Neuron, Vol. 33, 1–20, January 31, 2002, Copyright 2002 by Cell Press Theta Oscillations in the Hippocampus Review

Gyo¨ rgy Buzsa´ ki1 1990; Vinogradova, 1995; Vertes and Kocsis, 1997). The Center for Molecular and Behavioral theta models discussed in previous reviews utilize neu- Rutgers rons with passive cable properties. During the past sev- The State University of New Jersey eral years, it has become increasingly clear that and Neurological Institute of New Jersey are endowed with a host of active conductancies and Newark, New Jersey 07102 intrinsic oscillatory properties (Llina´ s, 1988; Ha¨ usser et al., 2000). A further goal of this review therefore is to integrate this new knowledge with network level infor- Theta oscillations represent the “on-line” state of the mation and propose new research directions for under- hippocampus. The extracellular currents underlying standing the functional significance of the theta rhythm. theta waves are generated mainly by the entorhinal input, CA3 (Schaffer) collaterals, and voltage-depen- Brain Systems Involved in the Generation ؉ dent Ca2 currents in pyramidal cell dendrites. The of Theta Oscillations rhythm is believed to be critical for temporal coding/ Theta oscillations depend on ongoing behavior (Gras- decoding of active neuronal ensembles and the modi- tya´ n et al., 1959; Vanderwolf, 1969). Although no consen- fication of synaptic weights. Nevertheless, numerous sus has yet emerged regarding their specific behavioral critical issues regarding both the generation of theta correlates, theta waves are most consistently present oscillations and their functional significance remain during REM sleep (Jouvet, 1969) and during various challenges for future research. types of locomotor activities described by the subjective terms “voluntary,” “preparatory,” “orienting,” or “explor- Functions of a brain structure can be deduced from atory” (Vanderwolf, 1969). In general, theta waves are correlation and perturbation methods. Convergence of absent in the immobile animal but short epochs of theta these methods at different levels of analysis indicates trains can be elicited by noxious conditioned stimuli (cf. that the hippocampus and associated structures serve Bland, 1986). Theta waves have been also assumed to to generate long-term memory traces. Clinical evidence be carriers of mnemonic processes (Miller, 1989; Lisman indicates that damage to the hippocampus produces and Idiart, 1995; Raghavachari et al., 2001). It should be anterograde amnesia. At the cellular-molecular level, the noted here that field oscillations at theta/␣ frequencies intricacies of synaptic plasticity, a candidate model for have been observed in numerous cortical structures memory storage, are being studied in great detail (cf. (Steriade, 2000). Intracranial recordings in humans indi- Kandel and Squire, 2000). However, the link between cate, however, that these patterns occur in different single computation and computation at the net- behavioral states and are not coherent with hippocam- work level is poorly understood. It remains unclear how pal theta waves (Kahana et al., 1999; Raghavachari et neuronal cooperativity in intact networks relates to al., 2001). Thus, the relevance of these rhythms to theta memories or how network activity in the behaving animal oscillations present in the hippocampus and associated brings about synaptic modification. Because network structures has yet to be established. patterns arise from the collective action of neurons, In the discussion of the mechanisms of oscillations, some of their behavior-related changes can be studied it is useful to distinguish two terms. The term “current by recording the current flow in the extracellular space. generator” refers to the transmembrane currents re- A prominent network pattern in the hippocampus of all sponsible for the magnitude of the recorded field. On the mammals studied to date, including humans (Arnolds other hand, “rhythm generator” refers to mechanisms et al., 1980; Kahana et al., 1999; Tesche and Karhu, responsible for the emergence and control of the oscilla- 2000; Bo´ dizs et al., 2001), is a slow oscillation in the tory pattern and frequency. Brain regions with parallel theta-␣ frequency band. Key issues therefore are to un- arranged dendrites and afferents, such as cortical struc- derstand how theta oscillation can group and segregate tures, give rise to large amplitude extracellular poten- neuronal assemblies and to assign various computa- tials, whereas subcortical nuclei with less orderly spatial tional tasks to them. An equally important task is to organization generate “closed fields,” i.e., small-ampli- reveal the relationship between synaptic activity (as re- tude field events. flected globally by field theta) and the output of the Theta oscillation is most regular in frequency and active single cells (as reflected by action potentials). A largest amplitude in the str. lacunosum-moleculare of goal of this review is to summarize new knowledge about the hippocampal CA1 region. Both the amplitude and the cellular-synaptic generation of theta waves and the phase of theta waves change as a of depth (i. importance of theta oscillations in the coordination of e., different layers; Figure 1), whereas in the same layers neuronal networks and in the modification of synaptic they are robustly similar along the long axis of the hippo- connections. Several excellent reviews have been pub- campus (Bullock et al., 1990). Theta oscillations are also lished on the physiology and pharmacology of theta present in the dentate gyrus and the CA3 region. In oscillations (Bland, 1986; Vanderwolf, 1988; Lopes da addition to the hippocampal formation, theta frequency Silva et al., 1990; Buzsa´ ki et al., 1994; Stewart and Fox, oscillations and phase-locked discharge of neurons to theta waves have been observed in several other struc- 1Correspondence: buzsaki@.rutgers.edu tures, including the subicular complex, entorhinal cor- Neuron 2

Figure 1. Voltage-versus-Depth Profile of Theta Oscillation in the Rat (Left) A 16-site silicon probe in the CA1-den- tate gyrus axis. Numbers indicate recording sites (100 ␮m spacing). o, str. oriens; p, pyra- midal layer; r, str. radiatum; lm, str. lacuno- sum-molecular; g, granule cell layer; h, hilus. (Right) Theta waves recorded during explora- tion. Note gradual shift of theta phase from str. oriens to str. lacunosum-moleculare. Gamma waves superimposed on theta oscil- lation are marked by arrows. Vertical bar: 1 mV. (From Bragin et al., 1995.)

tex, perirhinal cortex, cingulate cortex, and amygdala and Magoun, 1949) rather than structures specifically (Adey, 1967; Mitchell and Ranck, 1980; Alonso and Gar- involved in theta rhythm generation. In summary, the cia-Austt, 1987; Leung and Borst, 1987; Pare´ and Col- minimum conditions necessary for the generation of os- lins, 2000). These structures are thus the main current cillating extracellular currents in the theta frequency generators of the extracellularly recorded theta field. band are the proper connections between the hippo- However, none of these cortical structures are capable campus and MS-DBB. Despite 40 years of research of generating theta activity on their own. (Petsche et al., 1962), however, the exact physiological Several subcortical nuclei have been postulated to mechanisms of these interactions have remained unre- be critically involved in the rhythm generation of theta. solved. Afferents from these nuclei may release neurotransmit- ters that may allow for the emergence of network oscilla- “Classic” Theta Model and Its Inadequacies tions in the hippocampus and associated structures In the first and simplest theta model, the MS-DBB has (“permissive” action) or may provide a coherent, theta been postulated to be the rhythm generator (pace- frequency output (“pacemaker” function). Because le- maker), which supplies phasic modulation to the hippo- sion or inactivation of medial septum-diagonal band of campus (Petsche et al., 1962). Subsequent models Broca (MS-DBB) neurons abolishes theta waves in all added new components that are summarized in Figure cortical targets, it has been regarded as the ultimate 2. On the assumption that the extracellular field is gener- rhythm generator of theta (Petsche et al., 1962). The MS- ated by the summed activity of IPSPs and EPSPs on DBB is reciprocally connected to the supramammillary the somata and dendrites of principal cells, respectively, region (Borhegyi and Freund, 1998; Leranth et al., 1999), these models utilized a single canonical CA1 pyramidal a second critical structure involved in pacing the theta cell with passive membrane properties. It has been as- rhythm (cf. Vertes and Kocsis, 1997). Whether the MS- sumed that all pyramidal cells receive coherent excit- DBB and supramammillary nucleus are true pacemakers atory (from perforant path) and inhibitory (from septum or rhythmic firing of their neurons depends on the hippo- to feed-forward inhibitory neuron) inputs. The interplay campal and entorhinal feedback has yet to be deter- between these two current generators (dipoles) is as- mined (Lee et al., 1994; Brazhnik and Fox, 1999; King sumed to be responsible for the unique amplitude/phase et al., 1998; Borisyuk and Hoppensteadt, 1999; Denham versus depth profiles of hippocampal theta oscillation. and Borisyuk, 2000; Wang, 2002). According to this scheme, the most strongly excited Neurons in several other subcortical structures are minority of the population would discharge at the same phase locked to hippocampal theta oscillation, including time when nonspiking neurons are maximally depolar- the dorsal raphe nucleus, ventral tegmental nucleus of ized, justifying the “lumped” model. As Figure 1 illus- Gudden, and anterior thalamic nuclei (cf. Bland, 1986; trates, theta waves in the waking rat show a gradual Vertes and Kocsis, 1997). Finally, stimulation of several phase reversal between CA1 str. oriens and the hippo- subcortical nuclei elicits hippocampal theta (cf. Bland, campal fissure (Winson, 1974), consistent with the coor- 1986). However, these latter structures also project to dinated activity of at least two current generators (di- the thalamus and/or as well, and they are poles). Because the largest amplitude theta waves are part of a common ascending activation system (Moruzzi observed at the hippocampal fissure, rhythmic excita- Review 3

excitatory and inhibitory inputs as postulated from field/ CSD studies and the actual discharge of the pyramidal cells and interneurons. If the major excitatory drive dur- ing theta is the entorhinal input, CA1 pyramidal cells are expected to discharge maximally at the peak of the sink in str. lacunosum-moleculare. However, the highest probability of discharge in the behaving rat occurs around the positive peak of theta recorded at the level of the distal dendrites, corresponding to the negative phase of the theta waves in the pyramidal layer (Figure 3A; Fox et al., 1986; Buzsa´ ki et al., 1986; Brankack et al., 1993; Csicsvari et al., 1999). (2) Another finding, which is at odds with the classic model, is that the theta phase relationship of pyramidal cells is not fixed but changes dynamically as a function of behavior (O’Keefe and Recce, 1993; Skaggs et al., 1996). (3) Recent work indi- cates that pyramidal neurons in various limbic structures are endowed to oscillate at theta frequency. These in- Figure 2. “Classic” Model of Extracellular Theta Current Generation trinsic mechanisms may be as important in the genera- The medial septum and diagonal band of Broca (MS-DBB) area are tion of transmembrane currents as postsynaptic poten- assumed to be the rhythm generator (pacemaker) of theta. Choliner- tials. (4) The complex interconnections among the gic neurons (c) provide slow depolarization of their target pyramidal numerous hippocampal interneuron classes pose fur- cell and basket interneuron (i). MS-DBB GABAergic neurons (g) ther questions about their involvement in rhythm genera- rhythmically hyperpolarize the basket interneuron (i). In turn, the tion. (5) Importantly, more recent knowledge indicates rhythmic IPSPs in the pyramidal cell are assumed to induce periso- that the recurrent circuit of the CA3 region may function matic currents (red; inhibitory theta dipole). Rhythmic EPSPs from the entorinal cortex (EC) are responsible for the active sink in the as an intrahippocampal theta oscillator. (6) Finally, the distal dendritic region (blue; excitatory theta dipole). The relative mutual connections between MS-DBB and the hippo- magnitude and phase relationship of the inhibitory (somatic) and campus/entorhinal cortex indicate that the “septal pace- excitatory (distal dendritic) dipoles are hypothesized to determine maker”-“hippocampal follower” model is overly simple. the unique amplitude and phase versus depth distribution of theta We will discuss these topics below. Some potential solu- waves in the CA1 region. (Adapted from Holsheimer et al., 1982; tions will be offered whereas answers to several other Buzsa´ ki et al., 1983; Leung, 1984; Stewart and Fox, 1990; Lee et al., 1994.) issues will require further experimentation.

Transmitters and Receptors Responsible for Theta tion of the distal dendrites by the entorhinal afferents is Oscillations: Roles for Cholinergic, assumed to play the most important role in the current GABAA, and NMDA Receptors generation of extracellular field theta. Drugs affecting the field theta waves may interfere with A second theta dipole in the CA1 region is assumed the rhythm and/or the current generators. For example, to be generated by somatic IPSPs. These IPSPs are blockade or potentiation of GABAA receptors during pic- brought about by the ␥ frequency discharge of basket rotoxin-induced epilepsy or pentobarbital anesthesia, and chandelier cells, repeated at theta frequency (Ar- respectively, eliminates theta by affecting both rhythm temenko, 1972; Fox, 1989; Leung and Yim, 1986; Ylinen generation and current generation. In principle, every et al., 1995; Kamondi et al., 1998a). The rhythmic drive afferent pathway with phase-locked activity to the global and/or inhibition of basket cells may arrive from the rhythm contributes to the extracellular field. Given the MS-DBB (Buzsa´ ki et al., 1983; Stewart and Fox, 1990), large number of structures affecting theta oscillations, although several other inputs, including the entorhinal it is not surprising that theta activity can be influenced afferents, CA3 afferents, CA1 recurrent collaterals, and by a variety of drugs (Vanderwolf, 1988). Vanderwolf other interneurons with intrinsic oscillatory properties, (Kramis et al., 1975) suggested that, on the basis of may also be involved. Current-source density (CSD) pharmacological sensitivity, two types of theta could be analysis (Mitzdorf, 1985), showing a strong sink in the distinguished: atropine-sensitive and atropine-resis- CA1 str. lacunusom-moleculare and a source in the py- tant. The idea of an atropine-sensitive form of theta ramidal layer also support the above scheme (Buzsa´ ki comes from early observations that muscarinic block- et al., 1986; Brankack et al., 1993). The gradual phase ers, such as atropine, completely eliminate theta oscilla- shift of the theta waves with depth (Winson, 1974) and tions in anesthetized animals. In contrast, in the awake, the lack of a clear “null” zone are explained by the phase- walking rat, the amplitude and frequency of theta oscilla- shifted nature of the somatic and dendritic dipoles (Buz- tion do not substantially change even after large doses sa´ ki et al., 1983; Leung, 1984). of systemically administered muscarinic blockers, al- Research over the past several years points out sev- though the wave shape and depth profile of theta under eral inadequacies of the above model. (1) Although both atropine are quantitatively different from those in the CSD analysis and unit recording studies suggest that drug-free animal (Buzsa´ ki et al., 1986). This persisting dendritic excitation is coupled to somatic inhibition (Fig- form of theta is referred to as “atropine resistant” (Van- ures 3 and 4; Csicsvari et al., 1999), in the classic theta derwolf, 1988). The neurotransmitter(s) and receptor(s) model a controversy exists between the timing of the responsible for the atropine-resistant type of theta have Neuron 4

Figure 3. Theta Phase Relationship of Pyra- midal Cell Spiking Is Variable (A) Averaged field theta wave and the pre- ferred phase distribution of single interneur- ons in the pyramidal layer [int(p)] and alveus- str. oriens [int(a/o)] and CA1 pyramidal cells of the rat. Only neurons with significant phase modulation are included. Note that most in- terneurons and pyramidal cells discharge on the negative phase of local (pyramidal layer) theta waves. (B) Relationship between firing rate and theta phase distribution of spikes for a single pyramidal neuron. Line: running circular mean. The phase was calculated for each firing rate bin separately. Note the half- cycle phase shift from Ͻ2HztoϾ15 Hz. (C) Illustration of spike-phase relationship for weakly and strongly activated neurons. The strongly activated neuron (due to hypothe- sized coactivation by the entorhinal and Schaffer afferents; red) discharges on the negative phase of local extracellular (EC) theta. Neurons with threshold activation dis- charge on the positive phase (blue, green, and purple cells). Nonspiking cells are black. The strongly active few pyramidal cells thus may contribute more action potentials per theta cycle than the remaining population combined. (From Csicsvari et al., 1999 and unpublished data from K. Harris, H. Hirase, X. Leinekugel, and G.B.)

never been identified conclusively (Vanderwolf, 1988; We hypothesize that activation of NMDA receptors is Vertes and Kocsis, 1997). critical for the atropine-resistant form of theta oscillation. Complete surgical removal of the entorhinal cortex Although the pharmacological action of urethane is not or surgical isolation of the entorhinal cortex from its well understood, experiments using more specific drugs nonhippocampal afferents eliminates the theta dipole provide support for this hypothesis. First, the depth versus localized on the banks of the hippocampal fissure (Fig- voltage profile of theta under the NMDA receptor blocker, ure 4). Importantly, such lesions render the remaining ketamine, is similar to that described under urethane theta oscillation atropine sensitive (Buzsa´ ki et al., 1983; (Soltesz and Desche´ nes, 1993). Second, combination of Vanderwolf and Leung, 1983) and its depth versus volt- ketamine or other NMDA receptor blockers (phencycli- age profile somewhat similar to that observed under dine, 3-(2-carboxypiperazin-4-yl) propyl-1-phosphonic urethane anesthesia (Ylinen et al., 1995; Kamondi et al., acid [CPP], DL-2-amino-5-phosphonovaleric acid [APV]) 1998a). Two important hypotheses can be deduced from and atropine or scopolamine abolished all theta activity these observations. First, that the receptors involved in in the hippocampus (Vanderwolf and Leung, 1983; Solt- atropine-resistant type of theta are urethane sensitive esz and Desche´ nes, 1993; Horvath et al., 1988). Earlier (Kramis et al., 1975). The second deduction is that the experiments, using anesthetic doses of urethane or atropine-resistant component of hippocampal theta is NMDA blockers, reported that the frequency of atropine- conveyed by layer III and layer II entorhinal cortical affer- sensitive theta is substantially less (2–5 Hz) than in the ents to the CA1 and dentate/CA3 neurons, respectively awake rat (6–9 Hz). However, subanesthetic doses of (Amaral and Witter, 1989). Since these pathways contain NMDA receptor blockers or their intracerebro-ventricu- glutamate (cf. Bland, 1986) and urethane attenuates glu- lar application are also effective in reducing the atro- tamate release from presynaptic vesicles (Moroni et al., pine-resistant theta component (Horvath et al., 1988; 1981), one might expect quantitative similarities be- L.S. Leung, unpublished observations). A candidate tar- tween the CSD maps of theta and those evoked by get of the NMDA blockers is the entorhinal afferent syn- electrical stimulation of the entorhinal afferents. How- apses on the distal apical dendrites of CA1 pyramidal ever, the theta dipoles mediated by the entorhinal cortex neurons. These synapses have at least two distinguish- cannot be fully explained by glutamate activation of ing features. First, they are larger than the synapses in pyramidal and granule cells via fast acting AMPA recep- str. oriens and radiatum and are frequently perforated tors only (Buzsa´ ki et al., 1986; Brankack et al., 1993; (Megias et al., 2001). Second, in addition to spines, the Bragin et al., 1995). terminals frequently contact dendritic shafts. Activation Review 5

Figure 4. Entorhinal Cortex-Mediated Theta Dipole in the Hippocampus The simultaneously obtained field potentials (see Figure 1) were converted to one-dimensional current-source density (CSD) maps. Left: intact rat. Voltage trace of theta recorded in the CA1 pyramidal layer is superimposed for reference (gray trace). Rhythmic sources (red) in the pyramidal layer (p) are coupled to rhythmic sinks (blue) in the stratum lacunosum-moleculare (lm) representing a putative inhibitory source (I ) and excitation by the perforant path input (PP). An additional sink, mediated by the Schaffer collaterals of CA3 pyramidal cells (Sch), is present in str. radiatum (r). Right: same animal 2 days after bilateral removal of the entorhinal cortex (EC). Note the absence of sink at the distal dendrites and the survival of the more proximal source-sink pair. The same time and color scales apply to both CSD maps. M, molecular layer; g, granule cell layer. (From Kamondi et al., 1998a.) of the NMDA receptors on the distal apical dendrites ally associated with synaptic plasticity (Kentros et al., may be brought about by theta phase-locked Ca2ϩ po- 1998), these findings indicate that NMDA receptors lo- tentials in distal dendrites of discharging CA1 pyramidal cated on the distal apical dendrites are also important neurons (Figure 5). Furthermore, in vitro experiments in spontaneous synaptic events and the maintenance also indicate that the entorhinal input has a greater of synaptic function (Kovalchuk et al., 2000). NMDA component than the associational (CA3) input The precise targets of systemic atropine in hippocampal to CA1 pyramidal cells (Otmakhova and Lisman, 2000). theta generation are unknown. The remaining theta sinks Although NMDA receptors in the hippocampus are usu- and sources after bilateral lesion of the entorhinal cortex

Figure 5. Voltage-Dependent Theta Oscilla- tion in Pyramidal Cell Dendrites In Vivo (A) Continuous recording of extracellular (ex- tra; CA1 pyr layer) and intradendritic (intra) activity in the rat. Holding potential was man- ually shifted to progressively more depolar- ized levels by intradendritic current injection. Inset: location of the dendritic penetration. The recording electrode contained QX 314 to block fast (Naϩ) spikes. Some of the high- threshold Ca2ϩ spikes are marked by arrows. (B) Another dendritic recording without QX 314 (370 ␮m from pyramidal layer). Note large amplitude Ca2ϩ spikes and small amplitude fast spikes. (Reprinted from Kamondi et al., 1998a.) Neuron 6

are more compatible with associational/commissural in- the theta generation in the entorhinal cortex depends puts and perisomatic inhibition than with the distribution on nicotinic receptors (but see Whishaw, 1976). of septo-hippocampal cholinergic terminals, which are present in all layers (Kamondi et al., 1998a). This obser- Contribution of the Dentate Gyrus and CA3 vation suggests that cholinergic-muscarinic receptors Recurrent Collateral System in the hippocampus are not responsible directly for the to Theta Generation extracellular theta currents. Furthermore, cholinergic Neurons in all hippocampal regions are phase modu- activation of M1 receptors of pyramidal cells is much lated by theta oscillation (Bland et al., 1975; Buzsa´ ki et too slow (Cole and Nicoll, 1983; Hasselmo and Fehlau, al., 1983; Fox et al., 1986). Therefore, both the numerous 2001) for the generation of theta-associated cyclic mem- granule cells and CA3 pyramidal cells are expected to brane potential changes. Release of acetylcholine may generate their own theta fields. simply depolarize pyramidal cells and interneurons Although theta phase does not fully reverse across (Madison et al., 1987) and/or affect voltage-dependent the granule cell layer, the waves recorded in the outer conductancies, such as IM,Ih, and IA (Hoffman et al., molecular layer and the hilus are phase shifted by ap- 1997). Muscarinic receptors are both synaptic and extra- proximately 90Њ (Buzsa´ ki et al., 1983). Furthermore, large synaptic and are present on presynaptic boutons of both phase jumps (180Њ) have been described in the molecu- glutamatergic and GABAergic (Hasselmo and lar layer (Buzsa´ ki et al., 1986). If the neighboring excit- Schnell, 1994; Vizi and Kiss, 1998; Ha´ jos et al., 1998). atory dipoles, formed presumably by afferents originat- Furthermore, muscarinic blockers can affect receptors ing in the medial and lateral entorhinal cortex and the in both MS-DBB and hippocampus. We therefore con- hilar mossy cells (Amaral and Witter, 1989), overlap in clude that EPSPs brought about by the MS-DBB cholin- phase, the extracellular currents may be cancelled or ergic neurons on hippocampal pyramidal cells cannot reduced. The relatively smaller contribution of granule be responsible for the atropine-sensitive form of theta. cells to the overall hippocampal theta field is illustrated An alternative mechanism to the direct excitation of further by the observation that elimination of granule pyramidal cells by acetylcholine in theta generation cells by neonatal X-ray irradiation does not change the (Petsche et al., 1962) is the cholinergic modulation of depth versus voltage profile of theta waves appreciably interneurons (Buzsa´ ki et al., 1983; Stewart and Fox, in the anesthetized rat (Whishaw et al., 1978). In contrast, 1990). Tonic cholinergic excitation of interneurons, cou- elimination of CA1 pyramidal neurons by forebrain isch- pled with their phasic septal GABAergic inhibition, has emia results in a dramatic decrease of theta amplitude been suggested to be responsible for the rhythmic dis- (Monmaur et al., 1986; Buzsa´ ki et al., 1990). charge of hippocampal interneurons (Freund and Antal, The CA3 pyramidal cells contribute to theta fields re- 1988; Stewart and Fox, 1990). In turn, the theta fre- corded in both the CA3 and CA1 regions. The CA3 output quency discharge of the interneurons imposes rhythmic contributes directly to the theta field recorded in the CA1 IPSPs on their target principal cells (Fox, 1989; Soltesz region, as demonstrated by a sink in CA1 str. radiatum and Desche´ nes, 1993; Ylinen et al., 1995; Cobb et al., (Figure 4; Buzsa´ ki et al., 1986; Brankack et al., 1993). 1995; To´ th et al., 1997; Kamondi et al., 1998a). According The small magnitude of the sink is explained by the to this hypothesis, the theta waves observed in the anes- observation that only a small percentage of CA3 pyrami- thetized animal and in the entorhinal cortex-damaged dal cells are active during a given theta cycle. Neverthe- brain should represent mainly somatic outward currents less, this minority appears to be critical for the discharge and would be localized close to the pyramidal layer of target CA1 neurons because CA3 pyramidal cells are (Figure 2; Buzsa´ ki et al., 1983; Leung, 1984). Alterna- active on the same phase of theta as CA1 pyramidal tively, the theta dipoles, which remain after removing the entorhinal inputs (Figure 4), can be accounted for by cells (Fox et al., 1986; J. Csicsvari and G.B., unpublished the more proximal dendritic excitation of CA1 pyramidal data). cells by the associational (Schaffer) collaterals. This in- Despite the similar cytoarchitecture of the CA1 and terpretation assumes that the atropine-sensitive form of CA3 regions, the extracellular theta currents in the CA3 theta emerges in the CA3 recurrent collateral system region are considerably smaller than in CA1. Several (Kocsis et al., 1999). factors may be responsible for this conspicuous differ- It should be spelled out that there is an unresolved ence. First, the distal dendritic arbor of CA3 pyramidal discrepancy between the pharmacological blockage of cells is considerably smaller than that of the CA1 pyrami- the muscarinic M1 receptors and the absence of MS- dal neurons (Turner et al., 1995; Pyapali et al., 1998). DBB cholinergic neurons. Whereas the amplitude of hip- Second, in addition to the inputs common to CA3 and pocampal theta is only modestly affected after atropine CA1, CA3 pyramidal neurons receive perisomatic excita- treatment, it is reduced several-fold after selective neu- tion near their somata from the large mossy terminals rotoxin elimination of MS-DBB cholinergic cells (Lee et of granule cells (cf. Henze et al., 2000). The currents al., 1994). This discrepancy suggests that muscarinic induced by the mossy boutons flow in a direction oppo- M2 (Ha´ jos et al., 1998) and/or nicotinic receptors (Ji and site to the currents generated by the perforant path and Dani, 2000) may also be involved in the regulation of the recurrent collaterals, leading to a potentially large theta. Because the atropine-resistant type of theta in reduction of extracellular currents. Finally, the excitatory the hippocampus is conveyed by the entorhinal input, drive from the entorhinal layer II and layer III neurons to the theta generated in the entorhinal cortex is likely not CA3 and CA1 pyramidal cells may be different. Re- affected by muscarinic blockade. On the other hand, cording from identified neurons in the entorhinal cortex lesion of the MS-DBB abolishes theta oscillations in the in the behaving rat will be required to address this latter entorhinal cortex as well. It is possible therefore that issue. Review 7

While the CA1 and dentate regions act only as current IA,IAHP,IM, and IK[Ca]); Madison et al., 1987; Fellous and generators of theta, recent findings indicate that the Sejnowski, 2000).When the oscillating cells communi- CA3 recurrent collateral system can also act as a rhythm cate with each other, a network oscillation is expected generator. After surgical removal of the entorhinal cor- to occur. However, in contrast to the continuous theta tex, all remaining theta appears to depend on the integ- in the intact brain, theta-like oscillations in vitro consist rity of the CA3 region, and theta signals in all layers are of only a limited number of cycles (Konopacki et al., highly coherent with each other (Bragin et al., 1995). 1987; MacVicar and Tse, 1989; Traub et al., 1992; Van In contrast, the coherence of theta signals in CA1 str. der Linden et al., 1999). A possible explanation for the radiatum and str. lacunosum-moleculare in the intact short trains is that muscarinic receptor activation may brain is low and the powers of theta signals in these attenuate the spread of collateral excitation. In support layers are inversely correlated with each other (Kocsis of this assertion, acetylcholine has been suggested to et al., 1999). On the other hand, theta waves in CA1 str. attenuate glutamate release from the recurrent and radiatum and in the inner third of the dentate molecular Schaffer collaterals (Hasselmo and Schnell, 1994). From layer are strongly related. Since these layers are the the in vitro observations, it is tempting to conclude that targets of the intrahippocampal associational projec- the necessary and sufficient condition for the production tions (Amaral and Witter, 1989), the findings suggest of atropine-sensitive type of theta is a relatively intact that the recurrent network of CA3 pyramidal cells and CA3 recurrent collateral system and diffuse release of possibly hilar mossy cells forms an intrahippocampal acetylcholine in the extracellular environment. However, oscillator. The intrahippocampal theta oscillator re- when these conditions were created in vivo by grafting quires cholinergic activation because in the absence of the fetal septal region into the subcortically denervated the entorhinal input, the remaining theta is abolished by hippocampus, theta oscillation failed to emerge (Segal atropine. et al., 1985; Buzsa´ ki et al., 1987) even though release of The notion that the cholinergic component of theta acetycholine from the graft reached physiological levels emerges in the CA3 recurrent collateral system implies (Leanza et al., 1993). In addition, the magnitude of carba- that this form of oscillation can also be studied in the chol-induced synchrony in the hippocampal slice is sev- isolated hippocampus. To this end, Konopacki et al. eral times larger than that observed in vivo (Traub et al., (1987; Bland et al., 1988) observed that when a hippo- 1992). Despite these differences, the in vivo and in vitro campal slice is bathed in a solution containing the mus- experiments provide support for the existence of an carinic drug carbachol, short bursts of field oscillations intrinsic hippocampal theta oscillator. A broader impli- at 4 to 15 Hz occur intermittently. Depending on the cation of these observations is that in the behaving ani- concentration of the drug, the magnitude of the field mal, the intrahippocampal (CA3) theta oscillator can oscillation and the synchrony of the neuronal population change its frequency and phase relatively independently vary from in vivo theta-like pattern to overt epileptic from the extrahippocampal (entorhinal) theta inputs Ͻ patterns (Williams and Kauer, 1997). At a low dose ( 5 (Kocsis et al., 1999). The phase differences, in turn, can ␮ M), field oscillation may not be evident, although some have a profound effect on the timing of action potentials CA3 pyramidal neurons may show sustained subthresh- in the activated principal neurons. old oscillation at 5 to 15 Hz. Moder- ate doses (15–40 ␮M) induce field oscillations and asso- The Involvement of GABAergic Interneurons ciated discharge of pyramidal neurons (Fellous and in Theta Activity Sejnowski, 2000). The carbachol-induced rhythm is gen- Ample evidence supports the critical involvement of hip- erated in the CA3 region and attenuated by AMPA recep- pocampal interneurons in theta oscillations (Buzsa´ ki et tor blockers. In the CA1 region, theta waves are reversed in phase in the str. oriens/pyramidale and str. radiatum. al., 1983). Hippocampal interneurons are the exclusive Although some interneurons are entrained by the net- targets of the GABAergic septo-hippocampal projection work rhythm (Pitler and Alger, 1992; McMahon et al., (Freund and Antal, 1988) as well as the sole hippocampal 1998), they are not essential for the maintenance of the output to the neurons of MS-DBB (To´ th et al., 1993). rhythm (MacVicar and Tse, 1989; Traub et al., 1992; Thus, they are in a strategic position to amplify the hy- Fellous and Sejnowski, 2000), a main difference from pothesized septo-hippocampal pacemaker connection. the in vivo situation. Similar oscillatory patterns have Nevertheless, the exact role of each of the numerous also been observed in the organotypic hippocampal interneuron classes (Freund and Buzsa´ ki, 1996) in this slices and in hippocampus-septum cocultures (Fischer process has yet to be determined. In the behaving rat, et al., 1999). In the combined cultures, enhancement of the majority of CA1 interneurons discharge on the de- spontaneously released acetylcholine levels, by block- scending phase of theta in the pyramidal cells layer ing the enzyme acetylcholine esterase, is sufficient for the (ascending phase in the apical dendritic layers), and are production of theta-like network oscillation (Fischer assumed to be responsible for the increased ␥ power et al., 1999). on this phase (Bragin et al., 1995). However, it is not The mechanisms by which theta-like oscillations known whether interneurons with different preferred emerge in the hippocampal slice preparation have yet theta phases represent different classes of cells or to be disclosed. Pyramidal neurons are endowed with whether interneurons of the same class can discharge at intrinsic properties to oscillate at theta frequency (see different phases. At least three classes of interneurons below). Appropriate activation of such intrinsic oscilla- deserve special attention: (1) basket and chandelier cells tory mechanism may elicit theta rhythm in single cells. with perisomatic targets, (2) O-LM (oriens lacunosum- Acetylcholine or carbachol induces dendritic depolar- moleculare) and HIPP (hilar interneuron with perforant ϩ ization and can affect several K and other channels (Ih, path axon projection) interneurons (Halasy and Somo- Neuron 8

gyi, 1993; Freund and Buzsa´ ki, 1996), which specifically of these dendritic segments is critical in this process innervate the termination zones of entorhinal afferents, (Katona et al., 1999; Maccaferri et al., 2000). O-LM and and (3) interneurons with feedback septal projection HIPP interneurons specifically innervate the termination (Alonso and Kohler, 1982; To´ th et al., 1993). zones of entorhinal afferents on pyramidal cells and Perisomatic inhibition plays an important role in timing granule cells, respectively (Freund and Buzsa´ ki, 1996). the action potentials of principal cells within the theta Since an important excitatory input to these interneur- cycle (Buzsa´ ki et al., 1983). Basket and chandelier cells ons is the local collaterals of CA1-CA3 pyramidal cells discharge rhythmically at ␥ frequency on the descending and granule cells (Sik et al., 1995, 1997; Blasco-Iban˜ ez phase of the pyramidal layer theta and provide rhythmic and Freund, 1995), the feedback dendritic inhibition can hyperpolarization to the perisomatic region of pyramidal provide a “winner-take-all” mechanism and thereby pre- cells via GABAA receptor-mediated IPSPs (Artemenko, vent the discharge of weakly activated pyramidal neu- 1972; Fox, 1989; Leung and Yim, 1991; Soltesz and rons and granule cells. Assuming that O-LM and HIPP Desche´ nes, 1993; Ylinen et al., 1995; Cobb et al., 1995; neurons fire on the same phase of theta waves as the To´ th et al., 1997). A functional consequence of such input entorhinal cells, the outward inhibitory currents perisomatic shunting is periodic “isolation” of the so- they bring about may effectively reduce the inward cur- matic and dendritic compartments of pyramidal neurons rents conveyed by the entorhinal afferents. The ratio of at the time of maximum somatic inhibition (Figure 3A). inhibitory versus excitatory terminals is 2- to 3-fold Because the equilibrium potential of ClϪ-mediated higher in str. lacunosum-moleculare as compared to the IPSPs is close to the resting membrane potential of str. radiatum (Megias et al., 2001). On the other hand, the principal cells, the contribution of inhibition to the selective reduction of distal dendritic inhibition in dis- extracellular field depends on the state of the individual charging pyramidal cells may allow for the backpropa- pyramidal cells. In the majority of pyramidal cells, the gation of action potentials from soma to dendrites

GABAA receptor-mediated transmembrane currents are (Spruston et al., 1995; Buzsa´ ki et al., 1996; Tsubokawa very small because most pyramidal cells are silent dur- and Ross, 1996). Temporal coordination of distal den- ing theta activity (Harris et al., 2000).Thus, the quantita- dritic excitation and active spike backpropagation, in tive contribution of inhibition to the extracellular theta turn, may contribute to the theta frequency-triggered field has yet to be determined. dendritic Ca2ϩ spikes (Figure 5).

The role of GABAB receptors in theta oscillation is not A third group of cells that may be critically involved clear. Activity of a single presynaptic interneuron rarely in the rhythm generation of theta oscillation is septally activates postsynaptic GABAB receptors (Buhl et al., projecting interneurons (Alonso and Kohler, 1982; To´ th 1994; Cobb et al., 1995). However, because at least 60% et al., 1993). This special group includes the “backpro- of putative basket cells discharge synchronously during jection” or widely projecting class of interneurons. CA1 the theta cycle (Csicsvari et al., 1999), activating more backprojection interneurons receive feedback excita- than half of the approximately 100 synapses on pyrami- tion from CA1 pyramidal cells and innervate interneur- dal cells (Megias et al., 2001), the amount of GABA re- ons in the hilar area and CA3 region (Sik et al., 1994; A.I. leased in the intact brain during theta may be sufficient Gulya´ s et al., 2001, Soc. Neurosci., abstract). Septally to activate postsynaptic GABAB receptors (Dvorak-Car- projecting interneurons discharge rhythmically during bone and Schuman, 1999; Scanziani, 2000). Neverthe- theta (G. Dragoi and G.B., unpublished observations). less, the in vivo observation that intrasomatic theta has Although small in numbers, backprojection interneurons an amplitude minimum and phase reversal between Ϫ60 with septal collaterals are in a unique position to coordi- and Ϫ75 mV (Soltesz and Desche´ nes, 1993; Ylinen et nate rhythmic discharge of large neuronal populations al., 1995) indicates that the main ion responsible for the in both hippocampus and septum. theta-related membrane hyperpolarization is ClϪ rather In addition to the above three groups, interneurons than Kϩ. with intrinsic oscillatory properties in the theta fre-

“Slow” GABAA receptors have also been implicated quency band may be also important for the maintenance in theta oscillation. According to a recent model (Banks of theta activity. Unfortunately, the exact anatomical et al., 2000), an undisclosed subgroup of interneurons identity of these interneurons is not known (Chapman rhythmically inhibits the apical dendrites of both pyrami- and Lacaille, 1999). They may project back to the MS- dal cells and basket neurons via a slow type of GABAA DBB or may amplify rhythmic inputs from various receptor. In turn, rhythmic suppression of basket cells sources and convey them to the principal cells and other and, consequently, disinhibition of the somata of pyra- interneuron partners. Finally, a small group of interneur- midal cells are hypothesized to give rise to the extracel- ons, termed “antitheta” cells, may also hold a key in lular theta field. An explicit prediction of this model is understanding the emergence of theta oscillations be- that basket cells and pyramidal neurons discharge on cause of their reciprocal firing relationship with all the the opposite phases of theta, as observed under anes- other interneuron classes. Antitheta cells are virtually thesia (Ylinen et al., 1995; Buzsa´ ki et al., 1983; Fox et silent during theta oscillations, but fire rhythmically at al., 1986) and in vitro (Cobb et al., 1995). However, as 15–25 Hz in the absence of theta (Buzsa´ ki et al., 1983; discussed above, pyramidal cells and putative basket Mizumori et al., 1990). Their anatomical identity is un- cells discharge on the same phase of the theta cycle in known. Clarification of the role of various classes of the freely behaving rat (Figure 3; Fox et al., 1986; Skaggs interneurons in theta generation will require recording et al., 1996; Csicsvari et al., 1999). from identified interneurons in the awake animal. In addi- Because current flow through the distal dendrites of tion, reversible “deletion” of the subclasses by molecu- principal cells is the most important event for the genera- lar biological techniques will further aid in understanding tion of extracellular theta currents, inhibitory regulation their contribution. Review 9

Intrinsic Resonant Properties of Neurons over, the phase fluctuation of spikes is not random and Contribute to Theta Oscillations correlates with behavioral variables (O’Keefe and Recce, Not only postsynaptic potentials but also intrinsic con- 1993). Let us first consider the consequences of simulta- ductance changes of the neuronal membrane can con- neous dendritic depolarization and somatic inhibition in tribute to the extracellularly recorded local field. Layer a pyramidal neuron, the conditions that lead to maxi- II stellate cells of the entorhinal cortex have been first mum extracellular current flow. Threshold depolariza- described to possess voltage-dependent oscillatory tion will induce an on the peak of the properties in the theta frequency range (Alonso and Lli- depolarizing phase of the theta cycle (Figure 6). We na´ s, 1989). In these neurons, activation of a persistent assume that this weak but threshold dendritic depolar- Naϩ current may contribute to the membrane potential ization is responsible for the single, sporadic spikes oscillation. Neurons of the MS-DBB are especially prone associated with the local positive peaks of pyramidal to voltage-dependent oscillations. Cholinergic cells dis- layer theta waves (Figure 3C). On the other hand, play bursts of action potentials riding on low-threshold stronger dendritic excitation will bring the neuron to spikes recurring at a theta frequency. GABAergic neu- threshold earlier in the theta cycle. Because of the longer rons, on the other hand, display nonadapting clusters suprathreshold depolarization, more action potentials of spikes interspersed with rhythmic subthreshold mem- will be induced. However, the action potentials are brane-potential oscillations (Alonso et al., 1996; Serafin asymmetrically distributed relative to the peak of intra- et al., 1996). Voltage-dependent oscillations have also cellular depolarization (Figure 6), presumably because ϩ been described in the somata (Leung and Yim, 1991; of Na channel inactivation and/or because of the depo- ϩ Strata, 1998) and dendrites (Figure 5; Kamondi et al., larization-induced activation of a slow K current (Ka- 1998a) of hippocampal pyramidal cells. The presence mondi et al., 1998a). The overall result is a phase ad- of membrane resonance and subthreshold oscillations vancement of action potentials relative to the negative in hippocampal neurons suggests that they are not pas- peak of the mean local field (i.e., the distributed “clock” sive relays of incoming synaptic events, but rather par- theta signal). ticipate in sculpting their final output, regulated by fac- Importantly, O’Keefe and Recce (1993) and Skaggs tors that control their resonant properties. The exact et al. (1996) have observed that the theta phase relation- mechanisms of the intrinsic dendritic oscillation have ship of place cells varies systematically as the rat tra- yet to be revealed. One possibility is that high threshold verses the place field of the recorded unit. As the animal dendritic Ca2ϩ currents amplify a subthreshold somatic moves to the center of the field (as defined by the maxi- oscillation. Another possibility is that activation of NMDA mum discharge rate of the neuron), the first action po- receptors is responsible for the rhythmic events (Peet tential and the median of action potentials can advance Њ et al., 1987). Furthermore, activation of the Ca2ϩ chan- at least 180 during successive theta cycles (O’Keefe nels in distal apical dendrites is also facilitated by ac- and Recce, 1993; Skaggs et al., 1996). Computational tively backpropagating action potentials (Kamondi et models have offered several potential solutions for the al., 1998a; Magee and Johnston, 1995). The type of Ca2ϩ phase precession of action potentials (Tsodyks et al., channels involved in the amplification of theta oscillation 1996; Jensen and Lisman, 1996; Wallenstein and Has- is not known. Distal apical dendrites express a higher selmo, 1997). The potential importance of the spike- density of R- and T-type channels and a lower density phase precession phenomenon is that it may be taken of L- and N-type channels than the soma (Westenbroek as an indication that neurons encode information by et al., 1990; Magee and Johnston, 1995). timing instead of or in addition to firing rate (Ko¨ nig et al., 1996; Hopfield, 1995; Buzsa´ ki and Chrobak, 1995; The role of other voltage-dependent conductancies Lisman and Idiart, 1995). Because the exact timing of has yet to be disclosed in the generation of theta waves. spikes within the theta cycle varies systematically when The transient Kϩ current (I ) is at a markedly higher den- A the rat crosses the place field of the recorded neuron, sity in the dendrites (Hoffman et al., 1997). The density it was suggested that place is encoded by the spike- of hyperpolarization-activated current (I ) increases over h theta phase relationship (Skaggs et al., 1996; Jensen 6-fold from soma to distal dendrites (Magee, 1999). Be- and Lisman, 2000). However, since the spatial position cause I has its largest impact in the subthreshold range h of the rat can be also deduced from the firing rates of of membrane voltage, it may be particularly important hippocampal pyramidal neurons (O’Keefe and Nadel, in theta oscillation. During the theta cycle, the magnitude 1978; Wilson and McNaughton, 1993), an important is- of the membrane potential can change considerably, sue is whether firing rate increase and phase advance- therefore numerous voltage-dependent conductancies ment of spikes are produced by the same or different may be activated sequentially. In turn, activation of these mechanisms. If a single mechanism is responsible for conductancies can exert an important effect on the firing both changes, then space or other behavioral variables patterns of the principal cells. (Eichenbaum, 2000) may be redundantly coded by both discharge frequency and timing of the action potentials. Timing of Action Potentials within the Theta Cycle On the other hand, if firing rate and spike-phase ad- A number of factors affect the precise timing of action vancement can be dissociated experimentally, it sug- potentials in depolarized pyramidal neurons. In the gests that these different mechanisms can code differ- awake animal, REM sleep, and anesthesia, pyramidal ent stimulus dimensions or representations (Hirase et neurons, on average, discharge on the negative phase al., 2000). of the theta cycle recorded from the CA1 pyramidal Large and systematic phase shifts can be brought cell layer (Figure 3A). However, considerable variability about by the interference of two oscillators with slightly exists in the phase of individual spikes (Figure 3B). More- different frequencies (O’Keefe and Recce, 1993; Jensen Neuron 10

Figure 6. Oscillation of Inhibitory Networks Provides a Clock Signal for Timing of Action Potentials of Pyramidal Neurons (A) The interconnected inhibitory interneu- rons rhythmically hyperpolarize spatially dis- tant pyramidal cells. The timing of the action potentials is determined by the magnitude of the excitatory drive to the pyramidal cells: more strongly driven neurons will discharge earlier during the theta cycle. Interneurons discharge regularly at each theta cycle (light blue ticks). (B) Phase advancement of action potentials in the anesthetized rat. Theta oscil- lation was mimicked by sinusoid somatic cur- rent injection (0.05 nA). Place field activation in the awake animal was mimicked by adding a 2 s triangular waveform (gray trace). Bot- tom: a single trial. Histogram : average of 50 events. Dotted lines: peaks of the sinus oscil- lation. Note that most action potentials occur on the rising phase of oscillation and at pro- gressively earlier phases upon increasing lev- els of depolarization. Note also the asymmet- ric distribution of spikes on the depolarizing and repolarizing parts of the ramp, resem- bling the asymmetric nature of place fields (Mehta et al., 2000). Reprinted from Buzsa´ ki and Chrobak (1995); Parts (B) and (D), A. Henze and G.B., unpublished data.

and Lisman, 1996). As discussed earlier, the extrahippo- peak of theta waves in str. oriens/pyramidale (Figures campal (entorhinal) and intrahippocampal (CA3) theta 1 and 4). It is possible that the low-frequency single oscillators can be relatively independent. If the average spikes that occur on the positive peaks of the locally field theta is determined primarily by the entorhinal input derived theta (Figure 3B) are brought about by the ento- but place-related discharge requires the CA3 input as rhinal input. On the other hand, stronger activation may well, then a faster CA3 theta oscillator can bring about require the cooperation of entorhinal and CA3 inputs a full theta cycle spike advancement. The observation and the more strongly activated neurons will discharge that CA3 and CA1 pyramidal neurons discharge on the earlier in phase. Place cells discharge fastest in the same phase of theta (Fox et al., 1986; J. Csicsvari and center of the field that corresponds to approximately G.B., unpublished data) provides support for the hypoth- 180Њ of phase advancement (Skaggs et al., 1996). esis that the CA3 input is the major driving source of Because Ͻ3% of CA1 pyramidal cells are active during the discharging CA1 pyramidal neurons. the theta cycle (Csicsvari et al., 1998) and because the The activity-dependent theta phase advancement of majority of spikes are emitted within the place field, the spike may shed light on the controversy between field mean spike-theta phase relationship of the cell popula- measurements of theta waves and the timing of spikes tion (Figure 3A) is therefore strongly biased by the during the theta cycle. The entorhinal input appears to phase-advanced place cells. The slowly discharging be relatively sustained (Brankack et al., 1993) because majority, in fact, may have an opposite phase relation- layer III neurons are active over very large spatial areas ship (Figure 3C). These observations suggest that while (Barnes et al., 1990). Although most pyramidal neurons most neurons contribute to the generation of field theta, are silent during theta (Harris et al., 2000), the voltage it is the strongly active minority that is responsible for fluctuation in their membrane, nevertheless, also con- the empirically derived spike-phase relationship (Figure tributes to theta. The maximum dendritic depolarization 3A). In turn, this relationship could resolve the contro- corresponds to the negative peak of theta waves present versy between results derived from CSD unit recording in str. lacunosum-moleculare, mirrored by the positive studies regarding the relationship between afferent exci- Review 11

tation and the theta cycle-related discharges of CA1 be responsible. When theta-like oscillation is induced pyramidal cells. in the slice by carbachol, part of the effect may be

Overall, these considerations suggest that modeling explained by the drug’s blocking effect of GABAA recep- the generation of the extracellular theta field with a single tors (MacVicar and Tse, 1989). Importantly, carbachol “average” pyramidal cell (Figure 2) is not adequate. The has been shown to facilitate the somadendritic propaga- timing of the action potentials within the theta cycle tion of action potentials through activation of muscarinic depends on the activity level of pyramidal cell and other receptors (Tsubokawa and Ross, 1997). Backpropagat- hitherto unknown factors. We suggest that the behavior- ing spikes into the synaptically activated dendritic seg- dependent variability of spike timing may explain the ments, in turn, have been shown to be important for the discrepancy between the empirical observations (Figure enhancement of synaptic weights (Magee and John- 3) and the expected timing of the spikes on the basis ston, 1995). LTP may occur because large amplitude ϩ of extracellular theta sinks and sources (Figure 4). fast spikes in dendrites can serve as a trigger of Ca2 spikes (Kamondi et al., 1998b). In turn, the large depolar- 2ϩ Hippocampal Plasticity during Theta Oscillations ization associated with Ca spikes can enable NMDA Although indirect, several observations suggest the pos- receptors. In vivo, theta-associated somatic hyperpolar- sible involvement of theta oscillations in synaptic plas- ization may provide silent periods necessary for the oc- ticity. A number of in vitro and in vivo studies have currence of complex spike bursts (Harris et al., 2001), reported that induction of long-term potentiation (LTP) a condition necessary for synaptic potentiation (Paulsen is optimal when the time interval between stimuli is ap- and Sejnowski, 2000). Thus, theta oscillation may pro- proximately 200 ms (Larson and Lynch, 1986; Greenstein vide a mechanism for bringing together in time afferent- et al., 1988). Rhythmic stimulation at theta frequency is induced depolarization of pyramidal cell dendrites and not needed though and two high-frequency bursts are dendritic invasion of fast spikes, the key elements for sufficient. When two identical bursts are applied repeat- the induction of synaptic plasticity. Assuming that these edly to two different inputs at 200 ms intervals, only the observations in vitro and in the anesthetized rat can be synapses activated by the second burst show potentia- generalized to the brains of behaving animals, we can tion (Larson and Lynch, 1986). In fact, the priming stimu- assign a specific physiological role to the intrahippo- lus can be a single pulse (Rose and Dunwiddie, 1986). campal trisynaptic pathway. During theta oscillation, re- The mechanism underlying the priming effect is largely peated pairing of distal dendritic depolarization by the entorhinal input and the trisynaptically activated CA3 unknown. It may block a normally present outward cur- recurrent/Schaffer collaterals to CA3 and CA1 pyramidal rent (e.g., I ; Hoffman et al., 1997) or activate an NMDA A cells can result in synaptic modification of the intrahip- receptor-mediated inward current (Larson and Lynch, pocampal associational pathways. As a result, these 1986). Alternatively, the priming stimulus-induced hy- modified pathways will give rise to endogenous popula- perpolarization may activate I conductance or remove h tion patterns in the absence of the entorhinal inputs the inactivation of T-type of Ca2ϩ channels followed by during non-theta activity. Reactivation of the same syn- a rebound depolarization after 100–200 ms (Cobb et al., apses can further strengthen their efficacy (Buzsa´ ki, 1995). A third possibility is that GABA released from 1989). The synaptic modification, brought about by the terminals or cannabiods released from the discharging two-step process, in turn, will allow memory retrieval. neuron in response to the first pulse may block further When repeated trains are delivered on the negative GABA release by the subsequent train (Wilson and Ni- peak of the theta recorded in the str. radiatum, the pre- coll, 2001; Ha´ jos et al., 2000). Thus, the priming stimulus viously potentiated pathway becomes depotentiated prepares the network so that all neurons will be optimally (Huerta and Lisman, 1996; Ho¨ schler et al., 1997). There- depolarized by the burst stimulus. However, it is not fore, timing of the dendritic excitatory inputs during the clear yet how these postulated events relate to the con- theta cycle is critical for the strengthening and weaken- ductance changes present during the in vivo theta cycle. ing of synapses. As discussed earlier, action potentials The priming stimulus may be substituted by a previous occur on progressively earlier phases of the theta cycle theta cycle in oscillating hippocampal networks. When as the rat traverses the place field of the recorded unit a single tetanic train (four pulses at 100 Hz) is delivered (O’Keefe and Recce, 1993; Skaggs et al., 1996). Thus, if during the positive peak of carbachol-induced theta- theta phase-dependent potentiation and depotentiation like waves in CA1 str. radiatum, LTP occurs (Huerta and also hold in the intact brain (Ho¨ schler et al., 1997), one Lisman, 1996). Importantly, a single burst (200 Hz) of might expect that the synapses which bring about place- five pulses was also sufficient for the potentiation of related activity undergo both potentiation and depoten- the Schaffer collateral response in the anesthetized rat tiation every time the rat visits the place field of the (Ho¨ schler et al., 1997). Because the positive peak of neuron. Alternatively, pyramidal cells representing the str. radiatum theta (i.e., negative phase of theta in the same part of the environment may recurrently discharge pyramidal layer) corresponds to the highest probability their common basket and chandelier neurons (Csicsvari of pyramidal cell discharge (Fox et al., 1986; Buzsa´ ki et et al., 1998); thus dendritic depolarization and somatic al., 1983), the phase-locked discharge of neurons may hyperpolarization will still occur in phase for this subnet. substitute for the priming pulse. Similar observations As discussed in the previous section, this hypothesis were made also in the dentate gyrus, although the rela- assumes that the activated subnet of pyramidal cells tionship between the phase of theta and neuronal excit- and interneurons (i.e., a cell assembly) can “step out” ability is less clear in this region (Pavlides et al., 1988). from the theta cycle represented by the nonspiking or Why is potentiation so efficient during theta? Various slowly discharging population (Marshall et al., 2002). factors that act in concert during the theta cycle may We hypothesize that an important function of the theta Neuron 12

oscillation is to assemble and segregate neuronal large volume of neurons, active neurons may contribute groups. The segregated assembles then can “wire to- disproportionally more to the field than nonspiking cells. gether” and exert a differential effect on their down- In future studies, large-scale recordings spanning stream neurons. across different hippocampal regions should be used to reveal whether subnets of active neurons behave Significance of Theta Oscillations differently from the “average” population. The coherent and Future Directions oscillations of cell assemblies during theta provide an Brain waves reflect the collective behavior of neurons ideal mechanism for temporal coding and decoding (Lis- and provide insight into the fast periodic changes of man and Idiart, 1995; Wallenstein and Hasselmo, 1997). the network. Theta oscillations reliably correlate with a Theta phase precession of action potentials in activated variety of behaviors. The theta versus non-theta dichot- neurons supports the idea that phase coding is ex- omy objectively groups behaviors into preparatory ver- ploited by the hippocampus and associated structures. sus consummatory classes (Sherrington, 1897). However, further research is required to determine Field theta waves and phase-locked discharge of neu- whether phase coding and rate coding are redundant rons are observed in a large number of brain structures. or whether they are used to register different types of These structures are members of the anatomically de- information. On the basis of the efficient and tunable fined limbic system (Broca, 1878). Thus, theta oscillation pyramidal cell-interneuron feedback (Markram et al., can be used to functionally determine parts of an ana- 1998; Marshall et al., 2002), we hypothesize that this tomical macrosystem. Consequently, “limbic theta os- mechanism allows active pyramidal cells and their com- cillation” is a more appropriate term than hippocampal mon interneurons to beat at slightly different frequency theta. A theta cycle may be considered as an information and phase than the “master clock” rhythm of theta. As quantum, allowing the exchange of information among a result, the phase differences can segregate assembl- the linked members in a phase-locked manner. This ies of neurons that are assigned to different representa- discontiguous (cyclic) mode of operation may be a tions. unique solution to temporarily segregate and link neu- Addressing many of these issues requires the analyti- ronal assemblies to perform various operations. From cal power of in vitro preparations. The dynamic relation- this perspective, theta oscillation is a computational ship between pyramidal cell discharge and the postsyn- process that brings together activity of sensory- and/or aptic response of the various interneuron classes should memory-activated neurons in time, thereby affecting be- be worked out in detail. On the other hand, a coordinated havioral output and plasticity. Examples of these opera- effort is needed to compare the physiological relevance tions include phase reset of the theta cycle by sensory of the in vitro observations to the intact brain. In addition stimulation (Buzsa´ ki et al., 1979; Givens, 1996; Tesche to physiological experiments, several topics could bene- and Karhu, 2000), theta phase locking of motor activity fit from computational models. For example, modeling (Bun˜ o and Velluti, 1977; Semba and Komisaruk, 1978), extracellular current flow using networks of neurons may memory encoding and retrieval (Hasselmo et al., 2002), yield insights into the relative contribution of spiking and synaptic potentiation of sequentially activated place and nonspiking neurons to local field potentials. Genetic neurons (Mehta et al., 2000). modification of channels and receptors can confront the Because neurons in the entorhinal cortex, hippocam- relationship between single cell and network properties. pus, and septum are endowed with numerous voltage- Theta oscillation is a physiologically relevant phenotype dependent channels and intrinsic resonant properties, that provides a sensitive assay for the measurement of sequential activation of voltage-dependent conduc- ensemble properties of neurons. tancies during the theta cycle may set constraints for Most known disease-related impairments of the hip- excitability and plasticity. For example, artificial stimula- pocampus are associated with structural damage of its tion of any hippocampal pathway results in either ac- neurons. Because theta oscillation is a major operational commodation or augmentation of the response. In con- mode of the hippocampus, it is expected that functional trast, physiological activation of the same inputs during impairment of its oscillatory mode may be equally detri- theta activity may maintain stable firing patterns for pro- mental to mental health. Noninvasive, clinically applica- longed periods (Czurko´ et al., 1999). Exploration of the ble methods (Tesche and Karhu, 2000) are now available significance of the cyclic conductance changes in future to examine whether and how this population pattern is experiments is a necessary step to provide an insight affected in schizophrenia, epilepsy, Alzheimer’s dis- into the physiological role of theta. During theta oscilla- ease, and other diseases. tion, the dendrites of pyramidal cells can depolarize to Ϫ45 mV, thus potentially activating/inactivating a host Acknowledgments of voltage-gated channels. The sequential activation I thank J. Csicsvari, G. Dragoi, T.F. Freund, A. Gulya´ s, D. Henze, K. and inactivation of the numerous conductancies in pyra- Harris, M. Hasselmo, L.S. Leung, J. Lisman, L. Marshall, A. Sirota, midal cells and interneurons may offer clues to the physi- R.D. Traub, and X.-J. Wang for their valuable comments on the ological significance of theta oscillation in synaptic manuscript. This work was supported by NIH (NS34994, MH54671), transmission and plasticity. These issues can be ex- Fogarty Senior Fellowship, and the Kirby Foundation. plored in vitro using sinusoid stimulation of the soma References and dendrites at theta frequency (Kamondi et al., 1998a; Magee, 2001). Adey, W.R. (1967). Hippocampal states and functional relations with Although the extracellularly recorded field reflects the corticosubcortical systems in attention and learning. Prog. Brain summed activity of membrane currents over a relatively Res. 27, 228–245. Review 13

Alonso, A., and Garcia-Austt, E. (1987). Neuronal sources of theta Bun˜ o, W., Jr., and Velluti, J.C. (1977). Relationships of hippocampal rhythm in the entohrinal cortex of the rat. I. Laminar distribution of theta cycles with bar pressing during self-stimulation. Physiol. Be- theta field potentials. Exp. Brain Res. 67, 493–501. hav. 19, 615–621. Alonso, A., and Kohler, C. (1982). Evidence for separate projections Buzsa´ ki, G. (1989). Two-stage model of memory trace formation: a of hippocampal pyramidal and non-pyramidal neurons to different role for “noisy” brain states. Neuroscience 31, 551–570. parts of the septum in the rat brain. Neurosci. Lett. 31, 209–214. Buzsa´ ki, G., and Chrobak, J.J. (1995). Temporal structure in spatially Alonso, A., and Llina´ s, R.R. (1989). Subthreshold Naϩ-dependent organized neuronal ensembles: a role for interneuronal networks. theta-like rhythmicity in stellate cells of entorhinal cortex layer II. Curr. Opin. Neurobiol. 5, 504–510. Nature 342, 175–177. Buzsa´ ki, G., Grastya´ n, E., Tveritskaya, I.N., and Czopf, J. (1979). Alonso, A., Khateb, A., Fort, P., Jones, B.E., and Muhlethaler, M. Hippocampal evoked potentials and EEG changes during classical (1996). Differential oscillatory properties of cholinergic and noncho- conditioning in the rat. Electroencephalogr. Clin. Neurophysiol. 47, linergic nucleus neurons in guinea pig brain slice. J. Neurosci. 8, 64–74. 169–182. Buzsa´ ki, G., Leung, L., and Vanderwolf, C.H. (1983). Cellular bases Amaral, D.G., and Witter, M.P. (1989). The three-dimensional organi- of hippocampal EEG in the behaving rat. Brain Res. Bran Res. Rev. zation of the hippocampal formation: a review of anatomical data. 6, 139–171. Neuroscience 31, 571–591. Buzsa´ ki, G., Czopf, J., Kondakor, I., and Kellenyi, L. (1986). Laminar Arnolds, D.E., Lopes da Silva, F.H., Aitink, J.W., Kamp, A., and Boei- distribution of hippocampal rhythmic slow activity (RSA) in the be- jinga, P. (1980). The spectral properties of hippocampal EEG related having rat: Current source density analysis, effects of urethane and to behaviour in man. Electroencephalogr. Clin. Neurophysiol. 50, atropine. Brain Res. 365, 125–137. 324–328. Buzsa´ ki, G., Gage, F.H., Kellenyi, L., and Bjorklund, A. (1987). Behav- Artemenko, D.P. (1972). Role of hippocampal neurons in theta-wave ior dependence of the electrical activity of intracerebrally trans- generation. Neurophysiologia 4, 409–415. planted fetal hippocampus. Brain Res. 400, 334–347. Banks, M.I., White, J.A., and Pearce, R.A. (2000). Interactions be- Buzsa´ ki, G., Freund, T.F., Bayardo, F., Leon, A., and Somogyi, P. tween distinct GABA(A) circuits in hippocampus. Neuron 25, (1990). Ischemia-induced changes in the electrical activity of the 449–457. hippocampus. Exp. Brain Res. 78, 268–278. Barnes, C.A., McNaughton, B.L., Mizumori, S.J., Leonard, B.W., and Buzsa´ ki, G., Bragin, A., Chrobak, J.J., Na´ dasdy, Z., Sik, A., Hsu, M., Lin, L.H. (1990). Comparison of spatial and temporal characteristics and Ylinen, A. (1994). Oscillatory and intermittent synchrony in the of neuronal activity in sequential stages of hippocampal processing. hippocampus: relevance to memory trace formation. In Temporal Prog. Brain Res. 83, 287–300. Coding in the Brain, G. Buzsa´ ki, R. Llina´ s, W. Singer, A. Bethoz, and Bland, B.H. (1986). Physiology and pharmacology of hippocampal Y. Christen, eds. (Berlin: Springer-Verlag), pp.145–172. formation theta rhythms. Prog. Neurobiol. 26, 1–54. Buzsa´ ki, G., Penttonen, M., Na´ dasdy, Z., and Bragin, A. (1996). Pat- Bland, B.H., Andersen, P., and Ganes, T. (1975). Two generators of tern and inhibition-dependent invasion of pyramidal cell dendrites hippocampal theta activity in rabbits. Exp. Brain Res. 94, 199–218. by fast spikes in the hippocampus in vivo. Proc. Natl. Acad. Sci. Bland, B.H., Colom, L.V., Konopacki, J., and Roth, S.H. (1988). Intra- USA 93, 9921–9925. cellular records of carbachol-induced theta rhythm in hippocampal Chapman, C.A., and Lacaille, J.C. (1999). Cholinergic induction of slices. Brain Res. 447, 364–368. theta-frequency oscillations in hippocampal inhibitory interneurons Blasco-Iban˜ ez, J.M., and Freund, T.F. (1995). Synaptic input of hori- and pacing of pyramidal cell firing. J. Neurosci. 19, 8637–8645. zontal interneurons in stratum oriens of the hippocampal CA1 sub- Cobb, S.R., Buhl, E.H., Halasy, K., Paulsen, O., and Somogyi, P. field: structural basis of feed-back activation. Eur. J. Neurosci. 7, (1995). Synchronization of neuronal activity in hippocampus by indi- 2170–2180. vidual GABAergic interneurons. Nature 378, 75–98. Bo´ dizs, R., Ka´ ntor, S., Szu¨ cs, A., Ero¨ ss, L., and Hala´ sz, P. (2001). Cole, A.E., and Nicoll, R.A. (1983). Acetylcholine mediates a slow Rhythmic hippocampal slow oscillation characterizes REM sleep in synaptic potential in hippocampal pyramidal cells. Science 221, humans. Hippocampus 11, 747–753. 1299–1301. Borhegyi, Z., and Freund, T.F. (1998). Dual projection from the medial Csicsvari, J., Hirase, H., Czurko´ , A., and Buzsa´ ki, G. (1998). Reliability septum to the supramammillary nucleus in the rat. Brain Res. Bull. and state dependence of pyramidal-interneuron synapses in the 46, 453–459. hippocampus: an ensemble approach in the behaving rat. Neuron Borisyuk, R., and Hoppensteadt, F. (1999). Oscillatory models of the 21, 179–189. hippocampus: a study of spatio-temporal patterns of neural activity. Csicsvari, J., Hirase, H., Czurko´ , A., Mamiya, A., and Buzsa´ ki, G. Biol. Cybern. 81, 359–371. (1999). Oscillatory coupling of hippocampal pyramidal cells and in- Bragin, A., Jando, G., Na´ dasdy, Z., Hetke, J., Wise, K., and Buzsa´ ki, terneurons in the behaving rat. J. Neurosci. 19, 274–287. G. (1995). Gamma frequency (40–100 Hz) patterns in the hippocam- Czurko´ , A., Hirase, H., Csicsvari, J., and Buzsa´ ki, G. (1999). Sus- pus of the behaving rat. J. Neurosci. 15, 47–60. tained activation of hippocampal pyramidal cells by “space clamp- Brankack, J., Stewart, M., and Fox, S.E. (1993). Current source den- ing” in a running wheel. Eur. J. Neurosci. 11, 344–353. sity analysis of the hippocampal theta rhythm: associated sustained Denham, M.J., and Borisyuk, R.M. (2000). A model of theta rhythm potentials and candidate synaptic generators. Brain Res. 615, production in the septal-hippocampal system and its modulation 310–327. by ascending brain stem pathways. Hippocampus 10, 698–716. Brazhnik, E.S., and Fox, S.E. (1999). Action potentials and relations Dvorak-Carbone, H., and Schuman, E.M. (1999). Patterned activity to the theta rhythm of medial septal neurons in vivo. Exp. Brain Res. in stratum lacunosum moleculare inhibits CA1 pyramidal firing. Neu- 127, 244–258. rophysiol. 82, 3213–3222. Broca, P. (1878). Anatomie comparee´ circonvolutions ce´ re´ brales; le grand lobe limbique et la scissure limbiquie dans la se´ rie des Eichenbaum, H. (2000). Hippocampus: mapping or memory? Curr. mammife´ res. Rev. Antropol. Series 2. 1, 385–498. Biol. 10, R785–R797. Buhl, E.H., Halasy, K., and Somogyi, P. (1994). Diverse sources Fellous, J.M., and Sejnowski, T.J. (2000). Cholinergic induction of of hippocampal unitary inhibitory postsynaptic potentials and the oscillations in the hippocampal slice in the slow (0.5–2 Hz), theta number of synaptic release sites. Nature 368, 823–828. (5–12 Hz), and gamma (35–70 Hz) bands. Hippocampus 10, 187–197. Bullock, T.H., Buzsa´ ki, G., and McClune, M.C. (1990). Coherence Fischer, Y., Gahwiler, B.H., and Thompson, S.M. (1999). Activation of compound field potentials reveals discontinuities in the CA1- of intrinsic hippocampal theta oscillations by acetylcholine in rat subiculum of the hippocampus in freely-moving rats. Neuroscience septo-hippocampal co-cultures. J. Physiol. (Lond.) 519, 405–413. 38, 609–619. Fox, S.E. (1989). Membrane potential and impedance changes in Neuron 14

hippocampal pyramidal cells during theta rhythm. Exp. Brain Res. Ho¨ schler, C., Anwyl, R., and Rowan, M.J. (1997). Stimulation on 77, 283–294. the positive phase of hippocampal theta rhythm induces long-term Fox, S.E., Wolfson, S., and Ranck, J.B., Jr. (1986). Hippocampal potentiation that can be depotentiated by stimulation on the nega- theta rhythm and the firing of neurons in walking and urethane tive phase in area CA1 in vivo. J. Neurosci. 17, 6470–6477. anesthetized rats. Exp. Brain Res. 62, 495–508. Huerta, P.T., and Lisman, J.E. (1996). Bidirectional synaptic plasticity Freund, T.F., and Antal, M. (1988). GABA-containing neurons in the induced by a single burst during cholinergic theta oscillation in CA1 septum control inhibitory interneurons in the hippocampus. Nature in vitro. Neuron 15, 1053–1063. 336, 170–173. Jensen, O., and Lisman, J.E. (1996). Hippocampal CA3 region pre- Freund, T.F., and Buzsa´ ki, G. (1996). Interneurons of the hippocam- dicts memory sequences: accounting for the phase precession of pus. Hippocampus 6, 347–470. place cells. Learn. Mem. 3, 279–287. Givens, B. (1996). Stimulus-evoked resetting of the dentate theta Jensen, O., and Lisman, J.E. (2000). Position reconstruction from rhythm: relation to working memory. Neuroreport 8, 159–163. an ensemble of hippocampal place cells: contribution of theta phase coding. J. Neurophysiol. 83, 2602–2609. Grastya´ n, E., Lissa´ k, K., Madara´ sz, I., and Donhoffer, H. (1959). Hippocampal electrical activity during the development of condi- Ji, D.Y., and Dani, J.A. (2000). Inhibition and disinhibition of pyrami- tioned reflexes. Electroencephal. Clin. Neurophysiol. 11, 409–430. dal neurons by activation of nicotinic receptors on hippocampal interneurons. J. Neurophysiol. 83, 2682–2690. Greenstein, Y.J., Pavlides, C., and Winson, J. (1988). Long-term potentiation in the dentate gyrus is preferentially induced at theta Jouvet, M. (1969). Biogenic amines and the states of sleep. Science periodicity. Brain Res. 438, 331–334. 163, 32–41. Ha´ jos, N., Papp, E.C., Acsa´ dy, L., Levey, A.I., and Freund, T.F. (1998). Kahana, M.J., Sekuler, R., Caplan, J.B., Kirschen, M., and Madsen, Distinct interneuron types express m2 muscarinic receptor immuno- J.R. (1999). Human theta oscillations exhibit task dependence during reactivity their dendrites or axon terminals in the hippocampus. virtual maze. J. Neurosci. 399, 781–784. Neuroscience 82, 355–376. Kamondi, A., Acsa´ dy, L., Wang, X.-J., and Buzsa´ ki, G. (1998a). Theta Ha´ jos, N., Katona, I., Naiem, S.S., MacKie, K., Ledent, C., Mody, I., oscillations in somata and dendrites of hippocampal pyramidal cells and Freund, T.F. (2000). Cannabinoids inhibit hippocampal GABA- in vivo: activity dependent phase-precession of action potentials. ergic transmission and network oscillations. Eur. J. Neurosci. 12, Hippocampus 8, 244–261. 3239–3249. Kamondi, A., Acsa´ dy, L., and Buzsa´ ki, G. (1998b). Dendritic spikes Halasy, K., and Somogyi, P. (1993). Subdivisions in the multiple are enhanced by cooperative network activity in the intact hippo- GABAergic innervation of granule cells in the dentate gyrus of the campus. J. Neurosci. 18, 3919–3928. rat hippocampus. J. Neurosci. 5, 411–429. Kandel, E.R., and Squire, L.R. (2000). Neuroscience: breaking down Harris, K., Henze, D.A., Hirase, H., Csicsvari, J., and Buzsa´ ki, G. scientific barriers to the study of brain and mind. Science 290, 1113– (2000). The accuracy of tetrode spike separation as determined 1120. by simultaneous intracellular and extracellular measurements. J. Katona, I., Acsa´ dy, L., and Freund, T.F. (1999). Postsynaptic targets Neurophysiol. 84, 401–414. of somatostatin-immunoreactive interneurons in the rat hippocam- Harris, K., Hirase, H., Leinekugel, X., Henze, D.A., and Buzsa´ ki, G. pus. Neuroscience 88, 37–55. (2001). Temporal interaction between single spikes and complex Kentros, C., Hargreaves, E., Hawkins, R.D., Kandel, E.R., Shapiro, spike bursts in hippocampal pyramidal cells. Neuron 32, 141–149. M., and Muller, R.V. (1998). Abolition of long-term stability of new Hasselmo, M.E., and Fehlau, B.P. (2001). Differences in time course hippocampal place cell maps by NMDA receptor blockade. Science of cholinergic and GABAergic modulation of excitatory synaptic 280, 2121–2126. potentials in rat hippocampal slice preparations. J. Neurophysiol. King, C., Recce, M., and O’Keefe, J. (1998). The rhythmicity of cells 86, 1792–1802. of the medial septum/diagonal band of Broca in the freely moving Hasselmo, M.E., and Schnell, E. (1994). Laminar selectivity of the rat: relationships with behaviour and hippocampal theta. Eur. J. cholinergic suppression of synaptic transmission in rat hippocampal Neurosci. 10, 464–477. region CA1: computational modeling and brain slice physiology. J. Kocsis, B., Bragin, A., and Buzsa´ ki, G. (1999). Interdependence of Neurosci. 14, 3898–3914. multiple theta generators in the hippocampus: a partial coherence Hasselmo, M.E., Bodelon, C., and Wyble, B.P. (2002). A proposed analysis. J. Neurosci. 19, 6200–6212. function for hippocampal theta rhythm: Separate phases of encod- Ko¨ nig, P., Engel, A.K., and Singer, W. (1996). Integrator or coinci- ing and retrieval enhance reversal of prior learning. Neural Comput., dence detector? The role of the cortical neuron revisited. Trends in press. Neurosci. 19, 130–137. Ha¨ usser, M., Spruston, N., and Stuart, G.J. (2000). Diversity and Konopacki, J., MacIver, M.B., Bland, B.H., and Roth, S.H. (1987). dynamics of dendritic signaling. Science 290, 739–744. Theta in hippocampal slices: relation to synaptic responses of den- Henze, D.A., Urban, N.N., and Barrionuevo, G. (2000). The multifari- tate neurons. Brain Res. Bull. 18, 25–27. ous hippocampal mossy fiber pathway: a review. Neuroscience 98, Kovalchuk, Y., Eilers, J., Lisman, J., and Konnerth, A. (2000). NMDA 407–427. receptor-mediated subthreshold Ca(2ϩ) signals in spines of hippo- Hirase, H., Czurko´ , A., Csicsvari, J., and Buzsa´ ki, G. (2000). Firing campal neurons. J. Neurosci. 20, 1791–1799. rate and theta-phase coding by hippocampal pyramidal neurons Kramis, R., Vanderwolf, C.H., and Bland, B.H. (1975). Two types of during “space clamping.“ Eur. J. Neurosci. 11, 4373–4380. hippocampal rhythmical slow activity in both the rabbit and the rat: Hoffman, D.A., Magee, J.C., Colbert, C.M., and Johnston, D. (1997). relations to behavior and effects of atropine, diethyl ether, urethane, Kϩ channel regulation of signal propagation in dendrites of hippo- and pentobarbital. Exp. Neurol. 49, 58–85. campal pyramidal neurons. Nature 387, 869–875. Larson, J., and Lynch, G. (1986). Induction of synaptic potentiation Holsheimer, J., Boer, J., Lopes da Silva, F., and Van Rotterdam, A. in hippocampus by patterned stimulation two events. Science 232, (1982). The double dipole model of theta rhythm generation: simula- 985–988. tion of laminar field potential profiles in dorsal hippocampus of rat. Leanza, G., Nilsson, O.G., and Bjorklund, A. (1993). Functional activ- Brain Res. 235, 31–50. ity of intrahippocampal septal grafts is regulated by host afferents Hopfield, J.J. (1995). Pattern recognition computation using action as studied by microdialysis of acetylcholine. Brain Res. 618, 47–56. potential timing for stimulus representation. Nature 376, 33–36. Lee, M.G., Chrobak, J.J., Sik, A., Wiley, R.G., and Buzsa´ ki, G. (1994). Horvath, Z., Kamondi, A., Czopf, J., Bliss, T.V.P., and Buzsa´ ki, G. Hippocampal theta following selective lesion of the septal choliner- (1988). NMDA receptors may be involved in generation of hippocam- gic system. Neuroscience 62, 1033–1047. pal theta rhythm. In Synaptic Plasticity in the Hippocampus, H.L. Leranth, C., Carpi, D., Buzsa´ ki, G., and Kiss, J. (1999). The entorhino- Haas and G. Buzsa´ ki, eds. (Berlin: Springer-Verlag), p. 45. septo-supramammillary nucleus connection in the rat: morphologi- Review 15

cal basis of a feedback mechanism regulating hippocampal theta release of endogenous GABA and glutamate from the cerebral cor- rhythm. Neuroscience 88, 701–718. tex in the rat. Naunun-Schmiedeberg’s Arch. Pharmacol. 316, Leung, L.S. (1984). Model of gradual phase shift of theta rhythm in 235–239. the rat. J. Neurophysiol. 52, 1051–1065. Moruzzi, G., and Magoun, H.W. (1949). Brainstem reticular formation Leung, L.W., and Borst, J.G. (1987). Electrical activity of the cingulate and activation of the EEG. Electroencephal. Clin. Neurophysiol. 1, cortex. I. Generating mechanisms and relations to behavior. Brain 455–473. Res. 407, 68–80. O’Keefe, J., and Nadel, L. (1978). The hippocampus as a Cognitive Leung, L.S., and Yim, C.Y. (1986). Intracellular records of theta Map (Clarendon, Oxford). rhythm in hippocampal CA1 cells of the rat. Brain Res. 367, 323–327. O’Keefe, J., and Recce, M.L. (1993). Phase relationship between Leung, L.S., and Yim, C.Y. (1991). Intrinsic membrane potential oscil- hippocampal place units and the EEG theta rhythm. Hippocampus lations in hippocampal neurons in vitro. Brain Res. 553, 261–274. 3, 317–330. Lisman, J.E., and Idiart, M.A. (1995). Storage of 7 Ϯ2 short-term Otmakhova, N.A., and Lisman, J.E. (2000). Dopamine, serotonin, and memories in oscillatory subcycles. Science 267, 1512–1515. noradrenaline strongly inhibit the direct perforant path-CA1 synaptic Llina´ s, R.R. (1988). The intrinsic electrophysiological properties of input, but have little effect on the Schaffer collateral input. Ann. N mammalian neurons: insights into central nervous system function. Y Acad. Sci. 911, 462–464. Science 242, 1654–1664. Pare´ , D., and Collins, D.R. (2000). Neuronal correlates of fear in the Lopes da Silva, F.H., Witter, M., Boeijinga, P.H., and Lohman, A. lateral amygdala: multiple extracellular recordings in conscious cats. (1990). Anatomic organization and physiology of the limbic cortex. J. Neurosci. 20, 2701–2710. Physiol. Rev. 70, 453–511. Paulsen, O., and Sejnowski, T.J. (2000). Natural patterns of activity Maccaferri, G., Roberts, J.D., Szu¨ cs, P., Cottingham, C.A., and So- and long-term synaptic plasticity. Curr. Opin. Neurobiol. 10, mogyi, P. (2000). Cell surface domain specific postsynaptic currents 172–179. evoked by identified GABAergic neurones in rat hippocampus in Pavlides, C., Greenstein, Y.J., Grudman, M., and Winson, J. (1988). vitro. J. Physiol. 524, 91–116. Long-term potentiation in the dentate gyrus is induced preferentially MacVicar, B.A., and Tse, F.W. (1989). Local neuronal circuitry under- on the positive phase of theta rhythm. Brain Res. 439, 383–387. lying cholinergic rhythmical slow activity in CA3 area of rat hippo- Peet, M.J., Curry, K., Magnuson, D.S., and McLennan, H. (1987). The campal slices. J. Physiol. 417, 197–212. N-methyl-D-aspartate receptor and burst firing of CA1 hippocampal Madison, D.V., Lancaster, B., and Nicoll, R.A. (1987). Voltage clamp pyramidal neurons. Neuroscience 22, 563–571. analysis of cholinergic action in the hippocampus. Neuroscience 7, Petsche, H., Stumpf, C., and Gogola´ k, G. (1962). The significance 733–741. of the rabbit’s septum as a relay station between midbrain and the Magee, J.C. (1999). Dendritic Ih normalizes temporal summation in hippocampus: I. The control of hippocampus arousal activity by the hippocampal CA1 neurons. Nat. Neurosci. 2, 508–514. septum cells. Electroencephalogr. Clin. Neurophysiol. 14, 202–211. Magee, J.C. (2001). Dendritic mechanisms of phase precession in Pitler, T.A., and Alger, B.E. (1992). Cholinergic excitation of GABA- hippocampal CA1 pyramidal neurons. J. Neurophysiol. 86, 528–532. ergic interneurons in the rat hippocampal. J. Physiol. 450, 127–142. Magee, J.C., and Johnston, D. (1995). Synaptic activation of voltage- Pyapali, G.K., Sik, A., Penttonen, M., Buzsa´ ki, G., and Turner, D.A. gated channels in the dendrites of hippocampal pyramidal neurons. (1998). Dendritic properties of hippocampal CA1 pyramidal neurons Science 268, 301–304. in the rat: intracellular staining in vivo and in vitro. J. Comp. Neurol. 391, 335–352. Markram, H., Gupta, A., Uziel, A., Wang, Y., and Tsodyks, M. (1998). Information processing with frequency-dependent synaptic con- Raghavachari, S., Kahana, M.J., Rizzuto, D.S., Caplan, J.B., nections. Neurobiol. Learn. Mem. 70, 101–112. Kirschen, M.P., Bourgeois, B., Madsen, J.R., and Lisman, J.E. (2001). Gating of human theta oscillations by a working memory task. J. Marshall, L., Hirase, H., Leinekugel, X., Dragoi, G., Henze, D.A., and Neurosci. 21, 3175–3183. Buzsa´ ki, G. (2002). Hippocampal pyramidal cell-interneuron spike transmission is reliable and responsible for place modulation of Rose, G.M., and Dunwiddie, T.V. (1986). Induction of hippocampal interneuron discharge. J. Neurosci. 22, RC197. long-term potentiation using physiologically patterned stimulation. Neurosci. Lett. 69, 244–248. McMahon, L.L., Williams, J.H., and Kauer, J.A. (1998). Functionally distinct groups of interneurons identified during rhythmic oscilla- Scanziani, M. (2000). GABA spillover activates postsynaptic tions in hippocampus in vitro. J. Neurosci. 18, 5640–5651. GABA(B) receptors to control rhythmic hippocampal activity. Neuron 25, 673–681. Megias, M., Emri, Z., Freund, T.F., and Gulya´ s, A.I. (2001). Total number and distribution of inhibitory and excitatory synapses on Segal, M., Bjo¨ rklund, A., and Gage, F.H. (1985). Transplanted septal hippocampal CA1 pyramidal cells. Neuroscience 102, 527–540. neurons make viable cholinergic synapses with a host. Brain Res. Mehta, M.R., Quirck, M.C., and Wilson, M.A. (2000). Experience- 336, 302–307. dependent asymmetric shape of hippocampal receptive fields. Neu- Semba, K., and Komisaruk, B.R. (1978). Phase of the theta wave in ron 25, 707–715. relation to different limb movements in awake rats. Electroencepha- Miller, R. (1989). Cortico-hippocampal interpay: self-organizing logr. Clin. Neurophysiol. 44, 61–71. phase-locked loops for indexing memories. Psychobiology 17, Serafin, M., Williams, S., Khateb, A., Fort, P., and Muhlethaler, M. 115–128. (1996). Rhythmic firing of medial septum non-cholinergic neurons. Mitchell, S.J., and Ranck, J.B. (1980). Generation of theta rhythm in Neuroscience 75, 671–675. medial entorhinal cortex of freely moving rats. Brain Res. 189, 49–66. Sherrington, C.S. (1897). The central nervous system. In A Text Book Mitzdorf, U. (1985). Current source-density method and application of Physiology, part 3, M. Foster, ed. (Macmillan: London). in cat : investigation of evoked potentials and EEG Sik, A., Ylinen, A., Penttonen, M., and Buzsa´ ki, G. (1994). Inhibitory phenomena. Physiol. Rev. 65, 37–100. CA1–CA3-hilar region feedback in the hippocampus. Science 265, Mizumori, S.J., Barnes, C.A., and McNaughton, B.L. (1990). Behav- 1722–1724. ioral correlates of theta-on and theta-off cells recorded from hippo- Sik, A., Penttonen, M., Ylinen, A., Buzsa´ ki, G. (1995). Hippocampal campal formation of mature young and aged rats. Exp. Brain Res. CA1 interneurons: an in vivo intracellular labeling study. J. Neurosci 80, 365–373. 15, 6651–6665. Monmaur, P., Thomson, M.A., and M’Harzi, M. (1986). Temporal Sik, A., Penttonen, M., and Buzsa´ ki, G. (1997). Interneurons in the changes in hippocampal theta activity following twenty minutes of hippocampal dentate gyrus: an in vivo itnracellular study. Eur. J. forebrain ischemia in the chronic rat. Brain Res. 378, 262–273. Neurosci. 9, 573–588. Moroni, F., Corradetti, F., Moneti, G., and Pepeu, G. (1981). The Skaggs, W.E., McNaughton, B.L., Wilson, M.A., and Barnes, C.A. Neuron 16

(1996). Theta phase precession in hippocampal neuronal popula- field development, and the phase precession effect. J. Neurophysiol. tions and the compression of temporal sequences. Hippocampus 78, 393–408. 6, 149–173. Wang, X.-J. (2002). Pacemaker neurons for the theta rhythm and Soltesz, I., and Desche´ nes, M. (1993). Low- and high-frequency their synchronization in the septo-hippocampal reciprocal loop. J. membrane potential oscillations during theta activity in CA1 and Neurophysiol., in press. CA3 pyramidal neurons of the rat hippocampus under ketamine- Westenbroek, R.E., Ahlijanian, M.K., and Catterall, W.A. (1990). Clus- xylazine anesthesia. J. Neurophysiol. 70, 97–116. tering of L-type Ca2ϩ channels at the base of major dendrites in Spruston, N., Schiller, Y., Stuart, G., and Sakmann, B. (1995). Activ- pyramidal neurons. Nature 347, 281–284. ity-dependent action potential invasion and calcium influx into CA1 Whishaw, I.Q. (1976). Neuromuscular blockade: the effects on two dendrites. Science 268, 297–300. hippocampal RSA (theta) systems and neocortical desynchroniza- Steriade, M. (2000). Corticothalamic resonance, states of vigilance tion. Brain Res. Bull. 1, 573–581. and mentation. Neuroscience 101, 243–276. Whishaw, I.Q., Bland, B.H., and Bayer, S.A. (1978). Postnatal hippo- Stewart, M., and Fox, S.E. (1990). Do septal neurons pace the hippo- campal granule cell agenesis in the rat: effects on two types of campal theta rhythm? Trends Neurosci. 13, 163–168. rhythmical slow activity (RSA) in two hippocampal generators. Brain Res. 146, 249–268. Strata, F. (1998). Intrinsic oscillations in CA3 hippocampal pyramids: physiological relevance theta rhythm generation. Hippocampus 8, Williams, J.H., and Kauer, J.A. (1997). Properties of carbachol- 666–679. induced oscillatory activity in rat hippocampus. J. Neurophysiol. 78, 2631–2640. Tesche, C.D., and Karhu, J. (2000). Theta oscillations index human hippocampal activation during a working memory task. Proc. Natl. Wilson, M.A., and McNaughton, B.L. (1993). Dynamics of the hippo- Acad. Sci. USA 97, 919–924. campal ensemble code for space. Science 261, 1055–1058. Tsubokawa, H., and Ross, W.N. (1996). IPSPs modulate spike back- Wilson, R.I., and Nicoll, R.A. (2001). Endogenous cannabinoids me- propagation and associated [Ca2ϩ]i changes in the dendrites of diate retrograde signalling at hippocampal synapses. Nature 410, hippocampal CA1 pyramidal neurons. J. Neurophysiol. 76, 2896– 588–592. 2906. Winson, J. (1974). Patterns of hippocampal theta rhythm in freely Tsubokawa, H., and Ross, W.N. (1997). Muscarinic modulation of moving rat. Electroencephalogr. Clin. Neurophysiol. 36, 291–301. spike backpropagation in the apical dendrites of hippocampal CA1 Ylinen, A., Soltesz, I., Bragin, A., Penttonen, M., Sik, A., and Buzsa´ ki, pyramidal neurons. J. Neurosci. 17, 5782–5791. G. (1995). Intracellular correlates of hippocampal theta rhythm in identified pyramidal cells, granule cells and basket cells. Hippocam- To´ th, K., Borhegyi, Z., and Freund, T.F. (1993). Postsynaptic targets pus 5, 78–90. of GABAergic hippocampal neurons in the medial septum-diagonal band of broca complex. J. Neurosci. 13, 3712–3724. To´ th, K., Freund, T.F., and Miles, R. (1997). Disinhibition of rat hippo- campal pyramidal cells by GABAergic afferents from the septum. J. Physiol. (Lond.) 500, 463–474. Traub, R.D., Miles, R., and Buzsa´ ki, G. (1992). Computer simulation of carbachol-driven rhythmic population oscillations in the CA3 re- gion of the in vitro rat hippocampus. J. Physiol. (Lond.) 451, 653–672. Tsodyks, M.V., Skaggs, W.E., Sejnowski, T.J., and McNaughton, B.L. (1996). Population dynamics and theta rhythm phase precession of hippocampal place cell firing: a spiking neuron model. Hippocampus 6, 271–280. Turner, D.A., Li, X.G., Pyapali, G.K., Ylinen, A., and Buzsa´ ki, G. (1995). Morphometric and electrical properties of reconstructed hippocam- pal CA3 neurons recorded in vivo. J. Comp. Neurol. 356, 580–594. Van der Linden, S., Panzica, F., and De Curtis, M. (1999). Carbachol induces fast oscillations in the medial but not in the lateral entorhinal cortex of the isolated guinea pig brain. J. Neurophysiol. 82, 2441– 2450. Vanderwolf, C.H. (1969). Hippocampal electrical activity and volun- tary movement in the rat. Electroencephalogr. Clin. Neurophysiol. 26, 407–418. Vanderwolf, C.H. (1988). Cerebral activity and behavior: control by central cholinergic and serotonergic systems. Int. Rev. Neurobiol. 30, 225–340. Vanderwolf, C.H., and Leung, L.S. (1983). Hippocampal rhythmical slow activity: a brief history and effects entorhinal lesions and phen- cycidine. The Neurobiology of the Hippocampus, W. Seifert, ed. (London: Academic Press), pp. 275–302.. Vertes, R.P., and Kocsis, B. (1997). Brainstem-diencephalo-septo- hippocampal systems controlling the theta rhythm of the hippocam- pus. Neuroscience 81, 893–926. Vinogradova, O.S. (1995). Expression, control, and probable func- tional significance of the neuronal theta-rhythm. Prog. Neurobiol. 45, 523–583. Vizi, E.S., and Kiss, J.P. (1998). Neurochemistry and pharmacology of the major hippocampal transmitter systems: synaptic and non- synaptic interactions. Hippocampus 8, 566–607. Wallenstein, G.V., and Hasselmo, M.E. (1997). GABAergic modula- tion of hippocampal population activity: sequence learning, place