Main

We first investigated the cellular composition of the TRN, characterizing the expression of parvalbumin (PV), calbindin (CB) and somatostatin (SOM)—three markers that are useful in differentiating functionally distinct neural types in the neocortex and elsewhere13. Brain sections spanning the somatosensory sector of the TRN4,7,8 were prepared from SOM-Cre mice crossed with Cre-dependent tdTomato (tdT) reporters, then stained immunohistochemically for CB and PV (Fig. 1a).

Fig. 1: The somatosensory TRN is composed of neurochemically distinct neurons located in separate zones.
figure 1

a, Confocal images of a 40-μm section, centred on the somatosensory TRN, from a SOM-Cre × tdTomato mouse. Immunohistochemical (IHC) staining was performed for PV and CB. SOM-Cre cells were genetically labelled (SOM-Cre × tdT). The outlines of the TRN are based on PV labelling. Scale bar, 100 μm. b, Confocal images of an 18-μm section from a wild-type mouse, showing fluorescence in situ hybridization experiments for the visualization of Pvalb, Sst and Calb1 mRNA (encoding PV, SOM and CB). Scale bar, 100 μm. c, Left, proportion of cells expressing PV, SOM–tdT or CB in the somatosensory TRN (outlined by the bracket in the leftmost image of a). Middle, proportions of cells expressing combinations of markers (n = 1,075 cells, 3 sections, 3 mice). Right, proportions of each marker in three zones of the somatosensory TRN. A total of 187 cells were in the medial zone (medial 20%), 746 in the central zone (central 60%) and 142 were in the lateral zone (lateral 20%). The leftmost images of a and b show the zones marked by dashed lines. Fractional SOM– tdT cell densities were higher in the edge zones than in the central zone, whereas CB cell densities were higher in the central zone (all P < 0.001, χ2, Yates’ correction). d, Left and middle, as for c but for Pvalb, Sst and Calb1 mRNA (n = 593 cells, 6 sections, 6 mice). Right, 98 cells were in the medial zone, 412 in the central zone and 83 in the lateral zone. Again, Sst-expressing and Calb1-expressing cells were differentially distributed across zones (all P < 0.001, χ2, Yates’ correction). e, Confocal image from a P23 SOM-Cre mouse. AAV9-DIO–GFP was injected into the TRN at P14 to assess SOM-Cre expression. GFP cells (pseudocoloured cyan) were largely absent in the central zone (replicated in 12 mice). Data are mean ± s.e.m. Scale bar, 200 μm.

Two neuron types in the somatosensory TRN

Nearly all somatosensory TRN cells expressed PV1,14, whereas only subsets expressed SOM–tdT (around 64%) or CB (around 48%) (Fig. 1c, Extended Data Fig. 1). The distributions of SOM–tdT cells and CB cells across the somatosensory TRN were complementary. The highest densities of SOM–tdT cells were near the medial and lateral edges of the sector27, whereas CB cells were concentrated near the centre and were almost absent along the edges (Fig. 1a, Extended Data Fig. 1). Quantitative comparisons between the medial 20%, lateral 20% and central 60% of the somatosensory TRN confirmed that the proportions of cells expressing SOM–tdT were higher in the edge zones than in the central zone (P < 0.001, χ2 test), whereas the reverse was true for CB (P < 0.001) (Fig. 1c).

To further investigate the organization of neurons across the TRN, we used in situ hybridization to assay expression of mRNA encoding SOM (Sst), CB (Calb1) and PV (Pvalb). The highest densities of SOM cells were again found in the medial and lateral edge zones, CB cells were clustered centrally, and nearly all cells were PV-positive (Fig. 1b, d, Extended Data Fig. 2). Notably, the segregation between edge SOM cells and central CB cells was more salient in the mRNA assay, mainly due to decreased proportions of SOM cells in the central zone (only 15.3% of central cells expressed Sst mRNA, whereas 54.7% expressed SOM–tdT; Fig. 1a–d).

The near absence of SOM-encoding mRNA in the central zone suggests that most neurons located there may not actually express SOM protein in mature mice, and that expression of tdT in the central cells of SOM-Cre × tdT mice could result from genetic recombination early in development and persistent tdT production thereafter15. To test for SOM expression in mature mice we initially attempted immunohistochemistry; however, we were unable to find SOM antibodies that were adequate for the TRN (not shown). As an alternative, we assayed Cre expression in mature SOM-Cre mice by injecting an adeno-associated virus driving Cre-dependent GFP into the TRN. Cre expression in the somatosensory TRN of these mice was almost entirely restricted to the edge zones27, consistent with the pattern of Sst mRNA expression (Fig. 1b, d, e, Extended Data Fig. 3). Together, these results indicate that the somatosensory TRN is composed of neurochemically distinct cell types segregated into separate zones: a core central zone composed mostly of CB-expressing neurons, flanked by edge zones of SOM-expressing neurons.

Primary and higher-order TRN subcircuits

The primary ventral posterior (VP) and higher-order posterior medial (POM) thalamocortical nuclei transmit distinct information to different targets in the neocortex and send collaterals to the TRN; the latter leads to both open-loop and closed-loop thalamic inhibition4,5,6,8. Clarifying the organization of these circuits, including how primary and higher-order thalamocortical nuclei synapse with subtypes of TRN neurons16, is essential to understanding thalamic information processing7. To this end, we selectively expressed ChR2–eYFP in the VP or the POM, then characterized their inputs to the TRN (Fig. 2). Notably, their projections were found to segregate topographically, in close alignment with the observed patterns of TRN cell types. VP axons terminated in the CB-rich central zone of the somatosensory TRN, whereas POM axons terminated along the SOM-dense medial and lateral edges (Fig. 2a, b, Extended Data Fig. 4a, c).

Fig. 2: The VP, a primary thalamic nucleus, targets the central somatosensory TRN, whereas the POM, a higher-order nucleus, targets the TRN edges.
figure 2

a, Representative confocal images illustrating the distinct projections from the POM and the VP to the TRN. AAV2-ChR2–eYFP was injected into the POM (left) or the VP (right) and projections to the TRN were characterized. Scale bars: 500 μm (main images); 100 μm (expanded images). b, Group data showing average thalamic projections to the TRN. Viral anterograde tracing as in a. The POM targeted the edges of the somatosensory TRN (n = 9 mice), and the VP targeted the central zone (n = 8 mice). Boundaries between the central and edge zones are drawn at 20% and 80% of the medial–lateral distance across the TRN. Right, average SOM–tdT and CB immunohistochemical profiles (n = 5 mice) aligned to the same reference TRN as the anterograde projection maps. Scale bar, 200 μm. c, Synaptic responses to POM input from cells across the TRN, testing how synaptic strength relates to anterograde fluorescence and topographical location. Left, image of live TRN (outlined) with ChR2–eYFP projections from the POM. Circles show the locations of recorded cells. Right, EPSCs evoked in TRN cells by optical activation of POM axons (−84 mV). The colours match the cells in the left image and the colours in d. Scale bar, 100 μm. d, Normalized fluorescence (black line) and synaptic charge (coloured dots) for cells in c, as a function of their medial–lateral position in the TRN. e, Group relationship between the location of the TRN soma and the evoked synaptic response. Somas close to the TRN edges responded strongly to input from the POM and those in the centre did not (for e, f, n = 21 cells, 5 slices, 5 mice; each preparation has a unique symbol). f, Synaptic responses of TRN cells to POM input (normalized charge) correlated with fluorescence from POM axons surrounding the cell (r = 0.75, P < 0.0001, two-tailed Pearson’s correlation). g, h, Same as c, d, except activating VP input. Scale bar, 100 μm. i, Group data showing that somas near the TRN centre respond more strongly to VP input than those near the edges (for i, j, n = 31 cells, 6 slices, 4 mice). j, Synaptic responses of TRN cells to VP input correlated with fluorescence from VP axons surrounding the cell (r = 0.68, P < 0.0001, two-tailed Pearson’s correlation).

Given this stark anatomical segregation of projections, it seemed likely that central and edge TRN cells would be selectively targeted by synapses from the VP and the POM, respectively. However, the dendrites of TRN neurons might extend into adjacent zones, leading to functional crosstalk among the circuits1,8,17. To address this, we mapped excitatory synaptic strengths of POM and VP inputs to cells located across the mediolateral axis of the somatosensory TRN. Consistent with the anatomy, synaptic responses to POM inputs were much stronger for TRN edge cells than for central cells, whereas VP inputs evoked the strongest responses in central cells (Fig. 2c–e, g–i, Extended Data Fig. 4b, d). Moreover, synaptic strengths for TRN cells correlated with fluorescence intensities of the afferent terminals near their soma (Fig. 2d, f, h, j). Together, our findings show that primary and higher-order somatosensory thalamic inputs to the TRN are topographically segregated and align with the neurochemical pattern of TRN cell types: the VP projects strongly to CB-expressing central cells, and the POM to SOM-expressing edge cells.

TRN subcircuits are functionally distinct

Primary and higher-order thalamocortical nuclei convey qualitatively different types of information10,12,18,19,20,21, and we asked whether VP and POM communication with the TRN might involve parallel differences in synaptic mechanisms. Thus, we compared dynamic features of the glutamatergic synaptic currents (kinetics and short-term synaptic depression) evoked by photostimulating the VP and POM inputs to the TRN. VP synaptic currents in central TRN cells were brief and depressed deeply during repetitive activation. Conversely, POM currents in TRN edge cells were longer-lasting and more stable, depressing significantly less (Fig. 3a–c, Extended Data Figs. 5, 6).

Fig. 3: Central and edge TRN cells receive thalamic inputs with different kinetics and synaptic depression, and have distinct bursting tendencies.
figure 3

a, Representative locations of TRN cells targeted by the VP (black dot) and by the POM (red dots). b, POM-evoked EPSCs of edge cells are more prolonged than VP-evoked EPSCs of central cells. Left, normalized optogenetically evoked example EPSCs for each cell type (stimulus 1 of c; membrane potential −84 mV). Right, POM-evoked edge cell responses have greater fractional area in later stages of the EPSCs (P < 0.0001, unpaired two-tailed t-test; central: 48 cells, 14 mice; edge: 21 cells, 14 mice). c, Short-term synaptic depression. Left, representative EPSCs from a central cell (VP inputs) and an edge cell (POM inputs), normalized to response peaks (around 2 nA) (10 Hz LED trains). Right, group comparison: short-term depression was greater for central cells than for edge cells (P < 0.0001, ANOVA stimulus 2–10; all P < 0.005 from stimulus 5–10, two-tailed Bonferroni’s t-test; same cells as b). d, Left, intrinsic bursting of example central and edge cells. Offset bursts were triggered by injecting negative current to reach approximately −95 mV, followed by abrupt current removal. Right, number of spikes per burst as a function of position in the TRN (open circles: individual cells; filled circles: zone averages; 34 central cells, 11 mice; 22 medial cells, 10 mice; 16 lateral cells, 10 mice). Edge cells (red) discharged fewer spikes per burst than central cells (black) (unpaired two-tailed t-test, P < 0.0001). Data are mean ± s.e.m.

To understand how these dynamically distinct inputs to TRN neurons might be integrated postsynaptically, we assessed the intrinsic physiological characteristics of these neurons. Central and edge neurons differed in terms of a range of passive and active membrane properties17. Edge cells had higher resistances, lower capacitances and smaller somata than central cells. Action potential kinetics, afterpotentials and threshold currents also differed between the two neuron types (Extended Data Fig. 7, Supplementary Information 1). One notable intrinsic distinction between the TRN cell types, which could powerfully influence responses to synaptic input during certain behavioural states1,2,5,6,22,23,24,25,26, was the much greater tendency of central cells to fire spikes in high-frequency bursts (Fig. 3d). The bursting differences were consistent with stronger T-type (low-threshold) calcium currents in central cells17,27,28,29,30. Thus, nearly all central cells fired ‘offset bursts’ after release from hyperpolarizing stimuli (around 10 spikes per burst). By contrast, under matched stimulus conditions, most edge cells either failed to burst or fired weak bursts (around 2 spikes per burst; Fig. 3d, Supplementary Information 1). Neither cell type exhibited bursting when excited from a more depolarized steady-state (around −74 mV; Extended Data Fig. 7).

We next examined how the observed pathway-specific synaptic and intrinsic properties combine to control the spiking responses of TRN cells to their excitatory thalamic inputs. The brief and depressing synaptic inputs from the VP to central TRN cells, together with the propensity of central cells to burst, suggest that they might respond phasically—initially strong but quickly decreasing with repeated activation. By contrast, the kinetically slow and more stable inputs from the POM to the less bursty edge cells predict initially weaker—but more sustained—spiking.

First we generated simulated synaptic currents that matched the excitatory postsynaptic currents (EPSCs) previously recorded from central and edge TRN cells in response to their respective inputs. We then characterized TRN spike responses elicited by intracellular injection of these currents, delivered while the TRN cells were at their resting potentials (around −84 mV). As predicted, central cells responded to simulated VP inputs with initial bursts that sharply depressed. By contrast, the spiking responses of edge cells to simulated POM inputs were initially much weaker but persisted more during repetitive activation (Fig. 4a, b).

Fig. 4: Thalamus-evoked spiking in central TRN cells is transient, whereas the responses of edge cells are more sustained.
figure 4

a, Simulated synaptic currents, modelled as average recorded EPSCs evoked by activation of VP or POM axons (as in Fig. 3, Methods), were applied to central and edge TRN cells. Left, response of a central cell (membrane potential, Vm) to simulated VP synaptic currents (Iinj). The first stimulus evoked a 12-spike burst, but responses depressed to 0 spikes by stimulus 8. Right, response of an edge cell to simulated POM synaptic currents. The first stimulus evoked 3 spikes, and spiking persisted during repetitive stimulation. b, Mean spike outputs of central cells to simulated VP currents, and edge cells to simulated POM currents (that is, to their native synaptic inputs). TRN cells were held in burst mode (−84 mV). Central cell responses were initially strong and depressed sharply, whereas edge responses were more stable (spike counts from stimulus 1 to stimulus 10 declined by 97% for central cells and 63% for edge cells; P < 0.0001, unpaired two-tailed t-test). For bd: 13 central cells, 11 edge cells, 8 mice. c, Same as b except cells were held at −74 mV to reduce intrinsic bursting. Central responses became more edge-like, yet significant differences remained (spike counts from stimulus 1 to stimulus 10 declined by 77% for central and 33% for edge cells; P < 0.0003, unpaired two-tailed t-test). d, Spike responses of central cells and edge cells (−74 mV) to identical simulated synaptic currents (average of VP and POM EPSCs). Central and edge cell responses did not differ significantly (P = 0.19, unpaired two-tailed t-test). Data are mean ± s.e.m.

We then asked whether differences in intrinsic bursting (Fig. 3d) contribute to these marked differences in synaptically evoked spiking. To address this, we depolarized the TRN cells to −74 mV to partially inactivate T-type calcium channels and reduce intrinsic bursting, and then recharacterized their responses. The responses of central cells became far less phasic and were more persistent, whereas edge cells were hardly affected. Initial spiking was reduced by 41% in central cells but by only 23% in edge cells, and responses to later stimuli in the trains were enhanced more for central cells (Fig. 4c). These results indicate that intrinsic bursting in central TRN cells has a powerful role in their responses to excitatory inputs when that input arrives during relatively hyperpolarized states (for example, during sleep or periods of strong inhibition). The far smaller effects of polarization on edge cells indicates less influence of T-type calcium bursting and, importantly, weaker modulation by the types of membrane potential shifts that are thought to occur during behavioural state transitions1,6,22,23,24,25,26,31.

Finally, we considered whether the dynamic features of VP synaptic inputs to central cells (which are faster and more depressing than those to edge cells; Fig. 3a–c) also contributed to their phasic spike outputs. For this we generated simulated synaptic currents that were the average, in terms of EPSC kinetics and short-term depression, of the VP → central cell and POM → edge cell synaptic currents. We then tested the effects on evoked spiking, with steady-state potentials set to −74 mV to minimize bursting. Notably, responses of the two cell types were almost identical when triggered by the averaged synaptic input; central responses became more sustained and edge responses slightly more phasic (Fig. 4c, d). Together, these results indicate that central and edge cells differentially transform their native excitatory thalamic inputs into distinct spiking outputs through differences both in the dynamics of their synaptic inputs and in their intrinsic burstiness.

TRN inhibitory outputs are subcircuit-specific

To better understand the consequences of the distinct output from the TRN subcircuits, we examined their projections and the inhibitory feedback they produced. We found that the two TRN cell types predominantly inhibited the thalamocortical nuclei that drive them. That is, CB-expressing TRN cells projected to and inhibited neurons of the ventral posterior medial nucleus (VPM), whereas SOM-expressing edge cells bypassed the VPM and instead inhibited the POM (Extended Data Fig. 8). Thus, the primary and higher-order segregation of somatosensory reticulo-thalamic subcircuits seems to be largely reciprocal8,16,32.

Visual and somatosensory TRN have similar subcircuits

To test whether other sensory systems might share the salient structural and functional organization that we have described for the somatosensory TRN, we examined the visual TRN and its associated thalamocortical nuclei—the primary dorsal lateral geniculate and the higher-order lateral posterior nucleus (pulvinar). We found that the organization of the visual TRN, in terms of its synaptic input patterns and the intrinsic physiological properties of its neurons, was very similar to that of the somatosensory system (Extended Data Fig. 9). This suggests that a primary central core, flanked by higher-order edge neurons, might be a widespread TRN motif.

Discussion

Our results imply that sensory regions of the TRN have two discrete subcircuits that are distinguished by their structure and function (Extended Data Fig. 10). Structurally, two types of TRN neurons are segregated into central and edge zones and receive inputs from different thalamocortical nuclei. Functionally, the subcircuits have distinct dynamics, determined by the intrinsic physiology of their respective neurons and the properties of their excitatory thalamic synapses. The subcircuits seem to be tuned to temporal characteristics of the signals they process—transient sensory signals in the primary systems and more temporally distributed signals in the higher-order systems10,11,18,21.

A longstanding hypothesis suggests that the TRN serves as a gatekeeper of information flow, permitting distinct thalamocortical circuits to regulate one another and enabling functions such as selective attention1,5,9,33. It has been proposed that inhibitory crosstalk between thalamic circuits32,34 may underlie such regulation8,35,36. However, the sharp and reciprocal segregation of the subcircuits we observed here suggests that intrathalamic crosstalk might have a minor role. Instead, cross-system regulation in the thalamus could be mediated by other means—including descending cortical control of reticulo-thalamic subcircuits2,37,38,39, the nature of which is only beginning to emerge40.

The distinct kinetics of the parallel TRN subsystems suggests that arriving signals will be filtered through circuit-specific temporal tuning mechanisms. For example, it is generally thought that TRN neurons undergo shifts from bursting to tonic spiking mode as animals transition between behavioural states of quiescence and aroused wakefulness1,22,24,26, owing to activity-dependent changes in membrane potentials and neuromodulatory tone2,6,23,25,28,41. Our results suggest that such state transitions alter the spike mode in primary (central) TRN cells much more strongly than in higher-order (edge) cells27,31. This—in addition to the other kinetic differences between subcircuits that we describe here—are likely to have profound implications for the thalamic processing of afferent signals5, both ascending and descending38.

Previous studies of several species and sensory systems have suggested separate TRN laminae for primary and higher-order connections1,7,8, generally with just a single higher-order layer16,32,42,43. By contrast, in the somatosensory TRN of the mouse, we observed a primary core of CB cells surrounded by a higher-order shell-like zone of SOM cells. In at least one respect, this organization is reminiscent of the visual TRN of the galago, a primate; both systems have SOM-expressing zones that receive inputs from higher-order thalamus, and non-SOM zones receiving primary inputs43. The association between SOM neurons of the TRN and higher-order processing may therefore be a conserved feature of these circuits44. This finding, together with the association between CB expression and first-order processing, are of particular interest because they enable powerful new strategies for investigating the behavioural and perceptual functions of these distinct TRN circuits45.

Methods

Data reporting

No statistical methods were used to predetermine sample size. The experiments were not randomized and the investigators were not blinded to allocation during experiments and outcome assessment.

Mice

All procedures were approved by, and complied with all ethical regulations of, the Brown University Institutional Animal Care and Use Committee. The following mouse lines were used: SOM-IRES-Cre (The Jackson Laboratory, 013044), PV-Cre (The Jackson Laboratory, 008069), Calb1-IRES-Cre-D (‘CB-Cre’; The Jackson Laboratory, 028532), Vglut2-IRES-Cre (The Jackson Laboratory, 016963), GPR26-Cre (STOCK Tg(Gpr26-cre)KO250Gsat/Mmucd, MMRRC), Ai14 (The Jackson Laboratory, 007908), ICR (Charles River, CD-1[ICR], Strain Code 022). To fluorescently target SOM or PV cells in TRN, we bred the respective homozygous Cre mice (SOM-IRES-Cre, PV-Cre) with homozygous Cre-dependent tdTomato reporter mice (Ai14). In some experiments we instead injected viruses carrying Cre-dependent GFP genes into Cre mice (specified in the descriptions of the individual experiments, below). Of the 124 mice used in this study, 91 had C57 genetic backgrounds, 10 had ICR genetic backgrounds, and 23 had mixed C57/ICR backgrounds. Mice were maintained on a 12 h:12 h light/dark cycle, group-housed, and provided food and water ad libitum.

Immunohistochemistry

Mice from postnatal day (P)22–26 were deeply anaesthetized with Beuthanasia-D and intracardially perfused with 0.1 M phosphate buffer (PB) followed by 4% paraformaldehyde (PFA; in PB). Brains were post-fixed overnight at 4 °C in the same fixative and then transferred to a 30% sucrose/0.1 M PB solution until sectioning (4 °C, 2–3 days). Brain sections (40-μm thick) were cut on a freezing microtome at a somatosensory thalamocortical plane (35° tilt from coronal)46 designed to contain the somatosensory thalamus, TRN, barrel cortex and many of their interconnections. Next the sections were immunostained. In brief, they were washed 5 times in 0.1 M phosphate buffer containing 0.15 M NaCl, pH 7.4 (PBS) (5 min per wash), pre-incubated for 2 h at room temperature with a blocking solution (10% normal goat serum, 2% Triton X-100, 0.1% Tween 20 in 0.1 M PB), then incubated with primary antibodies for 5 days at 4 °C. After the primary incubation, sections were washed 8 times in PBS (5 min per wash), pre-incubated for 2 h in blocking solution, incubated with a secondary antibody solution for 3 days at 4 °C, then washed 8 times in PBS and 3 times in PB (5 min per wash). Sections were mounted and coverslipped using Prolong Gold or Prolong Gold with DAPI (Molecular Probes P36930 or P36931). Primary antibodies were: mouse monoclonal anti-parvalbumin (1:1,000, Swant clone 235, Lot 10-11F), rabbit polyclonal anti-calbindin D-28k (1:1,000, Swant clone CB-38a, Lot 9.03), mouse monoclonal anti-calbindin D-28k (1:1,000, Swant clone 300, Lot 07(F)), mouse monoclonal anti-NeuN (1:1,000, Millipore MAB377, Clone A60, Lot 2549411). Secondary antibodies were goat anti-mouse IgG (H+L) cross-adsorbed secondary antibody Pacific Blue (1:250, Molecular Probes P31582), goat anti-rabbit IgG (H+L) cross-adsorbed secondary antibody Alexa Fluor 647 (1:500, Molecular Probes A21244), goat anti-mouse IgG (H+L) cross-adsorbed secondary antibody Alexa Fluor 488 (1:250, Molecular Probes A11001), goat anti-rabbit IgG (H+L) highly cross-adsorbed secondary antibody Alexa Fluor 488 (1:250, Molecular Probes A11034). To test the specificity of the antibodies, no-primary and no-secondary controls were conducted for each. In addition, for the calbindin antibodies, we performed assays in which the antibodies were pre-adsorbed with calbindin protein (3.33 μg/1 μl of antibody, Swant recombinant rat Calbindin D-28) before tissue incubation. There was no clear labelling under any of the control conditions, providing support for the effectiveness and specificity of the antibodies. Confocal image stacks were taken on a Zeiss LSM 800, 20× objective, 0.26 μm pixel diameter.

For analysis of the immunohistochemical material, we selected sections spanning the somatosensory sector of TRN (approximately −1.7 mm to −1.1 mm from bregma), generally focusing on the anterior–posterior centre of that range (that is, −1.4 mm from bregma). The CellCounter ImageJ plugin was used to count cells positive for the tested markers. For accurate alignment across subjects and sections, small differences in TRN shapes were minimized by warping them to an average reference TRN image using the bUnwarpJ plugin in ImageJ. At least 12 landmark points were applied around the outer boundaries of each TRN (evenly spaced, 100 μm between points) for alignment. The TRN boundaries were identified using either the SOM-Cre × tdTomato or PV-Cre × tdTomato channel. Resulting corrections were applied to all channels.

Fluorescence in situ hybridization

Mice (P21–P27) were deeply anaesthetized with isofluorane and decapitated. The brains were removed while submerged in 4 °C saline solution and fresh-frozen on liquid nitrogen. Frozen brains were sectioned (18-μm thick) at the somatosensory thalamocortical plane (described in the above section ‘Immunohistochemistry’) using a cryostat (Leica), adhered to SuperFrost Plus slides (VWR, 48311-703), and refrozen (−80 °C) until used. Samples were fixed (4% PFA), processed as instructed in the ACD RNAScope Multiplex Fluorescent v2 Assay, and then immediately incubated with DAPI and coverslipped with ProLong Gold (Molecular Probes P36930). Probes for PV (Mm-Pvalb-C3, 421931-C3), SOM (Mm-Sst, 404631), CB (Mm-Calb1-C2, 428431-C2), and tdTomato (tdTomato-C3, 317041-C3) were purchased from Advanced Cell Diagnostics. Probes were visualized using the TSA Plus Fluorescein (1:1,000, Perkin Elmer NEL741E001KT), TSA Plus Cyanine 3 (1:1,000, Perkin Elmer NEL744E001KT), and TSA Plus Cyanine 5 (1:1,000, Perkin Elmer NEL745E001KT) evaluation kits. Confocal image stacks were taken on a Zeiss LSM 800, 20× objective, 0.26 μm pixel diameter.

One section was selected from each of six mice for the fluorescence in situ hybridization (FISH) analysis. These sections were centred on the somatosensory sector of the TRN (approximately 1.4 mm posterior to bregma). Four of the six mice were C57 wild-types. The remaining two mice were PV-Cre × Ai14 mice that expressed tdTomato in PV-Cre cells (the tdTomato-C3 probe, 317041-C3, was used as a proxy for PV for these two mice). Images were quantified using ImageJ. Regions of interest (ROIs) were manually drawn for each cell on the basis of the fluorescence signal of all three channels (SOM, CB and either PV or tdTomato). For quantification, the mean fluorescence intensity for a cell’s ROI was expressed as a percentage of the intensity range across the TRN of the section: 100 × ((mean intensity for the cell’s ROI)/(maximum pixel intensity in the TRN − minimum pixel intensity in the TRN)). Cells were considered positive for an RNA marker if this normalized expression for the marker was greater than 7.5%, which best matched qualitative visual assessment of expression thresholds.

Probes for PV (Mm-Pvalb-C4, 421931-C4), Cre (CRE-C3, 312281-C3), and SOM (Mm-Sst, 404631) were used to assess correspondence between Cre and Sst mRNA expression in the SOM-Cre mouse line (4 SOM-Cre × ICR mice).

Cell count analysis

For both immunohistochemical and FISH analysis, cells were counted in a region centred on somatosensory TRN extending 300 μm along the dorsal–ventral axis of the nucleus (dorsal–ventral boundaries indicated by the brackets in Fig. 1a, b, or the boxes in Extended Data Figs. 1b, 2a). This region consistently received axonal projections from the S1 cortex (data not shown) and the somatosensory thalamus (VP and POM; Fig. 2, Extended Data Fig. 4). Central/edge boundaries were drawn at 20% and 80% of the medial–lateral distance across the TRN.

Stereotactic injection procedure

Mice were anaesthetized with a mixture of Ketaset and Dexdormitor diluted in sterile saline (Ketaset, 70 mg/kg; Dexdormitor, 0.25 mg/kg; intraperitoneally). Once deeply anaesthetized, mice were placed into a stereotactic frame, and a craniotomy was made over the VP, POM, dorsal lateral geniculate nucleus of the thalamus (dLGN), lateral posterior nucleus of the thalamus (LP) or TRN. Virus solution was then pressure-ejected into the brain via a glass micropipette attached to a Picospritzer pressure system at a maximum rate of around 0.05 μl/min (10–30 min total injection times). After injection, the pipette was left in place for around 10 min before being slowly withdrawn from the brain. After surgery, mice were given Antisedan (2.5 mg/kg) to reverse the effects of Dexdormitor, and they were allowed to recover on a heating pad for around 1 h before being returned to their home cage. Experiments were usually performed about 10 days after the virus injections to allow for sufficient GFP or opsin expression (mean, 10.4 ± 0.3; range, 8–19 days).

Adeno-associated viruses (AAVs) and lentiviruses were acquired from the University of North Carolina, Addgene or the University of Pennsylvania Vector Cores and used at the following titers: (1) rAAV2/hSyn-ChR2(H134R)-eYFP-WPREpA (titre = ~3.93 × 1012 vg/ml), (2) pLenti-Synapsin-hChR2(H134R)-eYFP-WPRE (titre = ~2.53 × 1010 vg/ml), (3) AAV2.Syn.DIO.hChR2(H134R)-eYFP.WP.hGH (titre = ~2.32 × 1013 vg/ml), (4) rAAV2/hSyn-ChR2(H134R)-mCherry-WPREpA (titre = ~2.2 × 1012 vg/ml), (5) AAV9.Syn.DIO.EGFP.WPRE.hGH (titre = ~6.25 × 1012 vg/ml), (6) rAAV2/Syn-Flex-ChrimsonR-TdT (titre = ~3.8 × 1012 vg/ml).

For experiments testing projections from specific thalamocortical nuclei to the TRN (or from subtypes of TRN cells to thalamocortical nuclei), relatively small volumes (0.16–0.33 μl) of either AAV2 or lentivirus (described above) were stereotactically injected into individual presynaptic nuclei using the following mouse strains and coordinates. To test VP projections to the TRN (Figs. 2, 3, Extended Data Figs. 46), virus was injected into SOM-Cre × Ai14 (n = 15), ICR (n = 4), Vglut2-Cre × ICR (n = 5), or PV-Cre × Ai14 (n = 1) mice between P12 and P16 (mean 13.1 ± 0.2 days). Average coordinates from bregma for VP were 1.99 mm lateral, −0.74 mm posterior, 3.08 mm depth. For POM projections to TRN (Figs. 23, Extended Data Figs. 46), virus was injected into SOM-Cre × Ai14 (n = 8), ICR (n = 6), PV-Cre × Ai14 (n = 2), or GPR26-Cre × ICR (n = 2) mice between P11 and P14 (mean 12.0 ± 0.3). Coordinates for POM were 1.36 mm lateral, −1.17 mm posterior, 2.86 depth. For dLGN projections to TRN (Extended Data Fig. 9), virus was injected into SOM-Cre × Ai14 (n = 2), PV-Cre × Ai14 (n = 2) or GPR26-Cre × ICR (n = 1) mice between P14–P19 (mean 16.0 ± 0.8). Coordinates for dLGN were 2.38 mm lateral, −1.45 mm posterior, 2.35 depth. For LP projections to TRN (Extended Data Fig. 9), virus was injected into SOM-Cre × Ai14 (n = 6), PV-Cre × Ai14 (n = 6) or CB-Cre × Ai14 (n = 1) mice between P14–P17 (mean 15 ± 0.3). Coordinates for LP were 1.63 mm lateral, −1.36 mm posterior, 2.3 mm depth. For tests of the inhibitory outputs of TRN cell subtypes to VP and POM neurons (Extended Data Fig. 8), virus was injected into SOM-Cre × ICR (n = 5), CB-Cre × Ai14 (n = 1) or CB-Cre × ICR (n = 2) mice between P12 and P15 (mean 13.7 ± 0.3). Coordinates for these TRN injections were 2.2 mm lateral, −0.53 mm posterior, 3.0 mm depth.

For assessments of the positions of TRN cell subtypes (Fig. 1e, Extended Data Figs. 1, 3), relatively large volumes (1.3–2 μl) of AAV9 (described above) were injected across the entire dorsal–ventral extent of the TRN in SOM-Cre × Ai14 (n = 10), SOM-Cre × ICR (n = 2), PV-Cre × Ai14 (n = 4) or ICR (n = 2) mice between P14 and P25 (mean 16.3 ± 0.9). Average coordinates for these TRN injections were 2.2 mm lateral; −0.53 mm posterior, with continuous outflow of virus solution from 2.2 mm to 4.4 mm depth.

Slice preparation

Brain slices were prepared from P22–P34 mice of either sex as previously described40. Mice were deeply anaesthetized with isofluorane and decapitated. The brains were removed while submerged in cold (4 °C) oxygenated slicing solution containing (in mM): 3.0 KCl, 1.25 NaH2PO4, 10.0 MgSO4, 0.5 CaCl2, 26.0 NaHCO3, 10.0 glucose and 234.0 sucrose. Brains were then mounted, using a cyanoacrylate adhesive, onto the stage of a vibrating tissue slicer (Leica VT1000 or VT1200S) and somatosensory thalamocortical brain slices (300 μm thick, 35° tilt from coronal46) containing VP, POM, TRN, S1 and portions of dLGN and LP were obtained. Slices were incubated for around 1 min in the cold sucrose-based slicing solution, then transferred for 20 min to a holding chamber containing warm (32 °C) oxygenated (5% CO2, 95% O2) artificial cerebrospinal fluid (ACSF) solution. Finally, the slices were allowed to equilibrate in ACSF for 60 min at room temperature before imaging or recording. The ACSF solution contained (in mM): 126.0 NaCl, 3.0 KCl, 1.25 NaH2PO4, 1.0 MgSO4, 1.2 CaCl2, 26.0 NaHCO3, and 10.0 glucose.

Live imaging

Live sections (300 μm) centred on the somatosensory sector of TRN (−1.4 mm from bregma) were imaged using Nikon or Zeiss upright microscopes with 2.5–5× objectives and Andor Zyla sCMOS cameras. Both epifluorescent and transmitted light (bright-field) images were obtained to characterize the topographical positions of the TRN cell types and their connections with thalamic relay nuclei.

To generate the group maps showing the average VP and POM projections to the TRN (Fig. 2b), the TRN of each live slice was first outlined using either tdTomato expression driven by SOM-Cre (n = 6 for POM, n = 4 for VP) or bright-field images (n = 3 for POM, n = 4 for VPM). Those outlines were then used to warp the slice images to a common reference TRN (with the bUnwarpJ plugin in ImageJ—described for the immunohistochemical analysis above), allowing precise alignment across mice for averaging. The central/edge boundaries were drawn at 20% and 80% of the medial–lateral distance across the TRN.

Whole-cell recording procedure

Brain slices (300 μm) were placed in a submersion-type recording chamber maintained at 32 ± 1 °C and continuously superfused with oxygenated ACSF (above). Neurons were visualized for recording using DIC-IR optics with 40× water immersion objectives. Patch pipettes had tip resistances of 3–6 MΩ when filled with a potassium-based internal recording solution containing (in mM): 130.0 K-gluconate, 4.0 KCl, 2.0 NaCl, 10 HEPES, 0.2 EGTA, 4.0 ATP-Mg, and 0.3 GTP-Tris, 14.0 phosphocreatine-K (pH 7.25, ~290 mOsm). During all recordings, pipette capacitances were neutralized. Series resistances (~12–32 MΩ) were compensated online (100% for current–clamp, 60–70% for voltage-clamp). Pharmacological agents (stated in the figure legends when used) were diluted in ACSF just before use and applied though the bathing solution. Voltages reported here were corrected for a 14 mV liquid junction potential. The reversal potential for GABAA receptor-mediated responses in thalamic relay cells was −91 mV.

Measurements of intrinsic physiological properties

Resting membrane potentials were measured within 2 min of break-in. Steady-state potentials were adjusted to −74 mV or −84 mV with intracellular current to test physiological properties in tonic or burst mode, respectively. Input resistances (Rin) and membrane time constants (τm) were calculated from voltage responses (~3 mV deflections) to small negative current injections (3–50 pA, 600–1,000 ms). For τm, the voltage responses were fitted with a single exponential to the initial 50 ms of the response, omitting the first ms. Rin values were measured using Ohms law and input capacitances Cin were calculated as τm/Rin.

Threshold (rheobase) currents were measured as the minimum injected currents required to discharge an action potential (AP; determined using 1 s duration currents, 5 pA step increments). All other AP and after-AP properties were measured from the first AP discharged at the threshold current, but APs were only analysed if discharged in the initial 200 ms). AP voltage thresholds were measured, and verified by visual inspection, as the potential at which the rate of rise became greater than 10 V/s. AP amplitudes and after-AP properties were measured relative to the threshold potential. AP widths were measured at half of the AP amplitude (threshold voltage to peak). Fast afterhyperpolarizations (fast AHPs) were measured as the most hyperpolarized potential immediately succeeding the AP. In tonic mode, afterdepolarizations (ADPs) were measured as the most depolarized potential within 20 ms of the AP, and slow AHPs (sAHPs) were measured as the most hyperpolarized potential within 100 ms after the ADP. In burst mode, the AHP following the burst was measured as the minimum potential within 150 ms following the burst. ADP and sAHP measurements were not considered if a second AP (or tonic APs after a burst) confounded the measurements.

Repetitive spiking properties in tonic and burst mode were measured using positive current steps (25–200 pA, 25 pA increments, 1 s duration). Spike frequency adaptation was quantified as an adaptation ratio (frequency of the last 2 APs divided by the frequency of the first 2 APs) averaged across all sweeps in which the frequency of the last 2 APs was 20–60 Hz.

To elicit offset bursts (Fig. 3, Extended Data Fig. 9, Supplementary Information 1), the steady-state potential was adjusted to −74 mV with intracellular current, then 1 s duration negative currents were injected. Offset bursts were measured from trials in which negative current (20–300 pA, 20–25 pA test increments) led to a voltage of approximately −94 mV at the end of the current step.

Photostimulation

Synaptic physiology experiments were performed on inputs to TRN cells from excitatory thalamic relay neurons (Figs. 2, 3), and on inhibitory outputs of TRN cells to thalamic relay neurons. In both cases, ChR2 was optically excited using white light-emitting diodes (LEDs) (Mightex LCS-5500-03-22) controlled by Mightex LED controllers (SLCAA02-US or BLS-1000-2). The light was collimated and reflected through a 40× water immersion objective, resulting in a spot diameter of ~400 μm and a maximum LED power at the focal plane of 29.2 mW. The stimuli, delivered as 10 Hz trains of 1 ms flashes, were typically directed at ChR2-expressing presynaptic terminals by centring the light spot over the recorded postsynaptic cells (the postsynaptic cells did not express ChR2).

In a subset of experiments (Extended Data Fig. 6a–c), within-cell comparisons were made between TRN cell responses to stimulation directed at opsin-expressing presynaptic axons/terminals (from VP or POM) and stimulation directed further upstream, at or near the VP or POM cell bodies of origin. These experiments tested whether short-term synaptic plasticity differed when optogenetically stimulating soma/proximal axons of presynaptic cells versus their terminal boutons, as has been shown for some pathways47.

Simulated synaptic current injections

Simulated EPSC waveforms for thalamic (VP and POM) inputs to the TRN were generated from averaged measurements of optogenetically evoked synaptic responses (recorded in voltage clamp at −84 mV) from central and edge TRN cells (as in Fig. 3). These simulated synaptic currents for VP and POM inputs were matched for total synaptic charge across the 10 Hz trains (49.5 pC) and injected into the central and edge TRN cells to test features of integration in each cell type (described in the text and in Fig. 4).

Data analyses

Analyses of electrophysiological data were performed using CED Signal 6, Molecular Devices Clampfit 10, MATLAB and Microsoft Excel. Analyses of anatomical data were performed using ImageJ (plugins used: CellCounter, ROIManager, bUnwarpJ) and Microsoft Excel.

Statistical analyses

Statistical comparisons were performed using GraphPad Prism7 or SigmaPlot. Statistical tests used are indicated in the main text and figure legends. For representation of group data, centre values are means and error bars show s.e.m. Statistical significance was defined as P < 0.05, unless otherwise noted.

Reporting summary

Further information on research design is available in the Nature Research Reporting Summary linked to this paper.