Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Sep 1.
Published in final edited form as: J Neuroimmune Pharmacol. 2012 Mar 23;7(3):499–518. doi: 10.1007/s11481-012-9352-5

The Role of Gap Junction Channels During Physiologic and Pathologic Conditions of the Human Central Nervous System

Eliseo A Eugenin 1,2,, Daniel Basilio 3, Juan C Sáez 4,5, Juan A Orellana 6, Cedric S Raine 7,8, Feliksas Bukauskas 9, Michael V L Bennett 10, Joan W Berman 11,12
PMCID: PMC3638201  NIHMSID: NIHMS406617  PMID: 22438035

Abstract

Gap junctions (GJs) are expressed in most cell types of the nervous system, including neuronal stem cells, neurons, astrocytes, oligodendrocytes, cells of the blood brain barrier (endothelial cells and astrocytes) and under inflammatory conditions in microglia/macrophages. GJs connect cells by the docking of two hemichannels, one from each cell with each hemichannel being formed by 6 proteins named connexins (Cx). Unapposed hemichannels (uHC) also can be open on the surface of the cells allowing the release of different intracellular factors to the extracellular space. GJs provide a mechanism of cell-to-cell communication between adjacent cells that enables the direct exchange of intracellular messengers, such as calcium, nucleotides, IP3, and diverse metabolites, as well as electrical signals that ultimately coordinate tissue homeostasis, proliferation, differentiation, metabolism, cell survival and death. Despite their essential functions in physiological conditions, relatively little is known about the role of GJs and uHC in human diseases, especially within the nervous system. The focus of this review is to summarize recent findings related to the role of GJs and uHC in physiologic and pathologic conditions of the central nervous system.

Keywords: Connexin, Hemichannels, NeuroAIDS, HIV, Alzheimer, Disease

General introduction

Gap junction (GJ) channels are formed by two hemichannels, each contributed by one cell, which are hexamers of homologous subunit proteins, termed connexins (Cxs), that connect the cytoplasm of adjacent cells (Bennett et al. 2003; Sáez et al. 2003a) (Fig. 1a). Connexin hemichannels (Cx HCs) can be formed by one (homomeric connexon) or several (heteromeric connexon) types of Cxs, while GJ channels can be formed by either two identical, homotypic, or different, heterotypic, hemichannel subunits. These different subunit combinations enable GJs to differ in their biophysical and permeability properties (Harris 2001, 2007). In addition, it was shown that unapposed hemichannels (uHC), before their cell-to-cell docking to form GJ, also open on the surface of the cell allowing exchange of small factors between the cytoplasm and the extracellular environment. Both GJ and uHC have an internal pore of approximately 12 A°, allowing ions and intracellular messengers up to ~1 kDa in molecular mass to diffuse between connected cells or from the cytoplasm to the extracellular space (Bennett et al. 2003; Sáez et al. 2003a) (Fig. 1b). The diffusion of these second messengers through GJ and uHC results in the coordination of multiple physiologic functions (Sáez et al. 2003a). Here, we will review the pattern of Cx expression in each cell type of the central nervous system (CNS) and the function of GJ and uHC under normal and pathologic conditions.

Fig. 1.

Fig. 1

Schematic diagram showing connexin (Cxs) membrane topology and a plaque of GJ channels. a Model showing membrane topology of Cx E1 and E2, represent the extracellular loops and IL, the intracellular domain. b Model of the GJ plaque between two cells (cell 1-cell 2) and its role in mediating communication by diffusion of second messengers smaller than 1 kDa, such as Ca2+, IP3 and nucleotides

Expression of connexins in different cell types in the CNS

Astrocytes

Astrocytes participate in many brain functions, including CNS differentiation, neuronal excitability, production of neurotrophic factors, control of extracellular synaptic metabolites, syncytial signaling, synaptic plasticity, formation of scar tissue after neuronal loss, immune activation, inflammation, and blood brain barrier (BBB) integrity. In all of these functions, GJ and uHC have a critical role (Rouach et al. 2002c; Kielian and Esen 2004; Sáez et al. 2005; Kielian 2008). Astrocytes form extended networks regions of the brain parenchyma by direct communication through GJs between coupled cells. Reverse transcription-PCR and protein analyses showed that Cx43 and Cx30 are the main Cxs expressed in astrocytes (Dermietzel et al. 1989, 1991; Giaume et al. 1991; Rash et al. 2001a, b; Nagy et al. 2003b; Nakase and Naus 2004). Cx43/Cx30 double-knockout mice only have minimal gap junctional communication (GJC) between astrocytes (Wallraff et al. 2006; Rouach et al. 2008), suggesting that both proteins are the main components of functional astroglial GJ channels. The original studies on Cx43 did not consider the expression of Cx30. However, recent data indicate that Cx30 plays an important role in hippocampus GJ coupling, cellular degeneration and cochlear function (Cohen-Salmon et al. 2007; Chang et al. 2008; Sun et al. 2009; Gosejacob et al. 2011). Thus, further studies are required to clarify the role of Cx43 and Cx30 in physiologic and pathologic conditions.

Synaptic molecules released in response to neuronal activity, including K+, glutamate and other neurotransmitters, are normally taken up by astrocytes and extensively diluted in the astrocytic network through GJ channels. However, in pathologic conditions, GJC is compromised and these molecules can be toxic in the absence of functional GJ channels (Orkand et al. 1966; Rose and Ransom 1997; Wallraff et al. 2006). High neuronal activity enhances astrocyte GJ communication (Marrero and Orkand 1996), and also induces vasodilation of pial arterioles through connexin-based channels (Xu et al. 2008), suggesting a relationship between neuronal activity and blood supply through GJ, in eliminating these toxic molecules by providing better perfusion in areas subjected to high neuronal activity. In addition, GJC enables the coordination of several signaling events, including intercellular Ca2+ waves that control Ca2+-dependent glutamate release, and metabolic and electrical synchronization among astrocytes and neurons and astrocyte and endothelial cells (ECs) to control synaptic plasticity, neuronal survival and/or vascular tone (Guthrie et al. 1999; Paemeleire and Leybaert 2000; Sáez et al. 2003b; Simard et al. 2003; Dale 2008). All these findings indicate that GJs play a key role in the coordination of astrocyte signaling and metabolic events usually altered in many brain-associated diseases.

The opening of Cx uHC allows exchange of diverse molecules between the cytoplasm and extracellular environment, especially in conditions of cellular stress, to mediate autocrine/paracrine signaling. More recently, another gene family similar to the Cx family, encoding a set of three membrane proteins, named pannexins (Panxs 1–3), has been identified in different cell types, including astrocytes (Bruzzone et al. 2003). To date, the absence of ultrastructural evidence for GJ like structures by pannexins in mammalian cells suggests that the main function of pannexin is as pannexin (Panx) uHC. However, a recent study indicated that pannexin 3 expressed in osteblast can work as a uHC in the ER and the plasma membrane and also as a GJ, suggesting that this protein may have multiple functions depending on it cellular localization (Ishikawa et al. 2011). In astrocytes, along with Cx43 uHC, astrocytes also express functional Panx1 uHC (Iglesias et al. 2008, 2009; Iwabuchi and Kawahara 2011). However, the regulation of Panx1 uHC and their role in the function of astrocytes is still under investigation. Astroglial uHC are potential regulators of homeostatic imbalances present in diverse brain diseases or cellular stress conditions, but normally uHC are closed due to their high permeability that can result in cell death when the channels are opened (Abrams et al. 2002; Contreras et al. 2002; Iglesias et al. 2009; Orellana et al. 2011a). Astroglial dysregulation induced by ischemia-like conditions and metabolic inhibition resulted in opening of Cx43 uHC (Contreras et al. 2002). It was also demonstrated that fibroblast growth factor-1 (FGF-1), cytokines, hypoxia, oxidative stress and changes in intracellular/extracellular calcium result in opening of Cx43 uHC, allowing release of ATP, glutamate, nicotinamide adenine dinucleotide (NADH) and prostaglandins to the extracellular space mainly in pathologic conditions (Contreras et al. 2002; Stout et al. 2002; Retamal et al. 2006, 2007; Froger et al. 2009; Sánchez et al. 2009; Schalper et al. 2009; Garre et al. 2010; Orellana et al. 2010; Retamal et al. 2010; Sáez et al. 2010). Astrocytes provide metabolic and framework support to neurons. Therefore, damage associated with the opening of uHC has been proposed to increase neuronal susceptibility to insults (Sánchez et al. 2009a; Orellana et al. 2010; Sáez et al. 2010b; Orellana et al. 2011b). In agreement with this concept, it was demonstrated that opening of astroglial uHC potentiates glutamate-induced neurotoxicity by pro-inflammatory cytokines (Froger et al. 2010) and it has been shown that glutamate can be released through astroglial Cx43 uHC (Ye et al. 2003; Parpura et al. 2004; Malarkey and Parpura 2008; Orellana et al. 2011a), which can lead to neuronal death (Orellana et al. 2011a). Cx43 is the major Cx that forms uHC in astrocytes and Fig. 2 shows an example of activation of Cx43 uHC using whole cell electrophysiological recordings in HeLaCx43-EGFP cells subjected to metabolic inhibition (MI). The cartoon in this figure summarizes some of the conditions that mediate opening/closing events of Cx43 uHCs, such as low pH, high calcium (high Ca2+), DTT, MI and changes to positive voltage (+Vm) (Fig. 2). Under control conditions, Cx43 uHC are open with relatively low probability (data not shown). MI increases the incidence of opening/closing events of Cx43 uHC with a conductance of ~220 pS (Fig. 2). Non-transfected HeLa cells do not show such activity of uHC in response to MI (Contreras et al. 2002).

Fig. 2.

Fig. 2

Cartoon representing some of the agents/conditions that open/close hemichannels and an electrophysiological recording of their activity. The arrows in the diagram show the different conditions that open hemichannels, such as positive voltages (+Vm), metabolic inhibition (MI) and changes in redox potential (dithio-treitol, DTT), and the conditions that close hemichannels, such as low pH and high extracellular calcium (high Ca2+). An example of an electrophysiological recording obtained in Hela cells transfected with Cx43-EGFP in whole cell patch, voltage clamp at +30 mV, is shown under metabolic inhibition conditions as described (Contreras et al. 2002)

Opening of uHC, in addition to mediating the release of several intracellular factors, also participates in the coordination of signaling of other receptors, such as purinergic receptors. However, most of these data were obtained in immune cells. The complex between Panx-1/Cx uHC and ATP receptors mediates enhanced recognition of bacterial molecules, likely by autocrine release of ATP through uHC and subsequent activation of ATP receptors, including P2Y1 and P2Y2 (Kanneganti et al. 2007). This loop, involving uHC, ATP and purinergic receptors, participates in HIV infection and cell-to-cell fusion of immune cells, suggesting an essential role during the pathogenesis of HIV CNS disease (Lemaire et al. 2011; Seror et al. 2011). In agreement, a positive loop involving activation of Panx-1 uHC and ATP receptors has been described in many cell types (Dahl and Locovei 2006; Locovei et al. 2006). In addition, down regulation of Panx-1 uHC prevents the amplification of calcium waves (Locovei et al. 2007), usually spread through GJs and purinergic receptors, suggesting a co-participation of Panx-1 uHC, ATP receptors and GJ channels. Thus, in the CNS there may be interactions between GJs and uHC as well purinergic ATP receptors, which regulate cellular activation and inflammation. However, limited studies on these interactions have been performed in astrocytes.

Under pathologic conditions, the role of astroglial GJs and UHCs is controversial. Contradictory data may be explained by differences in the models used, intensity of the injury and method of analysis. Some studies indicate that inhibition of GJ channels increases neuronal vulnerability to oxidative stress or ischemic insult (Blanc et al. 1998; Siushansian et al. 2001; Nakase et al. 2003; Nakase and Naus 2004; Nakase et al. 2006). In contrast, other studies indicate that functional GJ channels amplify ischemic damage (Lin et al. 1998), and apoptosis during HIV infection (Eugenín and Berman 2007; Eugenín et al. 2011). Our data obtained using HIV-infected astrocytes indicated that only ~5 % of cells are infected with minimal to undetectable viral production. However, bystander killing of neighboring uninfected astrocytes, neurons and endothelial cells occurs by a GJ-dependent mechanism (Eugenín and Berman 2007; Eugenín et al. 2011) (see supplemental Fig. 1). We also demonstrated that HIV infection of a low percentage of astrocytes mediated endothelial apoptosis and BBB disruption by a mechanism that involved dysregulation of several signaling pathways present at the end-feet of the astrocytes that under physiologic conditions regulate blood flow (Eugenín et al. 2011). This BBB disruption was GJ-dependent, because blocking these channels with uncouplers, including 18-α-glycyrrhetinic acid (AGA, 32 μM), or carbenoxolone (CBX; 10 μM), reduced the spread of toxic signals from HIV infected astrocytes to uninfected neighboring astrocytes (Eugenín and Berman 2007), neurons (Supplemental Fig. 1A), and endothelial BBB cells (Eugenín et al. 2011). These data indicate that GJs channels in pathologic conditions, such as HIV infection of the CNS, promote the spread of apoptotic signals among connected cells. Our data indicated that extracellular glutamate does not play a role in bystander apoptosis, because although HIV infection increased glutamate release, blocking GJs increased further its release (Supplemental Fig. 1B). This is the same condition that was neuroprotective in Supplemental Fig. 1A. Thus extracellular glutamate did not contribute directly to HIV-astrocyte-neuronal apoptosis.

In humans it is impossible to examine the dynamic role of GJ channels during the active process of different CNS diseases; therefore, most data addressing these processes were accrued using postmortem human tissue or rodent models. One of the best animal models to examine the role of Cx43 in astrocytes is the Cx43fl/fl;hGFAP-Cre mouse. These animals develop normally and upon activation of expression of Cre in astrocytes, Cx43 expression is abolished in these cells specifically. This Cx43 deletion in astrocytes resulted in increased spreading depression and locomotor activity in these animals (Zhuo et al. 2001; Theis et al. 2003). One of the most striking findings using this mouse is the relationship between glial proliferation and the generation of gliotic areas, as astrogliosis was less pronounced in the Cx43 Cre(+) mice (lacking Cx43). Cx43 is important for control of astroglial proliferation in the penumbra area of the damaged brain (Nakase et al. 2004). In agreement with these reports the inhibition of GJs has been associated with increased levels of cyclins D1, D3, P21 and p27, all of which support proliferation (Sanchez-Alvarez et al. 2006; Tabernero et al. 2006), suggesting a close correlation between GJs, cellular proliferation and hypertrophy.

Although there is a clear metabolic relationship between astrocytes and neurons, the role of GJs in this interaction is not fully understood. Cocultures of astrocyte-neurons increased Cx43 expression and GJC in astrocytes (Rouach et al. 2004a, b). Treatment of astrocyte-neuronal cultures with NMDA or acetylcholine resulted in a prominent reduction of astrocytic GJ channels (Rouach et al. 2002a), suggesting that neuronal activity can have different effects on GJ channels. Inflammatory and stress factors, such as IL-1β (John et al. 1999; Duffy et al. 2000), NO (Bolanos and Medina 1996), ATP (Meme et al. 2004), FGF-2 (Reuss et al. 1998), TGF-β (Reuss et al. 1998), arachidonic acid (Martinez and Saez 1999), endothelins (Giaume et al. 1992), glutamate/kainate (Muller et al. 1996), and acidification (H+ and lactic acid) (Morley et al. 1996, 1997; Dunina-Barkovskaya 1998; Trexler et al. 1999; Duffy et al. 2002; Yamaguchi and Ma 2003; Duffy et al. 2004; Gonzalez-Nieto et al. 2008) reduce Cx43 expression and opening of GJ channels. This indicates that the reduction in GJs and Cx expression perhaps reduces the spread of damage. In contrast, some reports indicated that mild depolarization with K+ (Granda et al. 1998; De Pina-Benabou et al. 2001), glutamate/kainate (Enkvist and McCarthy 1994; Robe et al. 2000) or TGF-β (Robe et al. 2000) enhances GJC. Therefore, the functional state of GJs is altered during inflammation or by neuronal activity depending on the intensity and nature of the damage.

Neurons

Neurons express mainly Cx36, 30.2 and 45 and other Cxs with limited expression and localization (Rouach et al. 2002c; Sohl et al. 2005). Most neurons express Cx36, and Cx30.2 is mainly detected in interneurons in the retina and hippocampus (Kreuzberg et al. 2008; Muller et al. 2010). Our experiments using human fetal neurons obtained from the cortex and hippocampus indicated that these neurons also express Cx43. Additionally, Cx45 has been detected in neurons in the retina and the hippocampus (Schubert et al. 2005; Li et al. 2008; Zlomuzica et al. 2010; Blankenship et al. 2011). Interestingly, the function of these specific Cxs can be compensated in mouse KO models suggesting that their function can be replaced by other Cxs regardless of the differences in sequences and pore permeability (Frank et al. 2010). Despite the fact that extensive research on the Cxs in neurons has been performed, only a few examples of functional GJs were detected in vivo. Some of the best examples were demonstrated in GABAergic neurons in the striatum (Venance et al. 2004), neonatal/developmental cortex neurons (Peinado et al. 1993a, b; Bittman et al. 2002), motor neurons in the spinal cord (Chang et al. 1999; Chang and Balice-Gordon 2000), neurons of the inferior olivary nucleus (Benardo and Foster 1986; Devor and Yarom 2002a, b, c), interneurons in CA3 (Condorelli et al. 1998; Condorelli et al. 2003), visual cortex and dentate regions (Venance et al. 2000; Hormuzdi et al. 2001), in cortex fast-spiking interneurons (Galarreta and Hestrin 1999; Gibson et al. 1999), in the cerebellum (Mann-Metzer and Yarom 1999) and in almost all types of cells in the retina (Vaney 2002; Sohl et al. 2005).

It is known that Cx36 GJs in the hippocampus are required for normal spatial coding and short term spatial memory between interneurons (Allen et al. 2011) and participate in neuronal remodeling by altering differentiation of neuronal stem cells (Hartfield et al. 2011). The proposed functions of these channels in all these systems are to coordinate neuronal firing, spike frequency modulation (Moortgat et al. 2000), fast oscillations (Friedman and Strowbridge 2003; Migliore et al. 2005), neuronal remodeling, and other synchronization properties required under physiologic conditions.

Oligodendrocytes

Oligodendrocytes express Cx29, Cx32, Cx31.3, Cx45 and Cx47 (Kunzelmann et al. 1997; Nagy and Rash 2000; Nagy et al. 2003a; Kleopa et al. 2004; Kamasawa et al. 2005; Rash et al. 2005; Ahn et al. 2008; Orthmann-Murphy et al. 2008; Sargiannidou et al. 2008; Sargiannidou et al. 2009; Maglione et al. 2010; Parenti et al. 2010; Magnotti et al. 2011). Cx29 is present along cell processes, especially in the juxtaparanodal region, but does not colocalize with Cx32 (Altevogt et al. 2002; Menichella et al. 2003; Nagy et al. 2003a; Meier et al. 2004). Cx32 is expressed in paranodal loops, Schmidt-Lanterman incisures and between the outer two layers of internodal myelin, between compact and uncompact myelin (Meier et al. 2004; Kamasawa et al. 2005). Cx47 colocalizes with Cx32 in GJ plaques (Menichella et al. 2003). Cx45 expression in oligodendrocytes is controversial (Dermietzel et al. 1997; Pastor et al. 1998; Maxeiner et al. 2003) and appears to be associated with the cerebral vasculature (Kleopa et al. 2004), most likely with smooth muscle cells close to vessels (Li and Simard 2001). It has also been proposed that oligodendrocytes form GJs with astrocytes (Butt and Ransom 1989; Menichella et al. 2003; Orthmann-Murphy et al. 2008), and perhaps with axons, although this has only been shown between axons and Schwann cells in sciatic nerve using electron microscopy (Dezawa and Nagano 1996). Our data clearly indicated that oligodendrocyte interaction have a unique membrane specialization, such as desmosomes and GJ plaques that likely coordinate heterocellular signaling (see Fig. 3). These GJ structures are particularly dense in EM, due to a more intense accumulation of glial filaments (Soffer and Raine 1980). Desmosomes and GJs are common between glial processes or between astrocytes and oligodendrocytes (Fig. 3). In these cases, GJ have also been proposed to reduce high extracellular concentrations of K+ by providing a mechanism of lateral diffusion and dispersion. Thus, alterations in GJ channels as well as desmosomes may contribute to the pathology observed in several demyelinating diseases.

Fig. 3.

Fig. 3

Electron microscopy in normal mice CNS sections between astrocytes and oligodendrocytes. a An oligodendrocyte in the optical nerve lies near the apical surface (*). Note the gap junctional complex along its lower surface (arrow). x12,500. b Details of the gap junctional plaque in A, between the oligodendrocyte (above), and an astrocyte process (As). The gap junction plaque is unusually long and is flanked by desmosome-like contacts (arrows). Adjacent astrocyte processes are rich in glial filaments and a small desmosome can be seen (below). x62,000. This is an extraordinary demonstration of large gap junctions between astrocytes and oligodendrocytes that support the concept that oligodendricyte and astrocyte communication is active and under pathological conditions, alterations in communication could result in oligodendrocyte damage

In animal models, Cx32 deletion resulted in abnormalities in the sciatic nerve, associated with degeneration (Nelles et al. 1996; Anzini et al. 1997). However, Cx43/Cx47 compensatory mechanisms in the absence of Cx32 have been described (Nagy et al. 2003a). It also remains to be determined whether alterations in Cx32 or other oligodendrocytic Cxs impact different regions of the myelin sheath, such as uncompacted, compacted and/or internodal areas. Cx32 and Cx47 KO mice showed major abnormalities in myelin formation, such as hypomyelination and axonal loss, and oligodendrocyte survival (Odermatt et al. 2003). Similar myelin problems were found in human diseases, such as multiple sclerosis and Pelizaeus-Merzbacher disease (Martini 2000; Kleopa and Scherer 2002; Kleopa et al. 2004) (see section about MS). Experiments in the Cx47 mouse indicated that a missense mutation in the Cx47 gene causes Pelizaeus-Merzbacher disease and results in a pathologic phenotype in this animal model, suggesting that Cx47 significantly participates in the pathogenesis of this disease (Tress et al. 2011). The proposed role for GJs in oligodendrocyte physiology is to provide a shortcut for nutrients and second messengers across the different myelin layers to the axon, and therefore alterations in GJs might contribute to the pathogenesis of myelin/axonal diseases.

Microglia

Microglia express low to undetectable levels of Cx43 and Cx36 under resting conditions (Eugenín et al. 2001; Parenti et al. 2002; Dobrenis et al. 2005; Garg et al. 2005; Lee et al. 2005). Expression of Cx43 and formation of GJ channels can be induced by treatment of rat/mouse microglia with LPS or TNF-α plus IFN-γ (Eugenín et al. 2001), calcium ionophore plus PMA (Martínez et al. 2002), or Staphylococcus aureus-derived peptidoglycan (Garg et al. 2005). Cx36 expression in microglia has been shown using immunohistochemistry and RT-PCR under resting conditions (Parenti et al. 2002). Cxs in microglia may enable these cells to establish direct communication with other cells in the CNS to increase inflammation or to promote the repair of damaged tissue. Interestingly, activation of microglia can down-regulate Cx43 expression and GJ channels among astrocytes when both cell types are in co-culture suggesting a cell-specific control of Cx43 expression (Rouach et al. 2002b; Faustmann et al. 2003; Meme et al. 2006). Findings in dendritic cells indicated that GJC could be used for the sharing of antigenic peptides (Neijssen et al. 2005; Matsue et al. 2006; Corvalan et al. 2007; Handel et al. 2007; Mendoza-Naranjo et al. 2007; Pang et al. 2009), suggesting the possibility that GJs between microglia also coordinate the CNS immune response. Our data using cultures of rat microglia showed that under control conditions, the release of TNF-α, IL-1β and IL-6 was minimal (Supplemental Fig. 2a and b, white bars). Treatment of microglia cultures with LPS plus IFN-γ for 1 to 9 h, a condition that increased GJC, increased secretion of TNF-α and IL-1β (Supplemental Fig. 2A), but not IL-6 (Supplemental Fig. 2B, cross line bands). The secretion of these cytokines was partially blocked by a GJ blocker, AGA (Supplemental Fig. 2A, cross line bars), suggesting that functional GJs are important in microglia cytokine secretion. Thus, GJs in microglia are induced by specific inflammatory factors and the proposed function of these channels is to help to coordinate the microglial mediated inflammation.

Blood–brain barrier (BBB)

The BBB is composed of dynamic vessels that are capable of responding to rapid changes in the brain or in the blood stream (Gloor et al. 2001; Ballabh et al. 2004). The BBB is composed of endothelial cells (EC) in close contact with astrocytic end-feet across a basal lamina, perivascular macrophages and pericytes. These cellular interactions function in combination with other systems specialized for the transport of metabolites required for brain function, as well as tight junction proteins (TJP), which seal the intercellular gaps between EC-EC and EC-astrocytes, to establish impermeability to most macromolecules and blood cells (Ballabh et al. 2004). EC isolated from blood vessels of different organs express Cxs 37, 40 and 43 (Haefliger et al. 2004; Burnier et al. 2009). Cx45 has been detected in cerebral blood vessels but its expression is associated with smooth muscle cells (Li and Simard 2001). However, deletion of Cx45 results in defective vascular development, suggesting a critical role of Cx45 in the development of brain vasculature (Kruger et al. 2000). In the systemic vasculature, deletion of Cx40 results in hypertension (Firouzi et al. 2006a; Firouzi et al. 2006b; Hauer et al. 2006). Endothelial-specific deletion of Cx43 results in hypotension and bradycardia (Liao et al. 2001), suggesting that expression of these Cxs participates in the regulation of blood pressure. In agreement, the control of blood flow by a GJ-dependent mechanism has been described in peripheral arterioles (de Wit et al. 2000; de Wit et al. 2003). Our data demonstrated that HIV infection of just a few astrocytes in a BBB tissue culture model amplifies endothelial apoptosis by dysregulation of astrocyte end-feet signaling in a GJ-dependent manner (Eugenín et al. 2011). However, little is known about the expression of Cxs and their role in the physiology and pathogenesis of the BBB under normal and pathologic conditions.

Neuronal stem cells

Neuronal precursors express Cx26, Cx30, Cx33, Cx36, Cx37, Cx40, Cx43 and Cx47 depending on the area of the brain from which these cells were isolated, the differentiation stage and the cell culture conditions (Rozental et al. 1998; Duval et al. 2002; Maxeiner et al. 2003; Trosko and Chang 2003; Elias et al. 2007; Elias and Kriegstein 2008; Wen et al. 2008; Cina et al. 2009). Expression of Cxs 30, 36, 37 and 43, but not Cxs 26, 32 or 47, has been reported in NT2/D1 progenitor cells. These cells were obtained from a teratocarcinoma progenitor line that can be induced to differentiate into hNT neurons and NT-G non-neuronal cells (Bani-Yaghoub et al. 1999). hNT/NT-G cells differentiated with retinoic acid, express Cxs 36, 37 and 47. However, only undifferentiated cells are capable of dye transfer to other cells (Boucher and Bennett 2003). Functional GJ are required to maintain cortical neural progenitor cells in a proliferative state (Cheng et al. 2004) as well as for their radial migration in the neocortex (Elias et al. 2007). It has also been described that the carboxyl-terminal domain of Cx43 regulates neuronal differentiation (Cina et al. 2009; Santiago et al. 2010). In addition, Panx-2 uHC are expressed in postnatal hippocampus neuronal progenitors and also modulate the differentiation of neurons (Swayne et al. 2010). Thus, it is clear that GJs and uHC are critical for multiple functions involving neuronal differentiation and migration. However, the question whether GJ and uHC are involved in diseases involved in migration and differentiation is still under investigation.

Gap junctions and connexin expression in human diseases

Cx32 mutations linked to the X-linked hereditary motor and sensory neuropathy Charcot-Marie-Tooth (CMTX)

More than 305 different Cx32 mutations have been associated with CMTX (see for details; http://www.molgen.ua.ac.be/CMTMutations/), but only a few results in evident pathologic conditions (Kleopa et al. 2002). One explanation is that not all Cx32 mutations destabilize the myelin sheath and/or compromise communication in a significant manner. An alternative or additional explanation is that other Cxs expressed in oligodendrocytes or Schwann cells can be compensatory. Little is known about the functions of Cx32 in the PNS and CNS, due to difficulties in evaluating the function(s) of these channels in vivo, or during the pathogenesis of CMTX disease (Bergoffen et al. 1993; Ionasescu et al. 1996; Ionasescu 1998). Cx32-deficient mice did not show any evident oligodendrocyte-dependent myelination (Scherer et al. 1998). Neurons isolated from Cx32-deficient mice did not display impaired excitability and synaptic formation/stability (Sutor et al. 2000). However, Cx32/Cx47 KO mice exhibited motor problems and a myelin defect (vacuolization) when compared to single Cx47 KO or Cx32 KO mice (Odermatt et al. 2003). This suggests that deficiencies in both Cxs are required to result in a pathologic phenotype. Cx32 is also important in mediating GJ communication between oligodendrocytes and astrocytes or neurons, and these interactions may be altered in CMTX (Angaut and Sotelo 1973; Sotelo and Angaut 1973; Kettenmann et al. 1983; Butt and Ransom 1989; Menichella et al. 2003; Orthmann-Murphy et al. 2008).

Cxs and multiple sclerosis

Multiple sclerosis (MS) is a primary demyelinating disease in humans. As described above, Cx32 plays a key role in maintaining the structure and function of the myelin sheath. Deletions or mutations in this gene in rodents and humans can result in myelin abnormalities, similar to those characterized in MS. In the EAE, an animal model of MS, Cx43 expression was down-regulated and Cx30 upregulated in sites in close association with inflammation (Brand-Schieber et al. 2005). Our data using human tissue sections indicated that in control non inflammatory conditions, colocalization between myelin basic protein (MBP) and Cx32 was observed (Fig. 4). Tissue sections obtained from individuals with chronic active and silent MS lesions demonstrated that Cx32 is down regulated in oligodendrocytes in chronic active lesions at sites of inflammation (Fig. 4). In contrast, upregulation of Cx32 was observed in silent lesions from individuals with chronic silent MS, characterized by the absence of inflammation and re-myelination (Fig. 4). In active MS lesions, Cx32 staining was reduced (Fig. 4). Cx43, which is mainly expressed in astrocytes, was upregulated in areas around the lesion, suggesting that scar tissue has high expression of Cx43, most likely due to astrocyte proliferation (Fig. 4). However, astrocytes in the center of the lesions were negative for Cx43 in active chronic lesions (Fig. 4). Thus, our data indicated that Cx32 expression is associated with the degree of damage and remyelination, while Cx43 is associated with astrocyte activation/proliferation as well as inflammation. Future studies will be essential to understand the role of other Cxs expressed in oligodendrocytes, Schwann cells, astrocytes and neurons during the pathogenesis of demyelinating diseases.

Fig. 4.

Fig. 4

Distribution of Cx43 and Cx32 in human spinal cord sections obtained from normal individuals and individuals with MS. Confocal microscopy of Cx43 (FITC, green) and MBP (Cy3, red) staining, colocalization of both proteins is represented as orange staining in the last panel. A, B and C, represents staining of human tissue sections for GFAP (FITC, green) and Cx43 (Cy3, red), a small insert shows the Cx43 staining alone, from normal (a), MS with chronic active lesions (b) and chronic silent lesions (c). d, e and f, represents staining for Cx32 (FITC, Green) and MBP (Cy3, red) in human sections obtained from spinal cords from individuals with normal tissue (d), MS with chronic active lesions (e) and silent lesions (f). The small inserts in each picture show the Cx32 staining alone. These tissue sections were already characterized for the kind of MS lesions and the damage in the lesion area (Calderon et al. 2006). Note that in MS tissue it is possible observe oligodendrocyte atrophy and disorganization of the brain parenchyma. Bar: 70 μm

HIV and NeuroAIDS

HIV infection of the CNS induces different degrees of cognitive and motor impairment, collectively termed HIV associated neurologic disorders (HAND). Approximately 60 % of individuals infected with HIV have HAND even in the current anti-retroviral era (González-Scarano and Martín-García 2005; Boisse et al. 2008; Kraft-Terry et al. 2009). However, the mechanisms that mediate HAND in most HIV infected individuals are not completely understood. The neuropathology of HIV-infection includes microglial nodules, multinucleated giant cells and astrogliosis, as well as neuronal injury and loss (Albright et al. 1999; Kaul et al. 2001; González-Scarano and Martín-García 2005). Macrophages and microglia support high viral replication within the CNS (González-Scarano and Martín-García 2005). HIV-infected astrocytes have also been detected in vivo and in vitro (Tontsch and Bauer 1991; Tornatore et al. 1991; Conant et al. 1994; Tornatore et al. 1994a; Tornatore et al. 1994b; Bagasra et al. 1996; Ohagen et al. 1999; Gorry et al. 2003; Churchill et al. 2006; Eugenín and Berman 2007; Churchill et al. 2009; Eugenín et al. 2011) and was characterized by low to undetectable viral replication and a low numbers of cells that are infected (Tornatore et al. 1994a; Ohagen et al. 1999; Schweighardt and Atwood 2001; Eugenín and Berman 2007; Eugenín et al. 2011). In general, inflammation and infectious agents reduce Cx43 expression and GJS (see details below). However, HIV is different because despite its inflammatory nature, Cx43 expression and GJ channels are maintained in astrocytes (Eugenín and Berman 2007). Functional GJ channels promote the spread of toxic signals from a few HIV-infected astrocytes to uninfected astrocytes, neurons and endothelial cells resulting in the spread of toxic mediators and dysregulation of glutamate and CCL2 secretion (Eugenín and Berman 2007; Eugenín et al. 2011). Interestingly, the few HIV infected astrocytes are protected from apoptosis by a viral-dependent mechanism, resulting in a viral reservoir within the CNS to perpetuate the presence of the virus.

Despite extensive evidence of pathological changes in the CNS of HIV-infected people, the role of GJs has been minimally examined. An accepted mechanism by which cognitive impairment and dementia occurs involves the transmigration of HIV-infected monocytes across the BBB into the CNS parenchyma and the accumulation of macrophages and microglia within the CNS in correlation with several inflammatory factors (Persidsky et al. 1997; Weiss et al. 1998; Eugenín et al. 2006; Roberts et al. 2010). Normally, macrophages/microglia express low to undetectable levels of Cxs, however, we and others demonstrated that macrophages/microglia express higher levels of Cxs under inflammatory conditions (Eugenín et al. 2001; Martínez et al. 2002; Parenti et al. 2002; Dobrenis et al. 2005; Garg et al. 2005). In Supplemental Fig. 3, our data demonstrate that during HIV associated dementia (HAD) or HIV encephalitis (HIVE), alterations in GJs occur in HIV-infected leukocytes, microglia and astrocytes. HIV-infected leukocytes after HIV infection begin to express high levels of Cx43 (Supplemental Fig. 3A) that may be necessary for transmigration across the BBB, as we previously described in uninfected peripheral blood mononuclear cells (Eugenín et al. 2003). In addition, confocal analyses of HIVE human tissue demonstrated that microglia/macrophages express Cx43 (Supplemental Fig. 3B) and Cx36 (data not shown) in close contact with neuronal cell bodies (Supplemental Fig. 3B). An example is shown in Supplemental Fig. 3, showing Cx43 expression between a neuron and microglia/macrophage (CD68 positive cells) in HIVE tissue. We did not detect either Cx in microglia/macrophages in brain tissue sections obtained from normal or non-encephalitic HIV positive individuals (data not shown). In addition, Cx43 expressed in astrocytes and Cx36 in neurons were down-regulated in HIVE tissue sections as compared to cells in tissue sections obtained from uninfected brains. In addition, our findings indicated that uHC are opened in response to HIV infection in astrocytes, suggesting their participation in the pathogenesis of HIV CNS disease. These findings that Cxs participate in the pathogenesis of NeuroAIDS open a new avenue of investigation to study the mechanisms by which HIV “hijacks” this communication system, GJs and perhaps uHCs, to spread toxicity, inflammation, and increase leukocyte transmigration into the CNS.

Viral and bacterial infections

In general, both viral and bacterial infections reduce GJ channels and Cx expression. For example, swine Flu virus down-regulates endothelial Cx43 expression by an ERK and increased degradation dependent mechanism (Hsiao et al. 2010). Borna virus also down regulates Cx36 in the CNS in specific brain cell types (Koster-Patzlaff et al. 2007, 2008, 2009). Influenza viral infection during pregnancy alters development of the brain of the fetus, suggesting that viruses can impact neuronal development by affecting GJs (Fatemi et al. 2008) required for development and function of the CNS. Studies in Vero cells demonstrated that infection with Herpes Simplex Virus-2 (HSV-2), down-regulated GJ channels and Cxs expression (Fischer et al. 2001; Musee et al. 2002; Knabb et al. 2007). It was reported that an increase in tyrosine phosphorylation by HSV-2 and Rous sarcoma virus leads to an inhibition of GJ channels and Cx43 expression (Crow et al. 1990; Filson et al. 1990; Crow et al. 1992). Bovine papillomavirus type 4 E8, when bound to ductin, causes loss of GJ channels between primary fibroblasts (Faccini et al. 1996; Ashrafi et al. 2000). In contrast, our preliminary data show that HIV infection of astrocytes is different because it maintains/increases Cx43, Cx30 and GJ channels allowing toxic signals generated in a few infected cells to spread through GJs to uninfected cells (Eugenín and Berman 2007; Eugenín et al. 2011). In addition, GJ channels increase invasion and dissemination of Shigella in epithelial cells (Tran Van Nhieu et al. 2003), suggesting the possibility that specific infectious agents, such as Shigella and HIV, can use GJ channels to sensitize uninfected cells and spread infection/toxicity to healthy cells. Future experiments are necessary to identify the signals generated by infected cells that cross through GJs to sensitize uninfected cells and enable them to become targets of these infectious agents.

Alternatively, GJs benefit the host immune system by mediating a phenomenon termed cross-antigen presentation. This enables coupled cells to share viral peptides (antigens) and trigger a response in CTL cells, even when some cells were never directly exposed to the pathogen (Neijssen et al. 2005). GJ-mediated immune coupling suggests the possibility that GJs expressed by monocytes/macrophages in inflammatory conditions cross-present antigens to lymphocytes and other inflammatory cells to maintain an immune memory in cells never exposed directly to specific antigen (Neijssen et al. 2005). In agreement, Cx43 is recruited to the immunologic synapse during T cell priming, suggesting that GJ and uHCs also participate in antigen presentation (Mendoza-Naranjo et al. 2011).

Alzheimer’s disease (AD)

Aβ peptide is required for CNS function. However, under certain conditions, this peptide aggregates and produces toxic effects (Palop and Mucke 2010; Parihar and Brewer 2010). Up-regulation of Cx43 has been detected in cortical astrocytes in the brain containing Aβ peptide plaques as compared to normal brains (Nagy et al. 1996; Mei et al. 2010). Experiments using rat astrocytes demonstrated that β/A4 amyloid resulted in an increase in the amplitude, velocity and travel distance of evoked calcium waves, by an ATP- and GJC-dependent mechanism (Haughey and Mattson 2003). Recently, it was shown that Aβ peptide induces the release of ATP and glutamate through glial uHC, which leads to further neuronal death by activation of Panx-1 uHC in neurons (Orellana et al. 2011b). Additionally, expression of β/A4 amyloid in PC12 cells increased Cx43 expression and dye coupling (Lynn et al. 1995), suggesting that Aβ peptide increases GJC in CNS cells by an unknown mechanism. However, in all of these studies it is unclear whether normal regulation of Cx is a product of a physiologic or pathogenic effect of Aβ or due to the damage observed in the end stages of the disease. Thus, further examination of the role of Cxs and GJ during the pathogenesis of AD is required.

Parkinson’s disease (PD)

PD is a neurodegenerative disease characterized by loss of dopaminergic neurons, especially in the substantia nigra-striatum, that results in progressive tremor, and muscle and gait abnormalities (Lim et al. 2002; Lotharius and Brundin 2002). The mouse MTPT model of PD showed increased levels of Cx43 in the stratium (Rufer et al. 1996), but this may be related to gliosis and not be specific to PD. The dysfunction of astrocytic Cx43 induced by rotenone to trigger a PD phenotype, as a model of PD, can be reversed by opening mitochondrial ATP-sensitive potassium channels, suggesting the involvement of these channels in the mitochondria regulated Cx43 expression (Kawasaki et al. 2009; Zhang et al. 2010).

Interestingly, treatment of the Cx36 KO mouse with harmaline, a beta-carboline derivative, induced tremors by altering rhythmogenesis in a similar way to PD, but did not show any differences when compared to wild type mice, suggesting that Cx36 may not play a role in the initial pathogenesis of the disease (Long et al. 2002a). However, a compensatory mechanism should not be ruled out, as it has already been described in the Cx36 KO mice in coordinating rhythmic activity between the neurons of the inferior olive and suprachiasmatic nucleus (Long et al. 2002b, 2005), which is believed to be the region that generates the tremors (Elble 1996). In conclusion, there are only a few studies that characterize the expression, function and localization of Cxs in PD brains or animal models and the participation of these channels in the pathogenesis of PD is unclear.

Huntington disease

Huntington disease is an autosomal dominant neurodegenerative disorder characterized by motor dysfunctions, cognitive impairment and personality changes that is due to mutations in the protein huntingtin (htt). Studies using human tissue sections obtained from individuals with Huntington disease showed a similar distribution of Cx32 and Cx26 as compared to normal brain tissue sections, but Cx43 immunoreactivity was increased in the caudate nucleus, especially in regions richer in GFAP staining (Vis et al. 1998), perhaps due to gliosis. Studies in retina using a transgenic R6/2 mice expressing the mutant form of htt demonstrated decreased expression of Cx45 while Cx36 and Cx43 were not significantly altered (Petrasch-Parwez et al. 2004). Further studies are required to dissect whether these changes in Cx expression participate in the pathogenesis of the disease or are a consequence of the cellular damage characteristic of Huntington disease.

Epilepsy

Epileptic seizures may be associated with abnormal stimulation of certain brain areas causing abnormal depolarization, which expands by spreading to neighboring areas (Ure and Perassolo 2000; Ure et al. 2006). This process can be mediated by glutamate (Meldrum 1994; Moldrich et al. 2003), auto antibodies that activate AMPA receptors (GluR3) (Levite et al. 1999; Levite and Hermelin 1999; Palmer et al. 1999), and aberrant activation or deactivation of NMDA receptors, GABA receptors, potassium channels, sodium channels and GJs (Naus et al. 1991; Carlen et al. 2000; Samoilova et al. 2003; Brooks-Kayal et al. 2009; Kang and Macdonald 2009; Planells-Cases and Jentsch 2009; Galanopoulou 2010). Astrocytic GJ channels increase in epileptic tissue of cats as demonstrated by morphological studies (Vaquero et al. 1978) and an increase in Cx43 mRNA was found in human epileptic tissue (Naus et al. 1991), suggesting a potential role of Cx43 in the pathogenesis of this disease. In addition, increased GJ channels have been shown between astrocytes isolated from human epileptic tissue as assayed by the enhanced propagation of a calcium wave in response to glutamate (Lee et al. 1995). Based on these data, it has been proposed that Cxs are involved in expanding the spread of an epileptic wave during seizures (Perez Velazquez and Carlen 2000; Perez Velazquez et al. 2001). This hypothesis is supported by studies showing that GJ blockers reduced seizures in different models of epilepsy in vitro and in vivo (Sohl et al. 2000; Jahromi et al. 2002; Samoilova et al. 2003; Gareri et al. 2004; Gajda et al. 2005; Gareri et al. 2005; Bostanci and Bagirici 2006; Nilsen et al. 2006; Meldrum and Rogawski 2007; Jacobson et al. 2010). Studies using the Cx36 KO mouse showed that loss of GJ channels reduced kainate- or 4-aminopyridine-induced seizures (Hormuzdi et al. 2001; Maier et al. 2002; Christie et al. 2005), suggesting a key role of GJ channels in the development and sustaining of epileptic seizures. In human tissue, upregulation of Cx43 and Cx32 has been detected in different types of epilepsy (Elisevich et al. 1997; Aronica et al. 2001; Li et al. 2001; Fonseca et al. 2002; Samoilova et al. 2008; Yao et al. 2009). Additionally, Cx30, which is normally expressed in astrocytes, was detected in neurons after kainate-induced seizures (Condorelli et al. 2002). All these data suggest that de novo synthesis of Cxs causes changes in permeability and distribution of Cxs during epileptogenesis. While these data in humans are consistent, results in animal models are more variable and difficult to interpret. This may be due to the use of different experimental approaches to generate the epileptic phenotype and the methods to examine the disease that may or may not be good models for epilepsy in humans resulting in different patterns of Cxs expression and functions of GJ channels. Thus, design of better animal models and studies in humans are required to compare both systems.

CNS tumors

Gliomas are the most common tumor in the brain (Behin et al. 2003), characterized by high intracranial pressure, brain edema and vessel occlusion (Behin et al. 2003). Cx expression and GJ channels are low in tumors, primary glioma cells or glioma cell lines (Huang et al. 1999; Zhang et al. 1999; Soroceanu et al. 2001; Lin et al. 2002; Pu et al. 2004; Oliveira et al. 2005; Kubota et al. 2006; Bates et al. 2007; Lai et al. 2007). Tumor cells can directly couple with normal cells through GJs (Zhang et al. 1999). This communication results in a phenotypic transformation of astrocytes that may contribute to the susceptibility of surrounding tissue to glioma invasion. There are reports indicating a beneficial role of Cxs in glioma treatment. Cx43 expression by itself is considered to be a tumor suppressor gene independent of formation of functional GJs (Huang et al. 1998; Yamasaki et al. 1999; Moorby and Patel 2001; Omori et al. 2001; Zhang et al. 2003a, b, c; Del Monte and Statuto 2004). In addition, transfection of Cx43 and herpes simplex virus thymidine kinase (HSVtk) resulted in bystander cell death in a Cx43 dependent manner upon treatment with ganciclovir (GCV), a nucleoside analogue (Shinoura et al. 1996; Cirenei et al. 1998; Grignet-Debrus et al. 2000; Marconi et al. 2000; Namba et al. 2001; Asklund et al. 2003; Huang et al. 2010). GVC is phosphorylated by HSVtk into a monophosphate form and subsequently to GCV-triphosphate by endogenous kinases, and is incorporated into the DNA of the target cell, leading to strand breaks and cell death. Interestingly, neighboring cells coupled by GJs also die, although these cells do not express the enzyme. This phenomenon is believed to be a bystander effect mediated by the transfer of toxic GVC-metabolites through GJs from the cell infected with HSVtk to uninfected neighboring cells (Dilber et al. 1997). Cx43 transfection into tumor cells results in functional coupling and in enhancement of the bystander effect in vivo (Mesnil et al. 1996; Cirenei et al. 1998; Andrade-Rozental et al. 2000) and in vitro (Mesnil et al. 1996; Cirenei et al. 1998; Andrade-Rozental et al. 2000; Marconi et al. 2000; Musee et al. 2002; Asklund et al. 2003; Huang et al. 2010). Currently, several groups continue to examine the potential role of this bystander effect in the treatment of different types of tumors. The discovery that glioma cells can express functional uHC also increases the possibility that during pathologic conditions, these channels contribute to the regulation of proliferation or in protecting tumor cells from damage induced by GVC or chemotherapy by pumping out toxic metabolites in similar ways to ABC transporters (Andrade-Rozental et al. 2000). In addition, future studies investigating the generation of other intracellular toxic metabolites, such as cytochrome C, which has been shown to cause bystander killing in the retina, could be beneficial.

Conclusions

This review summarizes much of the current data on the role of Cxs in different cell types and their involvement in human central nervous system diseases. A more complete understanding of the role of Cxs in different human diseases require a more detailed study of the function of connexin-and pannexin-based cell-cell channels, GJs and uHC, under normal and pathologic conditions. A critical point is to examine, in addition to the role of GJs and uHC in physiologic conditions, the role of these channels during the pathogenesis of human disease. Recent work indicated that GJs and uHC are key channels involved in resolving, but also in spreading, disease. Our study with HIV indicates that some pathogens use these communication systems to their advantage. Thus, a more complete understanding of these communication systems should provide essential information for the development of novel therapeutic approaches

Supplementary Material

1-3

Acknowledgments

We apologize to the authors and groups whose work we did not cite due to limitations on the numbers of references. This work was supported by the National Institute of Mental Health grants (MH076679 and MH096625 to E.A.E. and MH075679 and MH083497 to J.W.B.) and by the National Institute of Neurological Disorders and Stroke (NS072238 to FFB, and NS55363 to M.V.L.B, who is the Sylvia and Robert S. Olnick Professor of Neuroscience). We thank the National Multiple Sclerosis Society Grant RG-1001-K-11 and CSR was the Wollowick Family Foundation Professor for Multiple Sclerosis Research (to Dr. Cedric Raine). We thank the NIH Centers for AIDS Research Grant (CFAR) AI-051519, Anillo ATC-71 (JCS) and Centro interdisciplinario de Neurociencias P09-022-F (to JCS) and a CFAR pilot project at the Albert Einstein College of Medicine.

Footnotes

Electronic supplementary material The online version of this article (doi:10.1007/s11481-012-9352-5) contains supplementary material, which is available to authorized users.

Conflict of interest None

Contributor Information

Eliseo A. Eugenin, Department of Pathology, F727, Albert Einstein College of Medicine, 1300 Morris Park Ave., Bronx, NY 10461, USA eliseo.eugenin@einstein.yu.edu Public Health Research Institute (PHRI) and Department of Microbiology and Molecular Genetics, UMDN, Newark, USA.

Daniel Basilio, Department of Neuroscience, Albert Einstein College of Medicine, Bronx, NY, USA.

Juan C. Sáez, Centro Interdisciplinario de Neurociencias de Valparaíso, Instituto Milenio, Valparaíso, Chile Departmento de Fisiología, Pontificia Universidad Católica de Chile, Santiago, Chile.

Juan A. Orellana, Departamento de Neurología, Facultad de Medicina, Pontificia Universidad Católica de Chile, Santiago, Chile

Cedric S. Raine, Department of Pathology, F727, Albert Einstein College of Medicine, 1300 Morris Park Ave., Bronx, NY 10461, USA Department of Neurology, Albert Einstein College of Medicine, Bronx, NY, USA.

Feliksas Bukauskas, Department of Neuroscience, Albert Einstein College of Medicine, Bronx, NY, USA.

Michael V. L. Bennett, Department of Neuroscience, Albert Einstein College of Medicine, Bronx, NY, USA

Joan W. Berman, Department of Pathology, F727, Albert Einstein College of Medicine, 1300 Morris Park Ave., Bronx, NY 10461, USA Department of Microbiology and Immunology, Albert Einstein College of Medicine, Bronx, NY, USA.

References

  1. Abrams CK, Bennett MV, Verselis VK, Bargiello TA. Voltage opens unopposed gap junction hemichannels formed by a connexin 32 mutant associated with X-linked Charcot-Marie-Tooth disease. Proc Natl Acad Sci U S A. 2002;99:3980–3984. doi: 10.1073/pnas.261713499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Ahn M, Lee J, Gustafsson A, Enriquez A, Lancaster E, Sul JY, Haydon PG, Paul DL, Huang Y, Abrams CK, Scherer SS. Cx29 and Cx32, two connexins expressed by myelinating glia, do not interact and are functionally distinct. J Neurosci Res. 2008;86:992–1006. doi: 10.1002/jnr.21561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Albright AV, Shieh JT, Itoh T, Lee B, Pleasure D, O’Connor MJ, Doms RW, Gonzalez-Scarano F. Microglia express CCR5, CXCR4, and CCR3, but of these, CCR5 is the principal coreceptor for human immunodeficiency virus type 1 dementia isolates. J Virol. 1999;73:205–213. doi: 10.1128/jvi.73.1.205-213.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Allen K, Fuchs EC, Jaschonek H, Bannerman DM, Monyer H. Gap junctions between interneurons are required for normal spatial coding in the hippocampus and short-term spatial memory. J Neurosci Off J Soc Neurosci. 2011;31:6542–6552. doi: 10.1523/JNEUROSCI.6512-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Altevogt BM, Kleopa KA, Postma FR, Scherer SS, Paul DL. Connexin29 is uniquely distributed within myelinating glial cells of the central and peripheral nervous systems. J Neurosci. 2002;22:6458–6470. doi: 10.1523/JNEUROSCI.22-15-06458.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Andrade-Rozental AF, Rozental R, Hopperstad MG, Wu JK, Vrionis FD, Spray DC. Gap junctions: the “kiss of death” and the “kiss of life”. Brain Res Brain Res Rev. 2000;32:308–315. doi: 10.1016/s0165-0173(99)00099-5. [DOI] [PubMed] [Google Scholar]
  7. Angaut P, Sotelo C. The fine structure of the cerebellar central nuclei in the cat. II. Synaptic organization. Exp Brain Res. 1973;16:431–454. doi: 10.1007/BF00233433. [DOI] [PubMed] [Google Scholar]
  8. Anzini P, Neuberg DH, Schachner M, Nelles E, Willecke K, Zielasek J, Toyka KV, Suter U, Martini R. Structural abnormalities and deficient maintenance of peripheral nerve myelin in mice lacking the gap junction protein connexin 32. J Neurosci. 1997;17:4545–4551. doi: 10.1523/JNEUROSCI.17-12-04545.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Aronica E, Gorter JA, Jansen GH, Leenstra S, Yankaya B, Troost D. Expression of connexin 43 and connexin 32 gap-junction proteins in epilepsy-associated brain tumors and in the perilesional epileptic cortex. Acta Neuropathol. 2001;101:449–459. doi: 10.1007/s004010000305. [DOI] [PubMed] [Google Scholar]
  10. Ashrafi GH, Pitts JD, Faccini A, McLean P, O’Brien V, Finbow ME, Campo S. Binding of bovine papillomavirus type 4 E8 to ductin (16 K proteolipid), down-regulation of gap junction intercellular communication and full cell transformation are independent events. J Gen Virol. 2000;81:689–694. doi: 10.1099/0022-1317-81-3-689. [DOI] [PubMed] [Google Scholar]
  11. Asklund T, Appelskog IB, Ammerpohl O, Langmoen IA, Dilber MS, Aints A, Ekstrom TJ, Almqvist PM. Gap junction-mediated bystander effect in primary cultures of human malignant gliomas with recombinant expression of the HSVtk gene. Exp Cell Res. 2003;284:185–195. doi: 10.1016/s0014-4827(02)00052-6. [DOI] [PubMed] [Google Scholar]
  12. Bagasra O, Lavi E, Bobroski L, Khalili K, Pestaner JP, Tawadros R, Pomerantz RJ. Cellular reservoirs of HIV-1 in the central nervous system of infected individuals: identification by the combination of in situ polymerase chain reaction and immunohistochemistry. AIDS. 1996;10:573–585. doi: 10.1097/00002030-199606000-00002. [DOI] [PubMed] [Google Scholar]
  13. Ballabh P, Braun A, Nedergaard M. The blood–brain barrier: an overview: structure, regulation, and clinical implications. Neurobiol Dis. 2004;16:1–13. doi: 10.1016/j.nbd.2003.12.016. [DOI] [PubMed] [Google Scholar]
  14. Bani-Yaghoub M, Felker JM, Naus CC. Human NT2/D1 cells differentiate into functional astrocytes. Neuroreport. 1999;10:3843–3846. doi: 10.1097/00001756-199912160-00022. [DOI] [PubMed] [Google Scholar]
  15. Bates DC, Sin WC, Aftab Q, Naus CC. Connexin43 enhances glioma invasion by a mechanism involving the carboxy terminus. Glia. 2007;55:1554–1564. doi: 10.1002/glia.20569. [DOI] [PubMed] [Google Scholar]
  16. Behin A, Hoang-Xuan K, Carpentier AF, Delattre JY. Primary brain tumours in adults. Lancet. 2003;361:323–331. doi: 10.1016/S0140-6736(03)12328-8. [DOI] [PubMed] [Google Scholar]
  17. Benardo LS, Foster RE. Oscillatory behavior in inferior olive neurons: mechanism, modulation, cell aggregates. Brain Res Bull. 1986;17:773–784. doi: 10.1016/0361-9230(86)90089-4. [DOI] [PubMed] [Google Scholar]
  18. Bennett MV, Contreras JE, Bukauskas FF, Saez JC. New roles for astrocytes: gap junction hemichannels have something to communicate. Trends Neurosci. 2003;26:610–617. doi: 10.1016/j.tins.2003.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Bergoffen J, Scherer SS, Wang S, Scott MO, Bone LJ, Paul DL, Chen K, Lensch MW, Chance PF, Fischbeck KH. Connexin mutations in X-linked Charcot-Marie-Tooth disease. Science. 1993;262:2039–2042. doi: 10.1126/science.8266101. [DOI] [PubMed] [Google Scholar]
  20. Bittman K, Becker DL, Cicirata F, Parnavelas JG. Connexin expression in homotypic and heterotypic cell coupling in the developing cerebral cortex. J Comp Neurol. 2002;443:201–212. doi: 10.1002/cne.2121. [DOI] [PubMed] [Google Scholar]
  21. Blanc EM, Bruce-Keller AJ, Mattson MP. Astrocytic gap junctional communication decreases neuronal vulnerability to oxidative stress-induced disruption of Ca2+ homeostasis and cell death. J Neurochem. 1998;70:958–970. doi: 10.1046/j.1471-4159.1998.70030958.x. [DOI] [PubMed] [Google Scholar]
  22. Blankenship AG, Hamby AM, Firl A, Vyas S, Maxeiner S, Willecke K, Feller MB. The role of neuronal connexins 36 and 45 in shaping spontaneous firing patterns in the developing retina. J Neurosci Off J Soc Neurosci. 2011;31:9998–10008. doi: 10.1523/JNEUROSCI.5640-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Boisse L, Gill MJ, Power C. HIV infection of the central nervous system: clinical features and neuropathogenesis. Neurol Clin. 2008;26:799–819. x. doi: 10.1016/j.ncl.2008.04.002. [DOI] [PubMed] [Google Scholar]
  24. Bolanos JP, Medina JM. Induction of nitric oxide synthase inhibits gap junction permeability in cultured rat astrocytes. J Neurochem. 1996;66:2091–2099. doi: 10.1046/j.1471-4159.1996.66052091.x. [DOI] [PubMed] [Google Scholar]
  25. Bostanci MO, Bagirici F. The effects of octanol on penicillin induced epileptiform activity in rats: an in vivo study. Epilepsy Res. 2006;71:188–194. doi: 10.1016/j.eplepsyres.2006.06.010. [DOI] [PubMed] [Google Scholar]
  26. Boucher S, Bennett SA. Differential connexin expression, gap junction intercellular coupling, and hemichannel formation in NT2/D1 human neural progenitors and terminally differentiated hNT neurons. J Neurosci Res. 2003;72:393–404. doi: 10.1002/jnr.10575. [DOI] [PubMed] [Google Scholar]
  27. Brand-Schieber E, Werner P, Iacobas DA, Iacobas S, Beelitz M, Lowery SL, Spray DC, Scemes E. Connexin43, the major gap junction protein of astrocytes, is down-regulated in inflamed white matter in an animal model of multiple sclerosis. J Neurosci Res. 2005;80:798–808. doi: 10.1002/jnr.20474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Brooks-Kayal AR, Raol YH, Russek SJ. Alteration of epileptogenesis genes. Neurotherapeutics. 2009;6:312–318. doi: 10.1016/j.nurt.2009.01.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Bruzzone R, Hormuzdi SG, Barbe MT, Herb A, Monyer H. Pannexins, a family of gap junction proteins expressed in brain. Proc Natl Acad Sci U S A. 2003;100:13644–13649. doi: 10.1073/pnas.2233464100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Burnier L, Fontana P, Angelillo-Scherrer A, Kwak BR. Intercellular communication in atherosclerosis. Physiology (Bethesda) 2009;24:36–44. doi: 10.1152/physiol.00036.2008. [DOI] [PubMed] [Google Scholar]
  31. Butt AM, Ransom BR. Visualization of oligodendrocytes and astrocytes in the intact rat optic nerve by intracellular injection of lucifer yellow and horseradish peroxidase. Glia. 1989;2:470–475. doi: 10.1002/glia.440020609. [DOI] [PubMed] [Google Scholar]
  32. Calderon TM, Eugenin EA, Lopez L, Kumar SS, Hesselgesser J, Raine CS, Berman JW. A role for CXCL12 (SDF-1alpha) in the pathogenesis of multiple sclerosis: regulation of CXCL12 expression in astrocytes by soluble myelin basic protein. J Neuroimmunol. 2006;177:27–39. doi: 10.1016/j.jneuroim.2006.05.003. [DOI] [PubMed] [Google Scholar]
  33. Carlen PL, Skinner F, Zhang L, Naus C, Kushnir M, Perez Velazquez JL. The role of gap junctions in seizures. Brain Res Brain Res Rev. 2000;32:235–241. doi: 10.1016/s0165-0173(99)00084-3. [DOI] [PubMed] [Google Scholar]
  34. Chang Q, Balice-Gordon RJ. Gap junctional communication among developing and injured motor neurons. Brain Res Brain Res Rev. 2000;32:242–249. doi: 10.1016/s0165-0173(99)00085-5. [DOI] [PubMed] [Google Scholar]
  35. Chang Q, Gonzalez M, Pinter MJ, Balice-Gordon RJ. Gap junctional coupling and patterns of connexin expression among neonatal rat lumbar spinal motor neurons. J Neurosci. 1999;19:10813–10828. doi: 10.1523/JNEUROSCI.19-24-10813.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Chang Q, Tang W, Ahmad S, Zhou B, Lin X. Gap junction mediated intercellular metabolite transfer in the cochlea is compromised in connexin30 null mice. PLoS One. 2008;3:e4088. doi: 10.1371/journal.pone.0004088. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Cheng A, Tang H, Cai J, Zhu M, Zhang X, Rao M, Mattson MP. Gap junctional communication is required to maintain mouse cortical neural progenitor cells in a proliferative state. Dev Biol. 2004;272:203–216. doi: 10.1016/j.ydbio.2004.04.031. [DOI] [PubMed] [Google Scholar]
  38. Christie JM, Bark C, Hormuzdi SG, Helbig I, Monyer H, Westbrook GL. Connexin36 mediates spike synchrony in olfactory bulb glomeruli. Neuron. 2005;46:761–772. doi: 10.1016/j.neuron.2005.04.030. [DOI] [PubMed] [Google Scholar]
  39. Churchill MJ, Gorry PR, Cowley D, Lal L, Sonza S, Purcell DF, Thompson KA, Gabuzda D, McArthur JC, Pardo CA, Wesselingh SL. Use of laser capture microdissection to detect integrated HIV-1 DNA in macrophages and astrocytes from autopsy brain tissues. J Neurovirol. 2006;12:146–152. doi: 10.1080/13550280600748946. [DOI] [PubMed] [Google Scholar]
  40. Churchill MJ, Wesselingh SL, Cowley D, Pardo CA, McArthur JC, Brew BJ, Gorry PR. Extensive astrocyte infection is prominent in human immunodeficiency virus-associated dementia. Ann Neurol. 2009;66:253–258. doi: 10.1002/ana.21697. [DOI] [PubMed] [Google Scholar]
  41. Cina C, Maass K, Theis M, Willecke K, Bechberger JF, Naus CC. Involvement of the cytoplasmic C-terminal domain of connexin43 in neuronal migration. J Neurosci. 2009;29:2009–2021. doi: 10.1523/JNEUROSCI.5025-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Cirenei N, Colombo BM, Mesnil M, Benedetti S, Yamasaki H, Finocchiaro G. In vitro and in vivo effects of retrovirus-mediated transfer of the connexin 43 gene in malignant gliomas: consequences for HSVtk/GCV anticancer gene therapy. Gene Ther. 1998;5:1221–1226. doi: 10.1038/sj.gt.3300714. [DOI] [PubMed] [Google Scholar]
  43. Cohen-Salmon M, Regnault B, Cayet N, Caille D, Demuth K, Hardelin JP, Janel N, Meda P, Petit C. Connexin30 deficiency causes instrastrial fluid-blood barrier disruption within the cochlear stria vascularis. Proc Natl Acad Sci U S A. 2007;104:6229–6234. doi: 10.1073/pnas.0605108104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Conant K, Tornatore C, Atwood W, Meyers K, Traub R, Major EO. In vivo and in vitro infection of the astrocyte by HIV-1. Adv Neuroimmunol. 1994;4:287–289. doi: 10.1016/s0960-5428(06)80269-x. [DOI] [PubMed] [Google Scholar]
  45. Condorelli DF, Parenti R, Spinella F, Trovato Salinaro A, Belluardo N, Cardile V, Cicirata F. Cloning of a new gap junction gene (Cx36) highly expressed in mammalian brain neurons. Eur J Neurosci. 1998;10:1202–1208. doi: 10.1046/j.1460-9568.1998.00163.x. [DOI] [PubMed] [Google Scholar]
  46. Condorelli DF, Mudo G, Trovato-Salinaro A, Mirone MB, Amato G, Belluardo N. Connexin-30 mRNA is up-regulated in astrocytes and expressed in apoptotic neuronal cells of rat brain following kainate-induced seizures. Mol Cell Neurosci. 2002;21:94–113. doi: 10.1006/mcne.2002.1155. [DOI] [PubMed] [Google Scholar]
  47. Condorelli DF, Trovato-Salinaro A, Mudo G, Mirone MB, Belluardo N. Cellular expression of connexins in the rat brain: neuronal localization, effects of kainate-induced seizures and expression in apoptotic neuronal cells. Eur J Neurosci. 2003;18:1807–1827. doi: 10.1046/j.1460-9568.2003.02910.x. [DOI] [PubMed] [Google Scholar]
  48. Contreras JE, Sanchez HA, Eugenin EA, Speidel D, Theis M, Willecke K, Bukauskas FF, Bennett MV, Saez JC. Metabolic inhibition induces opening of unapposed connexin 43 gap junction hemichannels and reduces gap junctional communication in cortical astrocytes in culture. Proc Natl Acad Sci U S A. 2002;99:495–500. doi: 10.1073/pnas.012589799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Corvalan LA, Araya R, Branes MC, Saez PJ, Kalergis AM, Tobar JA, Theis M, Willecke K, Saez JC. Injury of skeletal muscle and specific cytokines induce the expression of gap junction channels in mouse dendritic cells. J Cell Physiol. 2007;211:649–660. doi: 10.1002/jcp.20971. [DOI] [PubMed] [Google Scholar]
  50. Crow DS, Beyer EC, Paul DL, Kobe SS, Lau AF. Phosphorylation of connexin43 gap junction protein in uninfected and Rous sarcoma virus-transformed mammalian fibroblasts. Mol Cell Biol. 1990;10:1754–1763. doi: 10.1128/mcb.10.4.1754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Crow DS, Kurata WE, Lau AF. Phosphorylation of connexin43 in cells containing mutant src oncogenes. Oncogene. 1992;7:999–1003. [PubMed] [Google Scholar]
  52. Dahl G, Locovei S. Pannexin: to gap or not to gap, is that a question? IUBMB Life. 2006;58:409–419. doi: 10.1080/15216540600794526. [DOI] [PubMed] [Google Scholar]
  53. Dale N. Dynamic ATP signalling and neural development. J Physiol. 2008;586:2429–2436. doi: 10.1113/jphysiol.2008.152207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. De Pina-Benabou MH, Srinivas M, Spray DC, Scemes E. Calmodulin kinase pathway mediates the K+-induced increase in Gap junctional communication between mouse spinal cord astrocytes. J Neurosci. 2001;21:6635–6643. doi: 10.1523/JNEUROSCI.21-17-06635.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. de Wit C, Roos F, Bolz SS, Kirchhoff S, Kruger O, Willecke K, Pohl U. Impaired conduction of vasodilation along arterioles in connexin40-deficient mice. Circ Res. 2000;86:649–655. doi: 10.1161/01.res.86.6.649. [DOI] [PubMed] [Google Scholar]
  56. de Wit C, Roos F, Bolz SS, Pohl U. Lack of vascular connexin 40 is associated with hypertension and irregular arteriolar vasomotion. Physiol Genomics. 2003;13:169–177. doi: 10.1152/physiolgenomics.00169.2002. [DOI] [PubMed] [Google Scholar]
  57. Del Monte U, Statuto M. Drop of connexins: a possible link between aging and cancer? Exp Gerontol. 2004;39:273–275. doi: 10.1016/j.exger.2003.10.010. [DOI] [PubMed] [Google Scholar]
  58. Dermietzel R, Traub O, Hwang TK, Beyer E, Bennett MV, Spray DC, Willecke K. Differential expression of three gap junction proteins in developing and mature brain tissues. Proc Natl Acad Sci U S A. 1989;86:10148–10152. doi: 10.1073/pnas.86.24.10148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Dermietzel R, Hertberg EL, Kessler JA, Spray DC. Gap junctions between cultured astrocytes: immunocytochemical, molecular, and electrophysiological analysis. J Neurosci. 1991;11:1421–1432. doi: 10.1523/JNEUROSCI.11-05-01421.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Dermietzel R, Farooq M, Kessler JA, Althaus H, Hertzberg EL, Spray DC. Oligodendrocytes express gap junction proteins connexin32 and connexin45. Glia. 1997;20:101–114. [PubMed] [Google Scholar]
  61. Devor A, Yarom Y. Coherence of subthreshold activity in coupled inferior olivary neurons. Ann N Y Acad Sci. 2002a;978:508. doi: 10.1111/j.1749-6632.2002.tb07594.x. [DOI] [PubMed] [Google Scholar]
  62. Devor A, Yarom Y. Electrotonic coupling in the inferior olivary nucleus revealed by simultaneous double patch recordings. J Neurophysiol. 2002b;87:3048–3058. doi: 10.1152/jn.2002.87.6.3048. [DOI] [PubMed] [Google Scholar]
  63. Devor A, Yarom Y. Generation and propagation of subthresh-old waves in a network of inferior olivary neurons. J Neurophysiol. 2002c;87:3059–3069. doi: 10.1152/jn.2002.87.6.3059. [DOI] [PubMed] [Google Scholar]
  64. Dezawa M, Nagano T. Immunohistochemical localization of cell adhesion molecules and cell-cell contact proteins during regeneration of the rat optic nerve induced by sciatic nerve auto-transplantation. Anat Rec. 1996;246:114–126. doi: 10.1002/(SICI)1097-0185(199609)246:1<114::AID-AR13>3.0.CO;2-S. [DOI] [PubMed] [Google Scholar]
  65. Dilber MS, Abedi MR, Christensson B, Bjorkstrand B, Kidder GM, Naus CC, Gahrton G, Smith CI. Gap junctions promote the bystander effect of herpes simplex virus thymidine kinase in vivo. Cancer Res. 1997;57:1523–1528. [PubMed] [Google Scholar]
  66. Dobrenis K, Chang HY, Pina-Benabou MH, Woodroffe A, Lee SC, Rozental R, Spray DC, Scemes E. Human and mouse microglia express connexin36, and functional gap junctions are formed between rodent microglia and neurons. J Neurosci Res. 2005;82:306–315. doi: 10.1002/jnr.20650. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Duffy HS, John GR, Lee SC, Brosnan CF, Spray DC. Reciprocal regulation of the junctional proteins claudin-1 and connexin43 by interleukin-1beta in primary human fetal astrocytes. J Neurosci. 2000;20:RC114. doi: 10.1523/JNEUROSCI.20-23-j0004.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Duffy HS, Sorgen PL, Girvin ME, O’Donnell P, Coombs W, Taffet SM, Delmar M, Spray DC. pH-dependent intramolecular binding and structure involving Cx43 cytoplasmic domains. J Biol Chem. 2002;277:36706–36714. doi: 10.1074/jbc.M207016200. [DOI] [PubMed] [Google Scholar]
  69. Duffy HS, Ashton AW, O’Donnell P, Coombs W, Taffet SM, Delmar M, Spray DC. Regulation of connexin43 protein complexes by intracellular acidification. Circ Res. 2004;94:215–222. doi: 10.1161/01.RES.0000113924.06926.11. [DOI] [PubMed] [Google Scholar]
  70. Dunina-Barkovskaya A. pH dependence of junctional conductance. Membr Cell Biol. 1998;11:793–801. [PubMed] [Google Scholar]
  71. Duval N, Gomes D, Calaora V, Calabrese A, Meda P, Bruzzone R. Cell coupling and Cx43 expression in embryonic mouse neural progenitor cells. J Cell Sci. 2002;115:3241–3251. doi: 10.1242/jcs.115.16.3241. [DOI] [PubMed] [Google Scholar]
  72. Elble RJ. Central mechanisms of tremor. J Clin Neurophysiol. 1996;13:133–144. doi: 10.1097/00004691-199603000-00004. [DOI] [PubMed] [Google Scholar]
  73. Elias LA, Kriegstein AR. Gap junctions: multifaceted regulators of embryonic cortical development. Trends Neurosci. 2008;31:243–250. doi: 10.1016/j.tins.2008.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Elias LA, Wang DD, Kriegstein AR. Gap junction adhesion is necessary for radial migration in the neocortex. Nature. 2007;448:901–907. doi: 10.1038/nature06063. [DOI] [PubMed] [Google Scholar]
  75. Elisevich K, Rempel SA, Smith BJ, Edvardsen K. Hippocampal connexin 43 expression in human complex partial seizure disorder. Exp Neurol. 1997;145:154–164. doi: 10.1006/exnr.1997.6467. [DOI] [PubMed] [Google Scholar]
  76. Enkvist MO, McCarthy KD. Astroglial gap junction communication is increased by treatment with either glutamate or high K+ concentration. J Neurochem. 1994;62:489–495. doi: 10.1046/j.1471-4159.1994.62020489.x. [DOI] [PubMed] [Google Scholar]
  77. Eugenín EA, Berman JW. Gap junctions mediate human immunodeficiency virus-bystander killing in astrocytes. J Neurosci. 2007;27:12844–12850. doi: 10.1523/JNEUROSCI.4154-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Eugenín EA, Eckardt D, Theis M, Willecke K, Bennett MV, Saez JC. Microglia at brain stab wounds express connexin 43 and in vitro form functional gap junctions after treatment with interferon-gamma and tumor necrosis factor-alpha. Proc Natl Acad Sci U S A. 2001;98:4190–4195. doi: 10.1073/pnas.051634298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Eugenín EA, Branes MC, Berman JW, Saez JC. TNF-alpha plus IFN-gamma induce connexin43 expression and formation of gap junctions between human monocytes/macrophages that enhance physiological responses. J Immunol. 2003;170:1320–1328. doi: 10.4049/jimmunol.170.3.1320. [DOI] [PubMed] [Google Scholar]
  80. Eugenín EA, Osiecki K, Lopez L, Goldstein H, Calderon TM, Berman JW. CCL2/monocyte chemoattractant protein-1 mediates enhanced transmigration of human immunodeficiency virus (HIV)-infected leukocytes across the blood–brain barrier: a potential mechanism of HIV-CNS invasion and NeuroAIDS. J Neurosci. 2006;26:1098–1106. doi: 10.1523/JNEUROSCI.3863-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Eugenín EA, Gonzalez HE, Sanchez HA, Branes MC, Saez JC. Inflammatory conditions induce gap junctional communication between rat Kupffer cells both in vivo and in vitro. Cell Immunol. 2007;247:103–110. doi: 10.1016/j.cellimm.2007.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Eugenín EA, Clements JE, Zink MC, Berman JW. Human immunodeficiency virus infection of human astrocytes disrupts blood–brain barrier integrity by a gap junction-dependent mechanism. J Neurosci Off J Soc Neurosci. 2011;31:9456–9465. doi: 10.1523/JNEUROSCI.1460-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Faccini AM, Cairney M, Ashrafi GH, Finbow ME, Campo MS, Pitts JD. The bovine papillomavirus type 4 E8 protein binds to ductin and causes loss of gap junctional intercellular communication in primary fibroblasts. J Virol. 1996;70:9041–9045. doi: 10.1128/jvi.70.12.9041-9045.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Fatemi SH, Folsom TD, Reutiman TJ, Sidwell RW. Viral regulation of aquaporin 4, connexin 43, microcephalin and nucleolin. Schizophr Res. 2008;98:163–177. doi: 10.1016/j.schres.2007.09.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Faustmann PM, Haase CG, Romberg S, Hinkerohe D, Szlachta D, Smikalla D, Krause D, Dermietzel R. Microglia activation influences dye coupling and Cx43 expression of the astrocytic network. Glia. 2003;42:101–108. doi: 10.1002/glia.10141. [DOI] [PubMed] [Google Scholar]
  86. Filson AJ, Azarnia R, Beyer EC, Loewenstein WR, Brugge JS. Tyrosine phosphorylation of a gap junction protein correlates with inhibition of cell-to-cell communication. Cell Growth Differ. 1990;1:661–668. [PubMed] [Google Scholar]
  87. Firouzi M, Bierhuizen MF, Kok B, Teunissen BE, Jansen AT, Jongsma HJ, Groenewegen WA. The human Cx40 promoter polymorphism -44 G—>A differentially affects transcriptional regulation by Sp1 and GATA4. Biochim Biophys Acta. 2006a;1759:491–496. doi: 10.1016/j.bbaexp.2006.09.002. [DOI] [PubMed] [Google Scholar]
  88. Firouzi M, Kok B, Spiering W, Busjahn A, Bezzina CR, Ruijter JM, Koeleman BP, Schipper M, Groenewegen WA, Jongsma HJ, de Leeuw PW. Polymorphisms in human connexin40 gene promoter are associated with increased risk of hypertension in men. J Hypertens. 2006b;24:325–330. doi: 10.1097/01.hjh.0000200512.40818.47. [DOI] [PubMed] [Google Scholar]
  89. Fischer NO, Mbuy GN, Woodruff RI. HSV-2 disrupts gap junctional intercellular communication between mammalian cells in vitro. J Virol Methods. 2001;91:157–166. doi: 10.1016/s0166-0934(00)00260-3. [DOI] [PubMed] [Google Scholar]
  90. Fonseca CG, Green CR, Nicholson LF. Upregulation in astrocytic connexin 43 gap junction levels may exacerbate generalized seizures in mesial temporal lobe epilepsy. Brain Res. 2002;929:105–116. doi: 10.1016/s0006-8993(01)03289-9. [DOI] [PubMed] [Google Scholar]
  91. Frank M, Eiberger B, Janssen-Bienhold U, de Sevilla Muller LP, Tjarks A, Kim JS, Maschke S, Dobrowolski R, Sasse P, Weiler R, Fleischmann BK, Willecke K. Neuronal connexin-36 can functionally replace connexin-45 in mouse retina but not in the developing heart. J Cell Sci. 2010;123:3605–3615. doi: 10.1242/jcs.068668. [DOI] [PubMed] [Google Scholar]
  92. Friedman D, Strowbridge BW. Both electrical and chemical synapses mediate fast network oscillations in the olfactory bulb. J Neurophysiol. 2003;89:2601–2610. doi: 10.1152/jn.00887.2002. [DOI] [PubMed] [Google Scholar]
  93. Froger N, Orellana JA, Cohen-Salmon M, Ezan P, Amigou E, Saez JC, Giaume C. Cannabinoids prevent the opposite regulation of astroglial connexin43 hemichannels and gap junction channels induced by pro-inflammatory treatments. J Neurochem. 2009;111:1383–1397. doi: 10.1111/j.1471-4159.2009.06407.x. [DOI] [PubMed] [Google Scholar]
  94. Froger N, Orellana JA, Calvo CF, Amigou E, Kozoriz MG, Naus CC, Saez JC, Giaume C. Inhibition of cytokine-induced connexin43 hemichannel activity in astrocytes is neuroprotective. Mol Cell Neurosci. 2010;45:37–46. doi: 10.1016/j.mcn.2010.05.007. [DOI] [PubMed] [Google Scholar]
  95. Gajda Z, Szupera Z, Blazso G, Szente M. Quinine, a blocker of neuronal cx36 channels, suppresses seizure activity in rat neocortex in vivo. Epilepsia. 2005;46:1581–1591. doi: 10.1111/j.1528-1167.2005.00254.x. [DOI] [PubMed] [Google Scholar]
  96. Galanopoulou AS. Mutations affecting GABAergic signaling in seizures and epilepsy. Pflugers Arch. 2010;460:505–523. doi: 10.1007/s00424-010-0816-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Galarreta M, Hestrin S. A network of fast-spiking cells in the neocortex connected by electrical synapses. Nature. 1999;402:72–75. doi: 10.1038/47029. [DOI] [PubMed] [Google Scholar]
  98. Gareri P, Condorelli D, Belluardo N, Russo E, Loiacono A, Barresi V, Trovato-Salinaro A, Mirone MB, Ferreri Ibbadu G, De Sarro G. Anticonvulsant effects of carbenoxolone in genetically epilepsy prone rats (GEPRs) Neuropharmacology. 2004;47:1205–1216. doi: 10.1016/j.neuropharm.2004.08.021. [DOI] [PubMed] [Google Scholar]
  99. Gareri P, Condorelli D, Belluardo N, Citraro R, Barresi V, Trovato-Salinaro A, Mudo G, Ibbadu GF, Russo E, De Sarro G. Antiabsence effects of carbenoxolone in two genetic animal models of absence epilepsy (WAG/Rij rats and lh/lh mice) Neuropharmacology. 2005;49:551–563. doi: 10.1016/j.neuropharm.2005.04.012. [DOI] [PubMed] [Google Scholar]
  100. Garg S, Md Syed M, Kielian T. Staphylococcus aureus-derived peptidoglycan induces Cx43 expression and functional gap junction intercellular communication in microglia. J Neurochem. 2005;95:475–483. doi: 10.1111/j.1471-4159.2005.03384.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Garre JM, Retamal MA, Cassina P, Barbeito L, Bukauskas FF, Saez JC, Bennett MV, Abudara V. FGF-1 induces ATP release from spinal astrocytes in culture and opens pannexin and connexin hemichannels. Proc Natl Acad Sci U S A. 2010;107:22659–22664. doi: 10.1073/pnas.1013793107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Giaume C, Fromaget C, el Aoumari A, Cordier J, Glowinski J, Gros D. Gap junctions in cultured astrocytes: single-channel currents and characterization of channel-forming protein. Neuron. 1991;6:133–143. doi: 10.1016/0896-6273(91)90128-m. [DOI] [PubMed] [Google Scholar]
  103. Giaume C, Cordier J, Glowinski J. Endothelins inhibit junctional permeability in cultured mouse astrocytes. Eur J Neurosci. 1992;4:877–881. doi: 10.1111/j.1460-9568.1992.tb00198.x. [DOI] [PubMed] [Google Scholar]
  104. Gibson JR, Beierlein M, Connors BW. Two networks of electrically coupled inhibitory neurons in neocortex. Nature. 1999;402:75–79. doi: 10.1038/47035. [DOI] [PubMed] [Google Scholar]
  105. Gloor SM, Wachtel M, Bolliger MF, Ishihara H, Landmann R, Frei K. Molecular and cellular permeability control at the blood–brain barrier. Brain Res Brain Res Rev. 2001;36:258–264. doi: 10.1016/s0165-0173(01)00102-3. [DOI] [PubMed] [Google Scholar]
  106. Gonzalez-Nieto D, Gomez-Hernandez JM, Larrosa B, Gutierrez C, Munoz MD, Fasciani I, O’Brien J, Zappala A, Cicirata F, Barrio LC. Regulation of neuronal connexin-36 channels by pH. Proc Natl Acad Sci U S A. 2008;105:17169–17174. doi: 10.1073/pnas.0804189105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. González-Scarano F, Martín-García J. The neuropathogenesis of AIDS. Nat Rev Immunol. 2005;5:69–81. doi: 10.1038/nri1527. [DOI] [PubMed] [Google Scholar]
  108. Gorry PR, Ong C, Thorpe J, Bannwarth S, Thompson KA, Gatignol A, Vesselingh SL, Purcell DF. Astrocyte infection by HIV-1: mechanisms of restricted virus replication, and role in the pathogenesis of HIV-1-associated dementia. Curr HIV Res. 2003;1:463–473. doi: 10.2174/1570162033485122. [DOI] [PubMed] [Google Scholar]
  109. Gosejacob D, Dublin P, Bedner P, Huttmann K, Zhang J, Tress O, Willecke K, Pfrieger F, Steinhauser C, Theis M. Role of astroglial connexin30 in hippocampal gap junction coupling. Glia. 2011;59:511–519. doi: 10.1002/glia.21120. [DOI] [PubMed] [Google Scholar]
  110. Granda B, Tabernero A, Sanchez-Abarca LI, Medina JM. The K-ATP channel regulates the effect of Ca2+ on gap junction permeability in cultured astrocytes. FEBS Lett. 1998;427:41–45. doi: 10.1016/s0014-5793(98)00390-1. [DOI] [PubMed] [Google Scholar]
  111. Grignet-Debrus C, Cool V, Baudson N, Velu T, Calberg-Bacq CM. The role of cellular- and prodrug-associated factors in the bystander effect induced by the Varicella zoster and Herpes simplex viral thymidine kinases in suicide gene therapy. Cancer Gene Ther. 2000;7:1456–1468. doi: 10.1038/sj.cgt.7700250. [DOI] [PubMed] [Google Scholar]
  112. Guthrie PB, Knappenberger J, Segal M, Bennett MV, Charles AC, Kater SB. ATP released from astrocytes mediates glial calcium waves. J Neurosci. 1999;19:520–528. doi: 10.1523/JNEUROSCI.19-02-00520.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Haefliger JA, Nicod P, Meda P. Contribution of connexins to the function of the vascular wall. Cardiovasc Res. 2004;62:345–356. doi: 10.1016/j.cardiores.2003.11.015. [DOI] [PubMed] [Google Scholar]
  114. Handel A, Yates A, Pilyugin SS, Antia R. Gap junction-mediated antigen transport in immune responses. Trends Immunol. 2007;28:463–466. doi: 10.1016/j.it.2007.08.006. [DOI] [PubMed] [Google Scholar]
  115. Harris AL. Emerging issues of connexin channels: biophysics fills the gap. Q Rev Biophys. 2001;34:325–472. doi: 10.1017/s0033583501003705. [DOI] [PubMed] [Google Scholar]
  116. Harris AL. Connexin channel permeability to cytoplasmic molecules. Prog Biophys Mol Biol. 2007 doi: 10.1016/j.pbiomolbio.2007.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Hartfield EM, Rinaldi F, Glover CP, Wong LF, Caldwell MA, Uney JB. Connexin 36 expression regulates neuronal differentiation from neural progenitor cells. PLoS One. 2011;6:e14746. doi: 10.1371/journal.pone.0014746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Hauer RN, Groenewegen WA, Firouzi M, Ramanna H, Jongsma HJ. Cx40 polymorphism in human atrial fibrillation. Adv Cardiol. 2006;42:284–291. doi: 10.1159/000092579. [DOI] [PubMed] [Google Scholar]
  119. Haughey NJ, Mattson MP. Alzheimer’s amyloid beta-peptide enhances ATP/gap junction-mediated calcium-wave propagation in astrocytes. Neuromol Med. 2003;3:173–180. doi: 10.1385/NMM:3:3:173. [DOI] [PubMed] [Google Scholar]
  120. Hormuzdi SG, Pais I, LeBeau FE, Towers SK, Rozov A, Buhl EH, Whittington MA, Monyer H. Impaired electrical signaling disrupts gamma frequency oscillations in connexin 36-deficient mice. Neuron. 2001;31:487–495. doi: 10.1016/s0896-6273(01)00387-7. [DOI] [PubMed] [Google Scholar]
  121. Hsiao HJ, Liu PA, Yeh HI, Wang CY. Classical swine fever virus down-regulates endothelial connexin 43 gap junctions. Arch Virol. 2010;155:1107–1116. doi: 10.1007/s00705-010-0693-1. [DOI] [PubMed] [Google Scholar]
  122. Huang RP, Fan Y, Hossain MZ, Peng A, Zeng ZL, Boynton AL. Reversion of the neoplastic phenotype of human glioblastoma cells by connexin 43 (cx43) Cancer Res. 1998;58:5089–5096. [PubMed] [Google Scholar]
  123. Huang RP, Hossain MZ, Sehgal A, Boynton AL. Reduced connexin43 expression in high-grade human brain glioma cells. J Surg Oncol. 1999;70:21–24. doi: 10.1002/(sici)1096-9098(199901)70:1<21::aid-jso4>3.0.co;2-0. [DOI] [PubMed] [Google Scholar]
  124. Huang Q, Liu XZ, Kang CS, Wang GX, Zhong Y, Pu PY. The anti-glioma effect of suicide gene therapy using BMSC expressing HSV/TK combined with overexpression of Cx43 in glioma cells. Cancer Gene Ther. 2010;17:192–202. doi: 10.1038/cgt.2009.64. [DOI] [PubMed] [Google Scholar]
  125. Iglesias R, Locovei S, Roque A, Alberto AP, Dahl G, Spray DC, Scemes E. P2X7 receptor-Pannexin1 complex: pharmacology and signaling. Am J Physiol Cell Physiol. 2008;295:C752–C760. doi: 10.1152/ajpcell.00228.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Iglesias R, Dahl G, Qiu F, Spray DC, Scemes E. Pannexin 1: the molecular substrate of astrocyte “hemichannels”. J Neurosci. 2009;29:7092–7097. doi: 10.1523/JNEUROSCI.6062-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Ionasescu VV. X-linked Charcot-Marie-Tooth disease and connexin32. Cell Biol Int. 1998;22:807–813. doi: 10.1006/cbir.1998.0387. [DOI] [PubMed] [Google Scholar]
  128. Ionasescu V, Ionasescu R, Searby C. Correlation between connexin 32 gene mutations and clinical phenotype in X-linked dominant Charcot-Marie-Tooth neuropathy. Am J Med Genet. 1996;63:486–491. doi: 10.1002/(SICI)1096-8628(19960614)63:3<486::AID-AJMG14>3.0.CO;2-I. [DOI] [PubMed] [Google Scholar]
  129. Ishikawa M, Iwamoto T, Nakamura T, Doyle A, Fukumoto S, Yamada Y. Pannexin 3 functions as an ER Ca(2+) channel, hemi-channel, and gap junction to promote osteoblast differentiation. J Cell Biol. 2011;193:1257–1274. doi: 10.1083/jcb.201101050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Iwabuchi S, Kawahara K. Functional significance of the negative-feedback regulation of ATP release via pannexin-1 hemi-channels under ischemic stress in astrocytes. Neurochem Int. 2011;58:376–384. doi: 10.1016/j.neuint.2010.12.013. [DOI] [PubMed] [Google Scholar]
  131. Jacobson GM, Voss LJ, Melin SM, Mason JP, Cursons RT, Steyn-Ross DA, Steyn-Ross ML, Sleigh JW. Connexin36 knockout mice display increased sensitivity to pentylenetetrazol-induced seizure-like behaviors. Brain Res. 2010;1360:198–204. doi: 10.1016/j.brainres.2010.09.006. [DOI] [PubMed] [Google Scholar]
  132. Jahromi SS, Wentlandt K, Piran S, Carlen PL. Anticonvulsant actions of gap junctional blockers in an in vitro seizure model. J Neurophysiol. 2002;88:1893–1902. doi: 10.1152/jn.2002.88.4.1893. [DOI] [PubMed] [Google Scholar]
  133. John GR, Scemes E, Suadicani SO, Liu JS, Charles PC, Lee SC, Spray DC, Brosnan CF. IL-1beta differentially regulates calcium wave propagation between primary human fetal astrocytes via pathways involving P2 receptors and gap junction channels. Proc Natl Acad Sci U S A. 1999;96:11613–11618. doi: 10.1073/pnas.96.20.11613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Kamasawa N, Sik A, Morita M, Yasumura T, Davidson KG, Nagy JI, Rash JE. Connexin-47 and connexin-32 in gap junctions of oligodendrocyte somata, myelin sheaths, paranodal loops and Schmidt-Lanterman incisures: implications for ionic homeostasis and potassium siphoning. Neuroscience. 2005;136:65–86. doi: 10.1016/j.neuroscience.2005.08.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Kang JQ, Macdonald RL. Making sense of nonsense GABA(A) receptor mutations associated with genetic epilepsies. Trends Mol Med. 2009;15:430–438. doi: 10.1016/j.molmed.2009.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Kanneganti TD, Lamkanfi M, Kim YG, Chen G, Park JH, Franchi L, Vandenabeele P, Nunez G. Pannexin-1-mediated recognition of bacterial molecules activates the cryopyrin inflammasome independent of Toll-like receptor signaling. Immunity. 2007;26:433–443. doi: 10.1016/j.immuni.2007.03.008. [DOI] [PubMed] [Google Scholar]
  137. Kaul M, Garden GA, Lipton SA. Pathways to neuronal injury and apoptosis in HIV-associated dementia. Nature. 2001;410:988–994. doi: 10.1038/35073667. [DOI] [PubMed] [Google Scholar]
  138. Kawasaki A, Hayashi T, Nakachi K, Trosko JE, Sugihara K, Kotake Y, Ohta S. Modulation of connexin 43 in rotenone-induced model of Parkinson’s disease. Neuroscience. 2009;160:61–68. doi: 10.1016/j.neuroscience.2009.01.080. [DOI] [PubMed] [Google Scholar]
  139. Kettenmann H, Orkand RK, Schachner M. Coupling among identified cells in mammalian nervous system cultures. J Neurosci. 1983;3:506–516. doi: 10.1523/JNEUROSCI.03-03-00506.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Kielian T. Glial connexins and gap junctions in CNS inflammation and disease. J Neurochem. 2008;106:1000–1016. doi: 10.1111/j.1471-4159.2008.05405.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Kielian T, Esen N. Effects of neuroinflammation on glia-glia gap junctional intercellular communication: a perspective. Neurochem Int. 2004;45:429–436. doi: 10.1016/j.neuint.2003.09.010. [DOI] [PubMed] [Google Scholar]
  142. Kleopa KA, Scherer SS. Inherited neuropathies. Neurol Clin. 2002;20:679–709. doi: 10.1016/s0733-8619(01)00016-0. [DOI] [PubMed] [Google Scholar]
  143. Kleopa KA, Yum SW, Scherer SS. Cellular mechanisms of connexin32 mutations associated with CNS manifestations. J Neurosci Res. 2002;68:522–534. doi: 10.1002/jnr.10255. [DOI] [PubMed] [Google Scholar]
  144. Kleopa KA, Orthmann JL, Enriquez A, Paul DL, Scherer SS. Unique distributions of the gap junction proteins connexin29, connexin32, and connexin47 in oligodendrocytes. Glia. 2004;47:346–357. doi: 10.1002/glia.20043. [DOI] [PubMed] [Google Scholar]
  145. Knabb MT, Danielsen CA, McShane-Kay K, Mbuy GK, Woodruff RI. Herpes simplex virus-type 2 infectivity and agents that block gap junctional intercellular communication. Virus Res. 2007;124:212–219. doi: 10.1016/j.virusres.2006.11.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Koster-Patzlaff C, Hosseini SM, Reuss B. Persistent Borna disease virus infection changes expression and function of astroglial gap junctions in vivo and in vitro. Brain Res. 2007;1184:316–332. doi: 10.1016/j.brainres.2007.09.062. [DOI] [PubMed] [Google Scholar]
  147. Koster-Patzlaff C, Hosseini SM, Reuss B. Layer specific changes of astroglial gap junctions in the rat cerebellar cortex by persistent Borna disease virus infection. Brain Res. 2008;1219:143–158. doi: 10.1016/j.brainres.2008.04.062. [DOI] [PubMed] [Google Scholar]
  148. Koster-Patzlaff C, Hosseini SM, Reuss B. Loss of connexin36 in rat hippocampus and cerebellar cortex in persistent Borna disease virus infection. J Chem Neuroanat. 2009;37:118–127. doi: 10.1016/j.jchemneu.2008.10.004. [DOI] [PubMed] [Google Scholar]
  149. Kraft-Terry SD, Buch SJ, Fox HS, Gendelman HE. A coat of many colors: neuroimmune crosstalk in human immunodeficiency virus infection. Neuron. 2009;64:133–145. doi: 10.1016/j.neuron.2009.09.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Kreuzberg MM, Deuchars J, Weiss E, Schober A, Sonntag S, Wellershaus K, Draguhn A, Willecke K. Expression of connexin30.2 in interneurons of the central nervous system in the mouse. Mol Cell Neurosci. 2008;37:119–134. doi: 10.1016/j.mcn.2007.09.003. [DOI] [PubMed] [Google Scholar]
  151. Kruger O, Plum A, Kim JS, Winterhager E, Maxeiner S, Hallas G, Kirchhoff S, Traub O, Lamers WH, Willecke K. Defective vascular development in connexin 45-deficient mice. Development. 2000;127:4179–4193. doi: 10.1242/dev.127.19.4179. [DOI] [PubMed] [Google Scholar]
  152. Kubota T, Sato K, Arishima H, Takeuchi H, Kitai R, Nakagawa T. Astroblastoma: immunohistochemical and ultrastructural study of distinctive epithelial and probable tanycytic differentiation. Neuropathology. 2006;26:72–81. doi: 10.1111/j.1440-1789.2006.00636.x. [DOI] [PubMed] [Google Scholar]
  153. Kunzelmann P, Blumcke I, Traub O, Dermietzel R, Willecke K. Coexpression of connexin45 and -32 in oligodendrocytes of rat brain. J Neurocytol. 1997;26:17–22. doi: 10.1023/a:1018555207379. [DOI] [PubMed] [Google Scholar]
  154. Lai CP, Bechberger JF, Thompson RJ, MacVicar BA, Bruzzone R, Naus CC. Tumor-suppressive effects of pannexin 1 in C6 glioma cells. Cancer Res. 2007;67:1545–1554. doi: 10.1158/0008-5472.CAN-06-1396. [DOI] [PubMed] [Google Scholar]
  155. Lee SH, Magge S, Spencer DD, Sontheimer H, Cornell-Bell AH. Human epileptic astrocytes exhibit increased gap junction coupling. Glia. 1995;15:195–202. doi: 10.1002/glia.440150212. [DOI] [PubMed] [Google Scholar]
  156. Lee IH, Lindqvist E, Kiehn O, Widenfalk J, Olson L. Glial and neuronal connexin expression patterns in the rat spinal cord during development and following injury. J Comp Neurol. 2005;489:1–10. doi: 10.1002/cne.20567. [DOI] [PubMed] [Google Scholar]
  157. Lemaire I, Falzoni S, Zhang B, Pellegatti P, Di Virgilio F. The P2X7 receptor and Pannexin-1 are both required for the promotion of multinucleated macrophages by the inflammatory cytokine GM-CSF. J Immunol. 2011;187:3878–3887. doi: 10.4049/jimmunol.1002780. [DOI] [PubMed] [Google Scholar]
  158. Levite M, Hermelin A. Autoimmunity to the glutamate receptor in mice–a model for Rasmussen’s encephalitis? J Autoimmun. 1999;13:73–82. doi: 10.1006/jaut.1999.0297. [DOI] [PubMed] [Google Scholar]
  159. Levite M, Fleidervish IA, Schwarz A, Pelled D, Futerman AH. Autoantibodies to the glutamate receptor kill neurons via activation of the receptor ion channel. J Autoimmun. 1999;13:61–72. doi: 10.1006/jaut.1999.0301. [DOI] [PubMed] [Google Scholar]
  160. Li X, Simard JM. Connexin45 gap junction channels in rat cerebral vascular smooth muscle cells. Am J Physiol Heart Circ Physiol. 2001;281:H1890–H1898. doi: 10.1152/ajpheart.2001.281.5.H1890. [DOI] [PubMed] [Google Scholar]
  161. Li J, Shen H, Naus CC, Zhang L, Carlen PL. Upregulation of gap junction connexin 32 with epileptiform activity in the isolated mouse hippocampus. Neuroscience. 2001;105:589–598. doi: 10.1016/s0306-4522(01)00204-4. [DOI] [PubMed] [Google Scholar]
  162. Li X, Kamasawa N, Ciolofan C, Olson CO, Lu S, Davidson KG, Yasumura T, Shigemoto R, Rash JE, Nagy JI. Connexin45-containing neuronal gap junctions in rodent retina also contain connexin36 in both apposing hemiplaques, forming bihomotypic gap junctions, with scaffolding contributed by zonula occludens-1. J Neurosci Off J Soc Neurosci. 2008;28:9769–9789. doi: 10.1523/JNEUROSCI.2137-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Liao Y, Day KH, Damon DN, Duling BR. Endothelial cell-specific knockout of connexin 43 causes hypotension and bradycardia in mice. Proc Natl Acad Sci U S A. 2001;98:9989–9994. doi: 10.1073/pnas.171305298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Lim KL, Dawson VL, Dawson TM. The genetics of Parkinson’s disease. Curr Neurol Neurosci Rep. 2002;2:439–446. doi: 10.1007/s11910-002-0071-9. [DOI] [PubMed] [Google Scholar]
  165. Lin JH, Weigel H, Cotrina ML, Liu S, Bueno E, Hansen AJ, Hansen TW, Goldman S, Nedergaard M. Gap-junction-mediated propagation and amplification of cell injury. Nat Neurosci. 1998;1:494–500. doi: 10.1038/2210. [DOI] [PubMed] [Google Scholar]
  166. Lin JH, Takano T, Cotrina ML, Arcuino G, Kang J, Liu S, Gao Q, Jiang L, Li F, Lichtenberg-Frate H, Haubrich S, Willecke K, Goldman SA, Nedergaard M. Connexin 43 enhances the adhesivity and mediates the invasion of malignant glioma cells. J Neurosci. 2002;22:4302–4311. doi: 10.1523/JNEUROSCI.22-11-04302.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Locovei S, Wang J, Dahl G. Activation of pannexin 1 channels by ATP through P2Y receptors and by cytoplasmic calcium. FEBS Lett. 2006;580:239–244. doi: 10.1016/j.febslet.2005.12.004. [DOI] [PubMed] [Google Scholar]
  168. Locovei S, Scemes E, Qiu F, Spray DC, Dahl G. Pannexin1 is part of the pore forming unit of the P2X(7) receptor death complex. FEBS Lett. 2007;581:483–488. doi: 10.1016/j.febslet.2006.12.056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Long MA, Deans MR, Paul DL, Connors BW. Rhythmicity without synchrony in the electrically uncoupled inferior olive. J Neurosci Off J Soc Neurosci. 2002a;22:10898–10905. doi: 10.1523/JNEUROSCI.22-24-10898.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Long MA, Deans MR, Paul DL, Connors BW. Rhythmicity without synchrony in the electrically uncoupled inferior olive. J Neurosci. 2002b;22:10898–10905. doi: 10.1523/JNEUROSCI.22-24-10898.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Long MA, Jutras MJ, Connors BW, Burwell RD. Electrical synapses coordinate activity in the suprachiasmatic nucleus. Nat Neurosci. 2005;8:61–66. doi: 10.1038/nn1361. [DOI] [PubMed] [Google Scholar]
  172. Lotharius J, Brundin P. Pathogenesis of Parkinson’s disease: dopamine, vesicles and alpha-synuclein. Nat Rev Neurosci. 2002;3:932–942. doi: 10.1038/nrn983. [DOI] [PubMed] [Google Scholar]
  173. Lynn BD, Marotta CA, Nagy JI. Propagation of intercellular calcium waves in PC12 cells overexpressing a carboxy-terminal fragment of amyloid precursor protein. Neurosci Lett. 1995;199:21–24. doi: 10.1016/0304-3940(95)12028-3. [DOI] [PubMed] [Google Scholar]
  174. Maglione M, Tress O, Haas B, Karram K, Trotter J, Willecke K, Kettenmann H. Oligodendrocytes in mouse corpus callosum are coupled via gap junction channels formed by connexin47 and connexin32. Glia. 2010;58:1104–1117. doi: 10.1002/glia.20991. [DOI] [PubMed] [Google Scholar]
  175. Magnotti LM, Goodenough DA, Paul DL. Functional heterotypic interactions between astrocyte and oligodendrocyte connexins. Glia. 2011;59:26–34. doi: 10.1002/glia.21073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Maier N, Guldenagel M, Sohl G, Siegmund H, Willecke K, Draguhn A. Reduction of high-frequency network oscillations (ripples) and pathological network discharges in hippocampal slices from connexin 36-deficient mice. J Physiol. 2002;541:521–528. doi: 10.1113/jphysiol.2002.017624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Malarkey EB, Parpura V. Mechanisms of glutamate release from astrocytes. Neurochem Int. 2008;52:142–154. doi: 10.1016/j.neuint.2007.06.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Mann-Metzer P, Yarom Y. Electrotonic coupling interacts with intrinsic properties to generate synchronized activity in cerebellar networks of inhibitory interneurons. J Neurosci. 1999;19:3298–3306. doi: 10.1523/JNEUROSCI.19-09-03298.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Marconi P, Tamura M, Moriuchi S, Krisky DM, Niranjan A, Goins WF, Cohen JB, Glorioso JC. Connexin 43-enhanced suicide gene therapy using herpesviral vectors. Mol Ther. 2000;1:71–81. doi: 10.1006/mthe.1999.0008. [DOI] [PubMed] [Google Scholar]
  180. Marrero H, Orkand RK. Nerve impulses increase glial intercellular permeability. Glia. 1996;16:285–289. doi: 10.1002/(SICI)1098-1136(199603)16:3<285::AID-GLIA11>3.0.CO;2-W. [DOI] [PubMed] [Google Scholar]
  181. Martinez AD, Saez JC. Arachidonic acid-induced dye uncoupling in rat cortical astrocytes is mediated by arachidonic acid byproducts. Brain Res. 1999;816:411–423. doi: 10.1016/s0006-8993(98)01016-6. [DOI] [PubMed] [Google Scholar]
  182. Martínez AD, Eugenin EA, Branes MC, Bennett MV, Saez JC. Identification of second messengers that induce expression of functional gap junctions in microglia cultured from newborn rats. Brain Res. 2002;943:191–201. doi: 10.1016/s0006-8993(02)02621-5. [DOI] [PubMed] [Google Scholar]
  183. Martini R. Animal models for inherited peripheral neuropathies: chances to find treatment strategies? J Neurosci Res. 2000;61:244–250. doi: 10.1002/1097-4547(20000801)61:3<244::AID-JNR2>3.0.CO;2-R. [DOI] [PubMed] [Google Scholar]
  184. Matsue H, Yao J, Matsue K, Nagasaka A, Sugiyama H, Aoki R, Kitamura M, Shimada S. Gap junction-mediated intercellular communication between dendritic cells (DCs) is required for effective activation of DCs. J Immunol. 2006;176:181–190. doi: 10.4049/jimmunol.176.1.181. [DOI] [PubMed] [Google Scholar]
  185. Maxeiner S, Kruger O, Schilling K, Traub O, Urschel S, Willecke K. Spatiotemporal transcription of connexin45 during brain development results in neuronal expression in adult mice. Neuroscience. 2003;119:689–700. doi: 10.1016/s0306-4522(03)00077-0. [DOI] [PubMed] [Google Scholar]
  186. Mei X, Ezan P, Giaume C, Koulakoff A. Astroglial connexin immunoreactivity is specifically altered at beta-amyloid plaques in beta-amyloid precursor protein/presenilin1 mice. Neuroscience. 2010;171:92–105. doi: 10.1016/j.neuroscience.2010.08.001. [DOI] [PubMed] [Google Scholar]
  187. Meier C, Dermietzel R, Davidson KG, Yasumura T, Rash JE. Connexin32-containing gap junctions in Schwann cells at the internodal zone of partial myelin compaction and in Schmidt-Lanterman incisures. J Neurosci. 2004;24:3186–3198. doi: 10.1523/JNEUROSCI.5146-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Meldrum BS. The role of glutamate in epilepsy and other CNS disorders. Neurology. 1994;44:S14–S23. [PubMed] [Google Scholar]
  189. Meldrum BS, Rogawski MA. Molecular targets for antiepileptic drug development. Neurotherapeutics. 2007;4:18–61. doi: 10.1016/j.nurt.2006.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Meme W, Ezan P, Venance L, Glowinski J, Giaume C. ATP-induced inhibition of gap junctional communication is enhanced by interleukin-1 beta treatment in cultured astrocytes. Neuroscience. 2004;126:95–104. doi: 10.1016/j.neuroscience.2004.03.031. [DOI] [PubMed] [Google Scholar]
  191. Meme W, Calvo CF, Froger N, Ezan P, Amigou E, Koulakoff A, Giaume C. Proinflammatory cytokines released from microglia inhibit gap junctions in astrocytes: potentiation by beta-amyloid. FASEB J. 2006;20:494–496. doi: 10.1096/fj.05-4297fje. [DOI] [PubMed] [Google Scholar]
  192. Mendoza-Naranjo A, Saez PJ, Johansson CC, Ramirez M, Mandakovic D, Pereda C, Lopez MN, Kiessling R, Saez JC, Salazar-Onfray F. Functional gap junctions facilitate melanoma antigen transfer and cross-presentation between human dendritic cells. J Immunol. 2007;178:6949–6957. doi: 10.4049/jimmunol.178.11.6949. [DOI] [PubMed] [Google Scholar]
  193. Mendoza-Naranjo A, Bouma G, Pereda C, Ramirez M, Webb KF, Tittarelli A, Lopez MN, Kalergis AM, Thrasher AJ, Becker DL, Salazar-Onfray F. Functional gap junctions accumulate at the immunological synapse and contribute to T cell activation. J Immunol. 2011 doi: 10.4049/jimmunol.1100378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. Menichella DM, Goodenough DA, Sirkowski E, Scherer SS, Paul DL. Connexins are critical for normal myelination in the CNS. J Neurosci. 2003;23:5963–5973. doi: 10.1523/JNEUROSCI.23-13-05963.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Mesnil M, Piccoli C, Tiraby G, Willecke K, Yamasaki H. Bystander killing of cancer cells by herpes simplex virus thymidine kinase gene is mediated by connexins. Proc Natl Acad Sci U S A. 1996;93:1831–1835. doi: 10.1073/pnas.93.5.1831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Migliore M, Hines ML, Shepherd GM. The role of distal dendritic gap junctions in synchronization of mitral cell axonal output. J Comput Neurosci. 2005;18:151–161. doi: 10.1007/s10827-005-6556-1. [DOI] [PubMed] [Google Scholar]
  197. Moldrich RX, Chapman AG, De Sarro G, Meldrum BS. Glutamate metabotropic receptors as targets for drug therapy in epilepsy. Eur J Pharmacol. 2003;476:3–16. doi: 10.1016/s0014-2999(03)02149-6. [DOI] [PubMed] [Google Scholar]
  198. Moorby C, Patel M. Dual functions for connexins: Cx43 regulates growth independently of gap junction formation. Exp Cell Res. 2001;271:238–248. doi: 10.1006/excr.2001.5357. [DOI] [PubMed] [Google Scholar]
  199. Moortgat KT, Bullock TH, Sejnowski TJ. Gap junction effects on precision and frequency of a model pacemaker network. J Neurophysiol. 2000;83:984–997. doi: 10.1152/jn.2000.83.2.984. [DOI] [PubMed] [Google Scholar]
  200. Morley GE, Taffet SM, Delmar M. Intramolecular interactions mediate pH regulation of connexin43 channels. Biophys J. 1996;70:1294–1302. doi: 10.1016/S0006-3495(96)79686-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  201. Morley GE, Ek-Vitorin JF, Taffet SM, Delmar M. Structure of connexin43 and its regulation by pHi. J Cardiovasc Electrophysiol. 1997;8:939–951. doi: 10.1111/j.1540-8167.1997.tb00856.x. [DOI] [PubMed] [Google Scholar]
  202. Muller T, Moller T, Neuhaus J, Kettenmann H. Electrical coupling among Bergmann glial cells and its modulation by glutamate receptor activation. Glia. 1996;17:274–284. doi: 10.1002/(SICI)1098-1136(199608)17:4<274::AID-GLIA2>3.0.CO;2-#. [DOI] [PubMed] [Google Scholar]
  203. Muller LP, Dedek K, Janssen-Bienhold U, Meyer A, Kreuzberg MM, Lorenz S, Willecke K, Weiler R. Expression and modulation of connexin 30.2, a novel gap junction protein in the mouse retina. Vis Neurosci. 2010;27:91–101. doi: 10.1017/S0952523810000131. [DOI] [PubMed] [Google Scholar]
  204. Musee J, Mbuy GN, Woodruff RI. Antiviral agents alter ability of HSV-2 to disrupt gap junctional intercellular communication between mammalian cells in vitro. Antiviral Res. 2002;56:143–151. doi: 10.1016/s0166-3542(02)00106-7. [DOI] [PubMed] [Google Scholar]
  205. Nagy JI, Rash JE. Connexins and gap junctions of astrocytes and oligodendrocytes in the CNS. Brain Res Brain Res Rev. 2000;32:29–44. doi: 10.1016/s0165-0173(99)00066-1. [DOI] [PubMed] [Google Scholar]
  206. Nagy JI, Li W, Hertzberg EL, Marotta CA. Elevated connexin43 immunoreactivity at sites of amyloid plaques in Alzheimer’s disease. Brain Res. 1996;717:173–178. doi: 10.1016/0006-8993(95)01526-4. [DOI] [PubMed] [Google Scholar]
  207. Nagy JI, Ionescu AV, Lynn BD, Rash JE. Connexin29 and connexin32 at oligodendrocyte and astrocyte gap junctions and in myelin of the mouse central nervous system. J Comp Neurol. 2003a;464:356–370. doi: 10.1002/cne.10797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Nagy JI, Ionescu AV, Lynn BD, Rash JE. Coupling of astrocyte connexins Cx26, Cx30, Cx43 to oligodendrocyte Cx29, Cx32, Cx47: Implications from normal and connexin32 knockout mice. Glia. 2003b;44:205–218. doi: 10.1002/glia.10278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Nakase T, Naus CC. Gap junctions and neurological disorders of the central nervous system. Biochim Biophys Acta. 2004;1662:149–158. doi: 10.1016/j.bbamem.2004.01.009. [DOI] [PubMed] [Google Scholar]
  210. Nakase T, Fushiki S, Naus CC. Astrocytic gap junctions composed of connexin 43 reduce apoptotic neuronal damage in cerebral ischemia. Stroke. 2003;34:1987–1993. doi: 10.1161/01.STR.0000079814.72027.34. [DOI] [PubMed] [Google Scholar]
  211. Nakase T, Sohl G, Theis M, Willecke K, Naus CC. Increased apoptosis and inflammation after focal brain ischemia in mice lacking connexin43 in astrocytes. Am J Pathol. 2004;164:2067–2075. doi: 10.1016/S0002-9440(10)63765-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Nakase T, Yoshida Y, Nagata K. Enhanced connexin 43 immunoreactivity in penumbral areas in the human brain following ischemia. Glia. 2006;54:369–375. doi: 10.1002/glia.20399. [DOI] [PubMed] [Google Scholar]
  213. Namba H, Iwadate Y, Kawamura K, Sakiyama S, Tagawa M. Efficacy of the bystander effect in the herpes simplex virus thymidine kinase-mediated gene therapy is influenced by the expression of connexin43 in the target cells. Cancer Gene Ther. 2001;8:414–420. doi: 10.1038/sj.cgt.7700317. [DOI] [PubMed] [Google Scholar]
  214. Naus CC, Bechberger JF, Paul DL. Gap junction gene expression in human seizure disorder. Exp Neurol. 1991;111:198–203. doi: 10.1016/0014-4886(91)90007-y. [DOI] [PubMed] [Google Scholar]
  215. Neijssen J, Herberts C, Drijfhout JW, Reits E, Janssen L, Neefjes J. Cross-presentation by intercellular peptide transfer through gap junctions. Nature. 2005;434:83–88. doi: 10.1038/nature03290. [DOI] [PubMed] [Google Scholar]
  216. Nelles E, Butzler C, Jung D, Temme A, Gabriel HD, Dahl U, Traub O, Stumpel F, Jungermann K, Zielasek J, Toyka KV, Dermietzel R, Willecke K. Defective propagation of signals generated by sympathetic nerve stimulation in the liver of connexin32-deficient mice. Proc Natl Acad Sci U S A. 1996;93:9565–9570. doi: 10.1073/pnas.93.18.9565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Nilsen KE, Kelso AR, Cock HR. Antiepileptic effect of gap-junction blockers in a rat model of refractory focal cortical epilepsy. Epilepsia. 2006;47:1169–1175. doi: 10.1111/j.1528-1167.2006.00540.x. [DOI] [PubMed] [Google Scholar]
  218. Odermatt B, Wellershaus K, Wallraff A, Seifert G, Degen J, Euwens C, Fuss B, Bussow H, Schilling K, Steinhauser C, Willecke K. Connexin 47 (Cx47)-deficient mice with enhanced green fluorescent protein reporter gene reveal predominant oligodendrocytic expression of Cx47 and display vacuolized myelin in the CNS. J Neurosci. 2003;23:4549–4559. doi: 10.1523/JNEUROSCI.23-11-04549.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Ohagen A, Ghosh S, He J, Huang K, Chen Y, Yuan M, Osathanondh R, Gartner S, Shi B, Shaw G, Gabuzda D. Apoptosis induced by infection of primary brain cultures with diverse human immunodeficiency virus type 1 isolates: evidence for a role of the envelope. J Virol. 1999;73:897–906. doi: 10.1128/jvi.73.2.897-906.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Oliveira R, Christov C, Guillamo JS, de Bouard S, Palfi S, Venance L, Tardy M, Peschanski M. Contribution of gap junctional communication between tumor cells and astroglia to the invasion of the brain parenchyma by human glioblastomas. BMC Cell Biol. 2005;6:7. doi: 10.1186/1471-2121-6-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Omori Y, Zaidan Dagli ML, Yamakage K, Yamasaki H. Involvement of gap junctions in tumor suppression: analysis of genetically-manipulated mice. Mutat Res. 2001;477:191–196. doi: 10.1016/s0027-5107(01)00120-8. [DOI] [PubMed] [Google Scholar]
  222. Orellana JA, Hernandez DE, Ezan P, Velarde V, Bennett MV, Giaume C, Saez JC. Hypoxia in high glucose followed by reoxygenation in normal glucose reduces the viability of cortical astrocytes through increased permeability of connexin 43 hemichannels. Glia. 2010;58:329–343. doi: 10.1002/glia.20926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Orellana JA, Froger N, Ezan P, Jiang JX, Bennett MV, Naus CC, Giaume C, Saez JC. ATP and glutamate released via astroglial connexin 43 hemichannels mediate neuronal death through activation of pannexin 1 hemichannels. J Neurochem. 2011a doi: 10.1111/j.1471-4159.2011.07210.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Orellana JA, Shoji KF, Abudara V, Ezan P, Amigou E, Sáez PJ, Jiang JX, Naus CC, Sáez JC, Giaume C. Amyloid β-induced death in neurons involves glial and neuronal hemichannels. J Neurosci. 2011b;31:4962–4977. doi: 10.1523/JNEUROSCI.6417-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Orkand RK, Nicholls JG, Kuffler SW. Effect of nerve impulses on the membrane potential of glial cells in the central nervous system of amphibia. J Neurophysiol. 1966;29:788–806. doi: 10.1152/jn.1966.29.4.788. [DOI] [PubMed] [Google Scholar]
  226. Orthmann-Murphy JL, Abrams CK, Scherer SS. Gap junctions couple astrocytes and oligodendrocytes. J Mol Neurosci. 2008;35:101–116. doi: 10.1007/s12031-007-9027-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Paemeleire K, Leybaert L. Ionic changes accompanying astrocytic intercellular calcium waves triggered by mechanical cell damaging stimulation. Brain Res. 2000;857:235–245. doi: 10.1016/s0006-8993(99)02436-1. [DOI] [PubMed] [Google Scholar]
  228. Palmer CA, Geyer JD, Keating JM, Gilliam F, Kuzniecky RI, Morawetz RB, Bebin EM. Rasmussen’s encephalitis with concomitant cortical dysplasia: the role of GluR3. Epilepsia. 1999;40:242–247. doi: 10.1111/j.1528-1157.1999.tb02082.x. [DOI] [PubMed] [Google Scholar]
  229. Palop JJ, Mucke L. Amyloid-beta-induced neuronal dysfunction in Alzheimer’s disease: from synapses toward neural networks. Nat Neurosci. 2010;13:812–818. doi: 10.1038/nn.2583. [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Pang B, Neijssen J, Qiao X, Janssen L, Janssen H, Lippuner C, Neefjes J. Direct antigen presentation and gap junction mediated cross-presentation during apoptosis. J Immunol. 2009;183:1083–1090. doi: 10.4049/jimmunol.0900861. [DOI] [PubMed] [Google Scholar]
  231. Parenti R, Campisi A, Vanella A, Cicirata F. Immunocytochemical and RT-PCR analysis of connexin36 in cultures of mammalian glial cells. Arch Ital Biol. 2002;140:101–108. [PubMed] [Google Scholar]
  232. Parenti R, Cicirata F, Zappala A, Catania A, La Delia F, Cicirata V, Tress O, Willecke K. Dynamic expression of Cx47 in mouse brain development and in the cuprizone model of myelin plasticity. Glia. 2010;58:1594–1609. doi: 10.1002/glia.21032. [DOI] [PubMed] [Google Scholar]
  233. Parihar MS, Brewer GJ. Amyloid-beta as a modulator of synaptic plasticity. J Alzheimers Dis. 2010;22:741–763. doi: 10.3233/JAD-2010-101020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Parpura V, Scemes E, Spray DC. Mechanisms of glutamate release from astrocytes: gap junction “hemichannels”, purinergic receptors and exocytotic release. Neurochem Int. 2004;45:259–264. doi: 10.1016/j.neuint.2003.12.011. [DOI] [PubMed] [Google Scholar]
  235. Pastor A, Kremer M, Moller T, Kettenmann H, Dermietzel R. Dye coupling between spinal cord oligodendrocytes: differences in coupling efficiency between gray and white matter. Glia. 1998;24:108–120. [PubMed] [Google Scholar]
  236. Peinado A, Yuste R, Katz LC. Extensive dye coupling between rat neocortical neurons during the period of circuit formation. Neuron. 1993a;10:103–114. doi: 10.1016/0896-6273(93)90246-n. [DOI] [PubMed] [Google Scholar]
  237. Peinado A, Yuste R, Katz LC. Gap junctional communication and the development of local circuits in neocortex. Cereb Cortex. 1993b;3:488–498. doi: 10.1093/cercor/3.5.488. [DOI] [PubMed] [Google Scholar]
  238. Perez Velazquez JL, Carlen PL. Gap junctions, synchrony and seizures. Trends Neurosci. 2000;23:68–74. doi: 10.1016/s0166-2236(99)01497-6. [DOI] [PubMed] [Google Scholar]
  239. Perez Velazquez JL, Carlen PL, Skinner FK. Artificial electrotonic coupling affects neuronal firing patterns depending upon cellular characteristics. Neuroscience. 2001;103:841–849. doi: 10.1016/s0306-4522(01)00019-7. [DOI] [PubMed] [Google Scholar]
  240. Persidsky Y, Stins M, Way D, Witte MH, Weinand M, Kim KS, Bock P, Gendelman HE, Fiala M. A model for monocyte migration through the blood–brain barrier during HIV-1 encephalitis. J Immunol. 1997;158:3499–3510. [PubMed] [Google Scholar]
  241. Petrasch-Parwez E, Habbes HW, Weickert S, Lobbecke-Schumacher M, Striedinger K, Wieczorek S, Dermietzel R, Epplen JT. Fine-structural analysis and connexin expression in the retina of a transgenic model of Huntington’s disease. J Comp Neurol. 2004;479:181–197. doi: 10.1002/cne.20327. [DOI] [PubMed] [Google Scholar]
  242. Planells-Cases R, Jentsch TJ. Chloride channelopathies. Biochim Biophys Acta. 2009;1792:173–189. doi: 10.1016/j.bbadis.2009.02.002. [DOI] [PubMed] [Google Scholar]
  243. Pu P, Xia Z, Yu S, Huang Q. Altered expression of Cx43 in astrocytic tumors. Clin Neurol Neurosurg. 2004;107:49–54. doi: 10.1016/j.clineuro.2004.03.006. [DOI] [PubMed] [Google Scholar]
  244. Rash JE, Yasumura T, Dudek FE, Nagy JI. Cell-specific expression of connexins and evidence of restricted gap junctional coupling between glial cells and between neurons. J Neurosci. 2001a;21:1983–2000. doi: 10.1523/JNEUROSCI.21-06-01983.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  245. Rash JE, Yasumura T, Davidson KG, Furman CS, Dudek FE, Nagy JI. Identification of cells expressing Cx43, Cx30, Cx26, Cx32 and Cx36 in gap junctions of rat brain and spinal cord. Cell Commun Adhes. 2001b;8:315–320. doi: 10.3109/15419060109080745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Rash JE, Davidson KG, Kamasawa N, Yasumura T, Kamasawa M, Zhang C, Michaels R, Restrepo D, Ottersen OP, Olson CO, Nagy JI. Ultrastructural localization of connexins (Cx36, Cx43, Cx45), glutamate receptors and aquaporin-4 in rodent olfactory mucosa, olfactory nerve and olfactory bulb. J Neurocytol. 2005;34:307–341. doi: 10.1007/s11068-005-8360-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  247. Retamal MA, Cortes CJ, Reuss L, Bennett MV, Saez JC. S-nitrosylation and permeation through connexin 43 hemichannels in astrocytes: induction by oxidant stress and reversal by reducing agents. Proc Natl Acad Sci U S A. 2006;103:4475–4480. doi: 10.1073/pnas.0511118103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Retamal MA, Froger N, Palacios-Prado N, Ezan P, Saez PJ, Saez JC, Giaume C. Cx43 hemichannels and gap junction channels in astrocytes are regulated oppositely by proinflammatory cytokines released from activated microglia. J Neurosci. 2007;27:13781–13792. doi: 10.1523/JNEUROSCI.2042-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Retamal MA, Yin S, Altenberg GA, Reuss L. Voltage-dependent facilitation of Cx46 hemichannels. Am J Physiol Cell Physiol. 2010;298:C132–C139. doi: 10.1152/ajpcell.00258.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Reuss B, Dermietzel R, Unsicker K. Fibroblast growth factor 2 (FGF-2) differentially regulates connexin (cx) 43 expression and function in astroglial cells from distinct brain regions. Glia. 1998;22:19–30. [PubMed] [Google Scholar]
  251. Robe PA, Rogister B, Merville MP, Bours V. Growth regulation of astrocytes and C6 cells by TGFbeta1: correlation with gap junctions. Neuroreport. 2000;11:2837–2841. doi: 10.1097/00001756-200009110-00003. [DOI] [PubMed] [Google Scholar]
  252. Roberts TK, Buckner CM, Berman JW. Leukocyte transmigration across the blood–brain barrier: perspectives on neuroAIDS. Front Biosci. 2010;15:478–536. doi: 10.2741/3631. [DOI] [PubMed] [Google Scholar]
  253. Rose CR, Ransom BR. Regulation of intracellular sodium in cultured rat hippocampal neurones. J Physiol. 1997;499(Pt 3):573–587. doi: 10.1113/jphysiol.1997.sp021951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  254. Rouach N, Tence M, Glowinski J, Giaume C. Costimulation of N-methyl-D-aspartate and muscarinic neuronal receptors modulates gap junctional communication in striatal astrocytes. Proc Natl Acad Sci U S A. 2002a;99:1023–1028. doi: 10.1073/pnas.022257499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Rouach N, Calvo CF, Glowinski J, Giaume C. Brain macrophages inhibit gap junctional communication and downregulate connexin 43 expression in cultured astrocytes. Eur J Neurosci. 2002b;15:403–407. doi: 10.1046/j.0953-816x.2001.01868.x. [DOI] [PubMed] [Google Scholar]
  256. Rouach N, Avignone E, Meme W, Koulakoff A, Venance L, Blomstrand F, Giaume C. Gap junctions and connexin expression in the normal and pathological central nervous system. Biol Cell. 2002c;94:457–475. doi: 10.1016/s0248-4900(02)00016-3. [DOI] [PubMed] [Google Scholar]
  257. Rouach N, Koulakoff A, Giaume C. Neurons set the tone of gap junctional communication in astrocytic networks. Neurochem Int. 2004a;45:265–272. doi: 10.1016/j.neuint.2003.07.004. [DOI] [PubMed] [Google Scholar]
  258. Rouach N, Calvo CF, Duquennoy H, Glowinski J, Giaume C. Hydrogen peroxide increases gap junctional communication and induces astrocyte toxicity: regulation by brain macrophages. Glia. 2004b;45:28–38. doi: 10.1002/glia.10300. [DOI] [PubMed] [Google Scholar]
  259. Rouach N, Koulakoff A, Abudara V, Willecke K, Giaume C. Astroglial metabolic networks sustain hippocampal synaptic transmission. Science. 2008;322:1551–1555. doi: 10.1126/science.1164022. [DOI] [PubMed] [Google Scholar]
  260. Rozental R, Morales M, Mehler MF, Urban M, Kremer M, Dermietzel R, Kessler JA, Spray DC. Changes in the properties of gap junctions during neuronal differentiation of hippocampal progenitor cells. J Neurosci. 1998;18:1753–1762. doi: 10.1523/JNEUROSCI.18-05-01753.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  261. Rufer M, Wirth SB, Hofer A, Dermietzel R, Pastor A, Kettenmann H, Unsicker K. Regulation of connexin-43, GFAP, and FGF-2 is not accompanied by changes in astroglial coupling in MPTP-lesioned, FGF-2-treated parkinsonian mice. J Neurosci Res. 1996;46:606–617. doi: 10.1002/(SICI)1097-4547(19961201)46:5<606::AID-JNR9>3.0.CO;2-N. [DOI] [PubMed] [Google Scholar]
  262. Sáez JC, Contreras JE, Bukauskas FF, Retamal MA, Bennett MV. Gap junction hemichannels in astrocytes of the CNS. Acta Physiol Scand. 2003a;179:9–22. doi: 10.1046/j.1365-201X.2003.01196.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Sáez JC, Berthoud VM, Branes MC, Martinez AD, Beyer EC. Plasma membrane channels formed by connexins: their regulation and functions. Physiol Rev. 2003b;83:1359–1400. doi: 10.1152/physrev.00007.2003. [DOI] [PubMed] [Google Scholar]
  264. Sáez JC, Retamal MA, Basilio D, Bukauskas FF, Bennett MV. Connexin-based gap junction hemichannels: gating mechanisms. Biochim Biophys Acta. 2005;1711:215–224. doi: 10.1016/j.bbamem.2005.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Sáez JC, Schalper KA, Retamal MA, Orellana JA, Shoji KF, Bennett MV. Cell membrane permeabilization via connexin hemichannels in living and dying cells. Exp Cell Res. 2010;316:2377–2389. doi: 10.1016/j.yexcr.2010.05.026. [DOI] [PubMed] [Google Scholar]
  266. Samoilova M, Li J, Pelletier MR, Wentlandt K, Adamchik Y, Naus CC, Carlen PL. Epileptiform activity in hippocampal slice cultures exposed chronically to bicuculline: increased gap junctional function and expression. J Neurochem. 2003;86:687–699. doi: 10.1046/j.1471-4159.2003.01893.x. [DOI] [PubMed] [Google Scholar]
  267. Samoilova M, Wentlandt K, Adamchik Y, Velumian AA, Carlen PL. Connexin 43 mimetic peptides inhibit spontaneous epileptiform activity in organotypic hippocampal slice cultures. Exp Neurol. 2008;210:762–775. doi: 10.1016/j.expneurol.2008.01.005. [DOI] [PubMed] [Google Scholar]
  268. Sánchez HA, Orellana JA, Verselis VK, Saez JC. Metabolic inhibition increases activity of connexin-32 hemichannels permeable to Ca2+ in transfected HeLa cells. Am J Physiol Cell Physiol. 2009;297:C665–C678. doi: 10.1152/ajpcell.00200.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Sanchez-Alvarez R, Paino T, Herrero-Gonzalez S, Medina JM, Tabernero A. Tolbutamide reduces glioma cell proliferation by increasing connexin43, which promotes the up-regulation of p21 and p27 and subsequent changes in retinoblastoma phosphorylation. Glia. 2006;54:125–134. doi: 10.1002/glia.20363. [DOI] [PubMed] [Google Scholar]
  270. Santiago MF, Alcami P, Striedinger KM, Spray DC, Scemes E. The carboxyl-terminal domain of connexin43 is a negative modulator of neuronal differentiation. J Biol Chem. 2010;285:11836–11845. doi: 10.1074/jbc.M109.058750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Sargiannidou I, Ahn M, Enriquez AD, Peinado A, Reynolds R, Abrams C, Scherer SS, Kleopa KA. Human oligodendrocytes express Cx31.3: function and interactions with Cx32 mutants. Neurobiol Dis. 2008;30:221–233. doi: 10.1016/j.nbd.2008.01.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Sargiannidou I, Vavlitou N, Aristodemou S, Hadjisavvas A, Kyriacou K, Scherer SS, Kleopa KA. Connexin32 mutations cause loss of function in Schwann cells and oligodendrocytes leading to PNS and CNS myelination defects. J Neurosci. 2009;29:4736–4749. doi: 10.1523/JNEUROSCI.0325-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Schalper KA, Orellana JA, Berthoud VM, Saez JC. Dysfunctions of the diffusional membrane pathways mediated by hemichannels in inherited and acquired human diseases. Curr Vasc Pharmacol. 2009;7:486–505. doi: 10.2174/157016109789043937. [DOI] [PubMed] [Google Scholar]
  274. Scherer SS, Xu YT, Nelles E, Fischbeck K, Willecke K, Bone LJ. Connexin32-null mice develop demyelinating peripheral neuropathy. Glia. 1998;24:8–20. doi: 10.1002/(sici)1098-1136(199809)24:1<8::aid-glia2>3.0.co;2-3. [DOI] [PubMed] [Google Scholar]
  275. Schubert T, Maxeiner S, Kruger O, Willecke K, Weiler R. Connexin45 mediates gap junctional coupling of bistratified ganglion cells in the mouse retina. J Comp Neurol. 2005;490:29–39. doi: 10.1002/cne.20621. [DOI] [PubMed] [Google Scholar]
  276. Schweighardt B, Atwood WJ. HIV type 1 infection of human astrocytes is restricted by inefficient viral entry. AIDS Res Hum Retroviruses. 2001;17:1133–1142. doi: 10.1089/088922201316912745. [DOI] [PubMed] [Google Scholar]
  277. Seror C, et al. Extracellular ATP acts on P2Y2 purinergic receptors to facilitate HIV-1 infection. J Exp Med. 2011 doi: 10.1084/jem.20101805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Shinoura N, Chen L, Wani MA, Kim YG, Larson JJ, Warnick RE, Simon M, Menon AG, Bi WL, Stambrook PJ. Protein and messenger RNA expression of connexin43 in astrocytomas: implications in brain tumor gene therapy. J Neurosurg. 1996;84:839–845. doi: 10.3171/jns.1996.84.5.0839. discussion 846. [DOI] [PubMed] [Google Scholar]
  279. Simard M, Arcuino G, Takano T, Liu QS, Nedergaard M. Signaling at the gliovascular interface. J Neurosci. 2003;23:9254–9262. doi: 10.1523/JNEUROSCI.23-27-09254.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Siushansian R, Bechberger JF, Cechetto DF, Hachinski VC, Naus CC. Connexin43 null mutation increases infarct size after stroke. J Comp Neurol. 2001;440:387–394. doi: 10.1002/cne.1392. [DOI] [PubMed] [Google Scholar]
  281. Soffer D, Raine CS. Morphologic analysis of axo-glial membrane specializations in the demyelinated central nervous system. Brain Res. 1980;186:301–313. doi: 10.1016/0006-8993(80)90977-4. [DOI] [PubMed] [Google Scholar]
  282. Sohl G, Guldenagel M, Beck H, Teubner B, Traub O, Gutierrez R, Heinemann U, Willecke K. Expression of connexin genes in hippocampus of kainate-treated and kindled rats under conditions of experimental epilepsy. Brain Res Mol Brain Res. 2000;83:44–51. doi: 10.1016/s0169-328x(00)00195-9. [DOI] [PubMed] [Google Scholar]
  283. Sohl G, Maxeiner S, Willecke K. Expression and functions of neuronal gap junctions. Nat Rev Neurosci. 2005;6:191–200. doi: 10.1038/nrn1627. [DOI] [PubMed] [Google Scholar]
  284. Soroceanu L, Manning TJ, Jr, Sontheimer H. Reduced expression of connexin-43 and functional gap junction coupling in human gliomas. Glia. 2001;33:107–117. doi: 10.1002/1098-1136(200102)33:2<107::aid-glia1010>3.0.co;2-4. [DOI] [PubMed] [Google Scholar]
  285. Sotelo C, Angaut P. The fine structure of the cerebellar central nuclei in the cat. I. Neurons and neuroglial cells. Exp Brain Res. 1973;16:410–430. doi: 10.1007/BF00233432. [DOI] [PubMed] [Google Scholar]
  286. Stout CE, Costantin JL, Naus CC, Charles AC. Intercellular calcium signaling in astrocytes via ATP release through connexin hemichannels. J Biol Chem. 2002;277:10482–10488. doi: 10.1074/jbc.M109902200. [DOI] [PubMed] [Google Scholar]
  287. Sun Y, Tang W, Chang Q, Wang Y, Kong W, Lin X. Connexin30 null and conditional connexin26 null mice display distinct pattern and time course of cellular degeneration in the cochlea. J Comp Neurol. 2009;516:569–579. doi: 10.1002/cne.22117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  288. Sutor B, Schmolke C, Teubner B, Schirmer C, Willecke K. Myelination defects and neuronal hyperexcitability in the neocortex of connexin 32-deficient mice. Cereb Cortex. 2000;10:684–697. doi: 10.1093/cercor/10.7.684. [DOI] [PubMed] [Google Scholar]
  289. Swayne LA, Sorbara CD, Bennett SA. Pannexin 2 is expressed by postnatal hippocampal neural progenitors and modulates neuronal commitment. J Biol Chem. 2010;285:24977–24986. doi: 10.1074/jbc.M110.130054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  290. Tabernero A, Sanchez-Alvarez R, Medina JM. Increased levels of cyclins D1 and D3 after inhibition of gap junctional communication in astrocytes. J Neurochem. 2006;96:973–982. doi: 10.1111/j.1471-4159.2005.03623.x. [DOI] [PubMed] [Google Scholar]
  291. Theis M, Jauch R, Zhuo L, Speidel D, Wallraff A, Doring B, Frisch C, Sohl G, Teubner B, Euwens C, Huston J, Steinhauser C, Messing A, Heinemann U, Willecke K. Accelerated hippocampal spreading depression and enhanced locomotory activity in mice with astrocyte-directed inactivation of connexin43. J Neurosci. 2003;23:766–776. doi: 10.1523/JNEUROSCI.23-03-00766.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  292. Tontsch U, Bauer HC. Glial cells and neurons induce blood–brain barrier related enzymes in cultured cerebral endothelial cells. Brain Res. 1991;539:247–253. doi: 10.1016/0006-8993(91)91628-e. [DOI] [PubMed] [Google Scholar]
  293. Tornatore C, Nath A, Amemiya K, Major EO. Persistent human immunodeficiency virus type 1 infection in human fetal glial cells reactivated by T-cell factor(s) or by the cytokines tumor necrosis factor alpha and interleukin-1 beta. J Virol. 1991;65:6094–6100. doi: 10.1128/jvi.65.11.6094-6100.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Tornatore C, Chandra R, Berger JR, Major EO. HIV-1 infection of subcortical astrocytes in the pediatric central nervous system. Neurology. 1994a;44:481–487. doi: 10.1212/wnl.44.3_part_1.481. [DOI] [PubMed] [Google Scholar]
  295. Tornatore C, Meyers K, Atwood W, Conant K, Major E. Temporal patterns of human immunodeficiency virus type 1 transcripts in human fetal astrocytes. J Virol. 1994b;68:93–102. doi: 10.1128/jvi.68.1.93-102.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Tran Van Nhieu G, Clair C, Bruzzone R, Mesnil M, Sansonetti P, Combettes L. Connexin-dependent inter-cellular communication increases invasion and dissemination of Shigella in epithelial cells. Nat Cell Biol. 2003;5:720–726. doi: 10.1038/ncb1021. [DOI] [PubMed] [Google Scholar]
  297. Tress O, Maglione M, Zlomuzica A, May D, Dicke N, Degen J, Dere E, Kettenmann H, Hartmann D, Willecke K. Pathologic and phenotypic alterations in a mouse expressing a connexin47 missense mutation that causes pelizaeus-merzbacher-like disease in humans. PLoS Genet. 2011;7:e1002146. doi: 10.1371/journal.pgen.1002146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. Trexler EB, Bukauskas FF, Bennett MV, Bargiello TA, Verselis VK. Rapid and direct effects of pH on connexins revealed by the connexin46 hemichannel preparation. J Gen Physiol. 1999;113:721–742. doi: 10.1085/jgp.113.5.721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  299. Trosko JE, Chang CC. Isolation and characterization of normal adult human epithelial pluripotent stem cells. Oncol Res. 2003;13:353–357. doi: 10.3727/096504003108748366. [DOI] [PubMed] [Google Scholar]
  300. Ure JA, Perassolo M. Update on the pathophysiology of the epilepsies. J Neurol Sci. 2000;177:1–17. doi: 10.1016/s0022-510x(00)00356-7. [DOI] [PubMed] [Google Scholar]
  301. Ure J, Baudry M, Perassolo M. Metabotropic glutamate receptors and epilepsy. J Neurol Sci. 2006;247:1–9. doi: 10.1016/j.jns.2006.03.018. [DOI] [PubMed] [Google Scholar]
  302. Vaney DI. Retinal neurons: cell types and coupled networks. Prog Brain Res. 2002;136:239–254. doi: 10.1016/s0079-6123(02)36020-5. [DOI] [PubMed] [Google Scholar]
  303. Vaquero J, Oya S, Manrique M, Lozano AP, Bravo G. Cytological alterations in alumina cream experimental epilepsy. Acta Neurochir (Wien) 1978;42:235–243. doi: 10.1007/BF01405339. [DOI] [PubMed] [Google Scholar]
  304. Venance L, Rozov A, Blatow M, Burnashev N, Feldmeyer D, Monyer H. Connexin expression in electrically coupled postnatal rat brain neurons. Proc Natl Acad Sci U S A. 2000;97:10260–10265. doi: 10.1073/pnas.160037097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Venance L, Glowinski J, Giaume C. Electrical and chemical transmission between striatal GABAergic output neurones in rat brain slices. J Physiol. 2004;559:215–230. doi: 10.1113/jphysiol.2004.065672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Vis JC, Nicholson LF, Faull RL, Evans WH, Severs NJ, Green CR. Connexin expression in Huntington’s diseased human brain. Cell Biol Int. 1998;22:837–847. doi: 10.1006/cbir.1998.0388. [DOI] [PubMed] [Google Scholar]
  307. Wallraff A, Kohling R, Heinemann U, Theis M, Willecke K, Steinhauser C. The impact of astrocytic gap junctional coupling on potassium buffering in the hippocampus. J Neurosci. 2006;26:5438–5447. doi: 10.1523/JNEUROSCI.0037-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Weiss JM, Downie SA, Lyman WD, Berman JW. Astrocyte-derived monocyte-chemoattractant protein-1 directs the transmigration of leukocytes across a model of the human blood–brain barrier. J Immunol. 1998;161:6896–6903. [PubMed] [Google Scholar]
  309. Wen CM, Cheng YH, Huang YF, Wang CS. Isolation and characterization of a neural progenitor cell line from tilapia brain. Comp Biochem Physiol A Mol Integr Physiol. 2008;149:167–180. doi: 10.1016/j.cbpa.2007.11.005. [DOI] [PubMed] [Google Scholar]
  310. Xu HL, Mao L, Ye S, Paisansathan C, Vetri F, Pelligrino DA. Astrocytes are a key conduit for upstream signaling of vasodilation during cerebral cortical neuronal activation in vivo. Am J Physiol Heart Circ Physiol. 2008;294:H622–H632. doi: 10.1152/ajpheart.00530.2007. [DOI] [PubMed] [Google Scholar]
  311. Yamaguchi DT, Ma D. Mechanism of pH regulation of connexin 43 expression in MC3T3-E1 cells. Biochem Biophys Res Commun. 2003;304:736–739. doi: 10.1016/s0006-291x(03)00633-8. [DOI] [PubMed] [Google Scholar]
  312. Yamasaki H, Omori Y, Krutovskikh V, Zhu W, Mironov N, Yamakage K, Mesnil M. Connexins in tumour suppression and cancer therapy. Novartis Found Symp. 1999;219:241–254. doi: 10.1002/9780470515587.ch15. discussion 254–260. [DOI] [PubMed] [Google Scholar]
  313. Yao LF, Wang ZK, Wang ZG, Sui D, Zhang LM. Expression and function of Cx32 and Cx43 junctions in medically intractable temporal lobe epilepsy in human. Zhonghua Yi Xue Za Zhi. 2009;89:3058–3060. [PubMed] [Google Scholar]
  314. Ye ZC, Wyeth MS, Baltan-Tekkok S, Ransom BR. Functional hemichannels in astrocytes: a novel mechanism of glutamate release. J Neurosci Off J Soc Neurosci. 2003;23:3588–3596. doi: 10.1523/JNEUROSCI.23-09-03588.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Zhang W, Couldwell WT, Simard MF, Song H, Lin JH, Nedergaard M. Direct gap junction communication between malignant glioma cells and astrocytes. Cancer Res. 1999;59:1994–2003. [PubMed] [Google Scholar]
  316. Zhang W, Nwagwu C, Le DM, Yong VW, Song H, Couldwell WT. Increased invasive capacity of connexin43-overexpressing malignant glioma cells. J Neurosurg. 2003a;99:1039–1046. doi: 10.3171/jns.2003.99.6.1039. [DOI] [PubMed] [Google Scholar]
  317. Zhang YW, Nakayama K, Morita I. A novel route for connexin 43 to inhibit cell proliferation: negative regulation of S-phase kinase-associated protein (Skp 2) Cancer Res. 2003b;63:1623–1630. [PubMed] [Google Scholar]
  318. Zhang YW, Kaneda M, Morita I. The gap junction-independent tumor-suppressing effect of connexin 43. J Biol Chem. 2003c;278:44852–44856. doi: 10.1074/jbc.M305072200. [DOI] [PubMed] [Google Scholar]
  319. Zhang S, Liang R, Zhou F, Huang X, Ding JH, Hu G. Reversal of rotenone-induced dysfunction of astrocytic connexin43 by opening mitochondrial ATP-sensitive potassium channels. Cell Mol Neurobiol. 2010 doi: 10.1007/s10571-010-9560-6. [DOI] [PubMed] [Google Scholar]
  320. Zhuo L, Theis M, Alvarez-Maya I, Brenner M, Willecke K, Messing A. hGFAP-cre transgenic mice for manipulation of glial and neuronal function in vivo. Genesis. 2001;31:85–94. doi: 10.1002/gene.10008. [DOI] [PubMed] [Google Scholar]
  321. Zlomuzica A, Reichinnek S, Maxeiner S, Both M, May E, Worsdorfer P, Draguhn A, Willecke K, Dere E. Deletion of connexin45 in mouse neurons disrupts one-trial object recognition and alters kainate-induced gamma-oscillations in the hippocampus. Physiol Behav. 2010;101:245–253. doi: 10.1016/j.physbeh.2010.05.007. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

1-3

RESOURCES