Washington University School of Medicine Digital Commons@Becker

Open Access Publications

12-2008 Many specialists for suppressing cortical excitation Andreas Burkhalter Washington University School of Medicine in St. Louis

Follow this and additional works at: https://digitalcommons.wustl.edu/open_access_pubs Part of the Medicine and Health Sciences Commons

Recommended Citation Burkhalter, Andreas, ,"Many specialists for suppressing cortical excitation." Frontiers in .,. 1-167. (2008). https://digitalcommons.wustl.edu/open_access_pubs/495

This Open Access Publication is brought to you for free and open access by Digital Commons@Becker. It has been accepted for inclusion in Open Access Publications by an authorized administrator of Digital Commons@Becker. For more information, please contact [email protected]. FOCUSED REVIEW published: 15 December 2008 doi: 10.3389/neuro.01.026.2008

Many specialists for suppressing cortical excitation

Andreas Burkhalter*

Department of Anatomy and Neurobiology, Washington University School of Medicine, St. Louis, MO, USA

Cortical computations are critically dependent on GABA-releasing neurons for dynamically balancing excitation with inhibition that is proportional to the overall level of activity. Although it is widely accepted that there are multiple types of interneurons, defining their identities based on qualitative descriptions of morphological, molecular and physiological features has failed to produce a universally accepted ‘parts list’, which is needed to understand the roles that interneurons play in cortical processing. A list of features has been published by the Petilla Interneurons Nomenclature Group, which represents an important step toward an unbiased classification of interneurons. To this end some essential features have recently been studied quantitatively and their association was examined using multidimensional cluster analyses. Edited by: These studies revealed at least 3 distinct electrophysiological, 6 morphological and 15 molecular Javier DeFelipe, Cajal Institute (CSIC), phenotypes. This is a conservative estimate of the number of interneuron types, which almost Spain certainly will be revised as more quantitative studies will be performed and similarities will Reviewed by: Kathleen S. Rockland, RIKEN Brain be defined objectively. It is clear that interneurons are organized with physiological attributes Science Institute, Japan representing the most general, molecular characteristics the most detailed and morphological Patrick R. Hof, Mount Sinai School features occupying the middle ground. By themselves, none of these features are sufficient of Medicine, USA to define classes of interneurons. The challenge will be to determine which features belong * Correspondence: together and how cell type-specific feature combinations are genetically specified.

Keywords: , interneurons, GABAergic neurons, inhibition

Introduction areas (Douglas and Martin, 2007a; Sugino et al., The cerebral cortex does not reveal its secrets eas- 2006). This elemental circuit consists of local Andreas Burkhalter studied Zoology at ily. To the naked eye it appears as little more than recurrent ­excitatory and inhibitory connections the University of Zürich where he worked with Rüdiger Wehner on ant navigation. a plain sheet of gray matter around the forebrain, within and between layers, which are capable of He obtained his Ph.D. with Michel and yet it is considered the cardinal achievement generating such diverse properties as direction Cuénod at the Brain Research Institute of evolution (Goldman-Rakic and Rakic, 1991). selectivity for visual motion in primary visual in Zürich. He then moved to the Caltech, All mammals depend on it, so much that in the cortex and eye movement commands in the fron- where he was a postdoctoral student words of David Hubel and ‘a man tal eye field (Douglas and Martin, 1991; Heinzle of David Van Essen and collaborator of Larry Katz. He then joined the group without a cortex is almost a vegetable, speech- et al., 2007). The key feature of this circuit is that of Robert Baughman at Harvard less, sightless [and] senseless’ (Hubel and Wiesel, most of its excitatory drive is generated by local University. In 1985 he joined the faculty 1979). The cortex’s ability to support sensory recurrent connections within cortex, whereas the at Washington University, where perception, reasoning, planning and execution connections that carry sensory inputs from the he is currently Professor. Burkhalter’s of behaviors is thought to be embodied in a outside world are relatively sparse (Ahmed et al., laboratory studies the functional organization of areas in mouse visual canonical circuit that is constructed from exci- 1994; Binzegger et al., 2004). The ingenuity of this cortex. tatory and inhibitory neurons with similar gene design is that weak afferent inputs are amplified [email protected] expression patterns across functionally distinct by local positive feedback, which selects stimu-

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 155 Burkhalter Cortical interneurons

lus features that resonate with response profiles visual cortex (Ramón y Cajal, 1899; described represented in the strengths of recurrent excita- in DeFelipe and Jones, 1988), long before it tory connections (Douglas and Martin, 2007b). became known that most of them use GABA as The potential risk of this organization is that it inhibitory transmitter (DeFelipe, 2002; Houser generates runaway excitation and limits stimulus et al., 1984). In the outgoing 19th century it was selectivity. An effective solution to this problem is thought that the axons of ‘short axon cells’ never to dynamically suppress excitation by inhibition leave gray areas or send axons through white that is proportional to the overall level of activity. matter (Lenhossék, 1895; described in Rakic, This role is played by aspiny, GABAergic ‘local 1976). More recently, however, we have learned circuit neurons’ (Rakic, 1976), which account of exceptions, which make long-range connec- for only 10–20% of cortical neurons (Beaulieu, tions and send axons through white matter tracts 1993), but make up for their underrepresentation (Higo et al., 2007; McDonald and Burkhalter, with a rich morphological, electrophysiological 1993; Peters et al., 1990; Tomioka and Rockland, and molecular diversity (Markram et al., 2004), 2007; Tomioka et al., 2005). However, the axons of suggesting that synaptic inhibition acts in many most interneurons that have been reconstructed different ways. This notion is strongly supported after Golgi staining or dye-filling terminate within by studies in the hippocampus, in which at least 16 the home area of the cell body. The list includes distinct types of interneurons have been identi- cells with chandelier, large basket, nest basket, fied that play specific roles in the timing and syn- small basket (i.e., clutch cells; Kisvárday, 1992), chronization of excitatory pyramidal cell activity double bouquet, bipolar, bitufted, Martinotti (Jinno et al., 2008; Klausberger and Somogyi, and neurogliaform (i.e., spiderweb or dwarf cells; 2008; Somogyi and Klausberger, 2005). Some Ramón y Cajal, 1904) morphologies (DeFelipe, of these interneurons generate different forms 2002; Fairén et al., 1984; Jones, 1984; Jones and of perisomatic inhibition and have been impli- Hendry, 1984; Lorente de No, 1992; Markram cated in epilepsy and anxiety disorders (Freund et al., 2004; Peters, 1984a,b; Somogyi and Cowey, and Katona, 2007). Others preferentially target 1984). Although these fanciful names have been pyramidal cell dendrites and appear to special- used for decades, it is still debated whether they ize in modulating synaptic integration (Somogyi designate distinct types, whether cortex contains and Klausberger, 2005). Compared to the hip- dozens or hundreds of different cell types or pocampus, classification of interneurons in the whether the morphological traits vary on a con- less neatly laminated cerebral cortex is lagging tinuum (Parra et al., 1998; Stevens, 1998). The behind. However, gene expression studies and only interneurons that were consistently classified developmental analyses of cell fate specification as distinct type are chandelier cells (ChCs), which strongly suggest that cerebral cortex contains single-mindedly innervate axon initial segments multiple distinct classes of interneurons, which of pyramidal cells and are easily identified by the are born at specific locations and times and that candle stick appearance of the terminal portion dysfunction of inhibitory systems are linked to of their axons (Figure 1) (Somogyi, 1977). These pathologies such as schizophrenia, and GABAergic cells are unusual in yet another way, Tourette syndrome (Butt et al., 2008; Huang et al., which enables them to depolarize pyramidal cells 2007; Kalanithi et al., 2005; Lewis and Gonzalez- because the chloride reversal potential of the axon Burgos, 2008; Sugino et al., 2006; Tabuchi et al., initial segment is more depolarized than that of 2007; Wonders and Anderson, 2006; Yuste, 2005). the cell soma (Szabadics et al., 2006). Recently, A clearer picture of these interneuron classes is significant progress has been made in classify- slowly beginning to emerge due to advances in ing interneurons with less clear cut properties by quantitatively describing morphological, physi- quantitative multivariate grouping using cluster ological and molecular features and applying analysis of various features of somatic, dendritic multidimensional cluster analyses for detecting and axonal morphologies including the distribu- significant associations between these features. tion of boutons. Applying this classification, basket Here, we discuss how some of these cell types cells can now be distinguished based on the pref- were identified in the cerebral cortex and how erential targeting of 15–30% of their synapses to they modulate excitation at different target loca- pyramidal cell somata (Figure 1) (Karube et al., Interneurons tions, times and levels of activation. 2004; Wang et al., 2002). Basket cells are further A morphologically and physiologically subdivided into large basket cells (LBCs) and small heterogeneous population of Different types of interneurons basket cells (SBCs) by characteristic relationships GABAergic neurons, which have mostly short locally projecting axons and rarely Morphological properties of soma size and the frequency of axonal branch- make long-range connections between Interneurons were first classified by Ramon y Cajal ing (Karube et al., 2004). Nonpyramidal cells with different cortical areas. as ‘short axon cells’ in Golgi studies of human <10% of their boutons terminating on somata, a

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 156 Burkhalter Cortical interneurons

spiking patterns. In these early studies, interneu- rons were identified as cells that respond to injec- tions of a brief pulse of depolarizing current with a train of narrow action potentials whose interspike interval remained constant throughout the dura- tion of the stimulus. Soon after, it was discovered that not all interneurons exhibit such fast spiking properties and that many cells show spike fre- quency adaptation, which is also found in pyrami- dal cells (Connors and Gutnick, 1990; Kawaguchi, 1993). These non-fast spiking interneurons were then further subdivided into cells in which the onset of the initial action potential is delayed (i.e., late spiking), a spike train begins with a brief burst (i.e., bursting nonpyramidal cells or low threshold neurons), or the steady-state response is irregu- lar (Karube et al., 2004; Kawaguchi and Kubota, 1997). More recently, all of these properties were integrated in a scheme that provides an even more detailed subclassification into fast spiking cells (with continuous, delayed and stuttering firing properties), non-adapting non-fast spiking neu- rons (with bursting and continuous firing prop- erties), adapting cells (with bursting, continuous and delayed firing properties), irregular spiking cells (which fire single or phasic bursts) and cells

Figure 1 | Six distinct morphological types of interneurons. Diagrams showing different with accelerating spike rates (Gupta et al., 2000; morphological types of interneurons (red), which were identified by cluster analyses of the Markram et al., 2004). This scheme turned out subcellular distribution of axonal projections onto somata and dendrites of postsynaptic pyramidal to be extremely useful for describing different neurons (gray). Chandelier cells (ChC), small basket cells (SBC), large basket cells (LBC), firing properties, however it was immediately neurogliaform cells (NGFC), double bouquet cells (DBC) and Martinotti cells (MC). For examples of recognized that a more definitive cell type clas- non-diagrammatic reconstructions of interneurons from rat somatosensory cortex and monkey prefrontal cortex see Krimer et al. (2005), Huang et al. (2007) and Ascoli et al. (2008). sification based on electrophysiological features requires unbiased cluster analyses. Such analyses have recently been performed in visual cortex of strong preference for dendrites and a high percent- anesthetized cats and in slices of monkey dorsola- age of projections toward the white matter side teral prefrontal cortex. These studies revealed fast of the soma, are classified as double bouquet cells spiking and non-fast spiking cells, which could be (DBCs) (Figure 1) (Karube et al., 2004). Dendrite further subclassified into cells with regular spik- targeting neurons with a high proportion of bou- ing (i.e., broad action potentials and spike fre- tons on the pial side of the cell body correspond quency adaptation) and intermediate spiking (IS) to Martinotti cells (MCs) (Figure 1) (Dumitriu (i.e., narrow action potentials and spike frequency et al., 2007). Finally, neurogliaform cells (NGCs) adaptation) properties (Figure 2) (Krimer et al., are distinguished by their high bouton density 2005; Nowak et al., 2003). Dye-filling of recorded and an axon arbor that is twice the size of the neurons showed unambiguously that ChCs have dendritic tree (Figure 1) (Karube et al., 2004). fast spiking properties (Krimer et al., 2005). Together, these quantitative studies demonstrate However, the electrophysiological classification that based on their axonal projection patterns did not cleanly distinguish between basket cells, there are at least six distinct morphological types DBCs and NGCs. Despite these shortcomings, the of neocortical interneurons. study by Krimer et al. (2005) shows a lax exclusion rule by which fast spiking properties are strongly Electrical properties are not tightly associated with basket and ChC morphologies, IS associated with distinct morphological properties are preferentially associated with cells cell types that have vertically oriented axons (i.e., DBCs, Unbiased cluster analyses The pioneering intracellular recordings by MCs), whereas NGCs never express fast spiking Statistical methods for classifying objects without preselection, McCormick et al. (1985) in rodent cerebral cortex properties (Figure 2). Although this classification using shared traits with set amounts have shown that pyramidal cells and interneurons represents progress, the discrimination of less of similarities. have different passive membrane properties and than a handful of distinct electrophysiological

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 157 Burkhalter Cortical interneurons

perform the grouping in a high-dimensional parameter space. A different approach was recently used by Ma et al. (2006), who studied a pre-selected group of somatostatin (SOM)-expressing interneurons in mouse somatosensory cortex, which revealed that these cells fall into two separate groups: stutter- ing cells, which project mainly to layer 4 and low threshold bursting MCs. Using a similar strategy, Dumitriu et al. (2007) have determined the syn- aptic properties of mouse visual cortical neurons, which express the green fluorescent protein in parvalbumin (PV)-, SOM- and neuropeptide Y (NPY) immunoreactive neurons and found that PV-expressing basket cells, SOM-expressing MCs

Figure 2 | Associations of electrophysiological and and NPY-expressing cells, with neither basket nor morphological properties of interneurons based on MC morphologies, can be distinguished based cluster analyses. Left side: Fast spiking (FS) properties on differences in the amplitude and kinetics (i.e., narrow action potentials, non-accommodating) are of spontaneous excitatory postsynaptic poten- strongly linked (indicated by thick lines) to chandelier tials of their inputs and spontaneous inhibitory cells (ChC) and basket cells (BC). Relatively few (indicated by thin lines) interneurons show regular synaptic potentials of their outputs (Figure 2). spiking (RS) properties (i.e., broad action potentials, Although, these studies are highly suggestive of accommodating), which are typically found in pyramidal distinct electrophysiological types, it is important cells. Intermediate spiking (IS) properties (i.e., narrow to note that the classification was done in pre- action potentials, accommodating) are strongly selected populations of interneurons and may associated with interneurons that have vertically oriented axons (e.g., double bouquet cells (DBC) and differ from classifications derived by unbiased Martinotti cells (MC)). Neurogliaform cells (NGC) never sampling of the entire population of GABAergic express FS properties. Right side: BCs receive high neurons. amplitude spontaneous excitatory postsynaptic potentials (sEPSP) and generate fast decaying Molecular properties are not tightly spontaneous inhibitory postsynaptic potentials (sIPSPs). Putative DBCs and/or NGCs receive medium amplitude associated with distinct morphological sEPSPs and more slowly decaying sIPSPs. Notice, and electrophysiological cell types because the cells have only been identified by the The vast majority of cerebral cortical interneurons absence of BC and MC morphologies, the associations use γ-amino butyric acid as inhibitory transmitter are indicated by dashed lines. MCs receive low and are therefore considered GABAergic (Houser amplitude sEPSPs and generate slowly decaying sIPSPs. et al., 1984). GABAergic neurons differ genetically from glutamatergic excitatory pyramidal neurons, classes is no match for the 14 types proposed from and comprise different populations that express qualitative evaluations by Markram et al. (2004). different sets of genes (Esumi et al., 2008; Sugino One way forward is to abandon subjective clas- et al., 2006). This heterogeneity has long been sification and use multidimensional hierarchical recognized by the expression and colocalization cluster analyses of quantitatively defined mor- of many different peptides and calcium bind- phological, electrical and molecular parameters ing proteins (DeFelipe, 1993). Considering the outlined by the Petilla Interneuron Nomenclature enormous variety of interneuron morphologies Group (Ascoli et al., 2008). The potential limita- (Lorente de No, 1992) and electrophysiological tion of this approach is that cluster algorithms properties (Markram et al., 2004) it is, however, tend to impose structure to data sets, whether it surprising that a short list of only six immu- exists or not. This highlights the importance of nocytochemically identified molecules, which choosing the correct sets and weights of param- includes PV, calretinin (CR), SOM, vasointestinal eters as well as cutoffs for linkage steps. Making polypeptide (VIP), cholecystokinin (CCK) and these determinations will require analytical tools the NMDA receptor binding protein, α-actinin-2, for finding the most significant clusters with some accounts for 95% of GABAergic neurons in the iterative procedure for randomly varying the cerebral cortex of VGAT (i.e., vesicular GABA weights of individual parameters and entire sets transporter)-Venus (i.e., fluorescent marker)- of features. For this analysis to succeed it will be expressing transgenic rats (Uematsu et al., 2008). important to study interneurons that have not A similar list, which in addition includes NPY been pre-selected for particular features and to and the enzyme choline acetyltransferase (ChAT)

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 158 Burkhalter Cortical interneurons

but excludes α-actinin-2, was found to account for >90% of GABA immunostained neurons in mouse visual cortex (Gonchar et al., 2008). These studies support the well established notion that the expression of ­calcium binding proteins and peptides are effective tools for classifying interneu- rons (DeFelipe, 1993, 1997). The problem with this approach, however, is that different markers are often coexpressed by the same interneurons and that the expressions depend on species, loca- tion, age and neuronal activity (DeFelipe, 1993; DeFelipe et al., 2002; Gonchar and Burkhalter, 1997; Gonchar et al., 2008; Jiao et al., 2006), mak- ing it a tricky classification tool. The most distinct group is represented by PV-immunoreactive neurons, that were discov- ered by Celio (1986) and which in rodents account for 40–50% of GABAergic neurons (Gonchar and Burkhalter, 1997; Gonchar et al., 2008; Kubota et al., 1994). A much higher percentage (74%) of PV neurons was found in macaque visual cortex (DeFelipe et al., 1999; Van Brederode et al., 1990). An important property of PV neurons is that they do not stain with antibodies against CR, SOM, VIP, CCK, NPY, ChAT and α-actinin-2 (Gonchar et al., 2008; Uematsu et al., 2008) (Figure 3). However, coexpression of calbindin (CB), cor- ticotropin releasing factor and substance P (SP) Figure 3 | Associations of molecular (protein, mRNA) and morphological properties of interneurons. Left immunoreactivity in subsets of PV neurons sug- side: Immunocytochemical expression of commonly gests that PV neurons are chemically heteroge- used markers [i.e., calbindin (CB), parvalbumin (PV), neous (Gonchar and Burkhalter, 1997; Karube calretinin (CR), neuropeptide Y (NPY), vasointestinal et al., 2004; Kubota et al., 1994; Kawaguchi and polypeptide (VIP), somatostatin (SOM), cholecystokinin Kubota, 1993; Lewis and Lund, 1990; Vruwink (CCK), choline acetyltransferase (ChAT), substance P (SP) corticotropin releasing factor (CRF)] in different et al., 2001) (Figure 3). Heterogeneity within the morphological types of interneurons [i.e., chandelier PV population was long known to exist by the cells (ChC), large basket cells (LBC), nest basket cells presence of basket cell and ChC morphologies, (NBC), small basket cells (SBC), bipolar cells (BPC), and the expression of non-accommodating fast bitufted cells (BTC), Martinotti cells (MC), neurogliaform spiking and burst firing properties (Blatow et al., cells (NGC)]. Right side: mRNAs expression in different types of interneurons. Notice that most substances are 2003; Chattopadhyaya et al., 2004; DeFelipe et al., coexpressed in multiple cell types and that the overlap in 1989; Zaitsev et al., 2005). mRNA expression is much more extensive than In contrast to the very limited overlap of PV immunostaining, suggesting that antibody staining is a with other immunocytochemically detected more discriminating classification tool. Thus, genotypic substances, the overlap between PV mRNA and and phenotypic molecular classification schemes are not interchangeable. mRNAs of other calcium binding proteins and peptides is much more extensive (Cauli et al., 1997; Toledo-Rodriguez et al., 2005; Wang et al., CCK constitute largely non-overlapping popu- 2002) (Figure 3), indicating that in many neurons lations (Gonchar and Burkhalter, 1997; Kubota PV mRNA is not translated. Thus, the genotypic et al., 1994). However, recent double- and triple and phenotypic molecular classification schemes immunolabeling studies in adult mouse cerebral are not interchangeable. cortex have found that, unlike in rat, SOM and Unlike PV-immunopositive neurons, CR are coexpressed in a substantial percentage PV-immunonegative cells coexpress a much (20–40%) of GABAergic neurons, which con- broader panel of calcium binding proteins and firms the notion that molecular expression pat- peptides (Gonchar and Burkhalter, 1997; Kubota terns are species-specific (Gonchar et al., 2008; et al., 1994). Initial double immunolabeling stud- Halabisky et al., 2006; Hof et al., 1999; Miyoshi ies in rat frontal and visual cortex have shown et al., 2007; Xu et al., 2006). Furthermore, these that SOM and CR, SOM and VIP, NPY and studies have shown extensive overlap between

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 159 Burkhalter Cortical interneurons

SOM and VIP, SOM and SP, VIP and ChAT, and basket cells (NBCs) and MCs (Toledo-Rodriguez NPY and CCK (Gonchar et al., 2008; Karube et al., 2005; Wang et al., 2002) (Figure 3). The mor- et al., 2004; Miyoshi et al., 2007; von Engelhardt phological diversity of VIP-expressing neurons is et al., 2007; Wolansky et al., 2007; Xu et al., 2006). mirrored by the electrophysiological heterogene- Importantly, however, the overlap of markers ity, which includes accommodating, bursting and for example in mouse visual cortex is incom- irregular spiking properties (Cauli et al., 2000; plete, and the immunocytochemical colocali- Kawaguchi and Kubota, 1996; Toledo-Rodriguez zation of VIP + NPY, VIP + CCK, SOM + CCK, et al., 2005; Uematsu et al., 2008). Whether the SOM + ChAT and NPY + ChAT has never been different morphological and electrophysiological observed, ­suggesting some kind of exclusion rule phenotypes represent distinct cell classes remains (Gonchar et al., 2008). A slightly different ­pattern to be established with unbiased cluster analy- exists in rat primary motor cortex, where Karube ses. However, the coexpression of several mol- et al. (2004) found a small amount of overlap ecules (i.e., VIP + NPY, VIP + SOM, VIP + CR, between VIP and CCK. Although considerable VIP + ChAT, VIP + SOM + CR; Gonchar et al., advances in cell type identification and fate map- 2008; Toledo-Rodriguez et al., 2005) that are ping (see, e.g., Wonders and Anderson, 2006) can known to influence neuronal excitabilities and be credited to antibody labeling of various pro- vascular tone (Baraban and Tallent, 2004; Cauli teins and peptides, the usefulness of this approach et al., 2004; Schwaller et al., 2002) suggest that VIP for cell type classification is limited due to the neurons play multiple different roles in cortical finite number of marker combinations that can processing and blood flow regulation. be tested in single cells. Clearly, more definitive SOM immunolabeling and mRNA was found answers can be expected from gene-expression in NBCs, DBCs, BTCs, BPCs with ascending axons profiling and analyses of the combinatorial code and MCs that do not express VIP, but tends to of transcription factor expression that may ulti- be excluded from LBCs and SBCs (Kawaguchi mately define different cell types (Flames et al., and Kubota, 1996, 1997; Toledo-Rodriguez et al., 2007; Nelson et al., 2006; Yuste, 2005). Based on 2005; Wang et al., 2002, 2004) (Figure 3). Most the most extensive quantitative immunohisto- SOM-expressing neurons have accommodating chemical studies presently available, we conclude firing properties, but stuttering and bursting fir- that there are at least three different subgroups of ing patterns have also been observed (Cauli et al., PV neurons and 12 subgroups of non-PV cells 1997, 2000; Goldberg et al., 2004; Kawaguchi and (Gonchar et al., 2008; Kubota et al., 1994; Lewis Kubota, 1996; Wang et al., 2002, 2004). Thus, and Lund, 1990; Vruwink et al., 2001). except for the preference for ascending axonal In contrast, classification based on mRNA projections, few differences were found between expression is much less effective. In fact, the SOM and VIP neurons and the identification of overlap of mRNA expression revealed by sin- subtypes of SOM neurons is tentative. Recently, gle cell multiplex RT-PCR is so extensive that it however cluster analyses of pre-selected GFP- shows colocalization of PV with CR, VIP, SOM, labeled SOM neurons in transgenic ‘GIN’ mice NPY and CCK, which has never been observed (Oliva et al., 2000) have identified four subtypes with immunostaining. Thus, message expression of SOM neurons with different passive membrane fails to classify interneurons even broadly into properties, spiking patterns, afterhyperpolariza- PV-positive and PV-negative groups (Cauli et al., tions and spontaneous excitatory postsynaptic 2000; Condé et al., 1994; Demeulemeester et al., currents (Halabisky et al., 2006). Furthermore, 1991; Gonchar and Burkhalter, 1997; Gonchar studies in two different lines of transgenic mice et al., 2008; Kawaguchi and Kubota, 1997; Kubota with GFP-labeled SOM neurons, Ma et al. (2006) et al., 1994; Toledo-Rodriguez et al., 2005; Wang have identified two additional subtypes, one with et al., 2002). Nevertheless, clear trends in mRNA ascending projections to layer 4 and stuttering fir- expression profiles have emerged, which show sig- ing properties, and another that resembles MCs nificantly lower probabilities for the coexpression and fires low threshold bursts of action potentials. of PV + SOM, PV + VIP and SOM + VIP (Toledo- Finally, Xu et al. (2006) have shown that SOM- Rodriguez et al., 2005). expressing MCs in mouse cerebral cortex consist VIP immunolabeling and mRNA was shown of two subtypes that either coexpress CR or are to be preferentially expressed in SOM-negative CR-immunonegative (Xu et al., 2006). Additional SBCs, DBCs, bipolar cells (BPCs) and bitufted support for the notion of multiple subtypes of MCs cells (BTCs) with vertically descending axons derives from fate mapping studies, which show (Kawaguchi and Kubota, 1996, 1997; Toledo- that CR-immunonegative/SOM-positive cells are Rodriguez et al., 2005) (Figure 3). In contrast, born before CR-immunopositive/SOM-negative VIP mRNA tends to be excluded from LBCs, nest neurons (Miyoshi et al., 2007). Thus, at the very

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 160 Burkhalter Cortical interneurons

least SOM-expressing neurons contain four and play a role in facilitating interactions between dif- perhaps as many as six different subtypes. ferent cortical areas and in visual attention (Fries Double- and triple immunolabeling studies in et al., 2008; Galarreta and Hestrin, 2002; Galarreta rat and mouse cerebral cortex have shown that et al., 2004; Womelsdorf et al., 2007). Thus, sup- significant numbers of interneurons express CR pression of inhibition by endocanabinoids may but do not coexpress SOM or VIP (Gonchar et al., provide a mechanism by which CCK neurons 2008; Uematsu et al., 2008). Recently, these cells modulate the synchrony of network oscillations were shown to coexpress α-actinin-2, exhibit (Freund and Katona, 2007). neurogliaform morphologies, have late spik-

ing properties and generate GABAA and GABAB Role of interneurons in cortical receptor mediated inhibition (Tamás et al., 2003; circuits Uematsu et al., 2008), suggesting that PV-negative, There is a long way to go to understand the role α-actinin-2-positive GABAergic neurons repre- of every type of interneuron in cortical process- sent a distinct type of interneuron. ing. However, with the development of new tools CCK immunolabeling was found in the finish line has moved closer (Kuhlman and GABAergic neurons of rat motor and mouse Huang, 2008). Impressive progress has already visual cortex (Gonchar et al., 2008; Karube et al., been made in regard to PV neurons and two 2004; Kawaguchi and Kondo, 2002; Kubota and subtypes of SOM neurons, which are the main Kawaguchi, 1997). Studies in rat frontal cortex focus here. showed that some of these neurons in addition contain VIP and CR but that they lack NPY Parvalbumin-expressing fast spiking cells (Karube et al., 2004). This expression pattern Several anatomical and physiological studies in differs from mouse visual cortex, in which CCK visual, somatosensory and motor cortex of juve- neurons were found to contain CR and NPY but nile and young adult rats and cats have shown do not stain for VIP (Gonchar et al., 2008). In that PV-immunoreactive, fast spiking neurons contrast to the molecular heterogeneity, CCK play a role in mediating feedforward inhibition neurons appear to be physiologically similar in thalamocortical, interlaminar and interareal in that they uniformly express non-fast spik- circuits (Angulo et al., 1999; Beierlein et al., ing, irregular firing properties (Galarreta et al., 2003; Gonchar and Burkhalter, 1999; Porter 2004; Karube et al. 2004; Kawaguchi and Kondo, et al., 2001; Reyes et al., 1998; Thomson and 2002). Morphologically, however, CCK neurons Deuchars, 1997; Thomson et al., 2002; West et al., are diverse, exhibiting small cell bodies with 2006). How this feedforward inhibitory circuit bipolar and bitufted dendritic arbors as well as works is most completely understood in the rat axonal features of SBCs and LBCs (Karube et al., thalamocortical somatosensory pathway. In this 2004; Kubota and Kawaguchi, 1997). Large CCK system, PV neurons receive input from thalamo- immunoreactive basket cells in rat somatosen- cortical axons (Staiger et al., 1996) and synapse sory cortex were shown to innervate pyramidal onto layer 4 spiny neurons, which receive direct cell somata, and similar to their PV-expressing connections from the thalamus (White, 1978) counterparts are thought to provide perisomatic (Figure 4). Thus, thalamocortical input evokes a inhibition (Bodor et al., 2005). However, unlike brief excitatory response in spiny neurons, which

PV neurons which express α1 GABAA receptors, is readily opposed by inhibition from PV neurons GABAergic transmission at CCK synapses is medi- (Agmon and Connors, 1992; Zhu and Connors,

ated mainly by α2 receptors (Freund and Katona, 1999). In the intact rabbit somatosensory cor- 2007). In addition, unlike the axon terminals of tex, this disynaptic inhibition very effectively fast spiking PV neurons, CCK immunoreactive suppresses sustained firing of spiny neurons to presynaptic terminals in contact with pyrami- input from the thalamus (Swadlow, 2003). The dal neurons have a much higher failure rate and effectiveness of this suppression has multiple express type 1 cannabinoid receptors, which when reasons. First, fast spiking neurons (i.e., puta- activated by the release of endogenous cannabi- tive PV neurons) receive very powerful inputs noids from postsynaptic pyramidal cells, suppress from a large number of thalamocortical axons, inhibition (Bodor et al., 2005; Galarreta et al., which makes them more excitable than spiny 2008; Hill et al., 2007). Interestingly, both PV cells (Cruikshank et al., 2007; Gabernet et al., and CCK neurons are electrically coupled by gap 2005). Second, fast spiking neurons evoke large junctions, which are responsible for synchroniz- inhibitory responses in spiny cells (Cruikshank ing the activity within each population as well as et al., 2007; Gabernet et al., 2005). Finally, the among interconnected pyramidal neurons. Such transmission time between fast spiking neurons synchronous network oscillations are thought to and spiny cells and the dynamics of excitatory

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 161 Burkhalter Cortical interneurons

Fast spiking PV-immunoreactive interneurons in layer 2/3 of rat and cat visual cortex receive strong excitatory input from layer 4 and provide feedforward inhibition to neighboring layer 2/3 pyramidal cells (Dantzker and Callaway, 2000; Thomson et al., 2002). Paired recordings from fast spiking neurons (i.e., putative PV neurons) and pyramidal cells in layer 2/3 of rat visual cor- tex have shown that these feedforward inhibi- tory connections are preferentially received by pyramidal neurons that project back to the fast spiking neuron that provides the inhibitory input (Yoshimura and Callaway, 2005). This demon- strates that inhibition by fast spiking neurons is applied to the same pyramidal cells that supply their excitation, which is the hallmark of feedback inhibitory circuits (Figure 4). Importantly, these feedback inhibitory circuits are formed prefer- entially between fast spiking cells and layer 2/3 pyramidal cells, which are reciprocally connected and share common excitatory feedforward and horizontal inputs from layer 4 and layer 2/3 neu- rons (Yoshimura and Callaway, 2005; Yoshimura et al., 2005). These results suggest that specific sets of fast spiking neurons are owned by indi- vidual excitatory subnetworks, and that these Figure 4 | Inhibitory networks involving fast spiking parvalbumin (PV)-expressing neurons PV-expressing interneurons play a role in setting in thalamocortical, interlaminar and interareal cortical circuits. Feedforward excitatory the gain and timing of responses within neuro- thalamocortical inputs to spiny neurons (▲) and fast spiking interneurons (•) in layers 2–4 (a). nal populations that may share similar stimulus Inputs to interneurons are stronger (large arrowheads) than inputs to spiny cells. PV neurons provide strong (large rectangular endings) feedforward inhibition (b) to spiny cells. Feedback selectivities (Ohki and Reid, 2007). inhibition (c) results from PV neurons that are excited by the same spiny neurons that they inhibit. Unlike the subnetwork-specific feedforward These reciprocally connected spiny neuron/PV neuron pairs share common inputs (e.g., excitatory input from layer 4 to reciprocally inter- cells in layer 4 from thalamus or cells in layer 2/3 from layer 4) creating recurrent excitatory (d) connected layer 2/3 pyramidal cells and fast spik- and inhibitory subnetworks (contained within blue shaded boxes). ‘Lateral’ inhibition (e) of these ing neurons, excitatory feedback input from layer subnetworks results from PV neurons that are driven by excitatory feedback connections (f) from outside the subnetworks (e.g., by layer 5 to layer 2/3 connections or feedback from higher cortical 5 in rat visual cortex is not specifically targeted to areas). Notice that ‘lateral’ inhibition is weaker (small rectangular endings) than feedforward and individual layer 2/3 subnetworks, but is distrib- feedback inhibition and impinges on multiple subnetworks. uted to pyramidal cells and fast spiking neurons, regardless of whether these cells are directly inter- connected (Yoshimura and Callaway, 2005). This thalamocortical excitation is extremely fast, so means that excitatory feedback connections have that fast spiking neurons reach firing thresh- access to several different excitatory subnetworks old before feedforward inhibition from nearby and activate feedforward inhibition to pyrami- interneurons can take effect (Cruikshank et al., dal cells, which do not provide excitation to fast 2007; Gabernet et al., 2005). Together, these spiking cells. Yoshimura and Callaway (2005) properties enable fast spiking neurons to exert have shown that such non-reciprocated ‘lateral’ tight inhibitory control over the timing of inhibition (Figure 4) is driven from outside the responses of spiny neurons to sensory inputs. subnetwork and is much weaker than feedforward Because thalamocortical synapses onto fast spik- and feedback inhibition originating from within ing neurons are depressed by repetitive activa- the subnetwork. Similar to layer 5, higher cortical tion, the feedforward inhibition is weaker at high areas are an important source of excitatory feed- stimulation frequencies (Gabernet et al., 2005). back input to layer 2/3 of primary visual cortex As a result, the integration window of spiny neu- (Coogan and Burkhalter, 1993). The main inhibi- rons becomes broader and responses shift in an tory targets of this input are PV-immunoreactive activity-dependent fashion from high temporal neurons (Gonchar and Burkhalter, 1999, 2003; resolution, which is optimal for coincidence Yamashita et al., 2003), suggesting that these cells detection, to low temporal resolution, which is provide weak ‘lateral’ inhibition to pyramidal better suited for sensory integration. neurons (Yoshimura and Callaway, 2005). Indeed,

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 162 Burkhalter Cortical interneurons

studies in rat and mouse visual cortex have shown recent studies in rat somatosensory cortex that in interareal feedback circuits, excitation is have shown that one of these cell types, SOM- opposed by much weaker inhibition than in inter- immunoreactive MCs, are selectively activated areal feedforward circuits, suggesting that the bal- by action potential trains from single pyrami- ance between excitation and inhibition is tipped dal cells and distribute inhibition to several toward depolarization, which may increase the neighboring pyramidal neurons (Kapfer et al., gain of sensory inputs to primary visual cortex by 2007; Silberberg and Markram, 2007). In sharp top-down connections from higher visual areas contrast, activation of recurrent inhibition by (Dong et al., 2004a; Shao and Burkhalter, 1996). fast spiking PV-expressing neurons, whose con- That fast spiking interneurons play a key role in nectivity rate with pyramidal cells is very high balancing recurrent excitation and inhibition, was (Thomson and Lamy, 2007), requires inputs directly demonstrated in ferret visual cortex (Shu from more than a single pyramidal cell (Kapfer et al., 2003). et al., 2007; Silberberg and Markram, 2007; Wang et al., 2002). The low threshold of MC activation Role of somatostatin-expressing neurons in and its high inhibitory impact is due to the high feedforward inhibition degree of convergence, divergence and reciproc- Until recently, it was widely held that fast spiking, ity of connections with pyramidal neurons. In PV-immunoreactive neurons are solely respon- fact, the connectivity pattern is so ingenious that sible for thalamocortical feedforward inhibition synchronous excitatory input from two pyrami- (Daw et al., 2007; Sun et al., 2006). Recordings dal neurons leads to a supralinear increase in the from GFP-labeled SOM-immunopositive neurons number of recruited MCs (Kapfer et al., 2007). (CB-immunonegative, NPY-immunonegative) in A distinctive feature of the inhibitory feedback the somatosensory barrel cortex of the X94 line circuit between pyramidal neurons and MCs is of transgenic mice, however, show that cells in that inhibition of pyramidal cells is targeted to layer 5 with dense projections to layer 4 receive the distal dendritic tufts in layer 1, suggesting direct thalamocortical input, suggesting that that it is not involved in gain control of firing they play a role in feedforward inhibition of but in the modulation of dendritic excitability thalamo-recipient spiny neurons in layer 4 (Tan (Silberberg and Markram, 2007). Because layer 1 et al., 2008). Unlike PV neurons, whose synaptic is an important target of interareal feedback con- responses to repetitive thalamocortical inputs nections (Coogan and Burkhalter, 1993; Dong are depressing and transient, responses of GFP- et al., 2004b), inhibition by MCs may therefore expressing X94 SOM neurons are strongly facili- modulate the impact of global top-down influ- tating and produce sustained firing (Tan et al., ences on the processing of local information 2008). As a result of these dynamic properties, within the columnar network. PV neuron-mediated feedforward inhibition is most effective at low frequencies, whereas SOM Conclusions neuron-mediated inhibition is strongest at high An impressive body of work on the morphologi- frequencies. Because PV neurons preferentially cal, electrophysiological and molecular proper- innervate somata whereas SOM neurons are pref- ties of qualitatively pre-selected populations of erentially targeted to dendrites (DiCristo et al., interneurons suggests that cerebral cortex con- 2004; Dumitriu et al., 2007; Karube et al., 2004), tains multiple types of interneurons. Recent, the frequency-dependent switch reflects a transi- unbiased quantitative analyses have revealed tion from somatic to dendritic inhibition, thereby that axon projections are powerful classification gaining control over the integration of sensory parameters for identifying interneurons into inputs with recurrent intracortical excitation. LBCs and SBCs, ChCs, NGCs, DBCs and MCs. However, quantitative definitions of electrical Role of Martinotti cells in feedback excitabilities were found to be only loosely asso- inhibition ciated with specific morphological properties, Paired recordings have shown that many differ- which means that fast spiking properties are not ent types of interneurons including fast-, late-, exclusively expressed by ChCs and basket cells low threshold-, irregular- and burst spiking cells and spike accommodation is not a unique prop- (Beierlein et al., 2003; Blatow et al., 2003; Chu erty of DBCs, MCs and NGCs. Similarly, cluster et al., 2003; Galarreta et al., 2004; Gupta et al., analyses have shown that electrical parameters 2000; Reyes et al., 1998; Thomson and Deuchars, only partially match molecular expression pat- 1997; Thomson et al., 2002), receive inputs from terns, such that fast spiking properties are not pyramidal neurons and generate feedback inhi- only found in PV-expressing neurons and spike bition in the same cells that excites them. Two accommodation is not a surrogate property of

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 163 Burkhalter Cortical interneurons

all PV-negative neurons. Immunolabeling studies that this goal may be achieved through unbiased have shown that calcium binding protein, peptide multidimensional cluster analyses of morpho- and transmitter synthesizing enzyme expressions logical, electrical and molecular features listed extensively overlap and that the staining patterns by the Petilla Interneurons Nomenclature Group vary across development, location and species. (Ascoli et al., 2008). Once the players are defined, However, in adult mouse visual cortex, PV is never an important next step will be to develop a better coexpressed with SOM, CR, VIP, ChAT, NPY and understanding of how local inhibitory circuits are CCK. Further, VIP does not associate with NPY integrated into interareal cortical networks. and CCK. SOM does not colocalize with CCK and ChAT, and NPY is absent in ChAT-expressing Acknowledgements cells. Importantly, these exclusion rules do not I would like to thank my colleagues Hongwei Dong, apply to mRNAs, whose expressions are much Enquan Gao, Yuri Gonchar, Ekaterina Hatch, more promiscuous. Thus, genotypic and phe- Jeanne M. Nerbonne, Katia Valkova, Quanxin notypic molecular classification schemes are not Wang and Jun Xiao for their collaboration and interchangeable. Clearly, we are far from being discussion. Financial support was generously pro- able to unequivocally identify different types of vided by NIH grants EY05935, EY016184 and the interneurons. There is however, great promise McDonnell Center for Higher Brain Function.

References Whittington, M. A., Caputi, A., and perisomatic GABAergic innervation DeFelipe, J. (1997). Types of neurons, Agmon, A., and Connors, B. W. (1992). Monyer, H. (2003). A novel network in primary visual cortex during a synaptic connections and chemical Correlation between intrinsic firing of multipolar bursting interneurons postnatal critical period. J. Neurosci. characteristics of cells immunoreac- patterns and thalamocortical synaptic generates theta frequency oscillations 24, 9598–9611. tive for calbindin-D28K, parvalbu- responses of neurons in mouse barrel in neocortex. Neuron 38, 805–817. Chu, Z., Galarreta, M., and Hestrin, S. min and calretinin in the neocortex. cortex. J. Neurosci. 12, 319–329. Bodor, A. L., Katona, I., Nyíri, G., (2003). Synaptic interactions of late- J. Chem. Neuroanat. 14, 1–19. Ahmed, B., Anderson, J. C., Douglas, R. J., Mackie, K., Ledent, C., Hájos, N., and spiking neocortical neurons in layer 1. DeFelipe, J. (2002). Cortical interneurons: Martin, K. A. C., and Nelson, J. C. Freund, T. F. (2005). Endocannabinoid J. Neurosci. 23, 96–102. from Cajal to 2001. Prog. Brain Res. (1994). Polyneuronal innervation of signaling in rat somatosensory cor- Condé, F., Lund, J. S., Jacobowitz, D. M., 136, 215–238. spiny stellate neurons in cat visual cor- tex: laminar differences and involve- Baimbridge, K. G., and Lewis, D. A. DeFelipe, J., Alonso-Nanclares, L., and tex. J. Comp. Neurol. 341, 39–49. ment of specific interneuron types. J. (1994). Local circuit neurons immuno- Arellano, J. L. (2002). Microstructure Angulo, M. C., Rossier, J., Staiger, J. F., and Neurosci. 25, 6845–6856. reactive for calretinin, calbindin D-28k of the neocortex: comparative aspects. Audinat, E. (1999). Postsynaptic gluta- Butt, S. J. B., Sousa, V. H., Fuccillo, M. V, or parvalbumin in monkey prefrontal J. Neurocytol. 31, 299–316. mate receptors and integrative prop- Hjerling-Leffler, J., Miyoshi, G., cortex: distribution and morphology. DeFelipe, J., González-Albo, M. C., Del erties of fast-spiking interneurons in Kimura, S., and Fishell, G. (2008). The J. Comp. Neurol. 341, 95–116. Río, R., and Elston, G. N. (1999). the rat neocortex. J. Neurophysiol. 82, requirement of Nkx2-1 in the tempo- Connors, B. W., and Gutnick, M. J. (1990). Distribution and patterns of con- 1295–1302. ral specification of cortical interneu- Intrinsic firing patterns of diverse neo- nectivity of interneurons containing Ascoli, G. A., Alonso-Nanclares, L., ron subtypes. Neuron 59, 722–732. cortical neurons. Trends Neurosci. 13, calbindin, calretinin, and parvalbumin Anderson, S. A., Barrinuevo, G., Cauli, B., Audinat, E., Lambolez, B., 99–104. in visual areas of the occipital and tem- Benavides-Piccone, R., Burkhalter, A. Angulo, M.C., Ropert, N., Tsuzuki, K., Coogan, T. A., and Burkhalter, A. (1993). poral lobes of the macaque monkey. et al. (2008). Petilla terminology: Hestrin, S., and Rossier, S. (1997). Hierarchical organization of areas J. Comp. Neurol. 412, 515–526. nomenclature of features of GABAergic Molecular and physiological diver- in rat visual cortex. J. Neurosci. 13, DeFelipe, J., Hendry, S. H. C., and interneurons of the cerebral cortex. sity of cortical nonpyramidal cells. J. 3749–3772. Jones, E. G. (1989). Visualization of Nat. Rev. Neurosci. 9, 1–12. Neurosci. 17, 3894-3906. Cruikshank, S. J., Lewis, T. J., and Connors, chandelier cell axons by parvalbumin Baraban, S. C., and Tallent, M. K. Cauli, B., Porter, J. T., Tsuzuki, K., B. W. (2007). Synaptic basis for intense immunoreactivity in monkey cerebral (2004). Interneurons diversity Lambolez, B., Rossier, J., Quenet, B., thalamocortical activation of feedfor- cortex. Proc. Natl. Acad. Sci. USA 86, series: Interneuronal neuropeptides and Audinat, E. (2000). Classification ward inhibitory cells in neocortex. Nat. 2093–2097. – endogenous regulators of neuro- of fusiform neocortical interneurons Neurosci. 10, 462–468. DeFelipe, J., and Jones, E. G. (1988). Cajal nal excitability. Trends Neurosci. 27, based on unsupervised clustering. Proc. Dantzker, J. L., and Callaway, E. M. (2000). on the Cerebral Cortex. New York, 135–142. Natl. Acad. Sci. USA 97, 6144–6149. Laminar sources of synaptic input to Oxford University Press, pp. 139–146. Beaulieu, C. (1993). Numerical data on Cauli, B., Tong, X.-K., Rancillac, A., cortical inhibitory interneurons and Demeulemeester, H., Arckens, L., neocortical neurons in adult rat with Serluca, N., Lambolez, B., Rossier, J., pyramidal neurons. Nat. Neurosci. 3, Vandesande, F., Orban, G. A., Heizmann, special reference to the GABA popula- and Hamel, E. (2004). Cortical GABA 701–707. C. W., and Pochet, R. (1991). Calcium tion. Brain Res. 609, 284–292. interneurons in neurovascular cou- Daw, M. I., Ashby, M. C., and Isaac, J. T. binding proteins and neuropeptides Beierlein, M., Gibson, J. R., and pling: relays for subcortical vasoactive R. (2007). Coordinated developmen- as molecular markers of GABAergic Connors, B. W. (2003). Two dynami- pathways. J. Neurosci. 24, 8940–8949. tal recruitment of latent fast spiking interneurons in the cat visual cortex. cally distinct inhibitory networks in Celio, M. R. (1986). Parvalbumin in most interneurons in layer IV barrel cortex. Exp. Brain Res. 84, 538–544. layer 4 of the neocortex. J. Neurophysiol. γ-aminobutyric acid-containing neu- Nat. Neurosci. 10, 453–461. DiCristo, G., Di Cristo, G., Wu, C., 90, 2987–3000. rons of the rat cerebral cortex. Science DeFelipe, J. (1993). Neocortical neurons Chattopadhyaya, B., Ango, F., Binzegger, T., Douglas, R. J., and 231, 995–997. diversity: chemical heterogeneity Knott, G., Welker, E., Svoboda, K., Martin, K. A. C. (2004). A quantitative Chattopadhyaya, B., Di Cristo, G., revealed by colocalization studies of and Huang, Z. J. (2004). Subcellular map of the circuit of primary visual Higashiyama, H., Knott, G. W., classic neurotransmitters, neuropep- domain-restricted GABAergic inner- cortex. J. Neurosci. 24, 8441–8453. Kuhlman, S. J., Welker, E., and tides, calcium-binding proteins, and vation in primary visual cortex in Blatow, M., Rozov, A., KAtona, I., Huang, J. Z. (2004). Experience and cell surface molecules. Cereb. Cortex the absence of sensory and thalamic Hormuzi, S. G., Meyer, A., activity-dependent maturation of 3, 273–289. inputs. Nat. Neurosci. 7, 1184–1186.

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 164 Burkhalter Cortical interneurons

Dong, H., Shao, Z., Nerbonne, J. M., and interneurons expressing cannabinoid Hof, P. R., Glezer, I., Conde, F., Flagg, R. Kawaguchi, Y., and Kondo, S. (2002). Burkhalter, A. (2004a). Differential receptors. J. Neurosci. 24, 9700–9778. A., Rubin, M. B., Nimchinsky, E. A., Parvalbumin, somatostatin and cho- depression of inhibitory synaptic Galarreta, M., Erdélyi, F., Szabo, G., and and Vogt Weisenhorn, D. M. (1999). lecystokinin as chemical markers for responses in feedforward and feed- Hestrin, H. (2008). Cannabinoid sen- Cellular distribution of the calcium- specific GABAergic interneuron types back circuits between different areas of sitivity and synaptic properties of 2 binding proteins parvalbumin, calbin- in the rat frontal cortex. J. Neurocytol. mouse visual cortex. J. Comp. Neurol. GABAergic networks in the neocortex. din, and calretinin in the neocortex of 31, 277–287. 475, 361–373. Cereb. Cortex 18, 2296–2305. mammals: phylogenetic and develop- Kawaguchi, Y., and Kubota, Y. (1993). Dong, H., Wang, Q., Valkova, K., Galarreta, M., and Hestrin, S. (2002). mental patterns. J. Chem. Neuroanat. Correlation of physiological sub- Gonchar, Y., and Burkhalter, A. Electrical and chemical synapses 16, 77–116. groupings of nonpyramidal cells with (2004b). Experience-dependent devel- among parvalbumin fast-spiking Houser, C. R., Vaughn, J. E., Hendry, S. H. C., specific physiological and morpholog- opment of feedforward ad feedback GABAergic interneurons in mouse Jones, E. G., and Peters, A. (1984). ical characteristics in rat frontal cortex. circuits between lower and higher neocortex. Proc. Natl. Acad. Sci. USA GABA neurons in the cerebral cortex. J. Neurophysiol. 70, 387–396. areas of mouse visual cortex. Vision 99, 12438–12443. In Cerebral Cortex, Vol. 2, E. G. Jones Kawaguchi, Y., and Kubota, Y. (1996). Res. 44, 3389–3440. Goldberg, J. H., Lacefield, C. O., and Yuste, R. and A. Peters, eds (New York, NY, Physiological and morphological iden- Douglas, R. J., and Martin, K. A. C. (1991). (2004). Global dendritic calcium spikes Plenum Press), pp. 63–89. tification of somatostatin- or vasoac- A functional microcircuit for cat visual in mouse layer 5 low threshold spiking Huang, Z. J., Di Cristo, G., and Ango, F. tive intestinal polypeptide-containing cortex. J. Physiol. 440, 735–769. interneurones: implications for control (2007). Development of GABA cells among GABAergic cell subtypes Douglas, R. J., and Martin, K. A. C. (2007a). of pyramidal cell bursting. J. Physiol. innervation in the cerebral and cer- in rat frontal cortex. J. Neurosci. 16, Mapping the matrix: the ways of neo- 558, 456–478. ebellar cortices. Nat. Rev. Neurosci. 8, 2701–2715. cortex. Neuron 56, 226–238. Goldman-Rakic, P. S., and Rakic, P. (1991). 673–686. Kawaguchi, Y., and Kubota, Y. (1997). Douglas, R. J., and Martin, K. A. C. Preface: cerebral cortex has come of Hubel, D. H., and Wiesel, T. N. (1979). GABAergic cell subtypes and their (2007b).Recurrent neuronal cir- age. Cereb. Cortex 1, V. Brain mechanisms of vision. Sci. Am. synaptic connections in rat frontal cuits in the neocortex. Curr. Biol. 17, Gonchar, Y., and Burkhalter, A. (1997). 241, 130–144. cortex. Cereb. Cortex 7, 476–486. R496–R500. Three families of GABAergic neurons Jiao, Y., Zhang, C., Yanagawa, Y., and Kisvárday, Z. (1992). GABAergic networks Dumitriu, D., Cossart, R., Huang, J., and in rat visual cortex. Cereb. Cortex 11, Sun, Q. Q. (2006). Major effects of of basket cells in the visual cortex. Prog. Yuste, R. (2007). Correlation between 347–358. sensory experiences on the neocorti- Brain Res. 90, 385–405. axonal morphologies and synaptic Gonchar, Y., and Burkhalter, A. (1999). cal inhibitory circuits. J. Neurosci. 26, Klausberger, T., and Somogyi, P. (2008). input kinetics of interneurons from Differential subcellular localization of 8691–8701. Neuronal diversity and temporal mouse visual cortex. Cereb. Cortex forward and feedback interareal inputs Jinno, S., Klausberger, T., Martin, L. F., dynamics: The unity of hippocam- 17, 81–91. to parvalbumin expressing GABAergic Dalezios, Y., Roberts, J. D. B., pal circuit operations. Science 321, Esumi, S., Wu, S.-X., Yanagawa, Y., neurons in rat visual cortex. J. Comp. Fuentealba, P., Bushong, E. A., Henze, D., 53–57. Obata, K., Sugimoto, Y., and Neurol. 406, 346–360. Buszáki, G., and Somogyi, P. (2008). Krimer, L. S., Zaitsev, A. V., Czanner, G., Tamamaki, N. (2008). Method for Gonchar, Y., and Burkhalter, A. (2003). Neuronal diversity in GABAergic long- Kröner, S., Gonzalez-Burgos, G., single-cell microarray analysis and Distinct GABAergic targets of feed- range projections from the hippocam- Povysheva, N. V., Iyengar, S., application to gene-expression profil- forward and feedback connections pus. J. Neurosci. 27, 8790–8804. Barrinuevo, G., and Lewis, D. A. ing of GABAergic neuron progenitors. between lower and higher areas of Jones, E. G. (1984). Neurogliaform or (2005). Cluster analysis-based physi- Neurosci. Res. 60, 439–451. rat visual cortex. J. Neurosci. 23, spiderweb cells. In Cerebral Cortex, ological classification and morpholog- Fairén, A., DeFelipe, J., and Regidor, J. 10904–10912. Vol. 1, A. Peters and E. G. Jones, eds ical properties of inhibitory neurons (1984). Nonpyramidal neurons: Gonchar, Y., Wang, Q., and Burkhalter, A. (New York, NY, Plenum Press), pp. in layers203 of monkey dorsolateral general account. In Cerebral Cortex, (2008). Multiple distinct subtypes of 409–417. prefrontal cortex. J. Neurophysiol. 94, Vol. 1, A. Peters and E. G. Jones, eds GABAergic neurons in mouse visual Jones, E. G., and Hendry, S. C. H. (1984). 3009–3022. (New York, NY, Plenum Press), pp. cortex identified by triple immunos- Basket cells. In Cerebral Cortex, Kubota, Y., Hattori, R., and Yui, Y. (1994). 201–245. taining. Front. Neuroanat. 1, 1–11. Vol. 1, A. Peters and E. G. Jones, eds Three distinct subpopulations of Flames, N., Pla, R., Gelman, D. M., Gupta, A., Wang, Y., and Markram, H. (New York, NY, Plenum Press), pp. GABAergic neurons in rat frontal Rubenstein, J. L. R., Puelles, L., and (2000). Organizing principles for a 309–334. agranular cortex. Brain Res. 649, Marin, O. (2007). Delineation of mul- diversity of GABAergic interneurons Kalanithi, P. S., Zheng, W., Katoka, Y., 159–173. tiple subpallial progenitor domains and synapse in the neocortex. Science DiFiglia, M., Grantz, H., Saper, C. B., Kubota, Y., and Kawaguchi, Y. (1997). Two by the combinatorial expression of 287, 273–278. Schwartz, M. L., Leckman, J. F., and distinct subgroups of cholecystokinin transcriptional codes. J. Neurosci. 27, Halabisky, B, Shen, F., Huguenard, J. R., Vaccarino, F. M. (2005). Altered immunoreactive cortical interneurons. 9682–9695. and Prince, D. A. (2006). parvalbumin-positive neuron distri- Brain Res. 752, 175–183. Freund, T. F., and Katona, I. (2007). Electrophysiological classification of bution in basal ganglia of individuals Kuhlman, S. J., and Huang, Z. J. (2008). Perisomatic inhibition. Neuron 56, somatostatin-positive interneurons with Tourette syndrome. Proc. Natl. High-resolution labeling and func- 33–42. in mouse sensorimotor cortex. J. Acad. Sci. USA 102, 13307–13312. tional manipulation of specific neuron Fries, P., Womelsdorf, T., Oostenveld, R., Neurophysiol. 96, 83–845. Kapfer, C., Glickfield, L. L. Atallah, B. V., types in mouse brain by Cre-activated and Desimone, R. (2008). The effects Heinzle, J., Hepp, K., and Martin, K. A. and Scanziani, M. (2007). Supralinear viral gene expression. PLoS ONE 3, of visual stimulation and selective (2007). A microcircuit model for increase of recurrent inhibition during 1–11. visual attention on rhythmic neuronal the frontal eye fields. J. Neurosci. 27, sparse activity in the somatosensory Lenhossék, M. V. (1895). Der feine Bau des synchronization in macaque area V4. 9341–9353. cortex. Nat. Neurosci. 10, 743–753. Nervensystems. Berlin, Fischer. J. Neurosci. 28, 4823–4835. Higo, S., Udaka, N., and Tamamaki, N. Karube, F., Kubota, Y., and Kawaguchi, Y. Lewis, D. A., and Gonzalez-Burgos, G. Gabernet, L., Jadhav, S. P., Feldman, D. E., (2007). Long-range GABAergic pro- (2004). Axon branching and synaptic (2008). Neuroplasticity of neo- Carandini, M., and Scanziani, M. jection neurons in the cat neocortex. bouton phenotypes in GABAergic cortical circuits in schizophrenia. (2005). Somatosensory integration J. Comp. Neurol. 503, 421–431. nonpyramidal cell subtypes. J. Neuropharmacology 33, 141–165. controlled by dynamic thalamocorti- Hill, E. L., Gallopin, T., Ferezou, I., Cauli, B., Neurosci. 24, 2853–2865. Lewis, D. A., and Lund, J. S. (1990). cal feed-forward inhibition. Neuron Rossier, J., Schweitzer, P., and Lambolez, Kawaguchi, Y. (1993). Groupings of non- Heterogeneity of chandelier neurons 48, 315–327. B. (2007). Functional CB1 receptors pyramidal and pyramidal cells with in monkey neocortex: corticotropin- Galarreta, M., Erdélyi, F., Szabo, G., and are broadly expressed in neocortical specific physiological and morpho- releasing factor- and parvalbumin- Hestrin, S. (2004). Electrical coupling GABAergic and glutamatergic neurons. logical characteristics in rat frontal immunoreactive populations. J. Comp. among irregular-spiking GABAergic J. Neurophysiol. 97, 2580–2589. cortex. J. Neurophysiol. 69, 416–431. Neurol. 293, 599–615.

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 165 Burkhalter Cortical interneurons

Lorente de No, R. (1992). The cerebral tex in the cat. J. Comp. Neurol. 302, Molecular taxonomy of major neu- key white and gray matter. J. Comp. cortex of the mouse. Somatosens. Mot. 124–142. ronal classes in the adult mouse fore- Neurol. 505, 526–538. Res. 9, 3–36. Porter, J. T., Johnson, C. K., and Agmon, A. brain. Nat. Neurosci. 9, 99–107. Uematsu, M., Hirai, Y., Karube, K., Ma, Y., Hu, A., Berrebi, A. S., Mathers, P. H., (2001). Diverse types of interneurons Sun, Q.-Q., Huguenard, J. R., and Ebihara, S., Kato, M., Abe, K., and Agmon, A. (2006). Distinct sub- generate thalamus-evoked feedfor- Prince, D. A. (2006). Barrel cortex Obata, K., Yoshida, S., Hirabayashi, M., types of somatostatin-containing ward inhibition in the mouse barrel microcircuits: thalamocortical feed- Yanagwa, Y., and Kawaguchi, Y. (2008). neocortical interneurons revealed cortex. J. Neurosci. 21, 2699–2710. forward inhibition in spiny stellate Quantitative chemical composition of in transgenic mice. J. Neurosci. 26, Rakic, P. (1976). Local Circuit Neurons, cells in mediated by a small number of cortical GABAergic neurons revealed 5069–5082. Cambridge, MIT press, pp. 5–19. fast-spiking interneurons. J. Neurosci. in transgenic venus-expressing rats. Markram, H., Toledo-Rodriguez, M., Ramón y Cajal, S. (1899). Apuntes para el 26, 1219–1230. Cereb. Cortex 18, 315–330. Wang, Y., Gupta, A., Silberberg, G., estudio estructural de la corteza visual Swadlow, H. A. (2003). Fast-spike Van Brederode, J. F. M., Mulligan, K. and Wu, C. (2004). Interneurons of del cerebro humano. Rev. Iberoam interneurons and feedforward inhi- A., and Hendrickson, A. E. (1990). the neocortical inhibitory system. Nat. Cienc. Med. 1, 1–14. bition in awake sensory neocortex. Calcium-binding proteins as mark- Rev. Neurosci. 5, 793–807. Ramón y Cajal, S. (1904). Textura del Cereb. Cortex 13, 25–32. ers for subpopulations of GABAergic McCormick, D. A., Connors, B. W., sistema nervioso del hombre y de Szabadics, J., Varga, C., Molnár, G., neurons in monkey striate cortex. J. Lighthall, J. W., and Prince, D. A. los vertebrados. Madrid, Imprenta Olah, S., Barzó, P., and Tamás, G. Comp. Neurol. 298, 1–22. (1985). Comparative electrophysiol- N. Moya. (2006). Excitatory effect of GABAergic von Engelhardt, J., Eliava, M., Meyer, A. H., ogy of pyramidal and sparsely spiny Reyes, A., Lujan, R., Rozov, A., axo-axonic cells in cortical microcir- Rozov, A., and Monyer, H. (2007). stellate neurons of the neocortex. J. Burnashev, N., Somogyi, P., and cuits. Science 311, 233–235. Functional characterization of intrin- Neurophysiol. 54, 782–806. Sakmann, B. (1998). Target-cell- Tabuchi, K., Blundell, J., Etherton, M. R., sic cholinergic interneurons in the cor- McDonald, C. T., and Burkhalter, A. ­specific facilitation and depression in Hammr, R. E., Liu, X., Powell, C. M., tex. J. Neurosci. 27, 5633–5642. (1993). Organization of long-range neocortical circuits. Nat. Neurosci. 1, and Südhof, T. C. (2007). A neuroli- Vruwink, M., Schmidt, H. H. H. W., inhibitory connections within 279–285. gin-3 mutation implicated in autism Weinberg, R. J., and Burette, A. (2001). rat visual cortex. J. Neurosci. 13, Schwaller, B., Meyer, M., and Schiffmann, S. increases inhibitory synaptic trans- Substance P and nitric oxide signal- 768–781. (2002). ‘New’ functions for ‘old’ pro- mission in ice. Science 318, 71–76. ing in cerebral cortex: anatomical evi- Miyoshi, G., Butt, S. J., Takebayashi, H., teins: the role of the calcium binding Tamás, G., Lörincz, Simon, A., and dence for reciprocal signaling between and Fishell, G. (2007). Physiologically proteins calbindin D-28k, calretinin Szabadics, J. (2003). Identified two classes of interneurons. J. Comp. distinct temporal cohorts of corti- and parvalbumin in cerebellar physi- sources and targets of slow inhibi- Neurol. 441, 288–301. cal interneurons arise from telen- ology. Studies with knockout mice. tion in the neocortex. Science 299, Wang, Y., Gupta, A., Toledo-Rodriguez, M., cephalic Olig2-expressing precursors. Cerebellum 1, 241–258. 1902–1905. Wu, C. Z., and Markram, H. (2002). J. Neurosci. 27, 7786–7798. Shao, Z., and Burkhalter, A. (1996). Tan, Z., Hu, H., Huang, J. Z., and Agmon, A. Anatomical, physiological, molecular Nelson, S. B., Sugino, K., and Different balance of excitation and (2008). Robust but delayed thalamo- and circuit properties of nest basket Hempel, C. M. (2006). The problem inhibition in forward and feedback cortical activation of dendritic-target- cells in the developing somatosensory of neuronal cell types: a physiological circuits of rat visual cortex. J. Neurosci. ing inhibitory interneurons. Proc. Natl. cortex. Cereb. Cortex 12, 395–410. genomics approach. Trends Neurosci. 16, 7353–7365. Acad. Sci. USA 105, 2187–2192. Wang, Y., Toledo-Rodriguez, M., Gupta, A., 29, 339–345. Shu, Y., Hasenstaub, A., and Thomson, A. M., and Deuchars, J. (1997). Wu, C., Silberberg, G., Luo, J., and Nowak, L. G., Azouz, R., Sanchez- McCormick, D. A. (2003). Turning on Synaptic interactions in neocorti- Markram, H. (2004). Anatomical, Vivez, M. V., Gray, C. M., and and off recurrent balanced a cortical cal local circuits: dual intracellular physiological and molecular proper- McCormick, D. A. (2003). activity. Nature 423, 288–293. recordings in vitro. Cereb. Cortex 7, ties of Martinotti cells in the soma- Electrophysiological classes of cat pri- Silberberg, G., and Markram, H. (2007). 510–522. tosensory cortex of the juvenile rat. J. mary visual cortical neurons in vivo Disynaptic inhibition between Thomson, A. M, and Lamy, C. (2007). Physiol. 561, 65–90. as revealed by quantitative analyses. J. neocortical pyramidal cells medi- Functional maps of neocortical local West, D. C., Mercer, A., Kirchhecker, S., Neurophysiol. 89, 1541–1566. ated by Martinotti cells. Neuron 53, circuits. Front. Neurosci. 1, 19–42. Morris, O. T., and Thomson, A. M. Ohki, K., and Reid, C. R. (2007). Specificity 735–746. Thomson, A. M., West, D. C., Wang, Y., (2006). Layer 6 cortico-thalamic and randomness in the visual cortex. Somogyi, P. (1977). A specific ‘axo-axonal’ and Bannister, A. P. (2002). Synaptic pyramidal cells preferentially innervate Curr. Opin. Neurobiol. 17, 401–407. interneuron in the visual cortex of the connections and small circuits involv- interneurons and generate facilitating Oliva, A. A., Jiang, M., Lam, T., Smith, K. L., rat brain. Brain Res. 136, 354–350. ing excitatory and inhibitory neurons EPSPs. Cereb. Cortex 16, 200–211. and Swann, J. W. (2000). Novel hip- Somogyi, P., and Cowey, A. (1984). Double in layers 2–5 of adult rat and cat neo- White, E. L. (1978). Identified neurons pocampal interneuronal subtypes bouquet cells. In Cerebral Cortex, cortex: triple intracellular recordings in mouse Sm1 cortex with are post- identified using transgenic mice that Vol. 1, A. Peters and E. G. Jones, eds and biocytin labeling in vitro. Cereb. synaptic to thalamocortical axon ter- express green fluorescent protein in (New York, NY, Plenum Press), pp. Cortex 12, 936–953. minals: a combined Golgi-electron GABAergic interneurons. J. Neurosci. 337–358. Toledo-Rodriguez, M., Goodman, P., microscopic and degeneration study. 20, 3354–3368. Somogyi, P., and Klausberger, T. (2005). Illic, M., Wu, C., and Markram, H. J. Comp. Neurol. 181, 627–661. Parra, P., Gulyás, A. I., and Miles, R. (1998). Defined types of cortical interneurone (2005). Neuropeptide and calcium- Wolansky, T., Pagliardini, S., Greer, J. J., How many subtypes of inhibitory structure space and spike timing in the binding protein gene expression and Dickson, C. T. (2007). cells in the hippocampus? Neuron 20, hippocampus. J. Physiol. 562, 9–25. profiles predict neuronal anatomical Immunohistochemical characteriza- 983–993. Staiger, J. F., Zilles, K., and Freund, T. F. type in the juvenile rat. J. Physiol. 567, tion of substance P receptor (NK(1) Peters, A. (1984a). Chandelier cells. In (1996). Distribution of GABAergic 401–413. R)-expressing interneurons in the Cerebral Cortex, Vol. 1, A. Peters and E. elements postsynaptic to ventropos- Tomioka, R., Okamoto, K., Furuta, T., entorhinal cortex. J. Comp. Neurol. G. Jones, eds (New York, NY, Plenum teromedial thalamic projections in Fujiyama, F., Iwasato, T., Yanagawa, Y., 502, 427–441. Press), pp. 361–379. layer IV of rat barrel cortex. Eur. J. Obata, K., Kaneko, T., and Tamamaki, N. Womelsdorf, T., Schoffelen, J. M., Peters, A. (1984b). Bipolar cells. In Neurosci. 8, 2273–2285. (2005). Demonstration of long-range Oostenveld, R., Singer, W., Cerebral Cortex, Vol. 1, A. Peters Stevens, C. F. (1998). Neuronal diversity: GABAergic connections distributed Desimone, R., Engel, A. K., and Fries, P. and E. G. Jones, eds (New York, NY, too many cell types for comfort. Curr. throughout mouse neocortex. Eur. J. (2007). Modulation of neuronal inter- Plenum Press), pp. 381–405. Biol. 8, R708–R710. Neurosci. 21, 1587–1600. actions through neuronal synchroni- Peters, A., Payne, B. R., and Josephson, K. Sugino, K., Hempel, C. M., Miller, M. N., Tomioka, R., and Rockland, K. S. (2007). zation. Science 316, 1609–1612. (1990). Transcallosal non-pyramidal Hattox, A. M., Shapiro, P., Wu, C., Long-distance corticocortical Wonders, C. P., and Anderson, S. A. (2006). cell projections from visual cor- Huang, Z. J., and Nelson, S. B. (2006). GABAergic neurons in the adult mon- The origin and specification of corti-

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 166 Burkhalter Cortical interneurons

cal interneurons. Nat. Rev. Neurosci. works depends on inhibitory cell teins in physiologically and morpho- Received: 06 July 2008; accepted: 18 7, 687–696. type and connectivity. Nat. Neurosci. logically characterized interneurons of September 2008; published: 15 December Xu, X., Roby, K. D., and Callaway, E. M. 8, 1552–1559. monkey dorsolateral prefrontal cortex. 2008. (2006). Mouse cortical inhibitory neu- Yoshimura, Y., Dantzker, J. M., and Cereb. Cortex 15, 1178–1186. Citation: Front. Neurosci. (2008) 2, 2: ron type that coexpresses somatostatin Callaway, E. M. (2005). Excitatory Zhu, J. J., and Connors, B. W. (1999). 155–167. doi: 10.3389/neuro.01.026.2008 and calretinin. J. Comp. Neurol. 499, cortical neurons form fine-scale Intrinsic firing patterns and whisker- Copyright: © 2008 Burkhalter. This is an 144–160. functional networks. Nature 433, evoked synaptic responses of neurons open-access article subject to an exclusive Yamashita, A., Valkova, K., Gonchar, Y., and 868–873. in the rat barrel cortex. J. Neurophysiol. license agreement between the authors and Burkhalter, A. (2003). Rearrangement Yuste, R. (2005). Origin and classification 81, 1171–1183. the Frontiers Research Foundation, which of synaptic connections with inhibi- of neocortical interneurons. Neuron permits unrestricted use, distribution, and tory neurons in developing mouse 48, 524–527. Conflict of Interest Statement: This reproduction in any medium, provided the visual cortex. J. Comp. Neurol. 464, Zaitsev, A. V., Gonzalez-Burgos, G., article was prepared in the absence of any original authors and source are credited. 426–437. Povysheva, N. V., Kroner, S., commercial financial relationships that Yoshimura, Y., and Callaway, E. M. (2005). Lewis, D. A., Krimer, L. S. (2005). could be construed as potential conflict Fine-scale specificity of cortical net- Localization of calcium-binding pro- of interest.

Frontiers in Neuroscience www.frontiersin.org December 2008 | Volume 2 | Issue 2 | 167