Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Nov 1.
Published in final edited form as: Nat Rev Neurosci. 2009 May;10(5):383–390. doi: 10.1038/nm2635

Phasic acetylcholine release and the volume transmission hypothesis: time to move on

Martin Sarter 1, Vinay Parikh 1, W Matthew Howe 1
PMCID: PMC2699581  NIHMSID: NIHMS118877  PMID: 19377503

Abstract

Traditional descriptions of the cortical cholinergic input system focused on the diffuse organization of cholinergic projections and the hypothesis that slowly changing levels of extracellular acetylcholine (ACh) mediate different arousal states. The ability of ACh to reach the extrasynaptic space (volume neurotransmission), as opposed to remaining confined to the synaptic cleft (wired neurotransmission), has been considered an integral component of this conceptualization. Recent studies demonstrated that phasic release of ACh, at the scale of seconds, mediates precisely defined cognitive operations. This characteristic of cholinergic neurotransmission is proposed to be of primary importance for understanding cholinergic function and developing treatments for cognitive disorders that result from abnormal cholinergic neurotransmission.


The entire cortical mantle is innervated by cholinergic neurons that originate in the nucleus basalis of Meynert, the substantia innominata and the horizontal limb of the diagonal band — all structures of the basal forebrain (BF) (FIG. 1). Traditionally, the cortical cholinergic input system has been categorized as the rostral component of the brain's ascending arousal systems, complementing the modulatory roles of, and interacting with, noradrenergic, serotonergic and other projection systems that broadly influence the readiness of the forebrain for input processing, wakefulness and somnolence1. However, more recent evidence has supported the more specific hypothesis that cortical cholinergic inputs mediate essential aspects of attentional information processing29. As a result, efforts to develop treatments for a wide range of cognitive disorders have focused on cholinomimetic approaches, particularly acetylcholinesterase (ACHE) inhibitors and agonists at muscarinic (m) and nicotinic (n) acetylcholine (ACh) receptors (AChRs)1012.

Figure 1. The cortical cholinergic input system.

Figure 1

a | Basal forebrain (BF) efferent cholinergic projections to the entire cortical mantle, and the main telencephalic afferent projection systems of the BF (view at a sagittal section). Cholinergic neurons originate from the nucleus basalis of Meynert, the substantia innominata and the vertical and horizontal nuclei of the diagonal band of Broca (collectively termed the BF) and innervate all cortical areas and layers. The prefrontal cortex (PFC) is the only cortical region, in rodents and primates, that is known to project back to the BF both directly and indirectly (through the nucleus accumbens (NAc)). The BF, PFC and NAc are also all innervated by dopaminergic neurons from the ventral tegmental area (VTA), and these dopaminergic neurons in turn are contacted by PFC projections. This organization suggests a profound control of the BF by the PFC. Not shown are brainstem projections to the BF. b | A composite map showing the three-dimensional distribution of cholinergic cells projecting to four arbitrarily defined mediolateral sectors of the neocortex. Cells projecting to different regions are colour-coded (medial: red; intermediary sector: blue and yellow; lateral parts of the neocortex: green). Note the relatively ordered rostromedial to caudolateral distribution of cells projecting to mediolaterally located cortical areas.c | A surface density-based render of the major organizational features in the BF (unit space: 400 × 400 × 50 μm; density threshold > 2 cells per voxel; the numbers along the z axis are the layers (sections) and the x and y values correspond to the voxel indices; for details see REF. 14). The colours of the units represent the brain regions that the cholinergic cells in those areas project to (blue: posteromedial cortex; yellow: medial prefrontal cortex; red: barrel cortex; green: posterior insular-perirhinal cortex; light blue: agranular insular-lateral orbital cortex; magenta: lateral frontal (motor) cortex). ACh, acetylcholine; GABA, γ-aminobutyric acid. Parts b and c are reproduced, with permission, from REF. 14 © (2002) Springer.

The anatomical organization of the cortical cholinergic input system seems to be largely consistent with the notion of a diffuse pathway (this article does not address the hippocampal cholinergic projection system or cholinergic projections to the amygdala). Tracing studies revealed a roughly ventrolateral, dorsomedial and rostrocaudal topographical organization of cholinergic BF projections but did not suggest a more precise topography that would indicate, for example, that adjacent neurons in the BF innervate adjacent regions in the cortex1316 (FIG. 1 b,c). Nearly all cortical layers and regions are innervated by BF cholinergic neurons17, although the distribution of choline acetyltransferase (CHAT)- or ACHE-positive fibres in the cortex indicates differences in the density of the cholinergic innervation of specific layers1821 (FIG. 2). This seemingly diffuse organization of the cortical cholinergic input system has supported descriptions that it exerts general, uniform effects across the cortical hemispheres20.

Figure 2. Cholinergic fibre distribution in the cortex.

Figure 2

Coronal sections of the medial prefrontal cortex of the rat, visualized using choline acetyl-transferase (CHAT) immunohistochemistry (a) or a histochemical method for revealing acetylcholinesterase (ACHE)-positive fibres (b), are shown to illustrate the distribution of cholinergic fibres in the cortex. The low resolution sections in parts a and b show the anterior cingulate cortex (AC), the prelimbic cortex (Prel) and the infralimbic cortex (Infral); the expansions show photomicrographs of the stippled areas, with the cortical layers indicated for part b (note that in the rat the Prel is agranular (there is no layer IV)). CHAT immunoreactivity reveals fine varicose fibres and darkly stained bipolar interneurons with axons and dendrites that are organized perpendicularly to the pial surface. The phenotype of these neurons remains elusive94: they do not express p75 receptors and thus are unaffected by local infusions of the cholinotoxin 192 immunoglobulin G–saporin. Similarly, visualization of ACHE-positive fibres reveals dense cholinergic input in all layers. Except for some minor layer-specific organizational differences, the two methods reveal essentially similar patterns of cholinergic input (see also REF. 18). The density of cholinergic inputs is similar throughout the cortex, except that there are higher densities of cholinergic input to entorhinal and olfactory regions17. Part b is modified, with permission, from REF. 72 © (2008) Elsevier.

In contrast to other diffusely organized ascending systems, such as the ascending reticular systems of the brainstem, the axons of corticopetal cholinergic neurons (subcortical afferents that project to both cerebral hemispheres) do not seem to be extensively collateralized: individual neurons innervate a relatively small cortical field2224. Thus, separate cortical regions, such as frontal and parietal regions, are not innervated by the same cholinergic neurons, suggesting that these regions may be differentially modulated by the cholinergic input system.

It has recently been proposed14,15,25 that the corticopetal cholinergic system is less diffusely organized than was traditionally assumed (FIG. 1 b,c). In support of this hypothesis, it has been demonstrated that there are clusters of cholinergic cells in the BF15,25,26 and that the BF receives modality-specific projections27. The morphological heterogeneity of BF cholinergic neurons (see REFS 28,29) and of their efferent and afferent projection systems, including the degree to which they exhibit a topographical organization, remains insufficiently understood13. For example, the finding that manipulations of the excitability of the nucleus accumbens affect prefrontal ACh release but not the release of ACh in parietal regions30,31 does not correspond with traditional descriptions of the organization of this system: it is more consistent with views suggesting a refined anatomical or functional topographical organization of the BF corticopetal projection system.

Cholinergic transmission modes

Central to the debate about the organization and function of the cortical cholinergic input system is the question of whether cholinergic neurotransmission is restricted to classical synapses (wired transmission) or is capable of escaping the synapse to stimulate distant, extrasynaptic mAChRs and nAChRs (diffuse, paracrine, non-junctional or volume transmission)32,33 (FIG. 3; TABLE 1). Despite the inconclusive evidence concerning the transmission mode that characterizes cortical ACh release, briefly reviewed below, contemporary models formalizing the functions of the BF cholinergic system and efforts to develop pro-cholinergic treatments have been based largely on the assumption that ACh is volume transmitted7,10. However, new evidence suggests that phasic transmission might have a central role in the cholinergic system.

Figure 3. Major steps in the synthesis, release and metabolism of ACh, and the main characteristics of wired and volume transmission.

Figure 3

Except for localized increases in choline resulting from acetylcholine (ACh) hydrolysis by acetylcholinesterase (ACHE), extracellular concentrations of choline are stable at ∼4.85 μM95. To synthesize ACh, choline is transported into the terminal through choline transporter (CHT)95. In the terminal, choline acetyltransferase (CHAT) catalyses the synthesis of ACh from choline and acetyl CoA (AcCoA). The capacity of CHT is the most significant determinant of the rate of ACh synthesis. ACh is packed into vesicles by vesicular acetylcholine transporter (VACHT) and released on depolarization of the terminal. Following release, ACh can bind to nicotinic (n) and muscarinic (m) ACh receptors (AChRs) and is rapidly hydrolysed by ACHE to yield choline and acetate. In the wired model of cholinergic neurotransmission (a), the presence and high catalytic activity of ACHE restricts the neurotransmission to classic synapses or junctional complexes. By contrast, in the volume model of cholinergic neurotransmission (b), most presynaptic cholinergic terminals in the cortex do not form junctional complexes and so neurotransmission is mediated by ACh that escapes hydrolysis because of insufficient or regulated availability and/or activity of ACHE. This ACh reaches the extracellular space and can stimulate non-junctional nAChRs and mAChRs. As discussed in the main text, the generation of second-scale cholinergic transients seems to represent a more important characteristic of cholinergic neurotransmission than either mode of neurotransmission.

Table 1. Main characteristics of volume and wired neurotransmission.

Volume Wired
Distribution of AChR and release sites Non-junctional complex Junctional complex
Transmission specificity Transmission ‘privacy’ is limited to the specificity of the neurotransmitter and the selectivity of receptors Transmission ‘privacy’ is based on the presence of a transmission channel dedicated to this neurotransmitter at this synapse
Ratio of pre- to postsynaptic sites One source of neurotransmitter release affects many targets One-to-one neurotransmission
Transmission timeline Long transmission delay Minimal transmission delay
Type of coded information Produces widespread, general effects on ‘arousal’, ‘readiness for processing’, vigilance or somnolence Inserts discrete and essential information into target circuits
Mimicking transmission mode Effects of released neurotransmitter can typically be reproduced by administering direct agonists at target receptors or drugs that elevate extracellular levels of the neurotransmitter (for example, uptake inhibitors) As direct receptor agonists or drugs that elevate extrasynaptic neurotransmitter levels do not reproduce the phasic pattern of neurotransmission that typically codes the information transferred to the target neuron, such drugs typically fail to reproduce the effects of endogenously released neurotransmitter
Energy costs Low energy demands; opportunistic transmission along energy gradients is used for other purposes Consumes more space and involves higher energy costs

AChR, acetylcholine receptor. Table is adapted from Agnati et al.33.

Wired versus volume transmission

Some studies concluded that the great majority of cholinergic terminals in the cortex of rats and humans form synaptic contacts34,35, supporting the notion that cholinergic transmission is mainly wired. By contrast, other studies that quantitatively analysed the ratio of cholinergic pre- to postsynaptic structures3640 demonstrated that mAChRs are present at non-cholinergic synapses41, providing strong support for volume transmission of ACh. Furthermore, measures of basal ACh release obtained using microdialysis have been interpreted to indicate an extracellular ambient level of ACh37, estimated to be in the high nanomolar to low micromolar range42.

The reasons for the discrepancy between these studies remain unclear, and other studies have also produced apparently contradictory results. For example, the presence of extrasynaptic Ml and M2 mAChRs in the cortex with high affinity for ACh suggests volume transmission41. However, other studies that analysed the relationships between cholinergic innervation and the distribution of mAChRs suggested a close correspondence, indicative of wired transmission18,43. With respect to nAChRs in the cortex, the predominant presence of nAChRs at presynaptic terminals of glutamatergic and other cortical afferents44 has been suggested to be indicative of an extrasynaptic role and therefore volume transmission4547. Furthermore, extrasynaptic α7 nAChRs have been demonstrated to be present in subcortical regions46. However, it is not clear whether the presence of heteroreceptors by default indicates the presence of non-junctional complexes, and whether the extrasynaptic location of these receptors, specifically α7 nAChRs, indicates volume transmission or unrelated signalling events4850.

As microdialysis probes are too large to enter the synaptic cleft, the recovery of ACh using microdialysis has been attributed to the presence of ACh in the extracellular space. However, the exquisite sensitivity of ACh levels collected by microdialysis to depolarization blockade by tetrodotoxin (for example, see REF. 51) suggests that ACh collected by this method both is tightly controlled by presynaptic activity and originates from synapses located extremely close to the microdialysis membrane. Furthermore, tetrodotoxin administration to ACHE-deficient mutant mice, with 100 times the normal level of basal ACh (4.6 nM instead of 556 nM), decreased ACh levels by 98%52. This finding illustrates that basal ACh levels measured by microdialysis closely reflect synaptic activity.

The interpretation that ACh recovered by microdialysis is indicative of volume transmission assumes that many of these presynaptic terminals form non-junctional release sites and/or are associated with concentrations of ACHE that are insufficient to completely hydrolyse newly released ACh. However, insertion of microdialysis probes results in oedema, haemorrhage, blood–brain barrier disruption, decreases in various enzymes associated with neurotransmitter synthesis and metabolism, intracellular changes, gliosis, neurodegeneration and lasting suppression of glucose metabolism5358. Therefore, the ACh is in essence recovered from scarred tissue57, and so the conclusion that microdialysis results demonstrate volume transmission must be treated with caution (see also REFS 59,60).

Acetylcholinesterase

The exceptional catalytic power of ACHE (one molecule of ACHE can hydrolyse 5,000 molecules of ACh per second6163) and the presence of ACHE clusters at the synapse64 have traditionally been cited in support of wired transmission of ACh. It has been suggested that as ACh dissociates from receptors it is so effectively hydrolysed by free ACHE units that it is virtually impossible that a single molecule of ACh may escape and activate another receptor65. The view that an enzyme that is characterized by such high catalytic power functions primarily to maintain a relatively stable extracellular ACh concentration contrasts boldly with the more traditional view that ACHE serves to rapidly and completely hydrolyse newly released ACh, thereby preventing spillover into the extrasynaptic space.

For volume transmission to occur, or for ACHE to have only a minor role in the immediate elimination of newly released ACh37, the ACHE must be assumed to be limited in concentration, compartmentalized and/or regulated, and to thereby limit ACh metabolism. Much of our understanding of ACHE localization and enzymatic activity has been derived from work at the muscular endplate, where clusters of ACHE can be found. It has been suggested that the characteristics of ACHE localization at synapses in the brain might be sufficiently different to allow volume transmission37. It is not yet known whether this is the case, as evidence concerning the exact localization of synaptic, membrane-bound ACHE in the forebrain remains scarce. Likewise, the regulation and function of the soluble forms of synaptic ACHE are also poorly understood. Secreted, freely diffusing ACHE, although it constitutes a minor proportion of the overall ACHE in the brain, could hydrolyse ACh even if membrane-bound ACHE became saturated following a massive release event66.

Using enzyme-selective microelectrodes and fixed-potential amperometry, it is possible to measure choline spikes that result from the hydrolysis of newly released ACh67-71. Likewise, by immobilizing ACHE on the surface of the recording electrode, in addition to choline oxidase (CHOX), ACh can be directly detected because it is hydrolysed on the recording surface. This adds to the choline signal detected by CHOX and yields larger currents than those obtained with recordings that detect choline only. As indicated by the results of calibration studies, electrodes coated with both ACHE and CHOX were able to detect low nanomolar concentrations of ACh, and co-immobilization of ACHE and CHOX did not affect the sensitivity of such recording sites to choline alone (for details see REF. 72). Massive depolarization events, triggered in vivo by application of KCl72, did not result in the detection of larger current amplitudes by these doubly coated recording sites. ACh amounts at the recording sites were estimated at > 150 fmol. Although it cannot be excluded that lower levels of extracellular ACh exist and that electrodes equipped with more potent enzymes73 would reveal this, the above studies should have detected a KCl-evoked extra-synaptic concentric wave of a millimolar concentration of ACh37,40 in support of volume transmission if this were the case.

In summary, the evidence concerning the presence and the degree of volume transmission of ACh remains inconclusive. The ongoing debate is further complicated by the insufficiently and variably defined criteria and characteristics for both modes — wired and volume — of transmission (TABLE 1). It is possible that extracellular ACh diffuses only over short distances, thereby maintaining or establishing a fast form of non-synaptic transmission45. As we explain next, the exact mode of cholinergic neurotransmission may be of minor importance compared with the implications of new evidence that illustrates the phasic characteristics of functional ACh release.

Phasic cholinergic signalling

Here we discuss the potential functional implications of evidence which indicates that phasic cholinergic signals lasting seconds underlie the functions of the cortical cholinergic input system. This evidence per se does not reject the possibility that extrasynaptic cholinergic transmission takes place, particularly over short distances as described above45. However, it does question the functional significance of persistent ambient levels of extracellular ACh37,74.

The temporal resolution of studies that use microdialysis to measure ACh release is on the scale of minutes (BOX 1). Consequently, conclusions based on microdialysis data were consistent with the conventional characteristics of ACh acting as a neuromodulator, including slowly changing release levels and volume transmission. With the advent of enzyme-selective microelectrodes, the measurement of ACh concentrations with a sub-second resolution has become possible68,69, and real-time measurements of ACh release in vivo have necessitated revisions of hypotheses concerning the regulation and function of the cortical cholinergic input system.

Box 1. ACh in attentional performance.

During the performance of a sustained attention task, acetylcholine (ACh) release typically increases by 120–140% over the pre-task baseline. Part a of the figure shows a paradigmatic illustration of prefrontal ACh release, measured using in vivo microdialysis of animals performing an attentional task (the inset illustrates the placement of a microdialysis probe on a coronal section of the rat medial prefrontal cortex). Such increases are not observed in animals performing various control tasks. Because of the low temporal resolution of this technique (10 min collections were required to produce a single detectable data point), data from such studies are consistent with the conventional description of slowly changing levels of ACh, mediating ‘arousal’ states. If ACh release indeed had these characteristics, the question of whether, and to what degree, ACh is volume transmitted would be crucially important to hypotheses concerning the functions of this neuronal system. However, as illustrated in part b of the figure, our recent experiments using choline-sensitive microelectrodes (the placement of an electrode with four platinum- and enzyme-coated recording sites fabricated into its tip is illustrated in the inset) indicated that cholinergic activity occurs at the scale of seconds, and that transient increases in ACh release mediate the detection component of attention tasks. (Detection is defined as a cognitive process that involves the incorporation of a cue into the ongoing cognitive and behavioural process and therefore allows the cue to control behaviour; for details see REF. 9.) The graph provides a paradigmatic illustration of the cholinergic transients that are evoked by a cue in trials that result in the detection of such a cue. Cues that fail to evoke such transients are missed (for actual data see REF. 9). We previously demonstrated that summing and averaging second-scale increases in cholinergic activity in task-performing animals statistically reproduced minute-scale ACh release data obtained using microdialysis (see the evidence described in the appendix of REF. 9). These highly orchestrated cholinergic transients are arguably a more important characteristic of cholinergic neurotransmission than the potential existence of volume transmission, rendering the question about the presence and degree of volume neurotransmission of secondary importance. Data in part a are modelled from data in REF. 93.

Box 1

Recording prefrontal ACh release in animals performing a cued appetitive response task, we found that transient increases in ACh at the scale of seconds mediate cue-evoked attention to the reward ports (termed cue detection9; see BOX 1 for a paradigmatic illustration of such cholinergic transients). Furthermore, we measured ACh release using microdialysis in task-performing animals and demonstrated that the results, collected over an 8 min period, could be reproduced using the second-scale cholinergic transients by summing up and averaging these transients over 8 min periods (see supplemental data in REF. 9). Thus, the differential functions of the cortical cholinergic input system can be described by using different temporal units of behavioural and cognitive processes8. This evidence did not substantiate the presence of a separate, independent mode of cholinergic neurotransmission characterized by slower, minute-scale changes in cholinergic activity.

Recent research has begun defining the local prefrontal circuitry that is responsible for generating phasic cholinergic signalling44. The evidence suggests that such cholinergic signals result from local intracortical glutamate–choline interactions, with glutamate signals that originate from thalamic afferents and stimulation of ionotropic glutamate receptors, presumably situated on cholinergic terminals, representing key steps in the generation of cholinergic transients.

Implications for neuropsychopharmacology

If fluctuations in basal extracellular ACh levels were functionally significant, ACHE inhibitors, by blocking the hydrolysis of ACh and thereby robustly elevating basal extracellular ACh levels, would be expected to markedly enhance cue detection processes and associated attentional performance. Likewise, drugs that act as mAChR agonists and mimick the tonic stimulation of these receptors that results from increased extracellular ACh levels should enhance the cognitive functions of healthy subjects and patients suffering from cognitive impairments. However, the efficacy of such cholinomimetic treatments has remained strikingly below expectation75-77. Although numerous mechanisms might contribute to the limited pro-cognitive effects of these compounds, such treatments are unlikely to augment, amplify or mimic the phasic characteristics of cholinergic activity that mediate cognitive operations, and thus they would not be expected to facilitate the cognitive functions that depend on cholinergic activity.

In contrast to pharmacological approaches that focus on elevating extracellular levels of ACh and direct stimulation of mAChRs, nAChR agonists, specifically α4β2*-selective nAChR agonists, generate phasic cholinergic signals and are proposed to robustly enhance cognitive functions44. Even the non-selective nAChR agonist nicotine, when given systemically to animals performing a cued appetitive response task, enhances the proportion of trials that involve cue detection and, mechanistically, augments the amplitude of the cholinergic transient that mediates cue detection78. Accumulating clinical evidence indicates that nAChR agonists enhance the cognitive abilities of a wide range of patients79-81. Although the overall evidence presently remains limited, nAChR agonists seem to act primarily by enhancing the temporally precisely orchestrated phasic cholinergic activity that mediates fundamental cognitive mechanisms.

Thus, the available neuropsychopharmacological evidence points to phasic ACh release being of crucial significance. The focus on developing cognition enhancers that amplify cholinergic transients represents a clear departure from the more traditional view that modifying persistent extracellular levels of ACh is a useful neuropsychopharmacological target.

Different modes — time to move on?

The presence or absence of volume transmission and the degree, in terms of distance and time, of extrasynaptic effects of ACh remain unresolved issues. Indeed, it is difficult to conceive of an experiment that would conclusively reject the possibility that a proportion of ACh spills into the extracellular space. A more constrained version of volume transmission, involving fast extrasynaptic actions of ACh (as proposed in REF. 50), might be more plausible than the extreme hypothesis that fluctuations in extrasynaptic ACh levels are key to understanding cholinergic function. However, this debate seems to be less crucial for future research on the functions of cholinergic systems and for the development of pro-cognitive therapies than the implications of the phasic signalling characteristics of ACh that were found in recent experiments8,9,44.

As the functional significance of neurotransmission mode has also been debated with respect to several other neuromodulator systems82-84, such as serotonergic and noradrenergic systems, our conclusions might generalize to these systems. For example, ultrastructural evidence indicated a high proportion of non-junctional neurotransmitter receptor complexes at serotonergic and noradrenergic terminals. Furthermore, the ability to measure release of these modulators using microdialysis has been interpreted as evidence for the existence of volume transmission85,86. However, it is intriguing to speculate that, similar to ACh, these neuromodulators code discrete information based on phasic release patterns. It may be further speculated that it is a common feature of all ascending neuromodulator systems that although their anatomical organization reflects a diffuse and not obviously structured projection system, the regulation of their terminal activity by local (cortical) circuitry, involving heteroreceptors situated on their terminals, allowed the evolving forebrain to use the neuromodulator input to generate precisely orchestrated signals and mediate defined cognitive operations.

Concerning the mesolimbic dopamine system, there is strong evidence for the presence of both phasic and tonic neuronal activity, interactions between the two, and associated patterns of dopamine release (for example, see REFS 87,88). Moreover, differential functions of dopamine have been attributed to it acting at different timescales, ranging from quasi hormonal functions of tonic levels of dopamine release to reward and outcome processing by phasic release89. However, tonic release can also result from asynchronous firing of groups of dopaminergic neurons90,91. Furthermore, as the timescales that are applied to describe the release of a neurotransmitter are necessarily confounded by the temporal and spatial resolution as well as the sensitivity of detection techniques, the increasing use of electrochemical techniques to measure dopamine release will provide interesting new insights into the nature of functional dopamine release at the scale of seconds (for example, see REF. 92).

If transient release patterns are sufficient to explain the target area-specific cognitive operations that are mediated by ascending ‘neuromodulator’ systems, we could finally unchain ourselves from the misleading constraints of outdated descriptions of ‘arousal’ systems and their associated dichotomies, including phasic versus tonic and volume versus wired neurotransmission, and the classification of signalling molecules as neurotransmitters versus neuromodulators. Instead we could focus on determining how evolving forebrain circuits usurped these massive input systems. By modulating neurotransmission of these inputs based on heteroreceptors situated on their terminals and local microcircuitry contacting these terminals, forebrain target areas can use seemingly diffusely organized input systems to generate function-specific, transient changes in neurotransmitter release that initiate or foster the mediation of specific cognitive operations.

Acknowledgments

The authors' research was supported by Public Health Service (PHS) grants KO2MH10172, MH080426 and MH080332. W.M.H. was supported by the PHS Training Grant T32 DA007267. We thank S. Baran for comments on a draft of this paper and an anonymous reviewer for suggesting the final title of this paper.

Footnotes

DATABASES

Entrez Gene: http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?db=gene

ACHE | CHAT

FURTHER INFORMATION

Martin Sarter's homepage: http://sitemaker.umich.edu/martin.sarter

References

  • 1.Briand LA, Gritton H, Howe WM, Young DA, Sarter M. Modulators in concert for cognition: modulator interactions in the prefrontal cortex. Prog Neurobiol. 2007;83:69–91. doi: 10.1016/j.pneurobio.2007.06.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Everitt BJ, Robbins TW. Central cholinergic systems and cognition. Annu Rev Psychol. 1997;48:649–684. doi: 10.1146/annurev.psych.48.1.649. [DOI] [PubMed] [Google Scholar]
  • 3.Sarter M, Gehring WJ, Kozak R. More attention must be paid: the neurobiology of attentional effort. Brain Res Rev. 2006;51:145–160. doi: 10.1016/j.brainresrev.2005.11.002. [DOI] [PubMed] [Google Scholar]
  • 4.Sarter M, Givens B, Bruno JP. The cognitive neuroscience of sustained attention: where top-down meets bottom-up. Brain Res Rev. 2001;35:146–160. doi: 10.1016/s0165-0173(01)00044-3. [DOI] [PubMed] [Google Scholar]
  • 5.Sarter M, Hasselmo ME, Bruno JP, Givens B. Unraveling the attentional functions of cortical cholinergic inputs: interactions between signal-driven and top-down cholinergic modulation of signal detection. Brain Res Rev. 2005;48:98–111. doi: 10.1016/j.brainresrev.2004.08.006. [DOI] [PubMed] [Google Scholar]
  • 6.Yu AJ, Dayan P. Acetylcholine in cortical inference. Neural Netw. 2002;15:719–730. doi: 10.1016/s0893-6080(02)00058-8. [DOI] [PubMed] [Google Scholar]
  • 7.Hasselmo ME, McGaughy J. High acetylcholine levels set circuit dynamics for attention and encoding and low acetylcholine levels set dynamics for consolidation. Prog Brain Res. 2004;145:201–231. doi: 10.1016/S0079-6123(03)45015-2. [DOI] [PubMed] [Google Scholar]
  • 8.Parikh V, Sarter M. Cholinergic mediation of attention: contributions of phasic and tonic increases in prefrontal cholinergic activity. Ann NY Acad Sci. 2008;1129:225–235. doi: 10.1196/annals.1417.021. [DOI] [PubMed] [Google Scholar]
  • 9.Parikh V, Kozak R, Martinez V, Sarter M. Prefrontal acetylcholine release controls cue detection on multiple timescales. Neuron. 2007;56:141–154. doi: 10.1016/j.neuron.2007.08.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Wilkinson DG, Francis PT, Schwam E, Payne-Parrish J. Cholinesterase inhibitors used in the treatment of Alzheimer's disease: the relationship between pharmacological effects and clinical efficacy. Drugs Aging. 2004;21:453–478. doi: 10.2165/00002512-200421070-00004. [DOI] [PubMed] [Google Scholar]
  • 11.Messer WSJ. Cholinergic agonists and the treatment of Alzheimer's disease. Curr Top Med Chem. 2002;2:353–358. doi: 10.2174/1568026024607553. [DOI] [PubMed] [Google Scholar]
  • 12.Stip E, Chouinard S, Boulay LJ. On the trail of a cognitive enhancer for the treatment of schizophrenia. Prog Neuropsychopharmacol Biol Psychiatry. 2005;29:219–232. doi: 10.1016/j.pnpbp.2004.11.004. [DOI] [PubMed] [Google Scholar]
  • 13.Sarter M. The substantia innominata remains incognita: pressing research themes on basal forebrain neuroanatomy. Brain Struct Funct. 2008;213:11–15. doi: 10.1007/s00429-007-0165-x. [DOI] [PubMed] [Google Scholar]
  • 14.Zaborszky L, et al. In: Computational Neuroanatomy: Principles and Methods. Ascoli A, editor. Humana; Totowa, New Jersey; 2002. pp. 171–197. [Google Scholar]
  • 15.Zaborszky L. The modular organization of brain systems. Basal forebrain: the last frontier. Prog Brain Res. 2002;136:359–372. doi: 10.1016/s0079-6123(02)36030-8. [DOI] [PubMed] [Google Scholar]
  • 16.Zaborszky L. In: Neurotransmitter Interactions and Cognitive Functions. Levin ED, Decker MW, Butcher LL, editors. Birkauser; Boston: 1992. pp. 27–65. [Google Scholar]
  • 17.Eckenstein FP, Baughman RW, Quinn J. An anatomical study of cholinergic innervation in rat cerebral cortex. Neuroscience. 1988;25:457–474. doi: 10.1016/0306-4522(88)90251-5. [DOI] [PubMed] [Google Scholar]
  • 18.Lysakowski A, Wainer BH, Bruce G, Hersh LB. An atlas of the regional and laminar distribution of choline acetyltransferase immunoreactivity in rat cerebral cortex. Neuroscience. 1989;28:291–336. doi: 10.1016/0306-4522(89)90180-2. [DOI] [PubMed] [Google Scholar]
  • 19.Satoh K, Armstrong DM, Fibiger HC. A comparison of the distribution of central cholinergic neurons as demonstrated by acetylcholinesterase pharmacohistochemistry and choline acetyltransferase immunohistochemistry. Brain Res Bull. 1983;11:693–720. doi: 10.1016/0361-9230(83)90013-8. [DOI] [PubMed] [Google Scholar]
  • 20.Mesulam MM, Hersh LB, Mash DC, Geula C. Differential cholinergic innervation within functional subdivisions of the human cerebral cortex: a choline acetyltransferase study. J Comp Neurol. 1992;318:316–328. doi: 10.1002/cne.903180308. [DOI] [PubMed] [Google Scholar]
  • 21.Raghanti MA, et al. Cholinergic innervation of the frontal cortex: differences among humans, chimpanzees, and macaque monkeys. J Comp Neurol. 2008;506:409–424. doi: 10.1002/cne.21546. [DOI] [PubMed] [Google Scholar]
  • 22.Price JL, Stern R. Individual cells in the nucleus basalis-diagonal band complex have restricted axonal projections to the cerebral cortex in the rat. Brain Res. 1983;269:352–356. doi: 10.1016/0006-8993(83)90145-2. [DOI] [PubMed] [Google Scholar]
  • 23.Walker LC, Kitt CA, DeLong MR, Price DL. Noncollateral projections of basal forebrain neurons to frontal and parietal neocortex in primates. Brain Res Bull. 1985;15:307–314. doi: 10.1016/0361-9230(85)90156-x. [DOI] [PubMed] [Google Scholar]
  • 24.Koliatsos VE, et al. Topographic, non-collateralized basal forebrain projections to amygdala, hippocampus, and anterior cingulate cortex in the rhesus monkey. Brain Res. 1988;463:133–139. doi: 10.1016/0006-8993(88)90535-5. [DOI] [PubMed] [Google Scholar]
  • 25.Zaborszky L, Buhl DL, Pobalashingham S, Bjaalie JG, Nadasdy Z. Three-dimensional chemoarchitecture of the basal forebrain: spatially specific association of cholinergic and calcium binding protein-containing neurons. Neuroscience. 2005;136:697–713. doi: 10.1016/j.neuroscience.2005.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Zaborszky L, et al. Stereotaxic probabilistic maps of the magnocellular cell groups in human basal forebrain. Neuroimage. 2008;42:1127–1141. doi: 10.1016/j.neuroimage.2008.05.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Golmayo L, Nunez A, Zaborszky L. Electrophysiological evidence for the existence of a posterior cortical-prefrontal-basal forebrain circuitry in modulating sensory responses in visual and somatosensory rat cortical areas. Neuroscience. 2003;119:597–609. doi: 10.1016/s0306-4522(03)00031-9. [DOI] [PubMed] [Google Scholar]
  • 28.Loopuijt LD, Zahm DS. Synaptologic and fine structural features distinguishing a subset of basal forebrain cholinergic neurons embedded in the dense intrinsic fiber network of the caudal extended amygdala. J Comp Neurol. 2006;498:93–111. doi: 10.1002/cne.21044. [DOI] [PubMed] [Google Scholar]
  • 29.Gastard M, Jensen SL, Martin JR, Williams EA, Zahm DS. The caudal sublenticular region/anterior amygdaloid area is the only part of the rat forebrain and mesopontine tegmentum occupied by magnocellular cholinergic neurons that receives outputs from the central division of extended amygdala. Brain Res. 2002;957:207–222. doi: 10.1016/s0006-8993(02)03513-8. [DOI] [PubMed] [Google Scholar]
  • 30.Zmarowski A, Sarter M, Bruno JP. Glutamate receptors in nucleus accumbens mediate regionally selective increases in cortical acetylcholine release. Synapse. 2007;61:115–123. doi: 10.1002/syn.20354. [DOI] [PubMed] [Google Scholar]
  • 31.Zmarowski A, Sarter M, Bruno JP. NMDA and dopamine interactions in the nucleus accumbens modulate cortical acetylcholine release. Eur J Neurosci. 2005;22:1731–1740. doi: 10.1111/j.1460-9568.2005.04333.x. [DOI] [PubMed] [Google Scholar]
  • 32.Fuxe K, et al. Dynamics of volume transmission in the brain. Focus on catecholamine and opioid peptide communication and the role of uncoupling protein 2. J Neural Transm. 2005;112:65–76. doi: 10.1007/s00702-004-0158-3. [DOI] [PubMed] [Google Scholar]
  • 33.Agnati LF, et al. Volume transmission and wiring transmission from cellular to molecular networks: history and perspectives. Acta Physiol (Oxf) 2006;187:329–344. doi: 10.1111/j.1748-1716.2006.01579.x. [DOI] [PubMed] [Google Scholar]
  • 34.Turrini P, et al. Cholinergic nerve terminals establish classical synapses in the rat cerebral cortex: synaptic pattern and age-related atrophy. Neuroscience. 2001;105:277–285. doi: 10.1016/s0306-4522(01)00172-5. [DOI] [PubMed] [Google Scholar]
  • 35.Smiley JF, Morrell F, Mesulam MM. Cholinergic synapses in human cerebral cortex: an ultrastructural study in serial sections. Exp Neurol. 1997;144:361–368. doi: 10.1006/exnr.1997.6413. [DOI] [PubMed] [Google Scholar]
  • 36.Umbriaco D, Watkins KC, Descarries L, Cozzari C, Hartman BK. Ultrastructural and morphometric features of the acetylcholine innervation in adult rat parietal cortex: an electron microscopic study in serial sections. J Comp Neurol. 1994;348:351–373. doi: 10.1002/cne.903480304. [DOI] [PubMed] [Google Scholar]
  • 37.Descarries L, Gisiger V, Steriade M. Diffuse transmission by acetylcholine in the CNS. Prog Neurobiol. 1997;53:603–625. doi: 10.1016/s0301-0082(97)00050-6. [DOI] [PubMed] [Google Scholar]
  • 38.Descarries L, Mechawar N. Ultrastructural evidence for diffuse transmission by monoamine and acetylcholine neurons of the central nervous system Prog. Brain Res. 2000;125:27–47. doi: 10.1016/S0079-6123(00)25005-X. [DOI] [PubMed] [Google Scholar]
  • 39.Mechawar N, Watkins KC, Descarries L. Ultrastructural features of the acetylcholine innervation in the developing parietal cortex of rat. J Comp Neurol. 2002;443:250–258. doi: 10.1002/cne.10114. [DOI] [PubMed] [Google Scholar]
  • 40.Mechawar N, Cozzari C, Descarries L. Cholinergic innervation in adult rat cerebral cortex: a quantitative immunocytochemical description. J Comp Neurol. 2000;428:305–318. doi: 10.1002/1096-9861(20001211)428:2<305::aid-cne9>3.0.co;2-y. [DOI] [PubMed] [Google Scholar]
  • 41.Mrzljak L, Levey AI, Goldman-Rakic PS. Association of m1 and m2 muscarinic receptor proteins with asymmetric synapses in the primate cerebral cortex: morphological evidence for cholinergic modulation of excitatory neurotransmission. Proc Natl Acad Sci USA. 1993;90:5194–5198. doi: 10.1073/pnas.90.11.5194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Vinson PN, Justice JB., Jr Effect of neostigmine on concentration and extraction fraction of acetylcholine using quantitative microdialysis. J Neurosci Methods. 1997;73:61–67. doi: 10.1016/s0165-0270(96)02213-3. [DOI] [PubMed] [Google Scholar]
  • 43.Frey KA, Ehrenkaufer RL, Agranoff BW. Quantitative in vivo receptor binding. II. Autoradiographic imaging of muscarinic cholinergic receptors. J Neurosci. 1985;5:2407–2414. doi: 10.1523/JNEUROSCI.05-09-02407.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Parikh V, Man K, Decker MW, Sarter M. Glutamatergic contributions to nicotinic acetylcholine receptor agonist-evoked cholinergic transients in the prefrontal cortex. J Neurosci. 2008;28:3769–3780. doi: 10.1523/JNEUROSCI.5251-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Lendvai B, Vizi ES. Nonsynaptic chemical transmission through nicotinic acetylcholine receptors. Physiol Rev. 2008;88:333–349. doi: 10.1152/physrev.00040.2006. [DOI] [PubMed] [Google Scholar]
  • 46.Jones IW, Wonnacott S. Precise localization of α7 nicotinic acetylcholine receptors on glutamatergic axon terminals in the rat ventral tegmental area. J Neurosci. 2004;24:11244–11252. doi: 10.1523/JNEUROSCI.3009-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Hill JA, Jr, Zoli M, Bourgeois JP, Changeux JP. Immunocytochemical localization of a neuronal nicotinic receptor: the β2-subunit. J Neurosci. 1993;13:1551–1568. doi: 10.1523/JNEUROSCI.13-04-01551.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Alkondon M, Albuquerque EX. Subtype-specific inhibition of nicotinic acetylcholine receptors by choline: a regulatory pathway. J Pharmacol Exp Ther. 2006;318:268–275. doi: 10.1124/jpet.106.103135. [DOI] [PubMed] [Google Scholar]
  • 49.Wu HQ, Rassoulpour A, Schwarcz R. Kynurenic acid leads, dopamine follows: a new case of volume transmission in the brain? J Neural Transm. 2007;114:33–41. doi: 10.1007/s00702-006-0562-y. [DOI] [PubMed] [Google Scholar]
  • 50.Vizi ES, Lendvai B. Modulatory role of presynaptic nicotinic receptors in synaptic and non-synaptic chemical communication in the central nervous system. Brain Res Rev. 1999;30:219–235. doi: 10.1016/s0165-0173(99)00016-8. [DOI] [PubMed] [Google Scholar]
  • 51.Fadel J, Moore H, Sarter M, Bruno JP. Trans-synaptic stimulation of cortical acetylcholine release after partial 192 IgG-saporin-induced loss of cortical cholinergic afferents. J Neurosci. 1996;16:6592–6600. doi: 10.1523/JNEUROSCI.16-20-06592.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Hartmann J, Kiewert C, Duysen EG, Lockridge O, Klein J. Choline availability and acetylcholine synthesis in the hippocampus of acetylcholinesterase-deficient mice. Neurochem Int. 2008;52:972–978. doi: 10.1016/j.neuint.2007.10.008. [DOI] [PubMed] [Google Scholar]
  • 53.Robinson TE, Camp DM. The effects of four days of continuous striatal microdialysis on indices of dopamine and serotonin neurotransmission in rats. J Neurosci Methods. 1991;40:211–222. doi: 10.1016/0165-0270(91)90070-g. [DOI] [PubMed] [Google Scholar]
  • 54.Westerink BH, Tuinte MH. Chronic use of intracerebral dialysis for the in vivo measurement of 3,4-dihydroxyphenylethylamine and its metabolite 3,4-dihydroxyphenylacetic acid. J Neurochem. 1986;46:181–185. doi: 10.1111/j.1471-4159.1986.tb12942.x. [DOI] [PubMed] [Google Scholar]
  • 55.Benveniste H, Diemer NH. Cellular reactions to implantation of a microdialysis tube in the rat hippocampus. Acta Neuropathol. 1987;74:234–238. doi: 10.1007/BF00688186. [DOI] [PubMed] [Google Scholar]
  • 56.Schiffer WK, et al. Serial microPET measures of the metabolic reaction to a microdialysis probe implant. J Neurosci Methods. 2006;155:272–284. doi: 10.1016/j.jneumeth.2006.01.027. [DOI] [PubMed] [Google Scholar]
  • 57.Westerink BH. Brain microdialysis and its application for the study of animal behaviour. Behav Brain Res. 1995;70:103–124. doi: 10.1016/0166-4328(95)80001-8. [DOI] [PubMed] [Google Scholar]
  • 58.Clapp-Lilly KL, et al. An ultrastructural analysis of tissue surrounding a microdialysis probe. J Neurosci Methods. 1999;90:129–142. doi: 10.1016/s0165-0270(99)00064-3. [DOI] [PubMed] [Google Scholar]
  • 59.Chen KC. In: Handbook of Microdialysis. Westerink BHC, Cremers TIFH, editors. Elsevier Amsterdam; 2007. pp. 47–70. [Google Scholar]
  • 60.Borland LM, Shi G, Yang H, Michael AC. Voltammetric study of extracellular dopamine near microdialysis probes acutely implanted in the striatum of the anesthetized rat. J Neurosci Methods. 2005;146:149–158. doi: 10.1016/j.jneumeth.2005.02.002. [DOI] [PubMed] [Google Scholar]
  • 61.Cooper JR, Bloom FE, Roth RH. The Biochemical Basis of Neuropharmacology. Oxford Univ. Press; Oxford: 2003. [Google Scholar]
  • 62.Quinn DM. Acetylcholinesterase: enzyme structure, reaction dynamics, and virtual transition states. Chem Rev. 1987;87:955–979. [Google Scholar]
  • 63.Lawler HC. Turnover time of acetylcholinesterase. J Biol Chem. 1961;236:2296–2301. [PubMed] [Google Scholar]
  • 64.Peng HB, Xie H, Rossi SG, Rotundo RL. Acetylcholinesterase clustering at the neuromuscular junction involves perlecan and dystroglycan. J Cell Biol. 1999;145:911–921. doi: 10.1083/jcb.145.4.911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Zimmerman G, Soreq H. Termination and beyond: acetylcholinesterase as a modulator of synaptic transmission. Cell Tissue Res. 2006;326:655–669. doi: 10.1007/s00441-006-0239-8. [DOI] [PubMed] [Google Scholar]
  • 66.Appleyard ME. Secreted acetylcholinesterase: non-classical aspects of a classical enzyme. Trends Neurosci. 1992;15:485–490. doi: 10.1016/0166-2236(92)90100-m. [DOI] [PubMed] [Google Scholar]
  • 67.Burmeister JJ, Moxon K, Gerhardt GA. Ceramic-based multisite microelectrodes for electrochemical recordings. Anal Chem. 2000;72:187–192. doi: 10.1021/ac9907991. [DOI] [PubMed] [Google Scholar]
  • 68.Burmeister JJ, Palmer M, Gerhardt GA. Ceramic-based multisite electrode array for rapid choline measures in brain tissue. Anal Chim Acta. 2003;481:65–74. [Google Scholar]
  • 69.Parikh V, et al. Rapid assessment of in vivo cholinergic transmission by amperometric detection of changes in extracellular choline levels. Eur J Neurosci. 2004;20:1545–1554. doi: 10.1111/j.1460-9568.2004.03614.x. [DOI] [PubMed] [Google Scholar]
  • 70.Parikh V, Apparsundaram S, Kozak R, Richards JB, Sarter M. Reduced expression and capacity of the striatal high-affinity choline transporter in hyperdopaminergic mice. Neuroscience. 2006;141:379–389. doi: 10.1016/j.neuroscience.2006.03.055. [DOI] [PubMed] [Google Scholar]
  • 71.Bruno JP, et al. Second-by-second measurement of acetylcholine release in prefrontal cortex. Eur J Neurosci. 2006;24:2749–2757. doi: 10.1111/j.1460-9568.2006.05176.x. [DOI] [PubMed] [Google Scholar]
  • 72.Giuliano C, Parikh V, Ward JR, Chiamulera C, Sarter M. Increases in cholinergic neurotransmission measured by using choline-sensitive microelectrodes: enhanced detection by hydrolysis of acetylcholine on recording sites? Neurochem Int. 2008;52:1343–1350. doi: 10.1016/j.neuint.2008.02.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Burmeister JJ, et al. Ceramic-based multisite microelectrode arrays for simultaneous measures of choline and acetylcholine in CNS. Biosens Bioelectron. 2008;23:1382–1389. doi: 10.1016/j.bios.2007.12.013. [DOI] [PubMed] [Google Scholar]
  • 74.Descarries L. The hypothesis of an ambient level of acetylcholine in the central nervous system. J Physiol (Paris) 1998;92:215–220. doi: 10.1016/s0928-4257(98)80013-2. [DOI] [PubMed] [Google Scholar]
  • 75.Voelker R. Guideline: dementia drugs' benefits uncertain. JAMA. 2008;299:1763. doi: 10.1001/jama.299.15.1763. [DOI] [PubMed] [Google Scholar]
  • 76.Thal LJ, Forrest M, Loft H, Mengel H. Lu 25–109, a muscarinic agonist, fails to improve cognition in Alzheimer's disease. Lu25–109 Study Group. Neurology. 2000;54:421–426. doi: 10.1212/wnl.54.2.421. [DOI] [PubMed] [Google Scholar]
  • 77.Yesavage JA, et al. Acetylcholinesterase inhibitor in combination with cognitive training in older adults. J Gerontol B Psychol Sci Soc Sci. 2008;63:P288–P294. doi: 10.1093/geronb/63.5.p288. [DOI] [PubMed] [Google Scholar]
  • 78.Howe WM, et al. Prefrontal cholinergic transients indicating cue detection as a target for cognition enhancers. Soc Neurosci Ann Meet. 2008;388:26. [Google Scholar]
  • 79.Wilens TE, Decker MW. Neuronal nicotinic receptor agonists for the treatment of attention-deficit/hyperactivity disorder: focus on cognition. Biochem Pharmacol. 2007;74:1212–1223. doi: 10.1016/j.bcp.2007.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Newhouse PA, Potter A, Kelton M, Corwin J. Nicotinic treatment of Alzheimer's disease. Biol Psychiatry. 2001;49:268–278. doi: 10.1016/s0006-3223(00)01069-6. [DOI] [PubMed] [Google Scholar]
  • 81.Wilens TE, Verlinden MH, Adler LA, Wozniak PJ, West SA. ABT-089, a neuronal nicotinic receptor partial agonist, for the treatment of attention-deficit/hyperactivity disorder in adults: results of a pilot study. Biol Psychiatry. 2006;59:1065–1070. doi: 10.1016/j.biopsych.2005.10.029. [DOI] [PubMed] [Google Scholar]
  • 82.Rice ME, Cragg SJ. Dopamine spillover after quantal release: rethinking dopamine transmission in the nigrostriatal pathway. Brain Res Rev. 2008;58:303–313. doi: 10.1016/j.brainresrev.2008.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Nishiyama H, Linden DJ. Pure spillover transmission between neurons. Nature Neurosci. 2007;10:675–677. doi: 10.1038/nn0607-675. [DOI] [PubMed] [Google Scholar]
  • 84.Vargova L, Sykova E. Extracellular space diffusion and extrasynaptic transmission. Physiol Res. 2008;57(Suppl 3):S89–S99. doi: 10.33549/physiolres.931603. [DOI] [PubMed] [Google Scholar]
  • 85.Bunin MA, Wightman RM. Paracrine neurotransmission in the CNS: involvement of 5-HT. Trends Neurosci. 1999;22:377–382. doi: 10.1016/s0166-2236(99)01410-1. [DOI] [PubMed] [Google Scholar]
  • 86.Ciranna L. Serotonin as a modulator of glutamate- and GABA-mediated neurotransmission: implications in physiological functions and in pathology. Curr Neuropharmacol. 2006;4:101–114. doi: 10.2174/157015906776359540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Goto Y, Otani S, Grace AA. The yin and yang of dopamine release: a new perspective. Neuropharmacology. 2007;53:583–587. doi: 10.1016/j.neuropharm.2007.07.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Floresco SB. Dopaminergic regulation of limbic-striatal interplay. J Psychiatry Neurosci. 2007;32:400–411. [PMC free article] [PubMed] [Google Scholar]
  • 89.Schultz W. Multiple dopamine functions at different time courses. Annu Rev Neurosci. 2007;30:259–288. doi: 10.1146/annurev.neuro.28.061604.135722. [DOI] [PubMed] [Google Scholar]
  • 90.Arbuthnott GW, Wickens J. Space, time and dopamine. Trends Neurosci. 2007;30:62–69. doi: 10.1016/j.tins.2006.12.003. [DOI] [PubMed] [Google Scholar]
  • 91.Roitman MF, Day JJ, Seipel A, Carelli RM, Wightman RM. A steady-state concentration of dopamine is comprised of time-averaged, phasic dopamine release events. Monitoring Molecules in Neuroscience; 11th International Conference on In Vivo Methods; 2006. [Google Scholar]
  • 92.Roitman MF, Wheeler RA, Wightman RM, Carelli RM. Real-time chemical responses in the nucleus accumbens differentiate rewarding and aversive stimuli. Nature Neurosci. 2008;11:1376–1377. doi: 10.1038/nn.2219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Kozak R, et al. Toward a neuro-cognitive animal model of the cognitive symptoms of schizophrenia: disruption of cortical cholinergic neurotransmission following repeated amphetamine exposure in attentional task-performing, but not non-performing, rats. Neuropsychopharmacology. 2007;32:2074–2086. doi: 10.1038/sj.npp.1301352. [DOI] [PubMed] [Google Scholar]
  • 94.von Engelhardt J, Eliava M, Meyer AH, Rozov A, Monyer H. Functional characterization of intrinsic cholinergic interneurons in the cortex. J Neurosci. 2007;27:5633–5642. doi: 10.1523/JNEUROSCI.4647-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Parikh V, Sarter M. Cortical choline transporter function measured in vivo using choline-sensitive microelectrodes: clearance of endogenous and exogenous choline and effects of removal of cholinergic terminals. J Neurochem. 2006;97:488–503. doi: 10.1111/j.1471-4159.2006.03766.x. [DOI] [PubMed] [Google Scholar]
  • 96.Sarter M, Parikh V. Choline transporters, cholinergic transmission and cognition. Nature Rev Neurosci. 2005;6:48–56. doi: 10.1038/nrn1588. [DOI] [PubMed] [Google Scholar]

RESOURCES