biber, K., Burgeson, R.E., Hunter, D.D., Brunken, W.J., 2000. Laminin expression in adult
Heikenwalder, M., Geissmann, F., Rosenbauer, F., Prinz, M., 2013. Microglia and developing retinae: evidence of two novel CNS laminins. J. Neurosci. 20,
emerge from erythromyeloid precursors via Pu.1- and Irf8-dependent path- 6517e6528.
ways. Nat. Neurosci. 16, 273e280. Light, J.G., Fransen, J.W., Adekunle, A.N., Adkins, A., Pangeni, G., Loudin, J.,
Kim, I.B., Kim, K.Y., Joo, C.K., Lee, M.Y., Oh, S.J., Chung, J.W., Chun, M.H., 1998. Re- Mathieson, K., Palanker, D.V., McCall, M.A., Pardue, M.T., 2014. Inner retinal
action of Muller cells after increased intraocular pressure in the rat retina. Exp. preservation in rat models of retinal degeneration implanted with subretinal
Brain Res. 121, 419e424. photovoltaic arrays. Exp. Eye Res. 128C, 34e42.
Kim, J.H., Kim, J.H., Park, J.A., Lee, S.W., Kim, W.J., Yu, Y.S., Kim, K.W., 2006. Blood- Limb, G.A., Salt, T.E., Munro, P.M., Moss, S.E., Khaw, P.T., 2002. In vitro character-
neural barrier: intercellular communication at glio-vascular interface. ization of a spontaneously immortalized human Muller cell line (MIO-M1).
J. Biochem. Mol. Biol. 39, 339e345. Invest. Ophthalmol. Vis. Sci. 43, 864e869.
Kim, S.Y., Yang, H.J., Chang, Y.S., Kim, J.W., Brooks, M., Chew, E.Y., Wong, W.T., Lindqvist, N., Liu, Q., Zajadacz, J., Franze, K., Reichenbach, A., 2010. Retinal glial
Fariss, R.N., Rachel, R.A., Cogliati, T., Qian, H., Swaroop, A., 2014. Deletion of aryl (Müller) cells: sensing and responding to tissue stretch. Invest. Ophthalmol. Vis.
hydrocarbon receptor AHR in mice leads to subretinal accumulation of micro- Sci. 51, 1683e1690.
glia and RPE atrophy. Invest. Ophthalmol. Vis. Sci. 55, 6031e6040. Linser, P., Moscona, A.A., 1981. Carbonic anhydrase C in the neural retina: transition
Kingma, P.B., Bok, D., Ong, D.E., 1998. Bovine epidermal fatty acid-binding protein: from generalized to glia-specific cell localization during embryonic develop-
determination of ligand specificity and cellular localization in retina and testis. ment. Proc. Natl. Acad. Sci. U. S. A. 78, 7190e7194.
Biochemistry 37, 3250e3257. Liu, B., Hunter, D.J., Smith, A.A., Chen, S., Helms, J.A., 2014a. The capacity of neural
Kinkl, N., Ruiz, J., Vecino, E., Frasson, M., Sahel, J., Hicks, D., 2003. Possible crest-derived stem cells for ocular repair. Birth Defects Res. Part C Embryo
involvement of a fibroblast growth factor 9 (FGF9)-FGF receptor-3-mediated Today Rev. 102, 299e308.
pathway in adult pig retinal ganglion cell survival in vitro. Mol. Cell Neurosci. Liu, Q., Londraville, R.L., Azodi, E., Babb, S.G., Chiappini-Williamson, C., Marrs, J.A.,
23, 39e53. Raymond, P.A., 2002. Up-regulation of cadherin-2 and cadherin-4 in regener-
Kinoshita, J.H., 1986. Aldose reductase in the diabetic eye. XLIII Edward Jackson ating visual structures of adult zebrafish. Exp. Neurol. 177, 396e406.
memorial lecture. Am. J. Ophthalmol. 102, 685e692. Liu, R.T., Gao, J., Cao, S., Sandhu, N., Cui, J.Z., Chou, C.L., Fang, E., Matsubara, J.A., 2013.
Kitahara, A., Toyota, T., Kakizaki, M., Goto, Y., 1978. Activities of hepatic enzymes in Inflammatory mediators induced by amyloid-beta in the retina and RPE in vivo:
spontaneous diabetes rats produced by selective breeding of normal Wistar implications for inflammasome activation in age-related macular degeneration.
rats. Tohoku J. Exp. Med. 126, 7e11. Invest. Ophthalmol. Vis. Sci. 54, 2225e2237.
Kitano, S., Morgan, J., Caprioli, J., 1996. Hypoxic and excitotoxic damage to cultured Liu, X., Ye, F., Xiong, H., Hu, D., Limb, G.A., Xie, T., Peng, L., Yang, W., Sun, Y., Zhou, M.,
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 37

Song, E., Zhang, D.Y., 2014b. IL-1beta upregulates IL-8 production in human embryonic development throughout adulthood. J. Immunol. 188, 29e36.
Muller Cells through activation of the p38 MAPK and ERK1/2 signaling path- Morcos, Y., Chan-Ling, T., 2000. Concentration of astrocytic filaments at the retinal
ways. Inflammation 37, 1486e1495. optic nerve junction is coincident with the absence of intra-retinal myelination:
Liu, X., Ye, F., Xiong, H., Hu, D.-N., Limb, G.A., Xie, T., Peng, L., Zhang, P., Wei, Y., comparative and developmental evidence. J. Neurocytol. 29, 665e678.
Zhang, W., Wang, J., Wu, H., Leei, P., Songa, E., Zhangb, D.Y., 2014c. IL-1b induces Morishita, S., Oku, H., Horie, T., Tonari, M., Kida, T., Okubo, A., Sugiyama, T., Takai, S.,
IL-6 production in retinal Müller cells predominantly through the activation of Hara, H., Ikeda, T., 2014. Systemic simvastatin rescues retinal ganglion cells from
P38 MAPK/NF-kB signaling pathway. Exp. Cell Res. optic nerve injury possibly through suppression of astroglial NF-kappaB acti-
Ljubimov, A.V., Burgeson, R.E., Butkowski, R.J., Couchman, J.R., Zardi, L., Ninomiya, Y., vation. PLoS One 9, e84387.
Sado, Y., Huang, Z.S., Nesburn, A.B., Kenney, M.C., 1996. Basement membrane Morissette, N., Carbonetto, S., 1995. Laminin alpha 2 chain (M chain) is found within
abnormalities in human eyes with diabetic retinopathy. J. Histochem. Cyto- the pathway of avian and murine retinal projections. J. Neurosci. 15,
chem. 44, 1469e1479. 8067e8082.
Lopez-Colome, A.M., Martinez-Lozada, Z., Guillem, A.M., Lopez, E., Ortega, A., 2012. Morrison, J.C., 2006. Integrins in the optic nerve head: potential roles in glau-
Glutamate transporter-dependent mTOR phosphorylation in Muller glia cells. comatous optic neuropathy (an American Ophthalmological Society thesis).
ASN Neuro 4, e00095. Trans. Am. Ophthalmol. Soc. 104, 453e477.
Lorber, B., Berry, M., Douglas, M.R., Nakazawa, T., Logan, A., 2009. Activated retinal Moscona, A.A., Fox, L., Smith, J., Degenstein, L., 1985. Antiserum to lens antigens
glia promote neurite outgrowth of retinal ganglion cells via apolipoprotein E. immunostains Muller glia cells in the neural retina. Proc. Natl. Acad. Sci. U. S. A.
J. Neurosci. Res. 87, 2645e2652. 82, 5570e5573.
Lorber, B., Berry, M., Logan, A., 2008. Different factors promote axonal regeneration Muller, A., Hauk, T.G., Fischer, D., 2007. Astrocyte-derived CNTF switches mature
of adult rat retinal ganglion cells after lens injury and intravitreal peripheral RGCs to a regenerative state following inflammatory stimulation. Brain 130,
nerve grafting. J. Neurosci. Res. 86, 894e903. 3308e3320.
Lorber, B., Berry, M., Logan, A., Tonge, D., 2002. Effect of lens lesion on neurite Nagelhus, E.A., Ottersen, O.P., 2013. Physiological roles of aquaporin-4 in brain.
outgrowth of retinal ganglion cells in vitro. Mol. Cell. Neurosci. 21, 301e311. Physiol. Rev. 93, 1543e1562.
Lorber, B., Guidi, A., Fawcett, J.W., Martin, K.R., 2012. Activated retinal glia mediated Nahirnyj, A., Livne-Bar, I., Guo, X., Sivak, J.M., 2013. ROS detoxification and proin-
axon regeneration in experimental glaucoma. Neurobiol. Dis. 45, 243e252. flammatory cytokines are linked by p38 MAPK signaling in a model of mature
Lu, Y.-B., Pannicke, T., Wei, E.-Q., Bringmann, A., Wiedemann, P., Habermann, G., astrocyte activation. PLoS One 8, e83049.
Buse, E., Ka €s, J.A., Reichenbach, A., 2013. Biomechanical properties of retinal Nakazawa, T., Matsubara, A., Noda, K., Hisatomi, T., She, H., Skondra, D., Miyahara, S.,
glial cells: comparative and developmental data. Exp. Eye Res. 113, 60e65. Sobrin, L., Thomas, K.L., Chen, D.F., Grosskreutz, C.L., Hafezi-Moghadam, A.,
Lu, Y.B., Franze, K., Seifert, G., Steinhauser, C., Kirchhoff, F., Wolburg, H., Guck, J., Miller, J.W., 2006. Characterization of cytokine responses to retinal detachment
Janmey, P., Wei, E.Q., Kas, J., Reichenbach, A., 2006. Viscoelastic properties of in rats. Mol. Vis. 12, 867e878.
individual glial cells and neurons in the CNS. Proc. Natl. Acad. Sci. U. S. A. 103, Neugebauer, K.M., Tomaselli, K.J., Lilien, J., Reichardt, L.F., 1988. N-cadherin, NCAM,
17759e17764. and integrins promote retinal neurite outgrowth on astrocytes in vitro. J. Cell
Lu, Y.B., Iandiev, I., Hollborn, M., Korber, N., Ulbricht, E., Hirrlinger, P.G., Pannicke, T., Biol. 107, 1177e1187.
Wei, E.Q., Bringmann, A., Wolburg, H., Wilhelmsson, U., Pekny, M., Newman, E., Reichenbach, A., 1996. The Muller cell: a functional element of the
Wiedemann, P., Reichenbach, A., Kas, J.A., 2011. Reactive glial cells: increased retina. Trends Neurosci. 19, 307e312.
stiffness correlates with increased intermediate filament expression. FASEB J. Newman, E.A., 2003. New roles for astrocytes: regulation of synaptic transmission.
25, 624e631. Trends Neurosci. 26, 536e542.
Luna, G., Lewis, G.P., Banna, C.D., Skalli, O., Fisher, S.K., 2010. Expression profiles of Newman, E.A., Frambach, D.A., Odette, L.L., 1984. Control of extracellular potassium
nestin and synemin in reactive astrocytes and Muller cells following retinal levels by retinal glial cell Kþ siphoning. Science 225, 1174e1175.
injury: a comparison with glial fibrillar acidic protein and vimentin. Mol. Vis. 16, Ng, S.K., Wood, J.P., Chidlow, G., Han, G., Kittipassorn, T., Peet, D.J., Casson, R.J., 2014.
2511e2523. Cancer-like metabolism of the mammalian retina. Clin. Exp. Ophthalmol.
Lundkvis, A., Reichenbach, A., Betsholtz, C., Carmeliet, P., Wolburg, H., Pekny, M., Nguyen, J.V., Soto, I., Kim, K.Y., Bushong, E.A., Oglesby, E., Valiente-Soriano, F.J.,
2004. Under stress, the absence of intermediate filaments from Müller cells in Yang, Z., Davis, C.H., Bedont, J.L., Son, J.L., Wei, J.O., Buchman, V.L., Zack, D.J.,
the retina has structural and functional consequences. J. Cell. Sci. 117, Vidal-Sanz, M., Ellisman, M.H., Marsh-Armstrong, N., 2011. Myelination transi-
3481e3488. tion zone astrocytes are constitutively phagocytic and have synuclein depen-
Luo, C., Chen, M., Xu, H., 2011. Complement gene expression and regulation in dent reactivity in glaucoma. Proc. Natl. Acad. Sci. U. S. A. 108, 1176e1181.
mouse retina and retinal pigment epithelium/choroid. Mol. Vis. 17, 1588e1597. Noailles, A., Fernandez-Sanchez, L., Lax, P., Cuenca, N., 2014. Microglia activation in a
Lye-Barthel, M., Sun, D., Jakobs, T.C., 2013. Morphology of astrocytes in a glau- model of retinal degeneration and TUDCA neuroprotective effects.
comatous optic nerve. Invest. Ophthalmol. Vis. Sci. 54, 909e917. J. Neuroinflammation 11, 186.
Ma, W., Zhao, L., Fontainhas, A.M., Fariss, R.N., Wong, W.T., 2009. Microglia in the Oberheim, N.A., Takano, T., Han, X., He, W., Lin, J.H., Wang, F., Xu, Q., Wyatt, J.D.,
mouse retina alter the structure and function of retinal pigmented epithelial Pilcher, W., Ojemann, J.G., Ransom, B.R., Goldman, S.A., Nedergaard, M., 2009.
cells: a potential cellular interaction relevant to AMD. PLoS One 4, e7945. Uniquely hominid features of adult human astrocytes. J. Neurosci. 29,
Ma, W., Zhao, L., Wong, W.T., 2012. Microglia in the outer retina and their relevance 3276e3287.
to pathogenesis of age-related macular degeneration. Adv. Exp. Med. Biol. 723, Ogden, T.E., 1978. Nerve fiber layer astrocytes of the primate retina: morphology,
37e42. distribution, and density. Invest. Ophthalmol. Vis. Sci. 17, 499e510.
Ma, Y., Qu, Y., Fei, Z., 2011. Vascular endothelial growth factor in cerebral ischemia. Ola, M.S., Nawaz, M.I., Siddiquei, M.M., Al-Amro, S., Abu El-Asrar, A.M., 2012. Recent
J. Neurosci. Res. 89, 969e978. advances in understanding the biochemical and molecular mechanism of dia-
Mansford, K.R., Opie, L., 1968. Comparison of metabolic abnormalities in diabetes betic retinopathy. J. Diabetes Complicat. 26, 56e64.
mellitus induced by streptozotocin or by alloxan. Lancet 1, 670e671. Oosthuyse, B., Moons, L., Storkebaum, E., Beck, H., Nuyens, D., Brusselmans, K., Van
Marc, R.E., 1992. Structural organization of GABAergic circuitry in ectotherm retinas. Dorpe, J., Hellings, P., Gorselink, M., Heymans, S., Theilmeier, G., Dewerchin, M.,
Prog. Brain Res. 90, 61e92. Laudenbach, V., Vermylen, P., Raat, H., Acker, T., Vleminckx, V., Van Den
Mastellos, D.C., 2014. Complement emerges as a masterful regulator of CNS ho- Bosch, L., Cashman, N., Fujisawa, H., Drost, M.R., Sciot, R., Bruyninckx, F.,
meostasis, neural synaptic plasticity and cognitive function. Exp. Neurol. 261, Hicklin, D.J., Ince, C., Gressens, P., Lupu, F., Plate, K.H., Robberecht, W.,
469e474. Herbert, J.M., Collen, D., Carmeliet, P., 2001. Deletion of the hypoxia-response
Matteucci, A., Gaddini, L., Villa, M., Varano, M., Parravano, M.C., Monteleone, V., element in the vascular endothelial growth factor promoter causes motor
Cavallo, F., Leo, L., Mallozzi, C., Malchiodi-Albedi, F., Pricci, F., 2014. Neuro- neuron degeneration. Nat. Genet. 28, 131e138.
protection by rat Müller glia against high glucose-induced neurodegeneration Ooto, S., Akagi, T., Kageyama, R., Akita, J., Mandai, M., Honda, Y., Takahashi, M., 2004.
through a mechanism involving ERK1/2 activation. Exp. Eye Res. 125, 20e29. Potential for neural regeneration after neurotoxic injury in the adult mamma-
Matyash, V., Kettenmann, H., 2010. Heterogeneity in astrocyte morphology and lian retina. Proc. Natl. Acad. Sci. U. S. A. 101, 13654e13659.
physiology. Brain Res. Rev. 63, 2e10. Overby, D.R., Bertrand, J., Tektas, O.Y., Boussommier-Calleja, A., Schicht, M.,
Mauch, D.H., Nagler, K., Schumacher, S., Goritz, C., Muller, E.C., Otto, A., Ethier, C.R., Woodward, D.F., Stamer, W.D., Lutjen-Drecoll, E., 2014. Ultrastruc-
Pfrieger, F.W., 2001. CNS synaptogenesis promoted by glia-derived cholesterol. tural changes associated with dexamethasone-induced ocular hypertension in
Science 294, 1354e1357. mice. Invest. Ophthalmol. Vis. Sci. 55, 4922e4933.
Medzhitov, R., 2008. Origin and physiological roles of inflammation. Nature 454, Ozaki, H., Seo, M.S., Ozaki, K., Yamada, H., Yamada, E., Okamoto, N., Hofmann, F.,
428e435. Wood, J.M., Campochiaro, P.A., 2000. Blockade of vascular endothelial cell
Medzhitov, R., 2010. Inflammation 2010: new adventures of an old flame. Cell 140, growth factor receptor signaling is sufficient to completely prevent retinal
771e776. neovascularization. Am. J. Pathol. 156, 697e707.
Mesquida, M., Leszczynska, A., Llorenc, V., Adan, A., 2014. Interleukin-6 blockade in Pannicke, T., Wurm, A., Iandiev, I., Hollborn, M., Linnertz, R., Binder, D.K., Kohen, L.,
ocular inflammatory diseases. Clin. Exp. Immunol. 176, 301e309. Wiedemann, P., Steinha €user, C., Reichenbach, A., Bringmann, A., 2010. Deletion
Mi, X.S., Yuan, T.F., So, K.F., 2014. The current research status of normal tension of aquaporin-4 renders retinal glial cells more susceptible to osmotic stress.
glaucoma. Clin. Interv. Aging 9, 1563e1571. J. Neurosci. Res. 88, 2877e2888.
Miyamoto, A., Wake, H., Moorhouse, A.J., Nabekura, J., 2013. Microglia and synapse Paolicelli, R.C., Bolasco, G., Pagani, F., Maggi, L., Scianni, M., Panzanelli, P.,
interactions: fine tuning neural circuits and candidate molecules. Front. Cell. Giustetto, M., Ferreira, T.A., Guiducci, E., Dumas, L., Ragozzino, D., Gross, C.T.,
Neurosci. 7, 70. 2011. Synaptic pruning by microglia is necessary for normal brain development.
Mizutani, M., Pino, P.A., Saederup, N., Charo, I.F., Ransohoff, R.M., Cardona, A.E., Science 333, 1456e1458.
2012. The fractalkine receptor but not CCR2 is present on microglia from Park, S., Lee, Y.J., 2013. Nano-mechanical compliance of Müller cells investigated by
38 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

atomic force microscopy. Int. J. Biol. Sci. 9, 702e706. J. Chem. Neuroanat. 6, 201e213.
Parysek, L.M., del Cerro, M., Olmsted, J.B., 1985. Microtubule-associated protein 4 Reichenbach, A., Wurm, A., Pannicke, T., Landiev, I., Wiedemann, P., Bringmann, A.,
antibody: a new marker for astroglia and oligodendroglia. Neuroscience 15, 2007. Muller cells as players in retinal degeneration and edema. Graefe Arch.
869e875. Clin. Exp. Ophthalmol. 245, 627e636.
Pasti, L., Volterra, A., Pozzan, T., Carmignoto, G., 1997. Intracellular calcium oscilla- Reyes-Aguirre, L.I., Ferraro, S., Quintero, H., Sa nchez-Serrano, S.L., Go mez-
tions in astrocytes: a highly plastic, bidirectional form of communication be- Montalvo, A., Lamas, M., 2013. Glutamate-induced epigenetic and morpholog-
tween neurons and astrocytes in situ. J. Neurosci. 17, 7817e7830. ical changes allow rat Müller cell dedifferentiation but not further acquisition of
Paulson, O.B., Newman, E.A., 1987. Does the release of potassium from astrocyte a photoreceptor phenotype. Neuroscience 254, 347e360.
endfeet regulate cerebral blood flow? Science 237, 896e898. Rezaie, P., Male, D., 1999. Colonisation of the developing human brain and spinal
Pekny, M., 2001. Astrocytic intermediate filaments: lessons from GFAP and vimentin cord by microglia: a review. Microsc. Res. Tech. 45, 359e382.
knock-out mice. Prog. Brain Res. 132, 23e30. Rhee, K.D., Nusinowitz, S., Chao, K., Yu, F., Bok, D., Yang, X.J., 2013. CNTF-mediated
Pekny, M., Nilsson, M., 2005. Astrocyte activation and reactive gliosis. Glia 50, protection of photoreceptors requires initial activation of the cytokine receptor
427e434. gp130 in Muller glial cells. Proc. Natl. Acad. Sci. U. S. A. 110, E4520eE4529.
Pekny, M., Pekna, M., 2014. Astrocyte reactivity and reactive astrogliosis: costs and Rhee, K.D., Yang, X.J., 2010. Function and mechanism of CNTF/LIF signaling in ret-
benefits. Physiol. Rev. 94, 1077e1098. inogenesis. Adv. Exp. Med. Biol. 664, 647e654.
Pekny, M., Wilhelmsson, U., Pekna, M., 2014. The dual role of astrocyte activation Rieck, J., 2013. The pathogenesis of glaucoma in the interplay with the immune
and reactive gliosis. Neurosci. Lett. 565, 30e38. system. Invest. Ophthalmol. Vis. Sci. 54, 2393e2409.
Pena, J.D., Varela, H.J., Ricard, C.S., Hernandez, M.R., 1999. Enhanced tenascin Riehl, R., Johnson, K., Bradley, R., Grunwald, G.B., Cornel, E., Lilienbaum, A., Holt, C.E.,
expression associated with reactive astrocytes in human optic nerve heads with 1996. Cadherin function is required for axon outgrowth in retinal ganglion cells
primary open angle glaucoma. Exp. Eye Res. 68, 29e40. in vivo. Neuron 17, 837e848.
Peng, B., Xiao, J., Wang, K., So, K.F., Tipoe, G.L., Lin, B., 2014. Suppression of microglial Roberge, F.G., Caspi, R.R., Chan, C.C., Kuwabara, T., Nussenblatt, R.B., 1985. Long-term
activation is neuroprotective in a mouse model of human retinitis pigmentosa. culture of Muller cells from adult rats in the presence of activated lymphocytes/
J. Neurosci. 34, 8139e8150. monocytes products. Curr. Eye Res. 4, 975e982.
Perez-Alvarez, M.J., Isiegas, C., Santano, C., Salazar, J.J., Ramirez, A.I., Trivino, A., Roberge, F.G., Caspi, R.R., Chan, C.C., Nussenblatt, R.B., 1991. Inhibition of T
Ramirez, J.M., Albar, J.P., de la Rosa, E.J., Prada, C., 2008. Vimentin isoform lymphocyte proliferation by retinal glial Muller cells: reversal of inhibition by
expression in the human retina characterized with the monoclonal antibody glucocorticoids. J. Autoimmun. 4, 307e314.
3CB2. J. Neurosci. Res. 86, 1871e1883. Roesch, K., Jadhav, A.P., Trimarchi, J.M., Stadler, M.B., Roska, B., Sun, B.B., Cepko, C.L.,
Pfrieger, F.W., Barres, B.A., 1996. New views on synapse-glia interactions. Curr. Opin. 2008. The transcriptome of retinal Muller glial cells. J. Comp. Neurol. 509,
Neurobiol. 6, 615e621. 225e238.
Politi, L.E., Rotstein, N.P., Carri, N.G., 2001. Effect of GDNF on neuroblast proliferation Roesch, K., Stadler, M.B., Cepko, C.L., 2012. Gene expression changes within Muller
and photoreceptor survival: additive protection with docosahexaenoic acid. glial cells in retinitis pigmentosa. Mol. Vis. 18, 1197e1214.
Invest. Ophthalmol. Vis. Sci. 42, 3008e3015. Rogers, R.S., Dharsee, M., Ackloo, S., Sivak, J.M., Flanagan, J.G., 2012. Proteomics
Pollak, J., Wilken, M.S., Ueki, Y., Cox, K.E., Sullivan, J.M., Taylor, R.J., Levine, E.M., analyses of human optic nerve head astrocytes following biomechanical strain.
Reh, T.A., 2013. ASCL1 reprograms mouse Müller glia into neurogenic retinal Mol. Cell. Proteomics 11 (M111), 012302.
progenitors. Development 140, 2619e2631. Rohen, J.W., Futa, R., Lutjen-Drecoll, E., 1981. The fine structure of the cribriform
Ponsioen, T.L., van der Worp, R.J., van Luyn, M.J., Hooymans, J.M., Los, L.I., 2005. meshwork in normal and glaucomatous eyes as seen in tangential sections.
Packages of vitreous collagen (type II) in the human retina: an indication of Invest. Ophthalmol. Vis. Sci. 21, 574e585.
postnatal collagen turnover? Exp. Eye Res. 80, 643e650. Rojas, B., Gallego, B.I., Ramirez, A.I., Salazar, J.J., de Hoz, R., Valiente-Soriano, F.J.,
Ponsioen, T.L., van Luyn, M.J., van der Worp, R.J., Pas, H.H., Hooymans, J.M., Los, L.I., Aviles-Trigueros, M., Villegas-Perez, M.P., Vidal-Sanz, M., Trivino, A.,
2008. Human retinal Muller cells synthesize collagens of the vitreous and vit- Ramirez, J.M., 2014. Microglia in mouse retina contralateral to experimental
reoretinal interface in vitro. Mol. Vis. 14, 652e660. glaucoma exhibit multiple signs of activation in all retinal layers.
Porter, J.T., McCarthy, K.D., 1997. Astrocytic neurotransmitter receptors in situ and J. Neuroinflammation 11, 133e2094, 2011-2133.
in vivo. Prog. Neurobiol. 51, 439e455. Rowan, S., Cepko, C.L., 2004. Genetic analysis of the homeodomain transcription
Prada, I., Furlan, R., Matteoli, M., Verderio, C., 2013. Classical and unconventional factor Chx10 in the retina using a novel multifunctional BAC transgenic mouse
pathways of vesicular release in microglia. Glia 61, 1003e1017. reporter. Dev. Biol. 271, 388e402.
Prinz, M., Priller, J., 2014. Microglia and brain macrophages in the molecular age: Ruilin, Z., Kin-Sang, C., Dong Feng, C., Liu, Y., 2014. Ephrin-A2 and -A3 are negative
from origin to neuropsychiatric disease. Nat. Rev. Neurosci. 15, 300e312. regulators of the regenerative potential of Müller cells. Chin. Med. J. 127,
Prinz, M., Priller, J., Sisodia, S.S., Ransohoff, R.M., 2011. Heterogeneity of CNS myeloid 3438e3442.
cells and their roles in neurodegeneration. Nat. Neurosci. 14, 1227e1235. Ruiz-Ederra, J., Garcia, M., Hernandez, M., Urcola, H., Hernandez-Barbachano, E.,
Qin, Y., Ren, H., Hoffman, M.R., Fan, J., Zhang, M., Xu, G., 2012. Aquaporin changes Araiz, J., Vecino, E., 2005. The pig eye as a novel model of glaucoma. Exp. Eye
during diabetic retinopathy in rats are accelerated by systemic hypertension Res. 81, 561e569.
and are linked to the renin-angiotensin system. Invest. Ophthalmol. Vis. Sci. 53, Ruiz-Ederra, J., Hitchcock, P.F., Vecino, E., 2003. Two classes of astrocytes in the
3047e3053. adult human and pig retina in terms of their expression of high affinity NGF
Quigley, H.A., 2011. Glaucoma. Lancet 377, 1367e1377. receptor (TrkA). Neurosci. Lett. 337, 127e130.
Ramirez, A.I., Salazar, J.J., de Hoz, R., Rojas, B., Gallego, B.I., Salinas-Navarro, M., Rungger-Brandle, E., Messerli, J.M., Niemeyer, G., Eppenberger, H.M., 1993. Confocal
Alarcon-Martinez, L., Ortin-Martinez, A., Aviles-Trigueros, M., Vidal-Sanz, M., microscopy and computer-assisted image reconstruction of astrocytes in the
Trivino, A., Ramirez, J.M., 2010. Quantification of the effect of different levels of mammalian retina. Eur. J. Neurosci. 5, 1093e1106.
IOP in the astroglia of the rat retina ipsilateral and contralateral to experimental Ruzafa, N., Rey-Santano, C., Mielgo, V., Rodríguez-Fernandez, D., Vecino, E., 2014.
glaucoma. Invest. Ophthalmol. Vis. Sci. 51, 5690e5696. SIRCOVA-OFTARED-RIG Joined Congress Abstracts: effect of hypoxia in neonatal
Ramirez, M., Lamas, M., 2009. NMDA receptor mediates proliferation and CREB pig retina and Superior Colliculus. Ophthalmic Res. 52, 175e197.
phosphorylation in postnatal Müller glia-derived retinal progenitors. Mol. Vis. Ryskamp, D.A., Witkovsky, P., Barabas, P., Huang, W., Koehler, C., Akimov, N.P.,
15, 713e721. Lee, S.H., Chauhan, S., Xing, W., Renteria, R.C., Liedtke, W., Krizaj, D., 2011. The
Ransohoff, R.M., Brown, M.A., 2012. Innate immunity in the central nervous system. polymodal ion channel transient receptor potential vanilloid 4 modulates cal-
J. Clin. Invest. 122, 1164e1171. cium flux, spiking rate, and apoptosis of mouse retinal ganglion cells.
Ransohoff, R.M., Perry, V.H., 2009. Microglial physiology: unique stimuli, specialized J. Neurosci. 31, 7089e7101.
responses. Annu. Rev. Immunol. 27, 119e145. Saari, J.C., 2000. Biochemistry of visual pigment regeneration: the Friedenwald
Reichenbach, A., Bringmann, A., 2010. Müller Cells in the Healthy and Diseased lecture. Invest. Ophthalmol. Vis. Sci. 41, 337e348.
Retina. Springer, New York. Saederup, N., Cardona, A.E., Croft, K., Mizutani, M., Cotleur, A.C., Tsou, C.L.,
Reichenbach, A., Bringmann, A., 2013. New Funct. Müller Cells Glia 61, 651e678. Ransohoff, R.M., Charo, I.F., 2010. Selective chemokine receptor usage by central
Reichenbach, A., Bringmann, A., 2015. Retinal Glia. Morgan & Claypool Publishers. nervous system myeloid cells in CCR2-red fluorescent protein knock-in mice.
Reichenbach, A., Hagen, E., Schippel, K., Bruckner, G., Reichelt, W., Leibnitz, L., 1988. PLoS One 5, e13693.
Cytotopographical specialization of enzymatically isolated rabbit retinal Muller Salazar, J.J., Gallego-Pinazo, R., de Hoz, R., Pinazo-Duran, M.D., Rojas, B.,
(glial) cells: structure, ultrastructure, and 3H-ouabain binding sites. Z. Mikrosk. Ramirez, A.I., Serrano, M., Ramirez, J.M., 2013. “Super p53” mice display retinal
Anat. Forsch 102, 897e912. astroglial changes. PLoS One 8, e65446.
Reichenbach, A., Robinson, S.R., 1995. Phylogenetic constraints on retinal organi- Sanderson, J., Dartt, D.A., Trinkaus-Randall, V., Pintor, J., Civan, M.M., Delamere, N.A.,
zation and development. In: Osborne, N., Chader, G. (Eds.), Progress in Retinal Fletcher, E.L., Salt, T.E., Grosche, A., Mitchell, C.H., 2014. Purines in the eye:
and Eye Research. Pergamon Press., Oxford. recent evidence for the physiological and pathological role of purines in the
Reichenbach, A., Schneider, H., Leibnitz, L., Reichelt, W., Schaaf, P., Schumann, R., RPE, retinal neurons, astrocytes, Müller cells, lens, trabecular meshwork, cornea
1989. The structure of rabbit retinal Muller (glial) cells is adapted to the sur- and lacrimal gland. Exp. Eye Res. 127, 270e279.
rounding retinal layers. Anat. Embryol. (Berl.) 180, 71e79. Santiago, A.R., Baptista, F.I., Santos, P.F., Cristovao, G., Ambrosio, A.F., Cunha, R.A.,
Reichenbach, A., Siegel, A., Rickmann, M., Wolff, J.R., Noone, D., Robinson, S.R., 1995. Gomes, C.A., 2014. Role of microglia adenosine A(2A) receptors in retinal and
Distribution of Bergmann glial somata and processes: implications for function. brain neurodegenerative diseases. Mediat. Inflamm. 2014, 465694.
J. Hirnforsch. 36, 509e517. Santos-Bredariol, A.S., Belmonte, M.A., Kihara, A.H., Santos, M.F., Hamassaki, D.E.,
Reichenbach, A., Stolzenburg, J.U., Eberhardt, W., Chao, T.I., Dettmer, D., Hertz, L., 2006. Small GTP-binding protein RhoB is expressed in glial Muller cells in the
1993. What do retinal muller (glial) cells do for their neuronal 'small siblings'? vertebrate retina. J. Comp. Neurol. 494, 976e985.
E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40 39

Sarthy, P.V., Fu, M., 1990. Localization of laminin B1 mRNA in retinal ganglion cells Prog. Retin. Eye Res. 18, 689e735.
by in situ hybridization. J. Cell Biol. 110, 2099e2108. Sullivan, L.S., Koboldt, D.C., Bowne, S.J., Lang, S., Blanton, S.H., Cadena, E., Avery, C.E.,
Sarthy, V., Ripps, H., 2001. Role in retinal pathophysiology. In: Blakemore, C. (Ed.), Lewis, R.A., Webb-Jones, K., Wheaton, D.H., Birch, D.G., Coussa, R., Ren, H.,
The Retinal Müller Cell. Structure and Function. Kluwer Academic/Plenum Lopez, I., Chakarova, C., Koenekoop, R.K., Garcia, C.A., Fulton, R.S., Wilson, R.K.,
Publishers, New York, pp. 181e215. Weinstock, G.M., Daiger, S.P., 2014. A dominant mutation in hexokinase 1 (HK1)
Schafer, D.P., Lehrman, E.K., Kautzman, A.G., Koyama, R., Mardinly, A.R., Yamasaki, R., causes retinitis pigmentosa. Invest. Ophthalmol. Vis. Sci. 55, 7147e7158.
Ransohoff, R.M., Greenberg, M.E., Barres, B.A., Stevens, B., 2012. Microglia sculpt Takeda, M., Takamiya, A., Jiao, J.W., Cho, K.S., Trevino, S.G., Matsuda, T., Chen, D.F.,
postnatal neural circuits in an activity and complement-dependent manner. 2008. Alpha-aminoadipate induces progenitor cell properties of Müller glia in
Neuron 74, 691e705. adult mice. Invest. Ophthalmol. Vis. Sci. 49, 1142e1150.
Schafer, D.P., Lehrman, E.K., Stevens, B., 2013. The “quad-partite” synapse: Tang, J., Du, Y., Petrash, J.M., Sheibani, N., Kern, T.S., 2013. Deletion of aldose
microglia-synapse interactions in the developing and mature CNS. Glia 61, reductase from mice inhibits diabetes-induced retinal capillary degeneration
24e36. and superoxide generation. PLoS One 8, e62081.
Scheef, E., Wang, S., Sorenson, C.M., Sheibani, N., 2005. Isolation and characteriza- Tezel, G., Yang, X., Luo, C., Cai, J., Powell, D.W., 2012. An astrocyte-specific proteomic
tion of murine retinal astrocytes. Mol. Vis. 11, 613e624. approach to inflammatory responses in experimental rat glaucoma. Invest.
Schubert, H.D., 1989. Cystoid macular edema: the apparent role of mechanical Ophthalmol. Vis. Sci. 53, 4220e4233.
factors. Prog. Clin. Biol. Res. 312, 277e291. Thanos, S., Kacza, J., Seeger, J., Mey, J., 1994. Old dyes for new scopes: the
Schulz, C., Gomez Perdiguero, E., Chorro, L., Szabo-Rogers, H., Cagnard, N., phagocytosis-dependent long-term fluorescence labelling of microglial cells
Kierdorf, K., Prinz, M., Wu, B., Jacobsen, S.E., Pollard, J.W., Frampton, J., Liu, K.J., in vivo. Trends Neurosci. 17, 177e182.
Geissmann, F., 2012. A lineage of myeloid cells independent of Myb and he- Toft-Kehler, A.K., Skytt, D.M., Poulsen, K.A., Braendstrup, C.T., Gegelashvili, G.,
matopoietic stem cells. Science (New York, N. Y. 336, 86e90. Waagepetersen, H., Kolko, M., 2014. Limited energy supply in Muller cells alters
Sheppard, A.M., Onken, M.D., Rosen, G.D., Noakes, P.G., Dean, D.C., 1994. Expanding glutamate uptake. Neurochem. Res. 39, 941e949.
roles for alpha 4 integrin and its ligands in development. Cell Adhes. Commun. Tolentino, M.J., McLeod, D.S., Taomoto, M., Otsuji, T., Adamis, A.P., Lutty, G.A., 2002.
2, 27e43. Pathologic features of vascular endothelial growth factor-induced retinopathy
Sheridan, G.K., Murphy, K.J., 2013. Neuron-glia crosstalk in health and disease: in the nonhuman primate. Am. J. Ophthalmol. 133, 373e385.
fractalkine and CX3CR1 take centre stage. Open Biol. 3, 130181. Tran, T.L., Bek, T., la Cour, M., Nielsen, S., Prause, J.U., Hamann, S., Heegaard, S., 2014.
Shin, E.S., Huang, Q., Gurel, Z., Sorenson, C.M., Sheibani, N., 2014. High glucose alters Altered aquaporin expression in glaucoma eyes. APMIS 122, 772e780.
retinal astrocytes phenotype through increased production of inflammatory Tremblay, M.E., Lowery, R.L., Majewska, A.K., 2010. Microglial interactions with
cytokines and oxidative stress. PLoS One 9, e103148. synapses are modulated by visual experience. PLoS Biol. 8, e1000527.
Siddiqui, S., Horvat-Brocker, A., Faissner, A., 2008. The glia-derived extracellular Tserentsoodol, N., Sztein, J., Campos, M., Gordiyenko, N.V., Fariss, R.N., Lee, J.W.,
matrix glycoprotein tenascin-C promotes embryonic and postnatal retina axon Fliesler, S.J., Rodriguez, I.R., 2006. Uptake of cholesterol by the retina occurs
outgrowth via the alternatively spliced fibronectin type III domain TNfnD. primarily via a low density lipoprotein receptor-mediated process. Mol. Vis. 12,
Neuron Glia Biol. 4, 271e283. 1306e1318.
Siddiqui, S., Horvat-Broecker, A., Faissner, A., 2009. Comparative screening of glial Tuo, J., Bojanowski, C.M., Zhou, M., Shen, D., Ross, R.J., Rosenberg, K.I., Cameron, D.J.,
cell types reveals extracellular matrix that inhibits retinal axon growth in a Yin, C., Kowalak, J.A., Zhuang, Z., Zhang, K., Chan, C.C., 2007. Murine ccl2/cx3cr1
chondroitinase ABC-resistant fashion. Glia 57, 1420e1438. deficiency results in retinal lesions mimicking human age-related macular
Sidman, R.L., 1961. Histogenesis of mouse retina studied with thymidine-H3. In: degeneration. Invest. Ophthalmol. Vis. Sci. 48, 3827e3836.
Smelser, G.K. (Ed.), The Structure of the Eye. Academic Press, New York and Turner, D.L., Cepko, C.L., 1987. A common progenitor for neurons and glia persists in
London, pp. 487e506. rat retina late in development. Nature 328, 131e136.
Singh, P.K., Shiha, M.J., Kumar, A., 2014. Antibacterial responses of retinal Müller Uckermann, O., Vargova, L., Ulbricht, E., Klaus, C., Weick, M., Rillich, K.,
glia: production of antimicrobial peptides, oxidative burst and phagocytosis. Wiedemann, P., Reichenbach, A., Sykova, E., Bringmann, A., 2004. Glutamate-
J. Neuroinflammation 11, 33. evoked alterations of glial and neuronal cell morphology in the guinea pig
Singhal, S., Bhatia, B., Jayaram, H., Becker, S., Jones, M.F., Cottrill, P.B., Khaw, P.T., retina. J. Neurosci. 24, 10149e10158.
Salt, T.E., Limb, G.A., 2012. Human Muller glia with stem cell characteristics Uckermann, O., Wolf, A., Kutzera, F., Kalisch, F., Beck-Sickinger, A.G., Wiedemann, P.,
differentiate into retinal ganglion cell (RGC) precursors in vitro and partially Reichenbach, A., Bringmann, A., 2006. Glutamate release by neurons evokes a
restore RGC function in vivo following transplantation. Stem Cells Transl. Med. purinergic inhibitory mechanism of osmotic glial cell swelling in the rat retina:
1, 188e199. activation by neuropeptide Y. J. Neurosci. Res. 83, 538e550.
Skoff, R.P., 1990. Gliogenesis in rat optic nerve: astrocytes are generated in a single Ueki, Y., Karl, M.O., Sudar, S., Pollak, J., Taylor, R.J., Loeffler, K., Wilken, M.S.,
wave before oligodendrocytes. Dev. Biol. 139, 149e168. Reardon, S., Reh, T.A., 2012. P53 is required for the developmental restriction in
Smedowski, A., Pietrucha-Dutczak, M., Kaarniranta, K., Lewin-Kowalik, J., 2014. A rat Muller glial proliferation in mouse retina. Glia 60, 1579e1589.
experimental model of glaucoma incorporating rapid-onset elevation of intra- Ueki, Y., Reh, T.A., 2013. EGF stimulates Muller glial proliferation via a BMP-
ocular pressure. Sci. Rep. 4, 5910. dependent mechanism. Glia 61, 778e789.
Sobrado-Calvo, P., Vidal-Sanz, M., Villegas-Perez, M.P., 2007. Rat retinal microglial Uga, S., Smelser, 1973. Comparative study of the fine structure of retinal Muller cells
cells under normal conditions, after optic nerve section, and after optic nerve in various vertebrates. Invest. Ophthalmol. 12, 434e448.
section and intravitreal injection of trophic factors or macrophage inhibitory Ullian, E.M., Sapperstein, S.K., Christopherson, K.S., Barres, B.A., 2001. Control of
factor. J. Comp. Neurol. 501, 866e878. synapse number by glia. Science 291, 657e661.
Sofroniew, M.V., 2009. Molecular dissection of reactive astrogliosis and glial scar Unterlauft, J.D., Claudepierre, T., Schmidt, M., Müller, K., Yafai, Y., Wiedemann, P.,
formation. Trends Neurosci. 32, 638e647. Reichenbach, A., Eichler, W., 2014. Enhanced survival of retinal ganglion cells is
Song, W.T., Zhang, X.Y., Xia, X.B., 2013. Atoh7 promotes the differentiation of retinal mediated by Müller glial cell-derived PEDF. Exp. Eye Res. 127, 206e214.
stem cells derived from Muller cells into retinal ganglion cells by inhibiting Unterlauft, J.D., Eichler, W., Kuhne, K., Yang, X.M., Yafai, Y., Wiedemann, P.,
Notch signaling. Stem. Cell Res. Ther. 4, 94. Reichenbach, A., Claudepierre, T., 2012. Pigment epithelium-derived factor
Stasi, K., Nagel, D., Yang, X., Wang, R.F., Ren, L., Podos, S.M., Mittag, T., Danias, J., released by Muller glial cells exerts neuroprotective effects on retinal ganglion
2006. Complement component 1Q (C1Q) upregulation in retina of murine, cells. Neurochem. Res. 37, 1524e1533.
primate, and human glaucomatous eyes. Invest. Ophthalmol. Vis. Sci. 47, Valamanesh, F., Monnin, J., Morand-Villeneuve, N., Michel, G., Zaher, M., Miloudi, S.,
1024e1029. Chemouni, D., Jeanny, J.C., Versaux-Botteri, C., 2013. Nestin expression in the
Steele, M.R., Inman, D.M., Calkins, D.J., Horner, P.J., Vetter, M.L., 2006. Microarray retina of rats with inherited retinal degeneration. Exp. Eye Res. 110, 26e34.
analysis of retinal gene expression in the DBA/2J model of glaucoma. Invest. Vecino, E., 2008. Animal models in the study of the glaucoma: past, present and
Ophthalmol. Vis. Sci. 47, 977e985. future. Arch. Soc. Espanola Oftalmol. 83, 517e519.
Stephan, A.H., Barres, B.A., Stevens, B., 2012. The complement system: an unex- Vecino, E., Acera, A., 2015. Development and programed cell death in the eye. Int. J.
pected role in synaptic pruning during development and disease. Annu. Rev. Dev. Biol. http://dx.doi.org/ijdb.150070ev (in press).
Neurosci. 35, 369e389. Vecino, E., Caminos, E., Becker, E., Martin-Zanca, D., Osborne, N., 1998a. Expression
Stevens, B., 2008. Neuron-astrocyte signaling in the development and plasticity of of neurotrophins and their receptors within the glial cells of retina and optic
neural circuits. Neurosignals 16, 278e288. nerve. In: Castellano, B., Gonz alez, B., Nieto-Sampedro, M. (Eds.), Understanding
Stevens, B., Allen, N.J., Vazquez, L.E., Howell, G.R., Christopherson, K.S., Nouri, N., Glial Cells. Kluwer Academic Publisher, pp. 149e166.
Micheva, K.D., Mehalow, A.K., Huberman, A.D., Stafford, B., Sher, A., Litke, A.M., Vecino, E., Caminos, E., Ugarte, M., Martin-Zanca, D., Osborne, N.N., 1998b. Immu-
Lambris, J.D., Smith, S.J., John, S.W., Barres, B.A., 2007. The classical complement nohistochemical distribution of neurotrophins and their receptors in the rat
cascade mediates CNS synapse elimination. Cell 131, 1164e1178. retina and the effects of ischemia and reperfusion. Gen. Pharmacol. 30,
Stone, J., Dreher, Z., 1987. Relationship between astrocytes, ganglion cells and 305e314.
vasculature of the retina. J. Comp. Neurol. 255, 35e49. Vecino, E., Galdos, M., Bayon, A., Rodriguez, F.D., Mico, C., Sharma, S.C., 2015a.
Stone, J., Itin, A., Alon, T., Pe'er, J., Gnessin, H., Chan-Ling, T., Keshet, E., 1995. Elevated intraoucular pressure induces ultrastructural changes in the trabecular
Development of retinal vasculature is mediated by hypoxia-induced vascular meshwork. J. Citol. Histol. http://dx.doi.org/10.4172/2157-7099.S3-007 (in press).
endothelial growth factor (VEGF) expression by neuroglia. J. Neurosci. 15, Vecino, E., Garcia-Crespo, D., Garcia, M., Martinez-Millan, L., Sharma, S.C.,
4738e4747. Carrascal, E., 2002. Rat retinal ganglion cells co-express brain derived neuro-
Stone, J., Maslim, J., Valter-Kocsi, K., Mervin, K., Bowers, F., Chu, Y., Barnett, N., trophic factor (BDNF) and its receptor TrkB. Vis. Res. 42, 151e157.
Provis, J., Lewis, G., Fisher, S.K., Bisti, S., Gargini, C., Cervetto, L., Merin, S., Peer, J., Vecino, E., Heller, J.P., Veiga-Crespo, P., Martin, K.R., Fawcett, J., 2015b. Influence of
1999. Mechanisms of photoreceptor death and survival in mammalian retina. different extracellular matrix components on the expression of integrins and
40 E. Vecino et al. / Progress in Retinal and Eye Research 51 (2016) 1e40

regeneration of adult retinal ganglion cells. PLoS One (in press). Xue, L.P., Lu, J., Cao, Q., Hu, S., Ding, P., Ling, E.A., 2006a. Muller glial cells express
Vecino, E., Sharma, S., 2011. Glaucoma animal models. In: Rumelt, S. (Ed.), Glaucoma nestin coupled with glial fibrillary acidic protein in experimentally induced
e Basic and Clinical Concepts. In Tech, pp. 319e334. glaucoma in the rat retina. Neuroscience 139, 723e732.
Vecino, E., Ugarte, M., Nash, M.S., Osborne, N.N., 1999. NMDA induces BDNF Xue, L.P., Lu, J., Cao, Q., Kaur, C., Ling, E.A., 2006b. Nestin expression in Muller glial
expression in the albino rat retina in vivo. Neuroreport 10, 1103e1106. cells in postnatal rat retina and its upregulation following optic nerve tran-
Vessey, K.A., Greferath, U., Aplin, F.P., Jobling, A.I., Phipps, J.A., Ho, T., De Iongh, R.U., section. Neuroscience 143, 117e127.
Fletcher, E.L., 2014. Adenosine triphosphate-induced photoreceptor death and Xue, W., Cojocaru, R.I., Dudley, V.J., Brooks, M., Swaroop, A., Sarthy, V.P., 2011. Ciliary
retinal remodeling in rats. J. Comp. Neurol. 522, 2928e2950. neurotrophic factor induces genes associated with inflammation and gliosis in
Villacampa, N., Almolda, B., Gonzalez, B., Castellano, B., 2013. Tomato lectin histo- the retina: a gene profiling study of flow-sorted, Muller cells. PLoS One 6,
chemistry for microglial visualization. Methods Mol. Biol. 1041, 261e279. e20326.
Vogler, S., Grosche, A., Pannicke, T., Ulbricht, E., Wiedemann, P., Reichenbach, A., Yafai, Y., Iandiev, I., Lange, J., Unterlauft, J.D., Wiedemann, P., Bringmann, A.,
Bringmann, A., 2013. Hypoosmotic and glutamate-induced swelling of bipolar Reichenbach, A., Eichler, W., 2014. Müller glial cells inhibit proliferation of
cells in the rat retina: comparison with swelling of Müller glial cells. retinal endothelial cells via TGF-b2 and smad signaling. Glia 62, 1476e1485.
J. Neurochem. 126, 372e381. Yafai, Y., Iandiev, I., Lange, J., Yang, X.M., Wiedemann, P., Bringmann, A., Eichler, W.,
Voskuhl, R.R., Peterson, R.S., Song, B., Ao, Y., Morales, L.B., Tiwari-Woodruff, S., 2013. Basic fibroblast growth factor contributes to a shift in the Angioregulatory
Sofroniew, M.V., 2009. Reactive astrocytes form scar-like perivascular barriers activity of retinal glial (Müller cells). PLoS One 8, e68773.
to leukocytes during adaptive immune inflammation of the CNS. J. Neurosci. 29, Yafai, Y., Lange, J., Wiedemann, P., Reichenbach, A., Eichler, W., 2007. Pigment
11511e11522. epithelium-derived factor acts as an opponent of growth-stimulatory factors in
Wahl, V., Vogler, S., Grosche, A., Pannicke, T., Ueffing, M., Wiedemann, P., retinal glial-endothelial cell interactions. Glia 55, 642e651.
Reichenbach, A., Hauck, S.M., Bringmann, A., 2013. Osteopontin inhibits osmotic Yan, Q., Li, Y., Hendrickson, A., Sage, E.H., 2001. Regulation of retinal capillary cells
swelling of retinal glial (Muller) cells by inducing release of VEGF. Neuroscience by basic fibroblast growth factor, vascular endothelial growth factor, and hyp-
246, 59e72. oxia. In Vitro Cell Dev. Biol. Anim. 37, 45e49.
Wang, A.G., Yen, M.Y., Hsu, W.M., Fann, M.J., 2006. Induction of vitronectin and Yang, L., Xu, Y., Li, W., Yang, B., Yu, S., Zhou, H., Yang, C., Xu, F., Wang, J., Gao, Y.,
integrin alphav in the retina after optic nerve injury. Mol. Vis. 12, 76e84. Huang, Y., Lu, L., Liang, X., 2014. Diacylglycerol Kinase (DGK) Inhibitor II
Wang, B., Xiao, Z., Chen, B., Han, J., Gao, Y., Zhang, J., Zhao, W., Wang, X., Dai, J., 2008. (R59949) could suppress retinal neovascularization and protect retinal astro-
Nogo-66 promotes the differentiation of neural progenitors into astroglial cytes in an oxygen-induced retinopathy model. J. Mol. Neurosci.
lineage cells through mTOR-STAT3 pathway. PLoS One 3, e1856. Yao, J., Sun, X., Wang, Y., Xu, G., Qian, J., 2007. Math5 promotes retinal ganglion cell
Wang, K., Peng, B., Lin, B., 2014. Fractalkine receptor regulates microglial neuro- expression patterns in retinal progenitor cells. Mol. Vis. 13, 1066e1072.
toxicity in an experimental mouse glaucoma model. Glia 62, 1943e1954. Ye, X., Xu, G., Chang, Q., Fan, J., Sun, Z., Qin, Y., Jiang, A.C., 2010. ERK1/2 signaling
Wang, L., Chen, K., Liu, K., Zhou, Y., Zhang, T., Wang, B., Mi, M., 2015. DHA inhibited pathways involved in VEGF release in diabetic rat retina. Invest. Ophthalmol.
AGEs-induced retinal microglia activation via suppression of the PPARgamma/ Vis. Sci. 51, 5226e5233.
NFkappaB pathway and reduction of signal transducers in the AGEs/RAGE Axis Yoshida, S., Sotozono, C., Ikeda, T., Kinoshita, S., 2001. Interleukin-6 (IL-6) produc-
recruitment into lipid rafts. Neurochem. Res. tion by cytokine-stimulated human Muller cells. Curr. Eye Res. 22, 341e347.
Wang, L.L., Chen, H., Huang, K., Zheng, L., 2012. Elevated histone acetylations in Yu, A.L., Welge-Lussen, U., 2013. Antioxidants reduce TGF-beta2-induced gene ex-
Muller cells contribute to inflammation: a novel inhibitory effect of minocy- pressions in human optic nerve head astrocytes. Acta Ophthalmol. 91, e92e98.
cline. Glia 60, 1896e1905. Yu, D.Y., Cringle, S.J., Balaratnasingam, C., Morgan, W.H., Yu, P.K., Su, E.N., 2013.
Wang, M., Wong, W.T., 2014. Microglia-Muller cell interactions in the retina. Adv. Retinal ganglion cells: energetics, compartmentation, axonal transport, cyto-
Exp. Med. Biol. 801, 333e338. skeletons and vulnerability. Prog. Retin. Eye Res. 36, 217e246.
Watanabe, T., Raff, M.C., 1988. Retinal astrocytes are immigrants from the optic Yu, J., Chen, C., Wang, J., Cheng, Y., Wu, Q., Zhong, Y., Shen, X., 2012. In vitro effect of
nerve. Nature 332, 834e837. adenosine on the mRNA expression of Kir 2.1 and Kir 4.1 channels in rat retinal
Wieghofer, P., Knobeloch, K.P., Prinz, M., 2014. Genetic Targeting of Microglia. Wiley Muller cells at elevated hydrostatic pressure. Exp. Ther. Med. 3, 617e620.
Periodicals, Inc. Zack, D.J., 2000. Neurotrophic rescue of photoreceptors: are Müller cells the me-
Winkler, B.S., Arnold, M.J., Brassell, M.A., Puro, D.G., 2000. Energy metabolism in diators of survival? Neuron 26, 285e286.
human retinal Muller cells. Invest. Ophthalmol. Vis. Sci. 41, 3183e3190. Zamanian, J.L., Xu, L., Foo, L.C., Nouri, N., Zhou, L., Giffard, R.G., Barres, B.A., 2012.
Wohl, S.G., Schmeer, C.W., Isenmann, S., 2012. Neurogenic potential of stem/ Genomic analysis of reactive astrogliosis. J. Neurosci. 32, 6391e6410.
progenitor-like cells in the adult mammalian eye. Prog. Retin. Eye Res. 31, Zeng, H., Ding, M., Chen, X.X., Lu, Q., 2014. Microglial NADPH oxidase activation
213e242. mediates rod cell death in the retinal degeneration in rd mice. Neuroscience
Wolburg, H., 1981. Myelination and remyelination in the regenerating visual system 275, 54e61.
of the goldfish. Exp. Brain Res. 43, 199e206. Zeng, H.Y., Zhu, X.A., Zhang, C., Yang, L.P., Wu, L.M., Tso, M.O., 2005. Identification of
Wolter, J.R., 1956. Special astrocytes in the inner surface of the retina of the human sequential events and factors associated with microglial activation, migration,
eye. Klin. Monatsblatter Augenheilkd. Augenarztl. Fortbild. 129, 224e230. and cytotoxicity in retinal degeneration in rd mice. Invest. Ophthalmol. Vis. Sci.
Wong, V.H., Vingrys, A.J., Bui, B.V., 2011. Glial and neuronal dysfunction in 46, 2992e2999.
streptozotocin-induced diabetic rats. J. Ocul. Biol. Dis. Infor. 4, 42e50. Zhang, S., Li, W., Wang, W., Zhang, S.S., Huang, P., Zhang, C., 2013. Expression and
Wurm, A., Iandiev, I., Hollborn, M., Wiedemann, P., Reichenbach, A., activation of STAT3 in the astrocytes of optic nerve in a rat model of transient
Zimmermann, H., Bringmann, A., Pannicke, T., 2008a. Purinergic receptor acti- intraocular hypertension. PLoS One 8, e55683.
vation inhibits osmotic glial cell swelling in the diabetic rat retina. Exp. Eye Res. Zhang, X., Cheng, M., Chintala, S.K., 2004. Kainic acid-mediated upregulation of
87, 385e393. matrix metalloproteinase-9 promotes retinal degeneration. Invest. Ophthalmol.
Wurm, A., Pannicke, T., Iandiev, I., Francke, M., Hollborn, M., Wiedemann, P., Vis. Sci. 45, 2374e2383.
Reichenbach, A., Osborne, N.N., Bringmann, A., 2011. Purinergic signaling Zhao, J.J., Ouyang, H., Luo, J., Patel, S., Xue, Y., Quach, J., Sfeir, N., Zhang, M., Fu, X.,
involved in Müller cell function in the mammalian retina. Prog. Retin. Eye Res. Ding, S., Chen, S., Zhang, K., 2014a. Induction of retinal progenitors and neurons
30, 324e342. from mammalian Müller glia under defined conditions. J. Biol. Chem. 289,
Wurm, A., Pannicke, T., Wiedemann, P., Reichenbach, A., Bringmann, A., 2008b. Glial 11945e11951.
cell derived glutamate mediates autocrine cell volume regulation in the retina: Zhao, Z., Xu, P., Jie, Z., Zuo, Y., Yu, B., Soong, L., Sun, J., Chen, Y., Cai, J., 2014b.
activation by VEGF. J. Neurochem. 104, 386e399. Gammadelta T cells as a major source of IL-17 production during age-
Xie, B., Jiao, Q., Cheng, Y., Zhong, Y., Shen, X., 2012. Effect of pigment epithelium- dependent RPE degeneration. Invest. Ophthalmol. Vis. Sci. 55, 6580e6589.
derived factor on glutamate uptake in retinal Muller cells under high-glucose Zheng, X.R., Zhang, S.S., Yin, F., Tang, J.L., Yang, Y.J., Wang, X., Zhong, L., 2012.
conditions. Invest. Ophthalmol. Vis. Sci. 53, 1023e1032. Neuroprotection of VEGF-expression neural stem cells in neonatal cerebral
Xu, F., Wei, Y., Lu, Q., Zheng, D., Zhang, F., Gao, E., Wang, N., 2009. Immunohisto- palsy rats. Behav. Brain Res. 230, 108e115.
chemical localization of sortilin and p75(NTR) in normal and ischemic rat Zhou, X., Li, F., Kong, L., Chodosh, J., Cao, W., 2009. Anti-inflammatory effect of
retina. Neurosci. Lett. 454, 81e85. pigment epithelium-derived factor in DBA/2J mice. Mol. Vis. 15, 438e450.
Xu, H., Dawson, R., Forrester, J.V., Liversidge, J., 2007. Identification of novel den- Zieger, M., Ahnelt, P.K., Uhrin, P., 2014. CX3CL1 (fractalkine) protein expression
dritic cell populations in normal mouse retina. Invest. Ophthalmol. Vis. Sci. 48, in normal and degenerating mouse retina: in vivo studies. PLoS One 9,
1701e1710. e106562.

You might also like