You are on page 1of 15

Fast-spiking, parvalbumin+ GABAergic interneurons: From cellular

design to microcircuit function


Hua Hu et al.
Science 345, (2014);
DOI: 10.1126/science.1255263

This copy is for your personal, non-commercial use only.

If you wish to distribute this article to others, you can order high-quality copies for your
colleagues, clients, or customers by clicking here.

Downloaded from www.sciencemag.org on August 1, 2014


Permission to republish or repurpose articles or portions of articles can be obtained by
following the guidelines here.

The following resources related to this article are available online at


www.sciencemag.org (this information is current as of August 1, 2014 ):

Updated information and services, including high-resolution figures, can be found in the online
version of this article at:
http://www.sciencemag.org/content/345/6196/1255263.full.html
This article cites 131 articles, 58 of which can be accessed free:
http://www.sciencemag.org/content/345/6196/1255263.full.html#ref-list-1
This article appears in the following subject collections:
Neuroscience
http://www.sciencemag.org/cgi/collection/neuroscience

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by the
American Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005. Copyright
2014 by the American Association for the Advancement of Science; all rights reserved. The title Science is a
registered trademark of AAAS.
RESEARCH

Recent results also provide a better un-


REVIEW SUMMARY derstanding of how PV+ interneurons oper-
ate in neuronal networks. Not only are PV+
interneurons involved in basic microcircuit
INTERNEURONS functions, such as feedforward and feedback
inhibition or gamma-frequency oscillations,
but they also play a role in complex network
Fast-spiking, parvalbumin+ GABAergic operations, including expansion of dynamic
activity range, pattern separation, modula-

interneurons: From cellular design to tion of place and grid field shapes, phase
SOURCE (LEFT TO RIGHT): DATA FROM LAPRAY ET AL. (2012), WITH PERMISSION FROM THE PUBLISHER; DATA FROM HU ET AL. (2010); DATA FROM WILSON AND MCNAUGHTON (1993) AND BUETFERING ET AL. (2014), WITH PERMISSION FROM THE PUBLISHER

precession, and gain modulation of sensory


responses. Thus, PV+ interneurons are criti-
microcircuit function cally involved in advanced computations in
microcircuits and neuronal networks.
Hua Hu, Jian Gan, Peter Jonas*
OUTLOOK: Parvalbumin-expressing inter-
BACKGROUND: Neuronal networks in output signal within a millisecond, but it is neurons may also play a key role in numer-
the brain include glutamatergic principal unclear how these signaling properties are ous brain diseases. These include epilepsy,
neurons and GABAergic interneurons implemented at the molecular and cellular but also complex psychiatric disea ses
(GABA, -aminobutyric acid). The latter levels, nor how PV+ interneurons shape such as schizophrenia.
ON OUR WEBSITE
may be a minority cell type, but they are complex network functions. Thus, PV+ interneurons
vital for normal brain function because Read the full article may become important
they regulate the activity of principal neu- ADVANCES: Recent work sheds light on at http://dx.doi therapeutic targets in
rons. If interneuron function is impaired, the subcellular signaling properties of PV+ .org/10.1126/ the f u t ur e . However ,
science.1255263
higher brain function can be damaged interneurons. PV+ cells show a high degree much needs to be
and seizures may result. The fast-spiking, of polarity. The weakly excitable dendrites learned about the basic
parvalbumin-positive interneurons (PV+ allow PV+ interneurons to sample activity function of these interneurons before clini-
interneurons) are readily characterized in the surrounding network, whereas the cal neuroscientists will have a chance to
and, consequently, have been adopted as a highly excitable axons enable analog-to-dig- successfully use PV+ interneurons for thera-
research model for systematic and quantita- ital conversion and fast propagation of the peutic purposes.
tive investigations. These cells contribute digital signal to a large number of target
to feedback and feedforward inhibition and cells. Additionally, tight coupling of Ca2+ IST Austria (Institute of Science and Technology Austria),
are critically involved in the generation of channels and release sensors at GABAergic Am Campus 1, A-3400 Klosterneuburg, Austria.
*Corresponding author. E-mail: peter.jonas@ist.ac.at
network oscillations. They can convert an output synapses increases the efficacy and Cite this article as H. Hu et al., Science 345, 1255263 (2014).
excitatory input signal into an inhibitory speed of the inhibitory output. DOI: 10.1126/science.1255263

A B C
Soma

Dendrite Putative
DG +
PV
CA1 INs

CA1

SR

SP
+
PV IN
EC
SO

PN

PV+ interneurons: from cellular design to microcircuit function. (A) Morphological properties of PV+ interneurons. Reconstruction of a
PV+ basket cell in the CA1 region of the hippocampus previously recorded in a freely moving rat. Soma and dendrites are shown in black; axon
is depicted in red. CA1, cornu ammonis region 1; SR, stratum radiatum; SP, stratum pyramidale; SO, stratum oriens. For details, see Fig. 1B. (B)
Subcellular physiology of PV+ interneurons. Simultaneous recording from the soma and dendrite of a PV+ interneuron in the dentate gyrus (DG)
in vitro. (Bottom left) Somatic recording pipette (used for fluorescent dye labeling); (top right) dendritic recording pipette. For details, see Fig.
2A. (C) Network function of PV+ interneurons. (Top) Place cell firing of hippocampal neurons; (bottom) grid cell firing of entorhinal cortex (EC)
neurons in vivo. Warm colors indicate high action potential frequency. Putative PV+ interneurons (PV+ INs) have broader spatial fields than prin-
cipal neurons (PNs). For details, see Fig. 6B.

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 529


Published by AAAS
R ES E A RC H

also expressing PV (5) (Box 2). Furthermore, we


REVIEW focus on the hippocampus and the neocortex.
For in vitro analysis of PV+ interneurons, the
advantages of the hippocampus are evident,
INTERNEURONS especially the clearly defined layering and the
availability of elaborate classification schemes

Fast-spiking, parvalbumin+ GABAergic (5). For in vivo analysis, the advantages of the
neocortex become apparent, including the super-
ficial localization of cells in the brain and the
interneurons: From cellular design to opportunity to easily define adequate behav-
ioral stimuli.

microcircuit function Morphological properties and


connectomics of PV+ interneurons
Hua Hu, Jian Gan, Peter Jonas* How can we understand the function of PV+
interneurons at the molecular, cellular, and net-
The success story of fast-spiking, parvalbumin-positive (PV+) GABAergic interneurons work levels? Following Francis Cricks statement
(GABA, g-aminobutyric acid) in the mammalian central nervous system is noteworthy. In If you want to understand function, study struc-
1995, the properties of these interneurons were completely unknown. Twenty years later, ture (16), let us first take a look at the structure
thanks to the massive use of subcellular patch-clamp techniques, simultaneous multiple- of PV+ interneurons, particularly their input do-
cell recording, optogenetics, in vivo measurements, and computational approaches, our mains, the dendrites, and their output domains,
knowledge about PV+ interneurons became more extensive than for several types of the axons.
pyramidal neurons. These findings have implications beyond the small world of basic The morphological properties of the den-
research on GABAergic cells. For example, the results provide a first proof of principle that drites of PV+ interneurons are notable in sev-
neuroscientists might be able to close the gaps between the molecular, cellular, network, and eral ways (1). PV+ interneurons have multiple
behavioral levels, representing one of the main challenges at the present time. Furthermore, dendrites that often cross layers (1720), which
the results may form the basis for PV+ interneurons as therapeutic targets for brain permits the interneurons to receive input from
disease in the future. However, much needs to be learned about the basic function of these different afferent pathways, such as feedfor-
interneurons before clinical neuroscientists will be able to use PV+ interneurons for ward and feedback pathways. The cumulative
therapeutic purposes. dendritic length of a single PV+ interneuron

I
ranges from 3.1 to 9 mm (1720). Long den-
n a reductionists view of the brain, neu- How can we systematically and quantitatively drites allow PV+ interneurons to sample input
ronal networks are composed of two types study the function of such an enormously diverse from a large population of principal cells. Fi-
of neurons: glutamatergic principal neu- population of cells? One approach is to focus nally, the somata and dendrites of PV+ inter-
rons and GABAergic interneurons (GABA, on models, identifiable on the basis of standard- neurons are densely covered with synapses. PV+
g-aminobutyric acid). Across all cortical cir- ized criteria. That is why, around 1990, several interneurons in the hippocampal CA3 or CA1
cuits, glutamatergic neurons form ~80 to 90% laboratories started to work on one particular region have ~16,000 to 34,000 synapses, 94%
of the neuronal population, whereas GABAergic type of GABAergic interneuron: the fast-spiking, of which are excitatory and 6% are inhibitory
neurons constitute the remaining 10 to 20% (13). parvalbumin-positive (PV+) interneuron (Box 1). (17, 20). A large proportion of inhibitory syn-
Thus, in terms of neuron numbers, GABAergic In the hippocampal CA1 region, 11% of neurons apses is PV+ (17), but inhibitory inputs from
cells represent only a minority. Nevertheless, are GABAergic, and 24% of those are PV+; thus, vasoactive intestinal peptide- and somatostatin-
these GABAergic neurons serve important func- PV+ interneurons represent only 2.6% of the expressing interneurons are also present (21, 22).
tions. Most notably, they control the activity total neuronal population (8). Accordingly, PV+ Thus, PV+ interneurons receive convergent excit-
level of principal neurons in the entire brain. If interneurons appear to be somewhat exotic. atory input from principal neurons, and inhibi-
GABAergic interneuron function breaks down, However, numerous technical advantages out- tory input primarily from other PV+ interneurons.
excitation takes over, leading to seizures and weigh this potential disadvantage (Fig. 1). The Because the dendrites of PV+ interneurons are
failure of higher brain functions (4). selective expression of the Ca2+-binding pro- largely aspiny, excitatory synapses are formed
A hallmark of interneurons is their structural tein PV allows unequivocal post hoc labeling by on dendritic shafts. This may facilitate the gen-
and functional diversity (Fig. 1A). Twenty-one highly specific antibodies (9, 10) (Fig. 1B). Both eration of fast excitatory postsynaptic poten-
different classes of interneurons have been dis- the short action potential duration and the fast- tials (EPSPs) (23).
tinguished in the CA1 region of the hippocampus spiking action potential phenotype make it easy The morphological properties of the axon of
(5), and it is likely that an even larger number to identify these cells under experimental con- PV+ interneurons are also intriguing (1). In the
of types can be dissected in the neocortex (6). ditions (Fig. 1D). The high selectivity of the pro- classical anatomical literature, GABAergic inter-
These interneurons can be distinguished on moter of the PV gene can be used to genetically neurons were sometimes referred to as short
the basis of three sets of criteria: (i) morpholog- target these cells by enhanced green fluorescent axon cells. However, for many PV+ interneurons,
ical properties, particularly the target selectiv- protein and optogenetic methods (11, 12) (Fig. 1, this seems entirely incorrect. The axon shows
ity of the axon; (ii) expression of molecular E and F). The specific developmental trajectory extensive arborization, and the cumulative axo-
markers, such as neuropeptides (somatostatin, of cortical PV+ interneurons, which are born in nal length of a single PV+ interneuron is 30 to
cholecystokinin, vasoactive intestinal peptide, the medial ganglionic eminence and depend on 50 mm [33 mm in the dentate gyrus (18), 46 mm
and neuropeptide-Y) and Ca2+-binding proteins specific sets of transcription factors (i.e., Nkx2-1 in the hippocampal CA1 region (24), 20 and 24 mm
(parvalbumin, calretinin, and calbindin); and and Lhx6), may be exploited for labeling (1315). in the frontal cortex (25)]. A huge number of en
(iii) functional characteristicsmost importantly, In this Review, we summarize current knowl- passant (in passing, upon literal translation) ter-
the action potential phenotype (7). edge about fast-spiking, PV+ interneurons at the minals emerge from the extensive axonal arbor
molecular, cellular, and network levels. We con- [10400 in CA1 (24), 3200 and 3800 in the frontal
IST Austria (Institute of Science and Technology Austria),
centrate on basket cells (the classical PV+ inter- cortex (25)]. Thus, PV+ interneurons generate a
Am Campus 1, A-3400 Klosterneuburg, Austria. neurons) but include information about axo-axonic massively divergent inhibitory output (8). PV+
*Corresponding author. E-mail: peter.jonas@ist.ac.at cells or other types of GABAergic interneurons interneurons innervate postsynaptic target cells in

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-1


R ES E A RC H | R E V IE W

A D

Neocortex
CA1
50 mV

3 nA

25 ms

E
Targeted Parvalbumin-locus
B ATG STOP
IRES-CRE TK-NEO-Hygro
Ex1 Ex2 Ex3 Ex4 HindIII Ex5 HindIII HindIII
Neurobiotin
CA3 mCherry PV overlay
CA1
5 cm

Video tracking Parvalbumin

SR
25 m
CA1

SP
CA3

SO
100 m

50 m

Soma
Dendrite F
0.25 m
PFC CA1

C 6,000 Light AP frequency (Hz)


Energy 4 (V2)

Spont
13

MC Layer 5 0
6

500
0
500 9,000 2 1 0 1 2 3
150 m Energy 2 (V 2)
Time (s)

Fig. 1. The PV+ interneuron: An interneuron subtype with salient proper- train of action potentials in the intracellularly recorded neuron. Data are from
ties and distinct experimental identifiability. (A) The placement of PV+ (124). (E) Genetic fluorescent protein labeling of PV+ interneurons, using mice
interneurons in interneuron diversity schemes (122). (Left) Scheme showing a expressing Cre recombinase under the control of the PV promoter. (Left)
subset of the 21 types of GABAergic interneurons currently known in the mCherry labeling after adenoassociated virus infection; (center) PV immuno-
hippocampal CA1 region. PV+ axo-axonic cells and basket cells are located on reactivity; (right), overlay. Inset on top shows targeted PV gene. Scheme on
the left side. (Right) A PV+ basket cell (probably the most abundant type of top is from (125). (F) Functional labeling of PV+ interneurons, using mice
PV+ interneuron). Soma and dendrites are shown in orange; axons are expressing Cre recombinase and channelrhodopsin or halorhodopsin under
depicted in yellow. (B) A PV+ basket cell in the CA1 region of the hippocampus the control of the PV promoter in vivo. (Left) Identification of unit activity in
recorded in a freely moving rat. SR, stratum radiatum; SP, stratum PV+ interneurons expressing channelrhodopsin in an awake, freely moving
pyramidale; SO, stratum oriens. (Upper left inset) Movement trajectory of mouse recorded with optrodes. Data are from (126). (Right) Activity of a
the animal. (Upper right inset) PV immunoreactivity. (Lower right inset) PV+ interneuron expressing halorhodopsin in an awake mouse moving on a
Electron micrograph of output synapses. Data are from (87). (C) A basket linear track. Note that the light pulse abolishes action potential (AP) ini-
cell in layer five of the motor cortex (MC). Data are from (123). Color code in tiation in the PV+ interneuron (top) but increases the action potential fre-
(B) and (C): Soma and dendrites are shown in black; axon is depicted in red. quency in a simultaneously recorded pyramidal neuron (bottom). Data are
(D) Fast-spiking action potential phenotype of a putative PV+ interneuron in from (92). CA1, cornu ammonis region 1; CA3, cornu ammonis region 3;
the neocortex in vitro. A long somatic current pulse evoked a high-frequency PFC, prefrontal cortex.

1255263-2 1 AUGUST 2014 VOL 345 ISSUE 6196 sciencemag.org SCIENCE


RE S EAR CH | R E V I E W

the perisomatic domain. In basket cells, the axon tude and fast time course of the AMPA receptor period of temporal summation and promoting
forms basket-like arrangements around principal mediated postsynaptic conductance at excitatory action potential initiation with high speed and
cell somata and proximal dendrites. In axo-axonic input synapses (23, 36, 37). In the thin dendrites temporal precision (27, 38) (Fig. 2E) (Table 1).
cells, the axon of the interneuron follows the axon of PV+ interneurons, AMPA receptormediated K+ channel activation conveys sublinear inte-
initial segment of the principal cell, resulting in a conductances generate local EPSPs with large gration (27), which may allow PV+ interneurons
chandelier-like configuration (5). These morpho- peak amplitude (18), resulting in efficient activa- to accurately sample principal neuron activity
logical characteristics suggest that PV+ interneur- tion of dendritic Kv3 channels (7, 27) (Fig. 2D). over a wide range. Additionally, K+ channel ac-
ons generate a particularly powerful inhibition, This has profound functional consequences: Den- tivation makes PV+ cells less sensitive to clustered
because they innervate a large number of target dritic K+ channel activation accelerates the decay excitatory input, which activates K+ channels effi-
cells near the site of action potential initiation. time course of the EPSP, shortening the time ciently, but relatively more sensitive to distributed
However, these properties also raise new questions:
For example, one may wonder how reliable action
potential propagation is achieved in the highly Box 1. The steep scientific career of PV+ interneurons.
branching interneuron axon (26) and how the
functions of signal propagation and transmit- 1986: Celio (9) suggests that PV is expressed in the majority of GABAergic neurons in the
ter release are integrated into a single structure. cortex.
To address these questions, we must examine 1987: Kawaguchi et al. (132) suggest that PV is selectively expressed in fast-spiking
the function of PV+ interneurons directly at the interneurons.
subcellular level with micrometer spatial and micro- 1995: Geiger et al. (36) demonstrate that fast-spiking, PV+ interneurons express AMPA-type
second temporal resolution. Subcellular patch- glutamate receptors with high Ca2+ permeability and fast gating, caused by a low relative
clamp recording now allows researchers to obtain abundance of GluA2 subunit mRNA.
such measurements in both dendrites and axons 1996: Du et al. (133) demonstrate that Kv3 K+ channel subunits are selectively expressed in
of PV+ interneurons. PV+ cells, providing the first suggestion of a molecular mechanism underlying the fast-spiking
action potential phenotype.
The subcellular physiology 1997: Geiger et al. (23) show that fast-spiking PV+ interneurons in the dentate gyrus
of PV+ interneurons: Dendrites receive fast excitatory synaptic inputs.
Direct dendritic recordings have provided a de- 2001: Pouille and Scanziani (72) demonstrate that feedforward inhibition, presumably provided
tailed quantitative picture of the electrical events by PV+ interneurons, shortens the coincidence detection window in pyramidal neurons.
in PV+ interneuron dendrites (Fig. 2, A and B). 2005: Hippenmeyer et al. (125) generate a PV-Cre mouse line that specifically expresses
First, action potentials backpropagate into the Cre recombinase in PV+ interneurons. This opened the door for both selective labeling
dendrites in a highly decremental manner (27), and manipulation.
confirming findings of previous Ca2+ imaging ex- 2009: Cardin et al. and Sohal et al. (11, 12) show that rhythmic optogenetic stimulation of
periments (2831) (Fig. 2C). Similar results were PV+ interneurons results in the generation of gamma oscillations, whereas inhibition of
obtained in both basket and axo-axonic cells PV+ interneurons reduces gamma power.
(27). These properties differ from those of py- 2010: Hu et al. (27) provide the first recordings from subcellular processes of PV+
ramidal neurons, where backpropagation is interneurons (dendrites and, later, axons). This now results in a complete mapping of the
active (32). Second, dendritic spikes cannot be functional properties of these cells along the dendrite-soma-axon axis.
initiated, neither by dendritic current injection 2012: Wilson et al., Atallah et al., and Lee et al. (94, 95, 102) show that PV+ interneurons
nor by synaptic stimulation (27), although a re- control gain of sensory responses. This is probably the first demonstration that a specific
cent study suggested that dendritic spikes may aspect of signal processing in neuronal networks can be attributed to a distinct cell type.
be evoked by massive glutamate uncaging (31). June 2014: A PubMed search for parvalbumin interneuron returns 1644 hits, with many
Again, these properties differ from those of py- recent papers published in Science and Nature.
ramidal neurons, where dendritic spikes are
abundant (33). Third, the dendrites of PV+ in-
terneurons contain only a low density of voltage-
gated Na+ channels; Na+ channels are almost
Box 2. The several caveats of PV+ interneuron identification.
absent at distances >100 mm from the soma (27).
Fourth, the dendrites of hippocampal PV+ inter-
How can we identify the fast-spiking, PV-expressing basket cell? Ideally, for rigorous interneuron
neurons contain a high density of voltage-gated
identification, one would like to see complete morphological visualization, expression analysis for
K+ channels, consistent with the results of pre-
the most important interneuron markers, and functional characterization. In practice, identification
vious Ca2+ imaging experiments with K+ chan-
often relies on a subset of parameters. In the past, identification was often based on the action
nel blockers (28). The high dendritic ratio of K+
potential phenotype (124) (fast spiking; action potential frequency > 50 Hz at 22C and > 150 Hz
to Na+ channels distinguishes PV+ interneurons
at 34C) and the morphology of the axon (basket cells; ~90% of collaterals in the cell body layer)
from pyramidal cells and also from other inter-
(5). However, cholecystokinin- and vasoactive intestinal peptide-expressing basket cells have
neuron subtypes (3235). Lastly, analysis of gating
maximal action potential frequencies that are less than a factor of 2 lower than those of PV+
and pharmacological properties revealed that
neurons (57). More recently, identification increasingly exploited PV expression, for example, in
these channels are primarily of the Kv3 type, one
optogenetic experiments (11, 12). However, PV is expressed not only in basket cells, but also in a
of the four main subfamilies of voltage-gated
subset of axo-axonic cells, bistratified cells, and even in oriens-alveuslacunosum-moleculare
K+ channels (7). These channels show high ac-
(OLM) interneurons (5). Furthermore, one needs to consider regional differences. In the prefrontal
tivation threshold, fast activation, and fast de-
cortex, only a subset of axo-axonic cells may express PV (14); in the dentate gyrus, analogs of
activation (7).
bistratified cells may not be present. Finally, PV expression levels matter because OLM interneurons
Why should PV+ interneurons express high-
express PV at lower concentrations than basket or axo-axonic cells (10). More work is needed to
threshold K+ channels in the dendrites if the
elucidate the functional differences between PV+ cell types, mainly basket versus axo-axonic cells.
amplitude of backpropagated action potentials
This distinction is particularly important because basket cells have inhibitory effects on their
is too small to activate them? As it turns out, den-
postsynaptic target cells, whereas axo-axonic cells may be excitatory (134).
dritic Kv3 channels work synergistically with the
small diameter of dendrites and the large ampli-

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-3


R ES E A RC H | R E V IE W

input, which activates these channels only mini- but are connected to adjacent interneurons (37). from pyramidal neurons, where the initiation
mally (27) (Fig. 2E). Finally, gap junctions may boost the efficacy of site is more remote, sometimes even beyond the
Dendrites of PV+ interneurons in both the distal inputs and increase the average action po- axon initial segment (44). (ii) Action potentials
hippocampus and the neocortex are highly tential frequency after repetitive synaptic stimu- propagate with high reliability; propagation fail-
interconnected by gap junctions (3942). Such lation of distal synapses (35). ures occur only rarely. (iii) The orthodromic action
a syncytial organization of dendritic trees will potential propagation velocity is ~1.5 m s1 at near-
also affect synaptic integration. Gap junctions The subcellular physiology physiological temperature, notable for a thin,
will lead to speeding of the EPSP time course, of PV+ interneurons: Axons largely unmyelinated axon (43). The propagation
because excitatory charge can escape into ad- Direct recordings also revealed several surpris- velocity is faster than that of principal neuron
jacent dendrites. Furthermore, gap junctions ing properties of axons of hippocampal PV+ in- axons under comparable conditions (45, 46).
may widen the spatial range of detection of prin- terneurons (43) (Fig. 3, A and B): (i) The action (iv) PV+ interneurons exhibit a distinct Na+ chan-
cipal neuron activity, including input synapses potential is initiated very proximally, ~20 mm nel distribution: a stepwise density increase from
that are unconnected to a given PV+ interneuron from the soma (43) (Fig. 3A). This is different the soma to the proximal axon, followed by a

A C E Passive
Kv3

10 m 50 m

Half-duration (ms)
0.05 10
DG CA1 G/R 1 mV 9
8
10 ms 7
6
20 mV 5
200 ms 0 2 4 6 8 10
Gsyn (nS)
Peak G/R

0.12

0.08

EPSPmax /EPSP1

EPSPmax /EPSP1
2x
1.8 1.8
0.04 1.6 1.6
6
20 m 1.4 1.4
.4
4
0.00
1.2 11.2
0 20 40 60 80 100
Distance (m) 0 10 20 -20 -10 0 10 20
t (ms) t (ms)
B
Soma D F
DG Pyramidal neuron
Apical dendrite (124 m)
GK PV+ interneuron
20 mV
max
0.5 ms
Apical tuft
DG
AP peak amplitude (mV)

x Gain
80 Soma
60 Basal and
Axon obliques
40 0
Excit
Soma
Inhit
20

0
Axon
-50 0 50 100 150 200 250 300 350
Dendritic distance (m)

Fig. 2. The in and out of PV+ interneurons: dendrites. (A) Direct soma. Error bars indicate SEM. R, red; G, green fluorescent signal. Data are
patch-clamp recording from subcellular processes of PV+ interneurons in the from (30). (D) Dendritic Kv3-type K+ channels in a PV+ interneuron model
dentate gyrus (DG), using confocally targeted patch-clamp recording in a are locally activated by synaptic input (apical dendrite, arrow). Pseudocolor
brain slice in vitro. The cell was first loaded with fluorescent dye (Alexa Fluor code indicates the activated K+ conductance (GK). (E) Dendritic Kv3-type K+
488) via a somatic recording pipette. A dendritic recording was subsequently channels in PV+ interneurons accelerate the EPSP time course (top) and
obtained on the distal dendrite. (B) Decremental action potential back- enhance the ability of the neuron to detect temporally coincident, but spa-
propagation into dendrites. Peak amplitude of the action potential at the tially distributed inputs (bottom). Blue, passive dendrites; black, K+ channels
dendrite was plotted against distance, with recording sites at basal dendrites in synapse-containing dendrites. Gsyn, synaptic peak conductance. Data in
(negative distance) and apical dendrites (positive distance). Solid circles, (A), (B), (D), and (E) are from (27). (F) Schematic illustration of the different
somatic current injection; open circles, dendritic current injection. Inset on rules of dendritic integration in PV+ interneurons (left) and pyramidal neurons
top shows action potentials at the soma (black) and dendrite (red), with the (right). In PV+ interneurons, the high ratio of K+ channels to Na+ channels in
current pulse applied to the soma. (C) Dendritic Ca2+ transients in PV+ dendrites confers linear or sublinear integration (S, linear summation mecha-
interneurons in the hippocampal CA1 region in vitro. (Top) Dendritic Ca2+ nism). In pyramidal neurons, the high ratio of Na+ channels to K+ channels in
transients at 10 and 50 mm distance from the soma. (Bottom) Decline of dendrites enriches the repertoire of single-neuron computations [, sigmoidal
amplitude of dendritic Ca2+ transients as a function of distance from the threshold mechanism (127)].

1255263-4 1 AUGUST 2014 VOL 345 ISSUE 6196 sciencemag.org SCIENCE


RE S EAR CH | R E V I E W

further gradual increase to the distal axon (43, 47) What is the molecular identity of Na+ and K+ But how is the electrical signal in the axon con-
(Fig. 3B). In the distal axon, the Na+ conduct- channels in the axon of PV+ interneurons? For verted into GABA release (55)? Several factors
ance density is ~600 pS mm2, which is compa- voltage-gated Na+ channels, NaV1.1 and NaV1.6 are important for this conversion, including the
rable to values in invertebrate axons (43). Thus, immunoreactivity is abundantly present in the duration of the presynaptic action potential,
PV+ interneurons show a weakly excitable an- axons (49, 50). However, the contribution of other the gating of the presynaptic Ca2+ channels, the
alog somatodendritic domain (with graded syn- subunits cannot be excluded, because NaV1.2, -1.4, coupling between Ca2+ channels and release sen-
aptic potentials) and a highly excitable digital and -1.7 mRNAs are also detectable in PV+ inter- sors, and the Ca2+ binding and unbinding rates of
axonal domain (with all-or-none action poten- neurons (51). For voltage-gated K+ channels, Kv3 the release sensor. Collectively, these factors will
tials), separated by a steep transition zone. As- subunits are heavily expressed in PV+ interneurons, determine the synaptic delay, that is, the time
suming that the axon represents ~74% of the and Kv3 immunoreactivity has been localized to interval between the action potential in the pre-
surface area (18), ~99% of the Na+ channels would axons (7, 52). Furthermore, the pharmacological synaptic terminal and the event of exocytosis.
be located in the axon. Hence, the excitability and gating properties of axonal K+ channels Many of these factors in PV+ interneuron out-
mechanism of PV+ cells is almost entirely axo- imply that they are primarily of the Kv3 subtype put synapses are optimized for speed (Fig. 3C).
nal. (v) The axon of PV+ interneurons contains (48). The high activation threshold and the fast Direct recordings revealed that axonal action
voltage-gated K+ channels with properties simi- deactivation of these channels may ensure fast potentials are brief, comparable to those at the
lar to those in the dendrites of PV+ cells (43, 48). action potential repolarization in the axon. Fi- soma (43). Because presynaptic terminals are of
Why should PV+ interneurons express an ex- nally, Kv1 channels are present in the axon initial the en passant type, this will directly translate
cessively high density of Na+ channels in the segment of hippocampal and neocortical PV+ in- into fast and synchronous transmitter release.
axon? The first guess was that the high density terneurons (50, 53, 54). The low activation thresh- Consistent with this hypothesis, broadening of
guarantees reliability of action potential prop- old and the slower gating of these channels may the presynaptic action potential by K+ channel
agation (26). However, experiments and simula- define characteristic input-output conversion prop- blockers enhances both presynaptic Ca2+ tran-
tions indicate that Na+ channels are expressed erties in PV+ cells. Long current pulses will activate sients and peak amplitudes of postsynaptic cur-
at higher density values than the critical value these channels, suppressing the initiation of action rents (48).
required for reliability (43). The supercritical potentials (53). In contrast, fast EPSPs will bypass Whereas several types of synapses use mixtures
Na+ channel density has two additional advan- Kv1 channel activation, leading to action potential of P/Q-, N-, and R-type Ca2+ channels for trans-
tages: It increases the speed of action potential initiation with short delay (53). Thus, Kv1 channels mitter release (56), the output synapses of PV+
propagation and the maximal action potential may implement a fast coincidence detection mech- cells in both the hippocampus and the neocortex
frequency during sustained somatic current in- anism in PV+ interneurons (Table 1). In addition, exclusively rely on P/Q-type channels (5760). As
jection (43). Hence, the high Na+ channel den- the profound inactivation of these channels may P/Q-type Ca2+ channels show the fastest gating
sity in the axon contributes to rapid signaling explain delayed spiking during long-lasting de- among all Ca2+ channel subtypes (56), the spe-
in PV+ interneurons (Table 1). In relation to polarizations (53). cific usage of these channels will contribute to
propagation speed, the high channel density both the shortening of the synaptic delay and
compensates for the unfavorable morphological From axons to presynaptic terminals: the increase in the temporal precision of trans-
properties of interneuron axons (small segmen- Fast GABA release mitter release.
tal diameter, extensive branching, and high bou- Multiple properties of dendrites and axons of PV+ Another specific property of transmission is
ton density). interneurons are specialized for rapid signaling. the tight (nanodomain) coupling between Ca2+

Table 1. Fast signaling properties of PV+ interneurons.

Property Functional consequence Molecular mechanism Reference(s)


Fast EPSCs Fast EPSP time course; GluA1 and GluA4 AMPA (23, 37)
coincidence detection receptor subunits
Fast IPSCs High-frequency GABRA1 GABAA (129, 130)
network oscillations receptor subunits
Dendritic K+ channels Sublinear summation; Kv3 (27, 28)
(high activation threshold, sensitivity to distributed
fast deactivation) inputs
Perisomatic K+ channels Short action potentials; Kv3 (7, 131)
(high activation threshold, fast-spiking AP phenotype
fast deactivation)
D-type K+ channels in axon Coincidence detection; Kv1 (50, 53, 54)
initial segment delayed spiking during
long current pulses
Supercritical Na+ channel Fast axonal action potential NaV1.1 (43, 49, 50)
density in axon propagation; fast-spiking NaV1.6
AP phenotype
Fast presynaptic Ca2+ channels Speed and temporal Cav2.1, P/Q-type (5760)
precision of GABA release Ca2+ channel
Tight coupling between Speed and temporal precision Unknown (57, 59)
presynaptic Ca2+ channels of synaptic transmission
and release sensors
Fast release sensors Speed and temporal precision Synaptotagmin 2 (63, 64, 66, 67)
of synaptic transmission

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-5


R ES E A RC H | R E V IE W

channels and release sensors of exocytosis (61, 62) in extreme cases, the generation of Ca2+ spikes on synaptotagmin 1 (6365). Recent studies have
(Fig. 3C). Tight coupling increases the efficacy in presynaptic terminals. Thus, the small num- used synaptotagmin 2 immunolabeling to selec-
of release, shortens the synaptic delay, and in- ber of Ca2+ channels per release site at PV+ in- tively visualize PV+ boutons in the visual cortex
creases the temporal precision of release (61). terneuron output synapses contributes to fast and (66). As synaptotagmin 2 has the fastest Ca2+-
Furthermore, GABA release at presynaptic ter- temporally precise transmitter release. binding kinetics throughout the synaptotagmin
minals of PV+ interneurons is initiated by a small Finally, a subset of PV+ interneurons in the hip- family (67), the expression of this synaptotagmin
number of Ca2+ channels, probably only two or pocampus and the neocortex uses synaptotagmin isoform may also contribute to rapid signaling.
three per release site (59). The usage of a small 2 (1 out of the 15 members of the synaptotagmin Direct measurement of Ca2+-binding rates of dif-
number of Ca2+ channels could help avoid the family) as a release sensor for synaptic transmis- ferent synaptotagmin isoforms will be needed to
broadening of presynaptic action potentials or, sion; in contrast, principal neurons primarily rely quantitatively test this hypothesis.

A B Dendrite Axon
(188 m) Soma (315 m)

Axon (24 m)
Axon (315 m) 0 mV 0 mV 0 mV
Soma
Soma
-120 mV -120 mV -120 mV
20 mV
0.5 ms 50 pA
1 ms

DG 0.6 DG 120
AP1 latency (ms)

100

density (m-2)
0.4

Na+ channel
Soma 80
first 0.2
60
0
40
Axon 0.2
first 20
0.4 0
0 100 200 300 400 -200 -100 0 100 200 300 400 500
Distance (m) Distance (m)
Initiation site
C D Parvalbumin
Low PV
CA3
Intermediate high PV
Intermediate low PV High PV
EF domain 100
BAPTA
CD domain

+
80

Fraction of PV
neurons(%)
60
20
EGTA
20
0
AB domain Ctrl EE cFC
DG
Half-
Delay duration
DG Cerebellum
Release rate (norm.)

1 20 1.5
Time (ms)

Mean [PV] (M)

20 nm 1 750 0.8
Mean [PV] (M)

200 nm 15
0.5 0.6
0.5 500
CV

CV

10
0.4
0.5 250
0 5 0.2
0
0 50 100 150 200
0 1 2 3 4 5
Time (ms) Coupling distance (nm) 0 0 0 0

Fig. 3. The in and out of PV+ interneurons: axons and presynaptic EGTA used to probe the coupling configuration. (Top right) Concentration
terminals. (A) Proximal initiation of action potentials. (Top) Action potentials dependence of the effects of the two chelators on GABA release. Curves
at the soma (black) and axon (blue) in two different soma-axon recordings represent linearized models fit to the experimental data, revealing a coupling
with different axonal distance. (Bottom) Latency between action potentials in distance of 12 nm. IPSC, inhibitory postsynaptic current; r, coupling distance;
axon and soma was plotted against distance, with negative values indicating d, mean coupling distance; s, SD. (Bottom left) Dependence of simulated time
axon-first behavior and positive values representing soma-first behavior. course of release on coupling distance in a source-sensor model. (Bottom
Note the sharp initiation site in the axon ~20 mm from the soma. (B) Voltage- right) Simulated synaptic delay and half-duration of the time course of
gated Na+ channel spatial distribution profile. Channel density measured in release as a function of coupling distance. Data are from (61). (D) Action of
the outside-out patch configuration is plotted against distance, with negative the endogenous Ca2+ binding protein PV on transmitter release. (Top left)
values indicating dendritic location and positive values indicating axonal Secondary structure of PV, with the spheres indicating bound Ca2+ ions. Data
location. Note the stepwise increase of Na+ channel density from the soma to are from (128). (Top right) Activity-dependent regulation of PV concentration
the proximal axon, followed by a gradual increase to the distal axon. Data in in interneurons of the hippocampal CA3 region. EE, enriched environment;
(A) and (B) are from (43). (C) Tight coupling between Ca2+ channels and cFC, contextual fear conditioning. Data are from (69). (Bottom) Estimation of
release sensors at the output synapses of PV+ interneurons. (Top left) Chem- absolute PV concentration in the soma of different types of inhibitory
ical structure of the fast Ca2+ chelator BAPTA and the slow Ca2+ chelator interneurons by calibrated immunocytochemistry. Data are from (10).

1255263-6 1 AUGUST 2014 VOL 345 ISSUE 6196 sciencemag.org SCIENCE


RE S EAR CH | R E V I E W

What are the effects of PV on GABA release transmitter release, for example, by acting as neurons reveal a distance-dependent decline
at PV+ interneuron output synapses? The EF- an antifacilitation factor (10). of inhibition (72). Furthermore, this inhibition
hand domains of PV bind both Ca2+ and Mg2+ is primarily mediated by fast-spiking PV+ cells,
ions (Fig. 3D). Therefore, it is generally thought The role of PV+ interneurons in because these interneurons fire early after stim-
that Mg2+ must leave before Ca2+ can bind, con- microcircuits: Beyond simple inhibition ulation, before pyramidal cells and regularly
ferring slow Ca2+ binding properties to this GABAergic interneurons are involved in both spiking interneurons (74) (Fig. 4C).
protein (68). How can such a Ca2+ buffer act in feedforward and feedback inhibition (7073) Feedforward microcircuits incorporating PV+
nanodomain coupling regimes? The high PV (Box 3). But what is the specific contribution of cells may have several functions beyond simple
concentration may provide an answer to this PV+ cells, and what is the functional relevance inhibition. For example, feedforward inhibition
question. If the PV concentration is at milli- of their fast signaling mechanisms? Experimen- by PV+ interneurons narrows the window for tem-
molar levels, as found in cerebellar basket cells tal evidence indicates that PV+ interneurons in poral summation of EPSPs and action potential
(10), or if the PV concentration is up-regulated the hippocampus are involved in feedforward initiation in principal neurons (72) (Fig. 4D).
during behavior (e.g., contextual fear condition- inhibition (Fig. 4). In the hippocampal CA1 re- Feedforward inhibition by PV+ interneurons will
ing or learning completion), as observed in the gion, feedforward inhibition initiated by stimula- expand the dynamic range of activity in large prin-
hippocampal PV+ interneurons (69), the free apo tion of Schaffer collaterals is primarily mediated cipal neuron ensembles (74) (Fig. 4E). For both
form of PV may become functionally relevant (Fig. by perisomatic inhibitory interneurons, because functions, the fast signaling of PV+ interneurons
3D). Under these conditions, PV may modulate somatodendritic recordings from CA1 pyramidal is critically important. PV+ interneuron-mediated

A B CA1
C CA1 Stimulus
Rec
Rec Rec

feedforward IN
PC PC layer

Field
Subtraction 40 mV
0 1 2 3 3 mV
Delay (ms)* PC Loose patch PC d 500 ms
20 pA
PV+ IN * ax
Control
PN 3 mV
FS
Bicuculline 50 pA
2.5 ms 200 pA

3 mV
RS

40 pA
5 ms
D 1
CA1 200 m

E CA1
Control
Population activation curve Activation curves
Control 100 100
95 95
Bicuculline
Recruited PCs (%)

Recruited PCs (%)


Spikes (norm. prob.)

100
2
50 50
50
Bicu-
culline Control
5 5 Gabazine
0 0 0
-40 -30 -20 -10 0 10 20 30 40 0 0.2 0.4 0.6 0.8 1.0 0 0.2 0.4 0.6 0.8 1.0
20 pA t (ms) Input strength Input strength
10 ms

Fig. 4. The role of PV+ interneurons in feedforward inhibitory micro- stimulation of two Schaffer collateral inputs at different time intervals
circuits. (A) Schematic illustration of the feedforward inhibitory micro- under control conditions (top) and after a block of inhibition (bottom).
circuit. IN, interneuron; PN, principal neuron. [Scheme courtesy of G. Buzski.] Arrows indicate time points of first and second stimulus. (Right) Spike
(B) Time course of disynaptic feedforward inhibition in the hippocampal CA1 probability after stimulation of two excitatory inputs with time interval Dt in
region. Traces indicate control signal (monosynaptic EPSC and disynaptic the presence of inhibition (gray bars) and after a block of inhibition (white
IPSC), pharmacologically isolated EPSC, and analog subtraction. Note the bars). Error bars denote SEM. Data are from (72). (E) Feedforward in-
short latency of the disynaptic IPSCs (inset). The asterisk indicates delay. hibition expands the dynamic range of principal neuron population activity.
Data are from (72). (C) Activation of pyramidal neurons (top), fast-spiking (Left) Plot of proportion of recruited CA1 pyramidal cells (PCs) against
interneurons (center), and regularly spiking interneurons (bottom) during synaptic input strength. Filled data points and the black curve show the
feedforward inhibition. Upper traces show local field potentials; lower traces results for the entire population of CA1 pyramidal neurons; gray curves
illustrate loose-patch recordings. Right panels show corresponding morpho- show the results for individual cells. (Right) Comparison of population
logical properties of the recorded cells. Inset on top schematically illustrates input-output curves in the presence of inhibition (solid circles) and after a
stimulation procedure. Data are from (74). (D) Feedforward inhibition shortens block of inhibition (open circles). Although PV+ interneurons are likely to
the coincidence detection window in CA1 pyramidal neurons. (Left) Traces play an important role, other interneuron types may also be involved. Data
indicate action potentials recorded in the cell-attached configuration after are from (74).

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-7


R ES E A RC H | R E V IE W

inhibition has to be fast enough to ensure that a (independently of the distribution of excitation), interneurons in the hippocampus substantially
substantial inhibitory conductance is generated and this percentage is determined by the ratio changes during network oscillations, such as
before action potentials are initiated in princi- of the delay of disynaptic inhibition over the theta (4 to 10 Hz), gamma (40 to 100 Hz), and
pal neurons. membrane time constant of the principal cells ripple activity (140 to 200 Hz). In the absence of
PV+ interneurons are also involved in feedback (76) (Fig. 5E). Thus, the fast signaling properties oscillatory activity, action potential frequency is
(recurrent and lateral) inhibition (Fig. 5). Recip- of PV+ interneurons, which define the delay of low (6.5 Hz) (87). During theta oscillations, action
rocal coupling between principal neurons and disynaptic inhibition, are critically important for potential frequency markedly increases (21 Hz).
fast-spiking interneurons has been demonstrated the winner-takes-all mechanism. During sharp wave ripples (SWRs), the firing fre-
in several circuits, including the hippocampus It is often assumed that connectivity in feed- quency increases by more than one order of mag-
and the entorhinal cortex (23, 75) (Fig. 5, B and forward or feedback inhibitory microcircuits is nitude to 122 Hz [(87, 88), see (89) for similar
C). Antidromic activation of hippocampal CA1 random and that the properties of inhibitory in vitro data] (Fig. 6A). The massive activation of
pyramidal neurons by alveus stimulation gener- synaptic transmission are uniform (83). How- PV+ interneurons during SWRs (87) may be ex-
ates substantial inhibition after both single ever, this may not be the case. PV+ basket cells plained by the ability of these cells to efficiently
stimuli and high-frequency trains (73) (Fig. 5D). in the hippocampal CA1 region receive stronger respond to synchronous distributed input, which
Early inhibition is primarily mediated by periso- excitatory input from superficial pyramidal neu- will be generated by pyramidal cell activity during
matic inhibitory interneurons, whereas late in- rons but provide more powerful output to deep SWRs (90). Further, the high action potential fre-
hibition during trains is primarily mediated by pyramidal neurons (84). Inhibition at the output quency in PV+ interneurons in this network state
dendritic inhibitory interneurons, as shown by synapses of PV+ interneurons in the dentate gyrus (87) demonstrates that PV+ interneurons make
the different slope of the IPSP at somatic and has distance-dependent properties, with stron- use of their fast-spiking phenotype (7) under in
dendritic recording sites (73) (Fig. 5D). Quadruple ger and faster inhibition at short distances and vivo conditions.
recording in the entorhinal cortex also suggests weaker and slower inhibition at long distances An alternative way to approach the question
a substantial contribution of PV+ interneurons (85). In addition, inhibition at the output synapses is to examine the activity of PV+ interneurons
to both recurrent and lateral inhibition (75). on PV+ interneurons is stronger in target cells following adequate stimuli. In the hippocampus,
Feedback and lateral microcircuits involving with strong synaptic excitation and high acti- a substantial proportion of neurons are place
PV+ cells may have several functions beyond sim- vity levels (86). Thus, PV+ interneurons are sub- cells, so the location of the animal appears to be
ple inhibition. Feedback inhibition implements a stantially more than simple network stabilizers. the adequate stimulus. Whereas pyramidal neu-
winner-takes-all mechanism (76, 77) (Fig. 5E): They contribute to advanced computations in mi- rons show narrow place fields, PV+ interneurons
Once the principal cells with the strongest input crocircuits and neuronal networks. Furthermore, have much broader place fields (91, 92) (Fig. 6B).
fire, action potential initiation in the remaining they are not randomly and uniformly connected Similarly, whereas a large proportion of stellate
cells is inhibited. This computation could be par- but are embedded in microcircuits according to cells in the entorhinal cortex are grid cells (80),
ticularly important in the dentate gyrus, where specific rules. PV+ interneurons show substantially broader spa-
it may contribute to sparsification of activity tial tuning (81, 82) (Fig. 6B). In the primary visual
(78), pattern separation (79), and grid-to-place Activity of PV+ interneurons in vivo cortex, neurons are often sensitive to both orien-
code conversion (77, 80). Under certain condi- A central question in neuroscience is how specific tation and contrast of the stimulus, for example,
tions, network models with recurrent inhibitory neuron types shape higher brain functions, up to when animals are exposed to drifting gratings.
connectivity are also able to generate grid cell the level of animal behavior. The distinct exper- However, PV+ interneurons exhibit broader ori-
response patterns (75, 81) (Fig. 5F), which may imental accessibility and the detailed knowledge entation tuning and weaker contrast specificity
suggest that PV+ interneurons contribute to the about the cellular properties of PV+ interneurons than pyramidal neurons [(9395), but see (96)]
grid cell activity of stellate cells in the ento- may give us a chance, for the first time, to rigo- (Fig. 6C). In all of these cases, the broad tuning of
rhinal cortex [but see (82)]. rously address this question. PV+ interneurons may be explained by the con-
In the winner-takes-all feedback inhibitory One way to approach the problem is to ex- vergent input from a large number of principal
microcircuit (76, 77), it has been suggested that amine the activity of PV+ interneurons in vivo in neurons with a wide range of spatial or orienta-
the cells that receive excitation within a cer- awake, behaving animals during ongoing net- tion preferences. The action potential frequency
tain percentage of the maximum excitation fire work activity (Fig. 6, A to C). The activity of PV+ of PV+ interneurons in both hippocampal and
neocortical circuits increases with running (88, 97).
Thus, PV+ interneurons may receive a velocity-
modulated input that could be used for path in-
Box 3. Feedforward and feedback microcircuits. tegration, for instance.

In feedforward inhibition, afferent glutamatergic axons activate principal cells and interneurons The role of PV+ interneurons in network
in parallel (Fig. 4A). In feedback (recurrent and lateral) inhibition, afferent glutamatergic axons function and animal behavior in vivo
activate principal cells, which then activate interneurons in series (Fig. 5A). A more direct approach to bridge the gap be-
Feedback inhibition must be further subdivided into recurrent and lateral inhibition. tween cellular and network level is to interfere
However, the distinction between these two forms often remains fuzzy. Multiple cell recording with the activity of PV+ interneurons in vivo and
provides an elegant way to quantitatively distinguish the two forms. Recurrent inhibition in a to examine the behavioral consequences in awake
network can be defined as the proportion of principal cells that provide excitation to and animals (Fig. 6, D to F). In principle, optogenetics
receive inhibition from a given PV+ interneuron, whereas lateral inhibition is the proportion of [e.g., channelrhodopsin (11), halorhodopsin
principal cells that do not provide excitation to but receive inhibition from a given PV+ (12), or archaerhodopsin] or pharmacogenetics
interneuron (75) (Fig. 5C). [e.g., pharmacologically selective effector mole-
The speed of both feedforward and feedback inhibition is impressive; the latency of cules in combination with chimeric ligand-gated
disynaptic inhibition under physiological conditions is only 2 ms or less (71, 72). This high ion channels (98)] allow researchers to interfere
speed may be unexpected, because inhibition is composed of several steps (IN excitation via with the function of PV+ interneurons in either
PN-IN synapses propagation of EPSPs to the soma action potential initiation in the axon the positive or negative direction.
initial segment action potential propagation into the IN axon GABA release PN Such experiments are beginning to provide
inhibition via GABAergic output synapses). The fast signaling properties of PV+ interneurons insight into the function of PV+ interneurons at
(Table 1) play a key role in minimizing the delay. both network and behavioral levels. Optogenetic
manipulation of PV+ interneuron activity showed

1255263-8 1 AUGUST 2014 VOL 345 ISSUE 6196 sciencemag.org SCIENCE


RE S EAR CH | R E V I E W

that they are necessary and sufficient for the lations (12). This confirms a major role of PV+ toward the center of a place field, the action
generation of network oscillations in both the interneurons in network oscillations, as previ- potentials in pyramidal neurons shift to earlier
hippocampus and neocortex. Stimulation of PV+ ously inferred from in vitro data (100). phases of the theta cycle (101). When PV+ inter-
cells at theta frequency induces theta spike res- Experimental manipulation of PV+ interneu- neurons are inhibited, the steepness of the phase-
onance in CA1 pyramidal cells (99), and stim- ron activity in the hippocampus demonstrated position relation becomes reduced (92) (Fig. 6D).
ulation of PV+ cells at gamma frequency leads that PV+ interneurons regulate both the precise Thus, PV+ interneurons regulate the precise timing
to gamma oscillations in the local field potential shape of place fields and the phenomenon of of action potential initiation in a pyramidal neu-
in the somatosensory cortex (11, 12). Conversely, phase precession in CA1 pyramidal neurons (92) ron ensemble, probably making use of their fast
suppression of PV+ cells reduces gamma oscil- (Fig. 6D). As an animal moves from the periphery signaling properties.

A B C
DG EC 1 1 1
+ 3 4 3 4 3 4
PV IN 2 2 2
recurrent PN
1 1

2
H 2
CA3

3
100 m
GCL 3

4
4
lateral 40 mV
40 mV
200 ms 1 mV
200 ms
D E External
input
F
Alveus stimulation
CA1 Dendritic d Model
recording %max= max =
m
Somatic
GABAergic recording Model P1 P2 P3 Pn
interneuron
Pyramidal cell

1 stimulus I

15
Dendrite
10
k%

5
0
0 5 10 15 20 25 30
Soma Excitatory current (nA)
15 Theoretical value
E% max

10
60 cm
0.5 mV
50 ms 5
0
0 5 10 15 20 25 30
Excitatory current (nA)

Fig. 5. The role of PV+ interneurons in feedback inhibitory microcircuits. drite of a CA1 pyramidal neuron. The slope of rise is higher at the somatic
(A) Schematic illustration of the feedback inhibitory microcircuit. (Top) Re- versus the dendritic recording site (inset), indicating that inhibition is generated
current inhibition; (bottom) lateral inhibition. [Scheme courtesy of G. Buzski.] perisomatically. Data are from (73). (E) Feedback inhibition implements a
(B) Feedback inhibition in the dentate gyrus. Reciprocal coupling between a winner-takes-all mechanism. (Top) Scheme of the neuronal network with
fast-spiking (putative PV+) interneuron and a granule cell. Neurons were filled external input and feedback inhibition. (Bottom) Plot of the proportion of firing
with biocytin during recording and subsequently reconstructed. Black, soma cells (k%) and the proportion of cells that receive excitation within E% of the
and dendrites of PV+ interneuron; blue, axon of PV+ interneuron; green, soma maximal excitation (E% max). Line styles indicate different distributions of
and dendrites of granule cell; red, granule cell axon. Red dots, putative ex- inputs. Note that E% max is relatively independent of the stimulus current.
citatory synapses on interneurons; green dots, inhibitory synapses on granule Also note that E% max is d/tm, where d is the delay of disynaptic inhibition and
cells. H, hilus; GCL, granule cell layer. Data are from (23). (C) Feedback tm is the membrane time constant of the pyramidal cell. Thus, fast signaling in
inhibition in the entorhinal cortex (EC). Simultaneous recording from one fast- inhibitory interneurons is critically important for the computational properties
spiking (putative PV+) interneuron and three layer-two stellate cells. (Left) of the winner-takes-all mechanism. Data are from (76). (F) A continuous
Action potential phenotype of the four neurons. (Center) Action potentials in a attractor network model based on feedback inhibition may produce grid cell
stellate cell evoked EPSPs in the putative PV+ interneuron. (Right) Action po- activity patterns in excitatory neurons. (Top) Distribution of E-I and I-E synaptic
tentials in the interneuron evoked IPSPs in all three stellate cells. (Top) Schematic conductances in the model (E, excitatory; I, inhibitory). (Bottom) Pseudocolor
illustration of connectivity (circle, putative PV+ interneuron; hexagons, stellate representation of simulated activity of E cells in two-dimensional space. Cells
cells). Data are from (75). (D) Recurrent inhibition in the hippocampal CA1 receive background activation, theta-modulated input, velocity-modulated
region. (Top) Schematic illustration of the stimulation procedure. (Bottom) input, and place-cell input from the hippocampus. Synaptic output is shifted
Traces illustrating IPSPs simultaneously recorded in the soma and the den- according to the direction of preferred movement. Data are from (81).

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-9


R ES E A RC H | R E V IE W

A B C
V1

Putative
PV+
CA1 CA1 INs
PV+ IN

Frequency (Hz)

Frequency (Hz)
OSI=0.11
6 5

140
Frequency (Hz)

120
30 0 0
90 deg 0 1
Contrast
+
EC PV IN
15
PN

Frequency (Hz)

Frequency (Hz)
OSI=0.85
0 3 4
PN
Th

Sh

Lo wa
w
et

ar

os e
a

ci
l.
v

0 0
0 1
Optical probe 90 deg Contrast
Spines
Spots

D Treadmill
E F
Velcro

V1 AC
CA1
10 Light
Control

150 Arch ChR


Frequency (Hz)

0
Pyr cell response (%)

4 70
60

Freezing (%)
100
50
40
2 30
50 0 20
z-score

4 10
8 0
Baseline CS+
0
0
0 170 2 0 2 4 6
90 0 90 180 270
Position (cm) Direction (deg) Time (s)

Fig. 6. The role of PV+ interneurons in complex animal behavior. (A) Ac- running on a treadmill belt. (Top) Spike frequency of CA1 pyramidal neurons
tivity of PV+ interneurons in the hippocampal CA1 region of freely moving versus location. (Bottom) Theta phase of action potentials versus location.
rats. (Top) Firing of a PV+ interneuron during SWRs. Upper trace, local field Blue, control data; red, data after light pulses, leading to halorhodopsin-
potential (filtered 130230 Hz); lower trace, unit activity from a PV+ inter- mediated inhibition of PV+ interneurons. Data are from (92). (E) PV+ inter-
neuron. Asterisks indicate ripples. (Bottom) Summary bar graph indicating neuron activity regulates the gain of orientation-selective visual responses.
the mean action potential frequency. Note that the action potential fre- Tuning curves of pyramidal neuron activity in the primary visual cortex
quency of PV+ interneurons is >100 Hz during SWRs. Error bars denote during a visual stimulus (drifting grating) are shown. Black, control condi-
SEM. Data are from (87). (B) Spatial firing of PV+ interneurons and principal tions; red, after activation of PV+ interneurons with channelrhodopsin;
neurons. (Top) Place cell firing of hippocampal neurons; (bottom) grid cell green, after suppression of PV+ interneurons with archaerhodopsin. Error
firing of entorhinal cortex neurons. Warm colors indicate high action poten- bars indicate bootstrapped 95% confidence intervals. Data are from (95).
tial frequency. Note that putative PV+ interneurons (PV+ INs) have broader (F) Role of PV+ interneurons in associative auditory fear conditioning. (Left,
spatial fields than principal neurons (PNs). Data are from (82, 91). (C) top) Activity of PV+ interneuron in the auditory cortex during unconditional
Orientation selectivity (left) and contrast sensitivity (right) of layer two and stimuli (foot shocks, cell-attached recording). (Left, bottom) Summary of
three pyramidal neurons in the primary visual cortex. Top graphs, PV+ in- activity changes of PV+ interneurons during foot shocks (z score). (Right)
terneuron; bottom graphs, principal neuron. Note the broad orientation Behavioral freezing of the animal in baseline conditions and in the presence
specificity and shallow contrast sensitivity of PV+ interneurons in compar- of the conditioned stimulus (CS+, tone). Black, sham control; blue, with
ison to pyramidal neurons. V1, primary visual cortex; OSI, orientation selec- optogenetic stimulation of PV+ interneurons; red, after reconditioning with-
tivity index. Error bars indicate SEM. Data are from (95). (D) PV+ interneuron out optogenetic stimulation. Data are from (104). AC, auditory cortex. Error
activity regulates place field shape and phase precession in headfixed mice bars, SEM; asterisks, significance (P < 0.001).

1255263-10 1 AUGUST 2014 VOL 345 ISSUE 6196 sciencemag.org SCIENCE


RE S EAR CH | R E V I E W

Manipulation of PV+ interneuron activity re- phenotype (110113). Thus, haploinsufficiency 7. B. Rudy, C. J. McBain, Kv3 channels: Voltage-gated K+
vealed that PV+ interneurons regulate the gain of the SCN1A gene in PV+ interneurons appears channels designed for high-frequency repetitive firing.
Trends Neurosci. 24, 517526 (2001). doi: 10.1016/
of sensory responses (94, 95, 102) (Fig. 6E). In to be the cause of the disease. As NaV1.1 mRNA S0166-2236(00)01892-0; pmid: 11506885
the primary visual cortex, activation or inacti- is highly expressed in PV+ interneurons (51), 8. M. J. Bezaire, I. Soltesz, Quantitative assessment of CA1
vation of PV+ interneurons changes the gain of and NaV1.1 immunoreactivity is primarily present local circuits: Knowledge base for interneuron-pyramidal
orientation tuning curves in pyramidal neurons, in the axons of these cells (49), both fast-spiking cell connectivity. Hippocampus 23, 751785 (2013).
doi: 10.1002/hipo.22141; pmid: 23674373
whereas the width of these curves remains largely action potential phenotype and fast signal prop- 9. M. R. Celio, Parvalbumin in most g-aminobutyric acidcontaining
unchanged [(94, 95), but see (102)]. Such a gain agation in the axon will be affected. Further- neurons of the rat cerebral cortex. Science 231, 995997
modulation would be consistent with an under- more, missense mutations in the NaV1.1/SCN1A (1986). doi: 10.1126/science.3945815; pmid: 3945815
lying feedback inhibition mechanism (76). Sim- gene have been identified in GEFS+ patients. In 10. E. Eggermann, P. Jonas, How the slow Ca2+ buffer parvalbumin
affects transmitter release in nanodomain-coupling regimes.
ilarly, in the barrel cortex, activation of PV+ one of the most common mutations, the Na+ Nat. Neurosci. 15, 2022 (2012). doi: 10.1038/nn.3002;
interneurons changes the amplitude of sensory channel inactivation curve is left-shifted, which pmid: 22138646
responses evoked by whisker stimulation (11). will reduce the number of available Na+ chan- 11. J. A. Cardin et al., Driving fast-spiking cells induces
However, the changes are more complex in this nels (114). Thus, we increasingly understand the gamma rhythm and controls sensory responses. Nature
459, 663667 (2009). doi: 10.1038/nature08002;
system because the number of spikes, latency, relations between the fast signaling properties pmid: 19396156
and spike precision are affected (11). of PV+ interneurons and a neurological disease 12. V. S. Sohal, F. Zhang, O. Yizhar, K. Deisseroth, Parvalbumin
Finally, PV+ interneurons are involved in the phenotype. neurons and gamma rhythms enhance cortical
regulation of plasticity and learning (69, 103106) Another link leads from the receptor tyrosine circuit performance. Nature 459, 698702 (2009).
doi: 10.1038/nature07991; pmid: 19396159
(Fig. 6F). PV+ interneurons in the visual cortex are kinase ErbB4 (epidermal growth factor recep- 13. L. Tricoire et al., A blueprint for the spatiotemporal origins
transiently inhibited after monocular deprivation; tor 4)a protein selectively expressed in PV+ of mouse hippocampal interneuron diversity. J. Neurosci.
this down-regulation appears to be necessary to interneurons (basket cells and axo-axonic cells) 31, 1094810970 (2011). doi: 10.1523/JNEUROSCI.0323-
enable ocular dominance plasticity in the critical of several brain regions (115)to schizophrenia 11.2011; pmid: 21795545
14. H. Taniguchi, J. Lu, Z. J. Huang, The spatial and temporal
period (103, 106). Furthermore, PV+ interneurons (116). Mutations in both the ErbB4 gene and the origin of chandelier cells in mouse neocortex. Science
in the auditory cortex are inhibited by aversive gene of neuregulin 1, the putative ErbB4 ligand, 339, 7074 (2013). doi: 10.1126/science.1227622;
foot shocks in an auditory fear-conditioning pa- are frequently found in schizophrenic patients. pmid: 23180771
radigm, and this inhibition plays a critical role Furthermore, mouse models with general or 15. G. Bartolini, G. Ciceri, O. Marn, Integration of GABAergic
interneurons into cortical cell assemblies: Lessons
for associative fear learning (104, 105) (Fig. 6F). PV+ interneuron-specific genetic elimination of from embryos and adults. Neuron 79, 849864 (2013).
Additionally, stimulation of PV+ interneurons in ErbB4 replicate aspects of the disease pheno- doi: 10.1016/j.neuron.2013.08.014; pmid: 24012001
the prefrontal cortex accelerates extinction of type (117). However, even in such a clearly defined 16. F. Crick, in What Mad Pursuit: A Personal View of
reward seeking behavior (107). In summary, sup- case, the underlying mechanisms are highly Scientific Discovery (Basic Books, New York, 1988),
p. 150.
pression of PV+ interneurons (i.e., disinhibition of complex. Genetic elimination of ErbB4 impairs 17. A. I. Gulys, M. Megas, Z. Emri, T. F. Freund, Total number
pyramidal neurons) is necessary for certain forms both the excitatory synaptic input (115, 118, 119) and ratio of excitatory and inhibitory synapses converging
of learning, whereas activation of PV+ cells may and the inhibitory synaptic output of PV+ inter- onto single interneurons of different types in the CA1
promote extinction. Recent results further sug- neurons (115, 119). Consistent with the hypothesis area of the rat hippocampus. J. Neurosci. 19, 1008210097
(1999). pmid: 10559416
gest that not only do PV+ interneurons regulate of deficient PV+ interneuron excitation, ErbB4 18. A. Nrenberg, H. Hu, I. Vida, M. Bartos, P. Jonas, Distinct
learning, but learning also induces plastic changes immunoreactivity is located in postsynaptic den- nonuniform cable properties optimize rapid and efficient
in PV+ interneurons (69). Thus, the involvement sities of glutamatergic input synapses on dendrites activation of fast-spiking GABAergic interneurons. Proc. Natl.
of PV+ interneurons in learning is bidirectional. of PV+ interneurons (87, 115). Finally, neuregu- Acad. Sci. U.S.A. 107, 894899 (2010). doi: 10.1073/
pnas.0910716107; pmid: 20080772
lin 1 application impairs the fast-spiking action 19. Y. Kubota et al., Conserved properties of dendritic trees
The role of PV+ interneurons in potential phenotype by regulating Na+ and Kv1 in four cortical interneuron subtypes. Sci. Rep. 1, 89 (2011).
neurological and psychiatric diseases K+ channels (120, 121). Thus, we are beginning to doi: 10.1038/srep00089; pmid: 22355608
Another major challenge in neuroscience is to understand the relations between the fast sig- 20. J. J. Tukker et al., Distinct dendritic arborization and
in vivo firing patterns of parvalbumin-expressing basket
understand how specific neuron types are in- naling properties of PV+ interneurons and the cells in the hippocampal area CA3. J. Neurosci. 33,
volved in neurological or psychiatric diseases. extremely complex phenotype of a psychiatric 68096825 (2013). doi: 10.1523/JNEUROSCI.5052-12.2013;
What are the links between PV+ interneurons disease. pmid: 23595740
21. H. Hioki et al., Cell type-specific inhibitory inputs to dendritic
and brain diseases? The problem is not in
and somatic compartments of parvalbumin-expressing
finding the links, but instead that there are neocortical interneuron. J. Neurosci. 33, 544555 (2013).
RE FERENCES AND NOTES
too many! In several neurological and psychiat- doi: 10.1523/JNEUROSCI.2255-12.2013; pmid: 23303934
1. T. F. Freund, G. Buzski, Interneurons of the hippocampus.
ric diseases, the function of PV+ interneurons Hippocampus 6, 347470 (1996). doi: 10.1002/(SICI)1098-
22. C. K. Pfeffer, M. Xue, M. He, Z. J. Huang, M. Scanziani,
Inhibition of inhibition in visual cortex: The logic of
appears to be altered. These include epilepsy, 1063(1996)6:4<347::AID-HIPO1>3.0.CO;2-I; pmid: 8915675 connections between molecularly distinct interneurons.
schizophrenia, depression, autism, and Alzheimers 2. Y. Aika, J. Q. Ren, K. Kosaka, T. Kosaka, Quantitative analysis Nat. Neurosci. 16, 10681076 (2013). doi: 10.1038/nn.3446;
disease (108). The detailed knowledge about of GABA-like-immunoreactive and parvalbumin-containing pmid: 23817549
neurons in the CA1 region of the rat hippocampus using a
PV+ interneurons at the molecular, cellular, and stereological method, the disector. Exp. Brain Res. 99,
23. J. R. P. Geiger, J. Lbke, A. Roth, M. Frotscher, P. Jonas,
Submillisecond AMPA receptor-mediated signaling at a
network levels may now help us to focus on the 267276 (1994). doi: 10.1007/BF00239593; pmid: 7925807 principal neuron-interneuron synapse. Neuron 18,
most important relations, where the disease 3. K. Halasy, P. Somogyi, Distribution of GABAergic synapses 10091023 (1997). doi: 10.1016/S0896-6273(00)80339-6;
gene is selectively expressed in PV+ cells and the and their targets in the dentate gyrus of rat: A quantitative pmid: 9208867
immunoelectron microscopic analysis. J. Hirnforsch. 34, 24. A. Sik, M. Penttonen, A. Ylinen, G. Buzski, Hippocampal
phenotype can be replicated by restricted disease 299308 (1993). pmid: 8270784 CA1 interneurons: An in vivo intracellular labeling study.
gene expression in PV+ interneurons. 4. G. Westbrook, Seizures and epilepsy, in Principles of J. Neurosci. 15, 66516665 (1995). pmid: 7472426
One recently identified link leads from the Neural Science, E. Kandel, J. H. Schwartz, T. M. Jessell, 25. F. Karube, Y. Kubota, Y. Kawaguchi, Axon branching and
axonal NaV1.1 channel in PV+ interneurons to S. Siegelbaum, A. J. Hudspeth, Eds. (McGraw-Hill, New York, synaptic bouton phenotypes in GABAergic nonpyramidal
2013), pp. 11161139. cell subtypes. J. Neurosci. 24, 28532865 (2004).
Dravet syndrome [severe myoclonic epilepsy of 5. T. Klausberger, P. Somogyi, Neuronal diversity and temporal doi: 10.1523/JNEUROSCI.4814-03.2004; pmid: 15044524
infancy (SMEI)] and GEFS+ (generalized epi- dynamics: The unity of hippocampal circuit operations. 26. I. Parnas, I. Segev, A mathematical model for conduction of
lepsy with febrile seizures plus) (109). Trunca- Science 321, 5357 (2008). doi: 10.1126/science.1149381; action potentials along bifurcating axons. J. Physiol. 295,
tion mutations in the NaV1.1/SCN1A gene have pmid: 18599766 323343 (1979). pmid: 521942
6. J. DeFelipe et al., New insights into the classification and 27. H. Hu, M. Martina, P. Jonas, Dendritic mechanisms
been identified in SMEI patients, and mouse nomenclature of cortical GABAergic interneurons. Nat. underlying rapid synaptic activation of fast-spiking
models with general or PV+ interneuron-selective Rev. Neurosci. 14, 202216 (2013). doi: 10.1038/nrn3444; hippocampal interneurons. Science 327, 5258 (2010).
deletion of the SCN1A gene replicate the disease pmid: 23385869 doi: 10.1126/science.1177876; pmid: 19965717

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-11


R ES E A RC H | R E V IE W

28. J. H. Goldberg, G. Tamas, R. Yuste, Ca2+ imaging of mouse 25, 52305235 (2005). doi: 10.1523/JNEUROSCI.0722- properties in subsets of neurons. Neuron 54, 567581
neocortical interneurone dendrites: Ia-type K+ channels 05.2005; pmid: 15917463 (2007). doi: 10.1016/j.neuron.2007.05.004; pmid: 17521570
control action potential backpropagation. J. Physiol. 551, 49. I. Ogiwara et al., Nav1.1 localizes to axons of parvalbumin- 68. B. Schwaller, M. Meyer, S. Schiffmann, New functions for
4965 (2003). doi: 10.1113/jphysiol.2003.042580; positive inhibitory interneurons: A circuit basis for old proteins: The role of the calcium-binding proteins
pmid: 12844506 epileptic seizures in mice carrying an Scn1a gene mutation. calbindin D-28k, calretinin and parvalbumin, in cerebellar
29. Y. Aponte, J. Bischofberger, P. Jonas, Efficient Ca2+ buffering J. Neurosci. 27, 59035914 (2007). doi: 10.1523/ physiology. Studies with knockout mice. Cerebellum 1,
in fast-spiking basket cells of rat hippocampus. J. Physiol. JNEUROSCI.5270-06.2007; pmid: 17537961 241258 (2002). doi: 10.1080/147342202320883551;
586, 20612075 (2008). doi: 10.1113/jphysiol.2007.147298; 50. A. Lorincz, Z. Nusser, Cell-type-dependent molecular pmid: 12879963
pmid: 18276734 composition of the axon initial segment. J. Neurosci. 28, 69. F. Donato, S. B. Rompani, P. Caroni, Parvalbumin-expressing
30. O. Camir, L. Topolnik, Dendritic calcium nonlinearities 1432914340 (2008). doi: 10.1523/JNEUROSCI.4833- basket-cell network plasticity induced by experience
switch the direction of synaptic plasticity in fast-spiking 08.2008; pmid: 19118165 regulates adult learning. Nature 504, 272276 (2013).
interneurons. J. Neurosci. 34, 38643877 (2014). 51. B. W. Okaty, M. N. Miller, K. Sugino, C. M. Hempel, S. B. Nelson, doi: 10.1038/nature12866; pmid: 24336286
doi: 10.1523/JNEUROSCI.2253-13.2014; pmid: 24623765 Transcriptional and electrophysiological maturation of 70. G. Buzski, E. Eidelberg, Commissural projection to the
31. B. Chiovini et al., Dendritic spikes induce ripples in neocortical fast-spiking GABAergic interneurons. J. Neurosci. dentate gyrus of the rat: Evidence for feed-forward inhibition.
parvalbumin interneurons during hippocampal sharp waves. 29, 70407052 (2009). doi: 10.1523/JNEUROSCI.0105- Brain Res. 230, 346350 (1981). doi: 10.1016/0006-
Neuron 82, 908924 (2014). doi: 10.1016/j.neuron. 09.2009; pmid: 19474331 8993(81)90413-3; pmid: 7317783
2014.04.004; pmid: 24853946 52. M. Weiser et al., The potassium channel subunit KV3.1b is 71. R. Miles, Synaptic excitation of inhibitory cells by single
32. G. J. Stuart, B. Sakmann, Active propagation of somatic localized to somatic and axonal membranes of specific CA3 hippocampal pyramidal cells of the guinea-pig in vitro.
action potentials into neocortical pyramidal cell dendrites. populations of CNS neurons. J. Neurosci. 15, 42984314 J. Physiol. 428, 6177 (1990). pmid: 2231426
Nature 367, 6972 (1994). doi: 10.1038/367069a0; (1995). pmid: 7790912 72. F. Pouille, M. Scanziani, Enforcement of temporal fidelity in
pmid: 8107777 53. E. M. Goldberg et al., K+ channels at the axon initial pyramidal cells by somatic feed-forward inhibition. Science
33. N. L. Golding, N. Spruston, Dendritic sodium spikes are segment dampen near-threshold excitability of neocortical 293, 11591163 (2001). doi: 10.1126/science.1060342;
variable triggers of axonal action potentials in hippocampal fast-spiking GABAergic interneurons. Neuron 58, 387400 pmid: 11498596
CA1 pyramidal neurons. Neuron 21, 11891200 (1998). (2008). doi: 10.1016/j.neuron.2008.03.003; pmid: 18466749 73. F. Pouille, M. Scanziani, Routing of spike series by dynamic
doi: 10.1016/S0896-6273(00)80635-2; pmid: 9856473 54. E. Campanac et al., Enhanced intrinsic excitability in circuits in the hippocampus. Nature 429, 717723 (2004).
34. M. Martina, I. Vida, P. Jonas, Distal initiation and active basket cells maintains excitatory-inhibitory balance in doi: 10.1038/nature02615; pmid: 15170216
propagation of action potentials in interneuron dendrites. hippocampal circuits. Neuron 77, 712722 (2013). 74. F. Pouille, A. Marin-Burgin, H. Adesnik, B. V. Atallah,
Science 287, 295300 (2000). doi: 10.1126/ doi: 10.1016/j.neuron.2012.12.020; pmid: 23439123 M. Scanziani, Input normalization by global feedforward
science.287.5451.295; pmid: 10634782 55. U. Kraushaar, P. Jonas, Efficacy and stability of quantal inhibition expands cortical dynamic range. Nat. Neurosci. 12,
35. K. Vervaeke, A. Lorincz, Z. Nusser, R. A. Silver, Gap junctions GABA release at a hippocampal interneuron-principal 15771585 (2009). doi: 10.1038/nn.2441; pmid: 19881502
compensate for sublinear dendritic integration in an neuron synapse. J. Neurosci. 20, 55945607 (2000). 75. J. J. Couey et al., Recurrent inhibitory circuitry as a
inhibitory network. Science 335, 16241628 (2012). pmid: 10908596 mechanism for grid formation. Nat. Neurosci. 16, 318324
doi: 10.1126/science.1215101; pmid: 22403180 56. L. Li, J. Bischofberger, P. Jonas, Differential gating and (2013). doi: 10.1038/nn.3310; pmid: 23334580
36. J. R. P. Geiger et al., Relative abundance of subunit mRNAs recruitment of P/Q-, N-, and R-type Ca2+ channels in 76. L. de Almeida, M. Idiart, J. E. Lisman, A second function of
determines gating and Ca2+ permeability of AMPA receptors hippocampal mossy fiber boutons. J. Neurosci. 27, gamma frequency oscillations: An E%-max winner-take-all
in principal neurons and interneurons in rat CNS. Neuron 1342013429 (2007). doi: 10.1523/JNEUROSCI.1709-07.2007; mechanism selects which cells fire. J. Neurosci. 29,
15, 193204 (1995). doi: 10.1016/0896-6273(95)90076-4; pmid: 18057200 74977503 (2009). doi: 10.1523/JNEUROSCI.6044-08.2009;
pmid: 7619522 57. S. Hefft, P. Jonas, Asynchronous GABA release generates pmid: 19515917
37. M. Galarreta, S. Hestrin, Spike transmission and synchrony long-lasting inhibition at a hippocampal interneuron-principal 77. L. de Almeida, M. Idiart, J. E. Lisman, The input-output
detection in networks of GABAergic interneurons. Science neuron synapse. Nat. Neurosci. 8, 13191328 (2005). transformation of the hippocampal granule cells: From
292, 22952299 (2001). doi: 10.1126/science.1061395; doi: 10.1038/nn1542; pmid: 16158066 grid cells to place fields. J. Neurosci. 29, 75047512 (2009).
pmid: 11423653 58. A. V. Zaitsev, N. V. Povysheva, D. A. Lewis, L. S. Krimer, doi: 10.1523/JNEUROSCI.6048-08.2009; pmid: 19515918
38. D. Fricker, R. Miles, EPSP amplification and the precision of P/Q-type, but not N-type, calcium channels mediate GABA 78. A. J. Perna-Andrade, P. Jonas, Theta-gamma-modulated
spike timing in hippocampal neurons. Neuron 28, 559569 release from fast-spiking interneurons to pyramidal cells in synaptic currents in hippocampal granule cells in vivo define
(2000). doi: 10.1016/S0896-6273(00)00133-1; rat prefrontal cortex. J. Neurophysiol. 97, 35673573 (2007). a mechanism for network oscillations. Neuron 81, 140152
pmid: 11144364 doi: 10.1152/jn.01293.2006; pmid: 17329622 (2014). doi: 10.1016/j.neuron.2013.09.046; pmid: 24333053
39. M. Galarreta, S. Hestrin, A network of fast-spiking cells in 59. I. Bucurenciu, J. Bischofberger, P. Jonas, A small number 79. J. K. Leutgeb, S. Leutgeb, M. B. Moser, E. I. Moser, Pattern
the neocortex connected by electrical synapses. Nature 402, of open Ca2+ channels trigger transmitter release at a separation in the dentate gyrus and CA3 of the hippocampus.
7275 (1999). doi: 10.1038/47029; pmid: 10573418 central GABAergic synapse. Nat. Neurosci. 13, 1921 (2010). Science 315, 961966 (2007). doi: 10.1126/science.1135801;
40. J. R. Gibson, M. Beierlein, B. W. Connors, Two networks of doi: 10.1038/nn.2461; pmid: 20010820 pmid: 17303747
electrically coupled inhibitory neurons in neocortex. Nature 60. E. Rossignol, I. Kruglikov, A. M. van den Maagdenberg, 80. T. Hafting, M. Fyhn, S. Molden, M. B. Moser, E. I. Moser,
402, 7579 (1999). doi: 10.1038/47035; pmid: 10573419 B. Rudy, G. Fishell, CaV 2.1 ablation in cortical interneurons Microstructure of a spatial map in the entorhinal cortex.
41. G. Tams, E. H. Buhl, A. Lrincz, P. Somogyi, Proximally selectively impairs fast-spiking basket cells and causes Nature 436, 801806 (2005). doi: 10.1038/nature03721;
targeted GABAergic synapses and gap junctions synchronize generalized seizures. Ann. Neurol. 74, 209222 (2013). pmid: 15965463
cortical interneurons. Nat. Neurosci. 3, 366371 (2000). pmid: 23595603 81. H. Pastoll, L. Solanka, M. C. van Rossum, M. F. Nolan,
doi: 10.1038/73936; pmid: 10725926 61. I. Bucurenciu, A. Kulik, B. Schwaller, M. Frotscher, P. Jonas, Feedback inhibition enables y-nested g oscillations and grid
42. M. Bartos, I. Vida, M. Frotscher, J. R. P. Geiger, P. Jonas, Nanodomain coupling between Ca2+ channels and Ca2+ firing fields. Neuron 77, 141154 (2013). doi: 10.1016/
Rapid signaling at inhibitory synapses in a dentate gyrus sensors promotes fast and efficient transmitter release at a j.neuron.2012.11.032; pmid: 23312522
interneuron network. J. Neurosci. 21, 26872698 (2001). cortical GABAergic synapse. Neuron 57, 536545 (2008). 82. C. Buetfering, K. Allen, H. Monyer, Parvalbumin interneurons
pmid: 11306622 doi: 10.1016/j.neuron.2007.12.026; pmid: 18304483 provide grid cell-driven recurrent inhibition in the medial
43. H. Hu, P. Jonas, A supercritical density of Na+ channels 62. E. Eggermann, I. Bucurenciu, S. P. Goswami, P. Jonas, entorhinal cortex. Nat. Neurosci. 17, 710718 (2014).
ensures fast signaling in GABAergic interneuron axons. Nanodomain coupling between Ca2+ channels and sensors of doi: 10.1038/nn.3696; pmid: 24705183
Nat. Neurosci. 17, 686693 (2014). doi: 10.1038/nn.3678; exocytosis at fast mammalian synapses. Nat. Rev. Neurosci. 83. A. M. Packer, R. Yuste, Dense, unspecific connectivity of
pmid: 24657965 13, 721 (2012). doi: 10.1038/nrn3125; pmid: 22183436 neocortical parvalbumin-positive interneurons: A canonical
44. M. H. Kole et al., Action potential generation requires a high 63. Z. P. Pang et al., Synaptotagmin-2 is essential for survival microcircuit for inhibition? J. Neurosci. 31, 1326013271
sodium channel density in the axon initial segment. Nat. and contributes to Ca2+ triggering of neurotransmitter (2011). doi: 10.1523/JNEUROSCI.3131-11.2011;
Neurosci. 11, 178186 (2008). doi: 10.1038/nn2040; release in central and neuromuscular synapses. J. Neurosci. pmid: 21917809
pmid: 18204443 26, 1349313504 (2006). doi: 10.1523/JNEUROSCI.3519- 84. S. H. Lee et al., Parvalbumin-positive basket cells
45. J. P. Meeks, S. Mennerick, Action potential initiation and 06.2006; pmid: 17192432 differentiate among hippocampal pyramidal cells. Neuron
propagation in CA3 pyramidal axons. J. Neurophysiol. 97, 64. A. M. Kerr, E. Reisinger, P. Jonas, Differential dependence of 82, 11291144 (2014). doi: 10.1016/j.neuron.2014.03.034;
34603472 (2007). doi: 10.1152/jn.01288.2006; phasic transmitter release on synaptotagmin 1 at GABAergic pmid: 24836505
pmid: 17314237 and glutamatergic hippocampal synapses. Proc. Natl. Acad. 85. M. Strueber, P. Jonas, M. Bartos, Distance dependence of
46. C. Schmidt-Hieber, P. Jonas, J. Bischofberger, Action Sci. U.S.A. 105, 1558115586 (2008). doi: 10.1073/ efficacy and timing of synaptic inhibition in the hippocampal
potential initiation and propagation in hippocampal mossy pnas.0800621105; pmid: 18832148 network. program no. 655.10 2011, Neuroscience Meeting
fibre axons. J. Physiol. 586, 18491857 (2008). 65. M. Geppert et al., Synaptotagmin I: A major Ca2+ sensor for Planner (Society for Neuroscience, Washington, DC, 2011).
doi: 10.1113/jphysiol.2007.150151; pmid: 18258662 transmitter release at a central synapse. Cell 79, 717727 86. M. Xue, B. V. Atallah, M. Scanziani, Equalizing exciation
47. M. Martina, P. Jonas, Functional differences in Na+ channel (1994). doi: 10.1016/0092-8674(94)90556-8; pmid: 7954835 inhibition ratios across visual cortical neurons. Nature
gating between fast-spiking interneurones and principal 66. J. P. Sommeijer, C. N. Levelt, Synaptotagmin-2 is a reliable 511, 596600 (2014). doi: 10.1038/nature13321
neurones of rat hippocampus. J. Physiol. 505, 593603 marker for parvalbumin positive inhibitory boutons in the 87. D. Lapray et al., Behavior-dependent specialization of
(1997). doi: 10.1111/j.1469-7793.1997.593ba.x; pmid: 9457638 mouse visual cortex. PLOS ONE 7, e35323 (2012). identified hippocampal interneurons. Nat. Neurosci. 15,
48. E. M. Goldberg et al., Specific functions of synaptically doi: 10.1371/journal.pone.0035323; pmid: 22539967 12651271 (2012). doi: 10.1038/nn.3176; pmid: 22864613
localized potassium channels in synaptic transmission at the 67. J. Xu, T. Mashimo, T. C. Sdhof, Synaptotagmin-1, -2, and -9: 88. C. Varga, P. Golshani, I. Soltesz, Frequency-invariant
neocortical GABAergic fast-spiking cell synapse. J. Neurosci. Ca2+ sensors for fast release that specify distinct presynaptic temporal ordering of interneuronal discharges during

1255263-12 1 AUGUST 2014 VOL 345 ISSUE 6196 sciencemag.org SCIENCE


RE S EAR CH | R E V I E W

hippocampal oscillations in awake mice. Proc. Natl. Acad. 105. S. B. Wolff et al., Amygdala interneuron subtypes control fear 122. P. Somogyi, T. Klausberger, Defined types of cortical
Sci. U.S.A. 109, E2726E2734 (2012). doi: 10.1073/ learning through disinhibition. Nature 509, 453458 (2014). interneurone structure space and spike timing in the
pnas.1210929109; pmid: 23010933 doi: 10.1038/nature13258; pmid: 24814341 hippocampus. J. Physiol. 562, 926 (2005). doi: 10.1113/
89. N. Hjos et al., Input-output features of anatomically 106. Y. Yazaki-Sugiyama, S. Kang, H. Cteau, T. Fukai, T. K. Hensch, jphysiol.2004.078915; pmid: 15539390
identified CA3 neurons during hippocampal sharp wave/ Bidirectional plasticity in fast-spiking GABA circuits by visual 123. A. M. Thomson, D. C. West, J. Hahn, J. Deuchars, Single axon
ripple oscillation in vitro. J. Neurosci. 33, 1167711691 (2013). experience. Nature 462, 218221 (2009). doi: 10.1038/ IPSPs elicited in pyramidal cells by three classes of
doi: 10.1523/JNEUROSCI.5729-12.2013; pmid: 23843535 nature08485; pmid: 19907494 interneurones in slices of rat neocortex. J. Physiol. 496,
90. J. Csicsvari, H. Hirase, A. Mamiya, G. Buzski, Ensemble 107. D. R. Sparta et al., Activation of prefrontal cortical 81102 (1996). pmid: 8910198
patterns of hippocampal CA3-CA1 neurons during parvalbumin interneurons facilitates extinction of reward- 124. D. A. McCormick, B. W. Connors, J. W. Lighthall, D. A. Prince,
sharp wave-associated population events. Neuron 28, seeking behavior. J. Neurosci. 34, 36993705 (2014). Comparative electrophysiology of pyramidal and sparsely
585594 (2000). doi: 10.1016/S0896-6273(00)00135-5; doi: 10.1523/JNEUROSCI.0235-13.2014; pmid: 24599468 spiny stellate neurons of the neocortex. J. Neurophysiol. 54,
pmid: 11144366 108. O. Marn, Interneuron dysfunction in psychiatric disorders. 782806 (1985). pmid: 2999347
91. M. A. Wilson, B. L. McNaughton, Dynamics of the Nat. Rev. Neurosci. 13, 107120 (2012). pmid: 22251963 125. S. Hippenmeyer et al., A developmental switch in the
hippocampal ensemble code for space. Science 261, 109. W. A. Catterall, F. Kalume, J. C. Oakley, NaV1.1 channels and response of DRG neurons to ETS transcription factor
10551058 (1993). doi: 10.1126/science.8351520; epilepsy. J. Physiol. 588, 18491859 (2010). doi: 10.1113/ signaling. PLOS Biol. 3, e159 (2005). doi: 10.1371/journal.
pmid: 8351520 jphysiol.2010.187484; pmid: 20194124 pbio.0030159; pmid: 15836427
92. S. Royer et al., Control of timing, rate and bursts of 110. F. H. Yu et al., Reduced sodium current in GABAergic 126. D. Kvitsiani et al., Distinct behavioural and network correlates
hippocampal place cells by dendritic and somatic inhibition. interneurons in a mouse model of severe myoclonic of two interneuron types in prefrontal cortex. Nature 498,
Nat. Neurosci. 15, 769775 (2012). doi: 10.1038/nn.3077; epilepsy in infancy. Nat. Neurosci. 9, 11421149 (2006). 363366 (2013). doi: 10.1038/nature12176; pmid: 23708967
pmid: 22446878 doi: 10.1038/nn1754; pmid: 16921370 127. P. Poirazi, B. W. Mel, Impact of active dendrites and
93. S. B. Hofer et al., Differential connectivity and response 111. C. S. Cheah et al., Specific deletion of NaV1.1 sodium structural plasticity on the memory capacity of neural tissue.
dynamics of excitatory and inhibitory neurons in visual channels in inhibitory interneurons causes seizures and Neuron 29, 779796 (2001). doi: 10.1016/S0896-6273(01)
cortex. Nat. Neurosci. 14, 10451052 (2011). doi: 10.1038/ premature death in a mouse model of Dravet syndrome. 00252-5; pmid: 11301036
nn.2876; pmid: 21765421 Proc. Natl. Acad. Sci. U.S.A. 109, 1464614651 (2012). 128. C. A. Bottoms, J. P. Schuermann, S. Agah, M. T. Henzl,
94. N. R. Wilson, C. A. Runyan, F. L. Wang, M. Sur, Division and doi: 10.1073/pnas.1211591109; pmid: 22908258 J. J. Tanner, Crystal structure of rat a-parvalbumin at 1.05
subtraction by distinct cortical inhibitory networks in vivo. 112. S. B. Dutton et al., Preferential inactivation of Scn1a in resolution. Protein Sci. 13, 17241734 (2004). doi: 10.1110/
Nature 488, 343348 (2012). doi: 10.1038/nature11347; parvalbumin interneurons increases seizure susceptibility. ps.03571004; pmid: 15169955
pmid: 22878717 Neurobiol. Dis. 49C, 211220 (2012). pmid: 22926190 129. M. Bartos et al., Fast synaptic inhibition promotes
95. B. V. Atallah, W. Bruns, M. Carandini, M. Scanziani, 113. I. Ogiwara et al., Nav1.1 haploinsufficiency in excitatory synchronized gamma oscillations in hippocampal interneuron
Parvalbumin-expressing interneurons linearly transform neurons ameliorates seizure-associated sudden death networks. Proc. Natl. Acad. Sci. U.S.A. 99, 1322213227
cortical responses to visual stimuli. Neuron 73, 159170 in a mouse model of Dravet syndrome. Hum. Mol. Genet. (2002). doi: 10.1073/pnas.192233099; pmid: 12235359
(2012). doi: 10.1016/j.neuron.2011.12.013; pmid: 22243754 22, 47844804 (2013). doi: 10.1093/hmg/ddt331; 130. M. Galarreta, S. Hestrin, Electrical and chemical synapses
96. C. A. Runyan et al., Response features of parvalbumin- pmid: 23922229 among parvalbumin fast-spiking GABAergic interneurons in
expressing interneurons suggest precise roles for subtypes of 114. T. Mashimo et al., A missense mutation of the gene encoding adult mouse neocortex. Proc. Natl. Acad. Sci. U.S.A. 99,
inhibition in visual cortex. Neuron 67, 847857 (2010). voltage-dependent sodium channel (Nav1.1) confers 1243812443 (2002). doi: 10.1073/pnas.192159599;
doi: 10.1016/j.neuron.2010.08.006; pmid: 20826315 susceptibility to febrile seizures in rats. J. Neurosci. 30, pmid: 12213962
97. P. O. Polack, J. Friedman, P. Golshani, Cellular mechanisms 57445753 (2010). doi: 10.1523/JNEUROSCI.3360-09.2010; 131. M. Martina, J. H. Schultz, H. Ehmke, H. Monyer, P. Jonas,
of brain state-dependent gain modulation in visual cortex. pmid: 20410126 Functional and molecular differences between voltage-gated
Nat. Neurosci. 16, 13311339 (2013). doi: 10.1038/nn.3464; 115. P. Fazzari et al., Control of cortical GABA circuitry development K+ channels of fast-spiking interneurons and pyramidal
pmid: 23872595 by Nrg1 and ErbB4 signalling. Nature 464, 13761380 (2010). neurons of rat hippocampus. J. Neurosci. 18, 81118125
98. C. J. Magnus et al., Chemical and genetic engineering of doi: 10.1038/nature08928; pmid: 20393464 (1998). pmid: 9763458
selective ion channel-ligand interactions. Science 333, 116. D. A. Lewis, T. Hashimoto, D. W. Volk, Cortical inhibitory 132. Y. Kawaguchi, H. Katsumaru, T. Kosaka, C. W. Heizmann,
12921296 (2011). doi: 10.1126/science.1206606; neurons and schizophrenia. Nat. Rev. Neurosci. 6, 312324 K. Hama, Fast spiking cells in rat hippocampus (CA1 region)
pmid: 21885782 (2005). doi: 10.1038/nrn1648; pmid: 15803162 contain the calcium-binding protein parvalbumin. Brain Res.
99. E. Stark et al., Inhibition-induced theta resonance in 117. L. Wen et al., Neuregulin 1 regulates pyramidal neuron 416, 369374 (1987). doi: 10.1016/0006-8993(87)90921-8;
cortical circuits. Neuron 80, 12631276 (2013). doi: 10.1016/ activity via ErbB4 in parvalbumin-positive interneurons. pmid: 3304536
j.neuron.2013.09.033; pmid: 24314731 Proc. Natl. Acad. Sci. U.S.A. 107, 12111216 (2010). 133. J. Du, L. Zhang, M. Weiser, B. Rudy, C. J. McBain,
100. M. Bartos, I. Vida, P. Jonas, Synaptic mechanisms of doi: 10.1073/pnas.0910302107; pmid: 20080551 Developmental expression and functional characterization of
synchronized gamma oscillations in inhibitory interneuron 118. Y. Abe, H. Namba, T. Kato, Y. Iwakura, H. Nawa, Neuregulin-1 the potassium-channel subunit Kv3.1b in parvalbumin-
networks. Nat. Rev. Neurosci. 8, 4556 (2007). doi: 10.1038/ signals from the periphery regulate AMPA receptor sensitivity containing interneurons of the rat hippocampus. J. Neurosci.
nrn2044; pmid: 17180162 and expression in GABAergic interneurons in developing 16, 506518 (1996). pmid: 8551335
101. J. OKeefe, M. L. Recce, Phase relationship between neocortex. J. Neurosci. 31, 56995709 (2011). doi: 10.1523/ 134. J. Szabadics et al., Excitatory effect of GABAergic axo-axonic
hippocampal place units and the EEG theta rhythm. JNEUROSCI.3477-10.2011; pmid: 21490211 cells in cortical microcircuits. Science 311, 233235 (2006).
Hippocampus 3, 317330 (1993). doi: 10.1002/ 119. I. Del Pino et al., Erbb4 deletion from fast-spiking doi: 10.1126/science.1121325; pmid: 16410524
hipo.450030307; pmid: 8353611 interneurons causes schizophrenia-like phenotypes. Neuron
102. S. H. Lee et al., Activation of specific interneurons 79, 11521168 (2013). doi: 10.1016/j.neuron.2013.07.010; AC KNOWLED GME NTS
improves V1 feature selectivity and visual perception. pmid: 24050403 We thank J. Csicsvari, T. Freund, S. Hippenmeyer, T. Klausberger,
Nature 488, 379383 (2012). doi: 10.1038/nature11312; 120. M. J. Janssen, E. Leiva-Salcedo, A. Buonanno, Neuregulin J. Lisman, and I. Soltesz for critically reading the manuscript;
pmid: 22878719 directly decreases voltage-gated sodium current in A. Solymosi for text editing; and all colleagues at IST Austria for
103. S. J. Kuhlman et al., A disinhibitory microcircuit hippocampal ErbB4-expressing interneurons. J. Neurosci. 32, generating a stimulating scientific environment. This work was supported
initiates critical-period plasticity in the visual cortex. 1388913895 (2012). doi: 10.1523/JNEUROSCI.1420-12.2012; by the Fond zur Frderung der Wissenschaftlichen Forschung (grant
Nature 501, 543546 (2013). doi: 10.1038/nature12485; pmid: 23035098 P 24909-B24 to P.J.) and the European Union (European Research
pmid: 23975100 121. K. X. Li et al., Neuregulin 1 regulates excitability of fast- Council Advanced Grant 268548 to P.J.). We apologize for the fact that,
104. J. J. Letzkus et al., A disinhibitory microcircuit for associative spiking neurons through Kv1.1 and acts in epilepsy. Nat. owing to space constraints, not all relevant papers could be cited.
fear learning in the auditory cortex. Nature 480, 331335 Neurosci. 15, 267273 (2012). doi: 10.1038/nn.3006;
(2011). doi: 10.1038/nature10674; pmid: 22158104 pmid: 22158511 10.1126/science.1255263

SCIENCE sciencemag.org 1 AUGUST 2014 VOL 345 ISSUE 6196 1255263-13

You might also like