Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2012 Jan 4;590(Pt 4):715–724. doi: 10.1113/jphysiol.2011.224659

Cortical basket cell dysfunction in schizophrenia

Allison A Curley 1, David A Lewis 1
PMCID: PMC3381305  PMID: 22219337

Abstract

Schizophrenia, a debilitating illness affecting 0.5–1% of the world's population, is characterized by positive, negative and cognitive symptoms. The latter are the best predictor of functional outcome, though largely untreated by current pharmacotherapy; thus a better understanding of the mechanisms underlying cognitive deficits in schizophrenia is crucial. Higher order cognitive processes, such as working memory, are associated with θ (4–7 Hz) and γ (30–80 Hz) oscillations in the prefrontal cortex (PFC), and subjects with schizophrenia exhibit working memory impairments and reduced cortical θ and γ band power. Cortical θ and γ oscillations are dependent on perisomatic inhibition of pyramidal neurons from basket cells expressing cholecystokinin (CCKb cells) and parvalbumin (PVb cells), respectively. Thus, alterations in basket cells may underlie the cortical oscillation deficits and working memory impairments in schizophrenia. Recent findings from postmortem studies suggest that schizophrenia is associated with multiple molecular alterations that regulate signalling from CCKb and PVb cells. These alterations include lower CCK and cannabinoid 1 receptor (CB1R) in CCKb cells, and lower glutamic acid decarboxylase 67 (GAD67) and increased μ opioid receptor (μOR) in PVb cells, as well as lower GABAA receptor α1 subunit in pyramidal neurons postsynaptic to PVb cells. These changes are thought to lead to increased and decreased strength, respectively, of CCKb and PVb cell-mediated inhibition of postsynaptic pyramidal cells. Therefore, a convergence of evidence suggests a substantial shift in the relative strengths of PFC pyramidal cell inhibition from CCKb and PVb cells that may underlie cortical oscillation deficits and working memory impairments in schizophrenia.


David Lewis received his medical degree from the Ohio State University, completed residencies in internal medicine and in psychiatry at the University of Iowa, and received his research training at the Research Institute of the Scripps Clinic. He is currently the UPMC Professor in Translational Neuroscience and Chairman, Department of Psychiatry, University of Pittsburgh and Director of the Translational Neuroscience Program. His research activities focus on the neural circuitry of the prefrontal cortex, and the alterations of this circuitry in schizophrenia. Allison Curley completed her doctorate in the Department of Neuroscience at the University of Pittsburgh. She is currently a postdoctoral associate in Dr. Lewis's laboratory where she studies the role of interneuron and glial cell alterations in the prefrontal cortex in schizophrenia.

graphic file with name tjp0590-0715-m1.jpg

Schizophrenia is a devastating illness that affects 0.5–1% of the population worldwide (Lewis & Lieberman, 2000). Symptoms generally appear in late adolescence or early adulthood, and most affected individuals experience a life-long course of illness characterized by difficulties with employment, personal relationships and self-care. Schizophrenia is thought to be a neurodevelopmental disorder (Insel, 2010), caused by both genetic and environmental factors such as cannabis use during adolescence (van Os et al. 2010).

Symptoms of schizophrenia fall principally into three categories: positive, negative and cognitive. Positive symptoms are broadly characterized by an altered perception of reality, and include delusions, hallucinations, and disorganized thought and behaviour. Negative symptoms manifest as impaired emotional expression, motivation and social functioning. Cognitive impairments, such as alterations in working memory and executive function (Carter et al. 2008), are a core feature of schizophrenia (Elvevag & Goldberg, 2000) and the best predictor of functional outcome (Green et al. 2000). Unfortunately, current treatments for schizophrenia have little beneficial effects on cognitive symptoms (Goff et al. 2011). Thus, understanding the abnormalities in brain circuitry that give rise to the cognitive deficits is critical for developing new therapeutic interventions.

Higher order cognitive processes such as working memory are strongly associated with θ and γ oscillations, the synchronized activity of networks of pyramidal neurons at 4–7 and 30–80 Hz, respectively, in the prefrontal cortex (PFC) (Fries, 2009). For example, θ and γ band activity is induced during the delay period of working memory tasks in humans (Tallon-Baudry et al. 1998; Hsieh et al. 2011), and the power of θ and γ synchrony increases in proportion to working memory load (Gevins et al. 1997; Howard et al. 2003). In subjects with schizophrenia, cortical θ and γ band power and working memory performance are impaired (Schmiedt et al. 2005; Cho et al. 2006; Haenschel et al. 2009; Minzenberg et al. 2010).

Cortical θ and γ oscillations require strong inhibition provided by GABAergic inputs to the perisomatic region of pyramidal cells (Whittington et al. 1995; Bartos et al. 2007). Because a single interneuron may be highly divergent in its connections to pyramidal cells (e.g. in mouse neocortex, many parvalbumin (PV)-containing interneurons contact all pyramidal cells within their local field; Packer & Yuste, 2011), the firing of an interneuron can transiently silence many pyramidal cells. When the pyramidal cells are released from this inhibition, they fire in concert; the repetition of this cycle provides a synchronized oscillation of a neuronal network at a frequency that is inversely related to the duration of the pyramidal cell inhibition (Gonzalez-Burgos & Lewis, 2008).

Perisomatic inhibitory inputs to pyramidal neurons are furnished primarily by GABAergic basket cells that contain either the calcium-binding protein PV (PVb cells) or the neuropeptide cholecystokinin (CCKb cells). Although the activity of multiple interneuron subtypes is associated with both θ and γ oscillations, in the hippocampus the firing CCKb cells is most strongly coupled to the θ oscillation cycle (Klausberger et al. 2005; Klausberger & Somogyi, 2008), whereas the firing of PVb cells is most strongly coupled to the γ oscillation cycle (Hájos et al. 2004; Tukker et al. 2007; Gulyás et al. 2010). Indeed, recent studies in mice using optogenetic techniques have demonstrated that inhibition and activation of neocortical PV neurons suppresses and generates, respectively, cortical γ activity in vivo, indicating that PV neurons are both necessary and sufficient for the generation of γ oscillations (Cardin et al. 2009; Sohal et al. 2009). Additionally, because the duration of the inhibitory postsynaptic current (IPSC) in pyramidal cells is inversely related to oscillation frequency, the different GABAA receptor subunits that predominate at CCKb and PVb cell synapses with pyramidal neurons support the idea that the two subtypes contribute to oscillations of different frequencies (Traub et al. 1996). The slow IPSC decay mediated by GABAAα2-containing receptors (Lavoie et al. 1997) that, in rodent hippocampus, predominate at CCKb cell inputs to pyramidal cells (Nyíri et al. 2001), is consistent with low frequency θ oscillations; in contrast, the α1-containing receptors that predominate at PVb cell synapses (Klausberger et al. 2002) exhibit fast IPSC decay kinetics that are consistent with the high frequency of γ oscillations (Lavoie et al. 1997; Gonzalez-Burgos & Lewis, 2008). However, some studies have reported that measured IPSC durations from CCKb and PVb cells do not differ (Glickfeld & Scanziani, 2006; Galarreta et al. 2008). Taken together, these findings raise the possibility that alterations in CCKb and/or PVb cells, or in their inputs to pyramidal neurons, could contribute to the cortical oscillation deficits and working memory impairments in schizophrenia.

CCKb cell alterations in schizophrenia

In the primate neocortex, CCK-immunoreactive cell bodies are found predominately in layers 2–superficial 3 and their axon terminals predominantly innervate layers 2, 4 and 6 (Oeth & Lewis, 1993; Eggan et al. 2010) (Fig. 1). CCK mRNA is highly expressed in interneurons and is also present in pyramidal cells (Hökfelt et al. 2002), and CCK mRNA levels are lower in the PFC of subjects with schizophrenia (Hashimoto et al. 2008a; Fung et al. 2010) (Fig. 2). A subpopulation of CCKb (but not PVb) axon terminals contains the cannabinoid 1 receptor (CB1R) (Eggan et al. 2010) (CB1R is also present in some pyramidal cells at much lower levels; Freund et al. 2003), and lower CB1R mRNA and protein levels are present in the PFC in schizophrenia (Eggan et al. 2008, 2010; Hashimoto et al. 2008a) (Fig. 2). The deficits in CCK and CB1R mRNA levels are positively correlated (r = 0.81, P < 0.001 and r = 0.64, P = 0.001, respectively) with deficits in the mRNA encoding the 67 kDa isoform of glutamic acid decarboxylase (GAD67) (Hashimoto et al. 2008a; Eggan et al. 2008), the enzyme responsible for the majority of cortical GABA synthesis, leading to the suggestion that the GAD67 mRNA expression deficit in layers 2–3 includes CCKb cells (Eggan et al. 2008). However, levels of GAD67 have not yet been measured in CCKb cells in schizophrenia. Interestingly, in monkey PFC, CCK/CB1R co-expressing terminals have very low levels of GAD67 protein. For example, the ratio of GAD67 protein to the other isoform of the enzyme (GAD65) is only 0.2 in CCK/CB1R terminals relative to a ratio of 1.5 in PVb cell terminals (Fish et al. 2011). Consistent with the hypothesis that CCKb cells in monkey PFC principally rely on GAD65 for GABA synthesis (Fish et al. 2011), almost all CCK-containing, but very few PV-containing, cells in mouse neocortex also express GAD65 and appear to originate from the caudal ganglionic eminence (López-Bendito et al. 2004; Wierenga et al. 2010). Thus, CCKb cells may not contribute to the GAD67 deficit in schizophrenia.

Figure 1. Brightfield photomicrographs of CCK (left), CB1R (middle) and PV (right) immunoreactivity in area 46 of monkey PFC.

Figure 1

Numbers and hash marks denote the positions of the cortical layers, and the dashed lines indicate the layer 6–white matter border. Scale bar = 300 μm and applies to all panels. Left and middle panels are from Eggan et al. 2010, with permission from Elsevier; right panel is from Melchitzky et al. 1999, with permission from John Wiley & Sons.

Figure 2. Expression levels of CCK (top), CB1R (middle) and GAD67 (bottom) MRANs in the PFC of 23 pairs of matched comparison and schizophrenia subjects.

Figure 2

Mean values are indicated by the bar graphs. For all measures, the distribution within each subject group reflects the influence of other factors such as age and sex. Comparisons of individual pairs of matched subjects revealed that the schizophrenia subject had lower CCK, CB1R or GAD67 mRNA levels in 17, 21 and 20 of 23 pairs, respectively. Data are from Hashimoto et al. 2008a (CCK), Eggan et al. 2008 (CB1R) and Hashimoto et al. 2003, 2005 (GAD67).

The inputs from CCKb cells participate in a process known as depolarization-induced suppression of inhibition (DSI). When pyramidal cells are depolarized, the elevated calcium levels stimulate the retrograde release of endocannabinoids that bind to presynaptic CB1Rs located on CCKb cell terminals (Alger, 2004). Through inhibition of presynaptic calcium channels, activation of CB1R receptors results in suppressed GABA release from CCKb cell terminals, leading to reduced inhibition of the same and neighbouring pyramidal cells (Wilson & Nicoll, 2001, 2002). In the monkey PFC, DSI is present throughout development and into adulthood (Gonzalez-Burgos et al. 2010). Thus, in schizophrenia, lower levels of CB1Rs may result in less DSI, and subsequent greater inhibition of pyramidal cells by CCKb cells. Consistent with this idea, markers of endocannabinoid synthesis and degradation are not altered in schizophrenia (Volk et al. 2010), suggesting that CB1R expression is not downregulated in the illness in response to elevated levels of endocannabinoids. Interestingly, application of CCK also results in an endocannabinoid-mediated suppression of GABA release from CCKb cells (Földy et al. 2007); thus, lower levels of CCK in schizophrenia might also contribute to enhanced CCKb cell output. Alternatively, lower levels of CCK and CB1Rs might represent a compensatory response to augment GABA neurotransmission if GAD67 is lower in CCKb cells in schizophrenia (Eggan et al. 2008). Consistent with this interpretation, experimental reductions of GAD67 in mouse PFC are associated with lower CB1R mRNA expression, and similar to schizophrenia, the changes in GAD67 and CB1R mRNAs are significantly correlated (Eggan et al. 2012). However, since GAD67 protein levels are normally very low in CCKb axon terminals in monkey PFC (Fish et al. 2011), germ-line reductions in GAD67 expression may lead to lower CB1R mRNA levels in a non-cell-autonomous fashion.

PVb cell alterations in schizophrenia

In primate PFC, the highest densities of PV-labelled cell bodies and puncta (putative PVb axon terminals) are present in layers deep 3–4 (Conde et al. 1994; Melchitzky et al. 1999; Erickson & Lewis, 2002) (Fig. 1). A reduced density of PV-immunoreactive axon terminals in this laminar location in schizophrenia (Lewis et al. 2001) might reflect fewer PVb terminals. In addition, ∼50% of PV mRNA-positive neurons lack detectable levels of GAD67 in schizophrenia relative to matched control subjects (Hashimoto et al. 2003). Total GAD67 protein levels are also lower in schizophrenia (Guidotti et al. 2000; Curley et al. 2011), and this deficit is particularly pronounced in PVb axon terminals (Fig. 3). In a recent study (Curley et al. 2011), tissue sections were triple-labelled for GAD65, PV and GAD67. Because the axon terminals of the other major class of PV neurons, chandelier (PVch) cells, contain very low levels of GAD65 protein (Fish et al. 2011), the use of GAD65 to identify axon terminals facilitated the exclusion of PVch cell terminals. Quantitative fluorescent imaging demonstrated that GAD67 protein was ∼50% lower in the axon terminals of PVb cells, a deficit 5–10 times greater than the reduction in tissue levels of GAD67 protein in the same subjects (Curley et al. 2011) (Fig. 3). Together, these findings suggest that GAD67 mRNA and protein levels are lower predominantly in PVb cells and that this deficit may lead to reduced PVb cell-mediated inhibition of pyramidal cells in schizophrenia.

Figure 3. GAD67 protein in total grey matter (top; n = 19 pairs) and GAD67 protein in PVb axon terminals (bottom; n = 5 pairs) in the PFC of matched comparison and schizophrenia subjects.

Figure 3

Mean values are indicated by the bar graphs. For both measures, the distribution within each subject group reflects the influence of other factors such as age and sex. Comparisons of individual pairs of matched subjects revealed that the schizophrenia subject had lower GAD67 protein levels in 12/19 and 5/5 pairs, respectively. Data are from Curley et al. 2011.

PVb cells also exhibit other presynaptic alterations consistent with a weaker inhibition of pyramidal cells in the PFC of subjects with schizophrenia. For example, mRNA levels of the μ opioid receptor (μOR) are elevated in the PFC of subjects with schizophrenia (Volk et al. 2012). Activation of the μOR, which is localized to PV cells and other interneurons in the hippocampus (Drake & Milner, 2002; Stumm et al. 2004; Krook-Magnuson et al. 2011), diminishes the activity of PVb cells through two different mechanisms; perisomatic μORs hyperpolarize the cell body through the activation of inwardly rectifying potassium channels (Wimpey & Chavkin, 1991; Glickfeld et al. 2008), making the cell less likely to fire, and μORs on axon terminals lead to the suppression of vesicular GABA release (Capogna et al. 1993; Lupica, 1995). Thus, in schizophrenia, higher levels of μORs would be expected to reduce the activity of, and/or GABA release from, PVb cells resulting in weaker perisomatic inhibition of pyramidal cells.

On the postsynaptic side of the PVb-pyramidal cell synapse, lower expression of the GABAA receptor α1 subunit in schizophrenia may also result in reduced inhibition of pyramidal cells. Levels of α1 mRNA are lower in the PFC of subjects with schizophrenia (Akbarian et al. 1995; Hashimoto et al. 2008a,b; Beneyto et al. 2011; but see Duncan et al. 2010). In the same subjects, mRNA levels of the GABAAβ2 subunit, which preferentially co-assembles with the α1 subunit, are also lower to a similar degree in schizophrenia (Akbarian et al. 1995; Beneyto et al. 2011). Furthermore, the lower levels of GABAAα1 and β2 subunit mRNAs are most prominent in layers 3–4 (Beneyto et al. 2011), layers where PVb cells and axon terminals are most abundant (Conde et al. 1994). In layer 3, GABAAα1 mRNA levels are reduced by 40% in pyramidal cells, but are unaltered in GABA neurons, suggesting a selective reduction of pyramidal cell inhibition in schizophrenia (Glausier & Lewis, 2011).

PVb cells also receive input from CCKb cells (Karson et al. 2009). Application of CCK, acting at CCK2 receptors, can activate PV neurons (Földy et al. 2007; Karson et al. 2008). Thus, lower levels of CCK mRNA in schizophrenia (Hashimoto et al. 2008a; Fung et al. 2010) may also contribute to weaker PVb cell activity and lower inhibition of pyramidal neurons.

Although incomplete, the existing data suggest that the schizophrenia-associated abnormalities in PVb cells may differ from those in PVch neurons that synapse exclusively on the axon initial segment of pyramidal cells via a distinctive arrangement of axon terminals termed cartridges. PVch cells exhibit a reduced density of cartridges immunoreactive for the GABA membrane transporter 1 (GAT1) (Woo et al. 1998; Pierri et al. 1999). In contrast, the density of GAT1-IR puncta (presumably all other non-cartridge puncta, including those from PVb cells) is not altered in schizophrenia (Woo et al. 1998), suggesting that PVb terminals contain normal levels of GAT1. In addition, the density of axon initial segments immunoreactive for GABAA receptor α2 subunit is increased in schizophrenia (Volk et al. 2002), whereas the pyramidal cells postsynaptic to PVb cells exhibit lower levels of α1 mRNA (Glausier & Lewis, 2011). Since GAD67 mRNA is undetectable in ∼50% of PV neurons in schizophrenia (Hashimoto et al. 2003), and GAD67 mRNA levels per neuron in schizophrenia appear to be markedly reduced in a subpopulation of GABA neurons and normal in the remainder (Volk et al. 2000), the existing data raise the possibility that GAD67 expression is markedly reduced in PVb cells and unchanged in PVch cells in schizophrenia (Lewis et al. 2012). However, a direct assessment of GAD67 levels in PVch cells is needed.

Effect of basket cell alterations on pyramidal cell functioning in schizophrenia

In isolation, lower levels of CCK and CB1Rs in CCKb cells would be expected to strengthen the inhibition of postsynaptic pyramidal cells. Inhibition by CCKb cells might also be strengthened by increased expression of GABAAα2 receptors (Beneyto et al. 2011) in pyramidal cells, although whether these pre- and postsynaptic changes occur at the same synapses remains to be determined. In contrast, lower levels of GAD67 and increased levels of μORs in PVb cells would be expected to weaken the inhibition of postsynaptic pyramidal cells. This reduction in inhibition would be exacerbated by lower levels of GABAAα1 receptors in pyramidal cells and lower levels of CCK in CCKb cells that could decrease the activity of PVb cells. Thus, a convergence of evidence suggests that schizophrenia is associated with a substantial shift in the relative strengths of PFC pyramidal cell inhibition from CCKb and PVb cells (Table 1 and Fig. 4).

Table 1.

Comparison of the properties of CCKb and PVb cells in the primate PFC and in schizophrenia

Properties CCKb cells PVb cells Refs
Laminar location of cell bodies Predominately layer 2–superficial 3 Predominately layers 3–4 Oeth & Lewis (1993); Conde et al. (1994)
Laminar location of axon terminals Predominately layers 4 and 6 Predominately layers 3–4 Melchitzky et al. (1999)
Location of synapses on pyramidal neurons Dendritic shafts > dendritic spines > soma Dendritic spines > dendritic shafts > soma Melchitzky et al. (1999); Eggan et al. (2010)
GAD67/GAD65 ratio in axon terminals 0.2 1.5 Fish et al. (2011)
Postsynaptic GABAA receptor α subunit α2 α1 Nyíri et al. (2001); Klausberger et al. (2002)
Alterations in the PFC in schizophrenia GAD67 mRNA levels unknown ↓GAD67 mRNA Hashimoto et al. (2003)
GAD67 protein levels in axon terminals unknown ↓ GAD67 protein in axon terminals Curley et al. (2011)
Hashimoto et al. (2008a)
↓ CCK mRNA ↓ GABAAα1 mRNA in layers 3–4 Eggan et al. (2008)
↓ CB1R mRNA and protein ↓ GABAAα1 mRNA selectively in pyramidal cells in layer 3 Beneyto et al. (2011)
↑ GABAAα2 mRNA in layer 2 Glausier & Lewis (2011)
↑μ opioid receptor mRNA Volk et al. (2012)

Figure 4. Schematic drawing illustrating alterations in CCKb and PVb cells in schizophrenia.

Figure 4

CCKb cell-mediated inhibition of pyramidal neurons (grey cell) may be augmented due to lower CCK mRNA and CB1R mRNA and protein that enhances the GABA release from CCKb cells. In contrast, PVb cell-mediated inhibition of pyramidal neurons may be weakened due to lower GAD67 mRNA and protein resulting in less GABA release; increased μOR mRNA that diminishes PVb cell activity and suppresses GABA release; lower CCK mRNA that can lead to reduced activity of PVb cells; and lower GABAAα1 mRNA in pyramidal cells that diminishes the postsynaptic response. Levels of GAD67 in CCKb cells are not known.

An important question in determining the functional consequences of basket cell dysfunction in schizophrenia is whether the alterations in each cell type affect the same pyramidal cells. CCK mRNA levels in schizophrenia have not been examined in a laminar fashion. Although levels of CB1R immunoreactivity are lower in all PFC layers in schizophrenia, this difference is only statistically significant in layers deep 3–4 and 6 (Eggan et al. 2008). Additionally, PV-IR axon terminals exhibit a reduced density in layers deep 3–4, but not in layers 2–superficial 3 (Lewis et al. 2001). Thus, it appears that alterations in PVb and CCKb cell terminals both occur in the middle cortical layers, even though their cell bodies are located in different layers. However, whether the affected CCKb and PVb cells in schizophrenia target the same pyramidal neurons remains to be determined.

Although not conclusive, the current findings raise the possibility that schizophrenia is associated with an increase in the ratio of CCKb to PVb cell-mediated perisomatic inhibition of pyramidal cells. Disruption of the balance of these two sources of inhibition is hypothesized to perturb normal network functioning (Freund, 2003), and thus provides a plausible mechanism underlying both the θ and γ oscillation deficits in schizophrenia. In addition, the coupling of θ and γ oscillations during memory tasks (Schack et al. 2002; Lisman & Buzsáki, 2008) suggests that an impairment in one frequency, such as reduced γ oscillations resulting from deficient PVb-mediated inhibition, may result in corresponding alterations in θ oscillations. Thus, altered perisomatic inhibition of PFC pyramidal neurons by CCKb and PVb cells could provide a plausible mechanistic basis for the θ and γ oscillation impairments and cognitive deficits in schizophrenia (Schmiedt et al. 2005; Cho et al. 2006; Haenschel et al. 2009; Minzenberg et al. 2010).

Acknowledgments

The authors thank Mary Brady for technical assistance. Work by the authors cited in this review was supported by NIH grants MH084053 and MH043784.

Glossary

Abbreviations

CB1R

cannabinoid 1 receptor

CCK

cholecystokinin

CCKb

cholecystokinin basket

DSI

depolarization-induced suppression of inhibition

IPSC

inhibitory postsynaptic current

GAD65

glutamic acid decarboxylase 65

GAD67

glutamic acid decarboxylase 67

GAT1

GABA membrane transporter 1

μOR

mu opioid receptor

PFC

prefrontal cortex

PV

parvalbumin

PVb

parvalbumin basket

PVch

parvalbumin chandelier

Disclosures

David A. Lewis currently receives investigator-initiated research support from Bristol-Myers Squibb, Curridium Ltd and Pfizer and in 2009–2011 served as a consultant in the areas of target identification and validation and new compound development to BioLine RX, Bristol-Myers Squibb, Merck and SK Life Science. Allison A. Curley reports no competing interests.

References

  1. Akbarian S, Huntsman MM, Kim JJ, Tafazzoli A, Potkin SG, Bunney WE, Jones EG. GABAA receptor subunit gene expression in human prefrontal cortex: comparison of schizophrenics and controls. Cereb Cortex. 1995;5:550–560. doi: 10.1093/cercor/5.6.550. [DOI] [PubMed] [Google Scholar]
  2. Alger BE. Endocannabinoids: getting the message across. Proc Natl Acad Sci U S A. 2004;101:8512–8513. doi: 10.1073/pnas.0402935101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Bartos M, Vida I, Jonas P. Synaptic mechanisms of synchronized gamma oscillations in inhibitory interneuron networks. Nat Rev Neurosci. 2007;8:45–56. doi: 10.1038/nrn2044. [DOI] [PubMed] [Google Scholar]
  4. Beneyto M, Abbott A, Hashimoto T, Lewis DA. Lamina-specific alterations in cortical GABAA receptor subunit expression in schizophrenia. Cereb Cortex. 2011;21:999–1011. doi: 10.1093/cercor/bhq169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Capogna M, Gähwiler BH, Thompson SM. Mechanism of mu-opioid receptor-mediated presynaptic inhibition in the rat hippocampus in vitro. J Physiol. 1993;470:539–558. doi: 10.1113/jphysiol.1993.sp019874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Cardin JA, Carlen M, Meletis K, Knoblich U, Zhang F, Deisseroth K, Tsai LH, Moore CI. Driving fast-spiking cells induces gamma rhythm and controls sensory responses. Nature. 2009;459:663–667. doi: 10.1038/nature08002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Carter CS, Barch DM, Buchanan RW, Bullmore E, Krystal JH, Cohen J, Geyer M, Green M, Nuechterlein KH, Robbins T, Silverstein S, Smith EE, Strauss M, Wykes T, Heinssen R. Identifying cognitive mechanisms targeted for treatment development in schizophrenia: an overview of the first meeting of the Cognitive Neuroscience Treatment Research to Improve Cognition in Schizophrenia Initiative. Biol Psychiatry. 2008;64:4–10. doi: 10.1016/j.biopsych.2008.03.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Cho RY, Konecky RO, Carter CS. Impairments in frontal cortical gamma synchrony and cognitive control in schizophrenia. Proc Natl Acad Sci U S A. 2006;103:19878–19883. doi: 10.1073/pnas.0609440103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Conde F, Lund JS, Jacobowitz DM, Baimbridge KG, Lewis DA. Local circuit neurons immunoreactive for calretinin, calbindin D-28k or parvalbumin in monkey prefrontal cortex: distribution and morphology. J Comp Neurol. 1994;341:95–116. doi: 10.1002/cne.903410109. [DOI] [PubMed] [Google Scholar]
  10. Curley AA, Arion D, Volk DW, Asafu-Adjei JK, Sampson AR, Fish KN, Lewis DA. Cortical deficits of glutamic acid decarboxylase 67 expression in schizophrenia: clinical, protein, and cell type-specific features. Am J Psychiatry. 2011;168:921–929. doi: 10.1176/appi.ajp.2011.11010052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Drake CT, Milner TA. Mu opioid receptors are in discrete hippocampal interneuron subpopulations. Hippocampus. 2002;12:119–136. doi: 10.1002/hipo.1107. [DOI] [PubMed] [Google Scholar]
  12. Duncan CE, Webster MJ, Rothmond DA, Bahn S, Elashoff M, Shannon Weickert C. Prefrontal GABAA receptor alpha-subunit expression in normal postnatal human development and schizophrenia. J Psychiatr Res. 2010;44:673–681. doi: 10.1016/j.jpsychires.2009.12.007. [DOI] [PubMed] [Google Scholar]
  13. Eggan SM, Hashimoto T, Lewis DA. Reduced cortical cannabinoid 1 receptor messenger RNA and protein expression in schizophrenia. Arch Gen Psychiatry. 2008;65:772–784. doi: 10.1001/archpsyc.65.7.772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Eggan SM, Lazarus MS, Stoyak SR, Volk DW, Glausier JR, Huang ZJ, Lewis DA. Cortical glutamic acid decarboxylase 67 deficiency results in lower cannabinoid 1 receptor messenger RNA expression: implications for schizophrenia. Biol Psychiatry. 2012;71:114–119. doi: 10.1016/j.biopsych.2011.09.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Eggan SM, Melchitzky DS, Sesack SR, Fish KN, Lewis DA. Relationship of cannabinoid CB1 receptor and cholecystokinin immunoreactivity in monkey dorsolateral prefrontal cortex. Neuroscience. 2010;169:1651–1661. doi: 10.1016/j.neuroscience.2010.06.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Elvevag B, Goldberg TE. Cognitive impairment in schizophrenia is the core of the disorder. Crit Rev Neurobiol. 2000;14:1–21. [PubMed] [Google Scholar]
  17. Erickson SL, Lewis DA. Postnatal development of parvalbumin- and GABA transporter-immunoreactive axon terminals in monkey prefrontal cortex. J Comp Neurol. 2002;448:186–202. doi: 10.1002/cne.10249. [DOI] [PubMed] [Google Scholar]
  18. Fish KN, Sweet RA, Lewis DA. Differential distribution of proteins regulating GABA synthesis and reuptake in axon boutons of subpopulations of cortical interneurons. Cereb Cortex. 2011;21:2450–2460. doi: 10.1093/cercor/bhr007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Földy C, Lee SY, Szabadics J, Neu A, Soltesz I. Cell type-specific gating of perisomatic inhibition by cholecystokinin. Nat Neurosci. 2007;10:1128–1130. doi: 10.1038/nn1952. [DOI] [PubMed] [Google Scholar]
  20. Freund TF. Interneuron diversity series: Rhythm and mood in perisomatic inhibition. Trends Neurosci. 2003;26:489–495. doi: 10.1016/S0166-2236(03)00227-3. [DOI] [PubMed] [Google Scholar]
  21. Freund TF, Katona I, Piomelli D. Role of endogenous cannabinoids in synaptic signaling. Physiol Rev. 2003;83:1017–1066. doi: 10.1152/physrev.00004.2003. [DOI] [PubMed] [Google Scholar]
  22. Fries P. Neuronal gamma-band synchronization as a fundamental process in cortical computation. Annu Rev Neurosci. 2009;32:209–224. doi: 10.1146/annurev.neuro.051508.135603. [DOI] [PubMed] [Google Scholar]
  23. Fung SJ, Webster MJ, Sivagnanasundaram S, Duncan C, Elashoff M, Weickert CS. Expression of interneuron markers in the dorsolateral prefrontal cortex of the developing human and in schizophrenia. Am J Psychiatry. 2010;167:1479–1488. doi: 10.1176/appi.ajp.2010.09060784. [DOI] [PubMed] [Google Scholar]
  24. Galarreta M, Erdélyi F, Szabó G, Hestrin S. Cannabinoid sensitivity and synaptic properties of 2 GABAergic networks in the neocortex. Cereb Cortex. 2008;18:2296–2305. doi: 10.1093/cercor/bhm253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Gevins A, Smith ME, McEvoy L, Yu D. High-resolution EEG mapping of cortical activation related to working memory: effects of task difficulty, type of processing, and practice. Cereb Cortex. 1997;7:374–385. doi: 10.1093/cercor/7.4.374. [DOI] [PubMed] [Google Scholar]
  26. Glausier JR, Lewis DA. Selective pyramidal cell reduction of GABAA receptor α1 subunit messenger RNA expression in schizophrenia. Neuropsychopharmacology. 2011;36:2103–2110. doi: 10.1038/npp.2011.102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Glickfeld LL, Scanziani M. Distinct timing in the activity of cannabinoid-sensitive and cannabinoid-insensitive basket cells. Nat Neurosci. 2006;9:807–815. doi: 10.1038/nn1688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Glickfeld LL, Atallah BV, Scanziani M. Complementary modulation of somatic inhibition by opioids and cannabinoids. J Neurosci. 2008;28:1824–1832. doi: 10.1523/JNEUROSCI.4700-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Goff DC, Hill M, Barch D. The treatment of cognitive impairment in schizophrenia. Pharmacol Biochem Behav. 2011;99:245–253. doi: 10.1016/j.pbb.2010.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Gonzalez-Burgos G, Lewis DA. GABA neurons and the mechanisms of network oscillations: implications for understanding cortical dysfunction in schizophrenia. Schizophr Bull. 2008;34:944–961. doi: 10.1093/schbul/sbn070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Gonzalez-Burgos G, Yohsino H, Miyamae T, Pafundo D, Rotaru D, Lewis DA. 2010 Abstract Viewer/Itinerary Planner. Washington, DC: Society for Neuroscience; 2010. Postnatal development of retrograde endocannibinoid modulation of inhibitory transmission in monkey prefrontal cortex. Programme No. 742.18. [Google Scholar]
  32. Green MF, Kern RS, Braff DL, Mintz J. Neurocognitive deficits and functional outcome in schizophrenia: are we measuring the ‘right stuff’? Schizophr Bull. 2000;26:119–136. doi: 10.1093/oxfordjournals.schbul.a033430. [DOI] [PubMed] [Google Scholar]
  33. Guidotti A, Auta J, Davis JM, Di-Giorgi-Gerevini V, Dwivedi Y, Grayson DR, Impagnatiello F, Pandey G, Pesold C, Sharma R, Uzunov D, Costa E. Decrease in reelin and glutamic acid decarboxylase67 (GAD67) expression in schizophrenia and bipolar disorder: a postmortem brain study. Arch Gen Psychiatry. 2000;57:1061–1069. doi: 10.1001/archpsyc.57.11.1061. [DOI] [PubMed] [Google Scholar]
  34. Gulyás AI, Szabó GG, Ulbert I, Holderith N, Monyer H, Erdélyi F, Szabó G, Freund TF, Hájos N. Parvalbumin-containing fast-spiking basket cells generate the field potential oscillations induced by cholinergic receptor activation in the hippocampus. J Neurosci. 2010;30:15134–15145. doi: 10.1523/JNEUROSCI.4104-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Haenschel C, Bittner RA, Waltz J, Haertling F, Wibral M, Singer W, Linden DE, Rodriguez E. Cortical oscillatory activity is critical for working memory as revealed by deficits in early-onset schizophrenia. J Neurosci. 2009;29:9481–9489. doi: 10.1523/JNEUROSCI.1428-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Hashimoto T, Arion D, Unger T, Maldonado-Aviles JG, Morris HM, Volk DW, Mirnics K, Lewis DA. Alterations in GABA-related transcriptome in the dorsolateral prefrontal cortex of subjects with schizophrenia. Mol Psychiatry. 2008a;13:147–161. doi: 10.1038/sj.mp.4002011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Hashimoto T, Bazmi HH, Mirnics K, Wu Q, Sampson AR, Lewis DA. Conserved regional patterns of GABA-related transcript expression in the neocortex of subjects with schizophrenia. Am J Psychiatry. 2008b;165:479–489. doi: 10.1176/appi.ajp.2007.07081223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hashimoto T, Bergen SE, Nguyen QL, Xu B, Monteggia LM, Pierri JN, Sun Z, Sampson AR, Lewis DA. Relationship of brain-derived neurotrophic factor and its receptor TrkB to altered inhibitory prefrontal circuitry in schizophrenia. J Neurosci 25. 2005:372–383. doi: 10.1523/JNEUROSCI.4035-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Hashimoto T, Volk DW, Eggan SM, Mirnics K, Pierri JN, Sun Z, Sampson AR, Lewis DA. Gene expression deficits in a subclass of GABA neurons in the prefrontal cortex of subjects with schizophrenia. J Neurosci. 2003;23:6315–6326. doi: 10.1523/JNEUROSCI.23-15-06315.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Hájos N, Pálhalmi J, Mann EO, Németh B, Paulsen O, Freund TF. Spike timing of distinct types of GABAergic interneuron during hippocampal gamma oscillations in vitro. J Neurosci. 2004;24:9127–9137. doi: 10.1523/JNEUROSCI.2113-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Howard MW, Rizzuto DS, Caplan JB, Madsen JR, Lisman J, Aschenbrenner-Scheibe R, Schulze-Bonhage A, Kahana MJ. Gamma oscillations correlate with working memory load in humans. Cereb Cortex. 2003;13:1369–1374. doi: 10.1093/cercor/bhg084. [DOI] [PubMed] [Google Scholar]
  42. Hökfelt T, Blacker D, Broberger C, Herrera-Marschitz M, Snyder G, Fisone G, Cortés R, Morino P, You ZB, Ögren SO. Some aspects on the anatomy and function of central cholecystokinin systems. Pharmacol Toxicol. 2002;91:382–386. doi: 10.1034/j.1600-0773.2002.910617.x. [DOI] [PubMed] [Google Scholar]
  43. Hsieh LT, Ekstrom AD, Ranganath C. Neural oscillations associated with item and temporal order maintenance in working memory. J Neurosci. 2011;31:10803–10810. doi: 10.1523/JNEUROSCI.0828-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Insel TR. Rethinking schizophrenia. Nature. 2010;468:187–193. doi: 10.1038/nature09552. [DOI] [PubMed] [Google Scholar]
  45. Karson MA, Tang AH, Milner TA, Alger BE. Synaptic cross talk between perisomatic-targeting interneuron classes expressing cholecystokinin and parvalbumin in hippocampus. J Neurosci. 2009;29:4140–4154. doi: 10.1523/JNEUROSCI.5264-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Karson MA, Whittington KC, Alger BE. Cholecystokinin inhibits endocannabinoid-sensitive hippocampal IPSPs and stimulates others. Neuropharmacology. 2008;54:117–128. doi: 10.1016/j.neuropharm.2007.06.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Klausberger T, Somogyi P. Neuronal diversity and temporal dynamics: the unity of hippocampal circuit operations. Science. 2008;321:53–57. doi: 10.1126/science.1149381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Klausberger T, Marton LF, O’Neill J, Huck JH, Dalezios Y, Fuentealba P, Suen WY, Papp E, Kaneko T, Watanabe M, Csicsvari J, Somogyi P. Complementary roles of cholecystokinin- and parvalbumin-expressing GABAergic neurons in hippocampal network oscillations. J Neurosci. 2005;25:9782–9793. doi: 10.1523/JNEUROSCI.3269-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Klausberger T, Roberts JD, Somogyi P. Cell type- and input-specific differences in the number and subtypes of synaptic GABAA receptors in the hippocampus. J Neurosci. 2002;22:2513–2521. doi: 10.1523/JNEUROSCI.22-07-02513.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Krook-Magnuson E, Luu L, Lee SH, Varga C, Soltesz I. Ivy and neurogliaform interneurons are a major target of μ-opioid receptor modulation. J Neurosci. 2011;31:14861–14870. doi: 10.1523/JNEUROSCI.2269-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Lavoie AM, Tingey JJ, Harrison NL, Pritchett DB, Twyman RE. Activation and deactivation rates of recombinant GABAA receptor channels are dependent on α-subunit isoform. Biophys J. 1997;73:2518–2526. doi: 10.1016/S0006-3495(97)78280-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Lewis DA, Lieberman JA. Catching up on schizophrenia: natural history and neurobiology. Neuron. 2000;28:325–334. doi: 10.1016/s0896-6273(00)00111-2. [DOI] [PubMed] [Google Scholar]
  53. Lewis DA, Cruz DA, Melchitzky DS, Pierri JN. Lamina-specific deficits in parvalbumin-immunoreactive varicosities in the prefrontal cortex of subjects with schizophrenia: evidence for fewer projections from the thalamus. Am J Psychiatry. 2001;158:1411–1422. doi: 10.1176/appi.ajp.158.9.1411. [DOI] [PubMed] [Google Scholar]
  54. Lewis DA, Curley AA, Glausier JR, Volk DW. Cortical parvalbumin interneurons and cognitive function in schizophrenia. Trends Neurosci. 2012;35:57–67. doi: 10.1016/j.tins.2011.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Lisman J, Buzsáki G. A neural coding scheme formed by the combined function of gamma and theta oscillations. Schizophr Bull. 2008;34:974–980. doi: 10.1093/schbul/sbn060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. López-Bendito G, Sturgess K, Erdélyi F, Szabó G, Molnár Z, Paulsen O. Preferential origin and layer destination of GAD65-GFP cortical interneurons. Cereb Cortex. 2004;14:1122–1133. doi: 10.1093/cercor/bhh072. [DOI] [PubMed] [Google Scholar]
  57. Lupica CR. δ and μ enkephalins inhibit spontaneous GABA-mediated IPSCs via a cyclic AMP-independent mechanism in the rat hippocampus. J Neurosci. 1995;15:737–749. doi: 10.1523/JNEUROSCI.15-01-00737.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Melchitzky DS, Sesack SR, Lewis DA. Parvalbumin-immunoreactive axon terminals in macaque monkey and human prefrontal cortex: laminar, regional, and target specificity of type I and type II synapses. J Comp Neurol. 1999;408:11–22. [PubMed] [Google Scholar]
  59. Minzenberg MJ, Firl AJ, Yoon JH, Gomes GC, Reinking C, Carter CS. Gamma oscillatory power is impaired during cognitive control independent of medication status in first-episode schizophrenia. Neuropsychopharmacology. 2010;35:2590–2599. doi: 10.1038/npp.2010.150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Nyíri G, Freund TF, Somogyi P. Input-dependent synaptic targeting of α2-subunit-containing GABAA receptors in synapses of hippocampal pyramidal cells of the rat. Eur J Neurosci. 2001;13:428–442. doi: 10.1046/j.1460-9568.2001.01407.x. [DOI] [PubMed] [Google Scholar]
  61. Oeth KM, Lewis DA. Postnatal development of the cholecystokinin innervation of monkey prefrontal cortex. J Comp Neurol. 1993;336:400–418. doi: 10.1002/cne.903360307. [DOI] [PubMed] [Google Scholar]
  62. Packer AM, Yuste R. Dense, unspecific connectivity of neocortical parvalbumin-positive interneurons: a canonical microcircuit for inhibition? J Neurosci. 2011;31:13260–13271. doi: 10.1523/JNEUROSCI.3131-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Pierri JN, Chaudry AS, Woo TU, Lewis DA. Alterations in chandelier neuron axon terminals in the prefrontal cortex of schizophrenic subjects. Am J Psychiatry. 1999;156:1709–1719. doi: 10.1176/ajp.156.11.1709. [DOI] [PubMed] [Google Scholar]
  64. Schack B, Vath N, Petsche H, Geissler HG, Möller E. Phase-coupling of theta-gamma EEG rhythms during short-term memory processing. Int J Psychophysiol. 2002;44:143–163. doi: 10.1016/s0167-8760(01)00199-4. [DOI] [PubMed] [Google Scholar]
  65. Schmiedt C, Brand A, Hildebrandt H, Basar-Eroglu C. Event-related theta oscillations during working memory tasks in patients with schizophrenia and healthy controls. Brain Res Cogn Brain Res. 2005;25:936–947. doi: 10.1016/j.cogbrainres.2005.09.015. [DOI] [PubMed] [Google Scholar]
  66. Sohal VS, Zhang F, Yizhar O, Deisseroth K. Parvalbumin neurons and gamma rhythms enhance cortical circuit performance. Nature. 2009;459:698–702. doi: 10.1038/nature07991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Stumm RK, Zhou C, Schulz S, Höllt V. Neuronal types expressing μ- and δ-opioid receptor mRNA in the rat hippocampal formation. J Comp Neurol. 2004;469:107–118. doi: 10.1002/cne.10997. [DOI] [PubMed] [Google Scholar]
  68. Tallon-Baudry C, Bertrand O, Peronnet F, Pernier J. Induced gamma-band activity during the delay of a visual short-term memory task in humans. J Neurosci. 1998;18:4244–4254. doi: 10.1523/JNEUROSCI.18-11-04244.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Traub RD, Whittington MA, Colling SB, Buzsaki G, Jefferys JG. Analysis of gamma rhythms in the rat hippocampus in vitro and in vivo. J Physiol. 1996;493:471–484. doi: 10.1113/jphysiol.1996.sp021397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Tukker JJ, Fuentealba P, Hartwich K, Somogyi P, Klausberger T. Cell type-specific tuning of hippocampal interneuron firing during gamma oscillations in vivo. J Neurosci. 2007;27:8184–8189. doi: 10.1523/JNEUROSCI.1685-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. van Os J, Kenis G, Rutten BP. The environment and schizophrenia. Nature. 2010;468:203–212. doi: 10.1038/nature09563. [DOI] [PubMed] [Google Scholar]
  72. Volk DW, Austin MC, Pierri JN, Sampson AR, Lewis DA. Decreased glutamic acid decarboxylase67 messenger RNA expression in a subset of prefrontal cortical γ-aminobutyric acid neurons in subjects with schizophrenia. Arch Gen Psychiatry. 2000;57:237–245. doi: 10.1001/archpsyc.57.3.237. [DOI] [PubMed] [Google Scholar]
  73. Volk DW, Eggan SM, Lewis DA. Alterations in metabotropic glutamate receptor 1α and regulator of G protein signaling 4 in the prefrontal cortex in schizophrenia. Am J Psychiatry. 2010;167:1489–1498. doi: 10.1176/appi.ajp.2010.10030318. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Volk DW, Pierri JN, Fritschy JM, Auh S, Sampson AR, Lewis DA. Reciprocal alterations in pre- and postsynaptic inhibitory markers at chandelier cell inputs to pyramidal neurons in schizophrenia. Cereb Cortex. 2002;12:1063–1070. doi: 10.1093/cercor/12.10.1063. [DOI] [PubMed] [Google Scholar]
  75. Volk DW, Radchenkova PV, Walker EM, Sengupta EJ, Lewis DA. Cortical opioid markers in schizophrenia and across postnatal development. Cereb Cortex. 2012 doi: 10.1093/cercor/bhr202. (in press; doi: 10.1093/cercor/bhr202) [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Whittington MA, Traub RD, Jefferys JG. Synchronized oscillations in interneuron networks driven by metabotropic glutamate receptor activation. Nature. 1995;373:612–615. doi: 10.1038/373612a0. [DOI] [PubMed] [Google Scholar]
  77. Wierenga CJ, Müllner FE, Rinke I, Keck T, Stein V, Bonhoeffer T. Molecular and electrophysiological characterization of GFP-expressing CA1 interneurons in GAD65-GFP mice. PLoS One. 2010;5:e15915. doi: 10.1371/journal.pone.0015915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Wilson RI, Nicoll RA. Endogenous cannabinoids mediate retrograde signalling at hippocampal synapses. Nature. 2001;410:588–592. doi: 10.1038/35069076. [DOI] [PubMed] [Google Scholar]
  79. Wilson RI, Nicoll RA. Endocannabinoid signaling in the brain. Science. 2002;296:678–682. doi: 10.1126/science.1063545. [DOI] [PubMed] [Google Scholar]
  80. Wimpey TL, Chavkin C. Opioids activate both an inward rectifier and a novel voltage-gated potassium conductance in the hippocampal formation. Neuron. 1991;6:281–289. doi: 10.1016/0896-6273(91)90363-5. [DOI] [PubMed] [Google Scholar]
  81. Woo TU, Whitehead RE, Melchitzky DS, Lewis DA. A subclass of prefrontal γ-aminobutyric acid axon terminals are selectively altered in schizophrenia. Proc Natl Acad Sci U S A. 1998;95:5341–5346. doi: 10.1073/pnas.95.9.5341. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES