Activity-dependent tuning of intrinsic excitability in mouse and human neurogliaform cells

Ramesh Chittajallu, Kurt Auville, Vivek Mahadevan, Mandy Lai, Steven Hunt, Daniela Calvigioni, Kenneth A Pelkey, Kareem A Zaghloul, Chris J McBain, Ramesh Chittajallu, Kurt Auville, Vivek Mahadevan, Mandy Lai, Steven Hunt, Daniela Calvigioni, Kenneth A Pelkey, Kareem A Zaghloul, Chris J McBain

Abstract

The ability to modulate the efficacy of synaptic communication between neurons constitutes an essential property critical for normal brain function. Animal models have proved invaluable in revealing a wealth of diverse cellular mechanisms underlying varied plasticity modes. However, to what extent these processes are mirrored in humans is largely uncharted thus questioning their relevance in human circuit function. In this study, we focus on neurogliaform cells, that possess specialized physiological features enabling them to impart a widespread inhibitory influence on neural activity. We demonstrate that this prominent neuronal subtype, embedded in both mouse and human neural circuits, undergo remarkably similar activity-dependent modulation manifesting as epochs of enhanced intrinsic excitability. In principle, these evolutionary conserved plasticity routes likely tune the extent of neurogliaform cell mediated inhibition thus constituting canonical circuit mechanisms underlying human cognitive processing and behavior.

Trial registration: ClinicalTrials.gov NCT01273129.

Keywords: human; interneuron; intrinsic excitablity; mouse; neurogliaform cell; neuroscience; plasticity.

Conflict of interest statement

RC, KA, VM, ML, SH, DC, KP, KZ, CM No competing interests declared

Figures

Figure 1.. Barrage firing (BF) is evolutionary…
Figure 1.. Barrage firing (BF) is evolutionary conserved in human NGFCs.
(A,C) Biocytin reconstruction of mouse (dendrites in black; axon in gray) and human NGFCs (dendrites in black; axon in orange) located in superficial regions of hippocampus (stratum lacunosum moleculare; SLM) and cortex (Layer I). (B,D) Single example trace showing the induction of barrage firing (BF) in mouse and human NGFCs upon delivery of action potentials via 1 ms depolarizing steps (typically 1000–1200 pA @ 30 Hz and 10 action potentials every second). (E,F) Box plots of pooled data depicting the minimum number of action potentials required for emergence of BF and its duration under differing frequencies of inducing action potentials (for the box plots, left to right, the n values are 68, 15, 12, 12 and 7, respectively) (G,H) Single example waveforms and corresponding phase plots of spikelets nested within full spikes (the latter defined as those overshooting 0 mV) observed during BF episodes in a single mouse and human NGFC. Black and orange data represent that derived from mouse and human NGFCs, respectively.
Figure 1—figure supplement 1.. Privileged induction of…
Figure 1—figure supplement 1.. Privileged induction of barrage firing in mouse and human NGFCs.
(A, B) Confocal mages of biocytin filled cortico-hippocampal NGFCs and non-NGFCs in stratum lacunosum-moleculare (SLM) and Layer I. Insets show firing patterns of non-NGFCs to suprathreshold current injections. Note: reconstructed drawings and corresponding firing patterns of NGFCs are depicted in main Figure 1A,B. (C) Histogram showing the prevalence in the ability to induce BF in mouse (black; n = 124) and human (orange; n = 25) NGFCs. (D) Histogram showing the inability to induce BF in mouse (black; n = 26) and human (orange; n = 33) non-NGFCs.
Figure 1—figure supplement 2.. Induction of multiple…
Figure 1—figure supplement 2.. Induction of multiple barrage firing episodes in individual mouse and human NGFCs.
(A) Single trace example showing the inability to elicit prolonged BF episodes in single NGFC following immediate continuation of the induction protocol (30 × 10 Hz induction; see main Figure 1 for details of induction). (B) Single trace showing repeated BF episodes can by induced in a given NGFC following a determined elapsed period which in this example is 4.4 min. (C) Individual data plot showing the relationship between the duration of the inter-induction period and failure/success of eliciting a second BF episode (n = 61). (D) Line plots showing duration of BF for multiple subsequent BF episodes (BF1 thru BF4) in any given mouse or human NGFC. Plotted as combined mean for BF1,BF2 and BF3,BF4 (n = 16 and 6 for mouse and human NGFCs, respectively). (E) Corresponding box plots of the % change in the duration of BF elicited in any given NGFC induced by subsequent BF induction protocols (four induction protocols in each NGFC were delivered with inter induction interval greater than 1 min). Data are expressed as a percentage change in these parameters for the mean of BF3, BF4 vs. the mean of BF1, BF2 in each NGFC tested (n = 16 and 6 for mouse and human NGFCs, respectively). Paired t-test p values = 0.0077 and 0.162 for mouse and human NGFCs, respectively).
Figure 2.. Short-term potentiation of somatic-depolarization driven…
Figure 2.. Short-term potentiation of somatic-depolarization driven excitability (STP-SE) in mouse and human NGFCs after cessation of BF.
(A) Single example voltage traces in response to increasing square wave somatic depolarizing current injections before and after induction of BF. Values under each pair of traces (Pre-BF and Post-BF) indicate the size of the 1 s duration current injection. (B) Pooled data input/output curves of action potential number in response to increasing amplitudes of depolarizing current injections (50 pA – 220 pA; 25 pA increments; n = 12 and 3 for mouse and human NGFCs, respectively). Mouse data are comprised of 8 hippocampal and four cortical NGFCs (C) Line plots of AP output before and after BF (calculated as the total sum of APs in response to all current injection steps) in individual mouse (n = 12; paired t-test; p=0.0000083) and human NGFCs (n = 3; paired t-test; p=0.012) tested. (D) Single example current traces from NGFCs in response to depolarizing voltage steps (magnitude of steps from a holding potential of −70 mV are indicated under each pair of traces) showing action current output before and after BF. (E) Pooled data showing input/output curves depicting number of action currents in response to depolarizing voltage steps (5 mV to 30 mV; 5 mV increments) before and after BF induction (n = 5 and 3 for mouse and human NGFCs, respectively). Mouse data are comprised of 3 hippocampal and two cortical NGFCs (F) Single example time course of action current output in response to 25 mV depolarizing voltage steps (sweep interval 10 s) under baseline conditions and after BF induction. (G) Box plots depicting the duration of increased action current output to 25 mV depolarizing voltage steps after a single BF episode (post-BF1) and after 4 BF episodes (post BF-4) in mouse (n = 11 and 4 for mouse and human NGFCs, respectively; paired t-test, p=0.007). Mouse data are comprised of 8 hippocampal and three cortical NGFCs.
Figure 3.. STP-SE in mouse and human…
Figure 3.. STP-SE in mouse and human NGFCs occurs in the absence of BF.
(A–C) Single example time course of the conditioning protocol employed to elicit STP-SE in mouse and human NGFCs. Action potential output was elicited to a test rheobase depolarizing current injection (300 ms; typically 100 pA - 140 pA; baseline) in the absence (baseline) and during conditioning action potentials (induction; 10 depolarizing steps of 2 ms width; 1000 pA - 1200 pA). Conditioning action potentials were terminated upon visualization of a marked loss in 1st spike latency (doted vertical lines) invariably accompanied by increased action potential output in response to the test depolarizing current injection (post-induction). (D) Pooled time course of action potential output (measured as number of action potentials in response to the test rheobase current injection) during the entire STP-SE induction protocol. Vertical dotted line indicates the commencement of induction with its termination occurring at varying times for each tested cell. Time course plots are from mouse (black) and human (orange) NGFCs displaying STP-SE (n = 9 and 4 for mouse hippocampal and human cortical NGFCs, respectively). For NGFCs (n = 7 and 2 for mouse hippocampal and human cortical NGFCs, respectively) and non-NGFCs (n = 4 and 3 for mouse and human non-NGFCs, respectively) that did not show STP-SE data were pooled between species (dashed black and orange line). (E) Box plots comparing number of conditioning action potentials required for BF (note these data are replotted from Figure 1E; n = 68 and 12, for mouse hippocampal and human cortical NGFCs, respectively) and STP-SE (n = 9 and 4, for mouse hippocampal and human NGFCs, respectively). Paired t-test p-values were 0.0004 and 0.031 for mouse and human, respectively. (F,G) Phase plots of action potential output to the test depolarizing test pulse during baseline (solid lines) and after induction of STP-SE (dotted lines) in single example mouse and human NGFCs. (H,I) Line plots demonstrating hyperpolarization of action potential threshold and reduction of spike latency following induction of STP-SE in mouse hippocampal and human NGFCs. Paired t-test p-values for comparisons of mouse AP threshold and latency (baseline vs. STP-SE) were 0.0000028 and 0.000025, respectively (n = 9). Paired t-test p-values for comparisons of human AP threshold and latency (baseline vs. STP-SE) were 0.000085 and 0.003, respectively (n = 4).
Figure 3—figure supplement 1.. Inability to induce…
Figure 3—figure supplement 1.. Inability to induce STP-SE in a subpopulation of NGFCs and all non-NGFCs tested.
(A-D) Time course traces demonstrating the inability of a single example mouse NGFC and non-NGFC to express STP-SE. Traces in B and D are magnified for clarity at the time points in the induction protocol as indicated.
Figure 4.. Pharmacological block of a subthreshold…
Figure 4.. Pharmacological block of a subthreshold transient K+-conductance mimics the AP latency and threshold modulation seen in STP-SE.
(A) Single example voltage trace to a depolarizing current injection (120 pA) in the presence of 1 μm TTX and 200 μm NiCl2 to block voltage gated Na+ and low-threshold activated Ca2+ channels, respectively. (B) Representative voltage-gated outward K+-currents activated in response to a series of predominantly subthreshold depolarizing voltage steps. Transient (open circle) and sustained (closed circle) components are labeled (C) Pooled data of the current-voltage curve of outward peak and sustained K+-currents with gray shaded area depicting the mean action potential threshold measured in NGFCs (n = 6). (D) Box plot data depicting the pharmacological sensitivity of the subthreshold outward conductance to a subthreshold voltage step (−30 mV) by 100–200 nM DTX, 100 μm (low 4-AP) and 1–3 mM 4-AP (high 4-AP; n = 5, 7 and 11, respectively). p-values for DTX and high 4-AP datasets (Wilcoxin signed rank test) and paired t-test for low 4-AP were 0.44, 0.00098 and 0.02, respectively. (E) Single trace example of the pharmacological isolation of high 4-AP sensitive subthreshold outward K+-current elicited in NGFCs characterized by a transient, inactivating time course. (F,G) Single trace examples showing the effect of DTX and high 4-AP on action potential output upon rheobase current step injections (120 pA). (H) Single trace examples comparing the effect of action potential output to a rheobase current injection performed with K+-gluconate intracellular solution without (control; left trace) or with added 4-AP (intra-4-AP; right green trace; 3 mM). (I) Pooled data showing a comparison of the effects of DTX (n = 10), low 4-AP (n = 9), high 4-AP (n = 12), intra-4-AP (3 mM; n = 7) and STP-SE (n = 9) on action potential latency and threshold in relation to that seen with STP-SE (n = 9).
Figure 5.. Functional availability of I KA…
Figure 5.. Functional availability of IKA modulates intrinsic excitability and induction requirements of STP-SE.
(A) Single trace examples showing the effect of DTX (top traces), high 4-AP (middle traces) and with K+-gluconate intracellular solution without (control; bottom left trace) vs. 4-AP (intra-4-AP; bottom right green trace; 3 mM) on action potential output upon 1.5 x threshold current step injections (B) Pooled data showing the effects of STP-SE (n = 8), DTX (100–200 nM; n = 8), low 4-AP (100 μm; n = 9), high 4-AP (1 mM; n = 7) on intrinsic excitability measured the total number of action potentials to a series of depolarizing current injections. p-values from unpaired t-tests comparing increase in intrinsic excitability between STP-SE vs. DTX, STP-SE vs. low-4AP or STP-SE vs. high-4AP = 0.011, 0.045 and 0.38, respectively. (C) Pooled data comparing intrinsic excitability of NGFCs to series of depolarizing current injections with control intracellular K+-gluconate (n = 7; black trace) or K+-gluconate with added 4-AP (interleaved experiments; intra 4-AP 3 mM; n = 7; green trace). (D) Effect of high 4-AP (1 mM) on voltage response of NGFC to a subthreshold current injection step (120 pA). (E) Single trace examples illustrating recovery of subthreshold IKA current following successive depolarizing voltage steps of the same amplitude (typically 120 pA). (F) Corresponding action potential output of NGFCs to pairs of temporally separated threshold current injections (200 ms) shown previously in (E) to result in a range of IKA inactivation. (G) Pooled data showing comparison between the level of IKA inactivation induced measured as ratio of peak amplitude of IKA (n = 6) and ratio of latency (n = 6) of the first action potential between the two successive depolarizing voltage/current steps, respectively. (H) Pooled data showing relationship between the level of IKA inactivation induced (n = 6) and increase in action potential output (n = 6), the latter measured as percentage increase in number of action potentials of between the two successive depolarizing current steps. (I) Pooled time course of action potential output (measured as number of action potentials in response to the test rheobase current injection as percentage of baseline) during the entire STP-SE induction protocol in NGFCs with either control intracellular K+-gluconate (n = 7; black plot) or K+-gluconate with added 4-AP (interleaved experiments; intra 4-AP 3 mM; n = 7; green plot). Vertical dotted line indicates the commencement of induction with its termination occurring at varying times for each tested cell. (J) Box plot showing number of conditioning APs required for expression of STP-SE with control or intra-4AP K+-gluconate solution (interleaved experiments; n = 5 for both conditions; unpaired t-test p value = 0.00005).
Figure 6.. I KA in NGFCs is…
Figure 6.. IKA in NGFCs is predominantly mediated by Kv4-subunit containing channels.
(A) Uniform Manifold Approximation and Projection (UMAP) plot of RNAseq data publicly available from the Allen Brain Institute (see Materials and methods for details) highlighting clusters corresponding to CGE-NGFCs (green; 3626 cells) and hippocampal pyramidal cells (blue; 2530 cells). (B) Violin plots depicting comparison of mRNA levels in CGE-NGFCs that encode for subunits of IKA channels. Median values are depicted above each plot. (C) Top, middle; Voltage steps and corresponding single current traces illustrating the protocol employed to isolate IKA and IKDR. Bottom; Single traces illustrating the effects of TEA (2–10 mM) and AmmTx3 (500 nM) on the isolated IKA and IKDR. (D) Box plots depicting inhibition of IKA and IKDR by TEA (n = 8) and AmmTx3 (n = 4). (E) Scatterplot of median expression levels of Kcnd2 vs. Kcnd3 mRNA in CGE-NGFCs and hippocampal pyramidal cells. (F) Representative confocal images of Kcnd2 and Kcnd3 transcript expression in CA1 pyramidal cells revealed by RNAscope (see Materials and methods for details; scale bar = 100 μm) (G) Representative confocal image of the distribution of putative NGFCs (via expression of Ndnf mRNA transcripts) in CA1 SLM (scale bar = 100 μm). (H,I) Representative confocal images of mRNA expression of Kcnd2 and Kcnd3 in putative NGFCs (Ndnf-expressing) in CA1 SLM. (Scale bars = 50 μm for top panels; 20 μm for bottom panels) (J) Split-violin plots depicting comparison in the expression levels of mRNA transcripts encoding known auxiliary subunits of Kv4-containing channels (KChIPs and DPLPs) in CGE-NGFCs (green) vs. hippocampal pyramidal cells (blue).
Figure 6—figure supplement 1.. Delineation of mouse…
Figure 6—figure supplement 1.. Delineation of mouse cortico-hippocampal cell clusters obtained from publicly available Allen Brain Institute scRNAseq data.
(A) Uniform Manifold Approximation and Projection (UMAP) dimension reductionality of ~76,000 single-cell transcriptomes, depicting the clustering of broad cell types. (B) Clustering of CGE and MGE-derived GABAergic interneuron subtypes. Cell clusters were colored and annotated post hoc based on their transcriptional profile identities (see Materials and methods). (C) Violin plots showing the distribution of expression levels of established cell type-enriched marker genes across GABAergic interneuron subtypes (see Materials and methods). Boxed regions illustrate the CGE-NGFCs (3626 cells) and hippocampal pyramidal cell (2530 cells) clusters delineated via the combined expression of the specific genes highlighted (see Materials and methods).
Figure 6—figure supplement 2.. Comparison of subthreshold…
Figure 6—figure supplement 2.. Comparison of subthreshold Kv1 and Kv4 subunit gene expression between mouse and human NGFCs and other IN subtypes.
(A,B) mRNA expression levels of Kcna1, Kcnd2 and Kcnd3 in neural cell types. Data are from publicly available resources from the Allen Brain Institute and the McCarroll Laboratory, Harvard Medical School. Green shaded box in (A) denotes the cluster corresponding to human cortical NGFCs. ID2+ cells in (B) corresponds to a subpopulation of human cortical NGFCs. Panel (A), top: 2019, 2020 Allen Institute for Brain Science. Mouse Transcriptomic Cell Types Database – Whole cortex and hippocampus – smart-seq with 10X-smart-seq taxonomy. Available from http://celltypes.brain-map.org/rnaseq/mouse_ctx-hip_smart-seq. Panel (B) is used with permission from Drs. McCarroll and Krienen (http://interneuron.mccarrolllab.org/).
Figure 7.. STP-SE manifests as an increase…
Figure 7.. STP-SE manifests as an increase in afferent recruitment of NGFCs via enhanced E-S coupling.
(A) Single traces showing the effect of AmmTx3 (500 nM) on EPSC (top traces) and corresponding EPSP (bottom trace) in a hippocampal NGFC in response to electrical stimulation of SLM afferent fibers (2 × 30 Hz). (B) Box plots of EPSC amplitudes (EPSC1 vs. EPSC2) under baseline conditions and after AmmTx3 treatment (n = 7; p values for EPSC1 and EPSC2 comparisons are 0.11 and 0.27, respectively). (C) Box plots of paired pulse ratio (EPSC2 peak amplitude/EPSC1 peak amplitude) under baseline conditions and after AmmTx3 treatment (n = 7; p value = 0.36). (D) Box plots of EPSP summation measured as absolute Vm at peak of EPSP1 and EPSP2 under baseline conditions and after AmmTx3 treatment (n = 7;=7; p values for EPSP1 and EPSP2 comparisons are 0.005 and 0.016, respectively)). (E–G) Single trace examples depicting a time course of E-S coupling (SLM afferent stimulation delivered at 5 × 30 Hz) during baseline conditions and after STP-SE induction. (H) Line plot depicting increased E-S coupling measured as spike probability during 10 sweeps of baseline vs. 10 sweeps immediately after STP-SE induction (n = 9; p value = 0.002). (I) Box plot illustrating duration of enhanced E-S coupling following STP-SE induction (n = 9).
Figure 7—figure supplement 1.. E-S coupling potentiation…
Figure 7—figure supplement 1.. E-S coupling potentiation not attributed to modulation of spontaneous EPSC parameters.
(A) Traces showing spontaneous AMPA-receptor-mediated EPSCs (top) under baseline conditions (gray) and immediately after cessation of a 4th BF episode (post-BF4; black) elicited in a given NGFC. Bottom trace shows overlaid ensemble average of AMPA-receptor-mediated EPSCs during baseline and after 4th BF episode. (B,C,D) Box plots of AMPA-receptor-mediated sEPSC amplitude, frequency and decay time constant measured under baseline (n = 6) and immediately following cessation of BF, BF2, BF3 and BF4 (n = 6) episodes. Paired t-test p-values for comparison of sEPSC amplitude, frequency and tau decay between post-BF1 vs. post-BF4 were 0.02, 0.64 and 0.95, respectively.
Author response image 1.. No change in…
Author response image 1.. No change in absolute peak of action potentials before and during STP-SE.
Black and orange data are from mouse (n = 9) and human(n = 4) NGFCs, respectively.
Author response image 2.. Single example illustrating…
Author response image 2.. Single example illustrating a reversal in the modulation of AP output following STP-SE induction.
Author response image 3.. Changes in electrotonic…
Author response image 3.. Changes in electrotonic distance of AIS is not a prerequisite for heightened intrinsic excitability observed during STP-SE.
Values above 2nd derivative plots denote numbers of NGFCS displaying the changes highlighted. Black and orange data are from mouse (n = 9) and human(n = 4) NGFCs, respectively.

References

    1. Abs E, Poorthuis RB, Apelblat D, Muhammad K, Pardi MB, Enke L, Kushinsky D, Pu D-L, Eizinger MF, Conzelmann K-K, Spiegel I, Letzkus JJ. Learning-Related plasticity in Dendrite-Targeting layer 1 interneurons. Neuron. 2018;100:684–699. doi: 10.1016/j.neuron.2018.09.001.
    1. Alfaro-Ruíz R, Aguado C, Martín-Belmonte A, Moreno-Martínez AE, Luján R. Expression, cellular and subcellular localisation of Kv4.2 and Kv4.3 Channels in the Rodent Hippocampus. International Journal of Molecular Sciences. 2019;20:246. doi: 10.3390/ijms20020246.
    1. Armstrong C, Szabadics J, Tamás G, Soltesz I. Neurogliaform cells in the molecular layer of the dentate gyrus as feed-forward γ-aminobutyric acidergic modulators of entorhinal-hippocampal interplay. The Journal of Comparative Neurology. 2011;519:1476–1491. doi: 10.1002/cne.22577.
    1. Beaulieu-Laroche L, Toloza EHS, van der Goes MS, Lafourcade M, Barnagian D, Williams ZM, Eskandar EN, Frosch MP, Cash SS, Harnett MT. Enhanced dendritic compartmentalization in human cortical neurons. Cell. 2018;175:643–651. doi: 10.1016/j.cell.2018.08.045.
    1. Bekkers JM, Delaney AJ. Modulation of excitability by alpha-dendrotoxin-sensitive potassium channels in neocortical pyramidal neurons. The Journal of Neuroscience. 2001;21:6553–6560. doi: 10.1523/JNEUROSCI.21-17-06553.2001.
    1. Bezaire MJ, Soltesz I. Quantitative assessment of CA1 local circuits: knowledge base for interneuron-pyramidal cell connectivity. Hippocampus. 2013;23:751–785. doi: 10.1002/hipo.22141.
    1. Bliss TVP, Collingridge GL. Persistent memories of long-term potentiation and the N -methyl-d-aspartate receptor. Brain and Neuroscience Advances. 2019;3:239821281984821. doi: 10.1177/2398212819848213.
    1. Boldog E, Bakken TE, Hodge RD, Novotny M, Aevermann BD, Baka J, Bordé S, Close JL, Diez-Fuertes F, Ding SL, Faragó N, Kocsis ÁK, Kovács B, Maltzer Z, McCorrison JM, Miller JA, Molnár G, Oláh G, Ozsvár A, Rózsa M, Shehata SI, Smith KA, Sunkin SM, Tran DN, Venepally P, Wall A, Puskás LG, Barzó P, Steemers FJ, Schork NJ, Scheuermann RH, Lasken RS, Lein ES, Tamás G. Transcriptomic and morphophysiological evidence for a specialized human cortical GABAergic cell type. Nature Neuroscience. 2018;21:1185–1195. doi: 10.1038/s41593-018-0205-2.
    1. Bourdeau ML, Morin F, Laurent CE, Azzi M, Lacaille JC. Kv4.3-mediated A-type K+ currents underlie rhythmic activity in hippocampal interneurons. Journal of Neuroscience. 2007;27:1942–1953. doi: 10.1523/JNEUROSCI.3208-06.2007.
    1. Bourdeau ML, Laplante I, Laurent CE, Lacaille JC. KChIP1 modulation of Kv4.3-mediated A-type K(+) currents and repetitive firing in hippocampal interneurons. Neuroscience. 2011;176:173–187. doi: 10.1016/j.neuroscience.2010.11.051.
    1. Butler A, Hoffman P, Smibert P, Papalexi E, Satija R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nature Biotechnology. 2018;36:411–420. doi: 10.1038/nbt.4096.
    1. Butt SJ, Stacey JA, Teramoto Y, Vagnoni C. A role for GABAergic interneuron diversity in circuit development and plasticity of the neonatal cerebral cortex. Current Opinion in Neurobiology. 2017;43:149–155. doi: 10.1016/j.conb.2017.03.011.
    1. Campanac E, Gasselin C, Baude A, Rama S, Ankri N, Debanne D. Enhanced intrinsic excitability in basket cells maintains excitatory-inhibitory balance in hippocampal circuits. Neuron. 2013;77:712–722. doi: 10.1016/j.neuron.2012.12.020.
    1. Capogna M. Neurogliaform cells and other interneurons of stratum lacunosum-moleculare gate entorhinal-hippocampal dialogue. The Journal of Physiology. 2011;589:1875–1883. doi: 10.1113/jphysiol.2010.201004.
    1. Carrasquillo Y, Burkhalter A, Nerbonne JM. A-type K+ channels encoded by Kv4.2, Kv4.3 and Kv1.4 differentially regulate intrinsic excitability of cortical pyramidal neurons. The Journal of Physiology. 2012;590:3877–3890. doi: 10.1113/jphysiol.2012.229013.
    1. Cembrowski MS, Bachman JL, Wang L, Sugino K, Shields BC, Spruston N. Spatial Gene-Expression gradients underlie prominent heterogeneity of CA1 pyramidal neurons. Neuron. 2016a;89:351–368. doi: 10.1016/j.neuron.2015.12.013.
    1. Cembrowski MS, Wang L, Sugino K, Shields BC, Spruston N. Hipposeq: a comprehensive RNA-seq database of gene expression in hippocampal principal neurons. eLife. 2016b;5:e14997. doi: 10.7554/eLife.14997.
    1. Chittajallu R, Pelkey KA, McBain CJ. Neurogliaform cells dynamically regulate somatosensory integration via synapse-specific modulation. Nature Neuroscience. 2013;16:13–15. doi: 10.1038/nn.3284.
    1. Chittajallu R, Wester JC, Craig MT, Barksdale E, Yuan XQ, Akgül G, Fang C, Collins D, Hunt S, Pelkey KA, McBain CJ. Afferent specific role of NMDA receptors for the circuit integration of hippocampal neurogliaform cells. Nature Communications. 2017;8:152. doi: 10.1038/s41467-017-00218-y.
    1. Chittajallu R. Mouse-CGE-NGFCs. 204db8eGitHub. 2020
    1. Chu Z, Galarreta M, Hestrin S. Synaptic interactions of late-spiking neocortical neurons in layer 1. The Journal of Neuroscience. 2003;23:96–102. doi: 10.1523/JNEUROSCI.23-01-00096.2003.
    1. Cummings KA, Clem RL. Prefrontal somatostatin interneurons encode fear memory. Nature Neuroscience. 2020;23:61–74. doi: 10.1038/s41593-019-0552-7.
    1. Debanne D, Inglebert Y, Russier M. Plasticity of intrinsic neuronal excitability. Current Opinion in Neurobiology. 2019;54:73–82. doi: 10.1016/j.conb.2018.09.001.
    1. Deemyad T, Lüthi J, Spruston N. Astrocytes integrate and drive action potential firing in inhibitory subnetworks. Nature Communications. 2018;9:4336. doi: 10.1038/s41467-018-06338-3.
    1. Dehorter N, Ciceri G, Bartolini G, Lim L, del Pino I, Marín O. Tuning of fast-spiking interneuron properties by an activity-dependent transcriptional switch. Science. 2015;349:1216–1220. doi: 10.1126/science.aab3415.
    1. Dimidschstein J, Chen Q, Tremblay R, Rogers SL, Saldi G-A, Guo L, Xu Q, Liu R, Lu C, Chu J, Grimley JS, Krostag A-R, Kaykas A, Avery MC, Rashid MS, Baek M, Jacob AL, Smith GB, Wilson DE, Kosche G, Kruglikov I, Rusielewicz T, Kotak VC, Mowery TM, Anderson SA, Callaway EM, Dasen JS, Fitzpatrick D, Fossati V, Long MA, Noggle S, Reynolds JH, Sanes DH, Rudy B, Feng G, Fishell G. A viral strategy for targeting and manipulating interneurons across vertebrate species. Nature Neuroscience. 2016;19:1743–1749. doi: 10.1038/nn.4430.
    1. Elgueta C, Köhler J, Bartos M. Persistent discharges in Dentate Gyrus perisoma-inhibiting interneurons require hyperpolarization-activated cyclic nucleotide-gated channel activation. Journal of Neuroscience. 2015;35:4131–4139. doi: 10.1523/JNEUROSCI.3671-14.2015.
    1. Eyal G, Verhoog MB, Testa-Silva G, Deitcher Y, Lodder JC, Benavides-Piccione R, Morales J, DeFelipe J, de Kock CPJ, Mansvelder HD, Segev I. Unique membrane properties and enhanced signal processing in human neocortical neurons. eLife. 2016;5:e16553. doi: 10.7554/eLife.16553.
    1. Gainey MA, Aman JW, Feldman DE. Rapid disinhibition by adjustment of PV intrinsic excitability during whisker map plasticity in mouse S1. The Journal of Neuroscience. 2018;38:4749–4761. doi: 10.1523/JNEUROSCI.3628-17.2018.
    1. Goldberg JH, Tamas G, Yuste R. Ca2+ imaging of mouse neocortical interneurone dendrites: ia-type K+ channels control action potential backpropagation. The Journal of Physiology. 2003;551:49–65. doi: 10.1113/jphysiol.2003.042580.
    1. Hebb DO. The Organization of Behavior: A Neurophysiological Theory. John Wiley and Sons; 1949.
    1. Hodge RD, Bakken TE, Miller JA, Smith KA, Barkan ER, Graybuck LT, Close JL, Long B, Johansen N, Penn O, Yao Z, Eggermont J, Höllt T, Levi BP, Shehata SI, Aevermann B, Beller A, Bertagnolli D, Brouner K, Casper T, Cobbs C, Dalley R, Dee N, Ding SL, Ellenbogen RG, Fong O, Garren E, Goldy J, Gwinn RP, Hirschstein D, Keene CD, Keshk M, Ko AL, Lathia K, Mahfouz A, Maltzer Z, McGraw M, Nguyen TN, Nyhus J, Ojemann JG, Oldre A, Parry S, Reynolds S, Rimorin C, Shapovalova NV, Somasundaram S, Szafer A, Thomsen ER, Tieu M, Quon G, Scheuermann RH, Yuste R, Sunkin SM, Lelieveldt B, Feng D, Ng L, Bernard A, Hawrylycz M, Phillips JW, Tasic B, Zeng H, Jones AR, Koch C, Lein ES. Conserved cell types with divergent features in human versus mouse cortex. Nature. 2019;573:61–68. doi: 10.1038/s41586-019-1506-7.
    1. Imbrosci B, Neitz A, Mittmann T. Physiological properties of supragranular cortical inhibitory interneurons expressing retrograde persistent firing. Neural Plasticity. 2015;2015:1–12. doi: 10.1155/2015/608141.
    1. Jerng HH, Pfaffinger PJ. Modulatory mechanisms and multiple functions of somatodendritic A-type K+ channel auxiliary subunits. Frontiers in Cellular Neuroscience. 2014;8:82. doi: 10.3389/fncel.2014.00082.
    1. Jung SC, Hoffman DA. Biphasic somatic A-type K channel downregulation mediates intrinsic plasticity in hippocampal CA1 pyramidal neurons. PLOS ONE. 2009;4:e6549. doi: 10.1371/journal.pone.0006549.
    1. Keck T, Toyoizumi T, Chen L, Doiron B, Feldman DE, Fox K, Gerstner W, Haydon PG, Hübener M, Lee H-K, Lisman JE, Rose T, Sengpiel F, Stellwagen D, Stryker MP, Turrigiano GG, van Rossum MC. Integrating hebbian and homeostatic plasticity: the current state of the field and future research directions. Philosophical Transactions of the Royal Society B: Biological Sciences. 2017;372:20160158. doi: 10.1098/rstb.2016.0158.
    1. Kepecs A, Fishell G. Interneuron cell types are fit to function. Nature. 2014;505:318–326. doi: 10.1038/nature12983.
    1. Kim J, Hoffman DA. Potassium channels: newly found players in synaptic plasticity. The Neuroscientist. 2008;14:276–286. doi: 10.1177/1073858408315041.
    1. Krienen FM, Goldman M, Zhang Q, del Rosario R, Florio M, Machold R, McCarroll SA. Innovations in primate interneuron repertoire. bioRxiv. 2019 doi: 10.1101/709501.
    1. Krook-Magnuson E, Luu L, Lee SH, Varga C, Soltesz I. Ivy and neurogliaform interneurons are a major target of μ-opioid receptor modulation. Journal of Neuroscience. 2011;31:14861–14870. doi: 10.1523/JNEUROSCI.2269-11.2011.
    1. Kullmann DM, Moreau AW, Bakiri Y, Nicholson E. Plasticity of inhibition. Neuron. 2012;75:951–962. doi: 10.1016/j.neuron.2012.07.030.
    1. Lee S, Hjerling-Leffler J, Zagha E, Fishell G, Rudy B. The largest group of superficial neocortical GABAergic interneurons expresses ionotropic serotonin receptors. Journal of Neuroscience. 2010;30:16796–16808. doi: 10.1523/JNEUROSCI.1869-10.2010.
    1. Li K, Lu Y-M, Xu Z-H, Zhang J, Zhu J-M, Zhang J-M, Cao S-X, Chen X-J, Chen Z, Luo J-H, Duan S, Li X-M. Neuregulin 1 regulates excitability of fast-spiking neurons through Kv1.1 and acts in epilepsy. Nature Neuroscience. 2012;15:267–273. doi: 10.1038/nn.3006.
    1. Li G, Stewart R, Canepari M, Capogna M. Firing of hippocampal neurogliaform cells induces suppression of synaptic inhibition. Journal of Neuroscience. 2014;34:1280–1292. doi: 10.1523/JNEUROSCI.3046-13.2014.
    1. Lien CC, Martina M, Schultz JH, Ehmke H, Jonas P. Gating, modulation and subunit composition of voltage-gated K(+) channels in dendritic inhibitory interneurones of rat Hippocampus. The Journal of Physiology. 2002;538:405–419. doi: 10.1113/jphysiol.2001.013066.
    1. Lovett-Barron M, Kaifosh P, Kheirbek MA, Danielson N, Zaremba JD, Reardon TR, Turi GF, Hen R, Zemelman BV, Losonczy A. Dendritic inhibition in the Hippocampus supports fear learning. Science. 2014;343:857–863. doi: 10.1126/science.1247485.
    1. Maffie JK, Dvoretskova E, Bougis PE, Martin-Eauclaire MF, Rudy B. Dipeptidyl-peptidase-like-proteins confer high sensitivity to the scorpion toxin AmmTX3 to Kv4-mediated A-type K+ channels. The Journal of Physiology. 2013;591:2419–2427. doi: 10.1113/jphysiol.2012.248831.
    1. Malenka RC, Bear MF. LTP and LTD: an embarrassment of riches. Neuron. 2004;44:5–21. doi: 10.1016/j.neuron.2004.09.012.
    1. Mansvelder HD, Verhoog MB, Goriounova NA. Synaptic plasticity in human cortical circuits: cellular mechanisms of learning and memory in the human brain? Current Opinion in Neurobiology. 2019;54:186–193. doi: 10.1016/j.conb.2018.06.013.
    1. Marín O. Interneuron dysfunction in psychiatric disorders. Nature Reviews Neuroscience. 2012;13:107–120. doi: 10.1038/nrn3155.
    1. Mayer C, Hafemeister C, Bandler RC, Machold R, Batista Brito R, Jaglin X, Allaway K, Butler A, Fishell G, Satija R. Developmental diversification of cortical inhibitory interneurons. Nature. 2018;555:457–462. doi: 10.1038/nature25999.
    1. Mehta P, Kreeger L, Wylie DC, Pattadkal JJ, Lusignan T, Davis MJ, Turi GF, Li W-K, Whitmire MP, Chen Y, Kajs BL, Seidemann E, Priebe NJ, Losonczy A, Zemelman BV. Functional access to neuron subclasses in rodent and primate forebrain. Cell Reports. 2019;26:2818–2832. doi: 10.1016/j.celrep.2019.02.011.
    1. Menegola M, Misonou H, Vacher H, Trimmer JS. Dendritic A-type potassium channel subunit expression in CA1 hippocampal interneurons. Neuroscience. 2008;154:953–964. doi: 10.1016/j.neuroscience.2008.04.022.
    1. Mercier MS, Magloire V, Cornford JH, Kullmann DM. Long-term synaptic plasticity in hippocampal neurogliaform interneurons. bioRxiv. 2019 doi: 10.1101/531822.
    1. Mich JK, Graybuck LT, Hess EE, Mahoney JT, Kojima Y, Ding Y, Levi BP. Functional enhancer elements drive subclass-selective expression from mouse to primate neocortex. bioRxiv. 2020 doi: 10.1101/555318.
    1. Molnár G, Rózsa M, Baka J, Holderith N, Barzó P, Nusser Z, Tamás G. Human pyramidal to interneuron synapses are mediated by multi-vesicular release and multiple docked vesicles. eLife. 2016;5:e18167. doi: 10.7554/eLife.18167.
    1. Obermayer J, Heistek TS, Kerkhofs A, Goriounova NA, Kroon T, Baayen JC, Idema S, Testa-Silva G, Couey JJ, Mansvelder HD. Lateral inhibition by martinotti interneurons is facilitated by cholinergic inputs in human and mouse neocortex. Nature Communications. 2018;9:4101. doi: 10.1038/s41467-018-06628-w.
    1. Oláh S, Komlósi G, Szabadics J, Varga C, Tóth E, Barzó P, Tamás G. Output of neurogliaform cells to various neuron types in the human and rat cerebral cortex. Frontiers in Neural Circuits. 2007;1:4. doi: 10.3389/neuro.04.004.2007.
    1. Oláh VJ, Lukacsovich D, Winterer J, Arszovszki A, Lőrincz A, Nusser Z, Földy C, Szabadics J. Functional specification of CCK+ interneurons by alternative isoforms of Kv4.3 auxiliary subunits. eLife. 2020;9:e58515. doi: 10.7554/eLife.58515.
    1. Overstreet-Wadiche L, McBain CJ. Neurogliaform cells in cortical circuits. Nature Reviews Neuroscience. 2015;16:458–468. doi: 10.1038/nrn3969.
    1. Pelkey KA, Chittajallu R, Craig MT, Tricoire L, Wester JC, McBain CJ. Hippocampal GABAergic inhibitory interneurons. Physiological Reviews. 2017;97:1619–1747. doi: 10.1152/physrev.00007.2017.
    1. Poorthuis RB, Muhammad K, Wang M, Verhoog MB, Junek S, Wrana A, Mansvelder HD, Letzkus JJ. Rapid neuromodulation of layer 1 interneurons in human neocortex. Cell Reports. 2018;23:951–958. doi: 10.1016/j.celrep.2018.03.111.
    1. Povysheva NV, Zaitsev AV, Kröner S, Krimer OA, Rotaru DC, Gonzalez-Burgos G, Lewis DA, Krimer LS. Electrophysiological differences between neurogliaform cells from monkey and rat prefrontal cortex. Journal of Neurophysiology. 2007;97:1030–1039. doi: 10.1152/jn.00794.2006.
    1. Povysheva NV, Zaitsev AV, Rotaru DC, Gonzalez-Burgos G, Lewis DA, Krimer LS. Parvalbumin-positive basket interneurons in monkey and rat prefrontal cortex. Journal of Neurophysiology. 2008;100:2348–2360. doi: 10.1152/jn.90396.2008.
    1. Pozo K, Goda Y. Unraveling mechanisms of homeostatic synaptic plasticity. Neuron. 2010;66:337–351. doi: 10.1016/j.neuron.2010.04.028.
    1. Price CJ, Cauli B, Kovacs ER, Kulik A, Lambolez B, Shigemoto R, Capogna M. Neurogliaform neurons form a novel inhibitory network in the hippocampal CA1 area. Journal of Neuroscience. 2005;25:6775–6786. doi: 10.1523/JNEUROSCI.1135-05.2005.
    1. Rhodes KJ, Carroll KI, Sung MA, Doliveira LC, Monaghan MM, Burke SL, Strassle BW, Buchwalder L, Menegola M, Cao J, An WF, Trimmer JS. KChIPs and Kv4 alpha subunits as integral components of A-type potassium channels in mammalian brain. Journal of Neuroscience. 2004;24:7903–7915. doi: 10.1523/JNEUROSCI.0776-04.2004.
    1. Rózsa M, Toth M, Olah G, Baka J, Barzo P, Tamas G. Rhythmic Persistent Firing of Neurogliaform Interneurons in the Human and Rodent Neocortex. Program No. 202.07. 2017 Neuroscience Meeting Planner. Washington DC: Society for Neuroscience; 2017.
    1. Rudy B. Diversity and ubiquity of K channels. Neuroscience. 1988;25:729–749. doi: 10.1016/0306-4522(88)90033-4.
    1. Rudy B, McBain CJ. Kv3 channels: voltage-gated K+ channels designed for high-frequency repetitive firing. Trends in Neurosciences. 2001;24:517–526. doi: 10.1016/S0166-2236(00)01892-0.
    1. Serôdio P, Rudy B. Differential expression of Kv4 K+ channel subunits mediating subthreshold transient K+ (A-type) currents in rat brain. Journal of Neurophysiology. 1998;79:1081–1091. doi: 10.1152/jn.1998.79.2.1081.
    1. Sheffield MEJ, Best TK, Mensh BD, Kath WL, Spruston N. Slow integration leads to persistent action potential firing in distal axons of coupled interneurons. Nature Neuroscience. 2011;14:200–207. doi: 10.1038/nn.2728.
    1. Sheffield ME, Edgerton GB, Heuermann RJ, Deemyad T, Mensh BD, Spruston N. Mechanisms of retroaxonal barrage firing in hippocampal interneurons. The Journal of Physiology. 2013;591:4793–4805. doi: 10.1113/jphysiol.2013.258418.
    1. Simon A, Oláh S, Molnár G, Szabadics J, Tamás G. Gap-junctional coupling between neurogliaform cells and various interneuron types in the neocortex. Journal of Neuroscience. 2005;25:6278–6285. doi: 10.1523/JNEUROSCI.1431-05.2005.
    1. Storm JF. Temporal integration by a slowly inactivating K+ current in hippocampal neurons. Nature. 1988;336:379–381. doi: 10.1038/336379a0.
    1. Stuart T, Butler A, Hoffman P, Hafemeister C, Papalexi E, Mauck WM, Hao Y, Stoeckius M, Smibert P, Satija R. Comprehensive integration of Single-Cell data. Cell. 2019;177:1888–1902. doi: 10.1016/j.cell.2019.05.031.
    1. Sun QQ. Experience-dependent intrinsic plasticity in interneurons of barrel cortex layer IV. Journal of Neurophysiology. 2009;102:2955–2973. doi: 10.1152/jn.00562.2009.
    1. Suzuki N, Tang CS, Bekkers JM. Persistent barrage firing in cortical interneurons can be induced in vivo and may be important for the suppression of epileptiform activity. Frontiers in Cellular Neuroscience. 2014;8:76. doi: 10.3389/fncel.2014.00076.
    1. Szabadics J, Varga C, Molnár G, Oláh S, Barzó P, Tamás G. Excitatory effect of GABAergic axo-axonic cells in cortical microcircuits. Science. 2006;311:233–235. doi: 10.1126/science.1121325.
    1. Szegedi V, Paizs M, Csakvari E, Molnar G, Barzo P, Tamas G, Lamsa K. Plasticity in single axon glutamatergic connection to GABAergic interneurons regulates complex events in the human neocortex. PLOS Biology. 2016;14:e2000237. doi: 10.1371/journal.pbio.2000237.
    1. Szegedi V, Paizs M, Baka J, Barzó P, Molnár G, Tamas G, Lamsa K. Robust perisomatic GABAergic self-innervation inhibits basket cells in the human and mouse supragranular neocortex. eLife. 2020;9:e51691. doi: 10.7554/eLife.51691.
    1. Takesian AE, Hensch TK. Balancing plasticity/stability across brain development. Progress in Brain Research. 2013;207:3–34. doi: 10.1016/B978-0-444-63327-9.00001-1.
    1. Tasic B, Menon V, Nguyen TN, Kim TK, Jarsky T, Yao Z, Levi B, Gray LT, Sorensen SA, Dolbeare T, Bertagnolli D, Goldy J, Shapovalova N, Parry S, Lee C, Smith K, Bernard A, Madisen L, Sunkin SM, Hawrylycz M, Koch C, Zeng H. Adult mouse cortical cell taxonomy revealed by single cell transcriptomics. Nature Neuroscience. 2016;19:335–346. doi: 10.1038/nn.4216.
    1. Tasic B, Yao Z, Graybuck LT, Smith KA, Nguyen TN, Bertagnolli D, Goldy J, Garren E, Economo MN, Viswanathan S, Penn O, Bakken T, Menon V, Miller J, Fong O, Hirokawa KE, Lathia K, Rimorin C, Tieu M, Larsen R, Casper T, Barkan E, Kroll M, Parry S, Shapovalova NV, Hirschstein D, Pendergraft J, Sullivan HA, Kim TK, Szafer A, Dee N, Groblewski P, Wickersham I, Cetin A, Harris JA, Levi BP, Sunkin SM, Madisen L, Daigle TL, Looger L, Bernard A, Phillips J, Lein E, Hawrylycz M, Svoboda K, Jones AR, Koch C, Zeng H. Shared and distinct transcriptomic cell types across neocortical Areas. Nature. 2018;563:72–78. doi: 10.1038/s41586-018-0654-5.
    1. Tien NW, Kerschensteiner D. Homeostatic plasticity in neural development. Neural Development. 2018;13:9. doi: 10.1186/s13064-018-0105-x.
    1. Tricoire L, Pelkey KA, Erkkila BE, Jeffries BW, Yuan X, McBain CJ. A blueprint for the spatiotemporal origins of mouse hippocampal interneuron diversity. Journal of Neuroscience. 2011;31:10948–10970. doi: 10.1523/JNEUROSCI.0323-11.2011.
    1. Trimmer JS. Subcellular localization of K+ channels in mammalian brain neurons: remarkable precision in the midst of extraordinary complexity. Neuron. 2015;85:238–256. doi: 10.1016/j.neuron.2014.12.042.
    1. Varga C, Tamas G, Barzo P, Olah S, Somogyi P. Molecular and electrophysiological characterization of GABAergic interneurons expressing the transcription factor COUP-TFII in the adult human temporal cortex. Cerebral Cortex. 2015;25:4430–4449. doi: 10.1093/cercor/bhv045.
    1. Vormstein-Schneider D, Lin J, Pelkey K, Chittajallu R, Guo B, Garcia MA, Sakopoulos S, Stevenson O, Schneider G, Zhang Q, Sharma J, Franken TP, Smith J, Vogel I, Sanchez V, Ibrahim LA, Burbridge T, Favuzzi E, Saldi GA, Xu Q, Guo L, Yuan X, Zaghloul KA, Sabri E, Goldberg EM, Devinsky O, Batista-Brito R, Reynolds J, Feng G, Fu Z, McBain CJ, Fishell GJ, Dimidschstein J. Viral manipulation of functionally distinct neurons from mice to humans. bioRxiv. 2020 doi: 10.1101/808170.
    1. Wang Y, Toledo-Rodriguez M, Gupta A, Wu C, Silberberg G, Luo J, Markram H. Anatomical, physiological and molecular properties of martinotti cells in the somatosensory cortex of the juvenile rat. The Journal of Physiology. 2004;561:65–90. doi: 10.1113/jphysiol.2004.073353.
    1. Wang B, Yin L, Zou X, Ye M, Liu Y, He T, Deng S, Jiang Y, Zheng R, Wang Y, Yang M, Lu H, Wu S, Shu Y. A subtype of inhibitory interneuron with intrinsic persistent activity in human and monkey neocortex. Cell Reports. 2015;10:1450–1458. doi: 10.1016/j.celrep.2015.02.018.
    1. Williams SB, Hablitz JJ. Differential modulation of repetitive firing and synchronous network activity in neocortical interneurons by inhibition of A-type K(+) channels and ih. Frontiers in Cellular Neuroscience. 2015;9:89. doi: 10.3389/fncel.2015.00089.
    1. Yoshida M, Hasselmo ME. Persistent firing supported by an intrinsic cellular mechanism in a component of the head direction system. Journal of Neuroscience. 2009;29:4945–4952. doi: 10.1523/JNEUROSCI.5154-08.2009.
    1. Zaitsev AV, Povysheva NV, Gonzalez-Burgos G, Rotaru D, Fish KN, Krimer LS, Lewis DA. Interneuron diversity in layers 2-3 of monkey prefrontal cortex. Cerebral Cortex. 2009;19:1597–1615. doi: 10.1093/cercor/bhn198.

Source: PubMed

3
Sottoscrivi