Abstract
Electrophysiological recordings from behaving animals provide an unparalleled view into the functional role of individual neurons. Intracellular approaches can be especially revealing as they provide information about a neuron’s inputs and intrinsic cellular properties, which together determine its spiking output. Recent technical developments have made intracellular recording possible during an ever-increasing range of behaviors in both head-fixed and freely moving animals. These recordings have yielded fundamental insights into the cellular and circuit mechanisms underlying neural activity during natural behaviors in such areas as sensory perception, motor sequence generation, and spatial navigation, forging a direct link between cellular and systems neuroscience.
Introduction
Extracellular recordings in awake animals have yielded a rich set of discoveries involving patterns of spiking activity that underlie specific behaviors [1–7]. Intracellular recordings have been employed to investigate the cellular mechanisms underlying such firing patterns, but technical issues have limited the extent to which intracellular studies could be carried out in awake animals. Anesthetized or sleeping preparations, which in general have smaller and fewer mechanical disturbances caused by animal movements, are sufficient for investigating some aspects of neural processing, such as feature selectivity in primary sensory areas. However, for studying other phenomena, including active sensation, birdsong or place cell activity, the animal must be awake or even moving around an environment. Here we review technical advances that have allowed intracellular recordings involving increasing degrees of mobility and highlight some recent cases where a deeper understanding of underlying mechanisms has directly resulted from these recordings.
Why study the awake brain with intracellular recordings?
Electrophysiological recordings in anesthetized preparations have revealed a tremendous amount about the function and organization of neural processing centers. For instance, orientation tuning in primary visual cortex was discovered in an anesthetized preparation [8]. However, while some aspects of neural activity appear to be preserved across anesthesia and awake states, many features appear to uniquely correspond to the awake brain (Figure 1). On the one hand, recent work has shown that the shape of orientation [9–10*] and other [11] tuning is similar in the awake and anesthetized brain, thus detailed intracellular studies of the synaptic mechanisms underlying such tuning in anesthetized animals [12–15] should apply directly to awake conditions. On the other hand, it was also shown that response magnitudes are strongly modulated by behavior in the awake animal [10*,16–17**], highlighting the dynamic nature of processing in the awake state.
Figure 1.
Intracellular dynamics of an identified neuron type recorded in vivo across different conditions – anesthetized, naturally sleeping, and singing. Each trace is taken from a distinct antidromically-identified RA-projecting HVC [HVC(RA)] neuron in an adult zebra finch. (a) Under isoflurane anesthesia (1.5% in oxygen), large, infrequent postsynaptic potentials are observed, sometimes resulting in action potentials (truncated and demarcated with diagonal lines). (b) Membrane potential (Vm) of an HVC(RA) neuron in a naturally sleeping, head-fixed bird. (c) Vm of an HVC(RA) neuron in an awake, singing bird (spectrogram of song above). Both the (b) sleeping and (c) awake states were associated with considerably more frequent synaptic events than was observed under anesthesia. Although the general shape of the subthreshold activity surrounding spiking events was similar in these two conditions (note the shaded epochs), the fine structure (i-iii) of the spike-aligned subthreshold activity was uncorrelated during sleep and nearly perfectly correlated during singing. Panel (a) is courtesy of D. Vallentin; panels (b) and (c) are modified from [44].
A number of previous studies have demonstrated that anesthesia can significantly affect aspects of coding, such as the extent to which neurons respond to sensory stimuli [18–20]. Recently, the same neurons were recorded in barrel cortex during wakefulness and under the influence of the rapidly-reversible anesthetic isoflurane [21*]. Under anesthesia, neurons exhibited prominent slow-wave membrane potential (Vm) fluctuations, similar to those which occur during certain stages of natural sleep [22–23]. However, when the animal was awake, Vm bistability was abolished and the neuron entered a desynchronized state [22,24] (but see [25]). Electrophysiological characteristics during wakefulness may be affected by neuromodulators such as acetylcholine [26] or norepinephrine [21*], which have also been shown to play an important role in the generation of fast oscillations [27] as well as the gating of plasticity processes [28].
Given the aforementioned differences between anesthetized and awake states, many researchers have sought to use awake preparations. But because awake animals often exhibit significant movements beyond those associated with heartbeats and breathing, awake intracellular recordings are in general challenging, as the recording pipette must maintain direct physical contact with the recorded neuron’s membrane. The intracellular voltage, however, provides data that can distinguish between proposed models of circuit function. Examples (from awake animals) of such data unattainable with other methodologies include subthreshold excitatory and/or inhibitory responses to internal [29–31,32*,33] or external (i.e. sensory) [34–38] stimulation as well as spontaneous Vm dynamics [39] and intrinsic properties [40–41]. Intracellular methods also allow experimental control of Vm to further probe mechanisms [42**,43**,44**,45**].
Case study: Feature selectivity
An example that illustrates what intracellular recordings can tell us about an ethologically relevant sensory computation is complex auditory feature selectivity in frogs. Frog calls often consist of a series of regularly-timed sound pulses that females can distinguish based on differing pulse number. A possible neural substrate of this ability consists of single neurons in the auditory midbrain that begin to spike only after a threshold number of pulses has occurred with a specific interpulse interval (Figure 2a) [46]. An example “interval-counting cell” fires after exactly four pulses, but not after three (Figure 2b). What gives a neuron such remarkable specificity? There are many potential models using specific combinations of intrinsic and extrinsic (synaptic) conductances. For instance, neurons that respond to a smaller number of pulses could have a more hyperpolarized spike threshold or be more intrinsically excitable than those responding to a larger number of pulses. Alternately, short term dynamics of the synaptic inputs onto these neurons could determine their response properties. A direct approach to finding the actual solution involves obtaining the Vm from an interval-counting neuron during the presentation of a call. Indeed, such recordings support a mechanism in which successive excitatory postsynaptic potentials (EPSPs) are strongly facilitating, allowing the neuron to overcome a substantial inhibitory conductance (Figure 2c) [43**]. In contrast, the difference between resting Vm and spike threshold was not significantly correlated with the number of pulses needed to fire the neuron [43**]. This ability to explain the origin of a complex sensory representation within a neuron holds promise for understanding higher-order processing centers in other systems.
Figure 2.
Intracellular recordings reveal processes underlying feature selectivity for a complex sensory stimulus. A population of neurons within the torus semicircularis of the frog midbrain responds selectively to a specific number of sound pulses with a specific interpulse interval. (a) Probability of spiking versus the number of pulses for six such interval-counting neurons. (b) Raster plot showing the responses of an interval-counting neuron to three (no evoked spikes) and four (15 evoked spikes in 18 trials) sound pulses. (c) Whole-cell recording from another interval-counting neuron. At top are the mean intracellular responses to 1-4 sound pulses. Note that repeated pulses were associated with excitatory synaptic facilitation which eventually depolarized the neuron above its resting potential (represented by the dotted gray line). Below is an overlay of three individual trials showing suprathreshold responses to five sound pulses. Vertical lines above designate the timing of spiking responses. Panels (a) and (b) are modified from [46]; panel (c) is modified from [43] and augmented with additional data courtesy of G. Rose.
Experimental subjects as active participants
Although the previously mentioned recordings in frog midbrain were made in an awake (and paralyzed) preparation, the specific contributions of wakefulness to the underlying processes remain unclear in that example. However, in the following cases, the experimental subject’s own behavior is involved in shaping and/or generating the neuronal inputs themselves.
Active sensation
Animals do not exclusively experience their environment passively; they can also actively gather information by exploring the outside world with their senses. One example in which active sensation has been well-characterized is the electrosensory system of mormyrid fish. These fish possess an electric organ that generates a weak electrical field (called an electric organ discharge) and electroreceptors on the skin that respond to changes in the field caused by interaction with nearby objects. Intracellular recordings from neurons in the electrosensory lobe of paralyzed fish have allowed detailed studies of interactions between motor signals (i.e. efferent copies of the motor commands that drive the electric organ discharge) and sensory signals conveyed by electroreceptors [47]. While paralysis blocks the electric organ discharge itself, the fish continues to actively emit the motor commands that normally trigger discharge. This preparation allows one to observe responses due to motor signals either in isolation or in combination with artificial electrosensory stimuli that mimic the fish’s own electric organ discharge. The stability of this preparation has allowed for intracellular studies of the integration of sensory responses and motor commands in tiny (< 5 μm diameter) cerebellar-like granule cells [48]. The sparse response of these inhibitory neurons to combinations of sensory and motor inputs could enable them to form a specific “negative image” that subtracts the expected from actual signals from each discharge, thus allowing the fish to detect novel, behaviorally relevant features of the environment. In a conceptually similar series of elegant experiments, intracellular recordings have revealed a novel, single-neuron mechanism of corollary discharge in an invertebrate preparation [49–50**].
Active sensation strategies can also be seen in rodents, which rely heavily on whisker-based touch information to explore objects in the environment. To physically sample the surroundings, rodents sweep their whiskers forwards and backwards at 5-15 Hz [51–52]. Consistent with previous extracellular recordings of spiking phase [53], intracellular recordings in head-fixed, whisking rodents showed that the Vm of each barrel cortex neuron oscillates with a different preferred phase relative to the whisk cycle [54–55], which could underlie object position coding [56 ]. Furthermore, it has been shown by intracellular, extracellular, and imaging methods that in both anesthetized and awake rodents layer 2/3 cortical neurons display minimal amounts of spiking overall in response to whisker deflection, whether in passive or active sensing mode [35,45**,54,57–61], as well as spontaneously [62]. Recently, in order to investigate the cellular basis of the sparse response to object contact during whisking, in which only ~10% of layer 2/3 neurons fire, intracellular recordings in head-fixed mice were employed [45] (Figure 3a). Specifically, it was found that each object contact results in synaptic input that drives each neuron to a particular reversal potential that is a combination of excitatory and inhibitory conductances (Figure 3b). This potential differs between cells and for most neurons is below the spike threshold, explaining why only a fraction of cells respond to contact with spiking (Figure 3c).
Figure 3.
Mechanisms of sparse coding and persistent activity in awake behaving animals. (a-c) Intracellular recordings of neurons from the barrel cortex of head-fixed mice as they moved their whiskers into contact with an object. (a) The response to object contact of three neurons, all of which displayed a Vm deflection but only one of which (cell 1) fired spikes. (b) Experiments in which current injection was used to alter the Vm level before object contact, showing that contact drove this neuron to a particular reversal potential. (c) The distribution of contact-induced reversal potentials (open bars) and spike thresholds (gray bars) across neurons shows why spiking activity is sparse in response to object contact: the reversal potentials for most neurons are below their spike thresholds. (d,e) Intracellular recordings of neurons from the hindbrain of a head-fixed goldfish. (d) When the fish moved its eye from one position to another, there was a persistent change in firing rate and subthreshold Vm such that the activity level was correlated with eye position. (e) The effect of hyperpolarizing current injection on Vm and firing rate was transient, with a return to their original values when the current was removed. This shows that persistent activity is not maintained by an intrinsic mechanism involving plateau potentials; instead the result is consistent with a network mechanism in which synaptic inputs maintain the cell's activity level. Panels (a-c) are modified from [45]; panels (d) and (e) are modified from [42].
Persistent activity
In addition to sensory processing, mechanistic explanations for other neuronal phenomena have depended on awake intracellular recordings. One notable case in point is persistent activity, which is correlated with, and potentially required for, short-term memory [63–64]. Persistent activity has been described in goldfish hindbrain Area I neurons, in which each brief saccadic command is converted into a sustained change in firing rate that allows the eye to maintain a fixed position in space [65] (Figure 3d). How can a neural circuit generate this sustained activity? One possibility is that the neurons are capable of maintaining various stable Vm states due to mechanisms intrinsic to the cell [66], such as dendritic plateau potentials [67]. Another is that persistent activity occurs because of recurrent synaptic activity which results in the continual activation of the network from other neurons within that circuit [68]. To distinguish between these two models, intracellular recordings were performed in a goldfish that was performing horizontal eye movements while it was head-fixed and its body restrained in a water tank [42**]. Current injection into the recording pipette was unable to induce changes in Vm that outlasted the current pulse (Figure 3e), suggesting that membrane multistability does not explain the persistent firing activity of these neurons. In addition, the variance in Vm increased linearly with eye position, consistent with an increase in synaptic feedback from other integrator neurons.
Recordings in moving animals
In the examples described so far, the animals could perform the behaviors while fixed in place and sometimes even paralyzed or partially dissected (as in some invertebrate cases [69**]). However, there are situations where the topic of study clearly necessitates significant bodily movement, such as studies of locomotion. Furthermore, sensory responsiveness can be affected by animal movement. The firing rate of the visual orientation-tuned neurons discussed above increased when mice transitioned from being stationary to running, suggesting a dynamic modulation of “gain” dependent on different awake states [10*], and similar modulations have been described in walking versus still [16] and flying versus resting insects [17**,33]. For intracellular recordings to survive such movements requires special attention to mechanical stabilization of the electrode and nervous tissue.
Head-fixed animals
A general approach involves fixing a solid part of the body as near as possible to the recording site securely in place, but unlike in the fixation schemes described above, the animals are expected to move the rest of their body. Fixation of several vertebrae allowed intracellular recording from spinal cord neurons of cats walking on a standard, linear treadmill [70]. Along these lines, whole-cell recordings have recently been obtained during “tethered” flight in head-fixed flies [17**,33]. As an alternative to this passive stabilization approach, an active system for moving the electrode in synchrony with brain movement was developed, yielding sharp microelectrode recordings from rats running on a linear treadmill [71**].
One variant of tethering that has received renewed attention involves a “spherical treadmill” (usually a light ball floating on a thin layer of air inside a bowl), which allows more freedom of movement than other apparatuses such as linear treadmills while still keeping the animal in place. This allows insects [72–74] and rodents [75,76*,77**] to walk in two-dimensions. With the addition of a visual “virtual reality” system [78] – in which a visual scene surrounding the animal moves in response to rotation of the spherical treadmill – rodents can navigate through simulated spatial environments such as mazes [75,76*,77**]. Using such a system, it was recently demonstrated that intracellular recordings could be obtained in head-fixed mice running in a virtual maze [77**].
Freely moving animals
What about unrestrained movement? A few years ago, it was discovered that rigidly fixing the electrode to the skull could yield long-duration (up to one hour) whole-cell recordings in freely moving rats [79*,80,81*,82**]. Passive stabilization of electrodes in a chronically-implanted microdrive has also allowed repeated sharp intracellular recordings in freely moving birds over several days [44**]. We now describe in more detail studies that have relied on these methods that allow more freedom of movement, beginning with birdsong.
Motor sequence generation
The concept of sequential neuronal activity mediated through groups of synaptically connected neurons has existed for decades [83–84], and a role for synaptic chains has been proposed in several brain regions, including the neocortex [85], but demonstrating the existence of these chains has proven to be difficult [86–87]. One circuit proposed to contain chains of synaptic activity [44**,88–89] is HVC, a forebrain motor structure which acts as a “clock” for the production of song in the zebra finch [4,90–91]. To address the presence of synaptic chains within that nucleus, intracellular recordings were made in antidromically-identified premotor neurons within HVC [44**,92] (Figure 1) using a chronically implanted motorized microdrive which enabled recordings in the freely moving zebra finch [44**]. These recordings revealed rapidly rising Vm directly preceding spiking activity, consistent with neurons being driven by a large synaptic input rather than a slow ramp or subthreshold rhythm. Furthermore, steady current injection (both hyperpolarizing and depolarizing) did not significantly change the timing of the firing, consistent with the idea that the neurons were being driven by a transient burst of input from the previous link the synaptic chain [44**]. Both of these observations were consistent with the presence of premotor activity which is driven through direct synaptic connections within HVC, but additional experiments are needed to further characterize these interconnections. Despite this fact, these data represent the most compelling evidence of a chain-like synaptic organization within a forebrain structure, which may represent a generic circuit strategy in cortex.
Spatial navigation
A rodent’s ability to navigate to a specific location is dependent on the hippocampus [93–94]. Within a given environment (e.g. a maze), hippocampal CA1 pyramidal neurons divide into two groups: place cells, each of which fire when the animal is in a particular location (called the cell’s place field) and which are likely used to solve navigational tasks [6,93], and silent cells, which fire few or no spikes [95]. What makes a place cell spike where it does? Observing place cells requires an animal capable of moving around an environment (but see [96]), either real or simulated [76*,77**], and intracellular recordings of place cells have been obtained in both freely moving rats [81–82,97] and head-fixed mice in a spherical treadmill virtual reality system [77**]. These recordings revealed that a sustained subthreshold Vm depolarization underlies spiking within the place field (Figure 4a,b), which rules out several models of place field origin, for instance those involving solely a localized increase in fluctuations (in both depolarizing and hyperpolarizing directions) that cross spike threshold [77**]. Further analysis showed that the shape of this depolarization is asymmetric [77**], as predicted [98]. Additional findings detectable only with intracellular methods included spikelets [81*], which add another dimension to mechanisms of neuronal integration, and large, putatively calcium-based events [77**,82**,99–100] (Figure 4a,b), which could trigger plasticity [101–102] underlying spatial memory formation in the hippocampus [103–105]. Lastly, what makes a neuron a place cell versus a silent cell? Comparison of place (Figure 4c) and silent (Figure 4d) cells revealed unexpected differences in excitability suggesting that intrinsic properties determine which cells become place cells in a novel environment [82**]. Together with studies in which excitability was molecularly manipulated [106], this suggests that intrinsic excitability biases which cells will participate in new memory traces.
Figure 4.
Whole-cell recordings of hippocampal CA1 place cells in real and virtual environments. (a) Intracellularly-recorded CA1 place cell from a freely moving rat as it ran around an oval-shaped track (top) and through the cell's place field (gray). (b) Intracellularly-recorded CA1 place cell from a head-fixed mouse as it ran on a spherical treadmill linked to a virtual reality system that allowed it to move along a virtual linear track (top) and through the cell's place field (gray). (a,b) Vm during two passes (middle) through each place field shows a sustained subthreshold depolarization under the spiking inside the field (gray boxes correspond to periods when the animal was inside the place fields shown above). Inside the place fields, the cell also fired bursts with large, slow depolarizations (bottom). (c,d) Intrinsic firing pattern in response to a current step (top) for two CA1 pyramidal neurons in rats before they explored a novel maze. During freely moving exploration, the neuron with the intrinsically-bursting firing pattern (c) became a place cell with a place field (middle) and a sustained subthreshold depolarization under the spiking (bottom). The other neuron (d), which had a more regular-spiking firing pattern, became a silent cell that fired very few spikes in the maze (middle) and had a flat subthreshold Vm as a function of the animal's location in the maze as well as a higher spike threshold (dotted line) than the place cell (bottom). Panels (a), (c) and (d) are modified from [82]; panel (b) is modified from [77].
The virtues of reality and virtual reality
Virtual reality systems [75,76*,77**,78] hold the promise of studying many natural behaviors in preparations that allow the use of larger electrophysiology, imaging [107], and other equipment that need not be carried by the animal itself, combined with the manipulability of virtual environments. For instance, the existence of place cells in such preparations [76*,77**] could allow experiments that separate the contributions of self-motion and external sensory information to spatial firing. The absence of vestibular cues required for normal place cell activity in freely moving animals [108–109], though, suggests a need for further investigation into which additional features of place cells and spatial behaviors are preserved with virtual approaches [75,76*,77**]. For these and other (e.g. social interactions) cases that may require more degrees of behavioral freedom, freely moving methods for intracellular recording [44**,79*,80,82] and imaging [110–111] can be employed.
Conclusions and future directions
There are a number of exciting avenues that are just beginning to be explored using awake intracellular recordings. Recent technical achievements have allowed for intracellular recordings to be understood in the context of the anatomical circuit within which these neurons are embedded. For instance, the application of in vivo imaging to target neurons for electrophysiological recordings [112] enables investigators to study specific cell types in head-fixed animals during natural behaviors [17,113]. In addition, advanced methods have emerged for tract tracing [114–115] or circuit reconstruction [116–117] in order to establish the synaptic partners associated with the dynamics observed in vivo. Plasmids can be introduced within the recording pipette which label monosynaptically-connected neurons [118*,119]. If presynaptic neurons express a calcium indicator [120], one may (in theory) observe the input information that is presented to a single neuron. The combination of awake intracellular recording with such approaches could help achieve the ambitious goal of understanding the mechanisms underlying the input/output transformation of each cell in the circuit responsible for a specific behavior.
Acknowledgments
We would like to acknowledge helpful conversations with Vivek Jayaraman, Michael Reiser, Gaby Maimon, Nate Sawtell and Gary Rose. This work was funded by the NIH (M.L.) (DC009280 and NS075044) as well as the Howard Hughes Medical Institute (A.L.). We would also like to thank Daniela Vallentin, Daniel Okobi and Joshua Dudman for valuable comments on the manuscript.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Newsome WT, Britten KH, Movshon JA. Neuronal correlates of a perceptual decision. Nature. 1989;341:52–54. doi: 10.1038/341052a0. [DOI] [PubMed] [Google Scholar]
- 2.Salzman CD, Murasugi CM, Britten KH, Newsome WT. Microstimulation in visual area MT: effects on direction discrimination performance. J Neurosci. 1992;12:2331–2355. doi: 10.1523/JNEUROSCI.12-06-02331.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Georgopoulos AP, Schwartz AB, Kettner RE. Neuronal population coding of movement direction. Science. 1986;233:1416–1419. doi: 10.1126/science.3749885. [DOI] [PubMed] [Google Scholar]
- 4.Hahnloser RH, Kozhevnikov AA, Fee MS. An ultra-sparse code underlies the generation of neural sequences in a songbird. Nature. 2002;419:65–70. doi: 10.1038/nature00974. [DOI] [PubMed] [Google Scholar]
- 5.Yu AC, Margoliash D. Temporal hierarchical control of singing in birds. Science. 1996;273:1871–1875. doi: 10.1126/science.273.5283.1871. [DOI] [PubMed] [Google Scholar]
- 6.O'Keefe J, Dostrovsky J. The hippocampus as a spatial map. Preliminary evidence from unit activity in the freely-moving rat. Brain Res. 1971;34:171–175. doi: 10.1016/0006-8993(71)90358-1. [DOI] [PubMed] [Google Scholar]
- 7.Hafting T, Fyhn M, Molden S, Moser MB, Moser EI. Microstructure of a spatial map in the entorhinal cortex. Nature. 2005;436:801–806. doi: 10.1038/nature03721. [DOI] [PubMed] [Google Scholar]
- 8.Hubel DH, Wiesel TN. Receptive fields, binocular interaction and functional architecture in the cat's visual cortex. J Physiol. 1962;160:106–154. doi: 10.1113/jphysiol.1962.sp006837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Niell CM, Stryker MP. Highly selective receptive fields in mouse visual cortex. J Neurosci. 2008;28:7520–7536. doi: 10.1523/JNEUROSCI.0623-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- *10.Niell CM, Stryker MP. Modulation of visual responses by behavioral state in mouse visual cortex. Neuron. 2010;65:472–479. doi: 10.1016/j.neuron.2010.01.033. The authors investigated the response properties of visual cortex neurons in head-fixed awake mice on a spherical treadmill. Although neurons showed similar orientation tuning properties compared with anesthetized mice, the visually-evoked firing rate varied greatly depending on whether the animal was running or standing still. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Schumacher JW, Schneider DM, Woolley SM. Anesthetic state modulates excitability but not spectral tuning or neural discrimination in single auditory midbrain neurons. J Neurophysiol. 2011;106:500–514. doi: 10.1152/jn.01072.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Borg-Graham L, Monier C, Fregnac Y. Voltage-clamp measurement of visually-evoked conductances with whole-cell patch recordings in primary visual cortex. J Physiol Paris. 1996;90:185–188. doi: 10.1016/s0928-4257(97)81421-0. [DOI] [PubMed] [Google Scholar]
- 13.Anderson JS, Carandini M, Ferster D. Orientation tuning of input conductance, excitation, and inhibition in cat primary visual cortex. J Neurophysiol. 2000;84:909–926. doi: 10.1152/jn.2000.84.2.909. [DOI] [PubMed] [Google Scholar]
- 14.Wehr M, Zador AM. Balanced inhibition underlies tuning and sharpens spike timing in auditory cortex. Nature. 2003;426:442–446. doi: 10.1038/nature02116. [DOI] [PubMed] [Google Scholar]
- 15.Higley MJ, Contreras D. Integration of synaptic responses to neighboring whiskers in rat barrel cortex in vivo. J Neurophysiol. 2005;93:1920–1934. doi: 10.1152/jn.00917.2004. [DOI] [PubMed] [Google Scholar]
- 16.Chiappe ME, Seelig JD, Reiser MB, Jayaraman V. Walking modulates speed sensitivity in Drosophila motion vision. Curr Biol. 2010;20:1470–1475. doi: 10.1016/j.cub.2010.06.072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- **17.Maimon G, Straw AD, Dickinson MH. Active flight increases the gain of visual motion processing in Drosophila. Nat Neurosci. 2010;13:393–399. doi: 10.1038/nn.2492. Using a tethered flight preparation, Maimon and colleagues present the first whole-cell recordings from Drosophila engaged in active flight. They record from genetically labeled visually-responsive interneurons and demonstrate that these neurons have a more depolarized baseline Vm and are also more responsive to visual stimuli during flight compared with non-flight periods. Subsequent measurements of input resistance and Vm fluctuations suggest that the shift in baseline Vm is due to increased synaptic input during flight. [DOI] [PubMed] [Google Scholar]
- 18.Pack CC, Berezovskii VK, Born RT. Dynamic properties of neurons in cortical area MT in alert and anaesthetized macaque monkeys. Nature. 2001;414:905–908. doi: 10.1038/414905a. [DOI] [PubMed] [Google Scholar]
- 19.Rinberg D, Koulakov A, Gelperin A. Sparse odor coding in awake behaving mice. J Neurosci. 2006;26:8857–8865. doi: 10.1523/JNEUROSCI.0884-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Fontanini A, Katz DB. Behavioral states, network states, and sensory response variability. J Neurophysiol. 2008;100:1160–1168. doi: 10.1152/jn.90592.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- *21.Constantinople CM, Bruno RM. Effects and mechanisms of wakefulness on local cortical networks. Neuron. 2011;69:1061–1068. doi: 10.1016/j.neuron.2011.02.040. In their study, Constantinople and Bruno performed whole-cell recordings from rat barrel cortex neurons under isoflurane anesthesia and subsequent wakefulness. The slowly oscillating membrane potential commonly seen during anesthesia is abolished in the awake brain, and they show that this change requires noradernergic but not thalamic inputs. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Steriade M, Timofeev I, Grenier F. Natural waking and sleep states: a view from inside neocortical neurons. J Neurophysiol. 2001;85:1969–1985. doi: 10.1152/jn.2001.85.5.1969. [DOI] [PubMed] [Google Scholar]
- 23.Rudolph M, Pospischil M, Timofeev I, Destexhe A. Inhibition determines membrane potential dynamics and controls action potential generation in awake and sleeping cat cortex. J Neurosci. 2007;27:5280–5290. doi: 10.1523/JNEUROSCI.4652-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Mahon S, Vautrelle N, Pezard L, Slaght SJ, Deniau JM, Chouvet G, Charpier S. Distinct patterns of striatal medium spiny neuron activity during the natural sleep-wake cycle. J Neurosci. 2006;26:12587–12595. doi: 10.1523/JNEUROSCI.3987-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Poulet JF, Petersen CC. Internal brain state regulates membrane potential synchrony in barrel cortex of behaving mice. Nature. 2008;454:881–885. doi: 10.1038/nature07150. [DOI] [PubMed] [Google Scholar]
- 26.Metherate R, Cox CL, Ashe JH. Cellular bases of neocortical activation: modulation of neural oscillations by the nucleus basalis and endogenous acetylcholine. J Neurosci. 1992;12:4701–4711. doi: 10.1523/JNEUROSCI.12-12-04701.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Oren I, Hajos N, Paulsen O. Identification of the current generator underlying cholinergically induced gamma frequency field potential oscillations in the hippocampal CA3 region. J Physiol. 2010;588:785–797. doi: 10.1113/jphysiol.2009.180851. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Froemke RC, Merzenich MM, Schreiner CE. A synaptic memory trace for cortical receptive field plasticity. Nature. 2007;450:425–429. doi: 10.1038/nature06289. [DOI] [PubMed] [Google Scholar]
- 29.Yokota T, Reeves AG, MacLean PD. Intracellular olfactory response of hippocampal neurons in awake, sitting squirrel monkeys. Science. 1967;157:1072–1074. doi: 10.1126/science.157.3792.1072. [DOI] [PubMed] [Google Scholar]
- 30.Baranyi A, Szente MB, Woody CD. Electrophysiological characterization of different types of neurons recorded in vivo in the motor cortex of the cat. I. Patterns of firing activity and synaptic responses. J Neurophysiol. 1993;69:1850–1864. doi: 10.1152/jn.1993.69.6.1850. [DOI] [PubMed] [Google Scholar]
- 31.Matsumura M, Chen D, Sawaguchi T, Kubota K, Fetz EE. Synaptic interactions between primate precentral cortex neurons revealed by spike-triggered averaging of intracellular membrane potentials in vivo. J Neurosci. 1996;16:7757–7767. doi: 10.1523/JNEUROSCI.16-23-07757.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- *32.Jortner RA, Farivar SS, Laurent G. A simple connectivity scheme for sparse coding in an olfactory system. J Neurosci. 2007;27:1659–1669. doi: 10.1523/JNEUROSCI.4171-06.2007. In an insect olfactory preparation, the authors perform in vivo intracellular recordings of a postsynaptic neuron in conjunction with multielectrode recordings of its presynaptic neuron population to address the connectivity underlying the sparse firing of the postsynaptic neuron. Through spike-triggered averaging of the postsynaptic Vm with respect to spontaneous spikes of the presynaptic projection neurons, they determine that each of these postsynaptic neurons is connected to approximately half of its input population. They show that the effective size of each input is small and estimate the number of simultaneous inputs needed to fire a target cell. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Haag J, Wertz A, Borst A. Central gating of fly optomotor response. Proc Natl Acad Sci U S A. 2010;107:20104–20109. doi: 10.1073/pnas.1009381107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Covey E, Kauer JA, Casseday JH. Whole-cell patch-clamp recording reveals subthreshold sound-evoked postsynaptic currents in the inferior colliculus of awake bats. J Neurosci. 1996;16:3009–3018. doi: 10.1523/JNEUROSCI.16-09-03009.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Margrie TW, Brecht M, Sakmann B. In vivo, low-resistance, whole-cell recordings from neurons in the anaesthetized and awake mammalian brain. Pflugers Arch. 2002;444:491–498. doi: 10.1007/s00424-002-0831-z. [DOI] [PubMed] [Google Scholar]
- 36.Wilson RI, Turner GC, Laurent G. Transformation of olfactory representations in the Drosophila antennal lobe. Science. 2004;303:366–370. doi: 10.1126/science.1090782. [DOI] [PubMed] [Google Scholar]
- 37.Huston SJ, Krapp HG. Nonlinear integration of visual and haltere inputs in fly neck motor neurons. J Neurosci. 2009;29:13097–13105. doi: 10.1523/JNEUROSCI.2915-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Papadopoulou M, Cassenaer S, Nowotny T, Laurent G. Normalization for sparse encoding of odors by a wide-field interneuron. Science. 2011;332:721–725. doi: 10.1126/science.1201835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Okun M, Naim A, Lampl I. The subthreshold relation between cortical local field potential and neuronal firing unveiled by intracellular recordings in awake rats. J Neurosci. 2010;30:4440–4448. doi: 10.1523/JNEUROSCI.5062-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Swadlow HA, Beloozerova IN, Sirota MG. Sharp, local synchrony among putative feed-forward inhibitory interneurons of rabbit somatosensory cortex. J Neurophysiol. 1998;79:567–582. doi: 10.1152/jn.1998.79.2.567. [DOI] [PubMed] [Google Scholar]
- 41.Xie R, Gittelman JX, Li N, Pollak GD. Whole cell recordings of intrinsic properties and sound-evoked responses from the inferior colliculus. Neuroscience. 2008;154:245–256. doi: 10.1016/j.neuroscience.2008.02.039. [DOI] [PubMed] [Google Scholar]
- **42.Aksay E, Gamkrelidze G, Seung HS, Baker R, Tank DW. In vivo intracellular recording and perturbation of persistent activity in a neural integrator. Nat Neurosci. 2001;4:184–193. doi: 10.1038/84023. Sharp, whole-cell, and extracellular recordings were obtained from oculomotor neurons of awake, head-restrained goldfish with natural saccadic and fixation eye behaviors. The authors use Vm recordings as well as single-neuron perturbations as evidence in support of a network-based as opposed to single-neuron-based explanation for persistent activity in this eye control circuit. See text (“Persistent activity”) for more details. [DOI] [PubMed] [Google Scholar]
- **43.Edwards CJ, Leary CJ, Rose GJ. Counting on inhibition and rate-dependent excitation in the auditory system. J Neurosci. 2007;27:13384–13392. doi: 10.1523/JNEUROSCI.2816-07.2007. In this paper, Edwards and colleagues recorded intracellularly from “interval-counting” neurons in the frog auditory midbrain in vivo. They demonstrate that synaptic inhibition and short-term facilitation of excitation allow single neurons to respond invariantly to a specific temporal pattern of pulses within a vocal-like communication signal. See text (“Case Study: Feature selectivity”) for more details. [DOI] [PMC free article] [PubMed] [Google Scholar]
- **44.Long MA, Jin DZ, Fee MS. Support for a synaptic chain model of neuronal sequence generation. Nature. 2010;468:394–399. doi: 10.1038/nature09514. In this study, Long and colleagues used a chronically implanted, lightweight (1.6g) microdrive capable of recording the Vm of single neurons in freely behaving small animals in order to provide evidence for a synaptic chain within a pattern generating premotor nucleus of the songbird. See text (“Motor sequence generation”) for more details. [DOI] [PMC free article] [PubMed] [Google Scholar]
- **45.Crochet S, Poulet JF, Kremer Y, Petersen CC. Synaptic mechanisms underlying sparse coding of active touch. Neuron. 2011;69:1160–1175. doi: 10.1016/j.neuron.2011.02.022. Whole-cell recordings were performed in the barrel cortex of awake, head-fixed mice while they were actively contacting an object with their whiskers. Such active touch results in spiking by only a small fraction of layer 2/3 neurons. Each neuron is driven to a particular Vm level upon contact regardless of the pre-contact Vm level, reflecting activation of a specific combination of excitatory and inhibitory conductances. Only for neurons in which this potential is greater than the spike threshold will spikes consistently be fired in response to object contact. See text (“Active sensation”) for more details. [DOI] [PubMed] [Google Scholar]
- 46.Edwards CJ, Alder TB, Rose GJ. Auditory midbrain neurons that count. Nat Neurosci. 2002;5:934–936. doi: 10.1038/nn916. [DOI] [PubMed] [Google Scholar]
- 47.Bell CC. Memory-based expectations in electrosensory systems. Curr Opin Neurobiol. 2001;11:481–487. doi: 10.1016/s0959-4388(00)00238-5. [DOI] [PubMed] [Google Scholar]
- 48.Sawtell NB. Multimodal integration in granule cells as a basis for associative plasticity and sensory prediction in a cerebellum-like circuit. Neuron. 2010;66:573–584. doi: 10.1016/j.neuron.2010.04.018. [DOI] [PubMed] [Google Scholar]
- 49.Poulet JF, Hedwig B. A corollary discharge maintains auditory sensitivity during sound production. Nature. 2002;418:872–876. doi: 10.1038/nature00919. [DOI] [PubMed] [Google Scholar]
- **50.Poulet JF, Hedwig B. The cellular basis of a corollary discharge. Science. 2006;311:518–522. doi: 10.1126/science.1120847. In this pair of papers, the authors use in vivo intracellular recordings to identify the source of singing-related corollary discharge within the cricket as a single multisegmental interneuron. This cell feeds back onto auditory sensory neurons in order to protect them against desensitization during singing. [DOI] [PubMed] [Google Scholar]
- 51.Welker WI. Analysis of sniffing of the albino rat. Behaviour. 1964;22:223–244. [Google Scholar]
- 52.Berg RW, Kleinfeld D. Rhythmic whisking by rat: retraction as well as protraction of the vibrissae is under active muscular control. J Neurophysiol. 2003;89:104–117. doi: 10.1152/jn.00600.2002. [DOI] [PubMed] [Google Scholar]
- 53.Fee MS, Mitra PP, Kleinfeld D. Central versus peripheral determinants of patterned spike activity in rat vibrissa cortex during whisking. J Neurophysiol. 1997;78:1144–1149. doi: 10.1152/jn.1997.78.2.1144. [DOI] [PubMed] [Google Scholar]
- 54.Brecht M, Sakmann B. Whole-cell recordings in rat somatosensory cortex during whisking episodes. Society for Neuroscience Annual Meeting; San Diego, CA. 2001. [Google Scholar]
- 55.Crochet S, Petersen CC. Correlating whisker behavior with membrane potential in barrel cortex of awake mice. Nat Neurosci. 2006;9:608–610. doi: 10.1038/nn1690. [DOI] [PubMed] [Google Scholar]
- 56.Curtis JC, Kleinfeld D. Phase-to-rate transformations encode touch in cortical neurons of a scanning sensorimotor system. Nat Neurosci. 2009;12:492–501. doi: 10.1038/nn.2283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Brecht M, Roth A, Sakmann B. Dynamic receptive fields of reconstructed pyramidal cells in layers 3 and 2 of rat somatosensory barrel cortex. J Physiol. 2003;553:243–265. doi: 10.1113/jphysiol.2003.044222. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Kerr JN, de Kock CP, Greenberg DS, Bruno RM, Sakmann B, Helmchen F. Spatial organization of neuronal population responses in layer 2/3 of rat barrel cortex. J Neurosci. 2007;27:13316–13328. doi: 10.1523/JNEUROSCI.2210-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Sato TR, Gray NW, Mainen ZF, Svoboda K. The functional microarchitecture of the mouse barrel cortex. PLoS Biol. 2007;5:e189. doi: 10.1371/journal.pbio.0050189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.de Kock CP, Bruno RM, Spors H, Sakmann B. Layer- and cell-type-specific suprathreshold stimulus representation in rat primary somatosensory cortex. J Physiol. 2007;581:139–154. doi: 10.1113/jphysiol.2006.124321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.O’Connor DH, Peron SP, Huber D, Svoboda K. Neural activity in barrel cortex underlying vibrissa-based object localization in mice. Neuron. 2010;67:1048–1061. doi: 10.1016/j.neuron.2010.08.026. [DOI] [PubMed] [Google Scholar]
- 62.Golshani P, Goncalves JT, Khoshkhoo S, Mostany R, Smirnakis S, Portera-Cailliau C. Internally mediated developmental desynchronization of neocortical network activity. J Neurosci. 2009;29:10890–10899. doi: 10.1523/JNEUROSCI.2012-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Goldman-Rakic PS. Cellular basis of working memory. Neuron. 1995;14:477–485. doi: 10.1016/0896-6273(95)90304-6. [DOI] [PubMed] [Google Scholar]
- 64.Romo R, Brody CD, Hernandez A, Lemus L. Neuronal correlates of parametric working memory in the prefrontal cortex. Nature. 1999;399:470–473. doi: 10.1038/20939. [DOI] [PubMed] [Google Scholar]
- 65.Pastor AM, De la Cruz RR, Baker R. Eye position and eye velocity integrators reside in separate brainstem nuclei. Proc Natl Acad Sci U S A. 1994;91:807–811. doi: 10.1073/pnas.91.2.807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Egorov AV, Hamam BN, Fransen E, Hasselmo ME, Alonso AA. Graded persistent activity in entorhinal cortex neurons. Nature. 2002;420:173–178. doi: 10.1038/nature01171. [DOI] [PubMed] [Google Scholar]
- 67.Goldman MS, Levine JH, Major G, Tank DW, Seung HS. Robust persistent neural activity in a model integrator with multiple hysteretic dendrites per neuron. Cereb Cortex. 2003;13:1185–1195. doi: 10.1093/cercor/bhg095. [DOI] [PubMed] [Google Scholar]
- 68.Seung HS, Lee DD, Reis BY, Tank DW. Stability of the memory of eye position in a recurrent network of conductance-based model neurons. Neuron. 2000;26:259–271. doi: 10.1016/s0896-6273(00)81155-1. [DOI] [PubMed] [Google Scholar]
- **69.Gaudry Q, Kristan WB., Jr Behavioral choice by presynaptic inhibition of tactile sensory terminals. Nat Neurosci. 2009;12:1450–1457. doi: 10.1038/nn.2400. Both intracellular and extracellular recordings from a behaving medicinal leech were made in order to examine the changes that underlie the decreased response to mechanosensory stimuli during feeding. Specifically, the magnitude of mechanosensory inputs was reduced through presynaptic inhibition mediated primarily via serotonin. [DOI] [PubMed] [Google Scholar]
- 70.Glenn LL, Whitney JF, Rewitzer JS, Salamone JA, Mariash SA. Method for stable intracellular recordings of spinal alpha-motoneurons during treadmill walking in awake, intact cats. Brain Res. 1988;439:396–401. doi: 10.1016/0006-8993(88)91502-8. [DOI] [PubMed] [Google Scholar]
- *71.Fee MS. Active stabilization of electrodes for intracellular recording in awake behaving animals. Neuron. 2000;27:461–468. doi: 10.1016/s0896-6273(00)00057-x. This report describes a novel active stabilization system which was developed to compensate for relative movements between a sharp microelectrode and the brain in awake head-fixed animals. This system was designed to compensate for both periodic sources of biological motion (heart beat and respiration) as well as irregular movements of the animal. Specifically, in response to measured and predicted movement sources, a piezoelectric motor moved the electrode along its axis. [DOI] [PubMed] [Google Scholar]
- 72.Kerfoot WB. Orientometrer for study of insect behavior. Science. 1968;162:477. doi: 10.1126/science.162.3852.477. [DOI] [PubMed] [Google Scholar]
- 73.Buchner E. Elementary Movement Detectors in an Insect Visual System. Biol Cybern. 1976;24:85–101. [Google Scholar]
- 74.Seelig JD, Chiappe ME, Lott GK, Dutta A, Osborne JE, Reiser MB, Jayaraman V. Two-photon calcium imaging from head-fixed Drosophila during optomotor walking behavior. Nat Methods. 2010;7:535–540. doi: 10.1038/nmeth.1468. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Holscher C, Schnee A, Dahmen H, Setia L, Mallot HA. Rats are able to navigate in virtual environments. J Exp Biol. 2005;208:561–569. doi: 10.1242/jeb.01371. [DOI] [PubMed] [Google Scholar]
- *76.Schnee A. PhD thesis. Eberhard Karls Universität; Tübingen: 2008. Rats in virtual reality: The Development of an Advanced Method to Study Animal Behaviour. [Mallot HA (Advisor)]. vol PhD. A doctoral thesis in which a body-tethered, spherical treadmill-based virtual reality paradigm was used to investigate animal navigation in rats. Using extracellular methods, hippocampal neurons were shown to exhibit place cell activity, spiking whenever the animal was in a particular location within a virtual maze. [Google Scholar]
- **77.Harvey CD, Collman F, Dombeck DA, Tank DW. Intracellular dynamics of hippocampal place cells during virtual navigation. Nature. 2009;461:941–946. doi: 10.1038/nature08499. This paper presents a novel method for obtaining intracellular recordings in spatially-navigating rodents. Mice were trained to run on a spherical treadmill through a virtual reality environment. The authors recorded place cells intracellularly and showed that the subthreshold Vm of place fields were asymmetrically-skewed. They also simultaneously recorded the hippocampal LFP theta rhythm during running and compared it to the intracellular Vm oscillations, with implications for theta rhythm-related spiking. See text (“Spatial navigation”) for more details. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Dombeck DA, Reiser MB. Real neuroscience in virtual worlds. Curr Opin Neurobiol. 2012;22:3–10. doi: 10.1016/j.conb.2011.10.015. [DOI] [PubMed] [Google Scholar]
- *79.Lee AK, Manns ID, Sakmann B, Brecht M. Whole-cell recordings in freely moving rats. Neuron. 2006;51:399–407. doi: 10.1016/j.neuron.2006.07.004. The first report demonstrating intracellular recordings in freely moving animals. Methodologically, it showed that the principle of rigidly fixing the recording pipette with respect to the skull could result in long-lasting (up to one hour), high-quality (stable baseline Vm, low in vivo series resistance) whole-cell recordings in freely moving animals. [DOI] [PubMed] [Google Scholar]
- 80.Lee AK, Epsztein J, Brecht M. Head-anchored whole-cell recordings in freely moving rats. Nat Protoc. 2009;4:385–392. doi: 10.1038/nprot.2009.5. [DOI] [PubMed] [Google Scholar]
- *81.Epsztein J, Lee AK, Chorev E, Brecht M. Impact of spikelets on hippocampal CA1 pyramidal cell activity during spatial exploration. Science. 2010;327:474–477. doi: 10.1126/science.1182773. Spikelets were found to occur in hippocampal CA1 pyramidal neurons using whole-cell recordings in freely moving rats. These events were seen both in isolation as well as immediately preceeding spikes. Importantly, spikelets are too small to be detected by extracellular methods, but can strongly influence the spiking behavior of hippocampal neurons. See text (“Spatial navigation”) for more details. [DOI] [PubMed] [Google Scholar]
- **82.Epsztein J, Brecht M, Lee AK. Intracellular determinants of hippocampal CA1 place and silent cell activity in a novel environment. Neuron. 2011;70:109–120. doi: 10.1016/j.neuron.2011.03.006. Intracellular recordings in freely moving rats revealed that both input-based depolarizations and, unexpectedly, intrinsic properties (e.g. spike threshold) determine whether a hippocampal CA1 pyramidal neuron will be a place cell or remain silent within a given maze environment. Further, intrinsic properties appeared to predetermine which cells would become place cells, based on cellular responses to current step injection before exploration. Also, large, putative calcium-based “complex spikes” were found to fire preferentially within the center of the place field. See text (“Spatial navigation”) for more details. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Amari S. Learning patterns and pattern sequences by self-organizing nets of threshold elements. IEEE Transactions on Computers. 1972;c-21:1197–1206. [Google Scholar]
- 84.Abeles M. Corticonics: Neural Circuits of the Cerebral Cortex. Cambridge University Press; 1991. [Google Scholar]
- 85.Ikegaya Y, Aaron G, Cossart R, Aronov D, Lampl I, Ferster D, Yuste R. Synfire chains and cortical songs: temporal modules of cortical activity. Science. 2004;304:559–564. doi: 10.1126/science.1093173. [DOI] [PubMed] [Google Scholar]
- 86.Mokeichev A, Okun M, Barak O, Katz Y, Ben-Shahar O, Lampl I. Stochastic emergence of repeating cortical motifs in spontaneous membrane potential fluctuations in vivo. Neuron. 2007;53:413–425. doi: 10.1016/j.neuron.2007.01.017. [DOI] [PubMed] [Google Scholar]
- 87.Ikegaya Y, Matsumoto W, Chiou HY, Yuste R, Aaron G. Statistical significance of precisely repeated intracellular synaptic patterns. PLoS One. 2008;3:e3983. doi: 10.1371/journal.pone.0003983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Jin DZ, Ramazanoglu FM, Seung HS. Intrinsic bursting enhances the robustness of a neural network model of sequence generation by avian brain area HVC. J Comput Neurosci. 2007;23:283–299. doi: 10.1007/s10827-007-0032-z. [DOI] [PubMed] [Google Scholar]
- 89.Li M, Greenside H. Stable propagation of a burst through a one-dimensional homogeneous excitatory chain model of songbird nucleus HVC. Phys Rev E Stat Nonlin Soft Matter Phys. 2006;74:011918. doi: 10.1103/PhysRevE.74.011918. [DOI] [PubMed] [Google Scholar]
- 90.Long MA, Fee MS. Using temperature to analyse temporal dynamics in the songbird motor pathway. Nature. 2008;456:189–194. doi: 10.1038/nature07448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Fee MS, Long MA. New methods for localizing and manipulating neuronal dynamics in behaving animals. Curr Opin Neurobiol. 2011;21:693–700. doi: 10.1016/j.conb.2011.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Mooney R. Different subthreshold mechanisms underlie song selectivity in identified HVc neurons of the zebra finch. J Neurosci. 2000;20:5420–5436. doi: 10.1523/JNEUROSCI.20-14-05420.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.O'Keefe J, Nadel L. The Hippocampus as a Cognitive Map. Oxford University Press; 1978. [Google Scholar]
- 94.Morris RG, Garrud P, Rawlins JN, O'Keefe J. Place navigation impaired in rats with hippocampal lesions. Nature. 1982;297:681–683. doi: 10.1038/297681a0. [DOI] [PubMed] [Google Scholar]
- 95.Thompson LT, Best PJ. Place cells and silent cells in the hippocampus of freely-behaving rats. J Neurosci. 1989;9:2382–2390. doi: 10.1523/JNEUROSCI.09-07-02382.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Pastalkova E, Itskov V, Amarasingham A, Buzsaki G. Internally generated cell assembly sequences in the rat hippocampus. Science. 2008;321:1322–1327. doi: 10.1126/science.1159775. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Lee AK, Epsztein J, Brecht M. Whole-cell recordings of hippocampal CA1 place cell activity in freely moving rats. Society for Neuroscience Annual Meeting; Washington, D.C. 2008. [Google Scholar]
- 98.Mehta MR, Quirk MC, Wilson MA. Experience-dependent asymmetric shape of hippocampal receptive fields. Neuron. 2000;25:707–715. doi: 10.1016/s0896-6273(00)81072-7. [DOI] [PubMed] [Google Scholar]
- 99.Golding NL, Jung HY, Mickus T, Spruston N. Dendritic calcium spike initiation and repolarization are controlled by distinct potassium channel subtypes in CA1 pyramidal neurons. J Neurosci. 1999;19:8789–8798. doi: 10.1523/JNEUROSCI.19-20-08789.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Kandel ER, Spencer WA. Electrophysiology of hippocampal neurons. II. After-potentials and repetitive firing. J Neurophysiol. 1961;24:243–259. doi: 10.1152/jn.1961.24.3.243. [DOI] [PubMed] [Google Scholar]
- 101.Takahashi H, Magee JC. Pathway interactions and synaptic plasticity in the dendritic tuft regions of CA1 pyramidal neurons. Neuron. 2009;62:102–111. doi: 10.1016/j.neuron.2009.03.007. [DOI] [PubMed] [Google Scholar]
- 102.Sjostrom PJ, Nelson SB. Spike timing, calcium signals and synaptic plasticity. Curr Opin Neurobiol. 2002;12:305–314. doi: 10.1016/s0959-4388(02)00325-2. [DOI] [PubMed] [Google Scholar]
- 103.Morris RG. Synaptic plasticity and learning: selective impairment of learning rats and blockade of long-term potentiation in vivo by the N-methyl-D-aspartate receptor antagonist AP5. J Neurosci. 1989;9:3040–3057. doi: 10.1523/JNEUROSCI.09-09-03040.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Bannerman DM, Good MA, Butcher SP, Ramsay M, Morris RG. Distinct components of spatial learning revealed by prior training and NMDA receptor blockade. Nature. 1995;378:182–186. doi: 10.1038/378182a0. [DOI] [PubMed] [Google Scholar]
- 105.Tsien JZ, Huerta PT, Tonegawa S. The essential role of hippocampal CA1 NMDA receptor-dependent synaptic plasticity in spatial memory. Cell. 1996;87:1327–1338. doi: 10.1016/s0092-8674(00)81827-9. [DOI] [PubMed] [Google Scholar]
- 106.Zhou Y, Won J, Karlsson MG, Zhou M, Rogerson T, Balaji J, Neve R, Poirazi P, Silva AJ. CREB regulates excitability and the allocation of memory to subsets of neurons in the amygdala. Nat Neurosci. 2009;12:1438–1443. doi: 10.1038/nn.2405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Dombeck DA, Harvey CD, Tian L, Looger LL, Tank DW. Functional imaging of hippocampal place cells at cellular resolution during virtual navigation. Nat Neurosci. 2010;13:1433–1440. doi: 10.1038/nn.2648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Stackman RW, Clark AS, Taube JS. Hippocampal spatial representations require vestibular input. Hippocampus. 2002;12:291–303. doi: 10.1002/hipo.1112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Russell NA, Horii A, Smith PF, Darlington CL, Bilkey DK. Long-term effects of permanent vestibular lesions on hippocampal spatial firing. J Neurosci. 2003;23:6490–6498. doi: 10.1523/JNEUROSCI.23-16-06490.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Sawinski J, Wallace DJ, Greenberg DS, Grossmann S, Denk W, Kerr JN. Visually evoked activity in cortical cells imaged in freely moving animals. Proc Natl Acad Sci U S A. 2009;106:19557–19562. doi: 10.1073/pnas.0903680106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Flusberg BA, Nimmerjahn A, Cocker ED, Mukamel EA, Barretto RP, Ko TH, Burns LD, Jung JC, Schnitzer MJ. High-speed, miniaturized fluorescence microscopy in freely moving mice. Nat Methods. 2008;5:935–938. doi: 10.1038/nmeth.1256. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Margrie TW, Meyer AH, Caputi A, Monyer H, Hasan MT, Schaefer AT, Denk W, Brecht M. Targeted whole-cell recordings in the mammalian brain in vivo. Neuron. 2003;39 :911–918. doi: 10.1016/j.neuron.2003.08.012. [DOI] [PubMed] [Google Scholar]
- 113.Gentet LJ, Avermann M, Matyas F, Staiger JF, Petersen CC. Membrane potential dynamics of GABAergic neurons in the barrel cortex of behaving mice. Neuron. 2010;65:422–435. doi: 10.1016/j.neuron.2010.01.006. [DOI] [PubMed] [Google Scholar]
- 114.Wall NR, Wickersham IR, Cetin A, De La Parra M, Callaway EM. Monosynaptic circuit tracing in vivo through Cre-dependent targeting and complementation of modified rabies virus. Proc Natl Acad Sci U S A. 2010;107:21848–21853. doi: 10.1073/pnas.1011756107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Miyamichi K, Amat F, Moussavi F, Wang C, Wickersham I, Wall NR, Taniguchi H, Tasic B, Huang ZJ, He Z, et al. Cortical representations of olfactory input by trans-synaptic tracing. Nature. 2011;472:191–196. doi: 10.1038/nature09714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Briggman KL, Helmstaedter M, Denk W. Wiring specificity in the direction-selectivity circuit of the retina. Nature. 2011;471:183–188. doi: 10.1038/nature09818. [DOI] [PubMed] [Google Scholar]
- 117.Bock DD, Lee WC, Kerlin AM, Andermann ML, Hood G, Wetzel AW, Yurgenson S, Soucy ER, Kim HS, Reid RC. Network anatomy and in vivo physiology of visual cortical neurons. Nature. 2011;471:177–182. doi: 10.1038/nature09802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- *118.Rancz EA, Franks KM, Schwarz MK, Pichler B, Schaefer AT, Margrie TW. Transfection via whole-cell recording in vivo: bridging single-cell physiology, genetics and connectomics. Nat Neurosci. 2011;14:527–532. doi: 10.1038/nn.2765. In this study, a plasmid is introduced into individual neurons through the intracellular recording solution (whole-cell configuration) which can allow for the specific labeling of presynaptic neurons using viral techniques. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Wickersham IR, Lyon DC, Barnard RJ, Mori T, Finke S, Conzelmann KK, Young JA, Callaway EM. Monosynaptic restriction of transsynaptic tracing from single, genetically targeted neurons. Neuron. 2007;53:639–647. doi: 10.1016/j.neuron.2007.01.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Tian L, Hires SA, Mao T, Huber D, Chiappe ME, Chalasani SH, Petreanu L, Akerboom J, McKinney SA, Schreiter ER, et al. Imaging neural activity in worms, flies and mice with improved GCaMP calcium indicators. Nat Methods. 2009;6:875–881. doi: 10.1038/nmeth.1398. [DOI] [PMC free article] [PubMed] [Google Scholar]




