Abstract
The ventromedial motor thalamus (VM) is implicated in multiple motor functions and occupies a central position in the cortico-basal ganglia-thalamocortical loop. It integrates glutamatergic inputs from motor cortex (MC) and motor-related subcortical areas, and it is a major recipient of inhibition from basal ganglia. Previous in vitro experiments performed in mice showed that dopamine depletion enhances the excitability of thalamocortical (TC) neurons in VM due to reduced M-type potassium currents. To understand how these excitability changes impact synaptic integration in vivo, we constructed biophysically detailed mouse VM TC model neurons fit to normal and dopamine-depleted conditions, using the NEURON simulator. These models allowed us to assess the influence of excitability changes with dopamine depletion on the integration of synaptic inputs expected in vivo. We found that VM neuron models in the dopamine-depleted state showed increased firing rates with the same synaptic inputs. Synchronous bursting in inhibitory input from the substantia nigra pars reticulata (SNR), as observed in parkinsonian conditions, evoked a postinhibitory firing rate increase with a longer duration in dopamine-depleted than control conditions, due to different M-type potassium channel densities. With β oscillations in the inhibitory inputs from SNR and the excitatory inputs from cortex, we observed spike-phase locking in the activity of the models in normal and dopamine-depleted states, which relayed and amplified the oscillations of the inputs, suggesting that the increased β oscillations observed in VM of parkinsonian animals are predominantly a consequence of changes in the presynaptic activity rather than changes in intrinsic properties.
Significance Statement
The ventromedial motor thalamus is implicated in multiple motor functions. Experiments in vitro showed this area undergoes homeostatic changes following dopamine depletion (parkinsonian state). Here, we studied the expected impact of these changes in vivo, using biophysically detailed modeling. We found that dopamine depletion increased firing rate in the ventromedial thalamocortical neurons and changed their responses to synchronous inhibitory inputs from substantia nigra pars reticulata. All thalamocortical neuron models relayed and amplified β oscillations from substantia nigra reticulata and cortical/subcortical inputs, suggesting that increased β oscillations observed in parkinsonian animals predominantly reflect changes in presynaptic activity.
Introduction
Motor thalamus is a critical structure for multiple aspects of motor control (Bosch-Bouju et al., 2013). In rodents, it comprises the ventromedial (VM), the ventral-anterior (VA), and the ventrolateral nuclei (Kuramoto et al., 2011). In particular, VM and VA are the major recipients of inhibition from substantia nigra pars reticulata (SNR; Chevalier and Deniau, 1982; Kuramoto et al., 2011; Rovó et al., 2012), the output area of the basal ganglia. Excitatory inputs from primary and premotor cortices [motor cortex (MC)] play a key role in VM processing, supporting persistent motor preparatory activity in a closed loop (Z.V. Guo et al., 2017; K. Guo et al., 2018). In vivo experiments implicated VM in movement preparation and vigor control (Takahashi et al., 2021), while SNR mediates strong and temporally precise inhibition that controls movement direction and initiation in behaving animals (Morrissette et al., 2019; Catanese and Jaeger, 2021).
Parkinson’s disease (PD) is a neurodegenerative disorder that is primarily due to the death of dopamine neurons in the Substantia Nigra pars compacta. In rodents, dopamine depletion increases synchrony and bursting in the SNR (Lobb and Jaeger, 2015; Brazhnik et al., 2016; Willard et al., 2019). This maladaptive activity is conveyed to VM, and this pathway may be associated with deficits in movement selection and initiation (Kravitz et al., 2010; Morrissette et al., 2019; Takahashi et al., 2021). Moreover, local field potential (LFP) recordings showed increased β oscillations in SNR, MC, and VM of parkinsonian rats (Brazhnik et al., 2012, 2016; Nakamura et al., 2021), a typical hallmark of parkinsonian pathophysiology (Brown et al., 2001; Cassidy et al., 2002; Kühn et al., 2005; Sharott et al., 2018).
Previous experiments showed that dopamine depletion induced increased excitability in the VM of mice and increased their ability in generating rebound bursting on hyperpolarization (Bichler et al., 2021). Since there is little or no dopamine innervation to rodent VM (García-Cabezas et al., 2009), increased excitability is likely a homeostatic response due to increased inhibitory input from the SNR. As previously shown, such intrinsic excitability changes are broadly exhibited to compensate for a change in synaptic input balance (Desai et al., 1999; Turrigiano and Nelson, 2000). While experimental evidence suggests that the nigral synapses are well positioned to evoke rebound bursting in VM thalamocortical (TC) neurons (Bodor et al., 2008; Edgerton and Jaeger, 2014; Bichler et al., 2021), this hypothesis has not been tested in vivo. It remains unknown whether the effects of dopamine depletion on intrinsic VM properties may contribute to the generation of β oscillations in PD (Brazhnik et al., 2016), or exacerbate rebound bursting in vivo.
Because the contribution of intrinsic properties to firing patterns in vivo is hard to measure directly, we employed biophysically detailed modeling, using the NEURON simulator (Hines and Carnevale, 1997). Specifically, we fitted TC neuron models, replicating the different firing properties observed in normal and parkinsonian states (Bichler et al., 2021). Based on widely accepted thalamic literature (Reichova and Sherman, 2004; Bickford, 2016; Halassa and Acsády, 2016), we distinguished between driver (DRI), driver-like (DRI-l), and modulator (MOD) types of inputs. Specifically, we modeled four classes of afferent inputs to VM: glutamatergic MODs, approximating inputs from MC layer 6, glutamatergic DRI-l inputs from MC layer 5 and subcortical areas, and GABAergic inputs from SNR and reticular thalamic nucleus (RTN). Experimental evidence does not support the presence of excitatory drivers to VM (Rovó et al., 2012). Each afferent input replicated synaptic conductances and subcellular distributions observed experimentally (Bodor et al., 2008; Edgerton and Jaeger, 2014; Gornati et al., 2018), where the synapses were activated by artificial spike trains, replicating the activity during motor performance (Inagaki et al., 2022). This approach allowed us to control the firing patterns of the inputs, to reproduce either normal firing features or distinct firing features of the parkinsonian state. We then added varying amounts of synchronous bursting or β oscillations to these inputs. We found that TC neuron models in dopamine-depleted state responded at higher firing rate than TC neuron models in normal conditions to all in vivo-like input patterns. Synchronous nigral bursting was unable to evoke rebound bursting, as the synaptic excitation prevented sufficient hyperpolarization to de-inactivate T-type Ca2+ channels. However, synchronous nigral bursts still evoked a significant postinhibitory firing rate increase in TC neuron models, due to slow recovery of potassium currents on repolarization. This increase lasted longer in models fitted to dopamine-depleted conditions. Adding β oscillations in SNR inputs resulted in significant spike-phase locking in TC neuron model firing in both normal and parkinsonian states. This phase locking became stronger when excitatory inputs contained β oscillations at a 180-degree phase shift. These results suggest that the increased excitability induced by dopamine depletion does not affect β oscillations in VM TC neurons. In fact, in both normal and dopamine-depleted conditions our VM TC neuron models could amplify such oscillations.
Materials and Methods
Simulation
The morphology of a reconstructed TC neuron was divided in 40-μm-long compartments and used to construct multicompartmental biophysically detailed models with 11 Hodgkin–Huxley-style (HH) active conductances at varying densities (for details, see below, Neuron morphology and Intrinsic membrane properties).
Simulations were implemented in a fully integrated NEURON and Python environment (Hines and Carnevale, 1997). To fit the neuron models to physiological targets, we employed multiobjective optimization, based on genetic algorithms, using the BluePyOpt package (Van Geit et al., 2016). Here, each solution corresponds to a neuron model, encoded as an array of parameters associated with the intrinsic properties. The optimizer calculated model fitness by comparing the traces generated by each model against a set of features extracted from experimental recordings, simulating a battery of experiments with adaptive time steps, using the CVode solver for partial differential equations (for more details, see below, Fitting neuron models). The fitting process was executed on the supercomputer Expanse, managed by the San Diego Super Computer Center (SDSC), through the Neuroscience Gateway (NSG; http://www.nsgportal.org; Sivagnanam et al., 2013). Following model construction, simulations to evaluate model performance were executed using the NEURON version known as “CoreNEURON” (Kumbhar et al., 2019), and parallelized with the MPI4Py package. These simulations were executed on the supercomputer Expanse.
All simulation parameters such as temperature, reversal potentials of ion species, effects of ion channel blockers, and holding membrane voltage reflected the specifics of the experimental stimulation protocols and slice conditions (Table 1), while the number of active synapses reproduced experimentally observed values (Table 4). For each experiment, we calculated the values of reversal potential for sodium and potassium using the Nernst equation, accounting for the relative concentrations of ions in the artificial cerebrospinal fluid (aCSF) and pipette solutions. In some experiments, tetrodotoxin (TTX), a sodium channel blocker, was applied, and/or the pipette solution contained cesium, a potassium channel blocker. We accounted for their effects by turning off sodium and potassium HH-conductances in the simulations, respectively. To simulate in vivo-like conditions, we adjusted the temperature to 37°C (Figs. 3-8), which is higher than any simulated in vitro experiments (cf. 24–34°C; see Table 1). Increasing the temperature speeds up the kinetics of the ion channels in a way consistent with the Q10 rule (see below, Intrinsic membrane properties).
Code accessibility
The source code is publicly available on ModelDB and GitHub (https://github.com/FrancescoCavarretta/BGMT).
Modeling process overview
We designed a data-driven modeling pipeline (Fig. 1) to build realistic TC neuron models from VM of mice in normal and parkinsonian states (green boxes) along with afferent inputs (blue boxes). We defined a biophysically detailed model for TC neurons, comprising a full three-dimensional morphology from the VM of mouse (see below, Neuron morphology) and 11 HH models of ion channels (see below, Intrinsic membrane properties), with subcellular distributions based on experimental data of thalamic and pyramidal cortical neurons (see below, Ion channel distributions). The model comprised multiple free parameters, with values determined by multiobjective optimization (see below, Template data taken from in vitro experiments and Fitting neuron models). We then modeled the main synaptic inputs to VM, reproducing the subcellular distributions, numbers of synaptic contacts, and unitary conductances observed experimentally (see below, Modeling synaptic inputs), with realistic levels of presynaptic activity (see below, Presynaptic activity: artificial spike train generation).
Neuron morphology
We based our model on a full three-dimensional reconstruction of a TC neuron from the VM of mouse (id. AA0719; https://www.janelia.org/project-team/mouselight). For modeling purposes, we retained only the proximal 70-μm-long portion of the axon (i.e., axonal initial segment), approximated by two cylindrical compartments (diameter: 1.5 μm; length: 35 μm), as we did not further follow action potential propagation. The original reconstruction lacked estimates for dendritic diameters while the soma was represented by the cross-sectional perimeter only. We replaced the soma with a cylindrical section (diameter: 36.9 μm; length: 24 μm), yielding a cross-section surface of 293 μm2, consistent with the experimental estimates of VM TC neurons of rats (cf. 288 ± 13 and 298 ± 2 μm2; ±SE; Sawyer et al., 1989). To model the dendritic diameters, we defined a recursive formula based on Rall’s 2/3 Power Law:
If the dendrite has children:
with M1 = 0.515, Q1 = 0.182.
Otherwise:
The values of the free parameters (i.e., M*, Q*, A3, B3, C3, and DIAMmin) were estimated by calculating the least square regression curves from partial histologic reconstructions of VM TC neurons (n = 5) of mouse obtained in our laboratory, using biocytin and Neurolucida. Specifically, we estimated Q1 and M1 by linear regression between the diameters of parent dendrites versus the sum of the diameters of children to the 1.5 power; Q2 and M2 by linear regression of diameters versus branch depth for the dendrites that did not have branches; C3, B3, and A3 were estimated by fitting the exponential function C3 + A3 exp(-B3d) on the diameters of the primary dendrites normalized to the average as a function of the distance from soma.
Intrinsic membrane properties
NEURON models require the specification of passive properties, such as specific membrane specific membrane conductivity (g; i.e., inverse of resistivity), axial resistivity (ri; i.e., intracellular or cytoplasmic resistivity), specific membrane capacitance (cm), and resting potential (VRest). In our models, their values were uniform across all sections. In particular, cm was a constant using the canonical value (1 μF/cm2; Gentet et al., 2000), while the others were treated as free parameters, estimated by our genetic algorithm (see below, Fitting neuron models).
The active membrane properties of our models consisted of 11 HH-conductances (Extended Data Table 3-1), with states dependent on the membrane potential and/or intracellular Ca2+ concentration. They were transient (NaT) and persistent (NaP) sodium currents; delayed rectifier (KDR), A-type (KA), delaying (KD), and M-type (KM) potassium currents; H-type nonspecific cation current (IH); T-type (CaT) and L-type (CaL) Ca2+ currents; small conductance (SK) and big conductance (BK) Ca2+-dependent potassium currents. Most HH-conductances were specified as used in a previous model of ventrobasal TC neuron (NaT, KDR, KA, IH, CaT, CaL, SK; Iavarone et al., 2019), while KM was specified as used in a previous model of VM TC neuron (Bichler et al., 2021). We derived the NaP dynamics from the NaT model (Iavarone et al., 2019), decreasing the half-values of activation and inactivation curves by 14 mV, consistent with experimental estimates from pyramidal cells (Hu et al., 2009). We based the KD model on the dynamics of the shaker-related potassium channels observed in TC neurons of rats (Lioudyno et al., 2013), where the effects of these channels delayed the action potential (AP) onset (Lioudyno et al., 2013), similarly to the effects of KD currents observed in pyramidal neurons (Storm, 1988). We added the BK to induce firing rate adaptation in TC neurons (Ehling et al., 2013), observed also in VM (Fig. 2; Bichler et al., 2021). The BK dynamics were voltage and Ca2+ -dependent (Rothberg and Magleby, 2000), and accounted for the dependence on temperature (Q10 = 2.5; Yang and Zheng, 2014).
Extended Data Figure 2-1
Diversity of experimental and simulated responses of thalamocortical neurons from ventromedial motor thalamus in normal and parkinsonian states. Neuron responses to increasing depolarizing current injections for different neurons (n = 3) and models (n = 3), under the same experimental conditions as in Figure 2A. Both experimental and simulated recordings display variability in different action potential (AP) properties, such as AP amplitude, AHP depth, AP accommodation. Download Figure 2-1, TIF file.
Extended Data Figure 2-2
Comparison of action potential properties between experimental and simulation traces. Experimental and simulation traces displayed similar firing properties. A subset of experimental and simulation traces is shown in Figure 1A and Extended Data Figure 1-1. Each data point (black) is associated with a single TC neuron (Control: n = 9 and 6-OHDA: n = 7 neurons) and TC neuron model (Control: n = 16 and 6-OHDA: n = 17 models). We compared after-hyperpolarization (AHP) depth (top, left), action potential (AP) amplitude (top, right), firing (rate) adaptation (bottom, left), and sag amplitudes (bottom, right) of TC neurons (Control: n = 9 and 6-OHDA: n = 7 neurons; blue) and TC neuron models (Control: n = 16 and 6-OHDA: n = 17 models; orange). The same firing properties were considered as a target during the model fitting (see also Materials and Methods) and are described in Table 2. We used protocol 1 (i.e., the same stimulation protocol as in Figure 1A; for details, see Materials and Methods) to measure AHP depth, AP amplitude, and firing adaptation, while we used protocol 3 for sag amplitudes, in both experiments and simulations. The values of AHP depth and AP amplitude were estimated using threshold amplitudes of step current, while we used the minimal current evoking at least five APs for firing rate adaptation. As the threshold current depends on the input resistance of neurons, it was estimated for each real neuron and model. To measure the sag amplitudes, we used a hyperpolarizing step current (−200 pA). Comparing experimental and simulation traces, we found that most firing features were statistically indistinguishable in both normal and parkinsonian conditions (Mann–Whitney, p > 0.04), except the AP amplitudes (Mann–Whitney, Control: p < 0.01; 6-OHDA: p < 0.001), which were lower in simulations than experiments (Control: 21%; 6-OHDA: 33%). Download Figure 2-2, TIF file
Consistent with previous experimental studies (Womack et al., 2004), intracellular Ca2+ was structured in three separate microdomains, associated with distinct decay time constants. In particular, the Ca2+ flowing through CaL channels interacts with two microdomains that activate SK and BK channels separately, while Ca2+ flowing through CaT channels interacts with a third microdomain that does not activate other ion channels. In our models, the Ca2+ concentration of each microdomain determined the reversal potential for the related Ca2+ channel, calculated with the Goldman–Hodgkin–Katz flux equation.
Ion channel distributions
Based on experimental data of TC neurons (Budde et al., 1998; Williams and Stuart, 2000), we modeled the subcellular distributions of ion channels as follows:
The conductance density of voltage-dependent potassium channels (i.e., KDR, KA, KM, KD) is uniform throughout the somatodendritic shaft.
The conductance density of sodium channels (i.e., NaT, NaP) decreases at each branch point.
For T-type Ca2+ channels (i.e., CaT), the conductance density is double as that of the soma for primary dendrites, and half of the soma for all the other dendrites.
The conductance density of L-type Ca2+ channels (i.e., CaL) decreases to half of the soma within the proximal 10-μm-long portion of primary dendrites, and to a third in all other dendritic locations.
As the subcellular distribution of small-conductance and big-conductance Ca2+-dependent potassium channels (i.e., CaL, SK, BK) is unknown in TC neurons, and given the lack of data for the axon initial segment (AIS) of TC neurons, we modeled their distributions on data from pyramidal cells (Bowden et al., 2001; Hu et al., 2009; Battefeld et al., 2014):
As CaL and the Ca2+-dependent conductance densities (i.e., SK) are co-localized in pyramidal cells, we used the same subcellular distributions for SK and BK, while these channels were absent in the axon.
NaT and NaP conductance densities can be up to 19-fold higher than in the soma in the proximal and distal halves of the AIS, respectively.
KM conductance density in the distal half of the AIS can be up to 50 times higher than in the soma.
In the absence of specific information, we made the following assumptions:
H-type conductance density is uniform along the somatodendritic arbor but absent in the AIS.
KDR conductance density can be up to 5-fold higher in the proximal half of the AIS than in the soma to compensate for the effects of the high NaT conductance density, thus enabling full action potential repolarization.
Template data taken from in vitro experiments
We selected a subset of whole-cell recordings obtained from a previous study in vitro (Bichler et al., 2021), performed on adult mice of either sex in normal conditions and after unilateral 6-OHDA injections placed in the median forebrain bundle (Control: n = 9; 6-OHDA: n = 7). These recordings were obtained with different experimental protocols:
Protocol 1: 2-s-long depolarizing current steps (20–300 pA with 20-pA increments) on top of a bias current that held the membrane potential at −69 mV (Control: 93 pA; 6-OHDA: 67 pA), to estimate f-I curves.
Protocol 2: 2-s-long hyperpolarizing current steps (from –200 to –50 pA with 50-pA increments) on top of a bias current (same as in protocol 1), to estimate the rebound spike count versus current relationship.
Protocol 3: 2-s-long hyperpolarizing steps (–200 to –50 pA with 50-pA increments; without bias current), to measure sag amplitude and evoke rebound bursting versus current relationships.
Protocol 4: single pulse of –1 nA and 0.5 ms of duration, to measure membrane time constant.
Protocol 5: 100-ms-long current step of –10 pA, to estimate input resistance.
To measure firing properties from simulation and experimental traces, we used a custom version of the eFEL package (https://github.com/BlueBrain/eFEL). In particular, the targets for model performance were constituted by membrane and firing properties extracted from our experimental recordings obtained with protocols (see Table 2 for a complete list).
Fitting neuron models
We define a TC neuron model comprising 33 free parameters, including the conductance densities along with the decay time constants of intracellular Ca2+ concentration (see above, Intrinsic membrane properties). We also added voltage shift variables and multiplicative factors as free parameters to the equations describing the dynamics of NaT, KDR, KM, and BK currents, with different effects on the firing properties generated by our models. Specifically, we added:
Multiplicative factors to the equations of the rates for activation and inactivation of NaT and KDR, fixing the ranges in a way that slowed down their dynamics, and thus regulated the AP width.
A multiplicative factor to the equation of the rates for the activation of KM, fixing the ranges in a way that slows down the de-activation, and thus calibrated the contribution to firing rate adaptation.
A voltage shift variable to the equations of activation and inactivation, along with their rates, for KD, fixing the range in a way that increased the impact of the channel on the first-spike latency, yielding half-values of activation and inactivation more similar to the values observed in pyramidal cells (Storm, 1988).
Multiplicative factors and shift variables to the equations describing activation and inactivation, along with their time constants, for BK, mimicking the effects of β subunits expression on the channel dynamics (Behrens et al., 2000; Contreras et al., 2012), to modulate the impact of this channel on firing rate adaptation.
The values of the free parameters were determined by multiobjective optimization, using the BluePyOpt toolkit (Van Geit et al., 2016) to fit model traces to a set of target features (Table 2). These features described membrane potential dynamics (e.g., action potential amplitudes, after-hyperpolarization depth, spike count, firing adaptation index, and so on) in response to current injection paradigms and were extracted from our published whole-cell recordings in normal and parkinsonian states (Bichler et al., 2021).
Model candidates in normal and parkinsonian conditions were obtained with 45 and 15 optimization sessions, respectively, using a different random seed for each session, with 100 individuals and 100 generations per session. Technically, the optimizer minimized the error associated with each model as measured by the difference between electrophysiological features from simulations and experimental traces. The overall fitness of a given model resulted from the sum of the absolute errors associated with feature differences (passive features and active features; Table 2), each calculated as the deviation from the experimental mean normalized to the experimental standard deviation. To measure the firing properties of each model, the optimizer ran a battery of simulations reproducing experimental traces (Table 2). For final model selection, we first made a hall-of-fame as the population of models for which all errors fell within three standard deviations from mean (Control: n = 686; 6-OHDA: n = 4900). Of these models, we retained the ones that passed additional quality checks. First, we compared the entire f-I curves, obtained with protocol 1, along with the rebound spike count versus current relationships, obtained with protocols 2 and 3, retaining the hall-of-fames that yielded spike counts below three standard deviations from the experimental mean throughout the entire range of stimulation (Control: n = 78; 6-OHDA: n = 922). Second, we selected the models that replicated the effects of XE991 application (10–20 μm; Bichler et al., 2021), simulated by decreasing the M-type conductance density by 70% (Yeung and Greenwood, 2005). In the experiments, XE991 application in normal condition decreased the rheobase by ∼80 pA and shifted the f-I curves to the left, while the application did not alter the response in parkinsonian condition. For each model, we then compared the f-I curves with and without XE991, retaining normal and parkinsonian models generating distinguishable (Wilcoxon, p < 0.1; with shift in rheobase < 100 pA) and indistinguishable (Wilcoxon, p ≥ 0.9; with no shift in rheobase) curves (Control: n = 39; 6-OHDA: n = 21). Third, by inspecting traces, we rejected models that generated a nonphysiological after-hyperpolarization, with peaks below baseline (Control: n = 28; 6-OHDA: n = 19). Fourth, we tested the models with realistic synaptic inputs (see below, Modeling synaptic inputs), retaining the models that generated firing rates between 5 and 90 Hz (Control: n = 16; 6-OHDA: n = 17). The distributions of the parameters for normal and parkinsonian models are described in Table 3.
Modeling synaptic inputs
We modeled four groups of synaptic inputs: glutamatergic driver-like (DRI-l) inputs, glutamatergic modulators (MOD), GABAergic inputs from substantia nigra reticulata (SNR), and GABAergic inputs from reticular thalamic nuclei (RTN), each defined by postsynaptic location, numbers of terminals, and unitary conductance (Table 4), representing the presynaptic activity by artificial spike trains with realistic firing rates and irregularity (i.e., coefficient of variation of interspike interval).
Anatomical studies showed that vGluT1+ and vGluT2+ terminals in VM, which originate from layer 5 of cortex and subcortical regions, respectively, have small to medium size terminals, while large size terminals were not observed (Rovó et al., 2012), suggesting that the big excitatory driver synapses observed in other thalamic nuclei may be absent in VM. On the other hand, cerebellar terminals display “driver-like” physiology, whose activation evoked smaller EPSCs than classic drivers (Gornati et al., 2018). Therefore, we stayed with identical properties for all driver-like inputs to VM, with parameters matching cerebellothalamic terminals in VM.
Specifically, excitatory synapses comprised AMPA and NMDA components (Cavarretta et al., 2018), with a reversal potential of 0 mV, and NMDA/AMPA ratios of 0.6 for drivers and 1.91 for modulators as estimated from ventrobasal thalamus (Miyata and Imoto, 2006). For inhibitory synapses, we set the decay time constant to 14 ms (at 32°C; Edgerton and Jaeger, 2014), reversal potential to –81 mV (Ulrich and Huguenard, 1997), and Q10 to 2.1 (Otis and Mody, 1992). For each group of synapses, we determined subcellular distributions and numbers of contacts. We fixed the proportion of DRI-l in relation to MOD terminals to 10%, consistent with previous anatomic studies on drivers and modulators (Van Horn et al., 2000; Van Horn and Sherman, 2007). The DRI-l terminals were located at a distance from the soma that reproduced the distribution observed for deep cerebellar nuclei terminals in VM (Gornati et al., 2018). For the other synaptic inputs, we estimated the densities of synaptic terminals along the circumference of single dendritic sections obtained by electron microscopy (EM) of mouse VM (n = 2; Y. Smith, unpublished data). Specifically, we analyzed EM pictures of transversal dendrite profiles, obtained with magnifications between 20,000 and 40,000. The dendrites were then divided in three brackets based on cross-section diameters, namely small (0–0.5 μm; n = 222), medium (0.5–1 μm; n = 220), and large (>1 μm; n = 25). According to previous studies (Swain et al., 2020), synaptic terminals were categorized as asymmetric (AS) terminals, originating from MC, type-1 symmetric (S1) terminals, originating from RTN, and type-2 symmetric (S2) terminals, originating from SNR. Only terminals that showed an active zone at the level of a given EM section were counted. From these counts, we calculated the density of each type of terminal as density of terminals around the circumference of the dendritic profiles. We then assumed that independent samples of terminals along the dendrite would be obtained at distances equal to the average diameter of terminals, which is ∼0.8 μm for MOD and RTN terminals (Y. Smith, unpublished data), and 2.8 μm for SNR terminals (Bodor et al., 2008). Thus, the cross-sectional density of synapses around the circumference of dendrites was applied for every 0.8 μm (MOD and RTN) or 2.8 μm (SNR) of dendritic length in the model. Finally, we added SNR terminals contacting the TC neuron soma as a proportion of the dendritic ones (Bodor et al., 2008).
After defining the subcellular distributions, we estimated the unitary synaptic conductance for each group of synapses. To this end, we designed an optimizer that determined the optimal values matching the postsynaptic current amplitudes observed experimentally (Table 4, experiment). The optimizer explored the range of conductance values by using a convergence algorithm, testing each value by running simulations that reproduced the voltage-clamp experiments (Table 4, simulation).
We did not model short-term plasticity (STP) for any synapses, as an experimental characterization of STP parameters in VM is lacking. Moreover, a previous modeling work suggested that the main effect of short-term depression in a simulation with steady state rates of input is equivalent to a downscaled unitary conductance (Abbasi et al., 2017). We thus reduced the unitary conductances to approximate the effects of short-term depression in the DRI-l inputs (Gornati et al., 2018). Specifically, we considered an EPSC amplitude reduced by 25% as a target for estimating the conductance peak of the DRI-l synapses (see Table 4, experiment) so accounting for the paired-pulse ratio observed experimentally (Gornati et al., 2018).
By applying the subcellular distributions of synaptic terminals described above to the morphology (AA0719; MouseLight Archive), we obtained estimates of the number of terminals for each group of synapses (MOD: 3625; DRI-l: 350; SNR: 25; RTN: 400). In our simulations, we decreased the MOD terminals to 1450, i.e., ∼40% of the estimated value, accounting for the proportion of silent and nonsilent pyramidal neurons observed in layer 6 of MC of cats (Sirota et al., 2005). This configuration yielded a basal firing rate of 18.6 ± 9.4 Hz (±SD) in normal conditions, consistent with the values observed experimentally (Inagaki et al., 2022).
Presynaptic activity: artificial spike train generation
To represent presynaptic activity, we generated artificial spike trains (Abbasi et al., 2020), with firing rates consistent with experimental estimates (DRI-l: 30 Hz; MOD: 1.1 Hz; RTN: 10 Hz; SNR: 50 Hz; Sirota et al., 2005; Huh and Cho, 2016; Barrientos et al., 2019; Inagaki et al., 2022). Specifically, we used an algorithm that generates a random sequence of interspike intervals picked from a γ distribution with a refractory period and “shape” parameter set a priori (Abbasi et al., 2020). We targeted generic spike properties with a coefficient of variation of interspike intervals (CVISI) of 0.45 (as observed in SNR during in vivo recordings; Lobb and Jaeger, 2015), obtained with a shape of 5 (
Extended Data Table 3-1
Equations of active membrane properties. Bold literals indicate the free parameters (see Table 3). Download Figure 3-1, TIF file.
Extended Data Figure 6-1
β Modulation of different synaptic inputs individually. Related to Figure 6. Phase plots of the spiking activity for normal (black) and parkinsonian (red) models of thalamocortical (TC) neurons in presence of β modulation in the firing activity of substantia nigra reticulata (SNR), excitatory modulators (MOD), excitatory driver-like (DRI-l) inputs, and reticular thalamic nuclei (RTN). For each group of synaptic inputs, β modulation induces significant spike-phase locking in the activity of TC neuron models in both states (Rayleigh, p < 0.001). Each phase plot was obtained by averaging the responses of different TC neuron models (Control: n = 16; 6-OHDA: n = 17) and multiple simulations per model (n = 10). Compared to MOD and RTN inputs (128.1–154.9°), the circular SD of the spiking activity in TC neuron models achieved the lowest values with β modulation in SNR and DRI-l inputs (93.4–107.4°), suggesting that SNR and DRI-l inputs can induce the strongest spike-phase locking in TC neurons. Download Figure 6-1, TIF file.
Extended Data Figure 7-1
β Modulation of inhibitory inputs from substantia nigra reticulata and excitatory driver-like inputs. β Modulation in the activity of substantia nigra reticulata (SNR) and driver-like inputs (DRI-l) at 12.5 Hz (left) and 25 Hz (right). Modulators (MOD) were not modulated. The oscillations in DRI-l were shifted by 180° with respect to SNR inputs. The effects of β modulation in SNR only are shown in Figure 7. A, Spike histograms show the relative variation of instantaneous firing rate to average versus time (top) and exemplificative raster plots (bottom) of the presynaptic activity for MOD (magenta), DRI-l (orange), and SNR (blue). The spike histograms were calculated from multiple simulations (n = 10). The raster plots of each input were a subset of total (n = 25). B, Somatic voltage traces of thalamocortical (TC) neuron models in normal (Control; black) and parkinsonian (6-OHDA) states, with regular (red) and increased (+20%; cyan) SNR conductance for parkinsonian models. Each curve was obtained by averaging the voltage traces generated by models (Control: n = 16; 6-OHDA: n = 17) and multiple simulations per model (n = 10). C, Spike histograms showing the relative variation of instantaneous firing rate to average versus time (top) and exemplificative rastergrams (bottom) of TC neuron activity in normal (Control; black) and parkinsonian (6-OHDA) states, with regular (red) and increased (+20%; cyan) SNR conductance for parkinsonian models. The spike histograms were calculated from the pooled responses of multiple models (Control: n = 16; 6-OHDA: n = 17) and multiple simulations per model (n = 10). Each trial was associated with a random seed that determines the subcellular distributions and activation timing (i.e., artificial spike trains) of synapses (for details, see Materials and Methods). The rastergrams show the activity of TC neuron models (n = 15) in normal (black) and parkinsonian states (red, cyan), obtained with a single simulation. D, Phase plots of the spiking activity shown in C (spike histograms) for normal and parkinsonian models of TC neurons (Rayleigh, p < 0.001). Download Figure 7-1, TIF file.
Extended Data Figure 7-2
β Modulation of inhibitory inputs from substantia nigra reticulata and excitatory modulators. β Modulation in the activity of substantia nigra reticulata (SNR) and excitatory modulators (MOD) at 12.5 Hz (left) and 25 Hz (right). Driver-like inputs (DRI-l) were not modulated. The oscillations in MOD were shifted by 180° with respect to SNR inputs. The effects of β modulation in SNR only are shown in Figure 7. A, Spike histograms show the relative variation of instantaneous firing rate to average versus time (top) and exemplificative raster plots (bottom) of the presynaptic activity for MOD (magenta), DRI-l (orange), and SNR (blue). The spike histograms were calculated from multiple simulations (n = 10). The raster plots of each input were a subset of total (n = 25). B, Somatic voltage traces of thalamocortical (TC) neuron models in normal (Control; black) and parkinsonian (6-OHDA) states, with regular (red) and increased (+20%; cyan) SNR conductance for parkinsonian models. Each curve was obtained by averaging the voltage traces generated by models (Control: n = 16; 6-OHDA: n = 17) and multiple simulations per model (n = 10). C, Spike histograms showing the relative variation of instantaneous firing rate to average versus time (top) and exemplificative rastergrams (bottom) of TC neuron activity in normal (Control; black) and parkinsonian (6-OHDA) states, with regular (red) and increased (+20%; cyan) SNR conductance for parkinsonian models. The spike histograms were calculated from the pooled responses of multiple models (Control: n = 16; 6-OHDA: n = 17) and multiple simulations per model (n = 10). Each trial was associated with a random seed that determines the subcellular distributions and activation timing (i.e., artificial spike trains) of synapses (for details, see Materials and Methods). The rastergrams show the activity of TC neuron models (n = 15) in normal (black) and parkinsonian states (red, cyan), obtained with a single simulation. D, Phase plots of the spiking activity shown in C (spike histograms) for normal and parkinsonian models of TC neurons (Rayleigh, p < 0.001). Download Figure 7-2, TIF file.
Extended Data Figure 7-3
β Modulation of inhibitory inputs from substantia nigra reticulata, excitatory modulators and driver-like inputs, and reticular thalamic nuclei. A, Phase plots of the spiking activity shown in Figure 7 for normal (black) and parkinsonian (red) models of thalamocortical neurons (Rayleigh, p < 0.001). The spike histograms were calculated from the pooled responses of different models (Control: n = 16; 6-OHDA: n = 17) and multiple simulations per model (n = 10). B, Same as in A with β modulation of reticular inputs in phase with inputs from substantia nigra reticulata (SNR; Rayleigh, p < 0.001). C, Same as in A with β modulation of reticular inputs shifted by 180° with respect to SNR inputs (Rayleigh, p < 0.001). Download Figure 7-3, TIF file.
Extended Data Figure 8-1
Synaptic and membrane currents resulting from β modulation of synaptic inputs at 25 Hz. Same simulations as shown in Figure 8 with β modulation at 25 Hz, using the same configuration of inputs shown also in Figure 7, right. Download Figure 8-1, TIF file.
We enhanced the published method to generate bursts, treated as a special case of event in each spike train. Specifically, each burst sequence was generated separately and subsequently merged with an existing baseline firing spike train. A template described the time course of the intraburst firing rate, which can be manipulated to fit different accelerating or decelerating patterns. Specifically, we used different values of shape for the γ distributions of interspike intervals for regular spike trains (=5) and bursts (=3), which yielded bell shaped and more symmetric distributions for the former, and highly right skewed ones for the latter (Selinger et al., 2007). Furthermore, interburst intervals were governed by a γ distribution, associated with mean and regularity parameters that correlates with average (positively) and variance (negatively) of the interburst intervals, respectively.
We then generated bursting activity for SNR inputs (Fig. 5), as artificial spike trains of ∼150 ms in duration, with an average firing rate of ∼170 Hz and a refractory period of 1.5 ms. For the interburst activity, we decreased the baseline firing rate to 35 Hz and increased the regularity to 50. The interburst intervals were 1 s on average, with a minimum value of 150 ms and regularity values of 5 and 50,000 for spike trains with asynchronous and synchronous bursting, respectively. Overall, this configuration yielded artificial spike trains with firing rate of 55.7 ± 1.6 Hz (±SD) and coefficient of variation of interspike interval of 0.69 ± 0.02 (±SD), consistent with previous experimental recordings from SNR (Lobb and Jaeger, 2015).
Data analysis
We used a custom version of the eFEL python package for feature extraction from membrane potential traces (https://efel.readthedocs.io). The junction potential of the intracellular medium was estimated using JPcalc software (Barry, 1994). For each simulated experiment, reversal potentials of ion species were estimated by using the Nernst equation, accounting for the composition of the aCSF (see Table 1). For statistical analyses, we used Python3 with the SciPy and NumPy packages. To compare f-I curves (Fig. 2), we used the Wilcoxon matched-pairs signed rank test. To compare different sets of measures, we used Mann–Whitney U tests. To assess the significance of the Pearson correlation (Fig. 3), we used the Kowalski test. To assess the significance of spike-phase locking (Figs. 6, 7), we used the Rayleigh test on polar data. The confidence intervals were presented as mean ± SD, mean ± standard SE. We preferred the SD for the estimates of firing rates, the coefficient of variations for the interspike intervals, and the distributions of distance from soma and diameters of the postsynaptic dendrites (Fig. 3), to provide an immediate measure of the spread around the mean, while we used the SE in all the other cases. To smooth the curves of net current shown in Figure 8 and Extended Data Figure 8-1D, we applied a first-order filtering of Savitzky–Golay with a kernel of 101 points in size.
Results
The effects of unilateral 6-OHDA lesions on VM neural properties were previously recorded in vitro in mice (Bichler et al., 2021). The results showed that the excitability of TC neurons was increased in the dopamine-depleted hemisphere. This effect was attributed to M-type potassium current suppression. To understand how these excitability changes may impact synaptic integration in vivo, we constructed biophysically detailed VM TC neuron models in normal and 6-OHDA depleted conditions (Fig. 1; for details, see Materials and Methods, Fitting neuron models), using the NEURON simulator (Hines and Carnevale, 1997). We modeled the main afferent inputs to VM, i.e., excitatory modulators (MOD), excitatory driver-like inputs (DRI-l), inhibitory inputs from substantia nigra reticulata (SNR), and inhibitory inputs from reticular thalamic nuclei (RTN), to simulate in vivo-like conditions (Fig. 1; for details, see Simulation of in vivo-like conditions). This approach allowed us to assess the interaction of intrinsic thalamic neural dynamics with synaptic input properties expected in normal or parkinsonian waking conditions through measurements in the model that cannot be performed experimentally in vivo.
Simulation of in vitro experiments
Physiologically, VM TC neurons display the characteristic firing modes, e.g., tonic firing, low-threshold spiking, and rebound bursting (Edgerton and Jaeger, 2014; Bichler et al., 2021), commonly observed in other thalamic nuclei (Llinás and Steriade, 2006), while dopamine depletion increases tonic firing and prolongs rebound bursting in these neurons (Bichler et al., 2021). Here, we tested the ability of the VM TC neuron models in reproducing in vitro recordings and the changes following dopamine depletion, comparing simulation and experimental responses.
Specifically, we considered whole-cell recordings from slices of adult mice VM in normal conditions and after unilateral 6-OHDA application (Control: n = 9; 6-OHDA: n = 7), stimulated by depolarizing (20–300 pA; 2 s of duration; Fig. 2A,B) or hyperpolarizing (from –50 pA to –200 pA; 2 s of duration; Fig. 2C,D) current steps on top of a bias current (Control: 93.3 ± 18.7; 6-OHDA: 67.1 ± 12.7; ±SE). In particular, the bias current held the membrane potential to –69 mV, inactivating T-type Ca2+ channels, and thus enabled the generation of tonic firing in response to depolarizing steps (Fig. 2A, left), while hyperpolarizing steps evoked rebound bursting (Fig. 2C, left). 6-OHDA enhanced the excitability of TC neurons, increasing responses of TC neurons in tonic firing (Fig. 2A, left; compare black and red traces), assessed as a shift of the f-I curves to the left (Wilcoxon, p < 0.01; Fig. 2B, left; compare black and red curves), and prolonged rebound bursting (Fig. 2A, left, compare black and red traces), assessed as an increase in the rebound spike count (Mann–Whitney, 20–40 mV: p = 0.012; 40–60 mV: p < 0.01; Fig. 2D, left, compare black and red bars for each bracket).
To allow a direct comparison between the two sets of traces, the simulations reproduced the stimulation protocols (e.g., step and bias current amplitudes) and the environmental conditions of slices (e.g., temperature, reversal potential of ion species, holding membrane potential; Table 1). We thus observed that the models (Control: n = 16; 6-OHDA: n = 17) replicated the responses in tonic and burst firing modes observed experimentally, in response to depolarizing and hyperpolarizing steps, respectively, with realistic levels of firing rate adaptation and action potential properties (e.g., AP height, AHP depth, firing rate adaption). Both experimental and simulation traces displayed variability in multiple firing properties (Extended Data Figs. 2-1 and 2-2). Compared with normal models, parkinsonian models displayed stronger excitability, generating more action potentials in response to depolarizing and hyperpolarizing current steps (Fig. 2A,C, right, compare black and red traces). Parkinsonian models generated significantly higher spike rates throughout the entire range of stimulation using depolarizing currents (Wilcoxon, p < 0.01; Fig. 2B, right, compare black and red curves), while the spike rates were comparable to the values observed experimentally for both normal and parkinsonian models (Fig. 2B, compare experimental and simulation f-I curves in left and right panels, respectively). Likewise, rebound spike counts were statistically different in normal and parkinsonian models (Mann–Whitney, 20–40 mV: p < 0.01; 40–60 mV: p < 0.01; Fig. 2D, compare black and red bars for each bracket) but comparable to the corresponding experimental values (Fig. 2D, compare experimental and simulation spike counts for each bracket in left and right panels, respectively).
The models discussed above resulted from a selection process comprising several quality checks. One check concerned the ability of the models in replicating the effects of XE-991 application (10–20 μm; Bichler et al., 2021), simulated by reducing the M-type conductance density by 70% (Yeung and Greenwood, 2005), which shifted the f-I curves to the left significantly for normal but not parkinsonian models (not shown), in accord with our findings in vitro (Bichler et al., 2021). This yielded parkinsonian models with significantly lower expression of M-type potassium current than control (Mann–Whitney, p < 0.01; Control: 306.76 ± 29.24 μS/cm2; 6-OHDA: 9.69 ± 2.46 μS/cm2; ±SE; Table 3). Increasing M-type potassium current in parkinsonian models by ∼40 times rescued the effects of dopamine depletion (not shown), making the f-I curves of parkinsonian models indistinguishable from control (Wilcoxon, p = 0.30), and reducing their rebound spike count to similar values as observed for control (Mann–Whitney, 20–40 mV: p = 0.026; 40–60 mV: p = 0.013). Taken together, these simulations demonstrated that our models replicate the firing behaviors of VM TC neurons observed in vitro along with the alterations caused by dopamine depletion.
Simulation of in vivo-like conditions
After fitting the TC neuron models to in vitro recordings, we modeled the main afferent inputs to VM (Fig. 3): (1) modulators (MOD), approximating the glutamatergic inputs conveyed from layer 6 of MC; (2) driver-like inputs (DRI-l), approximating glutamatergic inputs from layer 5 of MC and motor-related subcortical areas; and GABAergic inputs from (3) substantia nigra reticulata (SNR) and (4) reticular thalamic nucleus (RTN).
For each group of synapses, we constrained synaptic physiology, subcellular distributions, number of terminals, and individual conductance to experimental data (for details, see Materials and Methods, Modeling synaptic inputs). In particular, the distribution of DRI-l terminals reproduced the distance from soma observed for deep cerebellar nuclei terminals in VM (Fig. 3A,B; Gornati et al., 2018), while the distributions for MOD, RTN, and SNR were modeled on our electron microscopy data (Fig. 3A,B; for details, see Materials and Methods, Modeling synaptic inputs). As an emergent property, we obtained a co-localization along the somato-dendritic arbor of VM TC neuron models (Fig. 3B) for DRI-l and SNR terminals, targeting proximal dendrites with larger diameters, as well as MOD and RTN terminals, targeting distal dendrites with small diameters, suggesting that a stronger interaction takes place in VM TC neurons between each pair of co-localized terminals.
We then tested the impact of these synaptic inputs on normal and parkinsonian TC neuron models (Fig. 3C). We represented the presynaptic activity with artificial spike trains replicating firing rates and irregularity observed experimentally (Fig. 3Ca; for details, see Materials and Methods, Modeling synaptic inputs). With these configurations of synaptic inputs, we observed that the highest values of total synaptic conductance over time were achieved by SNR and DRI-l inputs (Fig. 3Cb), suggesting that these two classes of synaptic inputs sustained the TC neuron firing in normal and parkinsonian states (see also below, Sensitivity analysis of synaptic response function). Although the two populations of models received the same synaptic inputs, they generated different firing responses (Fig. 3Cc). Compared with control, parkinsonian models displayed increased firing rates (Mann–Whitney, p < 0.0001; Control: 18.6 ± 9.4 Hz; 6-OHDA: 33.0 ± 17.3 Hz; ±SD) and baseline voltage (Mann–Whitney, p < 0.0001; Control: −51.4 ± 1.2 mV; 6-OHDA: −48.4 ± 1.3 mV; ±SD). Therefore, the models suggest that the increased excitability observed in vitro for parkinsonian states (Bichler et al., 2021) resulted in a strong firing increase with a balance of excitatory and inhibitory synaptic inputs as well.
Sensitivity analysis of synaptic response function
To establish the importance of each synaptic afferent, we performed a sensitivity analysis of their parameters, i.e., numbers of synapses (nsyn), unitary conductance (gsyn), and presynaptic firing rate. For each group of synapses, we plotted the thalamic firing rate for normal and parkinsonian models as a function of each parameter, varied by ±10% from control (Fig. 4A). For each parameter, we performed a linear regression between the percent of variation and the thalamic firing rate in normal and parkinsonian states (Fig. 4A, black and red lines), and used the absolute value of the Pearson correlation between thalamic firing rate and the varying parameter (Fig. 4B).
In normal and parkinsonian states, increasing synapse number, conductance, or firing rate for RTN and SNR inputs decreased firing rates of the TC neuron models (i.e., negative correlation), while increasing these parameters for MOD and DRI-l inputs resulted in the opposite effect (i.e., positive correlation; Fig. 4A). The fitted lines for each parameter were steeper for DRI-l and SNR inputs, whereas they appeared flat for MOD and RTN inputs (Fig. 4A), and the firing rate was higher in the parkinsonian state than the control state throughout the entire range of variation. By comparing the absolute values of correlation, we observed strong (0.38–0.50; Fig. 4B) and significant correlations for all the parameters related to DRI-l and SNR inputs (Kowalski test, p < 0.0001), whereas weak correlations were observed for MOD and RTN inputs (0.08–0.27; Fig. 4B), insignificant for all the parameters related to MOD and RTN inputs, except the number of RTN synapses, which however displayed a weaker significance than DRI-l and SNR inputs (Kowalski test, Control: p = 0.015; 6-OHDA: p = 0.014). Taken together, these simulations suggest that, under these configurations of afferent inputs, VM output in vivo is primarily driven by DRI-l and SNR but not MOD and RTN inputs, without any relevant difference between normal and parkinsonian states. This is consistent with the observation that synaptic inputs from SNR and DRI-l yielded the highest values of total synaptic conductance over time (Fig. 3). Additionally, synaptic inputs evoked stronger responses in parkinsonian states throughout the entire range of parameters, suggesting a robust effect of dopamine depletion on VM TC neuron excitability.
Effects of synchronous bursting inputs from substantia nigra reticulata
Previous speculations have described thalamic rebound bursting as a key mechanism of parkinsonian pathophysiology (Rubin and Terman, 2004; Meijer et al., 2011), and experiments showed that the occurrence of rebound bursting in ventrolateral thalamus coincides with the emergence of motor deficits (Kim et al., 2017). In vitro experiments showed that synchronous activation of nigral axons could evoke rebound bursting in VM, where dopamine depletion enhanced rebound bursting evoked by hyperpolarizing current steps (Bichler et al., 2021). Additionally, in vivo experiments showed that dopamine depletion increased spike synchrony and bursting in SNR (Wang et al., 2010; Anderson et al., 2015; Willard et al., 2019), an ideal mechanism for evoking rebound spiking in VM. While these findings support the possibility that rebound spiking in VM might contribute to parkinsonian dynamics, this hypothesis has not yet been directly tested in behaving animals.
To investigate the effects of synchronous nigral bursting on VM output, we generated trains of nigral bursts reproducing the rates and duration observed in our experimental recordings in dopamine-depleted mice (for details, see Materials and Methods, Presynaptic activity: artificial spike train generation). We observed that synchronous SNR bursts hyperpolarized the membrane of TC neuron models (Fig. 5A,B), blocking firing activity in normal (black) and parkinsonian (red) states, followed by a temporary firing rate increase at the offset of the nigral bursting. Comparing the thalamic activity evoked by SNR inputs with (Fig. 5B, black, red) and without bursts (Fig. 5B, gray, light red), we observed that the postinhibitory increase in firing rate occurred robustly at the offset of the SNR bursting, without affecting the firing activity over different time windows (Fig. 5B, left). Instead, asynchronous SNR bursting decreased the average firing rates throughout the entire course of firing activity (Fig. 5B, right). By comparing the poststimulus spike time histograms with (black, red) and without (gray, light red) synchronous nigral bursting (Fig. 5B), we observed that SNR bursting evoked a significant postinhibitory firing rate increase in most TC neuron models (Control: 16/16; 6-OHDA: 9/17; Fig. 5C), with different courses in normal and parkinsonian states, reaching a peak instantaneous firing rate of ∼80 and ∼110 Hz, respectively, and lasting up to 120 and 460 ms, respectively. Both the firing rate peak and the duration of the postinhibitory firing rate increase positively correlated with the percentage of synchronous SNR inputs (Fig. 5C), maintaining different peaks and durations in the two states (Fig. 5C,D). Therefore, our simulations suggest that synchronous SNR bursting evokes postinhibitory spiking activity in motor thalamus with different courses in normal and parkinsonian states, and that the intensity of this effect correlates with the percentage of synchronous SNR inputs.
To seek a mechanistic explanation for the difference between the postinhibitory responses observed in normal and parkinsonian models, we analyzed the variation in transmembrane current density induced by each HH-conductance over time (Fig. 5D). Surprisingly, the postinhibitory increase in firing rate following synchronous SNR bursting was not sustained by T-type Ca2+ current, as the hyperpolarization was not sufficient to de-inactivate these channels (Fig. 5D, brown lines). Instead, the early increase at the offset of the hyperpolarization coincided with a state of incomplete recovery for M-type potassium current (Fig. 5A, green lines), observed with both normal and parkinsonian models. Additionally, with parkinsonian models only, the incomplete recovery of BK potassium current, which maintained the current lower than baseline by ∼0.03–0.07 μS/cm2 for ∼800 ms, contributed to both early and late parts of postinhibitory firing rate increase (Fig. 5D, bottom, purple line). Therefore, these predictions suggest that synchronous SNR bursting in vivo cannot de-inactivate T-type Ca2+ channels enough to evoke rebound bursting in TC neurons, while it causes a postinhibitory increase in firing activity, due to the slow recovery of M-type potassium current for models in both states, along with the slow recovery of the BK potassium current in parkinsonian models only.
Effects of β modulation in the synaptic afferents to ventromedial thalamus
Another hallmark of parkinsonian pathophysiology is the exaggeration of β oscillations in VM, observed by local field potential (LFP) recordings (Brown et al., 2001; Cassidy et al., 2002; Levy et al., 2002; Kühn et al., 2005; Sharott et al., 2018), with increased spike-LFP coherence in VM, SNR, and MC (Brazhnik et al., 2012, 2016; Nakamura et al., 2021). Therefore, it has been hypothesized that oscillations might originate in SNR and then entrain VM, propagating globally throughout the entire motor cortico-thalamocortical loop. We tested this hypothesis by introducing fluctuations in the firing activity of SNR within the β band, i.e., at 12.5 and 25 Hz (Fig. 6A, blue curve and raster), with amplitude ±10% from the baseline. We observed that somatic membrane potential (Fig. 6B) and instantaneous firing rate (Fig. 6C) reflected the fluctuations in SNR, with significant spike-phase locking (Rayleigh, p < 0.001; Fig. 6D). Compared with SNR, the oscillations in firing rate of VM TC neuron models achieved higher amplitudes, with down and up phases of –90% and +148% on average, respectively, without any noticeable difference between normal and parkinsonian states, suggesting that VM does not only relay but also amplifies the β oscillations in SNR in either state.
Moreover, canonical models of PD suggest that excessive nigrothalamic inhibition, due to increased SNR firing rate, might impede thalamocortical transmission of motor signals, preventing the regular course of movements (DeLong, 1990). While our simulations suggested that increased firing rate and/or synaptic conductance of SNR inputs could be two ways of enhancing nigrothalamic inhibition (Fig. 3), studies on rodents led to inconsistent conclusions about the changes in SNR firing rate after dopamine depletion, with some experiments showing no changes or even a decrease in firing rate across different sets of SNR neurons (Díaz et al., 2003; Wang et al., 2010; Anderson et al., 2015; Lobb and Jaeger, 2015; Willard et al., 2019). Therefore, these findings suggest the possibility that enhanced SNR inhibition could be a consequence of increased synaptic conductance rather than firing rate. We tested this hypothesis by increasing unitary synaptic conductance by 20% for SNR on parkinsonian models, yielding a basal firing rate indistinguishable from control (Mann–Whitney, p = 0.41; 17.9 ± 10.1 vs 18.6 ± 9.4 Hz; ±SD), consistent with previous observations from basal ganglia receiving nuclei of motor thalamus (Anderson et al., 2015; Di Giovanni et al., 2020; Nakamura et al., 2021). The increase did not affect the oscillatory firing patterns in the VM (Fig. 6B–D, compare black, red, and cyan curves), suggesting that enhanced nigrothalamic inhibition by itself did not increase spike-phase locking.
Additionally, significant spike-phase locking was induced by β-modulating each synaptic input individually (Rayleigh, p < 0.001; Extended Data Fig. 6-1). To compare the levels of spike-phase locking in VM with β oscillations of each synaptic input, we calculated the circular standard deviation of spike times. In this analysis, the stronger spike-phase locking is revealed by a lower circular standard deviation of spike times. We thus observed that the circular standard deviation was lower with β modulation of SNR and DRI-l inputs (93.4−107.4°) than MOD and RTN inputs (128.1−154.9°), suggesting that β oscillations in VM are primarily conveyed from SNR and DRI-l inputs.
LFP recordings revealed that dopamine depletion increased β oscillations in layer 5/6 of MC (Brazhnik et al., 2012) and the spike-phase locking within SNR (Brazhnik et al., 2012; Nakamura et al., 2021). To study how these oscillations could entrain VM, we added β modulation in MOD and DRI-l synapses, approximating glutamatergic inputs to VM from layer 6 and 5 of MC in our simulations, respectively. Their phases were shifted by 180° with respect to SNR, consistent with previous experimental recordings (Brazhnik et al., 2012), which prevented a cancellation of inhibitory and excitatory input modulation. Adding β modulation in both the MOD and DRI-l inputs (Fig. 7A) yielded subthreshold oscillations in somatic voltage traces (Fig. 7B) as well as instantaneous firing rates (Fig. 7C) and significant spike-phase locking (Rayleigh, p < 0.001; Fig. 7D). In particular, the oscillations in firing rate displayed down and up phases of –92% and +152% on average, confirming that VM amplifies the β oscillations present in SNR, MODs, and DRI-l synaptic inputs. Compared with β modulation in SNR input alone, adding β modulation in the MOD synapses only slightly enhanced the oscillatory firing activity in VM (Extended Data Fig. 7-1), decreasing the circular standard deviation of spike times by 4.4−8.2° (Table 5). Similarly, adding β modulation only to RTN inputs had a minor impact on VM firing (Extended Data Fig. 7-3). Instead, adding β modulation in only the DRI-l inputs increased both spike-phase locking and amplification in firing rate oscillations (Extended Data Fig. 7-2). This could be assessed as a decrease in the circular standard deviation of spike times by 19.1−27.9°, which was substantially stronger than observed with β modulation in SNR and MOD inputs (Table 5, compare “SNR+DRI-l” with “SNR” and “SNR+MOD”). The increased effects on VM activity observed by adding β modulation in both MOD and DRI-l inputs were comparable with those observed by adding β modulation in DRI-l inputs only (Extended Data Fig. 7-2), suggesting that cortical inputs from layer 5 were more effective than those from layer 6 in transmitting β oscillations to VM. We observed that increasing SNR conductance in parkinsonian models slightly reduced the circular standard deviations of spike times by 6.7−8.6° (Table 5). Therefore, our simulations suggest that enhanced SNR inhibition refines the transmission of β oscillations from MOD and DRI-l inputs to VM.
Next, we addressed the mechanisms underlying enhanced β oscillations in the TC neuron models. Specifically, we pursued the same approach used in a previous publication (Jaeger et al., 1997), to determine the individual contributions of membrane and postsynaptic currents during β modulation of SNR, MOD, and DRI-l inputs at 12.5 Hz (Fig. 8; same configuration shown in Fig. 7, left) and 25 Hz (Extended Data Fig. 8-1; same configuration shown in Fig. 7, right). We observed that the total postsynaptic currents induced by each group of synapses followed the β oscillations in the presynaptic activity, with the strongest component due to DRI-l inputs (Fig. 8A). Although no β modulation took place in the RTN synapses, the postsynaptic current displayed fluctuations, similarly to the other postsynaptic currents, resulting from variations in their driving force. Changes in driving force in turn are a result of the oscillations in the voltage membrane of the TC neuron models conveyed from the other synaptic inputs, although the total conductance of RTN synapses was kept constant over time (excluding the fluctuations due to the randomness in the presynaptic spike trains; see Fig. 3Cb, cyan curve). All synaptic currents were similar in normal and parkinsonian states.
To understand the role of each voltage-gated channel type in contributing to the β modulated spike rate, we plotted the β phase modulated currents averaged over all models and 10 iterations (seeds) of each model (Fig. 8B). We find that the outward potassium currents all show increases when the neuron is more depolarized, i.e., fires action potentials at a higher frequency. This increase is due to both an increase in driving force for potassium current and an increased activation of such voltage-gated currents induced by depolarization and spiking. Note that the increase in outward current opposes the β oscillatory membrane dynamics. In contrast, the inward calcium and sodium currents are aiding the oscillatory dynamics by their increased activation during depolarization and spiking, which leads to an increased net inward current despite a reduction in driving force for sodium and calcium ions.
To characterize the relative impact of total synaptic input currents and total voltage-gated channel currents in relaying and amplifying β oscillations, we compared the net synaptic and the net membrane currents either individually (Fig. 8C, blue and green curves) or summed (Fig. 8D, teal curves), against the instantaneous firing rates of the TC neuron models (Fig. 8C,D, black and red spike histograms in the background). By determining the total synaptic and voltage-gated current separately (Fig. 8C), we observed that the net currents were outward for ion channels (blue) and inward for synapses (green), respectively, throughout the entire period of β oscillation (Fig. 8C). The β phase locked component of the total voltage-gated current opposed the β oscillations in the synaptic input (Fig. 8C), overall dampening the resulting voltage oscillation. By determining the total sum of membrane and synaptic currents (Fig. 8D, teal), we observed that the total net current displayed more inward current as the spike rate was increasing and more outward current as the spike rate was decreasing (Fig. 8D, compare teal curves in the foreground against black and red spike histograms). The phases of the up and down spike rate peaks differed from the phase peaks observed in the net synaptic and the net membrane currents individually (Fig. 8C,D; compare teal curves in D with blue and green curves in C) showing that currents precede and cause membrane voltage changes (Table 6). Note that the total currents are always inward, as we are not including the leak current, which balances the mean inward current shown here. Taken together, these simulations indicate that β phase locking in voltage and spike rate are caused by the membrane response to synaptic input current, and further that the magnitude of this phase-locking is dampened by a net counterbalancing voltage-gated current, which furthermore leads to a phase shift in the voltage oscillation.
Discussion
We used biophysically detailed simulations to investigate the consequences of dopamine depletion on synaptic integration in VM (Fig. 1). We constructed two populations of TC neuron models reproducing the firing properties observed in VM slices from control and 6-OHDA lesioned mice (Fig. 2; Bichler et al., 2021). Thus, we have developed the first biophysically detailed model of TC neuron from VM that reproduces the differing responses observed in normal conditions and after 6-OHDA treatment. This is a substantial innovation, as the previous biophysically-detailed models of TC neurons were fitted on data of other thalamic nuclei (Iavarone et al., 2019), and they did not account for any pathologic alteration of intrinsic properties. To simulate in vivo-like conditions, we modeled the main afferent inputs to VM (Fig. 3), i.e., glutamatergic MOD and DRI-l, GABAergic inputs from SNR and RTN. Our simulations showed that the increased excitability observed in vitro after 6-OHDA lesioning (Fig. 2) also resulted in a strong firing rate increase in vivo (Fig. 3). The sensitivity analysis of the models of afferent inputs suggested that VM output in vivo is primarily driven by DRI-l and SNR inputs (Fig. 4). We then systematically tested the effects of pathologic firing pattern changes observed in dopamine-depleted conditions: synchronous bursting (Fig. 5) and increased β oscillations (Figs. 6 and 7). Synchronous SNR bursting evoked postinhibitory spike rate increases in VM TC neuron models, depending on the number of synchronous SNR inputs. The time course of firing rate increase was different in models fit to parkinsonian conditions due to reduced M-type and increased BK potassium conductances (Fig. 5). β Frequency (12.5 or 25 Hz) modulations in SNR inputs were sufficient to induce spike-phase locking in VM (Fig. 6). The resulting β modulation depth of VM spiking was amplified by the VM transfer function and was stronger than that of the SNR inputs. Adding β oscillations in MOD and DRI-l inputs enhanced this effect (Fig. 7; Table 5), primarily due to the oscillations in the DRI-l inputs (Extended Data Fig. 7-1; Table 5). Therefore, VM TC neuron models relayed and amplified the oscillatory patterns in MOD, DRI-l, and SNR inputs (Figs. 6 and 7), without differences between normal and parkinsonian states. Finally, we addressed the basic mechanisms underlying these effects, observing that they resulted primarily from the high sensitivity of the model to synaptic input current without an amplification mechanism provided by voltage-gated currents (Fig. 8; Extended Data Fig. 8-1).
Effects of dopamine depletion in vitro
In vitro recordings showed that dopamine depletion increased excitability of VM as a result of M-type current suppression (Bichler et al., 2021). Comparing the parameters of normal and parkinsonian models, we determined differences in their T-type and M-type currents. In particular, the M-type conductance was 32-fold lower in parkinsonian models than control, and a subsequent 40-fold increase in M-type conductance yielded f-I curves and rebound spike counts in parkinsonian models indistinguishable from control. M-type current in thalamus is normally regulated by acetylcholine, where the activation of muscarinic receptors depolarizes thalamic neurons during high-attentional states (Llinás and Steriade, 2006). The main source of VM cholinergic input is from the pedunculopontine tegmentum and the laterodorsal tegmental nucleus (Huerta-Ocampo et al., 2020), and optogenetic stimulation of cholinergic neurons in this area switches sleep states from NREM to REM in mice (Van Dort et al., 2015). Indeed, optogenetic activation of VM neurons directly promotes arousal and wakes mice from NREM sleep (Honjoh et al., 2018). Therefore, we speculate that M-type current downregulation in VM would be causative of disturbed sleep states in PD that are characterized by insomnia, increased arousal and restless leg syndrome (for review, see Lajoie et al., 2021). In agreement with our simulations, muscarinic positive allosteric modulators that enhance M-type (Kv7) currents might have anti-parkinsonian effects on motor thalamic activity in PD and might improve associated sleep disorders.
Effects of dopamine depletion on thalamic firing rates in vivo
Our results indicated that firing rates in response to simulated in vivo-like inputs in the parkinsonian VM TC neuron models were robustly increased compared with control models due to M-type potassium current reduction. However, robust rate increases are not observed in thalamic recordings from dopamine-depleted rodents in vivo (Anderson et al., 2015; Di Giovanni et al., 2020; Nakamura et al., 2021). As classic models suggested that motor deficits in PD might be due to increased basal ganglia inhibition to motor thalamus, and that this might cause motor deficits (DeLong, 1990), we tested the hypothesis that increased inhibition in conjunction with increased excitability could lead to a normalization in firing rates. Indeed, increasing the conductance of SNR synapses by 20% yielded firing rates in parkinsonian models indistinguishable from control. Therefore, increased excitability of VM neurons after dopamine depletion might be a compensatory mechanism for increased inhibition in this state.
Our sensitivity analysis of the effect of different input sources on VM firing suggested that VM activity in vivo was primarily influenced by SNR inhibition, whereas the impact of RTN inhibition was weak (Fig. 4). This was primarily due to the smaller size and more distal location of RTN terminals that we adopted from the findings of anatomic studies (Ilinsky and Kultas‐Ilinsky, 2002; Wanaverbecq et al., 2008). This is consistent with experimental findings showing that the impact of non-nigral inhibition on VM of mouse was not substantial (Chevalier and Deniau, 1982).
Impact of nigral bursting on motor thalamic activity
In vitro studies demonstrated that synchronous SNR activation could evoke rebound bursting based on T-type Ca2+ current activation in VM (Edgerton and Jaeger, 2014). However, these experiments were conducted in the absence of background in vivo excitatory inputs that will limit the hyperpolarization achieved by nigral bursts. This effect of excitatory inputs was borne out by our simulations, where even 100% of synchronous SNR bursting in the presence of a tonic background of inhibition and excitation did not allow the development of T-type rebound bursts (Fig. 5). Instead, we found that under these conditions M-type and BK potassium currents caused a slower postinhibitory firing increase in VM neurons that was correlated with the percentage of synchronous SNR inputs (Fig. 5). Moreover, in our simulations, identical SNR inputs evoked rebound spiking with different courses in normal and parkinsonian states (Fig. 5), due to changes in M-type and BK current activation. This suggests that an altered course of postinhibitory spiking could be a marker of parkinsonian pathophysiology in VM as a consequence to excitability changes induced by dopamine depletion.
Impact of β oscillations in the afferent inputs to VM
In vivo recordings showed that dopamine depletion enhanced β oscillations in motor areas and that VM promoted their propagation to MC (Brazhnik et al., 2012, 2016). We simulated β oscillations at 12.5 or 25 Hz as a spike-phase locking in the activity of SNR input to VM, and/or cortical MOD and DRI-l inputs, consistent with the experimental recordings (Brazhnik et al., 2012, 2016). Particularly, recordings showed that β band oscillations were prevalent in nonattentive resting states whereas increased β oscillations occurred during treadmill walking (Brazhnik et al., 2012). In our simulations, low-amplitude β oscillations in the afferent inputs resulted in significant spike-phase locking in VM at both 12.5 and 25 Hz (Figs. 6 and 7), which not only relayed but also amplified the oscillatory patterns in input (Figs. 6 and 7). This amplification was not different between normal and parkinsonian conditions. Therefore, our simulation results suggest that the increased β oscillations and spike-phase locking observed in dopamine-depleted animals (Brazhnik et al., 2016; Nakamura et al., 2021) is a consequence of changes in the activity of the afferent inputs to VM and do not rely on intrinsic excitability changes in VM.
In vivo experiments showed the necessity of SNR-VM interactions for propagation of β oscillations to MC (Brazhnik et al., 2016). Our simulations showed that β oscillations in SNR alone were sufficient to induce significant spike-phase locking in VM (Fig. 6), supporting the possibility that β oscillations could originate in SNR and then entrain the cortico-thalamocortical loop through the VM. However, results in dopamine-depleted mice indicate a lack of β frequency increases in SNR (Lobb and Jaeger, 2015; Willard et al., 2019), suggesting a species-specific or state-dependent outcome in this regard.
In conclusion, modeling motor thalamic synaptic integration in normal and parkinsonian states indicates that parkinsonian firing pattern changes in the inputs are readily transmitted to the outputs. Increased intrinsic excitability may primarily act to compensate for changes in the excitatory/inhibitory input balance and lead to a homeostatic recalibration of firing rates. The increased excitability may come at the cost of accurate cholinergic modulation in motor thalamus, however, and contribute fluctuations in cortical arousal.
Acknowledgments
Acknowledgments: We thank Taylor M. Kahl for helping in proofreading the manuscript. Simulations were performed using the Neuroscience Gateway (NSG) resource and an XSEDE cluster allocation (NAT220001) granted to F.C.
Footnotes
The authors declare no competing financial interests.
This work was supported by National Institutes of Health Grants 1P50NS123103, P50-NS098685, and 1R01NS111470.
This is an open-access article distributed under the terms of the Creative Commons Attribution 4.0 International license, which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.