Skip to main content

Main menu

  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Collections
    • Podcast
  • ALERTS
  • FOR AUTHORS
    • Information for Authors
    • Fees
    • Journal Clubs
    • eLetters
    • Submit
    • Special Collections
  • EDITORIAL BOARD
    • Editorial Board
    • ECR Advisory Board
    • Journal Staff
  • ABOUT
    • Overview
    • Advertise
    • For the Media
    • Rights and Permissions
    • Privacy Policy
    • Feedback
    • Accessibility
  • SUBSCRIBE

User menu

  • Log out
  • Log in
  • My Cart

Search

  • Advanced search
Journal of Neuroscience
  • Log out
  • Log in
  • My Cart
Journal of Neuroscience

Advanced Search

Submit a Manuscript
  • HOME
  • CONTENT
    • Early Release
    • Featured
    • Current Issue
    • Issue Archive
    • Collections
    • Podcast
  • ALERTS
  • FOR AUTHORS
    • Information for Authors
    • Fees
    • Journal Clubs
    • eLetters
    • Submit
    • Special Collections
  • EDITORIAL BOARD
    • Editorial Board
    • ECR Advisory Board
    • Journal Staff
  • ABOUT
    • Overview
    • Advertise
    • For the Media
    • Rights and Permissions
    • Privacy Policy
    • Feedback
    • Accessibility
  • SUBSCRIBE
PreviousNext
TechSights

Beyond a Transmission Cable—New Technologies to Reveal the Richness in Axonal Electrophysiology

J. C. Mateus, M. M. Sousa, J. Burrone and P. Aguiar
Journal of Neuroscience 13 March 2024, 44 (11) e1446232023; https://doi.org/10.1523/JNEUROSCI.1446-23.2023
J. C. Mateus
1i3S- Instituto de Investigação e Inovação em Saúde, Universidade do Porto, 4200-135 Porto, Portugal
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for J. C. Mateus
M. M. Sousa
1i3S- Instituto de Investigação e Inovação em Saúde, Universidade do Porto, 4200-135 Porto, Portugal
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
J. Burrone
2MRC Centre for Neurodevelopmental Disorders, Institute of Psychiatry, Psychology and Neuroscience, King’s College London, London SE1 1UL, United Kingdom
3Centre for Developmental Neurobiology, Institute of Psychiatry, Psychology and Neuroscience, King’s College London, London SE1 1UL, United Kingdom
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for J. Burrone
P. Aguiar
1i3S- Instituto de Investigação e Inovação em Saúde, Universidade do Porto, 4200-135 Porto, Portugal
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • ORCID record for P. Aguiar
  • Article
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF
Loading

Abstract

The axon is a neuronal structure capable of processing, encoding, and transmitting information. This assessment contrasts with a limiting, but deeply rooted, perspective where the axon functions solely as a transmission cable of somatodendritic activity, sending signals in the form of stereotypical action potentials. This perspective arose, at least partially, because of the technical difficulties in probing axons: their extreme length-to-diameter ratio and intricate growth paths preclude the study of their dynamics through traditional techniques. Recent findings are challenging this view and revealing a much larger repertoire of axonal computations. Axons display complex signaling processes and structure–function relationships, which can be modulated via diverse activity-dependent mechanisms. Additionally, axons can exhibit patterns of activity that are dramatically different from those of their corresponding soma. Not surprisingly, many of these recent discoveries have been driven by novel technology developments, which allow for in vitro axon electrophysiology with unprecedented spatiotemporal resolution and signal-to-noise ratio. In this review, we outline the state-of-the-art in vitro toolset for axonal electrophysiology and summarize the recent discoveries in axon function it has enabled. We also review the increasing repertoire of microtechnologies for controlling axon guidance which, in combination with the available cutting-edge electrophysiology and imaging approaches, have the potential for more controlled and high-throughput in vitro studies. We anticipate that a larger adoption of these new technologies by the neuroscience community will drive a new era of experimental opportunities in the study of axon physiology and consequently, neuronal function.

  • axon computations
  • axon electrophysiology
  • axon guidance
  • functional imaging
  • microelectrode arrays

Significance Statement

Recent experimental evidence is revealing the rich electrophysiology of the axon and the importance of this structure for setting important neuronal computations. Most of these recent discoveries have been achieved at the intersection of the fields of neurotechnology and neurobiology. In particular, key recent findings in axon function have been made possible by the development of specific cutting-edge techniques that allow for the targeted probing of axons with high spatiotemporal resolution and signal-to-noise ratio. Here we review the state-of-the-art electrophysiological and optical toolsets for in vitro axonal experimentation and summarize the most recent and relevant discoveries they enabled. Altogether, we give the reader an up-to-date understanding of the neurotechnologies available to dissect axonal function.

Introduction

Historically, axons have been viewed as simple transmission cables of upstream decisions. In the central nervous system (CNS), this function is reduced to the reliable transmission of information from the somatodendritic compartments to the presynaptic terminals in the form of stereotypical “all-or-none” action potentials (APs). This perspective is embedded in the origin of the word axon (from Greek “áxōn,” meaning “axis”) and has been driven, at least partially, by the technical difficulties in probing axonal function, which bias most technical approaches to the recording of somatic APs alone. However, seven decades after the pioneering studies on the active properties of signal conduction along the axon (Hodgkin and Huxley, 1952), an increasing body of knowledge is challenging this limiting view (reviewed in (Bucher and Goaillard, 2011; Debanne et al., 2011; Rama et al., 2018; Alcami and el Hady, 2019). Key findings such as distal AP initiation (Dugladze et al., 2012; Deemyad et al., 2018; Mateus et al., 2021; Liu et al., 2022a; Rózsa et al., 2023); activity-dependent axon plasticity (Grubb and Burrone, 2010a; Kuba et al., 2010; Sheffield et al., 2011; Chéreau et al., 2017; Lezmy et al., 2017); and analog–digital facilitation of synaptic transmission (Alle and Geiger, 2006; Zbili et al., 2020) have sparked new interest in the study of axon electrophysiology and are decisively shifting the field away from the view of axons as simple transmission cables. Still, the full implication of this computational repertoire is far from understood.

Most neurons encode and transmit information via the conduction of APs along the axon arbor. AP propagation controls the reliability and the timing with which neuronal networks communicate; thus the modulation of this process has repercussions for, at least, temporal coding. Remarkably, most (if not all) parameters that influence AP propagation along the axon (e.g., axon diameter, ion channel density) can be modified by, at least, neuronal activity levels (Chéreau et al., 2017). Important AP characteristics, such as AP shape, can also be modulated locally via, for example, voltage-gated channels’ heterogeneous distribution, to convey information beyond timing (reviewed in Rama et al., 2018). Intriguingly, APs in the CNS may even initiate in distal parts of the axon, independently of somatodendritic integration, and propagate (antidromically) toward the soma, as well as toward axon terminals (Dugladze et al., 2012; Liu et al., 2022a; Rózsa et al., 2023). On the other hand, sensory neurons, which typically initiate APs in the peripheral terminals, can assemble an axon initial segment (AIS)-like structure in the proximal stem axon that acts as a source of spontaneous activity in neuropathic pain (Nascimento et al., 2022). This bidirectional axonal conduction, with notable implications for neuronal communication, can also be observed in vitro in both CNS and sensory neurons (Mateus et al., 2021). Many of these recent and unexpected findings have been made possible by a novel and extending repertoire of tools that allow for the probing of mammalian axons with unprecedented spatiotemporal resolution and signal-to-noise ratio (SNR).

Originally, most of our knowledge about axonal electrophysiology are derived from in silico studies and experiments on invertebrate models, despite their relatively limited morphological and functional complexity (Bean, 2007; Bucher and Goaillard, 2011). While the seminal work of Hodgkin and Huxley made use of millimeter-sized invertebrate axons that enabled experimental access to direct electrophysiological recording (Hodgkin and Huxley, 1952), the mammalian axons’ morphology long precluded the application of the same technical approach. Mammalian axons are typically the longest neuronal subcompartment (measuring up to meters) but are also extremely thin (≈0.1–20 µm) due to demanding energy requirements and space constraints (Perge et al., 2012; Costa et al., 2018). Axons can emerge from dendritic processes (Kole and Brette, 2018; Hodapp et al., 2022) and branch profusely into collaterals that are even thinner than the main axon trunk (Sasaki et al., 2012a; Mateus et al., 2021). Altogether, these morphological characteristics have hindered the study of mammalian axon's electrophysiology but have also provided clues about structure–function relationships unique to mammals.

Recent technological developments in nano-microfabrication and optical sensing have led to the emergence of new tools that allow for axon probing in vitro with high fidelity. Compared with the in vivo setting, in vitro models have the advantages of easier accessibility and manipulation without many of the confounding factors (e.g., surrounding tissue) and ethical challenges of animal experimentation. Moreover, they are compatible with emerging human-induced pluripotent stem cell (hiPSC) technology, which allows for the engineering of multiple cell types directly from patients (Girardin et al., 2022; Revah et al., 2022). Taken together, these developments have greatly increased the number of experimental possibilities in vitro.

Here, we aim to review the state-of-the-art electrophysiological and optical toolsets for in vitro experimentation and summarize the most recent discoveries in axon function that they have enabled. We will also review several microtechnologies that have been developed for controlling axon guidance in vitro and that can be readily combined with these toolsets. Our aim is not to provide an overview of the various biological processes involved in axon electrophysiology or guidance but instead to highlight fundamental methodological contributions that have increased our understanding of the axon. Finally, we provide an outlook of future research directions and outline our views on the prospects of combining technologies that promote axon guidance and enable the simultaneous recording of axon electrophysiology.

Measuring Axonal Electrophysiology

Classically, the study of axon function relied heavily on electrophysiological techniques which were not adapted to the study of the thin and complex arborizations of mammalian neurons. However, recent technological developments are driving a new era of experimental opportunities. Here we will address the most relevant: subcellular patch clamp, microelectrode arrays (MEAs), and functional imaging techniques. A schematic overview of these different techniques is shown in Figure 1.

Figure 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 1.

Methods for probing axonal activity. Schematic representation of the main experimental approaches to probe axonal activity and the respective stimulation and/or recording capabilities. The labels are merely illustrative and may not represent the signals/waveforms characteristic of each technique.

These approaches enable the probing of axonal activity on a variety of scales. Subcellular patch clamp allows for stimulation and extracellular-like (axon-attached configuration) or intracellular recordings (whole-cell configuration) from a particular point of a single axon. Standard MEAs can be adapted for the extracellular recording and stimulation of multiple points of multiple axons, while high-density MEAs can be used for the electrical mapping of axonal arbors. Functional imaging techniques (e.g., voltage imaging) can read out axonal arbors’ activity with the highest spatial resolution but, typically, at the cost of temporal resolution. Adding to this diversity of techniques, optogenetics is rapidly emerging as an optical tool for the manipulation of presynaptic terminals. A comparison of the key characteristics, advantages, and limitations of these different techniques is summarized in Table 1.

View this table:
  • View inline
  • View popup
Table 1.

Comparison of key characteristics, advantages, and limitations of different techniques for axon electrophysiology

Subcellular patch clamp

For a long time, patch clamp has been the gold standard for studying neurophysiology at single-cell or even single-ion channel resolution (Hamill et al., 1981). Typically, a recording micropipette is pressed against a patch of the membrane to form a strong seal resistance, which allows high-fidelity recordings in the cell-attached or whole-cell (if the patch is ruptured via suction, gaining intracellular access) configurations. However, most studies are limited to large compartments of the neuron (usually the soma) due to their easier accessibility.

Until recently, direct patch-clamp recordings from intact mammalian axons were limited to specific “giant” axonal structures (3–5 µm; e.g., calyx of Held, mossy fiber terminals; Bischofberger et al., 2006), which comprise <1% of all CNS synapses. This approach showed that hippocampal mossy fiber boutons act as sites of presynaptic AP amplification (by ≈40 mV; Engel and Jonas, 2005). However, recordings from smaller structures could only be obtained upon severing the axon, since the resulting swollen ends (i.e., blebs) were much larger (3–6 µm) than the intact axon (Shu et al., 2006; Kole et al., 2008). For the study of axon physiology, this may be problematic, as axonal lesions lead to an alteration of the passive properties of signal conduction (e.g., membrane capacitance), aberrant reorganization of the cytoskeleton, and, potentially, ion channel redistribution. Still, dual soma-axonal bleb recordings in the whole-cell configuration were the basis for the identification of the AP initiation zone (Debanne et al., 2011) and provide a useful method for recording signal propagation along the axon (Kramer et al., 2020; Zbili et al., 2020; Rózsa et al., 2023).

Beyond the aforementioned cases, new methodological approaches have now been developed to implement subcellular patch clamp in both brain slices (Sasaki et al., 2012b; Vandael et al., 2021) and cell culture preparations (Novak et al., 2013), as detailed below.

Axon-attached

Axon-attached (or loose-seal patch) recordings from intact unmyelinated axons (<1 µm diameter) have been made feasible by fluorescence-guided subcellular patch clamp (Sasaki et al., 2012b). The key advantage of this technique is that both the axon and the recording glass pipette are fluorescently labeled, which greatly facilitates live optical targeting. In combination with conventional whole-cell somatic recordings (i.e., dual recordings), this method allows for assessing the fidelity of AP propagation. Still, this technique is limited to single-site short-term recordings (<1 h) of extracellular-like AP waveforms (µV range), which are around three orders of magnitude lower than those recorded intracellularly (mV range), and as with other extracellular-based techniques (e.g., MEAs), subthreshold events are not detected (Sasaki et al., 2012a,b).

Sasaki et al. pioneered the application of fluorescence-guided subcellular patch clamp and demonstrated that the local activation of AMPA receptors by glutamate released from periaxonal astrocytes caused axonal AP broadening in hippocampal CA3 neurons. Broadened APs increased synaptic transmission to postsynaptic neurons by inducing greater calcium influx in presynaptic boutons (Sasaki et al., 2011). Hamada et al. performed recordings from axon collaterals’ boutons to characterize the temporal fidelity of presynaptic APs in a cuprizone-induced demyelination model. This approach revealed reduced conduction velocity, presynaptic AP broadening, and frequency-dependent AP failure in demyelinated L5 pyramidal neurons (Hamada et al., 2017). Recently, Lang-Ouellette et al. performed dual loose-seal patch recordings of young Purkinje cell somata and axons, which led to the surprising finding that the formation of axonal swellings (normally associated with axon degeneration) increases the fidelity of high-frequency AP propagation (Lang-Ouellette et al., 2021). This discovery suggests that axonal swellings in healthy states can represent a homeostatic form of axonal plasticity.

Dual somatic and axon-attached recordings have also revealed that the distal axons of CA3 pyramidal neurons (>600 µm from the soma) can fire at much higher rates (≈4–5 times) than the corresponding somata during gamma oscillations (Dugladze et al., 2012). These distally generated APs (i.e., ectopic APs) do not depolarize the soma due to the action of inhibitory axo-axonic cells that target the AIS. This implies a functional separation of axonal and somatic activity during this network behavior. Even though somatic recordings usually allow ectopic APs to be distinguished from APs initiated in the AIS (due to their characteristic waveform devoid of a somatodendritic component; Bähner et al., 2011), Dugladze et al. provide a remarkable example of how dual recordings can reveal insights unattainable with somatic recordings alone.

Whole-cell

Whole-cell recordings from varicosities (putative en passant boutons; Vivekananda et al., 2017; Ritzau-Jost et al., 2021; Vandael et al., 2021) or terminals (Vasylyev and Waxman, 2012; Kawaguchi and Sakaba, 2015; Zorrilla de San Martin et al., 2017; Vyleta and Snyder, 2023), as well as outside-out patch from the axon shaft (Hu and Jonas, 2014; Hu et al., 2018), have also been performed for several neuronal types. Novak et al. achieved a key technical breakthrough by combining subcellular patch clamp and super-resolution scanning ion conductance microscopy (<150 nm 3D resolution). This semiautomated approach uses a nano-pipette for identification, topographical imaging, and recording of presynaptic boutons and is compatible with all subcellular patch-clamp modalities in neuronal cultures (Novak et al., 2013).

Unlike in the axon-attached configuration, gaining intracellular access allows measurement of presynaptic AP shape, control of the cytoplasmic composition, and direct examination of biophysical properties of synaptic transmission. Several studies have now reported the facilitation of synaptic transmission via subthreshold depolarization of the presynaptic terminals (“analog–digital facilitation”; reviewed in Alpizar et al., 2019; Zbili and Debanne, 2019; Debanne et al., 2013), highlighting the importance of this type of axon signaling in neuronal function. Whole-cell electrophysiology is currently the only method capable of recording such subthreshold depolarizations. Paired recordings of presynaptic terminals and postsynaptic neurons are particularly well suited to measure synaptic strength and short-term plasticity. Still, obtaining the precise axonal AP waveform (one of the main advantages of conventional intracellular recordings) is compromised by technical difficulties. The necessary amplifier configurations and small pipette tip sizes (with increased resistance), adapted for very thin structures, inevitably distort axonal APs and require post hoc reconstitution (Oláh et al., 2021). Quartz glass pipettes (instead of conventional borosilicate) can be used to minimize this problem since they have lower resistance and capacitance at comparable sizes (Ritzau-Jost et al., 2021).

Recording an accurate AP shape is particularly important in the presynaptic terminals, where it modulates voltage-activated calcium currents and neurotransmitter-containing vesicle release. Direct recordings from boutons have allowed the dissection of Kv channels’ importance in presynaptic AP shape (Begum et al., 2016; Rowan et al., 2016; Vivekananda et al., 2017); the characterization of the presynaptic AP duration and amplitude in different cell types (Kawaguchi and Sakaba, 2015; Kramer et al., 2020; Ritzau-Jost et al., 2021); and the finding that presynaptic GABAA receptors mediate depolarization in Purkinje (Zorrilla de San Martin et al., 2017), dopaminergic (Kramer et al., 2020), and hippocampal mossy fiber neurons (Ruiz et al., 2010). Interestingly, Ritzau-Jost et al. demonstrated that presynaptic APs broaden during high-frequency firing in excitatory L5 pyramidal neurons, but not in inhibitory fast-spiking interneurons. In both cellular models, APs propagated reliably and at constant large amplitude (larger than at the soma) into axon collaterals, even during high-frequency firing (Ritzau-Jost et al., 2021).

Recently, two studies reported a nonconventional form of AP initiation in striatal dopamine axons which is independent of somatodendritic activity. Liu et al. and Kramer et al. performed direct recordings (whole-cell and perforated configurations) to reveal that these axons can initiate EPSPs and full-blown APs distally (near the synaptic terminals) in response to local acetylcholine release (Kramer et al., 2022; Liu et al., 2022a). Since presynaptic acetylcholine receptors are abundant throughout the brain, this finding raises the question of whether distal AP initiation is an important mechanism in physiological and pathological contexts beyond the dopaminergic system.

Overall, subcellular patch-clamp techniques, specifically the intracellular configurations, presently allow the highest possible fidelity in axonal recordings. Unfortunately, the technical complexity and invasiveness of these techniques limit the throughput and compatibility with long-term experiments. Furthermore, currently available intracellular methods are limited to relatively large structures of the axon (e.g., boutons). Critically, the inability to record from multiple sites along the axon simultaneously precludes subcellular patch clamp from tracking AP propagation throughout the axon arbor. Still, it is worth mentioning that several research groups have made efforts to automate patch clamping, which dramatically increased the throughput and accessibility of the technique in recent years (Suk et al., 2019; Koos et al., 2021; Ghovanloo et al., 2023). While these automated approaches are currently only applicable to somata, future endeavors may develop to study axons.

Microelectrode arrays (MEA)

MEA technology is at the forefront of technologies for recording electrical activity from large neuronal ensembles (reviewed in Obien et al., 2015; Xu et al., 2021). Typically, standard MEAs are composed of a fixed grid of multiple planar microelectrodes (most often 60–256 electrodes, interspaced by ≈30–500 µm), which are embedded in a transparent substrate (e.g., glass wafer) that functions as a cell culture vessel. After cell plating on MEAs, the activity of self-organized neuronal ensembles can be monitored and modulated (e.g., via stimulation) for over a year (Potter and DeMarse, 2001).

Currently, MEAs provide a versatile, noninvasive, and high-throughput functional assay at the network level. However, standard planar microelectrodes only allow for the recording of strongly attenuated extracellular signals (µV range) and are limited to the detection of suprathreshold activity (i.e., APs). Moreover, the recorded potentials are highly dependent on the positioning of the neuron relative to the electrode, and, typically, the microelectrodes only probe a very small fraction (<1%) of the whole substrate (Spira and Hai, 2013). These limitations make apparent the main reasons why conventional MEA technology alone is not adapted to the study of axon electrophysiology: (1) axons navigate the substrate freely; (2) the neuropil complexity hinders source–target identification (optically or electrophysiologically); (3) the SNR of axonal signals is so low that propagating extracellular APs can only be identified after averaging noise several times (Bakkum et al., 2013; Tovar et al., 2018; Abbott et al., 2020).

These limitations have recently been circumvented by either guiding axon growth over the electrodes (i.e., MEA and microfluidics combination, 3D-structured MEA) or dramatically increasing the spatial resolution (and the number of electrodes) of MEAs (i.e., high-density MEA), as detailed below.

MEA and microfluidics (μEF devices)

Several research groups, including ours, have merged the benefits of MEAs and microfluidics to attain simultaneous axon and network recordings (Pan et al., 2014; Lewandowska et al., 2015; Jang et al., 2016; Gladkov et al., 2017; Habibey et al., 2017; Hong et al., 2017; Gribi et al., 2018; Lopes et al., 2018; Moutaux et al., 2018; Shimba et al., 2021). We named this combination of microelectrodes and microfluidics μEF devices (Lopes et al., 2018).

Generally, microfluidics make use of subcellular microchannels to guide the axon to a separate culture compartment from the soma. μEF devices comprise tailor-made microfluidic devices (with matching microchannel-microelectrode spacings) to guide axons over multiple electrodes, allowing the study of signal propagation along multiple axons with high spatiotemporal resolution. To this end, we have developed specialized algorithms that enable, for example, the characterization of conduction direction and velocity within a user-friendly computational tool (Heiney et al., 2019). This combination creates advantages beyond axon guidance over electrodes. Conveniently, the microchannels’ small cross section creates an electrophysiological environment that increases the electrical coupling between axons and electrodes that greatly amplifies (by 1–2 orders of magnitude) the axonal signals (Dworak and Wheeler, 2009; Wang et al., 2012). It has been speculated that the amplitudes of axonal APs recorded in these confined settings (up to units of mV) may be large enough to modulate adjacent axons’ activity via ephaptic coupling (Pan et al., 2014; Narula et al., 2017), though no study has observed this phenomenon conclusively (Lewandowska et al., 2015). Importantly, the axon-specific compartmentalization allows selective treatment (e.g., of chemical blockers) and/or manipulations (e.g., axotomy), while recording network and axonal responses simultaneously (Habibey et al., 2015; Moutaux et al., 2018; Mateus et al., 2021; Shimba et al., 2021).

Probably due to the distinctive advantage of allowing for multisite, noninvasive (hence repeated and long-term) recordings with very high temporal resolution (up to 20 µs), most insights obtained via this combination so far have been related to conduction velocity. In particular, we have used μEF devices to reveal an increase in conduction velocity after the pharmacological inactivation of non-muscle myosin-II, an important protein in the structure of the axonal cytoskeleton. Super-resolution microscopy experiments showed that this inactivation leads to axon enlargement, providing one of the first demonstrations that axon cytoskeleton dynamics may act as mechanisms for regulating conduction velocity (Costa et al., 2020). Other research groups have used μEF devices to characterize changes in conduction velocity during network maturation (Habibey et al., 2017; Hong et al., 2017) or upon Nav1.2 blockade (Shimba et al., 2021).

Recently, μEF recordings revealed the unforeseen presence of ectopic/antidromic APs (Sheffield et al., 2011; Liu et al., 2022a) in hippocampal and sensory neurons. This activity was observed in basal conditions but could also be modulated by lesioning or chemical stimulation of distal axon portions (Mateus et al., 2021). Since these axons extended in empty microfluidic compartments (i.e., without synaptic targets), one may ask whether this form of subcellular initiation of electrical activity plays a role during development, when axons have not yet established synapses.

MEA and microfluidics can also be adapted to the study of explanted nerves. Gribi and colleagues used this “nerve on-a-chip” platform to characterize myelinated peripheral nerve fibers’ electrophysiology (Gribi et al., 2018) and test biomimetic stimulation strategies (high-frequency amplitude-modulated bursts) ex vivo (Formento et al., 2020).

While this combination of microtechnologies certainly provides unique benefits for the study of axon physiology, most studies have not moved beyond the proof-of-concept stage. Given its unique advantages, a wider adoption of μEF devices by the neuroscience community may help dissect the relevance of axonal electrophysiology in network function.

High-density MEA (HD-MEA)

A different type of MEA technology has also gained momentum in the study of axons (reviewed in more detail in Emmenegger et al., 2019). Complementary metal oxide semiconductor (CMOS)-based planar MEAs with higher spatial resolution than conventional MEAs have been used to probe axonal arbors at single-cell resolution in low-density (Bakkum et al., 2013, 2018; Radivojevic et al., 2017; Yuan et al., 2020) and high-density (Abbott et al., 2020) random cultures. These high-density MEAs (HD-MEAs) can pack up ≈7,000 electrodes/mm2 (up to ≈236.000 total) that can be read out simultaneously with high noise levels [active-pixel sensor (APS) architectures] or in subsets with low noise (switch-matrix architectures; Yuan et al., 2020; Suzuki et al., 2023). Until recently, due to SNR constraints, mapping axonal arbors was only possible following electrical stimulation of the axon and noise averaging of several stimulation trials (“stimulus-triggered averaging”; Bakkum et al., 2013). New-generation HD-MEAs, with improved SNR, now enable the mapping of axonal arbors based on their spontaneous activity (“spike-triggered averaging”; Abbott et al., 2020; Yuan et al., 2020).

HD-MEAs’ relatively high spatial resolution (≈0.25–20 µm) allows pure electrical mapping (“electrical imaging”) of the axon morphology (Abbott et al., 2020; Yuan et al., 2020; Radivojevic and Punga, 2023; Suzuki et al., 2023). This has allowed the discovery that the AIS, instead of the soma, is the main contributor to the extracellular potential (Bakkum et al., 2018). This electrical mapping is not limited to unmyelinated axons. Recently, Shimba et al. demonstrated the HD-MEAs’ capability to detect saltatory conduction along myelinated axons of sensory neurons (Shimba et al., 2022). In addition, Radivojevic et al. used HD-MEA technology to reveal that mammalian cortical axons conduct with high temporal precision (low jitter) and reliability (no conduction or branch-point failures) (Radivojevic et al., 2017). This finding demonstrates that, despite their intricate arborizations, mammalian axons seem to follow Rall's 3/2 power law describing the ideal geometrical ratio between mother and daughter branches to maintain propagation fidelity (Rall, 1959; Swadlow et al., 1980; Radivojevic et al., 2017).

A key disadvantage of HD-MEAs is their substrate opacity, which limits simultaneous probing with optical methods. Moreover, their recordings’ temporal resolution is, typically, lower than conventional MEA setups, thus prone to less precise measures of conduction velocity. Finally, while HD-MEAs can be combined with microfluidics for the control of axon guidance (Lewandowska et al., 2015; Duru et al., 2022), their nonplanar surface complicates the proper attachment of microfluidic devices. Still, these exciting new technological developments should accelerate the study of the conduction properties of single-cell axonal arbors.

3D-structured MEA

In recent years, many research groups have made efforts to improve the SNR of MEA recordings by fabricating 3D electrodes at the micro- or nanoscale (reviewed in Acarón Ledesma et al., 2019; Cho et al., 2021). In general, 3D-structured electrodes increase the electrical coupling between the electrode and the electrogenic compartment, which leads to an increase in the seal resistance at the electrode–membrane interface. The field was fueled by the discovery that large invertebrate Aplysia neurons spontaneously engulfed mushroom-like microelectrodes (Hai et al., 2010), which enabled the intracellular-like recording of APs and synaptic potentials (reviewed in Teixeira et al., 2020). Since then, intracellular-like recordings (units of mV) from the soma of mammalian neurons after spontaneous partial membrane engulfment (Shmoel et al., 2016) or after poration (Robinson et al., 2012; Dipalo et al., 2017; Liu et al., 2017; Abbott et al., 2019) have been demonstrated. However, due to their relatively small size, intracellular-like recordings from axons have yet to be achieved. Even the smallest reported 3D nanoelectrodes are in the same size order as mammalian axons (Cho et al., 2021); thus it is unlikely that this approach will soon allow for stable poration and probing of axonal structures. Although smaller 3D nanostructures (10–100 nm) can be fabricated, these are limited in recording capability due to the increased electrode impedance (Wu et al., 2019).

Even though the effects of topography on axon contact guidance are well known (Simitzi et al., 2017), surprisingly, the effects of 3D electrodes on network organization and axonal recordings have been largely neglected. A few studies have demonstrated the potential of incorporating discontinuous 3D features (e.g., nanopillars) for axonal guidance (Santoro et al., 2014; Amin et al., 2018; Mateus et al., 2019), but ultimately, these were either not integrated into MEAs or not capable of selective axon recordings. Theoretically, 3D MEAs can also be combined with microfluidic chambers, though the assembly procedure would need to be automated in a way that prevents damage to the fragile microstructures. Given their higher SNR, 3D-structured electrodes should allow for the reliable recording of axonal signals and contribute to the dissection of their importance in network function.

Interestingly, CMOS-based MEAs that integrate 4,096 3D-nanostructured electrodes have shown intracellular-like recording capabilities from thousands of somata simultaneously (in “pseudo-current-clamp” mode, although for limited time recordings), allowing for a partial connectome mapping (Abbott et al., 2019, 2020). In the future, the combination of higher spatial resolution HD-MEAs with 3D electrodes should allow for the extracellular probing of full axonal arbors, while recording somata intracellularly, hence gaining access to the subthreshold fluctuations (e.g., synaptic potentials) that precede and follow axonal conduction.

Functional imaging

The use of light to investigate axonal activity avoids the need for direct axon-electrode contact (i.e., subcellular patch clamp) or the close-apposition of extracellular electrodes (i.e., MEA). Functional imaging methods offer the advantage of probing axons at the highest possible spatial resolution, albeit at the cost of temporal resolution (essentially limited by the field of view/acquisition rate relationship; reviewed in Wang et al., 2019). The main optical methods for monitoring axon activity are calcium (Ca2+), sodium (Na+), and voltage imaging. These techniques can be considered passive, as they rely on sensors or indicators to measure ionic concentrations with no intended influence on the underlying physiological processes, although they may interfere in them (e.g., via calcium buffering, changes in membrane capacitance). Axonal optogenetic actuators, however, use light stimulation to manipulate ion fluxes and control neurotransmitter release.

Synaptic reporters are a different class of optical tools routinely used to measure the exocytosis of synaptic vesicles (e.g., pH sensors→pHluorin variants) and neurotransmitter release (e.g., glutamate via iGluSnFR)—events that are linked to the arrival of APs at the presynaptic terminals. These have been reviewed in depth elsewhere (Gobbo and Cattaneo, 2020; Jang et al., 2021) and are beyond the scope of this review.

One general limitation of functional imaging techniques is that the activity measured in each pixel may result from the superposition of multiple neurites or other electrogenic compartments. It is particularly difficult to resolve the axonal signal due to the limited surface area (hence few emitted photons). Moreover, the rapid kinetics of APs force very high acquisition rates (hence few contributing photons), which require ever more sensitive (i.e., larger dynamic range), bright (i.e., higher SNR), and photostable fluorophores.

Calcium imaging

APs reliably lead to Ca2+ influx (≈10-fold rise in intracellular concentration, which reverts to baseline within ≈100 ms; Koester and Sakmann, 2000); thus, Ca2+ imaging has long been used as an indirect proxy for neuronal activity (reviewed in Grienberger and Konnerth, 2012). Ca2+ signaling is particularly important in the presynaptic terminals, where it controls vesicle release probability. In fact, presynaptic Ca2+ entry and neurotransmitter release exhibit a nonlinear dependence, with small changes in the level of Ca2+ having a profound impact on quantal release (reviewed in Dolphin and Lee, 2020). Classical studies loaded neurons with Ca2+ indicators (e.g., BAPTA-1 dye) to image the Ca2+ dynamics at presynaptic terminals (Koester and Sakmann, 2000); however, the rapid progress in viral vector tools [particularly adeno-associated viruses (AAVs); Haggerty et al., 2020] has established genetically encoded Ca2+ indicators (GECIs) as the most prominent tool for imaging Ca2+ in neurons (Broussard et al., 2014). Most GECIs are engineered by fusing a fluorescent protein (e.g., GFP) to a Ca2+-sensing domain (e.g., calmodulin). When Ca2+ binds, the sensing domain undergoes a conformational change, resulting in a change in fluorescence.

Currently, GECIs are routinely used to assess neuronal activity at the network, cellular, and subcellular scales with high SNR (Dana et al., 2019). State-of-the-art GECIs allow detection of single axonal APs without averaging, if imaged with sufficiently high spatiotemporal resolution (Broussard and Petreanu, 2021; Huang et al., 2021). Thus, Ca2+ imaging can be used to monitor activity at levels ranging from multiple axons to single presynaptic terminals (or boutons; reviewed in Broussard and Petreanu, 2021). Since Cav channels are particularly concentrated at presynaptic structures, these regions allow higher SNR measurements than the axon shaft. Interestingly, the amplitude of Ca2+ transients can vary significantly across boutons of the same axonal branch in response to a single AP (Koester and Sakmann, 2000). Ca2+ imaging has helped demonstrate that presynaptic Ca2+ signaling can be modulated by several mechanisms beyond AP arrival, such as AP amplitude (Zbili et al., 2020), subthreshold signaling (Christie et al., 2010), and neuromodulation of Cav channels (Burke et al., 2018).

Recent advances in multicolor Ca2+ imaging facilitate simultaneous imaging of pre- and postsynaptic activity from differentially labeled neurons (Inoue et al., 2019). For the detection of fast Ca2+ transients, the sensors with faster responses are XcaMP-Gf (Inoue et al., 2019), jGCaMP7f (Dana et al., 2019), and the newest jGCaMP8 family (Zhang et al., 2023). The fastest sensor to date (jGCaMP8f) has reported half rise and decay times of ≈7 and ≈92 ms, respectively, which has greatly reduced the mismatch between Ca2+ sensor kinetics and actual Ca2+ transients in the axon. Still, it is important to note that untargeted GECIs preferentially label the somatodendritic compartment and diffuse poorly along the axonal arbor and thus are biased toward the detection of somatic APs (Knöpfel and Song, 2019; Broussard and Petreanu, 2021). To circumvent this limitation, several variants of GcaMP have been fused to presynaptic scaffolding proteins (e.g., synaptophysin) to enable presynaptic Ca2+ imaging (Dreosti et al., 2009; Zhang et al., 2018). These probes are known collectively as syGCaMPs and have allowed the dissection of different Cav channels’ contribution to presynaptic Ca2+ influx (Brockhaus et al., 2019). A red-shifted variant combined with pHluorin (SypHy-RGECO) has allowed researchers to directly relate presynaptic Ca2+ entry to vesicular release (Jackson and Burrone, 2016). For high SNR Ca2+ imaging of axonal arbors, an axon-targeted GECI (axon-GcaMP6) has been developed by fusing GcaMP6 to the growth-associated protein 43 (GAP43; Broussard et al., 2018).

Despite their ubiquitous use in today's research, some limitations in Ca2+ imaging approaches need to be considered. First, Ca2+ is a critical intracellular messenger; thus, Ca2+ concentration can change via cellular processes independent of firing activity. Moreover, the origin of axoplasmic Ca2+ transients in different portions of the axon is a matter of ongoing debate. While Ca2+ signaling at the presynapse is thought to be mediated by Cav channels (Dolphin and Lee, 2020), recent findings using high-speed Ca2+ and Na+ imaging have diverged regarding the origin of Ca2+ elevations in the AIS and nodes of Ranvier. Different studies have suggested that these AP-associated transients are mediated by Nav channels (Hanemaaijer et al., 2020; Filipis et al., 2023) or by Cav channels (Lipkin et al., 2021). Furthermore, Ca2+ sensors are (inherently) not sensitive to hyperpolarizing activity, and debate continues regarding whether they can faithfully report the whole extent of suprathreshold activity (especially in high-firing-rate conditions; Huang et al., 2021). Finally, it is important to consider that Ca2+ sensors need to sequester free Ca2+ to produce a signal, thus acting as Ca2+ buffers that reduce the transients’ magnitude and rate of change. In hippocampal axons filled with fura-2 AM, such buffering action has been estimated to produce 20-fold changes relative to no-sensor conditions (McMahon and Jackson, 2018). The effect of these perturbations in intracellular Ca2+ signaling is largely unknown and rarely acknowledged. Still, proper calibration of the setup against test samples of known concentration can lead to experiments that minimally interfere with endogenous signaling (Cherkas et al., 2018).

Sodium imaging

Sodium currents underlie key processes such as AP generation and conduction. High-speed fluorescence Na+ imaging allows precise recording of Na+ currents; thus, in the context of axon electrophysiology, it has primarily been used to characterize the spatiotemporal pattern of AP initiation in the AIS. Sodium imaging relies on dye indicators that exhibit an increase in fluorescence emission (and a slight shift in wavelength) upon binding Na+. In Na+ imaging, only a small fraction of Na+ binds to the indicator; thus, a linear relationship exists between the fluorescent change and the total Na+ transient for a wide range of indicator concentrations (Filipis and Canepari, 2021). This contrasts with Ca2+ imaging, where calibration of Ca2+ influx needs to consider binding to indicators and endogenous buffers.

In most central neurons, the AIS functions as a lower-threshold AP initiation zone, presumably due to the local high density of specialized Nav channel isoforms (Bean, 2007; Kole and Stuart, 2012), especially the Nav1.6 (in myelinated axons) and Nav1.2 (in unmyelinated axons; Kole et al., 2008; Hu et al., 2009; Lorincz and Nusser, 2010). These specialized channels mediate the rapid Na+ influx underlying AP initiation, and their voltage dependence for both activation and inactivation are shifted by ≈10 mV compared with their counterparts in the soma (Kole et al., 2008; Hu et al., 2009). Na+ imaging has been used to characterize the fast-activating and fast-inactivating transient Na+ currents associated with the AP, as well as subthreshold and persistent Na+ currents that originate from the AIS (Filipis and Canepari, 2021; Shvartsman et al., 2021).

Despite a consensus that Nav density is higher in the AIS than elsewhere in the neuron, the order of magnitude of this difference is yet to be determined and its importance in AP initiation is not clear (Colbert and Johnston, 1996; Kole et al., 2008; Baranauskas et al., 2013; Katz et al., 2018; Lazarov et al., 2018). Findings obtained via Na+ imaging of L5 pyramidal neurons have shown that the AP-associated Na+ flux is ≈3-fold and ≈8-fold higher in the AIS than in the soma and basal dendrites, respectively [suggesting a ≈3 (axon):1 (soma):0.3 (dendrite) Nav density ratio; Fleidervish et al., 2010]. However, thought-provoking experimental and theoretical evidence has suggested that high Nav density in the AIS is not required for AP initiation in the axon but instead is crucial for precise AP timing (Lazarov et al., 2018). In the study, even when axonal Nav density of hippocampal neurons decreased below somatic levels (through a mutation of the ankyrin G anchor βIV-spectrin), APs were still initiated in the axon. Future studies using Na+ imaging may help reveal where and how exactly APs initiate in analogous channelopathies.

Sodium imaging studies have also helped reveal a spatial mismatch between the highest Na+ flux, which occurs in the middle of the AIS, and the actual AP trigger zone (distal AIS; Baranauskas et al., 2013). This finding highlights the importance of axon morphology (i.e., lower capacitive load) and electric isolation from the soma in AP initiation, beyond Nav density. In another study, Katz et al. used Na+ imaging to demonstrate that Nav1.6 deficiency (Nav1.6-cKO) did not keep APs from being initiated in the distal AIS of L5 pyramidal neurons, as this loss was compensated by Nav1.2 channel expression (Katz et al., 2018), putting into question the role of particular Nav subtypes in AP initiation.

Since Na+ kinetics are particularly fast, an acquisition rate of ≈10 kHz is estimated to be necessary to characterize fast transients (Popovic et al., 2011). This acquisition rate is technically demanding; thus, researchers have interfered with Na+ channels’ kinetic properties (by lowering the experimental temperature to ≈21°C) to image the whole AIS at 0.5–2 kHz (Baranauskas et al., 2013; Katz et al., 2018). Recently, Filipis and Canepari reported 10 kHz recordings near physiological temperature (32–34°C) along ≈40-μm-long AISs (Filipis and Canepari, 2021; Filipis et al., 2023). Besides recording equipment difficulties, presently available indicators have limitations in sensitivity and SNR that prevent higher acquisition rates. Typically, these Na+-binding dyes (e.g., SBFI, ING-2) are loaded intracellularly in a ≈0.5–2 mM concentration range that allows a 0.5–5% change of fluorescence per AP after averaging several trials. Higher dye concentrations (that could lead to better SNR) are not well tolerated and alter electrophysiological properties (Filipis and Canepari, 2021). Thus future improvements of the Na+ imaging technique depend critically on the development of novel Na+ indicators.

Voltage imaging

Voltage-sensitive dyes (VSDs) and genetically encoded voltage indicators (GEVIs) have the potential to overcome some of the limitations of Ca2+ and Na+ imaging, because the changes in fluorescence are a direct proxy for changes in voltage (reviewed in Peterka et al., 2011). These indicators’ responses to voltage change can be linear (e.g., opsin-based indicators) or nonlinear (e.g., ASAP-based indicators), with linear indicators having generally faster on/off kinetics (time constant <1 ms), but less sensitivity (low %/mV). Consequently, linear and nonlinear indicators are more suitable for the recording of AP (fast kinetics but large voltage change) and subthreshold (slow kinetics but small voltage change) activity, respectively. Still, the conversion of fluorescence into absolute membrane potential (Vm) requires calibration via simultaneous current clamp (Cohen et al., 2020).

Early studies using VSDs revealed multiple AP initiation zones in invertebrate neurons (Zečević, 1996). Later studies with higher-sensitivity dyes (e.g., JPW3028, membrane potential imaging with a ≈11 µs delay) allowed determination of the location (20–40 µm from the soma) and length (≈20 µm) of the AP trigger zone in L5 pyramidal neurons, as well as the reliability of propagation along the axonal arbor, even at very high firing frequencies (≈400 Hz; Popovic et al., 2011). However, concerns regarding VSDs’ effects on the physiological properties of the cellular membrane (Peterka et al., 2011), as well as their lack of neuronal specificity, have stalled their use.

GEVIs circumvent the need for intracellular dye loading to achieve neuronal specificity via targeted expression and thus are currently the preferred tool for probing voltage changes in the axon. Still, due to their comparatively lower brightness, most studies need to average several stimulation trials to distinguish axonal Aps (Hoppa et al., 2014; Cho et al., 2017, 2020; Ma et al., 2017) and subthreshold activity (Rowan and Christie, 2017) from noise levels. Nevertheless, Cho et al. used QuasAr (an opsin-based GEVI) to reveal a novel role for Navβ2 subunits in the prevention of conduction failures at branching points (Cho et al., 2017). And recently, new-generation GEVIs (e.g., opsin-based Archon1 and 2, FRET-opsin Ace variants) have displayed the necessary kinetics and brightness to resolve the axonal AP in single trials (reviewed in Panzera and Hoppa, 2019), although fast photobleaching remains problematic (Bando et al., 2019). Furthermore, voltage imaging with Ace2N-mNeon at ≈3 kHz has been used to assess the direction and velocity of AP conduction, as well as to compare the AP waveform between different axonal segments (proximal vs distal) during Kv blockade (Gonzalez Sabater et al., 2021).

Due to the very high spatial resolution, voltage imaging is particularly useful for the study of voltage changes in structures difficult to probe with electrophysiological techniques, such as the axon shaft and presynaptic terminals. Several studies have used voltage imaging to help unveil the contribution of Kv channels to the presynaptic AP width and amplitude (Hoppa et al., 2014; Cho et al., 2020; Gonzalez Sabater et al., 2021). For example, Kv3 channels were demonstrated to cluster at boutons of stellate interneurons and influence presynaptic AP width with synapse specificity, even within the same axonal branch; this finding was validated with both voltage imaging and multisite bouton-attached patch clamp (Rowan et al., 2016).

Still, the adoption of GEVIs has so far been restricted by several limitations such as phototoxicity, photostability, low SNR, and poor performance under 2-photon microscopy. Moreover, depending on the sensor expression levels and intrinsic properties, GEVIs have been shown to increase membrane capacitance, reducing EPSP amplitude and increasing AP initiation thresholds (Akemann et al., 2009). Thus, the major challenge of voltage imaging is to develop sensors and methodologies that improve sensitivity, photostability, kinetics, and brightness, while minimally interfering with axon physiology. Recently, Liu et al. identified JEDI-2P as a 2-photon-compatible GEVI capable of reporting voltage transients of Drosophila interneurons’ axon terminals (Liu et al., 2022b). This and further developments have the potential to revolutionize how axonal activity is measured.

Optogenetics

Optogenetic actuators are gaining considerable momentum in the all-optical study of neuronal function (Hochbaum et al., 2014). These tools comprise genetically encoded photosensitive proteins (i.e., opsins) which can be used to control neuronal activity in a light-dependent manner via ion influx or efflux. Traditionally, their application has focused on activity control at the somatodendritic level, even though most opsins are not targeted to specific subcellular compartments. A notable exception is that of light-gated Cl− channels (e.g., GtACR2), intended for inhibition of neuronal activity, which require somatic targeting due to the antagonistic effect of Cl− influx in the soma (inhibition) and the axon (excitation; Mahn et al., 2018; Messier et al., 2018). Subcellular targeting of opsins to the axon is less common, but it is rapidly emerging as a way to directly control neuronal output and manipulate neurotransmitter release within the target areas of projection neurons (reviewed in Rost et al., 2022).

Understandably, the AIS is of particular interest for activity control by optogenetic actuators. However, attempts at subcellular targeting of opsins to the AIS have failed to meet expectations. For example, targeting channelrhodopsin-2 (ChR2) to the AIS, via an ankyrin G-binding domain of Nav channels, disrupts the local endogenous Nav clusters or fails to elicit APs (Grubb and Burrone, 2010b; Zhang et al., 2015). The presynaptic optogenetics toolbox, however, has greatly expanded and has proven useful in eliciting and inhibiting neurotransmitter release for different applications (Liu et al., 2019; Oldani et al., 2021).

Most axon-targeting actuators rely on protein fusion to synaptophysin, a protein abundant on both GABA-containing and glutamate-containing vesicles, with no perceived negative effect on synaptic function. Actuators that target specifically the presynaptic axonal plasma membrane are also emerging. For instance, Hamada et al. recently combined ChR2 with the mGluR2 C-terminal domain for presynaptic enrichment (S. Hamada et al., 2021). Optogenetic manipulation of presynaptic function represents an excellent tool for studying the role of individual synapses in information processing.

Still, it is important to note that presynaptic targeting of opsins is not a requirement for revealing new roles of axons via optogenetics. Recently, Hari et al. used GAD2CreER mice to express ChR2 (excitation) in GAD2+ neurons (axoaxonic GABAergic neurons that project onto myelinated proprioceptive sensory axons). This tour de force work demonstrated that, contrary to what was previously believed, activation of GAD2+ neurons does not result in presynaptic inhibition but rather prevents AP propagation failure at branch points by activating GABAA receptors at nodes of Ranvier (Hari et al., 2022). This study challenges, once again, our basic understanding of axonal GABA and shows a, potentially generalizable, way (i.e., nodal facilitation) of gating mediated by axo-axonic neurons.

Technique combinations

Virtually all the described techniques can be combined to extract more information from a single experiment. The combination of electrophysiological and optical techniques, in particular, allows probing of axon subcompartments’ morphological and functional traits with superior spatiotemporal resolution. All-electric (e.g., MEA and patch clamp) and all-optical (e.g., voltage and Ca2+ imaging) approaches can also be used complementarily, even though they are less common.

Probably the most used combination of techniques in the study of axonal activity is that of patch clamp and functional imaging dyes. Typically, patch pipettes are first used for filling neurons with the sensor of interest. Then, intracellular recording and stimulation of the soma are used as a “ground truth” for spiking activity, while spatiotemporal information (e.g., varicosities’ depolarization duration) is extracted from functional imaging. Sasaki et al. used a Ca2+ dye (BAPTA-1) to measure the dependence of AP-evoked Ca2+ elevation, in varicosities, on the axonal length and number of crossed branch points (Sasaki et al., 2012a). Recently, Zbili et al. used the same principle to demonstrate that the amplitude of the varicosities’ Ca2+ transients was increased when APs were evoked by a synchronous-like input (i.e., a depolarizing pulse directly from the resting Vm). This suggests that synaptic transmission may be facilitated when the presynaptic AP originates from synchronous inputs (Zbili et al., 2020). Dual soma–axon patch-clamp recordings combined with Na+ imaging revealed homeostatic changes (e.g., distal relocation of Na+ flux) in the AP trigger zone by acute blockade of M-type K+ channels (Lezmy et al., 2017). Cohen et al. performed voltage imaging of myelinated axons at 20 kHz to characterize saltatory conduction in the spatiotemporal domains. These optical recordings were voltage-calibrated via somatic patch clamp and revealed the existence of a periaxonal space that supports AP “saltation”—high-amplitude potentials at nodes of Ranvier precede attenuated potentials traveling in the internode (Cohen et al., 2020).

MEA electrophysiology has the benefit of allowing concurrent, noninvasive recording/stimulation and imaging of the axon. Even though simultaneous MEA electrophysiology and Ca2+ imaging of axon dynamics have been shown to be feasible (Moutaux et al., 2018; Mateus et al., 2021; Xue et al., 2022), new insights from this combination remain to be obtained. The recent development of fully transparent MEAs, compatible with high-resolution microscopy (Middya et al., 2021), should expedite the adoption of this technique combination.

All-optical approaches have the potential to allow monitoring and modulation of axonal subcompartments at will, with minimal invasiveness, but so far their use has been limited. The advent of newly engineered opsins that target the axon may enable new strategies where electrode-free setups can stimulate/inhibit (via optogenetics) and measure activity (via functional imaging) of targeted regions of the axon. Cho et al. coexpressed QuasAr and GCaMP6f to measure the shaft depolarizations and presynaptic Ca2 transients caused by single APs within the same axonal arbor (Cho et al., 2020). Lipkin et al. used Na+ and Ca2+ dyes to characterize AP-evoked Ca2+ transients and the distinct roles of Cav channels in the AIS (Lipkin et al., 2021). Theoretically, nonlinear and linear voltage indicators may also be combined to assess, respectively, sub- and suprathreshold activity (Panzera and Hoppa, 2019), although no study has so far reported such a combination.

All-electric combinations may allow precise control of Vm (via whole-cell patch clamp) and mapping of axonal arbor conduction (via HD-MEAs) concurrently. However, HD-MEAs are considerably harder to combine with other techniques due to their substrate opacity, which precludes experimental setups comprising inverted microscopes and hinders optical microscopy with upright microscopes. Still, HD-MEA recordings have been combined with somatic whole-cell patch-clamp recordings (Jäckel et al., 2017). In that study, somata could be patched by imaging the cultures with an upright microscope equipped with differential interference contrast. By taking advantage of the concurrent recording capabilities, this all-electric approach allowed the mapping of EPSPs evoked by stimulating the patched neuron. As this combination has the highest temporal resolution in axonal recordings, spatial clamping can be used to study propagation jitter and how it influences efficient information transmission. It is important to note that the introduction of propagation jitter may impact the coincidence detection of multiple EPSPs in the postsynaptic neuron, impacting synaptic transmission strength.

Controlling Axonal Guidance

The first step for analyzing axonal function is to identify/isolate its structure and growth path. In complex networks, such a task becomes daunting, because axons, dendrites, and glial processes form very intricate arbors within the neuropil. Still, this task can be simplified by patterning axon growth in vitro. Researchers have long taken advantage of the axon's growth-cone motility and cue-sensing capacity to control axon guidance in vitro (for historical perspectives, see Dupin et al., 2013; Roy et al., 2013; Raj et al., 2021). This control streamlines the study of axonal function by isolating axons and allowing the entire axon length (from the soma to the terminals) to be tracked.

A wide range of patterning techniques has been employed to control axon guidance (Raj et al., 2021). The most used can be divided into physical (e.g., microfluidics) and chemical (e.g., surface functionalization) patterning methods. Physical-based guidance strategies do not distinguish between axon, soma, and dendrites per se but may use topographical features and principles of size restriction to isolate the axon. Chemical-based strategies can use axon-specific permissive or repelling cues to control axon guidance more precisely. Both strategies take advantage of the growth-cone sensitivity to direct axon growth (i.e., taxis) and can be merged for enhanced effect. For example, growth-permissive chemical gradients (i.e., chemical patterning) can be established along microfluidic channels (i.e., physical patterning) to specify axon growth to target regions.

The most used methods for in vitro patterning are engineered substrates, microfluidics, and microcontact printing, which typically define fixed patterns for subsequent axon growth. These patterns can be regarded as static because they are created before cell seeding and cannot be easily changed thereafter. Recent developments, however, have used photo-reactive molecules and substrates that allow manipulation of axon guidance by creating and/or modifying paths in an in situ manner (Broguiere et al., 2020). Schematic representations of these methods are displayed in Figure 2.

Figure 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Figure 2.

Methods for controlling axon guidance. Schematic representations of approaches to control axon guidance via physical (top) and chemical (bottom) patterning. Detailed explanations of each method can be found in the respective subsection [engineered substrates; microfluidics; microcontact printing (µCP); in situ techniques]. For each method, the nature of the guiding feature (continuous; discontinuous, or custom) and possible direction for axon growth (unidirectional; bidirectional; multidirectional, or custom) are described.

The next sections detail the most recent developments in axon patterning in vitro. These methods are compatible with conventional neuronal culture methods, concurrent live imaging, and electrophysiology techniques and thus are amenable to high-throughput probing of axons.

Physical patterning

Engineered substrates

During development, axon pathfinding partially depends on the physical features of the environment (“topotaxis”), as well as the existence of guidepost cells where axons make turning decisions. The axon growth cone, via filopodial interactions, senses features (down to the order of tens of nanometers) to guide axon growth. Taking advantage of the recapitulation of this contact-mediated guidance in vitro and the advances in micro- and nanofabrication, researchers have engineered substrates that promote axon guidance. The outcome of these axon–substrate interactions can vary across neuron origin or age, but also across the topographical feature dimension, geometry, material, stiffness, and chemical properties (reviewed in greater detail elsewhere, Marcus et al., 2017; Simitzi et al., 2017; Leclech and Villard, 2020; Raj et al., 2021).

Different guiding features can be broadly separated into continuous (e.g., grooves, gratings) or discontinuous (e.g., an array of pillars, pits). Most of the structured substrates presented in the literature can also be considered anisotropic since they impose two (unidirectional) or more (multidirectional) symmetrical orientations (Leclech and Villard, 2020). Pure isotropic structures (e.g., nanowires) do not impose a direction on axon outgrowth but can influence aspects such as axon branching (Gautam et al., 2017; Seo et al., 2018). Continuous unidirectional features tend to increase the outgrowth rate and decrease the complexity of the branching pattern of the axon (Chua et al., 2014; Li et al., 2015), while discontinuous multidirectional features tend to increase the number of branches, which themselves are guided by the topography (Gautam et al., 2017; Seo et al., 2018; Milos et al., 2021).

Numerous studies have demonstrated that axons from different neuronal types and origins elongate along discontinuous or continuous topographical features in the direction of the pattern. The most pronounced effects are usually achieved via continuous structures, such as grooves with widths and depths in the µm scale (Chua et al., 2014). Recently, human-derived motor neurons’ axons were shown to grow over 1 cm along linear microgrooves (Hagemann et al., 2022). Discontinuous topographies, such as arrays of micropillars, can also promote axon guidance (Micholt et al., 2013; Park et al., 2016; Milos et al., 2021) but in this case, the critical parameter for strong guidance is the spacing between features, with ≈0.5–3 µm being optimal (Simitzi et al., 2017). These guidepost-like structures are of particular interest, as they can be adapted for the fabrication of 3D devices capable of recording neuronal activity with high-fidelity (e.g., MEA; Abbott et al., 2019; Mateus et al., 2019).

Microfluidics

Microfluidic devices follow a different rationale, which was pioneered in the 1970s with Campenot chambers (Campenot, 1977). Then, and now, the objective was to isolate axons in a separate culture compartment. Campenot's concept together with advances in microfabrication allowed a key breakthrough ∼20 years ago (Taylor et al., 2003, 2005). This microfluidic device comprised two compartments where neurons could be plated, interconnected by straight, high-aspect-ratio microchannels, into which only neurites could grow. If the microchannels were longer than 450 µm, only axons were able to cross the entire microchannel (Taylor et al., 2005). Thus, it allowed the spatial and fluidic compartmentalization of the somatodendritic and axon subcompartments. This simple, yet effective, design is widely used to date (reviewed in Neto et al., 2016; Holloway et al., 2021).

Several asymmetric microchannel geometries (e.g., arrowheads) that promote unidirectional axonal outgrowth have since been proposed in the literature to engineer neuronal circuits in vitro with defined connectivity (Peyrin et al., 2011; Forró et al., 2018; Holloway et al., 2019). These can be used to recreate specific circuits (e.g., hippocampus’ trisynaptic pathway), network motifs (e.g., feedforward excitation), or logical operators (e.g., “AND gates”; Feinerman et al., 2008), which rely on unidirectional connectivity. Microfluidic devices take advantage of soft lithography for replica molding the desired pattern using biocompatible silicones, typically polydimethylsiloxane (PDMS; Duffy et al., 1998). Besides neuronal compartmentalization, microfluidics’ advantages include the usage of fewer reagents and sample amounts, low cost, and increased high-throughput potential when compared with traditional in vitro systems.

Although most neuroscience applications make use of passive microfluidic systems, flow can be driven by capillarity, hydrostatic pressure differences, or external pumps. Hydrostatic pressure differences are routinely used to establish chemical gradients along the microchannels (via diffusion) or to treat target compartments (Park et al., 2006). Precise control of the fluid flow rate and angle can also be used to, for example, induce mechanical stress on the axon, leading to axonal injury (Pan et al., 2022).

Due to their versatility, microfluidic devices are currently the main tool for the structuring of “brain-on-a-chip” platforms (Holloway et al., 2021; Nikolakopoulou et al., 2021). Importantly, they allow isolation of large numbers of axons’ distal ends, which are easily accessible for targeted chemical or mechanical manipulations or coculturing with different cell types (e.g., glia). This characteristic has been used extensively in studies of axon development (reviewed in Millet and Gillette, 2012; Fantuzzo et al., 2019), regeneration (Hur et al., 2011), and degeneration (reviewed in Kim et al., 2012), among others. The study of axonal transport, in particular, has benefited greatly from microfluidic devices’ application. For example, Akin et al. used microfluidics, in combination with advanced live imaging modalities and new fluorescent tags, to demonstrate how different ion channels are produced in the soma, trafficked to the axon, and, finally, inserted into the axon membrane of sensory neurons (Akin et al., 2019, 2021).

We have recently shown that the limits of subcellular patterning using microfluidic compartmentalization can be pushed down to the nanoscale. By using a combination of electron beam lithography and photolithography, we were able to fabricate features ranging from 150 nm (nanochannels) to a few millimeters (microchannels). This allowed control of both axon and dendritic spine growth and thus the establishment of synapses in user-defined regions (Mateus et al., 2022). Such a degree of control over synapse formation should facilitate the study of synapse maintenance and plasticity in long-term experiments, as well as the consequences in information transmission.

Chemical patterning

Microcontact printing (µCP)

Chemical patterning takes advantage of the growth cone's sensitivity (“chemotaxis”) to substrate-bound (e.g., proteins) and/or soluble chemical cues [e.g., nerve growth factor (NGF)] for controlling axon guidance. Classical approaches (e.g., Bonhoeffer stripe assay) of chemical patterning helped to identify several attractive and repulsive molecules for axon guidance. Currently, microcontact printing (µCP) is the most used technique for substrate patterning, while soluble cues are often used to establish chemical gradients within microfluidic devices (reviewed in Aebersold et al., 2016; Raj et al., 2021).

Most often, in µCP, replica-molded PDMS stamps with microscale features are initially inked in a coating solution. Then, the stamps are used to imprint the desired patterns on a cell culture substrate. Upon plating, neurons tend to adhere to or migrate toward the areas that promote cell adhesion and then extend within the patterned surface (Kleinfeld et al., 1988). Like engineered substrates, µCP can be used to create continuous (e.g., lines) or discontinuous (e.g., spots) patterns that either create paths for axon growth (Vogt et al., 2005) or promote axon collateral branching (Kim et al., 2014). µCP can be used to control axon guidance more precisely, via the use of both promoting and repelling cues (Oliva et al., 2003; Weydert et al., 2019) or by the application of geometrical constraints (Roth et al., 2012).

Like microfluidics, µCP allows the isolation of large numbers of axons, with the advantage of easier axon accessibility due to the absence of a physical constraint (i.e., microchannel). Still, it is important to note that µCP alone does not establish fluidic compartmentalization, nor is it compatible with long-term control of axon guidance, as neurons typically start extending beyond the micropatterned surface after a few weeks in culture (Aebersold et al., 2016). For more reproducible and long-term control, physical barriers need to be considered.

In situ techniques

In recent years, several techniques have enabled the in situ patterning of mammalian axon outgrowth with great spatiotemporal resolution. These methods provide localized sources of guidance cues or provide paths for axon outgrowth on-the-fly, and thus they are not dependent on nano-/microfabrication of predefined patterns.

The classical micropipette assay initiated the in situ studies of axon guidance and has been critical in the discovery of chemical cues and associated intracellular pathways. In this method, a micropipette is placed close to a growth cone to create a chemical gradient. This allowed the pioneering discovery that sensory neurons’ axons are attracted to NGF (Gundersen and Barrett, 1979). Photorelease techniques, such as optical uncaging of caged molecules (Ellis-Davies, 2007) and photolysis of encapsulated guidance factors (Pinato et al., 2012), follow the same principle and provide interesting alternatives for creating local sources of guidance with no physical perturbation. However, these techniques share a limitation: they are not amenable to high-throughput experimentation. Typically, a single (and isolated) growth cone is exposed to the local cue each time (Dupin et al., 2013).

Emerging light-based techniques such as stereolithography or two-photon lithography are capable of additive manufacturing, material polymerization, or material ablation/etching in 2D and 3D. These techniques make use of photosensitive biomaterials (e.g., photoresins) to laser pattern with high precision (limited by the point spread function of the system) paths for axon growth. Broguiere et al. used two-photon patterning to dynamically control the growth of axons over 3D patterns of NGF (Broguiere et al., 2020). A different rationale was followed by Hong and Nam, who used photothermal etching to melt an axon-repulsive agarose hydrogel and allow axon outgrowth and manipulation of network connectivity (Hong and Nam, 2020). These proof-of-concept works demonstrate new avenues for in situ manipulations of axon guidance with high throughput. The maturation of these techniques and combination with functional readouts may lead to new insights into axon function. For example, given the degree of control over axon trajectories, these techniques can be used to study forms of direct axo-axonal coupling, such as electrical synapses (i.e., axonal gap junctions) or chemical synapses (e.g., GABAergic terminals onto the AIS of principal neurons), that are difficult to study in vivo.

Other in situ methods such as the application of electric fields have long been employed to guide axon growth in avian (Pan and Borgens, 2012) and amphibian (Patel and Poo, 1982) models, but their efficiency on mammalian axons remains controversial (Kim et al., 2016). Moreover, due to their incompatibility with concurrent direct axon recordings, they are outside the scope of this review.

Discussion and Future Perspectives

As wisely stated by Edgar Adrian in 1932, “The history of electrophysiology has been decided by the history of electrical recording instruments.” Almost a century later, it is still experimentally challenging to record from axons and, particularly, from multiple axons simultaneously. Consequently, most studies are still based on data derived from recordings of somatic APs. Unfortunately, this emphasis on somatic APs introduces several biases: it assumes that suprathreshold activity recorded in the soma serves as a direct proxy for neuronal output; it presumes that signaling is not modulated along the axon; and it disregards the fact that the axon is a plastic compartment. From the previous sections, it should be evident that somatic recordings do not provide a complete picture of neuronal output. In fact, in certain network behaviors (e.g., gamma oscillations), distal axons can generate APs at much higher rates than recorded in their corresponding somata (Dugladze et al., 2012). Another striking case is that of substantia nigra dopaminergic neurons, in which dopamine release is often defined by dopamine and acetylcholine receptors in the axon terminals and independent from somatic APs (Berke, 2018; Kramer et al., 2022; Liu et al., 2022a). Ultimately, in some cases, somatic APs may represent only a fraction of a given neuron output. Besides unforeseen AP initiation mechanisms, axons are also highly dynamic and can change structural (e.g., AIS length or position; Grubb and Burrone, 2010a), morphological (e.g., axon diameter; Chéreau et al., 2017), or functional (e.g., conduction velocity, presynaptic AP shape) properties (Zbili et al., 2020) depending on the excitability levels. The significance of this extended computational repertoire has yet to be fully characterized, but it is increasingly clear that both single-neuron and circuit functions cannot be understood based on somatic recordings alone.

As discussed here, cutting-edge technologies have provided means to unveil and characterize AP initiation mechanisms (Liu et al., 2022a; Rózsa et al., 2023), structure–function relationships (Hodapp et al., 2022; Nascimento et al., 2022), voltage-gated channels’ implications on AP shape and release probability (Rowan et al., 2016), signal conduction properties (Radivojevic et al., 2017; Mateus et al., 2021), synaptic transmission biophysics (Vyleta and Jonas, 2014), or presynaptic plasticity (Chéreau et al., 2017; Vandael et al., 2020). Although these findings likely impact neural computations, it has been difficult to determine how they affect network function since the majority were obtained through recordings of single axons. In systems neuroscience, the number of neurons recorded by microelectrodes has followed a trend of exponential growth, doubling approximately every 7 years (Stevenson and Kording, 2011). Unfortunately, this trend does not directly translate to the number of recorded axons due to several intrinsic limitations (e.g., electrode–axon size mismatch). The integration of new microtechnologies into in vitro axon research, however, offers exciting prospects and can help dramatically increase the experimental throughput. Moreover, patterning methods and in situ techniques, in particular, can enhance the degree of control over the desired axon path to dissect axon physiology. HD-MEAs can be used to study multiple axons per culture with a spatiotemporal resolution compatible with inquiries on axonal conduction (Yuan et al., 2020; Radivojevic and Punga, 2023). μEF devices currently allow for the extracellular recording of hundreds of axons simultaneously (Moutaux et al., 2018; Mateus et al., 2021) and can be further improved to achieve single-axon resolution recordings via clever nano-microfluidic design. Nano-microfluidics can control neuronal growth down to the dendritic spine scale, paving the way for inquiries with single-synapse resolution (Mateus et al., 2022). These technologies can also be used to create testbeds amenable to the study of elusive forms of axo-axonal communication such as gap junctional (i.e., electrical synapses; Alcamí and Pereda, 2019) or ephaptic coupling (Han et al., 2020). Importantly, microfluidics’ compartmentalization can be leveraged to study axon interactions with other cell types (e.g., glia) in coculture configurations or the short- and long-term effects of insults (e.g., axon lesioning) and treatments (Hyung et al., 2021). Despite several studies highlighting the influence of gliotransmission in axon electrophysiology (Yamazaki et al., 2007; Sasaki et al., 2011; Deemyad et al., 2018), glia–axon interactions are particularly understudied and can benefit from this technique combination.

Ultimately, “on-chip” technologies provide a highly controllable environment for experiments that are time-efficient, highly informative, and in accordance with the 3R's principle (replacement, reduction, and refinement of animals used) given the high number of axons studied per experiment. Still, this approach has not permeated the neuroscience community as expected (Nikolakopoulou et al., 2021). This is probably because such commitment requires interdisciplinary research between researchers coming from various backgrounds, ranging from materials science to neuroscience. Hopefully, the establishment of standardized protocols (Lopes et al., 2018) and specialized companies, that offer “key-in-hand” solutions, will help fill this gap.

Despite their crucial contributions to neuroscience, conventional microelectrode technologies are not fundamentally different from the first electrodes used by Galvani (circa 1770). These technologies suffer from intrinsic limitations such as the need for a nonbiological interface with the axon; lack of cell-type specificity; a limited number of electrodes; adverse scaling of reduced electrode size and SNR; and limited capability to simultaneously record and stimulate neural activity. Optical probes can circumvent these limitations but not without shortcomings. Consequently, although the replacement of the “electron” by the “photon” has been anticipated for over a decade (Scanziani et al., 2009), it has not yet taken place. In axon physiology studies, optical sensing is particularly challenging due to, mainly, the very limited volume/membrane available to pack in reporters (consequently limiting the number of contributing photons). The development of fluorescent sensors tagged with axon-specific protein tags will enable more precise labeling of structures of interest compared with untargeted sensors or indicators. Still, it is important to note that this drawback may already be circumvented in vitro via axon guidance techniques that isolate axons.

Future endeavors will also benefit from bright red or far-red VSDs, GEVIs, or GECIs that can be spectrally separated from green markers or sensors (e.g., GCaMP family) to monitor properties associated with voltage and Ca2+ transients. These may be coupled with genetically encoded sensors of neurotransmitter release (e.g., iGluSnFR) for the dissection of synaptic transmission and facilitation. Importantly, axon-specific imaging allows for multiscale probing of axon transmission—from entire projection bundles to particular synapses. This strategically positions imaging techniques as the go-to tools for many research questions. Nevertheless, even though the future of the field will increasingly depend on optical approaches, a synergism between imaging and electrophysiology will continue to be crucial. For instance, the combination of electrophysiology and structural imaging techniques, such as super-resolution microscopy, is only beginning to reveal how nanoscale structural changes impact axon function (Chéreau et al., 2017; Costa et al., 2020).

Despite the early examples that invertebrate axons do not serve as simple transmission lines (reviewed in Swadlow et al., 1980), the inaccessibility of the mammalian axon long hindered attempts of dissecting its electrophysiology. Recently, cutting-edge techniques reignited the interest of the neuroscience community in axon physiology and have made possible several unexpected findings that carry significant functional implications. While a steady stream of technological developments is to be expected, we cannot anticipate what the future will teach us concerning axon physiology. It will ultimately take several key breakthroughs to understand how the computational repertoire of the axon influences network-level function and, possibly, behavior. To this end, future studies will need to go beyond in vitro experimentation; but current research can undoubtedly gain from such highly controlled environments. In particular, we suggest that the combination of state-of-the-art technologies to guide and probe multiple axons can shed light on how axon signaling coordinates network activity. Hopefully, these approaches will help us better understand not only axon physiology but also, ultimately, pathologies that arise from axonal dysfunction.

Footnotes

  • This work was partially funded by national funds through Foundation for Science and Technology (FCT), under the project PTDC/NAN-MAT/4093/2021. J.C.M. was supported by FCT (PD/BD/135491/2018) in the scope of the BiotechHealth PhD Program (Doctoral Program on Cellular and Molecular Biotechnology Applied to Health Sciences).

  • The authors declare no competing financial interests.

  • Correspondence should be addressed to P. Aguiar at pauloaguiar{at}i3s.up.pt.

SfN exclusive license.

References

  1. ↵
    1. Abbott J,
    2. Ye T,
    3. Krenek K,
    4. Gertner RS,
    5. Ban S,
    6. Kim Y,
    7. Qin L,
    8. Wu W,
    9. Park H
    (2019) A nanoelectrode array for obtaining intracellular recordings from thousands of connected neurons. Nat Biomed Eng 4:232–241. doi:10.1038/s41551-019-0455-7
    OpenUrlCrossRef
  2. ↵
    1. Abbott J,
    2. Ye T,
    3. Krenek K,
    4. Gertner RS,
    5. Wu W,
    6. Jung HS,
    7. Ham D,
    8. Park H
    (2020) Extracellular recording of direct synaptic signals with a CMOS-nanoelectrode array. Lab Chip 17:3239–3248. doi:10.1039/D0LC00553C
    OpenUrlCrossRef
  3. ↵
    1. Acarón Ledesma H,
    2. Li X,
    3. Carvalho-de-Souza JL,
    4. Wei W,
    5. Bezanilla F,
    6. Tian B
    (2019) An atlas of nano-enabled neural interfaces. Nat Nanotechnol 14:645–657. doi:10.1038/s41565-019-0487-x
    OpenUrlCrossRef
  4. ↵
    1. Aebersold MJ,
    2. Dermutz H,
    3. Forró C,
    4. Weydert S,
    5. Thompson-Steckel G,
    6. Voros J,
    7. Demkó L
    (2016) “Brains on a chip”: towards engineered neural networks. Trends Analyt Chem 78:60–69. doi:10.1016/j.trac.2016.01.025
    OpenUrlCrossRef
  5. ↵
    1. Akemann W,
    2. Lundby A,
    3. Mutoh H,
    4. Knöpfel T
    (2009) Effect of voltage sensitive fluorescent proteins on neuronal excitability. Biophys J 96:3959–3976. doi:10.1016/j.bpj.2009.02.046
    OpenUrlCrossRefPubMed
  6. ↵
    1. Akin EJ,
    2. Alsaloum M,
    3. Higerd GP,
    4. Liu S,
    5. Zhao P,
    6. Dib-Hajj FB,
    7. Waxman SG,
    8. Dib-Hajj SD
    (2021) Paclitaxel increases axonal localization and vesicular trafficking of Nav1.7. Brain 144:1727–1737. doi:10.1093/brain/awab113
    OpenUrlCrossRefPubMed
  7. ↵
    1. Akin EJ,
    2. Higerd GP,
    3. Mis MA,
    4. Tanaka BS,
    5. Adi T,
    6. Liu S,
    7. Dib-Hajj FB,
    8. Waxman SG,
    9. Dib-Hajj SD
    (2019) Building sensory axons: delivery and distribution of Na V 1.7 channels and effects of inflammatory mediators. Sci Adv 5:eaax4755. doi:10.1126/sciadv.aax4755
    OpenUrlFREE Full Text
  8. ↵
    1. Alcamí P,
    2. Pereda AE
    (2019) Beyond plasticity: the dynamic impact of electrical synapses on neural circuits. Nat Rev Neurosci 20:253–271. doi:10.1038/s41583-019-0133-5
    OpenUrlCrossRefPubMed
  9. ↵
    1. Alcami P,
    2. el Hady A
    (2019) Axonal computations. Front Cell Neurosci 13:413.doi:10.3389/fncel.2019.00413
    OpenUrlCrossRef
  10. ↵
    1. Alle H,
    2. Geiger JRP
    (2006) Combined analog and action potential coding in hippocampal mossy fibers. Science 311:1290–1293. doi:10.1126/science.1119055
    OpenUrlAbstract/FREE Full Text
  11. ↵
    1. Alpizar SA,
    2. Cho IH,
    3. Hoppa MB
    (2019) Subcellular control of membrane excitability in the axon. Curr Opin Neurobiol 57:117–125. doi:10.1016/j.conb.2019.01.020
    OpenUrlCrossRef
  12. ↵
    1. Amin H,
    2. Dipalo M,
    3. De Angelis F,
    4. Berdondini L
    (2018) Biofunctionalized 3D nanopillar arrays fostering cell-guidance and promoting synapse stability and neuronal activity in networks. ACS Appl Mater Interfaces 10:15207–15215. doi:10.1021/acsami.8b00387
    OpenUrlCrossRef
  13. ↵
    1. Bähner F,
    2. Weiss EK,
    3. Birke G,
    4. Maier N,
    5. Schmitz D,
    6. Rudolph U,
    7. Frotscher M,
    8. Traub RD,
    9. Both M
    (2011) Cellular correlate of assembly formation in oscillating hippocampal networks in vitro. Proc Natl Acad Sci U S A 108:E607–E616. doi:10.1073/pnas.1103546108
    OpenUrlAbstract/FREE Full Text
  14. ↵
    1. Bakkum DJ,
    2. Frey U,
    3. Radivojevic M,
    4. Russell TL,
    5. Müller J,
    6. Fiscella M,
    7. Takahashi H,
    8. Hierlemann A
    (2013) Tracking axonal action potential propagation on a high-density microelectrode array across hundreds of sites. Nat Commun 4:2181. doi:10.1038/ncomms3181
    OpenUrlCrossRefPubMed
  15. ↵
    1. Bakkum DJ,
    2. Obien MEJ,
    3. Radivojevic M,
    4. Jäckel D,
    5. Frey U,
    6. Takahashi H,
    7. Hierlemann A
    (2018) The axon initial segment is the dominant contributor to the neuron’s extracellular electrical potential landscape. Adv Biosyst 3:2. doi:10.1002/adbi.201800308
    OpenUrlCrossRef
  16. ↵
    1. Bando Y,
    2. Sakamoto M,
    3. Kim S,
    4. Ayzenshtat I,
    5. Yuste R
    (2019) Comparative evaluation of genetically encoded voltage indicators. Cell Rep 26:802–813.e4. doi:10.1016/j.celrep.2018.12.088
    OpenUrlCrossRefPubMed
  17. ↵
    1. Baranauskas G,
    2. David Y,
    3. Fleidervish IA
    (2013) Spatial mismatch between the Na+ flux and spike initiation in axon initial segment. Proc Natl Acad Sci U S A 110:4051–4056. doi:10.1073/pnas.1215125110
    OpenUrlAbstract/FREE Full Text
  18. ↵
    1. Bean BP
    (2007) The action potential in mammalian central neurons. Nat Rev Neurosci 8:451–465. doi:10.1038/nrn2148
    OpenUrlCrossRefPubMed
  19. ↵
    1. Begum R,
    2. Bakiri Y,
    3. Volynski KE,
    4. Kullmann DiM
    (2016) Action potential broadening in a presynaptic channelopathy. Nat Commun 7:12102. doi:10.1038/ncomms12102
    OpenUrlCrossRefPubMed
  20. ↵
    1. Berke JD
    (2018) What does dopamine mean? Nat Neurosci 21:787–793. doi:10.1038/s41593-018-0152-y
    OpenUrlCrossRefPubMed
  21. ↵
    1. Bischofberger J,
    2. Engel D,
    3. Li L,
    4. Geiger JRP,
    5. Jonas P
    (2006) Patch-clamp recording from mossy fiber terminals in hippocampal slices. Nat Protoc 1:2075–2081. doi:10.1038/nprot.2006.312
    OpenUrlCrossRefPubMed
  22. ↵
    1. Brockhaus J,
    2. Brüggen B,
    3. Missler M
    (2019) Imaging and analysis of presynaptic calcium influx in cultured neurons using syngcamp6f. Front Synaptic Neurosci 11. doi:10.3389/fnsyn.2019.00012
    OpenUrlCrossRefPubMed
  23. ↵
    1. Broguiere N,
    2. Lüchtefeld I,
    3. Trachsel L,
    4. Mazunin D,
    5. Rizzo R,
    6. Bode JW,
    7. Lutolf MP,
    8. Zenobi-Wong M,
    9. Zenobi-Wong M
    (2020) Morphogenesis guided by 3D patterning of growth factors in biological matrices. Adv Mater 32:1908299. doi:10.1002/adma.201908299
    OpenUrlCrossRef
  24. ↵
    1. Broussard GJ,
    2. Liang Y,
    3. Fridman M,
    4. Unger EK,
    5. Meng G,
    6. Xiao X,
    7. Ji N,
    8. Petreanu L,
    9. Tian L
    (2018) In vivo measurement of afferent activity with axon-specific calcium imaging. Nat Neurosci 21:1272–1280. doi:10.1038/s41593-018-0211-4
    OpenUrlCrossRefPubMed
  25. ↵
    1. Broussard GJ,
    2. Liang R,
    3. Tian L
    (2014) Monitoring activity in neural circuits with genetically encoded indicators. Front Mol Neurosci 7:97. doi:10.3389/fnmol.2014.00097
    OpenUrlCrossRefPubMed
  26. ↵
    1. Broussard GJ,
    2. Petreanu L
    (2021) Eavesdropping wires: recording activity in axons using genetically encoded calcium indicators. J Neurosci Methods 360:109251. doi:10.1016/j.jneumeth.2021.109251
    OpenUrlCrossRef
  27. ↵
    1. Bucher D,
    2. Goaillard J-M
    (2011) Beyond faithful conduction: short-term dynamics, neuromodulation, and long-term regulation of spike propagation in the axon. Prog Neurobiol 94:307–346. doi:10.1016/j.pneurobio.2011.06.001
    OpenUrlCrossRefPubMed
  28. ↵
    1. Burke KJ,
    2. Keeshen CM,
    3. Bender KJ
    (2018) Two forms of synaptic depression produced by differential neuromodulation of presynaptic calcium channels. Neuron 99:969–984.e7. doi:10.1016/j.neuron.2018.07.030
    OpenUrlCrossRefPubMed
  29. ↵
    1. Campenot RB
    (1977) Local control of neurite development by nerve growth factor. Proc Natl Acad Sci U S A 74:4516–4519. doi:10.1073/pnas.74.10.4516
    OpenUrlAbstract/FREE Full Text
  30. ↵
    1. Chéreau R,
    2. Saraceno GE,
    3. Angibaud J,
    4. Cattaert D,
    5. Nägerl UV
    (2017) Superresolution imaging reveals activity-dependent plasticity of axon morphology linked to changes in action potential conduction velocity. Proc Natl Acad Sci U S A 114:1401–1406. doi:10.1073/pnas.1607541114
    OpenUrlAbstract/FREE Full Text
  31. ↵
    1. Cherkas V, et al.
    (2018) Measurement of intracellular concentration of fluorescently-labeled targets in living cells. PLoS One 13:1–21. doi:10.1371/journal.pone.0194031
    OpenUrlCrossRefPubMed
  32. ↵
    1. Cho IH,
    2. Panzera LC,
    3. Chin M,
    4. Alpizar SA,
    5. Olveda GE,
    6. Hill RA,
    7. Hoppa MB
    (2020) The potassium channel subunit Kvβ1 serves as a major control point for synaptic facilitation. Proc Natl Acad Sci U S A 117:29937–29947. doi:10.1073/pnas.2000790117
    OpenUrlAbstract/FREE Full Text
  33. ↵
    1. Cho IH,
    2. Panzera LC,
    3. Chin M,
    4. Hoppa MB
    (2017) Sodium channel β2 subunits prevent action potential propagation failures at axonal branch points. J Neurosci 37:9519–9533. doi:10.1523/JNEUROSCI.0891-17.2017
    OpenUrlAbstract/FREE Full Text
  34. ↵
    1. Cho YH,
    2. Park Y,
    3. Kim S,
    4. Park J
    (2021) 3D electrodes for bioelectronics. Adv Mater 2005805. doi:10.1002/adma.202005805
    OpenUrlCrossRef
  35. ↵
    1. Christie JM,
    2. Chiu DN,
    3. Jahr CE
    (2010) Ca2+-dependent enhancement of release by subthreshold somatic depolarization. Nat Neurosci 14:62–68. doi:10.1038/nn.2718
    OpenUrlCrossRef
  36. ↵
    1. Chua JS,
    2. Chng CP,
    3. Moe AAK,
    4. Tann JY,
    5. Goh ELK,
    6. Chiam KH,
    7. Yim EKF
    (2014) Extending neurites sense the depth of the underlying topography during neuronal differentiation and contact guidance. Biomaterials 35:7750–7761. doi:10.1016/j.biomaterials.2014.06.008
    OpenUrlCrossRef
  37. ↵
    1. Cohen CCH,
    2. Popovic MA,
    3. Klooster J,
    4. Weil MT,
    5. Möbius W,
    6. Nave KA,
    7. Kole MHP
    (2020) Saltatory conduction along myelinated axons involves a periaxonal nanocircuit. Cell 180:311–322.e15. doi:10.1016/j.cell.2019.11.039
    OpenUrlCrossRefPubMed
  38. ↵
    1. Colbert CM,
    2. Johnston D
    (1996) Axonal action-potential initiation and Na+ channel densities in the soma and axon initial segment of subicular pyramidal neurons. J Neurosci 16:6676–6686. doi:10.1523/JNEUROSCI.16-21-06676.1996
    OpenUrlAbstract/FREE Full Text
  39. ↵
    1. Costa AR, et al.
    (2020) The membrane periodic skeleton is an actomyosin network that regulates axonal diameter and conduction. Elife 9:e55471. doi:10.7554/eLife.55471
    OpenUrlCrossRef
  40. ↵
    1. Costa AR,
    2. Pinto-Costa R,
    3. Sousa SC,
    4. Sousa MM
    (2018) The regulation of axon diameter: from axonal circumferential contractility to activity-dependent axon swelling. Front Mol Neurosci 11:319. doi:10.3389/fnmol.2018.00319
    OpenUrlCrossRef
  41. ↵
    1. Dana H, et al.
    (2019) High-performance calcium sensors for imaging activity in neuronal populations and microcompartments. Nat Methods 16:649–657. doi:10.1038/s41592-019-0435-6
    OpenUrlCrossRefPubMed
  42. ↵
    1. Debanne D,
    2. Bialowas A,
    3. Rama S
    (2013) What are the mechanisms for analogue and digital signalling in the brain? Nat Rev Neurosci 14:63–69. doi:10.1038/nrn3361
    OpenUrlCrossRefPubMed
  43. ↵
    1. Debanne D,
    2. Campanac E,
    3. Bialowas A,
    4. Carlier E,
    5. Alcaraz G
    (2011) Axon physiology. Physiol Rev 91:555–602. doi:10.1152/physrev.00048.2009
    OpenUrlCrossRefPubMed
  44. ↵
    1. Deemyad T,
    2. Lüthi J,
    3. Spruston N
    (2018) Astrocytes integrate and drive action potential firing in inhibitory subnetworks. Nat Commun 9:1–13. doi:10.1038/s41467-018-06338-3
    OpenUrlCrossRefPubMed
  45. ↵
    1. Dipalo M,
    2. Amin H,
    3. Lovato L,
    4. Moia F,
    5. Caprettini V,
    6. Messina GC,
    7. Tantussi F,
    8. Berdondini L,
    9. De Angelis F
    (2017) Intracellular and extracellular recording of spontaneous action potentials in mammalian neurons and cardiac cells with 3D plasmonic nanoelectrodes. Nano Lett 17:3932–3939. doi:10.1021/acs.nanolett.7b01523
    OpenUrlCrossRef
  46. ↵
    1. Dolphin AC,
    2. Lee A
    (2020) Presynaptic calcium channels: specialized control of synaptic neurotransmitter release. Nat Rev Neurosci 21:213–229. doi:10.1038/s41583-020-0278-2
    OpenUrlCrossRefPubMed
  47. ↵
    1. Dreosti E,
    2. Odermatt B,
    3. Dorostkar MM,
    4. Lagnado L
    (2009) A genetically encoded reporter of synaptic activity in vivo. Nat Methods 6:883–889. doi:10.1038/nmeth.1399
    OpenUrlCrossRefPubMed
  48. ↵
    1. Duffy DC,
    2. McDonald JC,
    3. Schueller OJA,
    4. Whitesides GM
    (1998) Rapid prototyping of microfluidic systems in poly(dimethylsiloxane). Anal Chem 70:4974–4984. doi:10.1021/ac980656z
    OpenUrlCrossRefPubMed
  49. ↵
    1. Dugladze T,
    2. Schmitz D,
    3. Whittington MA,
    4. Vida I,
    5. Gloveli T
    (2012) Segregation of axonal and somatic activity during fast network oscillations. Science 336:1458–1461. doi:10.1126/science.1222017
    OpenUrlAbstract/FREE Full Text
  50. ↵
    1. Dupin I,
    2. Dahan M,
    3. Studer V
    (2013) Investigating axonal guidance with microdevice-based approaches. J Neurosci 33:17647. doi:10.1523/JNEUROSCI.3277-13.2013
    OpenUrlAbstract/FREE Full Text
  51. ↵
    1. Duru J,
    2. Küchler J,
    3. Ihle SJ,
    4. Forró C,
    5. Bernardi A,
    6. Girardin S,
    7. Hengsteler J,
    8. Wheeler S,
    9. Vörös J
    (2022) Engineered biological neural networks on high density CMOS microelectrode arrays. Front Neurosci 16:829884. doi:10.3389/fnins.2022.829884
    OpenUrlCrossRef
  52. ↵
    1. Dworak BJ,
    2. Wheeler BC
    (2009) Novel MEA platform with PDMS microtunnels enables the detection of action potential propagation from isolated axons in culture. Lab Chip 9:404–410. doi:10.1039/B806689B
    OpenUrlCrossRefPubMed
  53. ↵
    1. Ellis-Davies GCR
    (2007) Caged compounds: photorelease technology for control of cellular chemistry and physiology. Nat Methods 4:619. doi:10.1038/nmeth1072
    OpenUrlCrossRefPubMed
  54. ↵
    1. Emmenegger V,
    2. Obien MEJ,
    3. Franke F,
    4. Hierlemann A
    (2019) Technologies to study action potential propagation with a focus on HD-MEAs. Front Cell Neurosci 13:159. doi:10.3389/fncel.2019.00159
    OpenUrlCrossRef
  55. ↵
    1. Engel D,
    2. Jonas P
    (2005) Presynaptic action potential amplification by voltage-gated Na+ channels in hippocampal mossy fiber boutons. Neuron 45:405–417. doi:10.1016/j.neuron.2004.12.048
    OpenUrlCrossRefPubMed
  56. ↵
    1. Fantuzzo JA,
    2. Hart RP,
    3. Zahn JD,
    4. Pang ZP
    (2019) Compartmentalized devices as tools for investigation of human brain network dynamics. Dev Dyn 248:65–77. doi:10.1002/dvdy.24665
    OpenUrlCrossRef
  57. ↵
    1. Feinerman O,
    2. Rotem A,
    3. Moses E
    (2008) Reliable neuronal logic devices from patterned hippocampal cultures. Nat Phys 4:967–973. doi:10.1038/nphys1099
    OpenUrlCrossRef
  58. ↵
    1. Filipis L,
    2. Blömer LA,
    3. Montnach J,
    4. Loussouarn G,
    5. De Waard M,
    6. Canepari M
    (2023) Nav1.2 and BK channels interaction shapes the action potential in the axon initial segment. J Physiol 601:1957–1979. doi:10.1113/JP283801
    OpenUrlCrossRef
  59. ↵
    1. Filipis L,
    2. Canepari M
    (2021) Optical measurement of physiological sodium currents in the axon initial segment. J Physiol 599:49–66. doi:10.1113/JP280554
    OpenUrlCrossRefPubMed
  60. ↵
    1. Fleidervish IA,
    2. Lasser-Ross N,
    3. Gutnick MJ,
    4. Ross WN
    (2010) Na+ imaging reveals little difference in action potential-evoked Na+ influx between axon and soma. Nat Neurosci 13:852–860. doi:10.1038/nn.2574
    OpenUrlCrossRefPubMed
  61. ↵
    1. Formento E,
    2. D’Anna E,
    3. Gribi S,
    4. Lacour SP,
    5. Micera S
    (2020) A biomimetic electrical stimulation strategy to induce asynchronous stochastic neural activity. J Neural Eng 17:046019. doi:10.1088/1741-2552/aba4fc
    OpenUrlCrossRefPubMed
  62. ↵
    1. Forró C,
    2. Thompson-Steckel G,
    3. Weaver S,
    4. Weydert S,
    5. Ihle S,
    6. Dermutz H,
    7. Aebersold MJ,
    8. Pilz R,
    9. Demkó L,
    10. Vörös J
    (2018) Modular microstruture design to build neuronal networks of defined functional connectivity. Biosens Bioelectron 122:75–87. doi:10.1016/j.bios.2018.08.075
    OpenUrlCrossRef
  63. ↵
    1. Gautam V,
    2. Naureen S,
    3. Shahid N,
    4. Gao Q,
    5. Wang Y,
    6. Nisbet D,
    7. Jagadish C,
    8. Daria VR
    (2017) Engineering highly interconnected neuronal networks on nanowire scaffolds. Nano Lett 17:3369–3375. doi:10.1021/acs.nanolett.6b05288
    OpenUrlCrossRef
  64. ↵
    1. Ghovanloo MR,
    2. Tyagi S,
    3. Zhao P,
    4. Kiziltug E,
    5. Estacion M,
    6. Dib-Hajj SD,
    7. Waxman SG
    (2023) High-throughput combined voltage-clamp/current-clamp analysis of freshly isolated neurons. Cell Rep Methods 3:100385. doi:10.1016/j.crmeth.2022.100385
    OpenUrlCrossRef
  65. ↵
    1. Girardin S, et al.
    (2022) Topologically controlled circuits of human iPSC-derived neurons for electrophysiology recordings. Lab Chip 22:1386–1403. doi:10.1039/d1lc01110c
    OpenUrlCrossRef
  66. ↵
    1. Gladkov A,
    2. Pigareva Y,
    3. Kutyina D,
    4. Kolpakov V,
    5. Bukatin A,
    6. Mukhina I,
    7. Kazantsev V,
    8. Pimashkin A
    (2017) Design of cultured neuron networks in vitro with predefined connectivity using asymmetric microfluidic channels. Sci Rep 7:15625. doi:10.1038/s41598-017-15506-2
    OpenUrlCrossRef
  67. ↵
    1. Gobbo F,
    2. Cattaneo A
    (2020) Neuronal activity at synapse resolution: reporters and effectors for synaptic neuroscience. Front Mol Neurosci 13:184. doi:10.3389/fnmol.2020.572312
    OpenUrlCrossRef
  68. ↵
    1. Gonzalez Sabater V,
    2. Rigby M,
    3. Burrone J
    (2021) Voltage-gated potassium channels ensure action potential shape fidelity in distal axons. J Neurosci41:2020.09.15.297895. doi:10.1523/JNEUROSCI.2765-20.2021
    OpenUrlAbstract/FREE Full Text
  69. ↵
    1. Gribi S,
    2. du Bois de Dunilac S,
    3. Ghezzi D,
    4. Lacour SP
    (2018) A microfabricated nerve-on-a-chip platform for rapid assessment of neural conduction in explanted peripheral nerve fibers. Nat Commun 9:4403. doi:10.1038/s41467-018-06895-7
    OpenUrlCrossRef
  70. ↵
    1. Grienberger C,
    2. Konnerth A
    (2012) Imaging calcium in neurons. Neuron 73:862–885. doi:10.1016/j.neuron.2012.02.011
    OpenUrlCrossRefPubMed
  71. ↵
    1. Grubb MS,
    2. Burrone J
    (2010a) Activity-dependent relocation of the axon initial segment fine-tunes neuronal excitability. Nature 465:1070–1074. doi:10.1038/nature09160
    OpenUrlCrossRefPubMed
  72. ↵
    1. Grubb MS,
    2. Burrone J
    (2010b) Channelrhodopsin-2 localised to the axon initial segment. PLoS One 5:13761. doi:10.1371/journal.pone.0013761
    OpenUrlCrossRef
  73. ↵
    1. Gundersen RW,
    2. Barrett JN
    (1979) Neuronal chemotaxis: chick dorsal-root axons turn toward high concentrations of nerve growth factor. Science 206:1079–1080. doi:10.1126/science.493992
    OpenUrlAbstract/FREE Full Text
  74. ↵
    1. Habibey R,
    2. Golabchi A,
    3. Latifi S,
    4. Difato F,
    5. Blau A
    (2015) A microchannel device tailored to laser axotomy and long-term microelectrode array electrophysiology of functional regeneration. Lab Chip 15:4578–4590. doi:10.1039/C5LC01027F
    OpenUrlCrossRef
  75. ↵
    1. Habibey R,
    2. Latifi S,
    3. Mousavi H,
    4. Pesce M,
    5. Arab-Tehrany E,
    6. Blau A
    (2017) A multielectrode array microchannel platform reveals both transient and slow changes in axonal conduction velocity. Sci Rep 7:8558. doi:10.1038/s41598-017-09033-3
    OpenUrlCrossRef
  76. ↵
    1. Hagemann C,
    2. Moreno Gonzalez C,
    3. Guetta L,
    4. Tyzack G,
    5. Chiappini C,
    6. Legati A,
    7. Patani R,
    8. Serio A
    (2022) Axonal length determines distinct homeostatic phenotypes in human iPSC derived motor neurons on a bioengineered platform. Adv Healthc Mater 11:2101817. doi:10.1002/adhm.202101817
    OpenUrlCrossRef
  77. ↵
    1. Haggerty DL,
    2. Grecco GG,
    3. Reeves KC,
    4. Atwood B
    (2020) Adeno-associated viral vectors in neuroscience research. Mol Ther Methods Clin Dev 17:69–82. doi:10.1016/j.omtm.2019.11.012
    OpenUrlCrossRefPubMed
  78. ↵
    1. Hai A,
    2. Shappir J,
    3. Spira ME
    (2010) In-cell recordings by extracellular microelectrodes. Nat Methods 7:200–202. doi:10.1038/nmeth.1420
    OpenUrlCrossRefPubMed
  79. ↵
    1. Hamada S,
    2. Nagase M,
    3. Yoshizawa T,
    4. Hagiwara A,
    5. Isomura Y,
    6. Watabe AM,
    7. Ohtsuka T
    (2021) An engineered channelrhodopsin optimized for axon terminal activation and circuit mapping. Commun Biol 4:1–11. doi:10.1038/s42003-021-01977-7
    OpenUrlCrossRef
  80. ↵
    1. Hamada MS,
    2. Popovic MA,
    3. Kole MHP
    (2017) Loss of saltation and presynaptic action potential failure in demyelinated axons. Front Cell Neurosci 11:45. doi:10.3389/fncel.2017.00045
    OpenUrlCrossRefPubMed
  81. ↵
    1. Hamill OP,
    2. Marty A,
    3. Neher E,
    4. Sakmann B,
    5. Sigworth FJ
    (1981) Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflugers Arch 391:85–100. doi:10.1007/BF00656997
    OpenUrlCrossRefPubMed
  82. ↵
    1. Han K-S,
    2. Chen CH,
    3. Khan MM,
    4. Guo C,
    5. Regehr WG
    (2020) Climbing fiber synapses rapidly and transiently inhibit neighboring Purkinje cells via ephaptic coupling. Nat Neurosci 23:1399–1409. doi:10.1038/s41593-020-0701-z
    OpenUrlCrossRefPubMed
  83. ↵
    1. Hanemaaijer NAK,
    2. Popovic MA,
    3. Wilders X,
    4. Grasman S,
    5. Arocas OP,
    6. Kole MHP
    (2020) Ca2+ entry through nav channels generates submillisecond axonal Ca2+ signaling. Elife 9:54566. doi:10.7554/eLife.54566
    OpenUrlCrossRef
  84. ↵
    1. Hari K, et al.
    (2022) GABA facilitates spike propagation through branch points of sensory axons in the spinal cord. Nat Neurosci 25:1288–1299. doi:10.1038/s41593-022-01162-x
    OpenUrlCrossRefPubMed
  85. ↵
    1. Heiney K,
    2. Mateus JC,
    3. Lopes CDF,
    4. Neto E,
    5. Lamghari M,
    6. Aguiar P
    (2019) µSpikeHunter: an advanced computational tool for the analysis of neuronal communication and action potential propagation in microfluidic platforms. Sci Rep 9:5777. doi:10.1038/s41598-019-42148-3
    OpenUrlCrossRef
  86. ↵
    1. Hochbaum DR, et al.
    (2014) All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins. Nat Methods 11:825–833. doi:10.1038/nmeth.3000
    OpenUrlCrossRefPubMed
  87. ↵
    1. Hodapp A, et al.
    (2022) Dendritic axon origin enables information gating by perisomatic inhibition in pyramidal neurons. Science 377:1448–1452. doi:10.1126/science.abj1861
    OpenUrlCrossRefPubMed
  88. ↵
    1. Hodgkin AL,
    2. Huxley AF
    (1952) A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117:500–544. doi:10.1113/jphysiol.1952.sp004764
    OpenUrlCrossRefPubMed
  89. ↵
    1. Holloway PM,
    2. Hallinan GI,
    3. Hegde M,
    4. Lane SIR,
    5. Deinhardt K,
    6. West J
    (2019) Asymmetric confinement for defining outgrowth directionality. Lab Chip 19:1484–1489. doi:10.1039/C9LC00078J
    OpenUrlCrossRefPubMed
  90. ↵
    1. Holloway PM,
    2. Willaime-Morawek S,
    3. Siow R,
    4. Barber M,
    5. Owens RM,
    6. Sharma AD,
    7. Rowan W,
    8. Hill E,
    9. Zagnoni M
    (2021) Advances in microfluidic in vitro systems for neurological disease modeling. J Neurosci Res 00:1–32. doi:10.1002/jnr.24794
    OpenUrlCrossRef
  91. ↵
    1. Hong N,
    2. Joo S,
    3. Nam Y
    (2017) Characterization of axonal spikes in cultured neuronal networks using microelectrode arrays and microchannel devices. IEEE Trans Biomed Eng 64:492–498. doi:10.1109/TBME.2016.2567424
    OpenUrlCrossRef
  92. ↵
    1. Hong N,
    2. Nam Y
    (2020) Thermoplasmonic neural chip platform for in situ manipulation of neuronal connections in vitro. Nat Commun 11:1–12. doi:10.1038/s41467-019-13993-7
    OpenUrlCrossRefPubMed
  93. ↵
    1. Hoppa MB,
    2. Gouzer G,
    3. Armbruster M,
    4. Ryan TA
    (2014) Control and plasticity of the presynaptic action potential waveform at small CNS nerve terminals. Neuron 84:778–789. doi:10.1016/j.neuron.2014.09.038
    OpenUrlCrossRefPubMed
  94. ↵
    1. Hu H,
    2. Jonas P
    (2014) A supercritical density of Na+ channels ensures fast signaling in GABAergic interneuron axons. Nat Neurosci 17:686–693. doi:10.1038/nn.3678
    OpenUrlCrossRefPubMed
  95. ↵
    1. Hu H,
    2. Roth FC,
    3. Vandael D,
    4. Jonas P
    (2018) Complementary tuning of Na+ and K+ channel gating underlies fast and energy-efficient action potentials in GABAergic interneuron axons. Neuron 98:156–165.e6. doi:10.1016/j.neuron.2018.02.024
    OpenUrlCrossRefPubMed
  96. ↵
    1. Hu W,
    2. Tian C,
    3. Li T,
    4. Yang M,
    5. Hou H,
    6. Shu Y
    (2009) Distinct contributions of Nav1.6 and Nav1.2 in action potential initiation and backpropagation. Nat Neurosci 12:996–1002. doi:10.1038/nn.2359
    OpenUrlCrossRefPubMed
  97. ↵
    1. Huang L, et al.
    (2021) Relationship between simultaneously recorded spiking activity and fluorescence signal in gcamp6 transgenic mice. Elife 10:e51675. doi:10.7554/eLife.51675
    OpenUrlCrossRef
  98. ↵
    1. Hur E-M,
    2. Yang IH,
    3. Kim D-H,
    4. Byun J,
    5. Saijilafu XW-L,
    6. Nicovich PR,
    7. Cheong R,
    8. Levchenko A,
    9. Thakor N,
    10. Zhou F-Q
    (2011) Engineering neuronal growth cones to promote axon regeneration over inhibitory molecules. Proc Natl Acad Sci U S A 108:5057–5062. doi:10.1073/pnas.1011258108
    OpenUrlAbstract/FREE Full Text
  99. ↵
    1. Hyung S,
    2. Lee SR,
    3. Kim J,
    4. Kim Y,
    5. Kim S,
    6. Kim HN,
    7. Jeon NL
    (2021) A 3D disease and regeneration model of peripheral nervous system–on–a–chip. Sci Adv 7:9749–9778. doi:10.1126/sciadv.abd9749
    OpenUrlCrossRef
  100. ↵
    1. Inoue M, et al.
    (2019) Rational engineering of XCaMPs, a multicolor GECI suite for in vivo imaging of complex brain circuit dynamics. Cell 177:1346–1360.e24. doi:10.1016/j.cell.2019.04.007
    OpenUrlCrossRefPubMed
  101. ↵
    1. Jäckel D,
    2. Bakkum DJ,
    3. Russell TL,
    4. Müller J,
    5. Radivojevic M,
    6. Frey U,
    7. Franke F,
    8. Hierlemann A
    (2017) Combination of high-density microelectrode array and patch clamp recordings to enable studies of multisynaptic integration. Sci Rep 7:978. doi:10.1038/s41598-017-00981-4
    OpenUrlCrossRef
  102. ↵
    1. Jackson RE,
    2. Burrone J
    (2016) Visualizing presynaptic calcium dynamics and vesicle fusion with a single genetically encoded reporter at individual synapses. Front Synaptic Neurosci 8:21. doi:10.3389/fnsyn.2016.00021
    OpenUrlCrossRef
  103. ↵
    1. Jang Y,
    2. Kim SR,
    3. Lee SH
    (2021) Methods of measuring presynaptic function with fluorescence probes. Appl Microsc 51:2. doi:10.1186/s42649-021-00051-0
    OpenUrlCrossRef
  104. ↵
    1. Jang JM,
    2. Lee J,
    3. Kim H,
    4. Jeon NL,
    5. Jung W
    (2016) One-photon and two-photon stimulation of neurons in a microfluidic culture system. Lab Chip 16:1684–1690. doi:10.1039/C6LC00065G
    OpenUrlCrossRef
  105. ↵
    1. Katz E,
    2. Stoler O,
    3. Scheller A,
    4. Khrapunsky Y,
    5. Goebbels S,
    6. Kirchhoff F,
    7. Gutnick MJ,
    8. Wolf F,
    9. Fleidervish IA
    (2018) Role of sodium channel subtype in action potential generation by neocortical pyramidal neurons. Proc Natl Acad Sci U S A 115:E7184–E7192. doi:10.1073/pnas.1804655115
    OpenUrlAbstract/FREE Full Text
  106. ↵
    1. Kawaguchi S-y,
    2. Sakaba T
    (2015) Control of inhibitory synaptic outputs by low excitability of axon terminals revealed by direct recording. Neuron 85:1273–1288. doi:10.1016/j.neuron.2015.02.013
    OpenUrlCrossRefPubMed
  107. ↵
    1. Kim HJ, et al.
    (2012) Integrated microfluidics platforms for investigating injury and regeneration of CNS axons. Ann Biomed Eng 40:1268–1276. doi:10.1007/s10439-012-0515-6
    OpenUrlCrossRefPubMed
  108. ↵
    1. Kim WR,
    2. Jang MJ,
    3. Joo S,
    4. Sun W,
    5. Nam Y
    (2014) Surface-printed microdot array chips for the quantification of axonal collateral branching of a single neuron in vitro. Lab Chip 14:799–805. doi:10.1039/C3LC51169C
    OpenUrlCrossRef
  109. ↵
    1. Kim KM,
    2. Kim SY,
    3. Palmore GTR
    (2016) Axon outgrowth of rat embryonic hippocampal neurons in the presence of an electric field. ACS Chem Neurosci 7:1325–1330. doi:10.1021/acschemneuro.6b00191
    OpenUrlCrossRef
  110. ↵
    1. Kleinfeld D,
    2. Kahler KH,
    3. Hockberger PE
    (1988) Controlled outgrowth of dissociated neurons on patterned substrates. J Neurosci 8:4098–4120. doi:10.1523/JNEUROSCI.08-11-04098.1988
    OpenUrlAbstract/FREE Full Text
  111. ↵
    1. Knöpfel T,
    2. Song C
    (2019) Optical voltage imaging in neurons: moving from technology development to practical tool. Nat Rev Neurosci 20:719–727. doi:10.1038/s41583-019-0231-4
    OpenUrlCrossRefPubMed
  112. ↵
    1. Koester HJ,
    2. Sakmann B
    (2000) Calcium dynamics associated with action potentials in single nerve terminals of pyramidal cells in layer 2/3 of the young rat neocortex. J Physiol 529:625–646. doi:10.1111/j.1469-7793.2000.00625.x
    OpenUrlCrossRefPubMed
  113. ↵
    1. Kole MH,
    2. Brette R
    (2018) The electrical significance of axon location diversity. Curr Opin Neurobiol 51:52–59. doi:10.1016/j.conb.2018.02.016
    OpenUrlCrossRefPubMed
  114. ↵
    1. Kole MHP,
    2. Ilschner SU,
    3. Kampa BM,
    4. Williams SR,
    5. Ruben PC,
    6. Stuart GJ
    (2008) Action potential generation requires a high sodium channel density in the axon initial segment. Nat Neurosci 11:178–186. doi:10.1038/nn2040
    OpenUrlCrossRefPubMed
  115. ↵
    1. Kole MHP,
    2. Stuart GJ
    (2012) Signal processing in the axon initial segment. Neuron 73:235–247. doi:10.1016/j.neuron.2012.01.007
    OpenUrlCrossRefPubMed
  116. ↵
    1. Koos K, et al.
    (2021) Automatic deep learning-driven label-free image-guided patch clamp system. Nat Commun 12:1–11. doi:10.1038/s41467-021-21291-4
    OpenUrlCrossRef
  117. ↵
    1. Kramer PF,
    2. Brill-Weil SG,
    3. Cummins AC,
    4. Zhang R,
    5. Camacho-Hernandez GA,
    6. Newman AH,
    7. Eldridge MAG,
    8. Averbeck BB,
    9. Khaliq ZM
    (2022) Synaptic-like axo-axonal transmission from striatal cholinergic interneurons onto dopaminergic fibers. Neuron 110:2949–2960.e4. doi:10.1016/j.neuron.2022.07.011
    OpenUrlCrossRef
  118. ↵
    1. Kramer PF,
    2. Twedell EL,
    3. Shin JH,
    4. Zhang R,
    5. Khaliq ZM
    (2020) Axonal mechanisms mediating g-aminobutyric acid receptor type a (GABA-A) inhibition of striatal dopamine release. Elife 9:1–24. doi:10.7554/eLife.55729
    OpenUrlCrossRef
  119. ↵
    1. Kuba H,
    2. Oichi Y,
    3. Ohmori H
    (2010) Presynaptic activity regulates Na(+) channel distribution at the axon initial segment. Nature 465:1075–1078. doi:10.1038/nature09087
    OpenUrlCrossRefPubMed
  120. ↵
    1. Lang-Ouellette D, et al.
    (2021) Purkinje cell axonal swellings enhance action potential fidelity and cerebellar function. Nat Commun 12:1–13. doi:10.1038/s41467-021-24390-4
    OpenUrlCrossRefPubMed
  121. ↵
    1. Lazarov E,
    2. Dannemeyer M,
    3. Feulner B,
    4. Enderlein J,
    5. Gutnick MJ,
    6. Wolf F,
    7. Neef A
    (2018) An axon initial segment is required for temporal precision in action potential encoding by neuronal populations. Sci Adv 4:eaau8621. doi:10.1126/sciadv.aau8621
    OpenUrlFREE Full Text
  122. ↵
    1. Leclech C,
    2. Villard C
    (2020) Cellular and subcellular contact guidance on microfabricated substrates. Front Bioeng Biotechnol 8:1198. doi:10.3389/fbioe.2020.551505
    OpenUrlCrossRef
  123. ↵
    1. Lewandowska MK,
    2. Bakkum DJ,
    3. Rompani SB,
    4. Hierlemann A
    (2015) Recording large extracellular spikes in microchannels along many axonal sites from individual neurons. PLoS One 10:e0118514. doi:10.1371/journal.pone.0118514
    OpenUrlCrossRefPubMed
  124. ↵
    1. Lezmy J,
    2. Lipinsky M,
    3. Khrapunsky Y,
    4. Patrich E,
    5. Shalom L,
    6. Peretz A,
    7. Fleidervish IA,
    8. Attali B
    (2017) M-current inhibition rapidly induces a unique CK2-dependent plasticity of the axon initial segment. Proc Natl Acad Sci U S A 114:E10234–E10243. doi:10.1073/pnas.1708700114
    OpenUrlAbstract/FREE Full Text
  125. ↵
    1. Li W,
    2. Tang QY,
    3. Jadhav AD,
    4. Narang A,
    5. Qian WX,
    6. Shi P,
    7. Pang SW
    (2015) Large-scale topographical screen for investigation of physical neural-guidance cues. Sci Rep 5:8644. doi:10.1038/srep08644
    OpenUrlCrossRef
  126. ↵
    1. Lipkin AM,
    2. Cunniff MM,
    3. Spratt PWE,
    4. Lemke SM,
    5. Bender KJ
    (2021) Functional microstructure of CaV -mediated calcium signaling in the axon initial segment. J Neurosci 41:3764–3776. doi:10.1523/JNEUROSCI.2843-20.2021
    OpenUrlAbstract/FREE Full Text
  127. ↵
    1. Liu R, et al.
    (2017) High density individually addressable nanowire arrays record intracellular activity from primary rodent and human stem cell derived neurons. Nano Lett 17:2757–2764. doi:10.1021/acs.nanolett.6b04752
    OpenUrlCrossRef
  128. ↵
    1. Liu Q, et al.
    (2019) A photoactivatable botulinum neurotoxin for inducible control of neurotransmission. Neuron 101:863–875.e6. doi:10.1016/j.neuron.2019.01.002
    OpenUrlCrossRefPubMed
  129. ↵
    1. Liu Z, et al.
    (2022b) Sustained deep-tissue voltage recording using a fast indicator evolved for two-photon microscopy. Cell 185:3408–3425. doi:10.1016/J.CELL.2022.07.013
    OpenUrlCrossRef
  130. ↵
    1. Liu C,
    2. Cai X,
    3. Ritzau-Jost A,
    4. Kramer PF,
    5. Li Y,
    6. Khaliq ZM,
    7. Hallermann S,
    8. Kaeser PS
    (2022a) An action potential initiation mechanism in distal axons for the control of dopamine release. Science 375:1378–1385. doi:10.1126/science.abn0532
    OpenUrlCrossRefPubMed
  131. ↵
    1. Lopes CDF,
    2. Mateus JC,
    3. Aguiar P
    (2018) Interfacing microfluidics with microelectrode arrays for studying neuronal communication and axonal signal propagation. J Vis Exp 142:1–8. doi:10.3791/58878
    OpenUrlCrossRef
  132. ↵
    1. Lorincz A,
    2. Nusser Z
    (2010) Molecular identity of dendritic voltage-gated sodium channels. Science 328:906–909. doi:10.1126/science.1187958
    OpenUrlAbstract/FREE Full Text
  133. ↵
    1. Ma Y,
    2. Bayguinov PO,
    3. Jackson MB
    (2017) Action potential dynamics in fine axons probed with an axonally targeted optical voltage sensor. eNeuro 4:ENEURO.0146-17.2017. doi:10.1523/ENEURO.0146-17.2017
    OpenUrlAbstract/FREE Full Text
  134. ↵
    1. Mahn M,
    2. Gibor L,
    3. Patil P,
    4. Cohen-Kashi Malina K,
    5. Oring S,
    6. Printz Y,
    7. Levy R,
    8. Lampl I,
    9. Yizhar O
    (2018) High-efficiency optogenetic silencing with soma-targeted anion-conducting channelrhodopsins. Nat Commun 9:1–15. doi:10.1038/s41467-018-06511-8
    OpenUrlCrossRefPubMed
  135. ↵
    1. Marcus M,
    2. Baranes K,
    3. Park M,
    4. Choi IS,
    5. Kang K,
    6. Shefi O
    (2017) Interactions of neurons with physical environments. Adv Healthc Mater 6:1700267. doi:10.1002/adhm.201700267
    OpenUrlCrossRef
  136. ↵
    1. Mateus JC, et al.
    (2019) Improved in vitro electrophysiology using 3D-structured microelectrode arrays with a micro-mushrooms islets architecture capable of promoting topotaxis. J Neural Eng 16:036012. doi:10.1088/1741-2552/ab0b86
    OpenUrlCrossRef
  137. ↵
    1. Mateus JC, et al.
    (2021) Bidirectional flow of action potentials in axons drives activity dynamics in neuronal cultures. J Neural Eng 18:066045. doi:10.1088/1741-2552/ac41db
    OpenUrlCrossRef
  138. ↵
    1. Mateus JC,
    2. Weaver S,
    3. van Swaay D,
    4. Renz AF,
    5. Hengsteler J,
    6. Aguiar P,
    7. Vörös J
    (2022) Nanoscale patterning of in vitro neuronal circuits. ACS Nano 16:acsnano.1c10750. doi:10.1021/acsnano.1c10750
    OpenUrlCrossRef
  139. ↵
    1. McMahon SM,
    2. Jackson MB
    (2018) An inconvenient truth: calcium sensors are calcium buffers. Trends Neurosci 41:880. doi:10.1016/j.tins.2018.09.005
    OpenUrlCrossRefPubMed
  140. ↵
    1. Messier JE,
    2. Chen H,
    3. Cai ZL,
    4. Xue M
    (2018) Targeting light-gated chloride channels to neuronal somatodendritic domain reduces their excitatory effect in the axon. Elife 7:e38506. doi:10.7554/eLife.38506
    OpenUrlCrossRef
  141. ↵
    1. Micholt L,
    2. Gärtner A,
    3. Prodanov D,
    4. Braeken D,
    5. Dotti CG,
    6. Bartic C
    (2013) Substrate topography determines neuronal polarization and growth in vitro. PLoS One 8:e66170. doi:10.1371/journal.pone.0066170
    OpenUrlCrossRef
  142. ↵
    1. Middya S,
    2. Curto VF,
    3. Fernández-Villegas A,
    4. Robbins M,
    5. Gurke J,
    6. Moonen EJM,
    7. Kaminski Schierle GS,
    8. Malliaras GG
    (2021) Microelectrode arrays for simultaneous electrophysiology and advanced optical microscopy. AdvSci 8:2004434. doi:10.1002/advs.202004434
    OpenUrlCrossRef
  143. ↵
    1. Millet LJ,
    2. Gillette MU
    (2012) New perspectives on neuronal development via microfluidic environments. Trends Neurosci 35:752–761. doi:10.1016/j.tins.2012.09.001
    OpenUrlCrossRefPubMed
  144. ↵
    1. Milos F,
    2. Tullii G,
    3. Gobbo F,
    4. Lodola F,
    5. Galeotti F,
    6. Verpelli C,
    7. Mayer D,
    8. Maybeck V,
    9. Offenhäusser A
    (2021) High aspect ratio and light-sensitive micropillars based on a semiconducting polymer optically regulate neuronal growth. ACS Appl Mater Interfaces 13:23438–23451. doi:10.1021/acsami.1c03537
    OpenUrlCrossRef
  145. ↵
    1. Moutaux E,
    2. Charlot B,
    3. Genoux A,
    4. Saudou F,
    5. Cazorla M
    (2018) An integrated microfluidic/microelectrode array for the study of activity-dependent intracellular dynamics in neuronal networks. Lab Chip 18:3425–3435. doi:10.1039/C8LC00694F
    OpenUrlCrossRef
  146. ↵
    1. Narula U,
    2. Ruiz A,
    3. McQuaide M,
    4. DeMarse TB,
    5. Wheeler BC,
    6. Brewer GJ
    (2017) Narrow microtunnel technology for the isolation and precise identification of axonal communication among distinct hippocampal subregion networks. PLoS One 12:e0176868. doi:10.1371/journal.pone.0176868
    OpenUrlCrossRef
  147. ↵
    1. Nascimento AI,
    2. Da Silva TF,
    3. Fernandes EC,
    4. Luz LL,
    5. Mar FM,
    6. Safronov BV,
    7. Sousa MM
    (2022) Sensory neurons have an axon initial segment that initiates spontaneous activity in neuropathic pain. Brain 145:1632–1640. doi:10.1093/brain/awac078
    OpenUrlCrossRef
  148. ↵
    1. Neto E,
    2. Leitão L,
    3. Sousa DM,
    4. Alves CJ,
    5. Alencastre IS,
    6. Aguiar P,
    7. Lamghari M
    (2016) Compartmentalized microfluidic platforms: the unrivaled breakthrough of in vitro tools for neurobiological research. J Neurosci 36:11573–11584. doi:10.1523/JNEUROSCI.1748-16.2016
    OpenUrlAbstract/FREE Full Text
  149. ↵
    1. Nikolakopoulou P,
    2. Rauti R,
    3. Voulgaris D,
    4. Shlomy I,
    5. Maoz BM,
    6. Herland A
    (2021) Recent progress in translational engineered in vitro models of the central nervous system. Brain 143:3181–3213. doi:10.1093/brain/awaa268
    OpenUrlCrossRef
  150. ↵
    1. Novak P, et al.
    (2013) Nanoscale-targeted patch-clamp recordings of functional presynaptic ion channels. Neuron 79:1067–1077. doi:10.1016/j.neuron.2013.07.012
    OpenUrlCrossRefPubMed
  151. ↵
    1. Obien MEJ,
    2. Deligkaris K,
    3. Bullmann T,
    4. Bakkum DJ,
    5. Frey U
    (2015) Revealing neuronal function through microelectrode array recordings. Front Neurosci 8:423. doi:10.3389/fnins.2014.00423
    OpenUrlCrossRefPubMed
  152. ↵
    1. Oláh VJ,
    2. Tarcsay G,
    3. Brunner J
    (2021) Small size of recorded neuronal structures confines the accuracy in direct axonal voltage measurements. eNeuro 8:1–20. doi:10.1523/ENEURO.0059-21.2021
    OpenUrlCrossRef
  153. ↵
    1. Oldani S,
    2. Moreno-Velasquez L,
    3. Faiss L,
    4. Stumpf A,
    5. Rosenmund C,
    6. Schmitz D,
    7. Rost BR
    (2021) SynaptoPAC, an optogenetic tool for induction of presynaptic plasticity. J Neurochem 156:324–336. doi:10.1111/jnc.15210
    OpenUrlCrossRef
  154. ↵
    1. Oliva AA,
    2. James CD,
    3. Kingman CE,
    4. Craighead HG,
    5. Banker GA
    (2003) Patterning axonal guidance molecules using a novel strategy for microcontact printing. Neurochem Res 28:1639–1648. doi:10.1023/A:1026052820129
    OpenUrlCrossRefPubMed
  155. ↵
    1. Pan L,
    2. Alagapan S,
    3. Franca E,
    4. DeMarse T,
    5. Brewer GJ,
    6. Wheeler BC
    (2014) Large extracellular spikes recordable from axons in microtunnels. IEEE Trans Neural Syst Rehabil Eng 22:453–459. doi:10.1109/TNSRE.2013.2289911
    OpenUrlCrossRef
  156. ↵
    1. Pan L,
    2. Borgens RB
    (2012) Strict perpendicular orientation of neural crest-derived neurons in vitro is dependent on an extracellular gradient of voltage. J Neurosci Res 90:1335–1346. doi:10.1002/jnr.22809
    OpenUrlCrossRefPubMed
  157. ↵
    1. Pan X,
    2. Li J,
    3. Li W,
    4. Wang H,
    5. Durisic N,
    6. Li Z,
    7. Feng Y,
    8. Liu Y,
    9. Zhao C-X
    (2022) Axons-on-a-chip for mimicking non-disruptive diffuse axonal injury underlying traumatic brain injury. Lab Chip 22:4541–4555. doi:10.1039/d2lc00730d
    OpenUrlCrossRef
  158. ↵
    1. Panzera LC,
    2. Hoppa MB
    (2019) Genetically encoded voltage indicators are illuminating subcellular physiology of the axon. Front Cell Neurosci 13:52. doi:10.3389/fncel.2019.00052
    OpenUrlCrossRef
  159. ↵
    1. Park M,
    2. Oh E,
    3. Seo J,
    4. Kim M-H,
    5. Cho H,
    6. Choi JY,
    7. Lee H,
    8. Choi IS
    (2016) Control over neurite directionality and neurite elongation on anisotropic micropillar arrays. Small 12:1148–1152. doi:10.1002/smll.201501896
    OpenUrlCrossRef
  160. ↵
    1. Park JW,
    2. Vahidi B,
    3. Taylor AM,
    4. Rhee SW,
    5. Jeon NL
    (2006) Microfluidic culture platform for neuroscience research. Nat Protoc 1:2128–2136. doi:10.1038/nprot.2006.316
    OpenUrlCrossRefPubMed
  161. ↵
    1. Patel N,
    2. Poo MM
    (1982) Orientation of neurite growth by extracellular electric fields. J Neurosci 2:483–496. doi:10.1523/JNEUROSCI.02-04-00483.1982
    OpenUrlAbstract/FREE Full Text
  162. ↵
    1. Perge JA,
    2. Niven JE,
    3. Mugnaini E,
    4. Balasubramanian V,
    5. Sterling P
    (2012) Why do axons differ in caliber? J Neurosci 32:626–638. doi:10.1523/JNEUROSCI.4254-11.2012
    OpenUrlAbstract/FREE Full Text
  163. ↵
    1. Peterka DS,
    2. Takahashi H,
    3. Yuste R
    (2011) Imaging voltage in neurons. Neuron 69:9–21. doi:10.1016/j.neuron.2010.12.010
    OpenUrlCrossRefPubMed
  164. ↵
    1. Peyrin JM, et al.
    (2011) Axon diodes for the reconstruction of oriented neuronal networks in microfluidic chambers. Lab Chip 11:3663–3673. doi:10.1039/c1lc20014c
    OpenUrlCrossRefPubMed
  165. ↵
    1. Pinato G,
    2. Cojoc D,
    3. Lien LT,
    4. Ansuini A,
    5. Ban J,
    6. D’Este E,
    7. Torre V
    (2012) Less than 5 netrin-1 molecules initiate attraction but 200 Sema3A molecules are necessary for repulsion. Sci Rep 2:675. doi:10.1038/srep00675
    OpenUrlCrossRefPubMed
  166. ↵
    1. Popovic MA,
    2. Foust AJ,
    3. Mccormick DA,
    4. Zecevic D
    (2011) The spatio-temporal characteristics of action potential initiation in layer 5 pyramidal neurons: a voltage imaging study. J Physiol 589:4167–4187. doi:10.1113/jphysiol.2011.209015
    OpenUrlCrossRefPubMed
  167. ↵
    1. Potter SM,
    2. DeMarse TB
    (2001) A new approach to neural cell culture for long-term studies. J Neurosci Methods 110:17–24. doi:10.1016/S0165-0270(01)00412-5
    OpenUrlCrossRefPubMed
  168. ↵
    1. Radivojevic M,
    2. Franke F,
    3. Altermatt M,
    4. Müller J,
    5. Hierlemann A,
    6. Bakkum DJJ
    (2017) Tracking individual action potentials throughout mammalian axonal arbors. Elife 6:e30198. doi:10.7554/eLife.30198
    OpenUrlCrossRef
  169. ↵
    1. Radivojevic M,
    2. Punga AR
    (2023) Functional imaging of conduction dynamics in cortical and spinal axons. Elife 12:e86512. doi:10.7554/eLife.86512
    OpenUrlCrossRef
  170. ↵
    1. Raj V,
    2. Jagadish C,
    3. Gautam V
    (2021) Understanding, engineering, and modulating the growth of neural networks: an interdisciplinary approach. Biophys Rev 2:021303. doi:10.1063/5.0043014
    OpenUrlCrossRef
  171. ↵
    1. Rall W
    (1959) Branching dendritic trees and motoneuron membrane resistivity. Exp Neurol 1:491–527. doi:10.1016/0014-4886(59)90046-9
    OpenUrlCrossRefPubMed
  172. ↵
    1. Rama S,
    2. Zbili M,
    3. Debanne D
    (2018) Signal propagation along the axon. Curr Opin Neurobiol 51:37–44. doi:10.1016/j.conb.2018.02.017
    OpenUrlCrossRefPubMed
  173. ↵
    1. Revah O, et al.
    (2022) Maturation and circuit integration of transplanted human cortical organoids. Nature 610:319–326. doi:10.1038/s41586-022-05277-w
    OpenUrlCrossRef
  174. ↵
    1. Ritzau-Jost A,
    2. Tsintsadze T,
    3. Krueger M,
    4. Ader J,
    5. Bechmann I,
    6. Eilers J,
    7. Barbour B,
    8. Smith SM,
    9. Hallermann S
    (2021) Large, stable spikes exhibit differential broadening in excitatory and inhibitory neocortical boutons. Cell Rep 34:108612. doi:10.1016/j.celrep.2020.108612
    OpenUrlCrossRefPubMed
  175. ↵
    1. Robinson JT,
    2. Jorgolli M,
    3. Shalek AK,
    4. Yoon M-H,
    5. Gertner RS,
    6. Park H
    (2012) Vertical nanowire electrode arrays as a scalable platform for intracellular interfacing to neuronal circuits. Nat Nanotechnol 7:180–184. doi:10.1038/nnano.2011.249
    OpenUrlCrossRefPubMed
  176. ↵
    1. Rost BR,
    2. Wietek J,
    3. Yizhar O,
    4. Schmitz D
    (2022) Optogenetics at the presynapse. Nat Neurosci 25:948–998. doi:10.1038/s41593-022-01113-6
    OpenUrlCrossRef
  177. ↵
    1. Roth S,
    2. Bisbal M,
    3. Brocard J,
    4. Bugnicourt G,
    5. Saoudi Y,
    6. Andrieux A,
    7. Gory-Fauré S,
    8. Villard C
    (2012) How morphological constraints affect axonal polarity in mouse neurons. PLoS One 7:33623. doi:10.1371/journal.pone.0033623
    OpenUrlCrossRef
  178. ↵
    1. Rowan MJM,
    2. Christie JM
    (2017) Rapid state-dependent alteration in Kv3 channel availability drives flexible synaptic signaling dependent on somatic subthreshold depolarization. Cell Rep 18:2018–2029. doi:10.1016/j.celrep.2017.01.068
    OpenUrlCrossRefPubMed
  179. ↵
    1. Rowan MJM,
    2. DelCanto G,
    3. Yu JJ,
    4. Kamasawa N,
    5. Christie JM
    (2016) Synapse-level determination of action potential duration by K+ channel clustering in axons. Neuron 91:370–383. doi:10.1016/j.neuron.2016.05.035
    OpenUrlCrossRefPubMed
  180. ↵
    1. Roy J,
    2. Kennedy TE,
    3. Costantino S
    (2013) Engineered cell culture substrates for axon guidance studies: moving beyond proof of concept. Lab Chip 13:498. doi:10.1039/c2lc41002h
    OpenUrlCrossRef
  181. ↵
    1. Rózsa M,
    2. Tóth M,
    3. Oláh G,
    4. Baka J,
    5. Lákovics R,
    6. Barzó P,
    7. Tamás G
    (2023) Temporal disparity of action potentials triggered in axon initial segments and distal axons in the neocortex. Sci Adv 9:eade4511. doi:10.1126/sciadv.ade4511
    OpenUrlCrossRef
  182. ↵
    1. Ruiz A,
    2. Campanac E,
    3. Scott RS,
    4. Rusakov DA,
    5. Kullmann DM
    (2010) Presynaptic GABAA receptors enhance transmission and LTP induction at hippocampal mossy fiber synapses. Nat Neurosci 13:431–438. doi:10.1038/nn.2512
    OpenUrlCrossRefPubMed
  183. ↵
    1. Santoro F,
    2. Panaitov G,
    3. Andreas O
    (2014) Defined patterns of neuronal networks on 3D thiol-functionalized microstructures. Nano Lett 14:6906–6909. doi:10.1021/nl502922b
    OpenUrlCrossRef
  184. ↵
    1. Sasaki T,
    2. Matsuki N,
    3. Ikegaya Y
    (2011) Action-potential modulation during axonal conduction. Science 331:599–601. doi:10.1126/science.1197598
    OpenUrlAbstract/FREE Full Text
  185. ↵
    1. Sasaki T,
    2. Matsuki N,
    3. Ikegaya Y
    (2012a) Effects of axonal topology on the somatic modulation of synaptic outputs. J Neurosci 32:2868–2876. doi:10.1523/JNEUROSCI.5365-11.2012
    OpenUrlAbstract/FREE Full Text
  186. ↵
    1. Sasaki T,
    2. Matsuki N,
    3. Ikegaya Y
    (2012b) Targeted axon-attached recording with fluorescent patch-clamp pipettes in brain slices. Nat Protoc 7:1228–1234. doi:10.1038/nprot.2012.061
    OpenUrlCrossRefPubMed
  187. ↵
    1. Scanziani M,
    2. Häusser M,
    3. Hausser M
    (2009) Electrophysiology in the age of light. Nature 461:930–939. doi:10.1038/nature08540
    OpenUrlCrossRefPubMed
  188. ↵
    1. Seo J, et al.
    (2018) Nanotopography-promoted formation of axon collateral branches of hippocampal neurons. Small:1801763. doi:10.1002/smll.201801763
    OpenUrlCrossRef
  189. ↵
    1. Sheffield MEJ,
    2. Best TK,
    3. Mensh BD,
    4. Kath WL,
    5. Spruston N
    (2011) Slow integration leads to persistent action potential firing in distal axons of coupled interneurons. Nat Neurosci 14:200–207. doi:10.1038/nn.2728
    OpenUrlCrossRefPubMed
  190. ↵
    1. Shimba K,
    2. Asahina T,
    3. Sakai K,
    4. Kotani K,
    5. Jimbo Y
    (2022) Recording saltatory conduction along sensory axons using a high-density microelectrode array. Front Neurosci 16:537. doi:10.3389/fnins.2022.854637
    OpenUrlCrossRef
  191. ↵
    1. Shimba K,
    2. Kotani K,
    3. Jimbo Y
    (2021) Microfabricated device to record axonal conduction under pharmacological treatment for functional evaluation of axon ion channel. IEEE Trans Biomed Eng 68:3574–3581. doi:10.1021/nl502922b
    OpenUrlCrossRef
  192. ↵
    1. Shmoel N,
    2. Rabieh N,
    3. Ojovan SM,
    4. Erez H,
    5. Maydan E,
    6. Spira ME
    (2016) Multisite electrophysiological recordings by self-assembled loose-patch-like junctions between cultured hippocampal neurons and mushroom-shaped microelectrodes. Sci Rep 6:27110. doi:10.1038/srep27110
    OpenUrlCrossRef
  193. ↵
    1. Shu Y,
    2. Hasenstaub A,
    3. Duque A,
    4. Yu Y,
    5. McCormick DA
    (2006) Modulation of intracortical synaptic potentials by presynaptic somatic membrane potential. Nature 441:761–765. doi:10.1038/nature04720
    OpenUrlCrossRefPubMed
  194. ↵
    1. Shvartsman A,
    2. Kotler O,
    3. Stoler O,
    4. Khrapunsky Y,
    5. Melamed I,
    6. Fleidervish IA
    (2021) Subcellular distribution of persistent sodium conductance in cortical pyramidal neurons. Journal of Neuroscience 41:6190–6201. doi:10.1523/JNEUROSCI.2989-20.2021
    OpenUrlAbstract/FREE Full Text
  195. ↵
    1. Simitzi C,
    2. Ranella A,
    3. Stratakis E
    (2017) Controlling the morphology and outgrowth of nerve and neuroglial cells: the effect of surface topography. Acta Biomater 51:21–52. doi:10.1016/j.actbio.2017.01.023
    OpenUrlCrossRef
  196. ↵
    1. Spira ME,
    2. Hai A
    (2013) Multi-electrode array technologies for neuroscience and cardiology. Nat Nanotechnol 8:83–94. doi:10.1038/nnano.2012.265
    OpenUrlCrossRefPubMed
  197. ↵
    1. Stevenson IH,
    2. Kording KP
    (2011) How advances in neural recording affect data analysis. Nat Neurosci 14:139–142. doi:10.1038/nn.2731
    OpenUrlCrossRefPubMed
  198. ↵
    1. Suk HJ,
    2. Boyden ES,
    3. van Welie I
    (2019) Advances in the automation of whole-cell patch clamp technology. J Neurosci Methods 326:108357. doi:10.1016/j.jneumeth.2019.108357
    OpenUrlCrossRef
  199. ↵
    1. Suzuki I,
    2. Matsuda N,
    3. Han X,
    4. Noji S,
    5. Shibata M,
    6. Nagafuku N,
    7. Ishibashi Y
    (2023) Large-area field potential imaging having single neuron resolution using 236 880 electrodes CMOS-MEA technology. Adv Sci 10:e2207732. doi:10.1002/advs.202207732
    OpenUrlCrossRef
  200. ↵
    1. Swadlow HA,
    2. Kocsis JD,
    3. Waxman SG
    (1980) Modulation of impulse conduction along the axonal tree. Annu Rev Biophys Bioeng 9:143–179. doi:10.1146/annurev.bb.09.060180.001043
    OpenUrlCrossRefPubMed
  201. ↵
    1. Taylor AM,
    2. Blurton-Jones M,
    3. Rhee SW,
    4. Cribbs DH,
    5. Cotman CW,
    6. Jeon NL
    (2005) A microfluidic culture platform for CNS axonal injury, regeneration and transport. Nat Methods 2:599–605. doi:10.1038/nmeth777
    OpenUrlCrossRefPubMed
  202. ↵
    1. Taylor AM,
    2. Rhee SW,
    3. Tu CH,
    4. Cribbs DH,
    5. Cotman CW,
    6. Jeon NL
    (2003) Microfluidic multicompartment device for neuroscience research. Langmuir 19:1551–1556. doi:10.1021/la026417v
    OpenUrlCrossRefPubMed
  203. ↵
    1. Teixeira H,
    2. Dias C,
    3. Aguiar P,
    4. Ventura J
    (2020) Gold-mushroom microelectrode arrays and the quest for intracellular-like recordings: perspectives and outlooks. Adv Mater Technol 6:2000770. doi:10.1002/admt.202000770
    OpenUrlCrossRef
  204. ↵
    1. Tovar KR,
    2. Bridges DC,
    3. Wu B,
    4. Randall C,
    5. Audouard M,
    6. Jang J,
    7. Hansma PK,
    8. Kosik KS
    (2018) Action potential propagation recorded from single axonal arbors using multielectrode arrays. J Neurophysiol 120:306–320. doi:10.1152/jn.00659.2017
    OpenUrlCrossRef
  205. ↵
    1. Vandael D,
    2. Borges-Merjane C,
    3. Zhang X,
    4. Jonas P
    (2020) Short-term plasticity at hippocampal mossy fiber synapses is induced by natural activity patterns and associated with vesicle pool engram formation. Neuron 107:509–521.e7. doi:10.1016/j.neuron.2020.05.013
    OpenUrlCrossRefPubMed
  206. ↵
    1. Vandael D,
    2. Okamoto Y,
    3. Borges-Merjane C,
    4. Vargas-Barroso V,
    5. Suter BA,
    6. Jonas P
    (2021) Subcellular patch-clamp techniques for single-bouton stimulation and simultaneous pre- and postsynaptic recording at cortical synapses. Nat Protoc 16:2947–2967. doi:10.1038/s41596-021-00526-0
    OpenUrlCrossRef
  207. ↵
    1. Vasylyev DV,
    2. Waxman SG
    (2012) Membrane properties and electrogenesis in the distal axons of small dorsal root ganglion neurons in vitro. J Neurophysiol 108:729–740. doi:10.1152/jn.00091.2012
    OpenUrlCrossRefPubMed
  208. ↵
    1. Vivekananda U,
    2. Novak P,
    3. Bello OD,
    4. Korchev YE,
    5. Krishnakumar SS,
    6. Volynski KE,
    7. Kullmann DM
    (2017) Kv1.1 channelopathy abolishes presynaptic spike width modulation by subthreshold somatic depolarization. Proc Natl Acad Sci U S A 114:2395–2400. doi:10.1073/pnas.1608763114
    OpenUrlAbstract/FREE Full Text
  209. ↵
    1. Vogt AK,
    2. Wrobel G,
    3. Meyer W,
    4. Knoll W,
    5. Offenhäusser A
    (2005) Synaptic plasticity in micropatterned neuronal networks. Biomaterials 26:2549–2557. doi:10.1016/j.biomaterials.2004.07.031
    OpenUrlCrossRefPubMed
  210. ↵
    1. Vyleta NP,
    2. Jonas P
    (2014) Loose coupling between Ca2+ channels and release sensors at a plastic hippocampal synapse. Science 343:665–670. doi:10.1126/SCIENCE.1244811
    OpenUrlAbstract/FREE Full Text
  211. ↵
    1. Vyleta NP,
    2. Snyder JS
    (2023) Enhanced excitability but mature action potential waveforms at mossy fiber terminals of young, adult-born hippocampal neurons in mice. Commun Biol 6:1–9. doi:10.1038/s42003-023-04678-5
    OpenUrlCrossRef
  212. ↵
    1. Wang W,
    2. Kim CK,
    3. Ting AY
    (2019) Molecular tools for imaging and recording neuronal activity. Nat Chem Biol 15:101–110. doi:10.1038/s41589-018-0207-0
    OpenUrlCrossRef
  213. ↵
    1. Wang L,
    2. Riss M,
    3. Buitrago JO,
    4. Claverol-Tinturé E
    (2012) Biophysics of microchannel-enabled neuron–electrode interfaces. J Neural Eng 9:026010. doi:10.1088/1741-2560/9/2/026010
    OpenUrlCrossRefPubMed
  214. ↵
    1. Weydert S, et al.
    (2019) A versatile protein and cell patterning method suitable for long-term neural cultures. Langmuir 35:2966–2975. doi:10.1021/acs.langmuir.8b03730
    OpenUrlCrossRef
  215. ↵
    1. Wu Y,
    2. Chen H,
    3. Guo L
    (2019) Opportunities and dilemmas of in vitro nano neural electrodes. RSC Adv 10:187–200. doi:10.1039/c9ra08917a
    OpenUrlCrossRef
  216. ↵
    1. Xu L,
    2. Hu C,
    3. Huang Q,
    4. Jin K,
    5. Zhao P,
    6. Wang D,
    7. Hou W,
    8. Dong L,
    9. Hu S
    (2021) Trends and recent development of the microelectrode arrays (MEAs). Biosens Bioelectron 175:112854. doi:10.1016/j.bios.2020.112854
    OpenUrlCrossRef
  217. ↵
    1. Xue X,
    2. Buccino AP,
    3. Kumar SS,
    4. Hierlemann A,
    5. Bartram J
    (2022) Inferring monosynaptic connections from paired dendritic spine Ca2+ imaging and large-scale recording of extracellular spiking. J Neural Eng 19:046044. doi:10.1088/1741-2552/AC8765
    OpenUrlCrossRef
  218. ↵
    1. Yamazaki Y,
    2. Hozumi Y,
    3. Kaneko K,
    4. Sugihara T,
    5. Fujii S,
    6. Goto K,
    7. Kato H
    (2007) Modulatory effects of oligodendrocytes on the conduction velocity of action potentials along axons in the alveus of the rat hippocampal CA1 region. Neuron Glia Biol 3:325–334. doi:10.1017/S1740925X08000070
    OpenUrlCrossRefPubMed
  219. ↵
    1. Yuan X, et al.
    (2020) Versatile live-cell activity analysis platform for characterization of neuronal dynamics at single-cell and network level. Nat Commun 11:4854. doi:10.1038/s41467-020-18620-4
    OpenUrlCrossRef
  220. ↵
    1. Zbili M,
    2. Debanne D
    (2019) Past and future of analog–digital modulation of synaptic transmission. Front Cell Neurosci 13:160. doi:10.3389/fncel.2019.00160
    OpenUrlCrossRefPubMed
  221. ↵
    1. Zbili M,
    2. Rama S,
    3. Yger P,
    4. Inglebert Y,
    5. Boumedine-Guignon N,
    6. Fronzaroli-Moliniere L,
    7. Brette R,
    8. Russier M,
    9. Debanne D
    (2020) Axonal Na+ channels detect and transmit levels of input synchrony in local brain circuits. Sci Adv 6:eaay4313. doi:10.1126/sciadv.aay4313
    OpenUrlFREE Full Text
  222. ↵
    1. Zečević D
    (1996) Multiple spike-initiation zones in single neurons revealed by voltage-sensitive dyes. Nature 381:322–325. doi:10.1038/381322a0
    OpenUrlCrossRefPubMed
  223. ↵
    1. Zhang Y, et al.
    (2023) Fast and sensitive GCaMP calcium indicators for imaging neural populations. Nature 615:884–891. doi:10.1038/s41586-023-05828-9
    OpenUrlCrossRefPubMed
  224. ↵
    1. Zhang Z,
    2. Feng J,
    3. Wu C,
    4. Lu Q,
    5. Pan ZH
    (2015) Targeted expression of channelrhodopsin-2 to the axon initial segment alters the temporal firing properties of retinal ganglion cells. PLoS One 10:e0142052. doi:10.1371/JOURNAL.PONE.0142052
    OpenUrlCrossRef
  225. ↵
    1. Zhang QF,
    2. Li H,
    3. Chen M,
    4. Guo A,
    5. Wen Y,
    6. Poo MM
    (2018) Functional organization of intrinsic and feedback presynaptic inputs in the primary visual cortex. Proc Natl Acad Sci U S A 115:E5174–E5182. doi:10.1073/pnas.1719711115
    OpenUrlAbstract/FREE Full Text
  226. ↵
    1. Zorrilla de San Martin J,
    2. Trigo FF,
    3. Kawaguchi SY
    (2017) Axonal GABAA receptors depolarize presynaptic terminals and facilitate transmitter release in cerebellar Purkinje cells. J Physiol 595:7477–7493. doi:10.1113/JP275369
    OpenUrlCrossRefPubMed
Back to top

In this issue

The Journal of Neuroscience: 44 (11)
Journal of Neuroscience
Vol. 44, Issue 11
13 Mar 2024
  • Table of Contents
  • About the Cover
  • Index by author
  • Masthead (PDF)
Email

Thank you for sharing this Journal of Neuroscience article.

NOTE: We request your email address only to inform the recipient that it was you who recommended this article, and that it is not junk mail. We do not retain these email addresses.

Enter multiple addresses on separate lines or separate them with commas.
Beyond a Transmission Cable—New Technologies to Reveal the Richness in Axonal Electrophysiology
(Your Name) has forwarded a page to you from Journal of Neuroscience
(Your Name) thought you would be interested in this article in Journal of Neuroscience.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Print
View Full Page PDF
Citation Tools
Beyond a Transmission Cable—New Technologies to Reveal the Richness in Axonal Electrophysiology
J. C. Mateus, M. M. Sousa, J. Burrone, P. Aguiar
Journal of Neuroscience 13 March 2024, 44 (11) e1446232023; DOI: 10.1523/JNEUROSCI.1446-23.2023

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Respond to this article
Request Permissions
Share
Beyond a Transmission Cable—New Technologies to Reveal the Richness in Axonal Electrophysiology
J. C. Mateus, M. M. Sousa, J. Burrone, P. Aguiar
Journal of Neuroscience 13 March 2024, 44 (11) e1446232023; DOI: 10.1523/JNEUROSCI.1446-23.2023
Twitter logo Facebook logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Google Plus One

Jump to section

  • Article
    • Abstract
    • Significance Statement
    • Introduction
    • Measuring Axonal Electrophysiology
    • Controlling Axonal Guidance
    • Discussion and Future Perspectives
    • Footnotes
    • References
  • Figures & Data
  • Info & Metrics
  • eLetters
  • PDF

Keywords

  • axon computations
  • axon electrophysiology
  • axon guidance
  • functional imaging
  • microelectrode arrays

Responses to this article

Respond to this article

Jump to comment:

No eLetters have been published for this article.

Related Articles

Cited By...

More in this TOC Section

  • A Perspective on Neuroscience Data Standardization with Neurodata Without Borders
  • Why and How to Account for Sex and Gender in Brain and Behavioral Research
Show more TechSights
  • Home
  • Alerts
  • Follow SFN on BlueSky
  • Visit Society for Neuroscience on Facebook
  • Follow Society for Neuroscience on Twitter
  • Follow Society for Neuroscience on LinkedIn
  • Visit Society for Neuroscience on Youtube
  • Follow our RSS feeds

Content

  • Early Release
  • Current Issue
  • Issue Archive
  • Collections

Information

  • For Authors
  • For Advertisers
  • For the Media
  • For Subscribers

About

  • About the Journal
  • Editorial Board
  • Privacy Notice
  • Contact
  • Accessibility
(JNeurosci logo)
(SfN logo)

Copyright © 2025 by the Society for Neuroscience.
JNeurosci Online ISSN: 1529-2401

The ideas and opinions expressed in JNeurosci do not necessarily reflect those of SfN or the JNeurosci Editorial Board. Publication of an advertisement or other product mention in JNeurosci should not be construed as an endorsement of the manufacturer’s claims. SfN does not assume any responsibility for any injury and/or damage to persons or property arising from or related to any use of any material contained in JNeurosci.