Skip to main content

MINI REVIEW article

Front. Cell. Neurosci., 19 September 2018
Sec. Cellular Neurophysiology
This article is part of the Research Topic New Insights on Neuron and Astrocyte Function from Cutting-Edge Optical Techniques View all 17 articles

Gap Junctions in the Nervous System: Probing Functional Connections Using New Imaging Approaches

  • 1State Key Laboratory of Membrane Biology, Peking University School of Life Sciences, Beijing, China
  • 2PKU-IDG/McGovern Institute for Brain Research, Beijing, China
  • 3Peking-Tsinghua Center for Life Sciences, Beijing, China

Gap junctions are channels that physically connect adjacent cells, mediating the rapid exchange of small molecules, and playing an essential role in a wide range of physiological processes in nearly every system in the body, including the nervous system. Thus, altered function of gap junctions has been linked with a plethora of diseases and pathological conditions. Being able to measure and characterize the distribution, function, and regulation of gap junctions in intact tissue is therefore essential for understanding the physiological and pathophysiological roles that gap junctions play. In recent decades, several robust in vitro and in vivo methods have been developed for detecting and characterizing gap junctions. Here, we review the currently available methods with respect to invasiveness, signal-to-noise ratio, temporal resolution and others, highlighting the recently developed chemical tracers and hybrid imaging systems that use novel chemical compounds and/or genetically encoded enzymes, transporters, channels, and fluorescent proteins in order to map gap junctions. Finally, we discuss possible avenues for further improving existing techniques in order to achieve highly sensitive, cell type-specific, non-invasive measures of in vivo gap junction function with high throughput and high spatiotemporal resolution.

Introduction

Multicellular organisms rely on cell-cell communication to coordinate a wide range of physiological processes and maintain homeostasis. Most organisms have evolved a rich diversity of mechanisms to achieve this communication, including long-distance signaling through the release, and binding of hormones (Ansar Ahmed et al., 1985; Giustina and Veldhuis, 1998; Meier and Gressner, 2004), spatially confined synaptic transmission between two neurons (Krnjevic, 1974; Pereda, 2014), and gap junctional coupling between neighboring cells (Kumar and Gilula, 1996; Sohl et al., 2005; Mese et al., 2007). In the central nervous system, billions of neurons are intermingled and communicate with each other through a specialized structure called the synapse, forming a complex signaling network. Although synapses are predominantly chemical in nature, with neurotransmitters released from the presynaptic terminal and sensed by the postsynaptic neuron via surface receptors, gap junction–based electrical synapses are also widely distributed, and play an essential role in regulating both the development and function of the nervous system (Pereda, 2014).

Gap junctions, composed of connexins in vertebrates and innexins in invertebrates, are intercellular channel complexes between connected cells (Kumar and Gilula, 1996; Phelan et al., 1998). Pannexins are vertebrate homologs to the innexins (Baranova et al., 2004), form hemi-channels connecting cytosol and extracellular space (D’hondt et al., 2009), and could mediate gap junctional connection in cells when overexpressed (Bruzzone et al., 2003; Vanden Abeele et al., 2006; Lai et al., 2007; Ishikawa et al., 2011), although their in vivo role in forming functional gap junction is unclear (Sosinsky et al., 2011). Ions and other small molecules with a molecular mass up to approximately 1 to 2 kDa can freely diffuse through gap junctions (Loewenstein, 1981; Kumar and Gilula, 1996; Neijssen et al., 2005). Thus, signals such as action potentials can propagate directly between gap junction–coupled neurons, resulting in virtually no delay in signal transmission (Furshpan and Potter, 1957;Bennett and Zukin, 2004); in contrast, signal transmission via a chemical synapse has a delay on the order of milliseconds (Katz and Miledi, 1965; Sabatini and Regehr, 1996). Gap junctions therefore allow organisms to respond extremely rapidly under certain conditions, for example in the escape reflex in crayfish (Antonsen and Edwards, 2003) and the retina’s response to visual stimuli in vertebrates (Bloomfield and Volgyi, 2009). Gap junctions are also expressed in glial cell types, including astrocytes (Wallraff et al., 2004), microglia (Garg et al., 2005), oligodendrocytes and Schwann cells (Nualart-Marti et al., 2013), and insect blood-brain barrier glial cells (Speder and Brand, 2014), which is essential for the buffering of ions and transmitters, inflammatory response, myelination and neural stem cell proliferation. Gap junctions also connect glia and neurons (Dobrenis et al., 2005; Meng et al., 2016). Given their highly varied and important roles, it is therefore not surprising that malfunctions in gap junctions can disrupt communication among neurons and glia, thus giving rise to a variety of diseases and neurological disorders, including hereditary deafness (Martinez et al., 2009), uncorrelated motor neuron firing (Personius et al., 2007), and Charcot-Marie-Tooth disease (Kleopa, 2011).

Extensive studies of gap junctions in the nervous system have been carried out by various research groups over the past few decades; the expression of gap junction forming subunits were detected by northern blot (Paul et al., 1991; White et al., 1992), RT-PCR (Wrenzycki et al., 1996; Xia et al., 1998), western blot (Stauffer, 1995; Giepmans and Moolenaar, 1998), and immunohistochemistry (Beyer et al., 1989; Dermietzel et al., 1989). In this review, we focus on functional methods that can detect gap junctional coupling, first briefly summarizing current approaches relying on electrophysiological recording, tracer-based assays, and hybrid methods using genetic tools (Figure 1), mainly focusing on recently developed imaging methods. We summarize the performance and properties of these methods, including their invasiveness, throughput, feasibility, sensitivity, spatial resolution, and temporal resolution (Table 1). As new in vivo methods are being developed, new features of gap junction regulation will likely be revealed, yielding important new insights into the role that gap junctions play in both health and disease.

FIGURE 1
www.frontiersin.org

FIGURE 1. Schematic overview of the currently available methods for studying gap junctions. The principle behind each method is shown schematically on the left with a radar graph on the right that summarizes its corresponding performance index (e.g., sensitivity, throughput, resolution) in arbitrary units ranging from 0 to 4. Further details are provided in the text. FRAP, fluorescence recovery after photobleaching; LAMP, local activation of a molecular fluorescent probe; PLE, porcine liver esterase; Pept2, peptide transporter 2; βALA, AMCA-labeled dipeptides β-Ala-Lys; ChIEF, an engineered version of a channelrhodopsin.

TABLE 1
www.frontiersin.org

TABLE 1. Overview of the methods used to probe gap junctional communication.

Electrophysiological Recording

Gap junctional coupling can be measured using dual-electrode whole-cell current-clamp recordings (Furshpan and Potter, 1959). This method requires two microelectrodes; one electrode is used to inject current into one cell, and the other electrode is used to measure the resulting change in membrane potential in a connected neighboring cell. Because the two cells are electrically coupled, current injection leads to a change in the membrane potential of both cells (Figure 1A). Alternatively, dual-electrode whole-cell voltage clamp can also be used to measure gap junctional coupling; in this configuration, inducing a change in membrane potential between the two cells drives an electrical current through the gap junctions (Spray et al., 1979). Electrophysiological recording has millisecond resolution, picoampere current detection sensitivity, and the ability to measure conductance and rectification, both of which are important properties of electrical synapses in neurons. However, this method has obvious limitations, including the need for relatively high technical expertise, specialized equipment, high invasiveness due to disruption of the cell membrane integrity, relatively low throughput, and one-off recording. Moreover, this method by itself cannot discriminate distinct cell types, which is particularly problematic given the heterogeneous nature of the nervous system. In addition, the recordings are usually performed at the cell body, which does not take into account the subcellular localization of gap junctions, particularly in neurons and other cell types with complex morphology.

Transfer of Tracers

To assay the gap junction communications, tracers including fluorescent dyes, and bioactive small molecules can be introduced to one cell or a group of cells. The diffusion of tracers from the primary targeted cells to other connected cells reflects the gap junctional couplings.

Microinjection of the Tracer

The injection of a tracer, followed by measuring its transfer, is usually the first step in identifying the location and morphology of cells within a tissue. Because small molecules can pass freely through gap junctions, the diffusion of an injected small tracer molecule can be used to measure gap junctional coupling between cells (Figure 1B). With respect to the study of gap junction–mediated communication, the most commonly used fluorescent dye is Lucifer Yellow, with a molecular weight of 457 Da (Stewart, 1978), and the most commonly used bioactive small molecule is biocytin, with a molecular weight of 372 (Horikawa and Armstrong, 1988). The transfer (i.e., diffusion) of an injected tracer to neighboring cells can be observed either directly (in the case of a fluorescent dye) or post hoc using immunohistochemistry (in the case of a bioactive small molecule). The tracers used in these experiments are not membrane-permeable, thereby reducing non-specific diffusion through the cell membrane. This method is technically easier to perform compared to electrophysiology, which requires multiple electrodes and a sophisticated recording setup. However, it still lacks cell type specificity, requires the microelectrode, and the dye diffusion is irreversible, thus preventing the ability to study gap junctions repeatedly in the same cells. Moreover, the injection process requires either mechanical pressure or iontophoresis, and immunohistochemistry takes a relatively long time, thus reducing both the temporal resolution and the throughput.

Scrape Loading of the Tracer

In addition to the one-by-one injection, the tracer can be introduced into a large population of cells via the scrape (McNeil et al., 1984). Cultured cells in one layer are incubated with gap junction-permeable but cell membrane impermeable dyes as mentioned above and are scraped by a needle or a scalpel. Dye molecules therefore get into wounded cells, and can further diffuse to adjacent cells that are intact but coupled with the scraped cells by gap junctions (Figure 1C; el-Fouly et al., 1987). The scrape loading/dye transfer method is the easiest one to implement among all methods discussed here. Because of its simplicity, gap junctional communications were evaluated using this method in many cell types, including fibroblasts (Azzam et al., 2001), germ cells in testis (Decrouy et al., 2004), and astrocytes (Giaume and McCarthy, 1996). The limitations of this method include that it is mostly effective in adherent cells and therefore mainly applied in cultured cells or tissue slices in vitro. The scraping procedure in conjunction of cell fixation protocol offers qualitative rather than quantitative data for gap junctional connections.

Fluorescence Redistribution/Recovery After Photobleaching (FRAP)

To overcome the high invasiveness and technical expertise associated with microelectrode-based methods, Wade et al. (1986) developed the gap-FRAP technique (Figure 1D), an all-optical strategy used to study gap junctions. Rather than injecting fluorescent molecules into individual cells, cultured cells are incubated with membrane-permeable fluorescein-AM; upon entering the cell, the ester bond is hydrolyzed by intracellular esterases, leaving the hydrophilic fluorescein molecule trapped within the cell (Rotman and Papermaster, 1966). After an intense focused laser is used to bleach the fluorescein in one cell, the bleached fluorescein molecules and the unbleached fluorescein molecules in neighboring cells diffuse through the gap junctions, leading to the recovery of fluorescence in the original bleached cell. Compared to the methods described above, FRAP is less invasive and easier to perform. Importantly, this method provides both qualitative and quantitative information regarding the strength of the gap junctions, which is reflected by the kinetics of fluorescence recovery (Lee et al., 1995; Soroceanu et al., 2001). This method also provides satisfactory temporal resolution, as the photobleaching can be performed extremely rapidly using a high-power laser (Lippincott-Schwartz et al., 2003). One potential drawback of FRAP is that the intense laser illumination may damage the cell. In addition, in order to quantitatively characterize the FRAP kinetics which reflect the strength of the gap junctional communication, the recovery event needs to be monitored until the fluorescence recover to the plateau, which takes much longer time than the half-time for recovery (Lippincott-Schwartz et al., 2003). This requirement makes FRAP not suitable for measuring very fast dynamics of the gap junctions as can be done by electrophysiological recording (Abbaci et al., 2008). Finally, similar to tracer tracking methods, FRAP by itself lacks cell type specificity, constraining its application mainly to homogenous cell cultures.

Local Activation of a Molecular Fluorescent Probe (LAMP)

To avoid potential phototoxicity associated with photobleaching while still leaving the cell intact, Dakin et al. (2005) developed LAMP, which uses the caged fluorescent dye NPE-HCCC2-AM (Figure 1E; Zhao et al., 2004). After the cell is loaded as described above for NPE-HCCC2-AM, UV illumination is used to uncage NPE-HCCC2 in specific cells and release HCCC2, which has a molecular weight of 450 Da and emits blue fluorescence. The uncaged HCCC2 can then diffuse to neighboring cells connected via gap junctions. In a sense, LAMP is a combination between FRAP and tracer tracking in that it generates a fluorescent signal (the “tracer”) in one cell and then tracks the movement of the tracer through gap junctions, while maintaining cell integrity. In this respect, LAMP has the combined advantages of both methods in that it is non-invasive, provides quantitative data, and has relatively high temporal resolution. In addition, LAMP allows for multicolor imaging, as the uncaging of NPE-HCCC2 requires a small pulse of UV light and is therefore compatible with fluorescent indicators in the visible spectrum (Dakin et al., 2005; Dakin and Li, 2006; Abbaci et al., 2008). This method can be improved further by incorporating caged fluorescent molecules with higher uncaging efficiency and a more penetrable red-shifted emission spectrum. Unfortunately, LAMP still requires loading of an exogenous dye, which limits its applications in in vivo preparations. Moreover, uncaging of NPE-HCCC2 is irreversible, making it less suitable for studying the dynamics of gap junctions repeatedly over a prolonged period of time.

Hybrid Approaches Combined with Genetic Tools

In order to obtain more cell type-specific information, genetically encoded proteins can be incorporated into the method being used to map gap junctions; this is particularly important for studying gap junctions in a specific cell population within the heterogeneous central nervous system.

Esterase-Ester Pair

This enzyme-substrate pair has been used successfully to map gap junctions (Figure 1F). The enzyme is expressed in specific cell types; the substrate, which is water-soluble, membrane-permeable, unaffected by endogenous enzymes but catalyzed by the ectopically expressed enzyme, is able to diffuse through gap junctions. One example of this approach is the highly selective esterase-ester pair developed by Tian et al. (2012). In this approach, they synthesized a series of esters that fluoresce upon hydrolysis, and identified one substrate (called “substrate 6”) that was stable in several different cell types in a range of species, including flies, rodents, and humans. They also identified porcine liver esterase (PLE) (Lange et al., 2001) as the most potent at catalyzing the hydrolysis of substrate 6 and used the PLE-substrate 6 pair to map the distribution of gap junctions. PLE hydrolyzes substrate 6 to produce a fluorescent product, and the diffusion of this fluorescent product causes fluorescence in the cells that were connected via gap junctions to the PLE-expressing cells (Figure 1F). Thanks to genetic manipulation, this strategy provides higher cell type specificity using a relatively simple approach. The bio-specificity of substrate 6 ensures that this strategy can be used in a wide variety of organisms and cell types; however, it is still possible that endogenous esterases can cause a non-specific background signal under certain conditions. Thus, the system can be optimized by modifying the PLE enzyme and/or the substrate, or by identifying a more bioorthogonal enzyme-substrate pair (Sletten and Bertozzi, 2009; Ritter et al., 2015). At the same time, robust control experiments (for example, using knockout models) are an essential step in testing for non-specific background due to endogenous enzymes (Qiao and Sanes, 2015). Another drawback of this method is the relatively low temporal resolution, which requires up to 30 min of incubation in the substrate, thereby limiting its value in terms of studying the dynamics and regulation of gap junctions (Tian et al., 2012). In addition, this method has only been tested in cultured cell lines, and its feasibility in primary cells (e.g., neurons and cardiomyocytes) and in vivo applications has not been investigated.

Transporter-Substrate Pairs

An alternative strategy is to use a transporter-substrate pair in which a genetically encoded transporter is expressed in one cell, which then transports a fluorescent substrate into the cytoplasm; diffusion of the fluorescent substrate to neighboring cells indicates the presence and distribution of gap junctions. In 2015, Qiao and Sanes reported the use of human Pept2 (a peptide transporter) (Biegel et al., 2006) and the AMCA-labeled dipeptide β-Ala-Lys (βALA, the substrate) (Dieck et al., 1999) as a strategy for mapping gap junctions (Figure 1G; Qiao and Sanes, 2015). Using this innovative tool, they successfully mapped functional gap junctions in cultured HEK293 cells and quantified the diffusion properties of βALA, which reflects the strength of the gap junctions. Taking advantage of the CreER system and sparse labeling in Pept2 knockout mice, they then confirmed the presence of gap junctional communication between J-RGCs (a subset of retinal ganglion cells) and amacrine cells in the mouse retina (Bloomfield and Volgyi, 2009; Hoshi and Mills, 2009; Volgyi et al., 2009), and they demonstrated the presence of gap junctions in horizontal cells. Importantly, the authors also characterized the light-dependent electrical coupling of horizontal cells by mapping the pattern of gap junctional communication before and after illumination with light (Xin and Bloomfield, 1999). Thus, the Pept2-βALA pair provides a powerful tool for mapping the distribution and strength of gap junction connectivity both in cultured cells and in an ex vivo retinal preparation. On the other hand, a clear drawback associated with this method is that the temporal resolution (which is on the order of hours) is not sufficient to track the dynamics of the strength of gap junction connections.

Genetically Encoded Fluorescent Sensors/Optogenetics Combined With Patch-Clamp Recording

Genetically encoded fluorescent sensors provide another means to map gap junctions by monitoring the concentration change of a chemical during diffusion through gap junctions. In 2016, Kang and Baker reported the development of a novel genetically encoded fluorescent sensor called Pado, which can be used to track the diffusion of protons through gap junctions (Figure 1H; Kang and Baker, 2016). Pado is a dual-function protein created by fusing an engineered voltage-gated proton channel from Clonorchis sinesis with a pH-sensitive fluorescent protein (Super Ecliptic pHluorin 227A, or SE227A) (Jin et al., 2012). To demonstrate proof-of-principle, Kang and Baker expressed Pado in HEK293 cells, then used the whole-cell patch-clamp technique to depolarize one cell. The change in voltage opened the voltage-gated proton channels, facilitating the efflux of protons from the cell and creating an electrochemical gradient between this cell and neighboring cells connected via gap junctions. Protons then diffused from the neighboring cells down this electrochemical gradient, and the change in SE227A fluorescence was detected in both the clamped cell and the adjacent cells. While this method is promising, the data should be taken carefully and some calibrations allowing for quantitative analysis should be performed. A similar strategy utilizing a hybrid calcium indicator Calcium Green FlAsH could also enable detection of gap junctional couplings, by monitoring the intercellular propagation of calcium waves in gap junction coupled cells (Tour et al., 2007). Given that Calcium Green FlAsH needs to be applied exogenously, further improvements can be achieved by using pure genetically encoded calcium indicators such as GCaMP6 (Chen et al., 2013).

Given the electrical properties of gap junctions, optogenetics is yet another useful tool for mapping gap junctions, as an electrical signal generated by light-activated channelrhodopsins (Nagel et al., 2003, 2005) can propagate to coupled cells and be detected using patch clamp. Recently, Wang et al. (2014) combined an improved version of the channelrhodopsin ChIEF (Lin et al., 2009) with electrophysiology in order to map gap junction connections in the Drosophila olfactory system (Figure 1I). They performed patch-clamp recordings on cholinergic projection neurons (mPNs) while expressing ChIEF in mediolateral antennocerebral tract projection neurons (mIPNs) labeled by Mz699-Gal4 (Ito et al., 1997). Applying blue laser illumination to the mIPNs induced depolarization of some mPNs; this effect was not altered by the nicotinic receptor antagonist mecamylamine but was sensitive to the shakB2 mutation (which affects innexin-8) (Thomas and Wyman, 1984; Phelan et al., 1996; Zhang et al., 1999; Song and Tanouye, 2006), leading to the conclusion that mPNs and mIPNs are electrically coupled. The finding that blocking cholinergic receptors had no effect on the mIPN-mPN coupling indicates that when using this tool, it is important to distinguish chemical and electrical synapses using genetics and/or pharmacology, as ChIEF induced depolarization of presynaptic neurons can drive postsynaptic responses in both chemical and electrical synapses. Moreover, unlike the dual-electrode whole-cell patch-clamp technique, the ChIEF-based method is unidirectional and cannot be used to identify rectifying gap junctions. To overcome this limitation, a light-gated chloride pump such as the Halorhodopsin isolated from Natronomonas (NpHR) (Han and Boyden, 2007; Zhang et al., 2007) can be used to hyperpolarize the presynaptic terminal, thereby reversing the direction of the current across the electrical synapse.

Compared to previous methods, these two strategies (exemplified by Pado and ChIEF) do not require an exogenously applied substrate, which simplifies the experimental protocol and makes them more feasible for use in in vivo applications. In addition, because they have relatively faster kinetics (on the order of milliseconds to seconds), these methods can be used to collect repeated measurements, which is essential for studying the dynamics of the strength of gap junctional connections at high temporal resolution. On the other hand, these approaches require the use of glass micropipettes, reducing their throughput. Moreover, one needs to block chemical synapses when using ChIEF to detect electrical synapses, which may alter the normal state of the nervous system.

Perspectives

Gap junctions play an extremely important role in mediating cell-cell communication, and their distribution and dynamics are essential for maintaining normal physiological function and homeostasis. Although researchers have been able to link genetic mutations with these conditions, identifying precisely which cell populations are affected by these mutations has been far more difficult. In a more physiological context, single-cell transcriptomics has revealed that both neurons and glia are more heterogeneous than previously believed (Lake et al., 2016; Tasic et al., 2016). In addition, connexins and innexins are encoded by multiple genes, giving rise to a wide diversity of gap junctions. For example, the mouse and human genomes contain 20 and 21 connexin-encoding genes, respectively (Sohl and Willecke, 2003), and the Caenorhabditis elegans and Drosophila melanogaster genomes contain 25 and 8 innexin-coding genes, respectively (Starich et al., 2001; Stebbings et al., 2002). Therefore, investigating the function of gap junctions in distinct cell types and in an isoform-specific manner remains extremely challenging. To overcome these challenges, new methods providing improved genetic specificity, high spatial resolution, and functionally relevant temporal resolution are urgently needed. Ideally, these methods should be non-invasive and technically simple to perform, thereby facilitating their use in in vivo applications, allowing researchers to study gap junctions in a more physiological setting.

In principle, using genetically encoded tools provides a possible solution. For example, the PLE-ester and Pept2-βALA strategies discussed above eliminate the need to manually manipulate the cells with glass pipettes, while providing the advantages associated with fluorescence imaging (Tian et al., 2012; Qiao and Sanes, 2015). On the other hand, the patch-clamp–based Pado and ChIEF strategies provide faster kinetics and do not require an exogenous substrate, making the background signal easier to control by regulating the expression level (Wang et al., 2014; Kang and Baker, 2016). However, each of these methods includes a non-genetically encoded component (e.g., an exogenous substrate or whole-cell patch-clamp recording), which have inherent limitations as discussed above.

The vast majority of genetically encoded methods used to date are based on the diffusion of target molecules such as esters, ions, peptides, or synthetic dyes through gap junctions. In each case, the electrochemical gradient that drives this diffusion is generated exogenously (e.g., by patch clamp, tracer introduction, or substrate application), and the chemical transfer is usually detected using fluorescent probes. Thus, we can summarize the entire system as consisting of a “generator” and a “reporter”; the generator produces the electrochemical gradient between coupled cells, and the reporter reports the transfer of molecules through the gap junctions (Figure 2A). In a system comprised exclusively of genetically encoded optogenetics-based components, both the generator and the reporter would be proteins (e.g., a light-activated channel or transporter and a fluorescent sensor). In this idealized system, the generator would be controlled by light and would use the cel’s endogenous ions or chemicals to generate the electrochemical gradient, and the reporter would sense the change in concentration and change its fluorescence intensity. This non-invasive optogenetics-based system could be used to control and image a large number of cells simultaneously, and the background fluorescence could be minimized greatly by controlling the expression of the generator and reporter. More importantly, multicolor imaging could be achieved—at least in theory—by using a combination of generators with non-overlapping action wavelengths and/or reporters with non-overlapping excitation and emission spectra (Figure 2B). Given the wide range of clear benefits associated with this approach, genetically encoded optogenetics represents one of the most promising strategies for studying gap junctions in the future.

FIGURE 2
www.frontiersin.org

FIGURE 2. A proposed ideal optogenetics-based system for mapping gap junctions. (A) The principle behind the proposed optogenetics-based system shown on the left with its theoretical performance index on the right, similar to Figure 1. (B) A proposed multiplex, optogenetic system for mapping gap junction using two pairs of bio-orthogonal generators and reporters and its application in an intact tissue with heterogeneous cell types.

Author Contributions

YL, AD, and SL conceived and wrote the manuscript.

Funding

This work was supported by the National Basic Research Program of China (973 Program; grant 2015CB856402 to YL), the General Program of National Natural Science Foundation of China (projects 31671118 and 31371442 to YL), and the Junior Thousand Talents Program of China (to YL).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank the members of the Li lab for providing feedback on the manuscript.

References

Abbaci, M., Barberi-Heyob, M., Blondel, W., Guillemin, F., and Didelon, J. (2008). Advantages and limitations of commonly used methods to assay the molecular permeability of gap junctional intercellular communication. Biotechniques 45, 33–52, 56–62. doi: 10.2144/000112810

PubMed Abstract | CrossRef Full Text | Google Scholar

Ansar Ahmed, S., Penhale, W. J., and Talal, N. (1985). Sex hormones, immune responses, and autoimmune diseases. Mechanisms of sex hormone action. Am. J. Pathol. 121, 531–551.

Google Scholar

Antonsen, B. L., and Edwards, D. H. (2003). Differential dye coupling reveals lateral giant escape circuit in crayfish. J. Comp. Neurol. 466, 1–13. doi: 10.1002/cne.10802

PubMed Abstract | CrossRef Full Text | Google Scholar

Azzam, E. I., De Toledo, S. M., and Little, J. B. (2001). Direct evidence for the participation of gap junction-mediated intercellular communication in the transmission of damage signals from alpha -particle irradiated to nonirradiated cells. Proc. Natl. Acad. Sci. U.S.A. 98, 473–478. doi: 10.1073/pnas.011417098

PubMed Abstract | CrossRef Full Text | Google Scholar

Baranova, A., Ivanov, D., Petrash, N., Pestova, A., Skoblov, M., Kelmanson, I., et al. (2004). The mammalian pannexin family is homologous to the invertebrate innexin gap junction proteins. Genomics 83, 706–716. doi: 10.1016/j.ygeno.2003.09.025

PubMed Abstract | CrossRef Full Text | Google Scholar

Bennett, M. V., and Zukin, R. S. (2004). Electrical coupling and neuronal synchronization in the Mammalian brain. Neuron 41, 495–511. doi: 10.1016/S0896-6273(04)00043-1

CrossRef Full Text | Google Scholar

Beyer, E. C., Kistler, J., Paul, D. L., and Goodenough, D. A. (1989). Antisera directed against connexin43 peptides react with a 43-kD protein localized to gap junctions in myocardium and other tissues. J. Cell Biol. 108, 595–605. doi: 10.1083/jcb.108.2.595

PubMed Abstract | CrossRef Full Text | Google Scholar

Biegel, A., Knutter, I., Hartrodt, B., Gebauer, S., Theis, S., Luckner, P., et al. (2006). The renal type H + /peptide symporter PEPT2: structure-affinity relationships. Amino Acids 31, 137–156. doi: 10.1007/s00726-006-0331-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Bloomfield, S. A., and Volgyi, B. (2009). The diverse functional roles and regulation of neuronal gap junctions in the retina. Nat. Rev. Neurosci. 10, 495–506. doi: 10.1038/nrn2636

PubMed Abstract | CrossRef Full Text | Google Scholar

Bruzzone, R., Hormuzdi, S. G., Barbe, M. T., Herb, A., and Monyer, H. (2003). Pannexins, a family of gap junction proteins expressed in brain. Proc. Natl. Acad. Sci. U.S.A. 100, 13644–13649. doi: 10.1073/pnas.2233464100

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, T. W., Wardill, T. J., Sun, Y., Pulver, S. R., Renninger, S. L., Baohan, A., et al. (2013). Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300. doi: 10.1038/nature12354

PubMed Abstract | CrossRef Full Text | Google Scholar

Dakin, K., and Li, W. H. (2006). Local Ca2 + rise near store operated Ca2 + channels inhibits cell coupling during capacitative Ca2 + influx. Cell Commun. Adhes. 13, 29–39. doi: 10.1080/15419060600631425

PubMed Abstract | CrossRef Full Text | Google Scholar

Dakin, K., Zhao, Y., and Li, W. H. (2005). LAMP, a new imaging assay of gap junctional communication unveils that Ca2 + influx inhibits cell coupling. Nat. Methods 2, 55–62. doi: 10.1038/nmeth730

PubMed Abstract | CrossRef Full Text | Google Scholar

Decrouy, X., Gasc, J. M., Pointis, G., and Segretain, D. (2004). Functional characterization of Cx43 based gap junctions during spermatogenesis. J. Cell. Physiol. 200, 146–154. doi: 10.1002/jcp.10473

PubMed Abstract | CrossRef Full Text | Google Scholar

Dermietzel, R., Traub, O., Hwang, T. K., Beyer, E., Bennett, M. V., Spray, D. C., et al. (1989). Differential expression of three gap junction proteins in developing and mature brain tissues. Proc. Natl. Acad. Sci. U.S.A. 86, 10148–10152. doi: 10.1073/pnas.86.24.10148

PubMed Abstract | CrossRef Full Text | Google Scholar

D’hondt, C., Ponsaerts, R., De Smedt, H., Bultynck, G., and Himpens, B. (2009). Pannexins, distant relatives of the connexin family with specific cellular functions? Bioessays 31, 953–974. doi: 10.1002/bies.200800236

PubMed Abstract | CrossRef Full Text | Google Scholar

Dieck, S. T., Heuer, H., Ehrchen, J., Otto, C., and Bauer, K. (1999). The peptide transporter PepT2 is expressed in rat brain and mediates the accumulation of the fluorescent dipeptide derivative beta-Ala-Lys-Nepsilon-AMCA in astrocytes. Glia 25, 10–20. doi: 10.1002/(SICI)1098-1136(19990101)25:1<10::AID-GLIA2>3.0.CO;2-Y

PubMed Abstract | CrossRef Full Text | Google Scholar

Dobrenis, K., Chang, H. Y., Pina-Benabou, M. H., Woodroffe, A., Lee, S. C., Rozental, R., et al. (2005). Human and mouse microglia express connexin36, and functional gap junctions are formed between rodent microglia and neurons. J. Neurosci. Res. 82, 306–315. doi: 10.1002/jnr.20650

PubMed Abstract | CrossRef Full Text | Google Scholar

el-Fouly, M. H., Trosko, J. E., and Chang, C. C. (1987). Scrape-loading and dye transfer. A rapid and simple technique to study gap junctional intercellular communication. Exp. Cell Res. 168, 422–430. doi: 10.1016/0014-4827(87)90014-0

CrossRef Full Text | Google Scholar

Furshpan, E. J., and Potter, D. D. (1957). Mechanism of nerve-impulse transmission at a crayfish synapse. Nature 180, 342–343. doi: 10.1038/180342a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Furshpan, E. J., and Potter, D. D. (1959). Transmission at the giant motor synapses of the crayfish. J. Physiol. 145, 289–325. doi: 10.1113/jphysiol.1959.sp006143

CrossRef Full Text | Google Scholar

Garg, S., Md Syed, M., and Kielian, T. (2005). Staphylococcus aureus-derived peptidoglycan induces Cx43 expression and functional gap junction intercellular communication in microglia. J. Neurochem. 95, 475–483. doi: 10.1111/j.1471-4159.2005.03384.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Giaume, C., and McCarthy, K. D. (1996). Control of gap-junctional communication in astrocytic networks. Trends Neurosci. 19, 319–325. doi: 10.1016/0166-2236(96)10046-1

CrossRef Full Text | Google Scholar

Giepmans, B. N., and Moolenaar, W. H. (1998). The gap junction protein connexin43 interacts with the second PDZ domain of the zona occludens-1 protein. Curr. Biol. 8, 931–934. doi: 10.1016/S0960-9822(07)00375-2

CrossRef Full Text | Google Scholar

Giustina, A., and Veldhuis, J. D. (1998). Pathophysiology of the neuroregulation of growth hormone secretion in experimental animals and the human. Endocr. Rev. 19, 717–797. doi: 10.1210/er.19.6.717

PubMed Abstract | CrossRef Full Text | Google Scholar

Han, X., and Boyden, E. S. (2007). Multiple-color optical activation, silencing, and desynchronization of neural activity, with single-spike temporal resolution. PLoS One 2:e299. doi: 10.1371/journal.pone.0000299

PubMed Abstract | CrossRef Full Text | Google Scholar

Horikawa, K., and Armstrong, W. E. (1988). A versatile means of intracellular labeling: injection of biocytin and its detection with avidin conjugates. J. Neurosci. Methods 25, 1–11. doi: 10.1016/0165-0270(88)90114-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoshi, H., and Mills, S. L. (2009). Components and properties of the G3 ganglion cell circuit in the rabbit retina. J. Comp. Neurol. 513, 69–82. doi: 10.1002/cne.21941

PubMed Abstract | CrossRef Full Text | Google Scholar

Ishikawa, M., Iwamoto, T., Nakamura, T., Doyle, A., Fukumoto, S., and Yamada, Y. (2011). Pannexin 3 functions as an ER Ca(2 + ) channel, hemichannel, and gap junction to promote osteoblast differentiation. J. Cell Biol. 193, 1257–1274. doi: 10.1083/jcb.201101050

PubMed Abstract | CrossRef Full Text | Google Scholar

Ito, K., Sass, H., Urban, J., Hofbauer, A., and Schneuwly, S. (1997). GAL4-responsive UAS-tau as a tool for studying the anatomy and development of the Drosophila central nervous system. Cell Tissue Res. 290, 1–10. doi: 10.1007/s004410050901

PubMed Abstract | CrossRef Full Text | Google Scholar

Jin, L., Han, Z., Platisa, J., Wooltorton, J. R., Cohen, L. B., and Pieribone, V. A. (2012). Single action potentials and subthreshold electrical events imaged in neurons with a fluorescent protein voltage probe. Neuron 75, 779–785. doi: 10.1016/j.neuron.2012.06.040

PubMed Abstract | CrossRef Full Text | Google Scholar

Kang, B. E., and Baker, B. J. (2016). Pado, a fluorescent protein with proton channel activity can optically monitor membrane potential, intracellular pH, and map gap junctions. Sci. Rep. 6:23865. doi: 10.1038/srep23865

PubMed Abstract | CrossRef Full Text | Google Scholar

Katz, B., and Miledi, R. (1965). The measurement of synaptic delay, and the time course of acetylcholine release at the neuromuscular junction. Proc. R. Soc. Lond. B Biol. Sci. 161, 483–495. doi: 10.1098/rspb.1965.0016

PubMed Abstract | CrossRef Full Text | Google Scholar

Kleopa, K. A. (2011). The role of gap junctions in Charcot-Marie-Tooth disease. J. Neurosci. 31, 17753–17760. doi: 10.1523/JNEUROSCI.4824-11.2011

CrossRef Full Text | Google Scholar

Krnjevic, K. (1974). Chemical nature of synaptic transmission in vertebrates. Physiol. Rev. 54, 418–540. doi: 10.1152/physrev.1974.54.2.418

CrossRef Full Text | Google Scholar

Kumar, N. M., and Gilula, N. B. (1996). The gap junction communication channel. Cell 84, 381–388. doi: 10.1016/S0092-8674(00)81282-9

CrossRef Full Text | Google Scholar

Lai, C. P., Bechberger, J. F., Thompson, R. J., Macvicar, B. A., Bruzzone, R., and Naus, C. C. (2007). Tumor-suppressive effects of pannexin 1 in C6 glioma cells. Cancer Res. 67, 1545–1554. doi: 10.1158/0008-5472.CAN-06-1396

PubMed Abstract | CrossRef Full Text | Google Scholar

Lake, B. B., Ai, R., Kaeser, G. E., Salathia, N. S., Yung, Y. C., Liu, R., et al. (2016). Neuronal subtypes and diversity revealed by single-nucleus RNA sequencing of the human brain. Science 352, 1586–1590. doi: 10.1126/science.aaf1204

PubMed Abstract | CrossRef Full Text | Google Scholar

Lange, S., Musidlowska, A., Schmidt-Dannert, C., Schmitt, J., and Bornscheuer, U. T. (2001). Cloning, functional expression, and characterization of recombinant pig liver esterase. Chembiochem 2, 576–582. doi: 10.1002/1439-7633(20010803)2:7/8<576::AID-CBIC576>3.0.CO;2-Y

CrossRef Full Text | Google Scholar

Lee, S. H., Magge, S., Spencer, D. D., Sontheimer, H., and Cornell-Bell, A. H. (1995). Human epileptic astrocytes exhibit increased gap junction coupling. Glia 15, 195–202. doi: 10.1002/glia.440150212

PubMed Abstract | CrossRef Full Text | Google Scholar

Lin, J. Y., Lin, M. Z., Steinbach, P., and Tsien, R. Y. (2009). Characterization of engineered channelrhodopsin variants with improved properties and kinetics. Biophys. J. 96, 1803–1814. doi: 10.1016/j.bpj.2008.11.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Lippincott-Schwartz, J., Altan-Bonnet, N., and Patterson, G. H. (2003). Photobleaching and photoactivation: following protein dynamics in living cells. Nat. Cell Biol. 5(Suppl.), S7–S14.

Google Scholar

Loewenstein, W. R. (1981). Junctional intercellular communication: the cell-to-cell membrane channel. Physiol. Rev. 61, 829–913. doi: 10.1152/physrev.1981.61.4.829

PubMed Abstract | CrossRef Full Text | Google Scholar

Martinez, A. D., Acuna, R., Figueroa, V., Maripillan, J., and Nicholson, B. (2009). Gap-junction channels dysfunction in deafness and hearing loss. Antioxid. Redox Signal. 11, 309–322. doi: 10.1089/ars.2008.2138

PubMed Abstract | CrossRef Full Text | Google Scholar

McNeil, P. L., Murphy, R. F., Lanni, F., and Taylor, D. L. (1984). A method for incorporating macromolecules into adherent cells. J. Cell Biol. 98, 1556–1564. doi: 10.1083/jcb.98.4.1556

PubMed Abstract | CrossRef Full Text | Google Scholar

Meier, U., and Gressner, A. M. (2004). Endocrine regulation of energy metabolism: review of pathobiochemical and clinical chemical aspects of leptin, ghrelin, adiponectin, and resistin. Clin. Chem. 50, 1511–1525. doi: 10.1373/clinchem.2004.032482

PubMed Abstract | CrossRef Full Text | Google Scholar

Meng, L., Zhang, A., Jin, Y., and Yan, D. (2016). Regulation of neuronal axon specification by glia-neuron gap junctions in C. elegans. eLife 5:e19510. doi: 10.7554/eLife.19510

PubMed Abstract | CrossRef Full Text | Google Scholar

Mese, G., Richard, G., and White, T. W. (2007). Gap junctions: basic structure and function. J. Invest. Dermatol. 127, 2516–2524. doi: 10.1038/sj.jid.5700770

PubMed Abstract | CrossRef Full Text | Google Scholar

Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvili, N., Berthold, P., et al. (2003). Channelrhodopsin-2, a directly light-gated cation-selective membrane channel. Proc. Natl. Acad. Sci. U.S.A. 100, 13940–13945. doi: 10.1073/pnas.1936192100

PubMed Abstract | CrossRef Full Text | Google Scholar

Nagel, G., Szellas, T., Kateriya, S., Adeishvili, N., Hegemann, P., and Bamberg, E. (2005). Channelrhodopsins: directly light-gated cation channels. Biochem. Soc. Trans. 33, 863–866. doi: 10.1042/BST0330863

PubMed Abstract | CrossRef Full Text | Google Scholar

Neijssen, J., Herberts, C., Drijfhout, J. W., Reits, E., Janssen, L., and Neefjes, J. (2005). Cross-presentation by intercellular peptide transfer through gap junctions. Nature 434, 83–88. doi: 10.1038/nature03290

PubMed Abstract | CrossRef Full Text | Google Scholar

Nualart-Marti, A., Solsona, C., and Fields, R. D. (2013). Gap junction communication in myelinating glia. Biochim. Biophys. Acta 1828, 69–78. doi: 10.1016/j.bbamem.2012.01.024

PubMed Abstract | CrossRef Full Text | Google Scholar

Paul, D. L., Ebihara, L., Takemoto, L. J., Swenson, K. I., and Goodenough, D. A. (1991). Connexin46, a novel lens gap junction protein, induces voltage-gated currents in nonjunctional plasma membrane of Xenopus oocytes. J. Cell Biol. 115, 1077–1089. doi: 10.1083/jcb.115.4.1077

PubMed Abstract | CrossRef Full Text | Google Scholar

Pereda, A. E. (2014). Electrical synapses and their functional interactions with chemical synapses. Nat. Rev. Neurosci. 15, 250–263. doi: 10.1038/nrn3708

PubMed Abstract | CrossRef Full Text | Google Scholar

Personius, K. E., Chang, Q., Mentis, G. Z., O’donovan, M. J., and Balice-Gordon, R. J. (2007). Reduced gap junctional coupling leads to uncorrelated motor neuron firing and precocious neuromuscular synapse elimination. Proc. Natl. Acad. Sci. U.S.A. 104, 11808–11813. doi: 10.1073/pnas.0703357104

PubMed Abstract | CrossRef Full Text | Google Scholar

Phelan, P., Bacon, J. P., Davies, J. A., Stebbings, L. A., Todman, M. G., Avery, L., et al. (1998). Innexins: a family of invertebrate gap-junction proteins. Trends Genet. 14, 348–349. doi: 10.1016/S0168-9525(98)01547-9

CrossRef Full Text | Google Scholar

Phelan, P., Nakagawa, M., Wilkin, M. B., Moffat, K. G., O’kane, C. J., Davies, J. A., et al. (1996). Mutations in shaking-B prevent electrical synapse formation in the Drosophila giant fiber system. J. Neurosci. 16, 1101–1113. doi: 10.1523/JNEUROSCI.16-03-01101.1996

PubMed Abstract | CrossRef Full Text | Google Scholar

Qiao, M., and Sanes, J. R. (2015). Genetic method for labeling electrically coupled cells: application to retina. Front. Mol. Neurosci. 8:81. doi: 10.3389/fnmol.2015.00081

PubMed Abstract | CrossRef Full Text | Google Scholar

Ritter, C., Nett, N., Acevedo-Rocha, C. G., Lonsdale, R., Kraling, K., Dempwolff, F., et al. (2015). Bioorthogonal enzymatic activation of caged compounds. Angew. Chem. Int. Ed. Engl. 54, 13440–13443. doi: 10.1002/anie.201506739

PubMed Abstract | CrossRef Full Text | Google Scholar

Rotman, B., and Papermaster, B. W. (1966). Membrane properties of living mammalian cells as studied by enzymatic hydrolysis of fluorogenic esters. Proc. Natl. Acad. Sci. U.S.A. 55, 134–141. doi: 10.1073/pnas.55.1.134

PubMed Abstract | CrossRef Full Text | Google Scholar

Sabatini, B. L., and Regehr, W. G. (1996). Timing of neurotransmission at fast synapses in the mammalian brain. Nature 384, 170–172. doi: 10.1038/384170a0

PubMed Abstract | CrossRef Full Text | Google Scholar

Sletten, E. M., and Bertozzi, C. R. (2009). Bioorthogonal chemistry: fishing for selectivity in a sea of functionality. Angew. Chem. Int. Ed. Engl. 48, 6974–6998. doi: 10.1002/anie.200900942

PubMed Abstract | CrossRef Full Text | Google Scholar

Sohl, G., Maxeiner, S., and Willecke, K. (2005). Expression and functions of neuronal gap junctions. Nat. Rev. Neurosci. 6, 191–200. doi: 10.1038/nrn1627

PubMed Abstract | CrossRef Full Text | Google Scholar

Sohl, G., and Willecke, K. (2003). An update on connexin genes and their nomenclature in mouse and man. Cell Commun. Adhes. 10, 173–180. doi: 10.1080/cac.10.4-6.173.180

PubMed Abstract | CrossRef Full Text | Google Scholar

Song, J., and Tanouye, M. A. (2006). Seizure suppression by shakB2, a gap junction mutation in Drosophila. J. Neurophysiol. 95, 627–635. doi: 10.1152/jn.01059.2004

PubMed Abstract | CrossRef Full Text | Google Scholar

Soroceanu, L., Manning, T. J. Jr., and Sontheimer, H. (2001). Reduced expression of connexin-43 and functional gap junction coupling in human gliomas. Glia 33, 107–117. doi: 10.1002/1098-1136(200102)33:2<107::AID-GLIA1010>3.0.CO;2-4

CrossRef Full Text | Google Scholar

Sosinsky, G. E., Boassa, D., Dermietzel, R., Duffy, H. S., Laird, D. W., Macvicar, B., et al. (2011). Pannexin channels are not gap junction hemichannels. Channels 5, 193–197. doi: 10.4161/chan.5.3.15765

CrossRef Full Text | Google Scholar

Speder, P., and Brand, A. H. (2014). Gap junction proteins in the blood-brain barrier control nutrient-dependent reactivation of Drosophila neural stem cells. Dev. Cell 30, 309–321. doi: 10.1016/j.devcel.2014.05.021

PubMed Abstract | CrossRef Full Text | Google Scholar

Spray, D. C., Harris, A. L., and Bennett, M. V. (1979). Voltage dependence of junctional conductance in early amphibian embryos. Science 204, 432–434. doi: 10.1126/science.312530

PubMed Abstract | CrossRef Full Text | Google Scholar

Starich, T., Sheehan, M., Jadrich, J., and Shaw, J. (2001). Innexins in C. elegans. Cell Commun. Adhes. 8, 311–314. doi: 10.3109/15419060109080744

CrossRef Full Text | Google Scholar

Stauffer, K. A. (1995). The gap junction proteins beta 1-connexin (connexin-32) and beta 2-connexin (connexin-26) can form heteromeric hemichannels. J. Biol. Chem. 270, 6768–6772.

PubMed Abstract | Google Scholar

Stebbings, L. A., Todman, M. G., Phillips, R., Greer, C. E., Tam, J., Phelan, P., et al. (2002). Gap junctions in Drosophila: developmental expression of the entire innexin gene family. Mech. Dev. 113, 197–205. doi: 10.1016/S0925-4773(02)00025-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Stewart, W. W. (1978). Functional connections between cells as revealed by dye-coupling with a highly fluorescent naphthalimide tracer. Cell 14, 741–759. doi: 10.1016/0092-8674(78)90256-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Tasic, B., Menon, V., Nguyen, T. N., Kim, T. K., Jarsky, T., Yao, Z., et al. (2016). Adult mouse cortical cell taxonomy revealed by single cell transcriptomics. Nat. Neurosci. 19, 335–346. doi: 10.1038/nn.4216

PubMed Abstract | CrossRef Full Text | Google Scholar

Thomas, J. B., and Wyman, R. J. (1984). Mutations altering synaptic connectivity between identified neurons in Drosophila. J. Neurosci. 4, 530–538. doi: 10.1523/JNEUROSCI.04-02-00530.1984

PubMed Abstract | CrossRef Full Text | Google Scholar

Tian, L., Yang, Y., Wysocki, L. M., Arnold, A. C., Hu, A., Ravichandran, B., et al. (2012). Selective esterase-ester pair for targeting small molecules with cellular specificity. Proc. Natl. Acad. Sci. U.S.A. 109, 4756–4761. doi: 10.1073/pnas.1111943109

PubMed Abstract | CrossRef Full Text | Google Scholar

Tour, O., Adams, S. R., Kerr, R. A., Meijer, R. M., Sejnowski, T. J., Tsien, R. W., et al. (2007). Calcium Green FlAsH as a genetically targeted small-molecule calcium indicator. Nat. Chem. Biol. 3, 423–431. doi: 10.1038/nchembio.2007.4

PubMed Abstract | CrossRef Full Text | Google Scholar

Vanden Abeele, F., Bidaux, G., Gordienko, D., Beck, B., Panchin, Y. V., Baranova, A. V., et al. (2006). Functional implications of calcium permeability of the channel formed by pannexin 1. J. Cell Biol. 174, 535–546. doi: 10.1083/jcb.200601115

PubMed Abstract | CrossRef Full Text | Google Scholar

Volgyi, B., Chheda, S., and Bloomfield, S. A. (2009). Tracer coupling patterns of the ganglion cell subtypes in the mouse retina. J. Comp. Neurol. 512, 664–687. doi: 10.1002/cne.21912

PubMed Abstract | CrossRef Full Text | Google Scholar

Wade, M. H., Trosko, J. E., and Schindler, M. (1986). A fluorescence photobleaching assay of gap junction-mediated communication between human cells. Science 232, 525–528. doi: 10.1126/science.3961495

CrossRef Full Text | Google Scholar

Wallraff, A., Odermatt, B., Willecke, K., and Steinhauser, C. (2004). Distinct types of astroglial cells in the hippocampus differ in gap junction coupling. Glia 48, 36–43. doi: 10.1002/glia.20040

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, K., Gong, J., Wang, Q., Li, H., Cheng, Q., Liu, Y., et al. (2014). Parallel pathways convey olfactory information with opposite polarities in Drosophila. Proc. Natl. Acad. Sci. U.S.A. 111, 3164–3169. doi: 10.1073/pnas.1317911111

PubMed Abstract | CrossRef Full Text | Google Scholar

White, T. W., Bruzzone, R., Goodenough, D. A., and Paul, D. L. (1992). Mouse Cx50, a functional member of the connexin family of gap junction proteins, is the lens fiber protein MP70. Mol. Biol. Cell 3, 711–720. doi: 10.1091/mbc.3.7.711

PubMed Abstract | CrossRef Full Text | Google Scholar

Wrenzycki, C., Herrmann, D., Carnwath, J. W., and Niemann, H. (1996). Expression of the gap junction gene connexin43 (Cx43) in preimplantation bovine embryos derived in vitro or in vivo. J. Reprod. Fertil. 108, 17–24. doi: 10.1530/jrf.0.1080017

CrossRef Full Text | Google Scholar

Xia, J. H., Liu, C. Y., Tang, B. S., Pan, Q., Huang, L., Dai, H. P., et al. (1998). Mutations in the gene encoding gap junction protein beta-3 associated with autosomal dominant hearing impairment. Nat. Genet. 20, 370–373. doi: 10.1038/3845

PubMed Abstract | CrossRef Full Text | Google Scholar

Xin, D., and Bloomfield, S. A. (1999). Dark- and light-induced changes in coupling between horizontal cells in mammalian retina. J. Comp. Neurol. 405, 75–87. doi: 10.1002/(SICI)1096-9861(19990301)405:1<75::AID-CNE6>3.0.CO;2-D

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, F., Wang, L. P., Brauner, M., Liewald, J. F., Kay, K., Watzke, N., et al. (2007). Multimodal fast optical interrogation of neural circuitry. Nature 446, 633–639. doi: 10.1038/nature05744

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Z., Curtin, K. D., Sun, Y. A., and Wyman, R. J. (1999). Nested transcripts of gap junction gene have distinct expression patterns. J. Neurobiol. 40, 288–301. doi: 10.1002/(SICI)1097-4695(19990905)40:3<288::AID-NEU2>3.0.CO;2-O

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Y., Zheng, Q., Dakin, K., Xu, K., Martinez, M. L., and Li, W. H. (2004). New caged coumarin fluorophores with extraordinary uncaging cross sections suitable for biological imaging applications. J. Am. Chem. Soc. 126, 4653–4663. doi: 10.1021/ja036958m

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: gap junction, electrical synapse, fluorescence imaging, genetically encoded methods, nervous system

Citation: Dong A, Liu S and Li Y (2018) Gap Junctions in the Nervous System: Probing Functional Connections Using New Imaging Approaches. Front. Cell. Neurosci. 12:320. doi: 10.3389/fncel.2018.00320

Received: 01 July 2018; Accepted: 03 September 2018;
Published: 19 September 2018.

Edited by:

Bradley James Baker, Korea Institute of Science and Technology (KIST), South Korea

Reviewed by:

Bela Volgyi, University of Pécs, Hungary
Pablo Jose Saez, Institut Curie, France

Copyright © 2018 Dong, Liu and Li. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Ao Dong, aodong@pku.edu.cn Yulong Li, yulongli@pku.edu.cn, orcid.org/0000-0002-9166-9919

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.