U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Noebels JL, Avoli M, Rogawski MA, et al., editors. Jasper's Basic Mechanisms of the Epilepsies [Internet]. 4th edition. Bethesda (MD): National Center for Biotechnology Information (US); 2012.

  • This title is an author manuscript version first made accessible on the NCBI Bookshelf website July 2, 2012.

This title is an author manuscript version first made accessible on the NCBI Bookshelf website July 2, 2012.

Cover of Jasper's Basic Mechanisms of the Epilepsies

Jasper's Basic Mechanisms of the Epilepsies [Internet]. 4th edition.

Show details

Neuronal Synchronization and Thalamocortical Rhythms in Sleep, Wake and Epilepsy

, , , and .

Author Information and Affiliations

Neuronal synchronization can be defined as a correlated appearance in time of two or more events associated with various aspects of neuronal activity. Neuronal synchronization depends on chemical and electrical synaptic as well as ephaptic and non-specific interactions. We consider two distinct types of neuronal synchronization: local synchronization that is responsible for the generation of local field potentials; and long-range synchronization, detected with distantly located electrodes and mediated primarily via chemical synaptic interactions, which contributes to the EEG synchronization. Neocortical synchronization during sleep and wakefulness is often associated with rhythmic oscillations of neuronal activity: slow oscillation, delta, spindle, beta, gamma and ripples. Normal thalamocortical oscillations [sleep or wake oscillations] are generated as a result of both local and long-range synchronization. During paroxysmal [seizure] activity, the role of chemical synaptic interactions decreases because of alterations in ionic composition that impairs synaptic transmission. Synchronized activities in large population of neurons (such as neocortex) may occur as nearly simultaneous patterns across an entire population or as propagating waves. Neocortical synchronization is controlled by the activities in ascending systems: cholinergic, norepinephrinergic and serotoninergic. The presence of cortico-thalamo-cortical feedback loops contribute to the synchronization of cortical activities. We propose that all the types of neuronal interactions contribute to the generation of synchronous oscillatory activities, but the ratio of their contribution is different for different types of oscillations.

Neuronal synchronization can be distinguished into long-range and local synchronization. The long-range synchrony is usually detected with two or more electrodes placed at some distance apart. It leads to brain activity that is correlated at long distances and may be seen using both local filed potential (LFP) and electroencephalogram (EEG) recordings. The first tool (i.e., the LFP) provides a microscopic measure of brain activity summarizing electrical activities of possibly thousands of neurons 1–4. The second type of recording (i.e., the EEG) is a result of changes of electrical activity of multiple sources and ultimately represents activity patterns of large populations of neurons and glial cells in the brain. The local or short-range synchrony can be detected either with one relatively large field potential electrode or with two or more small, intracellular or extracellular unit (action potential) recording electrodes located at short (less than 1 mm) distances from each other. Synchronous activity of a few neurons not necessarily leads to measurable EEG signals, but can be seen using LFP recordings. Because of the low-pass filtering properties of the extracellular media 5, high frequency electric field associated with action potentials steeply attenuate and the large amplitude slow LFP and EEG potentials are mainly generated from nearly simultaneously occurring de- and hyperpolarizing events in a large number of neighboring cells with a major contribution of large pyramidal neurons 6.

In general, synchronization of electrical activity in the brain occurs as the result of interaction among sets of neurons. The following major types of interactions between brain cells (neurons and glia) are possible: (a) Chemical synaptic transmission, which takes place when an action potential that is fired by the presynaptic neuron, invades the nerve terminals, allowing Ca2+ entry, triggering neurotransmitter release, and finally causing de- or hyperpolarization of the postsynaptic neuron 7,8; (b) Interactions between electrically coupled cells via gap junctions 9; these interactions are bidirectional and work as a low-pass filter, i.e., fast processes (for example: action potentials) are filtered out, but slow processes (for example: slow synaptic potentials) are transmitted with minor attenuation 10; (c) Ephaptic interactions, which are mediated by large extracellular currents and field potentials generated by cortical structures in the absence of specialized contacts 11, 12 (it should be noted that some researchers separate ephaptic transmission from field effect interactions 13); (d) Non-specific interactions, commonly achieved by alteration in the extracellular ionic balance caused by the activity of brain cells 14–17.

MECHANISMS OF NEURONAL SYNCHRONIZATION

Neuronal synchronization is defined as correlated appearance in time of two or more events associated with various aspects of neuronal activity at different levels, from single cell to the whole brain.

A common scenario involves adjustment or phase locking of rhythms of two or more neurons leading to a stable phase difference of membrane voltage oscillations (periodic or not), e.g., coincidence of action potentials. Such coincidence of electrical activities of many neurons would lead to coordinated changes of extracellular ionic currents in close proximity to these neurons, which can be then detected as changes of the LFP, which are reflected on the scalp EEG. If individual neurons fire action potentials periodically, and these events are synchronized among many cells, one would observe periodic LFP oscillations. Finally, if long-range synchronization takes place among distinct brain areas, it would lead to global changes of electrical activity usually associated with EEG rhythms.

Chemical Synaptic Mechanisms of Synchronization

Synaptic interaction is a common mechanism of communication between neurons. Transmitter release by presynaptic terminal leads to receptor activation on the postsynaptic cell and the generation of inward or outward currents, which depolarize or hyperpolarize the postsynaptic cell. Summation of activities from synchronized inputs enhances transmission of information more effectively than raising discharge rates 18. Generally both excitatory and inhibitory synaptic interaction may contribute to synchronization of neuronal activity. For example, in a minimal circuit including synaptic connection from cell A to cell B the action potential in cell A would trigger excitatory postsynaptic potential (EPSP) in cell B. If this potential is ample enough, it can trigger an action potential in cell B, which may occur nearly simultaneously (after small delay) with action potential in cell A.

This process is more complex in real cortical networks. The amplitude of the single-axon EPSPs is small and variable. In vitro, it ranges from 0.1 to 2 mV (but sometimes up to 9 mV), with a mean of about 1 mV 19–23. In vivo, due to multiple factors associated with network activities (e.g., shunting or frequency dependent attenuation), the amplitude of successful single axon EPSPs is about 0.5 mV 15. Given the fact that the EPSP amplitude depends linearly on the number of synaptic contacts formed by all the presynaptic neurons 21, 23, only 40 to 50 synaptic contacts are needed to bring the postsynaptic neuron membrane potential from resting level of −75 mV to −80 mV to firing threshold of around −55 mV in vivo. On the other hand, the actual rate of EPSPs received by a neocortical neuron in vivo may be much higher. Indeed, spontaneous firing rates in vivo are at the order of 10 Hz 24, 25. Large neocortical neurons possess between 10,000 and 50,000 synapses 26. Assuming 25,000 synapses on average and considering that eighty percent of these synapses are excitatory 27, about 200,000 synapses may be activated each second, or 200 synapses each millisecond (80% from 25,000 synapses * 10 Hz = 200.000 per sec or 200 per msec). The half-width of EPSPs in vivo is about 10 ms and this time might be assumed as a period for efficient integration of presynaptic inputs. Thus, a neocortical neuron can receive up to 2000 inputs during 10 ms window for the efficient integration. Because the failure rates in vivo are very high (70–80%) 15 the neuron receives EPSPs in the order of 400–800 within integrating period. This is still 10–20 times more than the needed number of simultaneously firing presynaptic neurons to fire postsynaptic neuron (40–50 synapses [see above]). Therefore, the mechanisms of cortical synchronization based exclusively on excitatory intracortical connectivity would lead to over-excitation of the network and cannot be considered as an efficient mechanism of cortical synchronization.

Efficient way to synchronize population of excitatory cells involves inhibition. Imagine a group of excitatory neurons A1,…, An that generate action potentials, which are not synchronized across cells. Imagine now that all these neurons receive common inhibitory input from inhibitory cell B. Action potential in cell B would provide synchronous inhibitory postsynaptic potentials (IPSPs) in neurons A1,…, An, so, when the IPSP terminates, all these neurons A1,…, An can spike synchronously. If now some of these excitatory cells project back to inhibitory neuron B, synchronous spiking between A1,…, An would trigger a new spike in cell B therefore starting a new cycle of oscillation (Figure 1). This mechanism of synchronization commonly referred to as “feedback inhibition” (see 28 for review of this topic), is involved in many brain rhythms in different brain systems including, e.g., some types of gamma oscillations 29, 30, thalamic spindles 1, 31 and others.

Figure 1. Effect of feedback inhibition in spike synchronization.

Figure 1

Effect of feedback inhibition in spike synchronization. Top, eight pyramidal neurons (PYs, thin lines of different color) oscillate asynchronously in 20–25 Hz frequency range with no inhibitory feedback. About 200 ms after inhibitory GABAergic (more...)

Gap Junctions and Synchronization

Gap junctions (i.e., electrical synapses) allow direct flux of ions between connected cells therefore providing a mechanism of communication that is action potential independent 32, 33. If few cells are connected through electrical synapses, any change of membrane voltage in one neuron would trigger current flow between coupled neurons leading to corresponding changes of membrane voltage in the latter cells, therefore, providing synchronizing effect. Because of the relatively high resistance of electrical synapses, they act as low pass filters 10. In cortical networks, groups of GABA-releasing interneurons 34, 35 and glial cells 36 are interconnected via gap junctions. There is a set of indirect evidences suggesting that electrical coupling may occur between axons of pyramidal neurons 37, 38. The role of axonal gap junctions in the generation of electrical activities is unclear. Intra-axonal recordings did not reveal the presence of afterhyperpolarizing potentials (AHP) in axons 39, 40. Low-pass filter properties of gap junctions prevent efficient transmitting of fast spikes to coupled neurons and absence of AHP prevents slower synchronization. Because of this and because of extreme short-range of electrical coupling, while axo-axonal gap junctions may enhance already existing synchronization manifested by slow membrane voltage oscillations, it is unlikely that they play a primary role in establishing neocortical synchronization.

Ephaptic Interactions

Extracellular currents produced by electrical activity of neurons and constituting local field potentials may directly influence electrical properties of neurons 13. This effect may include depolarization or hyperpolarization of cell membrane, and therefore a change in excitability. While being relatively weak, these effects have a global influence and may provide significant impact when neuronal activity is already synchronized by means of other mechanisms. Examples include slow-wave sleep oscillations or epileptic activity when a certain degree of synchrony of electrical activity between neurons is already achieved through chemical or electrical interactions. In that case, weak but global effect of ephaptic interaction further enhances the synchrony. Not only internally generated fields can affect neuronal excitability. The application of external electric field with an amplitude within the range of in vivo endogenous fields modulates cortical slow oscillation 12, 41 and potentiates memory 41.

Changes of the Extracellular Ionic Concentrations

Electrical activity of neurons is associated with opening and closing of different ionic channels and therefore may lead to changes of extracellular ion concentrations. These effects include primarily ions whose extracellular concentrations are low, such as K+ and Ca2+. Opening of K+ channels (such as delayed rectifier voltage gated channels mediating hyperpolariziung phase of action potential) or Ca2+ channels increase extracellular K+ concentration and decrease extracellular Ca2+ concentration. While these local changes are normally compensated by ionic pumps and interaction with glia, significant alternations of extracellular ion concentrations triggered by electrical activity in one cell may diffuse to neighbor neurons leading to changes of their excitability and, therefore, contribution to synchronization of electrical activity between neighbor cells. Experimental evidences suggest that indeed a variety of ions undergo activity-dependent changes in concentration 14, 16. During active states of cortical slow oscillation, due to activation of synaptic currents (primarily NMDA receptor-mediated) and Ca2+ gated intrinsic channels, the extracellular concentration of Ca2+ decreases from 1.2 mM to 1.0 mM 15, 42. This leads to a dramatic increase in synaptic failure rates (up to 80 %) 15. Therefore, chemical synaptic interactions contributing to synchronization become largely impaired. On the other hand, lower concentrations of extracellular Ca2+ promote opening of hemichannels 43 that increases effectiveness of gap junctions and therefore, electrical coupling. The most dramatic changes in extracellular concentration of K+ and Ca2+ occur during paroxysmal discharges. Extracellular K+ concentration reaches 7–18 mM during spontaneous electrographic seizures in neocortex 44–46 and the extracellular Ca2+ concentration decreases to 0.4–0.7 mM 44, 47, 48. These changes dramatically reduce synaptic excitability basically abolishing synaptic responses 49. Changes in ionic concentrations affect reversal potential of a variety of intrinsic currents that contribute to neuronal excitability, synchronization and generation of oscillatory activities, potentially promoting synchronized oscillations during epileptic seizures 17, 50. For example, hyperpolarization activated depolarizing current (Ih) in neocortical neurons is relatively weak and in normal conditions it is unlikely to play any significant role in the generation of cortical oscillations. However, during seizure activity, when extracellular K+ concentration increases, the Ih reversal potential shifts to a more depolarized value, allowing for this current to depolarize a set of cortical neurons to the firing threshold, thus leading to the generation of spikes and onset of next cycles of spike-wave discharges 51.

THALAMOCORTICAL OSCILLATIONS

Periodic oscillatory activity is a common result of neuronal synchronization and it is an emerging property of the thalamocortical system. Neuronal oscillations enable selective and dynamic control of distributed functional cell assemblies 52. The patterns and the dominant frequencies of these oscillations are defined by the specific mechanisms involved and depend on the functional state of the brain. Normal oscillatory activities include slow (0.1–15 Hz, present mainly during slow-wave sleep or anesthesia), and fast (20–600 Hz) activities. The fast activities may be present in various states of vigilance and frequently coexist with slower rhythms. Each type of oscillation is generated by a particular set of intrinsic neuronal currents, synaptic interactions and extracellular factors. Spontaneous brain rhythms during different states of vigilance may lead to increased responsiveness and plastic changes in the strength of connections among neurons, thus affecting information flow in the thalamocortical system.

The various oscillatory rhythms generated in the thalamocortical system may be divided in two main classes: intrinsic, generated by a single neuron as a result of an interplay between specific intrinsic currents [e.g. thalamic delta oscillation], and extrinsic, or network oscillations, which require the interaction of excitatory and inhibitory cells within a neuronal population (e.g. spindle oscillation, reviewed in 53). Intrinsic neuronal currents contribute to the generation of network oscillations. Oscillations may also be generated in a population of non-pacemaker neurons coupled through gap junctions.

a. Near Steady or Infra-Slow Oscillation (<0.1 Hz)

This type of oscillatory activity has a period within the range of tens of seconds to a minute 54. Prolonged metabolic disturbance, irritation of brain structures or afferent stimulation intensify these activities, and phenobarbital or diethyl ether anesthesia as well as strong electrical stimulation of cerebral cortex dramatically decrease or abolish this brain activity 54. Very little is known about the underlying mechanisms of these oscillations. Hypercapnia induced negative scalp DC shifts (~ −0.3 mV/CO2%) and hypocapnia induced positive scalp DC shifts (~0.3 mV/CO2%) 55. Infra-slow activities likely have a cortical origin given that they can be recorded from neocortical slabs 56. Indirect evidence suggests that infra-slow oscillations (0.02–0.2 Hz) synchronize faster activities, modulate cortical excitability and contribute to the aggravation of epileptic activity during sleep 57.

b. Slow oscillation

During slow-wave sleep (SWS) and some types of anesthesia the dominant activity pattern is a slow oscillation occurring with a low frequency (0.3 – 1 Hz) 1, 25, 58–60. During slow oscillation the entire cortical network alternates between silent (Hyperpolarizing, or Down) and active (Depolarizing, or Up) states, each lasting 0.2–1 s (Figure 2). Silent periods are periods of disfacilitation, i.e. periods of absence of synaptic activity, while active periods are periods of intensive synaptic activity leading to the generation of fast oscillations within the thalamocortical system 25, 59, 61–64. The survival of slow oscillations after extensive thalamic lesions in vivo 65 and in cortical in vitro preparations 66, and the absence of slow oscillation in the thalamus of decorticated cats 67 point to an intracortical origin of this rhythm.

Figure 2. Cortical sleep slow oscillation.

Figure 2

Cortical sleep slow oscillation. At EEG level the slow oscillation appears as periodic alterations of positive and negative waves (indicated by + and − signs). During EEG depth-positivity cortical neurons remain in hyperpolarized, silent state. (more...)

A particular feature of slow oscillation is that each cortical slow wave originates in a particular location and propagates to involve other cortical regions. According to Massimini et al. 68, in young adult humans (20–25 years old males) the preferential sites of origin of slow waves are in frontal cortical regions. However, in this high density EEG study, it was unclear whether it was active or silent states that propagated. Multisite intracellular experiments on cats demonstrated that it was the onset of active state that propagated, but the onset of silent states occurred more synchronously 69. Distinct from the above mentioned study 68, the active states in cats more commonly start at the border between area 5 and 7 of parietal cortex 69. A very recent study 70 demonstrates that propagation of slow activity in the human brain undergoes progressive age-dependent changes. In preschool children (2–3 years old), the propagating waves start from the occipital cortex. From early childhood to late adolescence, the location on the scalp showing maximal slow wave activity undergoes a shift from posterior to anterior regions. This suggests a progressive change in excitability of various cortical areas and that slow oscillation in the brain of adult cat is generated in areas homotypical to human adolescent brain. A propagation of activity suggests: (a) the presence of cortical regions with different levels of excitability that start the active states and (b) the presence of active synaptic chains.

At least three distinct mechanisms for the origin of slow cortical oscillations were proposed. The first depends on spontaneous miniature synaptic activities (mPSP) 71 caused by the action potential-independent release of transmitter vesicles and regulated at the level of single synapses 72, 73. Occasional summation of the miniature EPSPs during the hyperpolarized (silent) phase of slow-sleep oscillations activates the persistent sodium current and depolarizes the membrane of cortical pyramidal cells, which is sufficient for spike generation 74–76. This triggers the active phase of the slow oscillation, which is maintained by synaptic activities and the persistent sodium current. Modeling experiments suggest that short-term synaptic depression and the activation of Ca2+-dependent K+ current were sufficient to terminate the active state 75. More recent electrophysiological experiments demonstrated that the active states terminate with high levels of synchrony 69. Because the short-term synaptic plasticity and activity of intrinsic currents are very specific for each synapse or neuron, the occurrence of nearly simultaneous termination of the active states over large distances suggests the presence of active widespread (inhibitory) mechanism for the termination of active states. Indeed, about 40% of inhibitory fast-spiking interneurons increase their firing rates at the end of spontaneous active states during slow oscillation 77. These cells can contribute to the synchronous termination of active states.

Another possible mechanism accounting for the generation of active states during slow-sleep oscillation may be the generation of spontaneous activity by layer V cortical neurons 66, 78. Using a cortical slice preparation, it was shown that oscillatory activity in the frequency range of slow sleep oscillation could be generated at a concentration of extracellular K+ of 3.5 mM, which is slightly higher than in vivo (i.e., 3.2 mM) 66. This activity was usually initiated in layer V and propagated over the whole slice.

The third hypothesis attributes transitions from silent to active states to the selective synchronization of spatially structured neuronal ensembles involving a small number of cells 79. The ‘‘selective synchronization’’ hypothesis predicts that even during the silent states, some neurons of the network still generate irregular spontaneous firing.

A group of recent in vitro studies suggests that thalamocortical neurons encompass an intrinsic mechanism that may act as secondary oscillator supporting slow rhythm 80. Cortical activity can recruit, through cortico-thalamic synapses, intrinsic oscillatory mechanisms in thalamocortical neurons 81, which could then increase synchrony among cortical inhibitory interneurons. Our current studies indicate that thalamic mechanisms may be responsible for the maintenance of active states with properties observed in vivo 82, 83.

A recent in vivo study on anesthetized or sleeping cats demonstrated that spontaneous active states could originate from any cortical layer, but most often they start from deep layers and propagate toward more superficial layers 76. It was shown indeed that numerous spontaneous, spike-independent synaptic events detected during network silent states summate prior to the onset of active states in all recorded neurons. The pyramidal layer V neurons are the biggest cortical neurons 26; they possess the biggest number of synapses that contribute to the highest likelihood of these neurons to summate spontaneous synaptic events to a threshold level for action potential generation and initiate the active network states. During silent network states, the levels of extracellular Ca2+ are increased and contribute to the increase in synaptic release of neurotransmitter 15, facilitating the onset of the active network states.

Regardless of the specific mechanisms involved in sleep slow oscillation generation, long-range excitatory and inhibitory synaptic connectivity likely plays a major role in synchronization of transitions between active and silent states of slow oscillation leading to periodic LFP and EEG oscillations observed in animals and humans during deep sleep.

c. Delta oscillation

Field potential recordings from neocortex in human and animal models during sleep reveal the presence of delta oscillations (1–4 Hz). The delta oscillation likely has two different components, one of which originates in the neocortex and the other in the thalamus. Neocortical delta activity was significantly enhanced after surgical removal of thalamus 84 and chronically isolated neocortical slabs 85. Little is known about the cellular mechanisms mediating cortical delta oscillation. One of the hypotheses suggests that cortical delta activity could be driven by the discharge of intrinsically bursting neurons 86. On the other hand, thalamic delta rhythm is a well known example of rhythmic activity generated intrinsically by thalamic relay neurons as a result of an interplay between their low-threshold Ca2+ current (IT) and hyperpolarization activated cation current (Ih) 86b. As such, the delta oscillation may be observed during deep sleep when thalamic relay neurons are hyperpolarized sufficiently to deinactivate IT 87–90. It was also shown that at a certain level of leak current (Ileak), the ‘window’ component of IT in thalamocortical neurons, may create oscillations similar in frequency to the intrinsic thalamic delta oscillation 91. While specific intrinsic mechanisms are responsible for the generation of oscillatory activity in individual thalamic neurons, these oscillations would not be synchronized and would not lead to the observed LFP and EEG rhythms unless spiking/bursting activity of individual neurons become synchronized by chemical and electrical synapses as described earlier in this chapter.

Slow wave sleep may be essential for memory consolidation and memory formation 92–95. It has been proposed that synaptic plasticity associated with slow and delta oscillations could contribute to the consolidation of memory traces acquired during wakefulness 96.

d. Sleep spindle oscillations

Sleep spindle oscillations consist of waxing-and-waning field potentials at 7–14 Hz, which last 1–3 seconds and recur every 5–15 seconds. In vivo, spindle oscillations are typically observed during the early stages of sleep or during active phases of sleep slow oscillation. In vivo, in vitro and in silico studies suggest that the minimal substrate accounting for spindle oscillations consists in the interaction between thalamic reticular and relay cells (Figure 3) 31, 97–100. Burst firing of reticular thalamic neurons induces inhibitory postsynaptic potentials in thalamocortical neurons. This deinactivates IT, inducing rebound burst firing in thalamocortical neurons, which, in turn, excite reticular thalamic neurons allowing the cycle to start again. During the early phase of spindles, the reticular nucleus single-handedly drives the spindle oscillation via intrinsic mechanisms 100–103. The second component of spindles, on the other hand, primarily develops as a result of interactions between reticular and relay neurons 104, 105. Additionally, cortical firing contributes to spindle synchronization through cortico-thalamic feedback connections, thereby imposing simultaneous excitation of reticular and relay neurons 106, 107. The waning phase of spindles occurs as a result of network desychronization 103, 107a and of Ca2+-induced cAMP-dependent up-regulation of the hyperpolarization activated cation current, Ih, in relay cells 108–110. It is important to mention that reticular thalamic neurons are interconnected via gap junctions, which aid in synchronizing spindles 111, 112.

Figure 3. Cellular basis of spindle activity.

Figure 3

Cellular basis of spindle activity. A, In vivo recordings. Three phases of a spindle sequence. Dual intracellular recording of cortical (area 4) and thalamocortical (TC) neurons from ventro-lateral (VL) nucleus of the thalamus. B, Computational model. (more...)

While spindle oscillations propagate in thalamic slices 113, thalamocortical synaptic interactions promote large-scale synchrony of spindle oscillations in vivo 114. Human sleep spindles are highly synchronous across the scalp when measured by EEG, but not when measured simultaneously with magnitoencephalogram (MEG) 115. MEG signals show low correlation and low coherence with each other or with EEG signals. Principle Components Analysis shows that the MEG field pattern is more complex than the EEG pattern implying that MEG signals are produced from multiple partially independent cortical generators, whereas EEG may instead be dominated by a weak, but widespread spindle generator. It was proposed that the differential activity patterns in the core (thalamocortical neurons forming specific, focused projections to cortical layers IV and VI) and matrix (thalamocortical neurons forming diffuse projections to layer I) subsystems may explain discrepancies between the temporal patterns of spindles simultaneously observed in EEG and MEG recordings 116.

Recent studies show that sleep related spindle oscillations are essential for memory formation 92 and demonstrate short- and middle term synaptic plasticity (reviewed in 96). Spindling may activate the protein kinase A molecular “gate”, thus opening the door for gene expression 117 and allowing long-term changes to take place following subsequent inputs.

e. Beta-gamma oscillation

The waking state of the brain is characterized by the predominance of frequencies in the beta (15–30 Hz) and gamma (30–80 Hz) ranges 118, 119. These fast rhythms are also synchronized between neighboring cortical sites during some forms of anesthesia, natural slow-wave sleep, and REM sleep 62, 120,120a when consciousness is suspended and thoughts are either absent or bizarre. During slow-wave sleep the fast rhythms follow the onset of depth-negative EEG wave (Figure 4). In vitro experiments and computational models demonstrated that gamma activity generation depends on the activity of inhibitory interneurons 121–126. Gamma activity can exist in transient and persistent forms. Experimentally, transient (hundreds of milliseconds) gamma oscillations can be induced by tetanic stimulation of the hippocampus 124, 127. Persistent gamma activity is found in CA3 128 and neocortex 129; this form of gamma oscillation can be induced by bath-application of carbachol or kainate and it can last for minutes to hours. During persistent gamma activity, the interneurons fire on every cycle or every two cycles and pyramidal cells fire at much lower frequencies suggesting that active inhibition is involved in the generation of this particular rhythm. Finally, it was found that GABAergic interaction in isolated interneuron networks may lead to network oscillation in the gamma frequency range 130, 131. In both computational models and in vitro experiments, it was shown that the frequency of these oscillations depends on the conductance and decay time of GABAA currents 130. Large-scale network simulations revealed that coherent gamma range oscillations may appear through occasional increases in spiking synchrony within local groups of cortical neurons 132. It was shown that even local synaptic connectivity may explain long-range synchrony of gamma oscillations 133.

Figure 4. Gamma oscillation is an important component of slow oscillation of sleep.

Figure 4

Gamma oscillation is an important component of slow oscillation of sleep. Depth- EEG recoding from area 5. Slow oscillation, spindles and gamma activities are indicated. Below, Fast Fourier Transformation of the recoding shown above. (Modified from ). (more...)

During waking, gamma activity is associated with attentiveness 134, 135, focused arousal 136, sensory perception 137, movement 138, 139 and prediction 140. It has been proposed that synchronization in the gamma frequency range is related to cognitive processing and is important for temporal binding of sensory stimuli 18, 141, 142. Gamma activity with low levels of coherence is present within active phases of slow-wave sleep 62, 64, 143, 144. However, the role of this activity is unknown. During REM sleep, highly coherent gamma activity 143 may contribute to the generation of dreams.

f. Ripples

Fast oscillations (>100 Hz), termed ripples, were initially described in CA1 hippocampal area and perirhinal cortex, where they were associated with bursts of sharp potentials during anesthesia, behavioral immobility, and natural sleep 145–149. In the neocortex, fast oscillations (>200 Hz, up to 600 Hz) have been found in sensory-evoked potentials in the rat barrel cortex 150, 151, during high-voltage spike-and-wave patterns 152. The neocortical network seems to be sufficient for the generation of ripples, as demonstrated using isolated cortical slab preparations 153. In addition to active inhibition 149, 153, the electrical coupling mediated by gap junctions contributes to ripple synchronization 154–156. The electrical coupling may occur between axons of principal cells 157 as well as via a network of inhibitory interneurons 10, 34, 35, 158, 159. Since ripples are also recorded in glial cells, the electrical coupling between gial cells could also play a role in the synchronization of ripples 155. The field potentials increase neuronal excitability, and by a positive feedback loop they could also be involved in the generation of neocortical ripples 160. In non-anesthetized brain the highest amplitude of ripples in neocortex occurs during active phases of slow-wave sleep, it is lower during both REM sleep and waking state, and it reaches lowest values during silent phases of slow-wave sleep 153.

Neocortical ripples are generated during large amplitude spontaneous or evoked field potential deflections. These ample changes in the field potential are associated with synchronous activity of many neurons. This suggests that ripples may “alarm” the brain network about the presence of a large firing neuronal constellation. The danger of such a focal synchronous excitation of a neuronal pool is that it may overcome a certain threshold of excitability, leading to the onset of seizures 155, 160.

Short- and Long-Range Synchrony during Experimental Epilepsy

Cortically generated electrographic seizures arise without discontinuity from sleep oscillations 161. These seizures are accompanied by the generation of large amplitude field potentials, suggesting the presence of enhanced local synchrony 161 and are characterized by spike-wave (SW) or spike-wave/polyspike-wave (SW/PSW) complexes at 1–2.5 Hz, intermingled with episodes of fast runs at ~7–16 Hz 162–165. The evolvement of these seizures from the cortically generated slow oscillations may be shaped by the thalamus 81, 163, 166–169. The electrographic pattern of SW/PSW seizures as well as their occurrence during slow-wave sleep resembles the seizures accompanying the Lennox-Gastaut syndrome in humans 170–173.

These seizures could be classified as primarily focal and secondarily generalized (see Figure 3 in 174); visual inspection suggests that on most occasions they start and stop almost simultaneously in all recorded areas. Precise measurements of synchrony during SW/PSW seizures suggest that long-range synchrony recorded on a wave-by-wave basis is rather loose 161, 168, 174, 175. The coefficient of correlation of field potentials recorded during SW complexes fluctuates between 0.3 and 0.8 176. However, the field potentials recorded during these seizures are of large amplitude, often exceeding the amplitude of slow waves recorded during natural slow-wave sleep 176 or during anesthesia-induced slow oscillation 161, 163, suggesting existence of high local synchrony. Within the fast runs, the patterns of synchronization recorded in different electrodes are as follows: (i) synchronous, in phase, (ii) synchronous, with phase shift, (iii) patchy, repeated in phase/phase shift transitions and (iv) non-synchronous, with slightly different frequencies in different recording sites or absence of oscillatory activity in one of the recording sites; the synchronous patterns (in phase or with phase shifts) were most common 174. All these patterns could be recorded in the same pair of electrodes during different seizures or even during different periods of fast runs within the same seizure. An example of such recording is shown in the figure 5. Here, during the first period of fast runs, which started and terminated almost simultaneously at both the intracellular and the EEG levels, the field potential and the neuron recorded 2 mm apart oscillated with different frequencies, and thus there was no synchronization. During the next period of fast run, both the neuron and the field potential signal oscillated in perfect synchrony, and the neuron fired constantly during the descending phase of the field potential deflection.

Figure 5. Variability in neuron – field synchronization during fast runs.

Figure 5

Variability in neuron – field synchronization during fast runs. A. Depth-EEG and simultaneous intracellular recordings during an electrographic seizure. The distance between field potential electrode and the intracellularly recorded neuron was (more...)

Why is the local synchrony during seizures enhanced as indicated by large amplitude field potentials, and why is the long-range synchrony impaired, as shown by a variety of synchrony patterns (Figure 5)? We propose the following scenario that explains both enhanced local and decreased long-range synchrony. Seizures are associated with activation of a large variety of neuronal currents 161, 177. During seizures, the extracellular levels of K+ increase, and the extracellular levels of Ca2+ decrease. While diffusion potentially reduces ion gradients in the extracellular space, a limited diffusion rate, further reduced by cell swelling during seizure, allows formation of local areas of “abnormal” ion concentrations. These changes are commonly found in different structures, including neocortex and hippocampus, and in both in vivo and in vitro recordings 44, 47, 178, 179. The changes in extracellular concentration of K+ and Ca2+ occur simultaneously 47, 180, 181. Low levels of extracellular levels of Ca2+ significantly dampen synaptic transmission 8, 182. Thus, chemical synaptic transmission cannot play a leading role in synaptic interactions during seizures. In addition, our recent data show that in high (12 mM) extracellular K+ conditions, the action potential generation is impaired and action potentials fail to propagate via axons (Figure 6), probably due to a depolarizing block and decreased input resistance 49. Therefore, action potentials do not reach target structures. All these factors lead to an impairment of long-range synchronization during seizures because the neurotransmitter mediating these interactions is not released. On the other hand, low Ca2+ conditions could favor local synchronization via gap junctions. Indeed, when extracellular levels of Ca2+ drops to 1 mM and lower, as it is observed during seizures, the hemichannels that form gap junctions remain open 43, while they are half closed at physiological Ca2+ conditions (1.2 mM). This strengthens electrical coupling between glial cells 183 and cortical interneurons 34, 35 that mediate local synchrony. Furthermore, elevated extracellular levels of K+ enhances cell excitability and promotes bursting activity 17, 50, 184 that is transmitted better through gap junctions than single action potentials. Potentially, electrical coupling between axons of pyramidal cells 37 could also contribute to local synchronization, although the major type of activities recorded in axons are action potentials, which could not be efficiently transferred via gap junctions, because they constitute a low pass filter. Accordingly, the use of halothane, a gap junction blocker, effectively blocks seizure generation 155.

Figure 6. High K+ concentration impairs generation of action potentials in neocortical neurons.

Figure 6

High K+ concentration impairs generation of action potentials in neocortical neurons. A. Responses of a cortical neuron to depolarizing current pulses in control (2.8 mM) and ‘paroxysmal’ (12 mM) K+ conditions. When 12 mM K+ was used the (more...)

Hence, the decrease in extracellular levels of Ca2+ and the increase in extracellular levels of K+ to those concentrations that occur during seizures, significantly impair long-range neuronal synchronization, due to a decrease in neuronal firing ability and a diminution of transmitter release. However, the same changes may increase local synchrony, mediated via gap junctions and possibly ephaptic interactions.

CONCLUSIONS

We propose that neocortical synchronization can be achieved via intracortical and thalamocortical mechanisms. Four leading mechanisms of synchronization play a distinct and important role: (i) chemical synaptic transmission, (ii) electrical coupling, (iii) ephaptic interactions and (iv) activity-dependent changes in ionic concentrations. Major types of cortical rhythmic activities (slow oscillation, delta, spindles, beta-gamma, ripples, paroxysmal spike-waves and fast runs) are generated using these mechanisms. It is difficult, however, to identify the exact contribution of each mechanism to a specific type of oscillation. It is clear that long-range synchronization is achieved using chemical synaptic interactions. Due to the low-pass filtering properties of gap junctions, electrical coupling should play a smaller role in synchronization of fast rhythmic activities as compared to slower activities. Nevertheless gamma rhythm depends on activity of interneurons, which are known to be interconnected via gap junctions; therefore, gap junctions should play a critical role in the generation of these oscillations. Weak electrical fields in the EEG range synchronize effectively cortical activities at low, but not at high frequencies. Finally, activity-dependent ionic changes may influence local synchronization by modulating neuronal excitability; however, since ionic changes also affect synaptic transmitter release, this could result in impairment of long-rage synchrony.

Acknowledgments

Authors are supported by Canadian Institutes of Health Research (MOP-37862, MOP-67175), National science and engineering research council of Canada (grant 298475) and National Institute of Neurological Disorders and Stroke (1R01NS060870, 1R01NS059740). I.T. is Fonds de la Recherche en Santé du Québec Research Scholar.

References

1.
Steriade M, McCormick DA, Sejnowski TJ. Thalamocortical oscillations in the sleeping and aroused brain. Science. 1993;262:679–685. [PubMed: 8235588]
2.
Bazhenov M, Timofeev I. Scholarpedia. Vol. 1. 2006. Thalamocortical oscillations; p. 1319.
3.
Niedermeyer E, Lopes da Silva F. Electroencephalography: Basic principles, clinical applications and related fields. 5th ed. Philadelphia: Lippincott Williams & Wilkins; 2005. p. 1309.
4.
Katzner S, Nauhaus I, Benucci A, Bonin V, Ringach DL, Carandini M. Local origin of field potentials in visual cortex. Neuron. 2009;61:35–41. [PMC free article: PMC2730490] [PubMed: 19146811]
5.
Bedard C, Kroger H, Destexhe A. Modeling extracellular field potentials and the frequency-filtering properties of extracellular space. Biophys J. 2004;86:1829–42. [PMC free article: PMC1304017] [PubMed: 14990509]
6.
Lopes da Silva F, Van Rotterdam A. Biophysical aspects of EEG and magnetoencephalogram generation, Electroencephalography: Basic principles, clinical applications, and related fields. Niedermeyer E, Lopes da Silva F, editors. Lippincott Williams & Wilkins; Philadelphia: 2005. pp. 107–125.
7.
Eccles JC. The physiology of synapses. Berlin: Springer; 1964. p. 316.
8.
Katz B, Miledi R. The role of calcium in neuromuscular facilitation. J Physiol. 1968;195:481–92. [PMC free article: PMC1351674] [PubMed: 4296699]
9.
Bennett MV. Gap junctions as electrical synapses. J Neurocytol. 1997;26:349–66. [PubMed: 9278865]
10.
Galarreta M, Hestrin S. Electrical synapses between GABA-releasing interneurons. Nat Rev Neurosci. 2001;2:425–33. [PubMed: 11389476]
11.
Taylor CP, Dudek FE. Synchronous neural afterdischarges in rat hippocampal slices without active chemical synapses. Science. 1982;218:810–2. [PubMed: 7134978]
12.
Frohlich F, McCormick DA. Endogenous electric fields may guide neocortical network activity. Neuron. 2010;67:129–143. [PMC free article: PMC3139922] [PubMed: 20624597]
13.
Jefferys JG. Nonsynaptic modulation of neuronal activity in the brain: Electric currents and extracellular ions. Physiol Rev. 1995;75:689–723. [PubMed: 7480159]
14.
Somjen GG. Ion regulation in the brain: Implications for pathophysiology. Neuroscientist. 2002;8:254–67. [PubMed: 12061505]
15.
Crochet S, Chauvette S, Boucetta S, Timofeev I. Modulation of synaptic transmission in neocortex by network activities. Eur J Neurosci. 2005;21:1030–1044. [PubMed: 15787708]
16.
Somjen GC. Ions in the brain: Normal function, seizures, and stroke. Oxford University Press; USA: 2004. p. 470.
17.
Frohlich F, Bazhenov M, Iragui-Madoz V, Sejnowski TJ. Potassium dynamics in the epileptic cortex: New insights on an old topic. Neuroscientist. 2008;14:422–33. [PMC free article: PMC2854295] [PubMed: 18997121]
18.
Singer W, Gray CM. Visual feature integration and the temporal correlation hypothesis. Annu Rev Neurosci. 1995;18:555–86. [PubMed: 7605074]
19.
Thomson AM, West DC, Deuchars J. Properties of single axon excitatory postsynaptic potentials elicited in spiny interneurons by action potentials in pyramidal neurons in slices of rat neocortex. Neuroscience. 1995;69:727–738. [PubMed: 8596643]
20.
Stratford KJ, Tarczy-Hornoch K, Martin KA, Bannister NJ, Jack JJ. Excitatory synaptic inputs to spiny stellate cells in cat visual cortex. Nature. 1996;382:258–61. [PubMed: 8717041]
21.
Markram H, Lubke J, Frotscher M, Roth A, Sakmann B. Physiology and anatomy of synaptic connections between thick tufted pyramidal neurones in the developing rat neocortex. J Physiol. 1997;500:409–440. [PMC free article: PMC1159394] [PubMed: 9147328]
22.
Buhl EH, Tamãs G, Szilãgyi T, Stricker C, Paulsen O, Somogyi P. Effect, number and location of synapses made by single pyramidal cells onto aspiny interneurones of cat visual cortex. J Physiol. 1997;500:689–713. [PMC free article: PMC1159419] [PubMed: 9161986]
23.
Krimer LS, Goldman-Rakic PS. Prefrontal microcircuits: Membrane properties and excitatory input of local, medium, and wide arbor interneurons. J Neurosci. 2001;21:3788–96. [PMC free article: PMC6762691] [PubMed: 11356867]
24.
Vyazovskiy VV, Olcese U, Lazimy YM, Faraguna U, Esser SK, Williams JC, Cirelli C, Tononi G. Cortical firing and sleep homeostasis. Neuron. 2009;63:865–878. [PMC free article: PMC2819325] [PubMed: 19778514]
25.
Steriade M, Timofeev I, Grenier F. Natural waking and sleep states: A view from inside neocortical neurons. J Neurophysiol. 2001;85:1969–85. [PubMed: 11353014]
26.
DeFelipe J, Farinas I. The pyramidal neuron of the cerebral cortex: Morphological and chemical characteristics of the synaptic inputs. Prog Neurobiol. 1992;39:563–607. [PubMed: 1410442]
27.
Gabbott PL, Somogyi J, Stewart MG, Hamori J. GABA-immunoreactive neurons in the rat cerebellum: A light and electron microscope study. J Comp Neurol. 1986;251:474–90. [PubMed: 3537020]
28.
Bazhenov M, Stopfer M. Forward and back: Motifs of inhibition in olfactory processing. Neuron. 2010;67:357–8. [PMC free article: PMC5337941] [PubMed: 20696373]
29.
Whittington MA, Traub RD, Jefferys JG. Synchronized oscillations in interneuron networks driven by metabotropic glutamate receptor activation. Nature. 1995;373:612–5. [PubMed: 7854418]
30.
Joho RH, Ho CS, Marks GA. Increased gamma- and decreased delta-oscillations in a mouse deficient for a potassium channel expressed in fast-spiking interneurons. J Neurophysiol. 1999;82:1855–64. [PubMed: 10515974]
31.
von Krosigk M, Bal T, McCormick DA. Cellular mechanisms of a synchronized oscillation in the thalamus. Science. 1993;261:361–4. [PubMed: 8392750]
32.
Dermietzel R, Spray DC. Gap junctions in the brain: Where, what type, how many and why. Trends Neurosci. 1993;16:186–92. [PubMed: 7685944]
33.
Perez Velazquez JL, Carlen PL. Gap junctions, synchrony and seizures. Trends Neurosci. 2000;23:68–74. [PubMed: 10652547]
34.
Galarreta M, Hestrin S. A network of fast-spiking cells in the neocortex connected by electrical synapses. Nature. 1999;402:72–5. [PubMed: 10573418]
35.
Gibson JR, Beierlein M, Connors BW. Two networks of electrically coupled inhibitory neurons in neocortex. Nature. 1999;402:75–9. [PubMed: 10573419]
36.
Mugnaini E. Cell junctions of astrocytes, ependyma, and related cells in the mammalian central nervous system, with emphasis on the hypothesis of a generalized functional syncytium of supporting cells, Astrocytes. Fedoroff S, Vernadakis A, editors. New York: Academic: New York; 1986. pp. 329–371.
37.
Schmitz D, Schuchmann S, Fisahn A, Draguhn A, Buhl EH, Petrasch-Parwez E, Dermietzel R, Heinemann U, Traub RD. Axo-axonal coupling a novel mechanism for ultrafast neuronal communication. Neuron. 2001;31:831–40. [PubMed: 11567620]
38.
Vigmond EJ, Perez Velazquez JL, Valiante TA, Bardakjian BL, Carlen PL. Mechanisms of electrical coupling between pyramidal cells. J Neurophysiol. 1997;78:3107–16. [PubMed: 9405530]
39.
Yu Y, Shu Y, McCormick DA. Cortical action potential backpropagation explains spike threshold variability and rapid-onset kinetics. J Neurosci. 2008;28:7260–7272. [PMC free article: PMC2664555] [PubMed: 18632930]
40.
Naundorf B, Wolf F, Volgushev M. Unique features of action potential initiation in cortical neurons. Nature. 2006;440:1060–3. [PubMed: 16625198]
41.
Marshall L, Helgadottir H, Molle M, Born J. Boosting slow oscillations during sleep potentiates memory. Nature. 2006;444:610–3. [PubMed: 17086200]
42.
Massimini M, Amzica F. Extracellular calcium fluctuations and intracellular potentials in the cortex during the slow sleep oscillation. J Neurophysiol. 2001;85:1346–50. [PubMed: 11248006]
43.
Thimm J, Mechler A, Lin H, Rhee S, Lal R. Calcium-dependent open/closed conformations and interfacial energy maps of reconstituted hemichannels. J Biol Chem. 2005;280:10646–10654. [PubMed: 15615707]
44.
Amzica F, Massimini M, Manfridi A. Spatial buffering during slow and paroxysmal sleep oscillations in cortical networks of glial cells in vivo. J Neurosci. 2002;22:1042–1053. [PMC free article: PMC6758489] [PubMed: 11826133]
45.
Moody WJ, Futamachi KJ, Prince DA. Extracellular potassium activity during epileptogenesis. Exp Neurol. 1974;42:248–63. [PubMed: 4824976]
46.
Prince DA, Lux HD, Neher E. Measurement of extracellular potassium activity in cat cortex. Brain Res. 1973;50:489–95. [PubMed: 4705519]
47.
Heinemann U, Lux HD, Gutnick MJ. Extracellular free calcium and potassium during paroxysmal activity in the cerebral cortex of the cat. Exp Brain Res. 1977;27:237–43. [PubMed: 880984]
48.
Pumain R, Kurcewicz I, Louvel J. Fast extracellular calcium transients: Involvement in epileptic processes. Science. 1983;222:177–9. [PubMed: 6623068]
49.
Seigneur J, Timofeev I. Synaptic impairment induced by paroxysmal ionic conditions in neocortex. Epilepsia. 2011;52(1):132–139. [PMC free article: PMC3151430] [PubMed: 21126243]
50.
Bazhenov M, Timofeev I, Frohlich F, Sejnowski TJ. Cellular and network mechanisms of electrographic seizures. Drug Discov Today Dis Models. 2008;5:45–57. [PMC free article: PMC2633479] [PubMed: 19190736]
51.
Timofeev I, Bazhenov M, Sejnowski T, Steriade M. Cortical hyperpolarization-activated depolarizing current takes part in the generation of focal paroxysmal activities. Proc Natl Acad Sci U S A. 2002;99:9533–7. [PMC free article: PMC123175] [PubMed: 12089324]
52.
Canolty RT, Ganguly K, Kennerley SW, Cadieu CF, Koepsell K, Wallis JD, Carmena JM. Oscillatory phase coupling coordinates anatomically dispersed functional cell assemblies. Proc Natl Acad Sci USA. 2010;107(40):17356–17361. [PMC free article: PMC2951408] [PubMed: 20855620]
53.
Timofeev I, Bazhenov M. Mechanisms and biological role of thalamocortical oscillations, Trends in chronobiology research. Columbus F, editor. Nova Science Publishers Inc; 2005. pp. 1–47.
54.
Aladjalova NA. Infra-slow rhythmic oscillations of the steady potential of the cerebral cortex. Nature. 1957;4567:957–959. [PubMed: 13430746]
55.
Nita DA, Vanhatalo S, Lafortune FD, Voipio J, Kaila K, Amzica F. Nonneuronal origin of CO2-related DC EEG shifts: An in vivo study in the cat. J Neurophysiol. 2004;92:1011–22. [PubMed: 15056689]
56.
Aladjalova NA. Slow electrical processes in the brain. Moscow: Acad Sci USSA; 1962.
57.
Vanhatalo S, Palva JM, Holmes MD, Miller JW, Voipio J, Kaila K. Infraslow oscillations modulate excitability and interictal epileptic activity in the human cortex during sleep. Proc Natl Acad Sci U S A. 2004;101:5053–7. [PMC free article: PMC387372] [PubMed: 15044698]
58.
Steriade M, Nunez A, Amzica F. A novel slow (<1 hz) oscillation of neocortical neurons in vivo : Depolarizing and hyperpolarizing components. J Neurosci. 1993;13:3252–3265. [PMC free article: PMC6576541] [PubMed: 8340806]
59.
Timofeev I, Grenier F, Steriade M. Disfacilitation and active inhibition in the neocortex during the natural sleep-wake cycle: An intracellular study. Proc Natl Acad Sci U S A. 2001;98:1924–1929. [PMC free article: PMC29358] [PubMed: 11172052]
60.
Achermann P, Borbely AA. Low-frequency (< 1 hz) oscillations in the human sleep electroencephalogram. Neuroscience. 1997;81:213–22. [PubMed: 9300413]
61.
Contreras D, Timofeev I, Steriade M. Mechanisms of long-lasting hyperpolarizations underlying slow sleep oscillations in cat corticothalamic networks. J Physiol. 1996;494:251–64. [PMC free article: PMC1160627] [PubMed: 8814619]
62.
Steriade M, Contreras D, Amzica F, Timofeev I. Synchronization of fast (30–40 hz) spontaneous oscillations in intrathalamic and thalamocortical networks. J Neurosci. 1996;16:2788–808. [PMC free article: PMC6578775] [PubMed: 8786454]
63.
Timofeev I, Contreras D, Steriade M. Synaptic responsiveness of cortical and thalamic neurones during various phases of slow sleep oscillation in cat. J Physiol. 1996;494:265–78. [PMC free article: PMC1160628] [PubMed: 8814620]
64.
Mukovski M, Chauvette S, Timofeev I, Volgushev M. Detection of active and silent states in neocortical neurons from the field potential signal during slow-wave sleep. Cereb Cortex. 2007;17:400–414. [PubMed: 16547348]
65.
Steriade M, Nunez A, Amzica F. Intracellular analysis of relations between the slow (<1 hz) neocortical oscillations and other sleep rhythms of electroencephalogram. J Neurosci. 1993;13:3266–3283. [PMC free article: PMC6576520] [PubMed: 8340807]
66.
Sanchez-Vives MV, McCormick DA. Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nat Neurosci. 2000;3:1027–34. [PubMed: 11017176]
67.
Timofeev I, Steriade M. Low-frequency rhythms in the thalamus of intact-cortex and decorticated cats. J Neurophysiol. 1996;76:4152–68. [PubMed: 8985908]
68.
Massimini M, Huber R, Ferrarelli F, Hill S, Tononi G. The sleep slow oscillation as a traveling wave. J Neurosci. 2004;24:6862–6870. [PMC free article: PMC6729597] [PubMed: 15295020]
69.
Volgushev M, Chauvette S, Mukovski M, Timofeev I. Precise long-range synchronization of activity and silence in neocortical neurons during slow-wave sleep. J Neurosci. 2006;26:5665–5672. [PMC free article: PMC6675259] [PubMed: 16723523]
70.
Kurth S, Ringli M, Geiger A, LeBourgeois M, Jenni OG, Huber R. Mapping of cortical activity in the first two decades of life: A high-density sleep electroencephalogram study. J Neurosci. 2010;30:13211–9. [PMC free article: PMC3010358] [PubMed: 20926647]
71.
Fatt P, Katz B. Spontaneous sub-threshold activity at motor-nerve endings. J Physiol. 1952;117:109–128. [PMC free article: PMC1392564] [PubMed: 14946732]
72.
Paré D, Lebel E, Lang EJ. Differential impact of miniature synaptic potentials on the soma and dendrites of pyramidal neurons in vivo, J Neurophysiol. Vol. 78. 1997. pp. 1735–9. [PubMed: 9310459]
73.
Salin PA, Prince DA. Spontaneous gabaa receptor-mediated inhibitory currents in adult rat somatosensory cortex. J Neurophysiol. 1996;75:1573–88. [PubMed: 8727397]
74.
Timofeev I, Grenier F, Bazhenov M, Sejnowski TJ, Steriade M. Origin of slow cortical oscillations in deafferented cortical slabs. Cereb Cortex. 2000;10:1185–1199. [PubMed: 11073868]
75.
Bazhenov M, Timofeev I, Steriade M, Sejnowski TJ. Model of thalamocortical slow-wave sleep oscillations and transitions to activated states. J Neurosci. 2002;22:8691–704. [PMC free article: PMC6757797] [PubMed: 12351744]
76.
Chauvette S, Volgushev M, Timofeev I. Origin of active states in local neocortical networks during slow sleep oscillation. Cereb Cortex. 2010;20:2660–2674. [PMC free article: PMC2951844] [PubMed: 20200108]
77.
Puig MV, Ushimaru M, Kawaguchi Y. Two distinct activity patterns of fast-spiking interneurons during neocortical up states. Proc Natl Acad Sci USA. 2008;105(24):8428–8433. [PMC free article: PMC2448853] [PubMed: 18550841]
78.
Compte A, Sanchez-Vives MV, McCormick DA, Wang XJ. Cellular and network mechanisms of slow oscillatory activity (<1 hz) and wave propagations in a cortical network model. J Neurophysiol. 2003;89:2707–25. [PubMed: 12612051]
79.
Cossart R, Aronov D, Yuste R. Attractor dynamics of network up states in the neocortex. Nature. 2003;423:283–8. [PubMed: 12748641]
80.
Crunelli V, Hughes SW. The slow (<1 hz) rhythm of non-rem sleep: A dialogue between three cardinal oscillators. Nat Neurosci. 2010;13:9–17. [PMC free article: PMC2980822] [PubMed: 19966841]
81.
Hughes SW, Cope DW, Blethyn KL, Crunelli V. Cellular mechanisms of the slow (<1 hz) oscillation in thalamocortical neurons in vitro. Neuron. 2002;33:947–58. [PubMed: 11906700]
82.
Lemieux M, Timofeev I. Impact of thalamic inactivation on slow oscillation in the suprasylvian gyrus of cat in Society for Neuroscience 40th Annual Meeting; San-Diego: Society for Neuroscience; 2010.
83.
Lonjers P, Timofeev I, Chen J-Y, Bazhenov M. The roles of thalamocortical input in generating slow sleep oscillations in Society for Neuroscience 40th Annual Meeting; San-Diego: Society for Neuroscience; 2010.
84.
Villablanca J. Role of the thalamus in sleep controle:Sleep-wakefulness studies in chronic diencephalic and athalamic cats, Basic sleep mechanisms. Petre-Quadens O, Schlag J, editors. Academic; New York: 1974. pp. 51–81.
85.
Grafstein B, Sastry PB. Some preliminary electrophysiological studies on chronic neuronally isolated cerebral cortex. Electroencephalogr Clin Neurophysiol. 1957;9:723–5. [PubMed: 13480247]
86.
Amzica F, Steriade M. Electrophysiological correlates of sleep delta waves. Electroencephalogr Clin Neurophysiol. 1998;107:69–83. [PubMed: 9751278]
86b.
McCormick DA, Pape HC. Properties of a hyperpolarization-activated cation current and its role in rhythmic oscillation in thalamic relay neurons. J Physiol. 1990;431:291–318. [PMC free article: PMC1181775] [PubMed: 1712843]
87.
Curró Dossi R, Nunez A, Steriade M. Electrophysiology of slow (0.5–4 hz) intrinsic oscillation of cat thalamocortical neurones in vivo. J Physiol. 1992;447:215–234. [PMC free article: PMC1176033] [PubMed: 1593448]
88.
Leresche N, Lightowler S, Soltesz I, Jassik-Gerschenfeld D, Crunelli V. Low-frequency oscillatory activities intrinsic to rat and cat thalamocortical cells. J Physiol. 1991;441:155–174. [PMC free article: PMC1180191] [PubMed: 1840071]
89.
McCormick DA, Pape HC. Noradrenergic and serotonergic modulation of a hyperpolarization-activated cation current in thalamic relay neurones. J Physiol. 1990;431:319–342. [PMC free article: PMC1181776] [PubMed: 1712844]
90.
Soltesz I, Lightowler S, Leresche N, Jassik-Gerschenfeld D, Pollard CE, Crunelli V. Two inward currents and the transformation of low-frequency oscillations of rat and cat thalamocortical cells. J Physiol. 1991;441:175–197. [PMC free article: PMC1180192] [PubMed: 1667794]
91.
Williams SR, Tóth TI, Turner JP, Hughes SW, Crunelli W. The ‘window’ component of the low threshold ca2+ current produces input signal amplification and bistability in cat and rat thalamocortical neurones. J Physiol. 1997;505:689–705. [PMC free article: PMC1160046] [PubMed: 9457646]
92.
Gais S, Plihal W, Wagner U, Born J. Early sleep triggers memory for early visual discrimination skills. Nat Neurosci. 2000;3:1335–1339. [PubMed: 11100156]
93.
Huber R, Ghilardi MF, Massimini M, Tononi G. Local sleep and learning. Nature. 2004;430:78–81. [PubMed: 15184907]
94.
Stickgold R, James L, Hobson JA. Visual discrimination learning requires sleep after training. Nat Neurosci. 2000;3:1237–8. [PubMed: 11100141]
95.
Maquet P. The role of sleep in learning and memory. Science. 2001;294:1048–52. [PubMed: 11691982]
96.
Steriade M, Timofeev I. Neuronal plasticity in thalamocortical networks during sleep and waking oscillations. Neuron. 2003;37:563–76. [PubMed: 12597855]
97.
Steriade M, Deschenes M. The thalamus as a neuronal oscillator. Brain Res Rev. 1984;8:1–63. [PubMed: 6440659]
98.
Steriade M, Llinas R. The functional states of the thalamus and the associated neuronal interplay. Physiol Reviews. 1988;68:649–742. [PubMed: 2839857]
99.
Steriade M, Jones EG, Llinas R. Thalmic oscillations and signaling, The neuroscience institute publications. New York: John Wiley & Sons; 1990. p. 431.
100.
Steriade M, Deschenes M, Domich L, Mulle C. Abolition of spindle oscillations in thalamic neurons disconnected from nucleus reticularis thalami. J Neurophysiol. 1985;54:1473–97. [PubMed: 4087044]
101.
Destexhe A, Contreras D, Sejnowski TJ, Steriade M. A model of spindle rhythmicity in the isolated thalamic reticular nucleus. J Neurophysiol. 1994;72:803–818. [PubMed: 7527077]
102.
Bazhenov M, Timofeev I, Steriade M, Sejnowski TJ. Self-sustained rhythmic activity in the thalamic reticular nucleus mediated by depolarizing GABAa receptor potentials. Nat Neurosci. 1999;2:168–74. [PubMed: 10195202]
103.
Timofeev I, Bazhenov M, Sejnowski T, Steriade M. Contribution of intrinsic and synaptic factors in the desynchronization of thalamic oscillatory activity. Thalamus & Related Systems. 2001;1:53–69.
104.
Destexhe A, Bal T, McCormick DA, Sejnowski TJ. Ionic mechanisms underlaying synchronized and propagating waves in a model of ferret thalamic slices. J Neurophysiol. 1996;76:2049–2070. [PubMed: 8890314]
105.
Bazhenov M, Timofeev I, Steriade M, Sejnowski T. Spiking-bursting activity in the thalamic reticular nucleus initiate sequences of spindle oscillations in thalamic network. J Neurophysiol. 2000;84:1076–87. [PubMed: 10938329]
106.
Contreras D, Destexhe A, Sejnowski TJ, Steriade M. Control of spatiotemporal coherence of a thalamic oscillation by corticothalamic feedback. Science. 1996;274:771–774. [PubMed: 8864114]
107.
Destexhe A, Contreras D, Steriade M. Mechanisms underlying the synchronizing action of corticothalamic feedback through inhibition of thalamic relay cells. J Neurophysiol. 1998;79:999–1016. [PubMed: 9463458]
107a.
Bonjean M, Baker T, Lemieux M, Timofeev I, Sejnowski T, Bazhenov M. Corticothalamic feedback controls sleep spindle duration in vivo. J Neurosci. 2011;31:9124–9134. [PMC free article: PMC3131502] [PubMed: 21697364]
108.
Bal T, McCormick DA. What stops synchronized thalamocortical oscillations. Neuron. 1996;17:297–308. [PubMed: 8780653]
109.
Budde T, Biella G, Munsch T, Pape H-C. Lack of regulation by intracellular ca2+ of the hyperpolarization-activated cation current in rat thalamic neurones. J Physiol (Lond). 1997;503:79–85. [PMC free article: PMC1159888] [PubMed: 9288676]
110.
Luthi A, Bal T, McCormick DA. Periodicity of thalamic spindle waves is abolished by ZD7288,a blocker of Ih. J Neurophysiol. 1998;79:3284–3289. [PubMed: 9636128]
111.
Fuentealba P, Crochet S, Timofeev I, Bazhenov M, Sejnowski TJ, Steriade M. Experimental evidence and modeling studies support a synchronizing role for electrical coupling in the cat thalamic reticular neurons in vivo. Eur J Neurosci. 2004;20:111–9. [PMC free article: PMC2905213] [PubMed: 15245484]
112.
Landisman CE, Long MA, Beierlein M, Deans MR, Paul DL, Connors BW. Electrical synapses in the thalamic reticular nucleus. J Neurosci. 2002;22:1002–9. [PMC free article: PMC6758490] [PubMed: 11826128]
113.
Kim U, Bal T, McCormick DA. Spindle waves are propagating synchronized oscillations in the ferret lgnd in vitro. J Neurophysiol. 1995;74:1301–1323. [PubMed: 7500152]
114.
Contreras D, Destexhe A, Sejnowski TJ, Steriade M. Spatiotemporal patterns of spindle oscillations in cortex and thalamus. J Neurosci. 1997;17:1179–1196. [PMC free article: PMC6573181] [PubMed: 8994070]
115.
Dehghani N, Cash SS, Chen CC, Hagler DJ Jr, Huang M, Dale AM, Halgren E. Divergent cortical generators of MEG and EEG during human sleep spindles suggested by distributed source modeling. PLoS One. 2010;5:e11454. [PMC free article: PMC2898804] [PubMed: 20628643]
116.
Baker T, Bonjean M, Bazhenov M, Cash SS, Halgren E, Sejnowski T. Role of thalamocortical matrix projections in sleep spindles synchronization, SFN. San-Diego: 2010. [PMC free article: PMC3342310] [PubMed: 22496571]
117.
Sejnowski TJ, Destexhe A. Why do we sleep. Brain Res. 2000;886:208–223. [PubMed: 11119697]
118.
Bressler SL. The gamma wave: A cortical information carrier. Trends Neurosci. 1990;13:161–2. [PubMed: 1693231]
119.
Freeman WJ. The physiology of perception. Sci Am. 1991;264:78–85. [PubMed: 2000483]
120.
Steriade M, Amzica F, Contreras D. Synchronization of fast (30–40 hz) spontaneous cortical rhythms during brain activation. J Neurosci. 1996;16:392–417. [PMC free article: PMC6578724] [PubMed: 8613806]
120a.
Chauvette S, Crochet S, Volgushev M, Timofeev I. Properties of Slow Oscillation during Slow-Wave Sleep and Anesthesia in Cats. J Neurosci. 2011;31:14998–15008. [PMC free article: PMC3209581] [PubMed: 22016533]
121.
Traub RD, Whittington MA, Buhl EH, Jefferys JG, Faulkner HJ. On the mechanism of the gamma --> beta frequency shift in neuronal oscillations induced in rat hippocampal slices by tetanic stimulation. J Neurosci. 1999;19:1088–1105. [PMC free article: PMC6782135] [PubMed: 9920671]
122.
Traub RD, Spruston N, Soltesz I, Konnerth A, Whittington MA, Jefferys GR. Gamma-frequency oscillations: A neuronal population phenomenon, regulated by synaptic and intrinsic cellular processes, and inducing synaptic plasticity. Prog Neurobiol. 1998;55:563–75. [PubMed: 9670218]
123.
Traub RD, Jefferys JG, Whittington MA. Simulation of gamma rhythms in networks of interneurons and pyramidal cells. J Comput Neurosci. 1997;4:141–50. [PubMed: 9154520]
124.
Traub RD, Whittington MA, Stanford IM, Jefferys JG. A mechanism for generation of long-range synchronous fast oscillations in the cortex. Nature. 1996;383:621–624. [PubMed: 8857537]
125.
Lytton WW, Sejnowski TJ. Simulations of cortical pyramidal neurons synchronized by inhibitory interneurons. J Neurophysiol. 1991;66:1059–79. [PubMed: 1661324]
126.
Borgers C, Kopell N. Synchronization in networks of excitatory and inhibitory neurons with sparse, random connectivity. Neural Comput. 2003;15:509–38. [PubMed: 12620157]
127.
Whittington MA, Stanford IM, Colling SB, Jefferys JG, Traub RD. Spatiotemporal patterns of gamma frequency oscillations tetanically induced in the rat hippocampal slice. J Physiol. 1997;502(Pt 3):591–607. [PMC free article: PMC1159531] [PubMed: 9279811]
128.
Fisahn A, Pike FG, Buhl EH, Paulsen O. Cholinergic induction of network oscillations at 40 hz in the hippocampus in vitro. Nature. 1998;394:186–189. [PubMed: 9671302]
129.
Buhl EH, Tamas G, Fisahn A. Cholinergic activation and tonic excitation induce persistent gamma oscillations in mouse somatosensory cortex in vitro. J Physiol (Lond). 1998;513:117–26. [PMC free article: PMC2231263] [PubMed: 9782163]
130.
Traub RD, Whittington MA, Colling SB, Buzsaki G, Jefferys JG. Analysis of gamma rhythms in the rat hippocampus in vitro and in vivo. J Physiol. 1996;493(Pt 2):471–84. [PMC free article: PMC1158931] [PubMed: 8782110]
131.
Wang XJ, Buzsaki G. Gamma oscillation by synaptic inhibition in a hippocampal interneuronal network model. J Neurosci. 1996;16:6402–13. [PMC free article: PMC6578902] [PubMed: 8815919]
132.
Rulkov NF, Timofeev I, Bazhenov M. Oscillations in large-scale cortical networks: Map-based model. J Comput Neurosci. 2004;17:203–23. [PubMed: 15306740]
133.
Bazhenov M, Rulkov NF, Timofeev I. Effect of synaptic connectivity on long-range synchronization of fast cortical oscillations. J Neurophysiol. 2008;100:1562–1575. [PMC free article: PMC2652170] [PubMed: 18632897]
134.
Bouyer JJ, Montaron MF, Rougeul A. Fast fronto-parietal rhythms during combined focused attentive behaviour and immobility in cat: Cortical and thalamic localozations. Electroencephalography and Clinical Neurophysiology. 1981;51:244–252. [PubMed: 6163613]
135.
Rougeul-Buser A, Bouyer JJ, Buser P. From attentiveness to sleep. A topographical analysis of localized "synchronized" activities on the cortex of normal cat and monkey. Acta Neurobiol Exp (Warsz). 1975;35:805–19. [PubMed: 813498]
136.
Sheer DE. Focused arousal and the cognitive 40-hz event-related potentials: Differential diagnosis of alzheimer's disease. Prog Clin Biol Res. 1989;317:79–94. [PubMed: 2690123]
137.
Gray CM, Konig P, Engel AK, Singer W. Oscillatory responses in cat visual cortex exhibit inter-columnar synchronization which reflects global stimulus properties. Nature. 1989;338:334–7. [PubMed: 2922061]
138.
Murthy VN, Fetz EE. Coherent 25- to 35-hz oscillations in the sensorimotor cortex of awake behaving monkeys. Proc Natl Acad Sci U S A. 1992;89:5670–4. [PMC free article: PMC49354] [PubMed: 1608977]
139.
Pfurtscheller G, Neuper C. Simultaneous eeg 10 hz desynchronization and 40 hz synchronization during finger movements. Neuroreport. 1992;3:1057–60. [PubMed: 1493217]
140.
Womelsdorf T, Fries P, Mitra PP, Desimone R. Gamma-band synchronization in visual cortex predicts speed of change detection. Nature. 2006;439:733–6. [PubMed: 16372022]
141.
Joliot M, Ribary U, Llinas R. Human oscillatory brain activity near 40 hz coexists with cognitive temporal binding. Proc Natl Acad Sci U S A. 1994;91:11748–11751. [PMC free article: PMC45309] [PubMed: 7972135]
142.
Llinas R, Ribary U. Coherent 40-hz oscillation characterizes dream state in humans. Proc Natl Acad Sci U S A. 1993;90:2078–2081. [PMC free article: PMC46024] [PubMed: 8446632]
143.
Achermann P, Borbely AA. Coherence analysis of the human sleep electroencephalogram. Neuroscience. 1998;85:1195–1208. [PubMed: 9681957]
144.
Steriade M, Amzica F, Contreras D. Synchronization of fast I30–40 hz) spontaneous cortical rhythms during brain activation. J Neurosci. 1996;16:392–417. [PMC free article: PMC6578724] [PubMed: 8613806]
145.
Chrobak JJ, Buzsáki G. High-friquency oscillations in the output networks of the hippocampal-entorhinal axis of the freely behaving rat. J Neurosci. 1996;16:3056–3066. [PMC free article: PMC6579047] [PubMed: 8622135]
146.
Collins DR, Lang EJ, Pare D. Spontaneous activity of the perirhinal cortex in behaving cats. Neuroscience. 1999;89:1025–39. [PubMed: 10362292]
147.
Csicsvari J, Hirase H, Czurko A, Mamiya A, Buzsaki G. Oscillatory coupling of hippocampal pyramidal cells and interneurons in the behaving rat. J Neurosci. 1999;19:274–287. [PMC free article: PMC6782375] [PubMed: 9870957]
148.
Csicsvari J, Hirase H, Czurko A, Mamiya A, Buzsaki G. Fast network oscillations in the hippocampal CA1 region of the behaving rat. J Neurosci. 1999;19:RC20 (1–4). [PMC free article: PMC6782850] [PubMed: 10436076]
149.
Ylinen A, Bragin A, Nadasdy Z, Jando G, Szabo I, Sik A, Buzsaki G. Sharp wave-associated high-frequency oscillation (200 hz) in the intact hippocampus: Network and intracellular mechanisms. J Neurosci. 1995;15:30–46. [PMC free article: PMC6578299] [PubMed: 7823136]
150.
Jones MS, Barth DS. Spatiotemporal organization of fast (>200 hz) electrical oscillations in rat vibrissa/barrel cortex. J Neurophysiol. 1999;82:1599–609. [PubMed: 10482773]
151.
Jones MS, MacDonald KD, Choi B, Dudek FE, Barth DS. Intracellular correlates of fast (>200 hz) electrical oscillations in rat somatosensory cortex. J Neurophysiol. 2000;84:1505–18. [PubMed: 10980023]
152.
Kandel A, Buzsaki G. Cellular-synaptic generation of sleep spindles, swpike-and-wave discharges, and evoked thalamocortical responses in the neocortex of the rat. J Neurosci. 1997;17:6783–6797. [PMC free article: PMC6573130] [PubMed: 9254689]
153.
Grenier F, Timofeev I, Steriade M. Focal synchronization of ripples (80–200 hz) in neocortex and their neuronal correlates. J Neurophysiol. 2001;86:1884–1898. [PubMed: 11600648]
154.
Draguhn A, Traub RD, Schmitz D, Jefferys JG. Electrical coupling underlies high-frequency oscillations in the hippocampus in vitro. Nature. 1998;394:189–192. [PubMed: 9671303]
155.
Grenier F, Timofeev I, Steriade M. Neocortical very fast oscillations (ripples, 80–200 hz) during seizures: Intracellular correlates. J Neurophysiol. 2003;89:841–852. [PubMed: 12574462]
156.
Traub RD, Schmitz D, Jefferys JG, Draguhn A. High-frequency population oscillations are predicted to occur in hippocampal pyramidal neuronal networks interconnected by axoaxonal gap junctions. Neuroscience. 1999;92:407–426. [PubMed: 10408594]
157.
Schmidt M, Sudkamp S, Wahle P. Characterization of pretectal-nuclear-complex afferents to the pulvinar in the cat. Exp Brain Res. 2001;138:509–19. [PubMed: 11465750]
158.
Galarreta M, Hestrin S. Spike transmission and synchrony detection in networks of GABAergic interneurons. Science. 2001;292:2295–9. [PubMed: 11423653]
159.
Gibson JR, Beierlein M, Connors BW. Functional properties of electrical synapses between inhibitory interneurons of neocortical layer 4. J Neurophysiol. 2005;93:467–80. [PubMed: 15317837]
160.
Grenier F, Timofeev I, Crochet S, Steriade M. Spontaneous field potentials influence the activity of neocortical neurons during paroxysmal activities in vivo. Neuroscience. 2003;119:277–91. [PubMed: 12763088]
161.
Timofeev I, Steriade M. Neocortical seizures: Initiation, development and cessation. Neuroscience. 2004;123:299–336. [PubMed: 14698741]
162.
Steriade M, Contreras D. Spike-wave complexes and fast components of cortically generated seizures. I. Role of neocortex and thalamus. J Neurophysiol. 1998;80:1439–1455. [PubMed: 9744951]
163.
Steriade M, Amzica F, Neckelmann D, Timofeev I. Spike-wave complexes and fast components of cortically generated seizures. II. Extra- and intracellular patterns. J Neurophysiol. 1998;80:1456–1479. [PubMed: 9744952]
164.
Neckelmann D, Amzica F, Steriade M. Spike-wave complexes and fast components of cortically generated seizures. III. Synchronizing mechanisms. J Neurophysiol. 1998;80:1480–1494. [PubMed: 9744953]
165.
Timofeev I, Grenier F, Steriade M. Spike-wave complexes and fast components of cortically generated seizures. IV. Paroxysmal fast runs in cortical and thalamic neurons. J Neurophysiol. 1998;80:1495–1513. [PubMed: 9744954]
166.
Steriade M, Contreras D. Relations between cortical and thalamic cellular events during transition from sleep patterns to paroxysmal activity. J Neurosci. 1995;15:623–42. [PMC free article: PMC6578306] [PubMed: 7823168]
167.
Steriade M, Timofeev I. Corticothalamic operations through prevalent inhibition of thalamocortical neurons. Thalamus & related systems. 2001;1:225–236.
168.
Meeren HK, Pijn JP, Van Luijtelaar EL, Coenen AM, Lopes da Silva FH. Cortical focus drives widespread corticothalamic networks during spontaneous absence seizures in rats. J Neurosci. 2002;22:1480–95. [PMC free article: PMC6757554] [PubMed: 11850474]
169.
Nita D, Cisse Y, Frohlich F, Timofeev I. Cortical and thalamic components of neocortical kindling-induced epileptogenesis in behaving cats. Exp Neurol. 2008;211:518–528. [PubMed: 18423621]
170.
Halasz P. Runs of rapid spikes in sleep: A characteristic EEG expression of generalized malignant epileptic encephalopathies. A conceptual review with new pharmacological data. Epilepsy Res Suppl. 1991;2:49–71. [PubMed: 1662048]
171.
Kotagal P. Multifocal independent spike syndrome: Relationship to hypsarrhythmia and the slow spike-wave (lennox-gastaut) syndrome. Clin Electroencephalogr. 1995;26:23–9. [PubMed: 7882540]
172.
Niedermeyer E. Abnormal EEG patterns: Epileptic and paroxysmal, Electroencephalography: Basic principles, clinical applications, and related fields. Niedermeyer E, Lopes de Silva F, editors. Lippincott Williams & Wilkins; Philadelphia: 2005. pp. 255–280.
173.
Niedermeyer E. Epileptic seizure disorders, Electroencephalography: Basic principles, clinical applications, and related fields. Niedermeyer E, Lopes de Silva F, editors. Lippincott Williams & Wilkins; Philadelphia: 2005. pp. 505–620.
174.
Boucetta S, Chauvette S, Bazhenov M, Timofeev I. Focal generation of paroxysmal fast runs during electrographic seizures. Epilepsia. 2008;49:1925–1940. [PMC free article: PMC2629507] [PubMed: 18616553]
175.
Derchansky M, Rokni D, Rick JT, Wennberg R, Bardakjian BL, Zhang L, Yarom Y, Carlen PL. Bidirectional multisite seizure propagation in the intact isolated hippocampus: The multifocality of the seizure "focus". Neurobiol Dis. 2006;23:312–28. [PubMed: 16815026]
176.
Nita D, Cisse Y, Timofeev I. State-dependent slow outlasting activities following neocortical kindling in cats. Exp Neurol. 2008;211:456–468. [PubMed: 18420200]
177.
Timofeev I, Grenier F, Steriade M. Contribution of intrinsic neuronal factors in the generation of cortically driven electrographic seizures. J Neurophysiol. 2004;92:1133–1143. [PubMed: 14749320]
178.
Avoli M, Barbarosie M, Lucke A, Nagao T, Lopantsev V, Kohling R. Synchronous GABA-mediated potentials and epileptiform discharges in the rat limbic system in vitro. J Neurosci. 1996;16:3912–24. [PMC free article: PMC6578615] [PubMed: 8656285]
179.
Jensen MS, Yaari Y. Role of intrinsic burst firing, potassium accumulation, and electrical coupling in the elevated potassium model of hippocampal epilepsy. J Neurophysiol. 1997;77:1224–33. [PubMed: 9084592]
180.
Krnjevic K, Morris ME, Reiffenstein RJ. Stimulation-evoked changes in extracellular K+ and CA2+ in pyramidal layers of the rat's hippocampus. Can J Physiol Pharmacol. 1982;60:1643–57. [PubMed: 7165860]
181.
Nicholson C, Bruggencate GT, Steinberg R, Stockle H. Calcium modulation in brain extracellular microenvironment demonstrated with ion-selective micropipette. Proc Natl Acad Sci U S A. 1977;74:1287–90. [PMC free article: PMC430669] [PubMed: 265573]
182.
Silver RA, Lubke J, Sakmann B, Feldmeyer D. High-probability uniquantal transmission at excitatory synapses in barrel cortex. Science. 2003;302:1981–4. [PubMed: 14671309]
183.
Giaume C, McCarthy KD. Control of gap-junctional communication in astrocytic networks. Trends Neurosci. 1996;19:319–25. [PubMed: 8843600]
184.
Frohlich F, Bazhenov M. Coexistence of tonic firing and bursting in cortical neurons. Phys Rev E Stat Nonlin Soft Matter Phys. 2006;74:031922. [PubMed: 17025682]
Copyright © 2012, Michael A Rogawski, Antonio V Delgado-Escueta, Jeffrey L Noebels, Massimo Avoli and Richard W Olsen.

All Jasper's Basic Mechanisms of the Epilepsies content, except where otherwise noted, is licensed under a Creative Commons Attribution-NonCommercial-NoDerivs 3.0 Unported license, which permits copying, distribution and transmission of the work, provided the original work is properly cited, not used for commercial purposes, nor is altered or transformed.

Bookshelf ID: NBK98144PMID: 22787672

Views

  • PubReader
  • Print View
  • Cite this Page
  • PDF version of this page (1.8M)
  • PDF version of this title (143M)

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Similar articles in PubMed

See reviews...See all...

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...