Skip to main content

Neuronal and synaptic adaptations underlying the benefits of deep brain stimulation for Parkinson's disease

Abstract

Deep brain stimulation (DBS) is a well-established and effective treatment for patients with advanced Parkinson's disease (PD), yet its underlying mechanisms remain enigmatic. Optogenetics, primarily conducted in animal models, provides a unique approach that allows cell type- and projection-specific modulation that mirrors the frequency-dependent stimulus effects of DBS. Opto-DBS research in animal models plays a pivotal role in unraveling the neuronal and synaptic adaptations that contribute to the efficacy of DBS in PD treatment. DBS-induced neuronal responses rely on a complex interplay between the distributions of presynaptic inputs, frequency-dependent synaptic depression, and the intrinsic excitability of postsynaptic neurons. This orchestration leads to conversion of firing patterns, enabling both antidromic and orthodromic modulation of neural circuits. Understanding these mechanisms is vital for decoding position- and programming-dependent effects of DBS. Furthermore, patterned stimulation is emerging as a promising strategy yielding long-lasting therapeutic benefits. Research on the neuronal and synaptic adaptations to DBS may pave the way for the development of more enduring and precise modulation patterns. Advanced technologies, such as adaptive DBS or directional electrodes, can also be integrated for circuit-specific neuromodulation. These insights hold the potential to greatly improve the effectiveness of DBS and advance PD treatment to new levels.

Background

Deep brain stimulation (DBS) is a promising treatment option for individuals suffering from advanced Parkinson’s disease (PD), especially those who initially respond well to dopamine replacement therapy but develop complications such as dyskinesia and "on–off" fluctuations over time [1, 2]. In the context of PD, DBS is thought to exert its effects on specific nuclei of the basal ganglia [3, 4]. The subthalamic nucleus (STN) and the globus pallidus internus (GPi) are the most common targets for DBS in PD. DBS at these two sites has demonstrated efficacy in the treatment of cardinal motor symptoms in the "off"-medicated state of PD [5,6,7,8], with an acceptable side effect profile, including weight gain, dysarthria, and mood changes [9]. In specific cases, the ventral intermediate thalamic nucleus (Vim) and the pedunculopontine nucleus (PPN) have been targeted to address tremor-dominant PD [10, 11] and gait and postural instability [12, 13]. Despite the proven efficacy of DBS in PD by numerous clinical trials and the recommendation of target selection based on clinical features [14,15,16,17,18], the common mechanisms underlying the effects remain elusive [19,20,21,22,23,24,25,26].

DBS stands as a neurosurgical intervention with the potential to leverage neural adaptations for therapeutic purposes [27]. Essentially, the therapeutic benefits and possible side effects of DBS are intricately linked to the response elicited by the stimulation of cell bodies, nerve terminals, and traversing axons (Fig. 1). This stimulation sets in motion a complex interplay of functional circuits, both during and after neuromodulation [28]. Notably, these perturbations of neuromodulation can be counteracted by homeostatic plasticity mechanisms, which aim to stabilize neuronal and circuit activity [29]. This adds a layer of complexity to the overall neuronal and synaptic adaptations brought about by DBS. A comprehensive understanding of the adaptations associated with PD-DBS is pivotal for enhancing therapeutic efficacy and laying the groundwork for future applications [30] and even for achieving long-term therapeutic benefits [31,32,33].

Fig. 1
figure 1

Differentiating mechanisms underlying the neuromodulation effects of DBS and optogenetics. Left: A hypothesis posits that single pulses of electrical stimuli activate all converging presynaptic inputs to stimulate target neurons. Responses at the target location are determined by the distribution of excitatory/inhibitory (E/I) afferent inputs [48, 49]. Repetitive high-frequency stimulation (HFS) can lead to neuronal suppression due to short-term synaptic depression [48, 50]. This results from rapid decreases in synaptic strength after brief bursts of activity, depleting presynaptic neurotransmitters [53]. Local action potentials (APs) evoked by the stimulus can propagate orthodromically to facilitate neurotransmitter release at the distal end of the soma and antidromically to activate upstream neurons [88]. Occasionally, the AP reaches the base of the axonal arbor first and then bifurcates at various branch points, eventually invading the entire axonal arbor and reaching all terminal points (Invasion). This leads to neurotransmitter release at terminal locations beyond the stimulation site [81, 87]. Right: In contrast, optogenetics (right) relies on genetically-encoded proteins that change conformation in response to a light stimulation, regulating cell activity [34]. Opsin tools expressed on membranes of specific neurons enable selective activation or inhibition of those neurons with light, leaving other non-opsin-expressing cells unaffected by the illumination [28]

Currently, optogenetics is increasingly being used in animal models to investigate the underlying mechanisms of DBS [28, 34]. Optogenetic approaches involve the introduction of light-sensitive channel proteins into target neural populations, which enables the mapping and control of neural circuits using precise light frequencies in animal models [34]. Unlike electrical stimulation, optogenetic stimulation can specifically target certain cell types, afferent and efferent projections without affecting other tissues within the stimulated area, even passing fibers (Fig. 1) [28]. In this review, we will present an overview of insights gained from opto-DBS studies regarding neural, synaptic, and circuitry aspects that explain the therapeutic benefits of DBS in PD. Our primary aim is to establish a conceptual framework for understanding the mechanisms underpinning DBS therapy.

Local neuronal adaptations to DBS

In PD, depletion of dopamine is thought to disrupt the balance of activity between the direct and indirect pathways within the striatum [35], leading to increased activity in the indirect pathway, resulting in reduced firing in the globus pallidus externus (GPe). This, in turn, elevates activity in the STN and ultimately increases thalamic inhibition via the GPi. Simultaneously, reduced firing in the direct pathway disinhibits GPi neurons, allowing them to suppress the thalamic and cortical activity. This abnormal firing pattern within the common PD targets, STN and GPi, results in excessive inhibition of the thalamus and cortex, giving rise to symptoms like bradykinesia, rigidity, and tremor. Furthermore, evidence suggests that increased activity in the cerebellothalamocortical circuit contributes to the resting tremor observed in PD [36, 37]. Consequently, DBS targeting the Vim is effective in alleviating tremors, likely through inhibiting thalamic neurons [36,37,38,39]. While it is indisputable that DBS electrode stimulation modulates the excitability of neural tissue surrounding the electrode, it is imperative to acknowledge the multifaceted and intricate local neuronal responses at the stimulation target site.

Complex effects of DBS on local neuronal activity

The question of whether DBS elicits neural excitation or inhibition during treatment remains a topic of discussion. Initial experimentation supported the "inhibition hypothesis", indicating that high-frequency electrical subthalamic and pallidal stimulation enhances PD motor symptoms by obstructing overactive ganglia output and aligns with the classical rate model [40, 41]. Nonetheless, subsequent studies have uncovered a more complicated scenario (Table 1). Some studies indicated neuronal inhibition around the electrode during high-frequency stimulation (HFS) [42,43,44], echoing the "inhibition hypothesis", while others reported increased neural activity [45]. For instance, the responses of GPi neurons to HFS-DBS differ, with some displaying facilitation, suppression, or no change in firing rate [46, 47]. In the study by Luo et al., high-frequency microstimulation of the human GPi resulted in a prolonged after-facilitation in over one-third of neurons. In addition, the neurons exhibited two types of facilitation: continuous facilitation and discontinuous facilitation. The type of facilitation likely depended on the HFS charge density [46].

Table 1 Effects of DBS on neuronal activity around the electrodes

The variations of local neural response to HFS-DBS at the same target can be explained by a biophysically realistic computational framework [48]. This framework proposes that single pulses of electrical stimulation activate all converging presynaptic inputs to stimulate target neurons simultaneously, and the resulting responses are determined by the relative distribution of excitatory and inhibitory (E/I) afferent inputs [48, 49]. The different E/I ratios of synaptic inputs in various nuclei could account for the variable responses to DBS [50]. This framework explains that thalamic structures which mainly receive excitatory inputs show excitatory neural responses, while basal ganglia structures that mainly receive inhibitory inputs exhibit inhibitory responses [48, 51]. A similar pattern has been found by Xiao et al. [52] suggesting opposite effects on neuron subtypes following unique inputs within the targeted brain region. The STN consists of microcircuits regulated by the expression of different subtypes of nicotinic acetylcholine receptors. Local electrical stimulation mainly excites ι4β2+ neurons and significantly inhibits 58% of ι7+ neurons, potentially due to ι4β2+ neurons receiving more glutamatergic inputs and ι7+ neurons receiving more GABAergic inputs within the STN [52]. Therefore, mixed subpopulations of neurons with diverse inputs result in inconsistent, and sometimes conflicting responses to DBS at the same location.

Furthermore, computational studies suggest that repetitive HFS results in decreased site-specificity and neuronal suppression mediated by short-term synaptic depression [48, 50]. This phenomenon consists of a rapid decline in synaptic strength after brief bursts of activity, followed by a return to initial strength after a short rest period [53]. In vitro electrophysiology experiments demonstrated that HFS with 100 pulses delivered at 100 Hz significantly decreased the amplitude of evoked excitatory post-synaptic currents (eEPSCs) of STN neurons in dopamine-depleted slices. On the contrary, low-frequency stimulation (LFS) with 10 pulses per second at 40 Hz did not exhibit the same effect [54]. This short-term synaptic plasticity is a result of the depletion of readily released neurotransmitter vesicle pools when delivering rapid successive stimuli. Reduced presynaptic Ca2+ conductivity or inactivation of neurotransmitter release sites causes a reversible decline in synaptic efficacy due to the delayed recovery of vesicle fusion events [55,56,57,58,59].

In summary, at stimulation frequencies below the synaptic depression threshold, the local neuronal responses to DBS depend on the relative distributions of excitatory and inhibitory afferent inputs (site-specific effects), while at stimulation frequencies above the synaptic depression threshold, the local neuronal responses to DBS are progressively reduced, due to the synaptic depression effect (frequency-dependent effects).

Conversion of neuronal firing patterns by DBS

In addition to the intermingling of neurons exhibiting diverse responses to HFS, the modulation of neuronal firing patterns linked to intrinsic excitability further complicates the DBS effects [60,61,62]. For example, in STN, subthalamic neurons possess the capability to transit between single-spike firing and burst firing modes under normal conditions, a transition governed by the activation of distinct sets of ion channels contingent on the membrane's potential state [63]. However, in PD models, dopamine deficiency leads to relative membrane hyperpolarization, promoting STN burst firing. This stands in contrast to dopamine's role in depolarizing STN neuronal membranes [64, 65]. The abnormally heightened STN burst firing is intricately linked to parkinsonian symptoms [66,67,68] and serves both as the electrophysiological hallmark of PD and a primary target for therapeutic intervention via DBS [69]. Bursting activity in the STN has been observed to precede the pathological local field potential (LFP) oscillation in most cases, suggesting its pivotal role in generating aggregate-level LFP oscillations [70]. Therefore, amelioration of excessive STN burst firing is emerging as a fundamental mechanism underlying the clinical benefits of DBS (Fig. 2), a premise substantiated by research in PD patients, demonstrating that STN-DBS mitigates excessive burst firing and ameliorates PD symptoms [71].

Fig. 2
figure 2

Schematic illustration of DBS suppressing abnormal burst firing in the STN. Under normal conditions, STN neurons are capable of transition between single-spike firing and burst firing by activating distinct sets of ion channels based on the membrane potential state [63]. However, in PD, dopamine deficiency results in relative membrane hyperpolarization, facilitating burst firing in the STN [64, 65]. This abnormal burst firing pattern is closely associated with the manifestation of parkinsonian symptoms [66,67,68]. HFS-DBS induces a transient depolarization of the neuronal membrane. Subsequently, it effectively blocks voltage-gated currents, with a notable impact on T- and L-type Ca2+ currents as well as Ca2+-activated inward currents. This suppression of abnormal burst firing in the STN contributes to the amelioration of PD symptoms [60, 73]

Similarly, experiments with HFS in PD animals have shown that most STN neurons are inhibited and burst firing is reduced [72] (Fig. 2). Electrophysiological recordings have also shown that HFS can suppress STN burst firing through transient neuronal membrane depolarization and subsequent inhibition of voltage-gated currents, particularly T- and L-type Ca2+ currents and Ca2+-activated inward currents [60, 73]. Notably, STN burst firing can be bidirectionally regulated by altering the neuronal membrane potential using different electrical stimuli—depolarizing current (with or without pulses) decreases burst firing, whereas application of reverse-polarizing hyperpolarizing currents increases it [61, 62]. In addition, the hyperpolarization-activated cyclic nucleotide-gated channel 2 (HCN2) channels coupled to histamine H2 receptors, the GluN2A subunit-containing N-methyl-D-aspartate ionotropic receptors (NMDARs), and the ether-à-go-go-related gene (ERG) K+ channels have been identified as factors that can regularize neuronal firing patterns [74,75,76].

Firing pattern conversions have also been observed in the GPe and GPi. Studies in unanesthetized patients have shown that DBS in the GPi does not uniformly silence local neuronal activity, but rather disrupts pathological firing patterns by loosely entraining neuronal activity [77]. Similarly, research in monkeys has reported that HFS in the GPe does not lead to complete inhibition, but instead induces a complex restructuring of the temporal structure of neuronal activity [78]. This complex pattern change may be related to the therapeutic effect of DBS, but the exact mechanism remains to be elucidated and verified.

In summary, these findings suggest that the local response to DBS is influenced by both variations in synaptic inputs and alterations of intrinsic neuronal excitability through manipulation of membrane potential and ion channels to normalize firing patterns (Fig. 2). The intrinsic excitability of a neuron and synaptic efficacy, which represents the capacity of a presynaptic input to influence postsynaptic output, often work together to modify neural circuit function [27, 79, 80]. For instance, HFS can modify the temporal firing pattern of neurons in GPe and GPi, which underpins the beneficial effects of STN-DBS in PD [81]. The cortex and the direct cortical-STN projections, known as the hyperdirect pathway, are also potential components of the therapeutic mechanisms of STN-DBS [82]. In the following sections, we will discuss the antidromic and orthodromic effects of DBS through synaptic adaptations.

Orthodromic and antidromic effects of DBS

As discussed previously, DBS can affect various neural elements, including soma, axons, and dendrites of neurons. Studies have shown that DBS activates axons and dendrites in the stimulation region, increasing the frequency of action potential (AP) output from the soma of neurons [83]. Computational models have suggested that axons and dendrites have lower stimulation thresholds than the soma [84], suggesting that stimulation primarily affects axons and dendrites in the vicinity of the electrode. Therefore, most somatic effects are likely due to the propagation of stimulation effects from their local dendritic membranes [85, 86].

A computational model [87] proposes that if the stimulus is strong enough, it triggers APs that propagate orthodromically to the distal end of the cell body facilitating neurotransmitter release, and also propagate antidromically to activate upstream neurons [88]. In addition, the AP first reaches the base of the axonal arbor and then bifurcates at various branching points, eventually invading the entire axonal arbors and reaching all terminal points (Fig. 1). This leads to neurotransmitter release not only at the stimulation site but also at other terminal sites, illustrating multiple effects of DBS [81, 87] (Table 2).

Table 2 Orthodromic and antidromic effects of DBS

As an intrinsic property, electrical stimulation propagates in multiple directions, thus DBS can modulate neural circuits in various disease states. The sustained changes in neural activity induced by DBS may trigger adaptive changes within the nervous system, including activity-dependent synaptic adaptations in clinical settings [31,32,33]. This involves the reconfiguration of neuronal and synaptic components and the homeostatic regulation of neural circuit function [27].

A substantial body of research supports this synaptic adaptation theory, with observations indicating that DBS normalizes the distribution of corticostriatal glutamatergic terminals, thereby altering striatal glutamatergic neurotransmission in animal models [89]. In addition, DBS has been shown to modulate key components of the motor cortico-striato-thalamo-cortical loop in humans [90]. The enhancement of inhibitory synaptic transmission [91] and the restoration of intracortical inhibition associated with motor improvements [92] have also been reported in DBS studies, highlighting the role of adaptation-related mechanisms in its clinical effects.

Orthodromic and antidromic effects of STN-DBS

The STN controls two basal ganglia output nuclei: the GPi and the substantia nigra pars reticulata (SNr) [35]. This suggests that the effects of STN-DBS are related to the regulation of STN neurons projecting to the GPi and the SNr (Fig. 3). Electrophysiological data have illustrated the differences between STN-SNr and STN-GPi neurons in terms of their synaptic inputs, responses to electrical stimulation, and adaptations under PD conditions [93]. A prevalence of inhibitory synaptic inputs is more evident in STN-GPi neurons than in STN-SNr neurons. In PD mice, 6-hydroxydopamine (6-OHDA) lesioning disrupted the inhibitory inputs to STN-GPi neurons. This alteration reversed the predominance of inhibitory over excitatory inputs in STN-GPi neurons but did not affect synaptic inputs in STN-SNr neurons. Prolonged electrical stimulation enhanced inhibition and reduced excitation in both STN-SNr and STN-GPi neurons [93]. Consistent with this, in vivo recordings confirmed that STN-DBS led to the inhibition of neurons in GPi and SNr [42, 94], consequently suppressing the basal ganglia output and relieving the ventrolateral motor thalamic nucleus activity, thereby ameliorating PD symptoms [42]. However, in an experiment with two Parkinsonian rhesus monkeys, subthalamic stimulation elicited short-latency excitatory responses that caused a tonic increase in the average firing rate in the GPi and the GPe [81]. Furthermore, GCaMP (genetically encoded calcium indicator) fiber photometry in PD mice showed increased SNr activity during STN-DBS [45]. Although different experimental conditions may lead to different conclusions, this evidence supports that STN-DBS likely acts by disrupting neuronal activity patterns within the STN rather than by direct inhibition or antidromic activation [95]. Notably, similar to the complex response of STN neurons during HFS, spontaneous spiking of neurons in the SNr also exhibits variability [96].

Fig. 3
figure 3

Schematic of common DBS targets and their connections in mice. The subthalamic nucleus (STN) and the globus pallidus internus (GPi) are the primary DBS targets in PD [5,6,7,8]. The STN receives excitatory input from the cortex, referred to as the hyperdirect pathway [35], and from thalamic areas like the parafascicular nucleus [120]. Its main inhibitory input comes from the globus pallidus externus (GPe), contributing to the indirect pathway. STN neurons are primarily glutamatergic and project efferent fibers to the GPi and substantia nigra pars reticulata (SNr) to convey motor information. Notably, STN neurons also project to the caudate putamen (CPu), as demonstrated by viral tracing experiments in mice [225]. The GPi primarily receives glutamatergic afferents from STN and GABAergic input from the GPe and CPu. GPi sends GABAergic efferents to the thalamus and lateral habenula (LHb) [130]. The pedunculopontine nucleus (PPN) is a component of the mesencephalic locomotor region and is targeted to address gait and postural instability issues [12, 13]. In addition to its descending projections to the medulla and spinal cord, PPN neurons project to multiple ascending targets, including the thalamus and several basal ganglia components. These projections comprise a mixture of cholinergic and noncholinergic afferents [226]. A significant portion of inputs to the PPN originates in brainstem and midbrain structures, including the substantia nigra pars compacta (SNc) and SNr. PPN neurons also receive direct input from the zona incerta (ZI) in the hypothalamus [163]

As downstream nuclei of the STN (Fig. 3), the SNr, GPi, and ventral pallidum (VP) play a role in mediating PD-related pain, a prevalent and distressing non-motor symptom affecting 30%–95% of patients [97]. In normal mice, unilateral optogenetic activation (channelrhodopsin-2, ChR2) of the STN-SNr projections reduces the thermal but not the mechanical pain threshold, whereas stimulation of the STN-GPi and STN-VP projections reduces the mechanical but not the thermal pain threshold [98]. In PD models, optogenetic inhibition (NpHR) of the STN-SNr projection attenuates mechanical and thermal pain hypersensitivity, whereas photoinhibition of STN-GPi and STN-VP projections reduces only mechanical but not thermal pain hypersensitivity. These results suggest that subpopulations of the STN projection neurons control the perception of different pain stimuli. While STN-DBS may partially alleviate pain by improving motor function in PD, it is also essential to modulate individual STN projections linked to the nociceptive network to achieve its analgesic effects.

In addition to its orthodromic effects, several studies have suggested that DBS can also modulate neuronal activity in upstream nuclei through antidromic stimulation (Fig. 3) [82, 99,100,101]. While synaptic inputs play a minor role in maintaining spontaneous firing in STN neurons [102, 103], they play a critical role in shaping the firing rate and pattern [104, 105]. Notably, afferent axons projecting to STN neurons, particularly those within the hyperdirect pathway, are directly stimulated and highly implicated in the therapeutic effects of STN-DBS [106, 107].

Studies have shown that the layer-5 primary motor cortex (M1) neurons project extensively to the STN [108, 109]. Optical HFS (ChR2-E123T/T159C) of these projections significantly ameliorates hypokinesia and bradykinesia [108]. The spontaneous firing rate of M1 pyramidal cells increases in rats lesioned with 6-OHDA [82]. Interestingly, DBS can normalize the pathophysiological hyperactivity of motor cortex pyramidal neurons while inhibiting PV+ and activating somatostatin (SST)-expressing (SST+) GABAergic interneurons [82]. This study further demonstrated that direct optogenetic activation (ChR2-H134R) of cortical SST+ interneurons can mimic the effects of DBS and alleviate motor symptoms in a PD mouse model, suggesting that STN-DBS may recruit cortical GABAergic networks to inhibit the hyperactivity of cortical-subthalamic projections.

Paradoxically, another study in rats found that the average spontaneous firing rate of corticofugal projection neurons (CxFn) was reduced after 6-OHDA treatment, accompanied by increased episodes of burst firing. STN-DBS at 125 Hz significantly increased the spontaneous firing of CxFn, disrupting the dominance of the beta rhythm [88]. Antidromic activation of M1 during STN-DBS has been reported to contribute to the disruption of synchronization in cortical neuronal populations in Parkinsonian non-human primates [110]. However, the antidromic activation diminished over time and was not observed during GPi-DBS, which had similar therapeutic effects as STN-DBS, raising doubts about the mechanisms underlying the therapeutic effect of DBS. These inconsistencies in conclusion may be due to differences in experimental animals, models, stimuli, and measurement methods. Nevertheless, these studies suggest that STN-DBS, to some extent, regulates cortical neuronal activity through antidromic transmission.

Within the STN, the hyperdirect and indirect pathways serve as the main motor inhibitory circuits in the basal ganglia [35]. The hyperdirect pathway predominantly conveys glutamatergic inputs from the motor cortex to the STN, while the indirect pathway primarily transmits GABAergic inputs to the STN from the GPe. An intrinsic homeostatic mechanism in the STN has been identified to balance cortical excitation by adjusting the strength of GPe inhibition [111]. Stimulation of the motor cortex-STN inputs by optogenetic activation (ChR2-H134R) of motor cortical projection neurons induces heterosynaptic long-term potentiation (LTP) of GPe-STN transmission through NMDARs. This process may promote pathological activity after dopamine depletion [111]. In conclusion, DBS normalizes pathological hyperactivity of the motor cortex and indirectly inhibits GPe-STN transmission through heterosynaptic regulation of the hyperdirect and indirect pathways. NMDAR-dependent processes in neurons receiving afferents from the STN are likely a cellular mechanism by which STN-DBS exerts its therapeutic effects.

The thalamus is a critical node in the basal ganglia network, and several studies indicate that STN-DBS can affect the firing rate and the bursting activity of thalamic neurons [112, 113]. Among the different thalamic nuclei, the parafascicular nucleus (Pf) has been identified as a critical player in modulating basal ganglia activity and mediating the therapeutic effects of STN-DBS [114, 115]. The Pf is involved in the regulation of striatal function and plays a critical role in learning, arousal, and behavioral flexibility [116, 117]. Pf neurons project to both the STN and striatum, suggesting that the Pf-STN pathway may contribute to the clinical benefits of STN-DBS [118, 119].

A study shows that optogenetic stimulation (ChR2-H134R) of Pf projections to the STN leads to improved motor function, whereas stimulation of the Pf-striatum cell body or terminal does not have the same effect [119]. Unilateral or bilateral optogenetic stimulation of the Pf-STN terminal significantly improves locomotion and alleviates severe akinesia in a bilateral 6-OHDA PD model [119]. In addition, optogenetic enhancement (oChIEF) of the Pf-STN circuit using the optical LTP approach restores motor learning. Notably, inhibition of PV+ STN neurons prevents this LTP-based recovery, highlighting the critical role of PV+ STN neurons in this rescue process [120]. These findings suggest that the Pf-STN pathway provides a circuit mechanism that may elucidate the clinical efficacy of STN-DBS in alleviating motor symptoms of PD.

Orthodromic and antidromic effects of GPi-DBS

GPi, one of the commonly targeted brain regions for DBS, has shown therapeutic effects like STN. Studies have demonstrated that GPi-DBS can reduce the activity of neurons within the STN. The suppression could be attributed to the activation of fibers that originate in the GPe and pass through the GPi. Recently, researchers have discovered a phenomenon called evoked resonant neural activity (ERNA) occurring in GPi-DBS [121, 122]. ERNA is a HFS-evoked response typically occurring at 200 to 500 Hz, and is associated with synchronized patterned neuronal inhibition. Additionally, ERNA has also been reported in STN-DBS [123,124,125,126,127,128], and a biophysical model suggests that it results from the reciprocal connections between the STN and GPe [125]. GPi-DBS has the potential to activate fibers within the STN-GPe loop or affect axon collaterals [122], leading to the possible indirect triggering of ERNA. It seems that GPi-DBS is effective in influencing the activity of STN but an alternative theory states that both STN-DBS and GPi-DBS produce comparable modulatory effects on an "overlapping" functional network in PD patients [122, 129]. This hypothesis is supported by practical research that discovered remarkably similar connectivity profiles associated with STN-DBS and GPi-DBS [129].

In practice, while both GPi and STN are recommended as potential DBS targets, their clinical outcomes differ. STN-DBS typically results in a greater reduction of levodopa usage, whereas GPi-DBS is linked with a lower frequency of neuropsychiatric side effects [15]. A viral genetic tracing study in mice showed that neurons in the entopeduncular nucleus (EP, analogous to GPi in human) gather inputs from both the striatum and GPe. They then relay the inputs prominently to the lateral habenula (LHb) and the ventro-anterior lateral thalamus/ventro-medial thalamus (VAL/VM) [130] (Fig. 3). The neurons situated in the EP provide inhibitory input to the VAL/VM thalamus to control movement. Conversely, when they are inhibited by upstream basal ganglia nuclei, the movement is allowed [131]. Electrophysiological studies in primates and humans with PD suggest that increasing the firing rates of GPi neurons could lead to the development of motor deficits associated with the condition, likely due to VAL/VM thalamus inhibition and decreased basal ganglia output [40, 132, 133]. LHb neurons are arranged to receive EP input and project to the rostral medial tegmental area, which innervates the ventral tegmental area and is involved in aversive conditioning [134]. Consistent with this study, electrophysiological studies in primates have demonstrated that GPi neurons, projecting to the LHb, respond to reward-related signals and some sensory stimuli [135].

Therefore, with regards to its orthodromic effects, GPi-DBS has the potential to regulate neuron activity in the VAL/VM thalamus, thus improving motor function. Additionally, it may also yield emotional and neuropsychological benefits by impacting the LHb activity.

Position-dependent therapeutic effects

Due to the position-dependent nature of DBS therapy, the positioning of electrodes and their active contacts is a crucial parameter in DBS treatment, which requires precise programming by physicians. Clinically, the electrodes are often placed in regions that yield the maximal DBS benefits, such as the dorsolateral STN [136,137,138,139], the posterolateral GPi [140, 141], and the caudal PPN [142,143,144,145]. These regions have distinct features of circuit connectivity and cellular composition, which should be seriously considered when tailoring electrode placement (Fig. 4).

Fig. 4
figure 4

Position-dependent therapeutic effects of DBS for PD. a Topologically determined position-dependent effects (exemplified by STN-DBS). In mice, the STN receives inputs from both the cortex and the GPe. These inputs exhibit a topographically graded organization, forming the hyperdirect and indirect pathways, respectively [150]. Furthermore, a topographical organization exists between the cortex and GPe. To elaborate, the posterolateral to anteromedial regions of the STN receive projections from various cortical areas, including sensorimotor, association, and limbic regions [136, 147, 148]. Within the STN, there is a distinct distribution of PV+ glutamatergic neurons, primarily clustered in the dorsolateral and middle regions. These neurons exhibit unique burst firing patterns and may contribute to excessive burst firing observed in PD [150]. Consequently, clinical benefits are typically observed when DBS electrodes are precisely positioned within the dorsolateral sensorimotor area [136,137,138,139]. b Neuronal population-determined position-dependent effects (exemplified by the PPN-DBS). The PPN is an integral component of the mesencephalic locomotor region, characterized by the spatial distribution of glutamatergic, GABAergic, and cholinergic neurons [159, 160]. Among these, glutamatergic neurons represent the major subpopulation. Activation of caudal glutamatergic neurons promotes locomotion [155, 162, 163], while their rostral counterparts induce locomotor arrest [162]. GABAergic neurons are slightly more concentrated in the rostral PPN [161] and tend to decrease the locomotor speed when activated [155, 163]. Cholinergic neurons outnumber GABAergic neurons, yet their influence on locomotion is less clear, with reported effects spanning from improvement to suppression of movement upon optogenetic activation [155, 163]. The specific distribution of these neuronal subpopulations likely underlies the rationale for targeting the caudal PPN as the optimal stimulation site [142,143,144,145]

Topological factor for position-dependent effects

The position-dependent effects are influenced by topological factors (Fig. 4a). Despite the STN and GPi being relatively small nuclei in primates and rodents, several human studies have revealed the presence of three territories within these regions: sensorimotor, associative, and limbic territories [136, 146, 147]. For instance, in the case of the STN, anatomical-functional subregions extending from the posterolateral to the anteromedial parts of the nucleus receive projections from sensorimotor, association, and limbic areas in the cortex [136, 147]. This organization aligns with findings from anterograde tracing studies in primates [148]. Single-cell recordings from PD patients have also identified neurons with sensorimotor responses in the dorsolateral region of the STN [149]. This topographical organization of the STN supports the structural basis for information processing from the cortex to the basal ganglia. Recent mouse studies have further confirmed this organization, revealing a graded distribution of cortical projections to the STN with a notable degree of overlap along its longitudinal axis [150], in line with earlier studies [148, 151]. The convergent projection patterns within the STN reinforce the clinical efficacy of DBS in the dorsolateral STN, which is associated with sensorimotor functions.

Similarly, the GPi also exhibits a topological distribution [146]. The anterior region of GPi is associated with limbic territories and associative connectivity, while the posterolateral regions of the nucleus are linked to sensorimotor functions [140, 152]. Individualized treatment planning, focusing on identifying the sensorimotor regions of the GPi, particularly its posterolateral aspect, has been demonstrated to enhance the alleviation of PD motor symptoms through DBS [140, 153]. This underscores the importance of the topological structure when determining electrode implantation sites in DBS therapy.

Neuronal population factor for position-dependent effects

The position-dependent effects of DBS are influenced not only by topological factor but also by the composition of neuronal populations within targeted regions (Fig. 4b). A close examination of the STN revealed heterogeneity in neuronal population, although it is generally considered a homogeneous glutamatergic nucleus. Serial multiplex single-molecule fluorescence in situ hybridization data showed the presence of two populations: PV+ and PV− neurons in the STN. The glutamatergic PV+ neurons predominantly occupy the dorsolateral and middle portions of the STN and exhibit characteristics of phasic burst firing compared to the PV− subpopulation [150]. Electrophysiological recordings in PD patients have further corroborated this finding, showing spatial distribution of burst activity mainly in the dorsal region of the STN [154]. This burst firing, along with beta oscillations, is a hallmark of PD. STN-DBS mitigates these pathophysiological patterns that are associated with motor symptoms in PD [69], suggesting that the glutamatergic PV+ neurons are the source of excessive burst firing in PD. Their distribution in the dorsolateral and middle parts of STN likely underlies the position-dependent effects observed in STN-DBS [12, 13].

Similarly, for PPN-DBS, the position-dependent effects are pivotal for addressing freezing gait and postural abnormalities in PD patients who are resistant to dopaminergic treatments [155,156,157,158]. However, clinical outcomes of PPN-DBS can be variable [158], possibly due to the nonspecific electrical stimulation of different PPN populations and regions. The PPN consists of spatially distributed glutamatergic, GABAergic, and cholinergic neurons [159, 160], with the glutamatergic neurons as the major neuronal subpopulation. These neurons are functionally diverse and produce different motor responses upon activation [161]. For instance, the caudal vesicular glutamate transporter 2 (VGLUT2)-expressing (VGLUT2+) neurons promote locomotion within an exploratory speed range [155, 162, 163], while those in the rostral PPN induce locomotor arrest. The GABAergic neurons are more concentrated in the rostral PPN [161] and decrease locomotor speed when optogenetically activated (ChR2) in rodents [155, 163]. The cholinergic neurons, while more abundant than GABAergic neurons, exhibit uncertain effects on locomotion, as both improvement and suppression of movement upon optogenetic activation have been reported in mice [155, 163].

Studies indicate that the motor-enhancing effect of PPN-DBS is specifically attributed to the caudal PPN [142,143,144,145]. Chemical genetic activation (hM3Dq) of the caudal VGLUT2+ PPN neurons can rescue movement deficits in PD mice [160], and DBS in the caudal PPN improves gait parameters in PD rats [142]. This is consistent with clinical findings that DBS in the caudal PPN enhances gait freezing and postural stability in PD patients [144]. Importantly, the glutamatergic PPN neurons projecting to different targets, such as the SNr or spinal cord, may underlie various DBS effects, influencing forelimb movements, behavior, or body extension, depending on the specific projection [156]. As a result, the therapeutic effects of PPN-DBS on freezing gait and postural balance may depend on the specific subsets of PPN neurons being stimulated.

In summary, both the topological structure and the composition of neuronal subpopulations contribute to the spatial distribution of inputs and outputs in targeted brain regions. These factors collectively determine the position-dependent therapeutic effects of DBS.

DBS programming-dependent therapeutic effects

DBS programming, the adjustment of electrical stimulation parameters to optimize clinical benefit for individual patients, is a critical aspect of DBS. Physicians must carefully fine-tune parameters such as frequency, pulse width, voltage, and electrode contact to achieve the best symptom relief with minimal side effects [164]. The selection of the active contact is predominantly influenced by position-dependent effects, and consequently, it affects communication within the pertinent neural circuits. The other programming parameters also play a vital role in modulating neural circuitry and synaptic plasticity.

Frequency-dependent therapeutic effects

Frequency is a crucial parameter in DBS programming. Studies have found that the magnitude of the beneficial effect in PD patients is most pronounced within the frequency range of 130–185 Hz, with a progressive improvement in motor symptoms as the frequency increases from 50 to 130 Hz [165]. Similar observations have been made in PD mice undergoing STN-DBS, showing that the movement speed scales linearly with frequency up to approximately 120 Hz. This phenomenon mirrors the response seen in PD patients undergoing STN-DBS. The rationale behind these observations lies in the synaptic depression caused by repetitive higher frequency stimulation, which weakens synaptic transmission strength and suppresses somatic firing in postsynaptic neurons. The application of halorhodopsin (NpHR) as an optogenetic tool for inhibiting postsynaptic neurons in the GPi or STN has shown remarkable promise in improving motor symptoms in hemiparkinsonian animal models [166,167,168,169]. These effects are consistent with the therapeutic benefits observed with HFS-DBS in PD patients.

However, when attempting to directly activate local excitatory STN neurons using optogenetic methods like ChR2, results were less favorable, with only minimal changes in rotational behavior and even motor deficits in the contralateral limb [107, 170]. These findings suggest that optogenetic stimulation, particularly with ChR2, cannot precisely replicate the effects of STN-DBS. This limitation is attributed to the relatively slow opening and closing kinetics of ChR2, which cannot generate firing rates > 100 Hz in the STN or drive glutamate release at rates greater than 100 Hz [171, 172]. To potentially bridge this gap, faster optogenetic actuators like fast channelrhodopsin (i.e., ChR2-E123T/T159C) or Chronos have been proposed [108, 173]. Optogenetic STN stimulation using Chronos at a frequency of 130 pulses per second demonstrated a reduction in pathological circling behavior and an improvement in forelimb stepping deficits, mirroring the effects of electrical DBS [173]. Faster optogenetic actuators have the potential to generate higher overall firing rates and greater firing rate fidelity than ChR2. Consequently, the therapeutic effects of DBS may be more closely tied to the stimulation frequency when using these faster actuators [171, 174]. Furthermore, a study by Yu et al. [173] noted significant differences in the absolute changes in the firing rates of responsive STN neurons across several stimulation frequencies using optogenetic stimulation with Chronos. In an optogenetic experiment, the targeted cells and their axons, rather than afferent or passing axonal fibers, were selectively activated [175]. This implies that different stimulation frequencies have varying effects on the postsynaptic neuronal intrinsic excitability.

Recent human studies have proposed that the shape and the amplitude of ERNA, generated by DBS, also depend on the frequency and duration of stimulation [124, 125]. The steady states of ERNA frequency and amplitude do not immediately return to baseline when STN-DBS is turned off, and it takes several seconds for these parameters to normalize [124]. Higher stimulation frequencies have been associated with significantly longer silent periods after stimulation [44]. These slow temporal dynamics in the recovery period may be linked to the time needed to replenish presynaptic vesicle pools, which affect the synaptic transmission fidelity [58]. In essence, the gradual changes observed in ERNA may be correlated with the progressive deterioration of synaptic transmission fidelity [124, 176]. While neurons and the axons of afferent and efferent pathways can potentially keep pace with HFS for an extended period, synaptic resources are more likely to be depleted within seconds to minutes. This depletion results in a functional disconnection between the STN and the broader basal ganglia network [124].

Another hypothesis regarding synaptic depression suggests the involvement of presynaptic metabotropic GABAB receptors, which lead to longer-lasting inhibitory effects at higher frequencies compared to lower frequencies [44]. This mechanism relies heavily on the regulation of Ca2+ conductance. Reductions in Ca2+ conductance, both on autoreceptors located on GABA-releasing terminals and on heteroreceptors in neighboring terminals, are thought to be induced by HFS. This ultimately results in the inhibition of neurotransmitter release [59, 177,178,179]. However, it is important to note that human studies have yet to definitively elucidate these adaptation mechanisms at the molecular and synaptic levels [31,32,33]. Consequently, these explanations remain speculative and need further validation through additional research involving animal models.

Pulse width/intensity-dependent therapeutic effects

The therapeutic efficacy of DBS is profoundly influenced by the spatial distribution of the stimulation field in relation to the brain's anatomy [180,181,182]. Clinical studies have lent support to this idea, indicating that the volume of tissue activated, a parameter modifiable through DBS settings, including pulse width and voltage or current intensity titration, can be instrumental in altering the range of stimulated nervous tissue, consequently impacting the clinical outcomes of DBS [137, 183, 184].

Preclinical data reinforce these findings. The effectiveness of STN-DBS, as measured by improvements in movement speed in PD mice, has exhibited a linear correlation with pulse width and current intensity [139]. Nevertheless, once pulse width values reach a certain threshold, lower stimulus intensities may be needed to achieve the desired clinical effect [185]. It is worth noting that dyskinesia can emerge as a side effect of HFS in humans [186], and similar effects have been demonstrated in mice that HFS, alongside increased pulse width and current, can lead to severe dyskinesia in vivo [187].

In vitro electrophysiological studies have illuminated the role of pulse width in determining the type of neuronal response elicited by HFS in the GPi. Low charge density HFS (60 Οs) primarily induces excitation, while high charge density HFS (400 Οs) triggers a distinct subtype of excitation characterized by late inhibition, which involves glutamatergic and cholinergic modulation, as well as Ca2+-activated non-specific cation channels [46]. Furthermore, elevating the intensity of HFS has been found to extend the duration of excitation in the excitation-only after-effect [46]. The amplitudes of ERNA and the silent periods observed during HFS-DBS also display a positive relationship with pulse width and intensity [124, 188]. These observations underscore the pivotal role of programming parameters in modulating synaptic adaptions.

Long-lasting effects mediated by optimized programming

While DBS has demonstrated symptomatic efficacy in PD, its effects are transient and vanish once the stimulation is discontinued, leading to a swift return of motor symptoms [139, 189]. However, recent investigations have shown that optimized programming, facilitated by patterned electrical stimulation, can yield enduring therapeutic benefits. Coordinated reset DBS (CR-DBS), an innovative DBS approach, is under investigation in preclinical and clinical studies [190] and have shown potential to induce sustained therapeutic improvements in Parkinsonian symptoms, even after stimulation cessation [191,192,193,194]. CR-DBS aims to reconfigure the neuronal connectivity therapeutically by modulating synaptic plasticity, particularly spike timing-dependent plasticity (STDP) [195,196,197]. This approach reduces the coincidence rates, resulting in a decrease of synaptic weights due to STDP, making the network unlearn pathological connectivity and synchronicity [198].

Furthermore, Spix and colleagues have delineated a precise DBS stimulation protocol with long-term efficacy in mice, highlighting distinct responses of two types of neurons in the GPe to electrical stimulation [199]. The GPe, a basal ganglia nucleus, maintains connections with various brain regions, including the thalamus, amygdala, brainstem, and cortex [200, 201], and plays a role in abnormal neural dynamics seen in PD [202]. Mastro et al. [201] have demonstrated that optogenetic activation of PV+ neurons using ChR2 and inhibition of Lim homeobox protein 6 (Lhx6)-expressing (Lhx6+) neurons via Arch (a light-activated inhibitory proton pump), two subpopulations in the GPe with distinctive intrinsic physiological and projection properties, ameliorate locomotor deficits in dopamine-depleted mice four hours following stimulation [203]. Building on this, Spix et al. employed a specific electrical stimulation mode (175 Hz, 200 ms) utilizing brief bursts to effectively segregate the responses of PV+ and Lhx6+ GPe neurons. Burst stimulation not only improves bradykinesia in 6-OHDA-lesioned mice but also provides long-lasting therapeutic benefits that persist for hours post-stimulation [199]. These findings suggest that the induction of sustained behavioral improvement arises from frequency-dependent, cell-type-specific activation or inhibition, specifically, an increase in the firing rate of PV+ GPe neurons relative to Lhx6+ GPe neurons.

The differing circuit properties of PV+ and Lhx6+ GPe neurons are one mechanism underpinning their distinct firing responses and the sustained effects observed. Despite receiving similar levels of excitatory input from the STN, Lhx6+ GPe neurons receive proportionally more inhibition than PV+ GPe neurons from the D1-dopamine receptor-expressing spiny-projection neuron (D1-SPN) afferents. The specific electrical stimulation applied is likely skewed toward the antidromic activation of D1-SPNs, thereby producing more potent inhibition of Lhx6+ GPe neurons than PV+ GPe neurons. This disruption in the network's balance through the stimulation of distinct neuronal subpopulations results in sustained therapeutic benefits (Fig. 5) [199]. These findings may serve as a basis for understanding the cell-type-specific mechanisms of DBS and exploring tailored stimulation strategies for potential clinical applications.

Fig. 5
figure 5

Optimized programming of DBS produces long-lasting effects. An example of GPe-DBS with population-specific neuromodulation that prolongs therapeutic benefits [199]. Both PV+ GPe and Lhx6+ GPe neurons receive excitatory inputs from STN to a similar degree. However, a distinction arises in their inhibition patterns originating from D1-SPN afferents. Lhx6+ GPe neurons experience proportionally greater inhibition from these afferents compared to PV+ GPe neurons. A highly precise electrical stimulation mode (175 Hz, 200 ms) with brief bursts is designed to bias towards antidromic activation of D1-SPNs, resulting in more potent inhibition of Lhx6+ GPe, while simultaneously exciting PV+ GPe neurons. Consequently, the firing rates of PV+ GPe neurons exceed those of Lhx6+ GPe neurons, which plays a crucial role in ameliorating bradykinesia in 6-OHDA-lesioned PD mice. Notably, these improvements persist long after stimulation. While the precise mechanism responsible for the extended therapeutic effects achieved through GPe-DBS with relative cell-specificity remains elusive, it is conceivable that this specific stimulation pattern bears similarities to certain forms of DBS, notably adaptive and coordinated reset DBS, both of which have shown the ability to produce enduring therapeutic benefits [191,192,193,194]

Optimization of DBS treatment

In recent decades, DBS is emerging as a pivotal treatment for various refractory movement and psychiatric disorders. While DBS has demonstrated efficacy and safety, it also has challenges including diminishing efficacy over time and the occurrence of adverse effects [187, 204]. To address these issues and enhance therapeutic outcomes, researchers have been exploring novel stimulation methods, including innovative waveform shapes and patterns. However, further evaluation and refinement are necessary to fully optimize DBS treatments [205].

Adaptive stimulation

Adaptive DBS (aDBS) is a promising avenue for DBS application, which seeks to enhance the effectiveness and safety of DBS by dynamically adjusting stimulation parameters based on real-time feedback signals. Unlike conventional open-loop DBS, aDBS operates as a closed-loop system, which is capable of bidirectional communication and automatic parameter adjustment [206, 207]. This feature makes aDBS a potential strategy to control the symptoms of PD [208, 209] and mitigate the levodopa-induced dyskinesia [210].

To effectively implement aDBS, it is essential to identify an electrophysiological biomarker that can accurately reflect the clinical characteristics of the disease and serve as a feedback for the system. The LFP, particularly beta oscillation across the motor network, has been widely employed as a biomarker for aDBS [209, 211, 212]. However, since beta oscillation is more closely associated with rigidity and bradykinesia than with tremor [213,214,215,216], there is a pressing need for the identification of biomarkers that can capture different PD symptoms to facilitate the development of effective aDBS algorithms.

One promising candidate is the narrowband gamma activity (60–90 Hz) observed in the motor cortex and STN during dyskinesia [217]. This gamma activity shows potential as a biomarker for aDBS [218], as it is less influenced by voluntary movements and exhibits sensitivity to stimulation-induced dyskinesia, displaying a higher signal amplitude and a more favorable signal-to-noise ratio compared to beta activity [217]. Furthermore, burst firing in the STN by individual neurons has been directly implicated in PD pathophysiology and the manifestation of PD symptoms [69]. Its role in aDBS strategies warrants further exploration [219].

In summary, further refinement of biomarkers is needed to advance aDBS application. Future research endeavors may explore new biomarkers to unlock even better therapeutic outcomes through aDBS treatments.

Directional stimulation

Directional stimulation technology represents a promising frontier in the evolution of DBS therapy. It introduces a level of precision previously unseen in DBS treatments, achieved by manipulating or configuring electrodes with radially segmented contacts, anodes, and cathodes to guide the flow of current in specific directions. This innovative approach offers a potential for a more nuanced and adaptable stimulation field [205, 220], capable of preventing unnecessary spread or "leakage" of stimulation, thereby expanding the therapeutic window in practical DBS applications. Adverse effects of DBS often stem from its non-selective stimulation that affects nearby neurons, including surrounding structures involved in various circuit connections with diverse physiological functions.

Clinical investigations have demonstrated that directional electrodes can deliver more efficient stimulation at a given amplitude compared to omnidirectional electrodes [221,222,223]. The directional electrodes hold promise for enhancing DBS effectiveness while minimizing adverse effects [220, 223, 224]. Improvement of the understanding of brain anatomy and circuit projections will guide the precise targeting of DBS stimulation. Therefore, directional stimulation stands out as a pivotal direction for the development of DBS therapy.

More precise stimulation, more effective treatment

Recent advances in neuroscience and brain function research have provided deeper insights into the topological connections between different brain regions and the distribution patterns of distinct neuronal subgroups within various nuclei. These advances have partially illuminated the electrophysiological and circuit mechanisms underpinning the clinical effects of DBS, while also directing the development of DBS technology. The intricate anatomical complexity and circuitry interconnections among numerous brain nuclei suggest that non-selective DBS stimulation may lead to unintended clinical side effects.

The emerging cutting-edge technologies such as aDBS and directional stimulation improve DBS therapies toward delivering more efficient and targeted interventions. It is increasingly evident that leveraging more specific DBS stimulations to achieve precise modulation of neural function is a future direction of development in this field.

Conclusions

DBS stands as a valuable treatment modality for advanced PD. In this review, we have delved into recent opto-DBS studies, shedding light on the potential mechanisms of neuronal and synaptic adaptations that underlie the efficacy of DBS in PD. Response of local neural circuits to DBS can be affected by a complicated interplay of many factors, including the distribution of presynaptic inputs, frequency-dependent synaptic depression, and the intrinsic excitability of postsynaptic neurons, which involves membrane potential dynamics and ion channel functionality. These factors collectively enable both antidromic and orthodromic modulation of neural circuits, laying the foundation for understanding the position- and programming-dependent therapeutic effects and side effects associated with DBS.

Availability of data and materials

The datasets used and/or analyzed during the current study are available from the corresponding author on reasonable request.

Abbreviations

6-OHDA:

6-Hydroxydopamine

aDBS:

Adaptive DBS

ChR2:

Channelrhodopsin-2

CR-DBS:

Coordinated reset DBS

CxFn:

Corticofugal projection neurons

D1-SPN:

D1-dopamine receptor-expressing spiny projection neuron

DBS:

Deep brain stimulation

E/I:

Excitatory/inhibitory

EP:

Entopeduncular nucleus

ERNA:

Evoked resonant neural activity

GABA:

Îł-Aminobutyric acid

GPe:

Globus pallidus externus

GPi:

Globus pallidus internus

HFS:

High-frequency stimulation

LFP:

Local field potential

LHb:

Lateral habenula

Lhx6:

Lim homeobox protein 6

LTP:

Long-term potentiation;

NMDAR:

N-methyl-D-aspartate ionotropic glutamate receptor

NpHR:

Halorhodopsin

PD:

Parkinson's disease

Pf:

Parafascicular nucleus

PPN:

Pedunculopontine nucleus

PV:

Parvalbumin

SNr:

Substantia nigra pars reticulata

SPN:

Spiny projection neuron

SST:

Somatostatin

STN:

Subthalamic nucleus

VAL:

Ventro-anterior lateral thalamus

VM:

Ventro-medial thalamus

VP:

Ventral pallidum

VGLUT2:

Vesicular glutamate transporter 2

References

  1. Poewe W, Seppi K, Tanner CM, Halliday GM, Brundin P, Volkmann J, et al. Parkinson disease. Nat Rev Dis Prim. 2017;3:17013.

    Article  PubMed  Google Scholar 

  2. Bloem BR, Okun MS, Klein C. Parkinson’s disease. Lancet. 2021;397(10291):2284–303.

    Article  CAS  PubMed  Google Scholar 

  3. Deniau JM, Degos B, Bosch C, Maurice N. Deep brain stimulation mechanisms: beyond the concept of local functional inhibition. Eur J Neurosci. 2010;32(7):1080–91.

    Article  PubMed  Google Scholar 

  4. Wichmann T, Bergman H, DeLong MR. Basal ganglia, movement disorders and deep brain stimulation: advances made through non-human primate research. J Neural Transm (Vienna). 2018;125(3):419–30.

    Article  PubMed  Google Scholar 

  5. Ramirez-Zamora A, Ostrem JL. Globus pallidus interna or subthalamic nucleus deep brain stimulation for Parkinson disease: a review. JAMA Neurol. 2018;75(3):367–72.

    Article  PubMed  Google Scholar 

  6. Follett KA, Weaver FM, Stern M, Hur K, Harris CL, Luo P, et al. Pallidal versus subthalamic deep-brain stimulation for Parkinson’s disease. N Engl J Med. 2010;362(22):2077–91.

    Article  CAS  PubMed  Google Scholar 

  7. Odekerken VJ, van Laar T, Staal MJ, Mosch A, Hoffmann CF, Nijssen PC, et al. Subthalamic nucleus versus globus pallidus bilateral deep brain stimulation for advanced Parkinson’s disease (NSTAPS study): a randomised controlled trial. Lancet Neurol. 2013;12(1):37–44.

    Article  PubMed  Google Scholar 

  8. Moro E, Lozano AM, Pollak P, Agid Y, Rehncrona S, Volkmann J, et al. Long-term results of a multicenter study on subthalamic and pallidal stimulation in Parkinson’s disease. Mov Disord. 2010;25(5):578–86.

    Article  PubMed  Google Scholar 

  9. Videnovic A, Metman LV. Deep brain stimulation for Parkinson’s disease: prevalence of adverse events and need for standardized reporting. Mov Disord. 2008;23(3):343–9.

    Article  PubMed  Google Scholar 

  10. Hariz MI, Krack P, Alesch F, Augustinsson LE, Bosch A, Ekberg R, et al. Multicentre European study of thalamic stimulation for parkinsonian tremor: a 6 year follow-up. J Neurol Neurosurg Psychiatry. 2008;79(6):694–9.

    Article  CAS  PubMed  Google Scholar 

  11. Helmich RC, Toni I, Deuschl G, Bloem BR. The pathophysiology of essential tremor and Parkinson’s tremor. Curr Neurol Neurosci Rep. 2013;13(9):378.

    Article  PubMed  Google Scholar 

  12. Yu K, Ren Z, Guo S, Li J, Li Y. Effects of pedunculopontine nucleus deep brain stimulation on gait disorders in Parkinson’s disease: a meta-analysis of the literature. Clin Neurol Neurosurg. 2020;198:106108.

    Article  PubMed  Google Scholar 

  13. Wilcox RA, Cole MH, Wong D, Coyne T, Silburn P, Kerr G. Pedunculopontine nucleus deep brain stimulation produces sustained improvement in primary progressive freezing of gait. J Neurol Neurosurg Psychiatry. 2011;82(11):1256–9.

    Article  PubMed  Google Scholar 

  14. Schuepbach WMM, Tonder L, Schnitzler A, Krack P, Rau J, Hartmann A, et al. Quality of life predicts outcome of deep brain stimulation in early Parkinson disease. Neurology. 2019;92(10):e1109–20.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Odekerken VJ, Boel JA, Schmand BA, de Haan RJ, Figee M, van den Munckhof P, et al. GPi vs STN deep brain stimulation for Parkinson disease: three-year follow-up. Neurology. 2016;86(8):755–61.

    Article  CAS  PubMed  Google Scholar 

  16. Ricchi V, Zibetti M, Angrisano S, Merola A, Arduino N, Artusi CA, et al. Transient effects of 80 Hz stimulation on gait in STN DBS treated PD patients: a 15 months follow-up study. Brain Stimul. 2012;5(3):388–92.

    Article  PubMed  Google Scholar 

  17. Honey CR, Hamani C, Kalia SK, Sankar T, Picillo M, Munhoz RP, et al. Deep brain stimulation target selection for Parkinson’s disease. Can J Neurol Sci. 2017;44(1):3–8.

    Article  PubMed  Google Scholar 

  18. Dallapiazza RF, De Vloo P, Fomenko A, Lee DJ, Hamani C, Munhoz RP, et al. Considerations for patient and target selection in deep brain stimulation surgery for Parkinson’s disease. In: Stoker TB, Greenland JC, editors., et al., Parkinson’s disease: pathogenesis and clinical aspects. Brisbane: Codon Publications; 2018.

    Google Scholar 

  19. Dostrovsky JO, Lozano AM. Mechanisms of deep brain stimulation. Mov Disord. 2002;17(Suppl 3):S63-68.

    Article  PubMed  Google Scholar 

  20. Vitek JL. Mechanisms of deep brain stimulation: excitation or inhibition. Mov Disord. 2002;17(Suppl 3):S69-72.

    Article  PubMed  Google Scholar 

  21. Benabid AL, Benazzous A, Pollak P. Mechanisms of deep brain stimulation. Mov Disord. 2002;17(Suppl 3):S73-74.

    Article  PubMed  Google Scholar 

  22. Perlmutter JS, Mink JW. Deep brain stimulation. Annu Rev Neurosci. 2006;29:229–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Kringelbach ML, Jenkinson N, Owen SL, Aziz TZ. Translational principles of deep brain stimulation. Nat Rev Neurosci. 2007;8(8):623–35.

    Article  CAS  PubMed  Google Scholar 

  24. Ashkan K, Rogers P, Bergman H, Ughratdar I. Insights into the mechanisms of deep brain stimulation. Nat Rev Neurol. 2017;13(9):548–54.

    Article  PubMed  Google Scholar 

  25. Yuan TF, Li WG, Zhang C, Wei H, Sun S, Xu NJ, et al. Targeting neuroplasticity in patients with neurodegenerative diseases using brain stimulation techniques. Transl Neurodegener. 2020;9(1):44.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Mahlknecht P, Foltynie T, Limousin P, Poewe W. How does deep brain stimulation change the course of Parkinson’s disease? Mov Disord. 2022;37(8):1581–92.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Ganguly K, Poo MM. Activity-dependent neural plasticity from bench to bedside. Neuron. 2013;80(3):729–41.

    Article  CAS  PubMed  Google Scholar 

  28. Gittis AH, Yttri EA. Translating insights from optogenetics to therapies for Parkinson’s disease. Curr Opin Biomed Eng. 2018;8:14–9.

    Article  PubMed  PubMed Central  Google Scholar 

  29. Turrigiano G. Homeostatic synaptic plasticity: local and global mechanisms for stabilizing neuronal function. Cold Spring Harb Perspect Biol. 2012;4(1):a005736.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Herrington TM, Cheng JJ, Eskandar EN. Mechanisms of deep brain stimulation. J Neurophysiol. 2016;115(1):19–38.

    Article  CAS  PubMed  Google Scholar 

  31. Prescott IA, Dostrovsky JO, Moro E, Hodaie M, Lozano AM, Hutchison WD. Levodopa enhances synaptic plasticity in the substantia nigra pars reticulata of Parkinson’s disease patients. Brain. 2009;132:309–18.

    Article  CAS  PubMed  Google Scholar 

  32. Ruge D, Cif L, Limousin P, Gonzalez V, Vasques X, Hariz MI, et al. Shaping reversibility? Long-term deep brain stimulation in dystonia: the relationship between effects on electrophysiology and clinical symptoms. Brain. 2011;134:2106–15.

    Article  PubMed  Google Scholar 

  33. Udupa K, Bahl N, Ni Z, Gunraj C, Mazzella F, Moro E, et al. Cortical plasticity induction by pairing subthalamic nucleus deep-brain stimulation and primary motor cortical transcranial magnetic stimulation in Parkinson’s disease. J Neurosci. 2016;36(2):396–404.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Emiliani V, Entcheva E, Hedrich R, Hegemann P, Konrad KR, LĂźscher C, et al. Optogenetics for light control of biological systems. Nat Rev Methods Primers. 2022;2:56.

    Article  Google Scholar 

  35. McGregor MM, Nelson AB. Circuit mechanisms of Parkinson’s disease. Neuron. 2019;101(6):1042–56.

    Article  CAS  PubMed  Google Scholar 

  36. Benabid AL, Pollak P, Gervason C, Hoffmann D, Gao DM, Hommel M, et al. Long-term suppression of tremor by chronic stimulation of the ventral intermediate thalamic nucleus. Lancet. 1991;337(8738):403–6.

    Article  CAS  PubMed  Google Scholar 

  37. Neudorfer C, Hinzke M, Hunsche S, El Majdoub F, Lozano A, Maarouf M. Combined deep brain stimulation of subthalamic nucleus and ventral intermediate thalamic nucleus in tremor-dominant Parkinson’s disease using a parietal approach. Neuromodulation. 2019;22(4):493–502.

    Article  PubMed  Google Scholar 

  38. Fayed I, Cobourn KD, Pivazyan G, Torres-Yaghi YA, Pagan FL, Lo SE, et al. Combination targeting of subthalamic nucleus and ventral intermediate thalamic nucleus with a single trajectory in deep brain stimulation for tremor-dominant Parkinson’s disease. J Clin Neurosci. 2021;85:92–100.

    Article  PubMed  Google Scholar 

  39. Milosevic L, Kalia SK, Hodaie M, Lozano AM, Popovic MR, Hutchison WD. Physiological mechanisms of thalamic ventral intermediate nucleus stimulation for tremor suppression. Brain. 2018;141(7):2142–55.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Boraud T, Bezard E, Bioulac B, Gross C. High frequency stimulation of the internal Globus Pallidus (GPi) simultaneously improves parkinsonian symptoms and reduces the firing frequency of GPi neurons in the MPTP-treated monkey. Neurosci Lett. 1996;215(1):17–20.

    Article  CAS  PubMed  Google Scholar 

  41. Benabid AL. Deep brain stimulation for Parkinson’s disease. Curr Opin Neurobiol. 2003;13(6):696–706.

    Article  CAS  PubMed  Google Scholar 

  42. Benazzouz A, Gao DM, Ni ZG, Piallat B, Bouali-Benazzouz R, Benabid AL. Effect of high-frequency stimulation of the subthalamic nucleus on the neuronal activities of the substantia nigra pars reticulata and ventrolateral nucleus of the thalamus in the rat. Neuroscience. 2000;99(2):289–95.

    Article  CAS  PubMed  Google Scholar 

  43. Dostrovsky JO, Levy R, Wu JP, Hutchison WD, Tasker RR, Lozano AM. Microstimulation-induced inhibition of neuronal firing in human globus pallidus. J Neurophysiol. 2000;84(1):570–4.

    Article  CAS  PubMed  Google Scholar 

  44. Milosevic L, Kalia SK, Hodaie M, Lozano AM, Fasano A, Popovic MR, Hutchison WD. Neuronal inhibition and synaptic plasticity of basal ganglia neurons in Parkinson’s disease. Brain. 2018;141:177–90.

    Article  PubMed  Google Scholar 

  45. Schor JS, Gonzalez Montalvo I, Spratt PWE, Brakaj RJ, Stansil JA, Twedell EL, et al. Therapeutic deep brain stimulation disrupts movement-related subthalamic nucleus activity in parkinsonian mice. Elife. 2022;11:e75253.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Luo F, Kim LH, Magown P, Noor MS, Kiss ZHT. Long-lasting electrophysiological after-effects of high-frequency stimulation in the globus pallidus: human and rodent slice studies. J Neurosci. 2018;38(50):10734–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. McCairn KW, Turner RS. Deep brain stimulation of the globus pallidus internus in the parkinsonian primate: local entrainment and suppression of low-frequency oscillations. J Neurophysiol. 2009;101(4):1941–60.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Milosevic L, Kalia SK, Hodaie M, Lozano AM, Popovic MR, Hutchison WD, Lankarany M. A theoretical framework for the site-specific and frequency-dependent neuronal effects of deep brain stimulation. Brain Stimul. 2021;14(4):807–21.

    Article  PubMed  Google Scholar 

  49. Bower KL, McIntyre CC. Deep brain stimulation of terminating axons. Brain Stimul. 2020;13(6):1863–70.

    Article  PubMed  PubMed Central  Google Scholar 

  50. Iremonger KJ, Anderson TR, Hu B, Kiss ZH. Cellular mechanisms preventing sustained activation of cortex during subcortical high-frequency stimulation. J Neurophysiol. 2006;96(2):613–21.

    Article  PubMed  Google Scholar 

  51. Neumann WJ, Steiner LA, Milosevic L. Neurophysiological mechanisms of deep brain stimulation across spatiotemporal resolutions. Brain. 2023;146(11):4456–68.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Xiao C, Miwa JM, Henderson BJ, Wang Y, Deshpande P, McKinney SL, Lester HA. Nicotinic receptor subtype-selective circuit patterns in the subthalamic nucleus. J Neurosci. 2015;35(9):3734–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. von Gersdorff H, Borst JG. Short-term plasticity at the calyx of Held. Nat Rev Neurosci. 2002;3(1):53–64.

    Article  Google Scholar 

  54. Yamawaki N, Magill PJ, Woodhall GL, Hall SD, Stanford IM. Frequency selectivity and dopamine-dependence of plasticity at glutamatergic synapses in the subthalamic nucleus. Neuroscience. 2012;203:1–11.

    Article  CAS  PubMed  Google Scholar 

  55. Rosenmund C, Stevens CF. Definition of the readily releasable pool of vesicles at hippocampal synapses. Neuron. 1996;16(6):1197–207.

    Article  CAS  PubMed  Google Scholar 

  56. Dittman JS, Regehr WG. Calcium dependence and recovery kinetics of presynaptic depression at the climbing fiber to Purkinje cell synapse. J Neurosci. 1998;18(16):6147–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Zucker RS, Regehr WG. Short-term synaptic plasticity. Annu Rev Physiol. 2002;64:355–405.

    Article  CAS  PubMed  Google Scholar 

  58. Rizzoli SO, Betz WJ. Synaptic vesicle pools. Nat Rev Neurosci. 2005;6(1):57–69.

    Article  CAS  PubMed  Google Scholar 

  59. Fioravante D, Regehr WG. Short-term forms of presynaptic plasticity. Curr Opin Neurobiol. 2011;21(2):269–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Beurrier C, Bioulac B, Audin J, Hammond C. High-frequency stimulation produces a transient blockade of voltage-gated currents in subthalamic neurons. J Neurophysiol. 2001;85(4):1351–6.

    Article  CAS  PubMed  Google Scholar 

  61. Tai CH, Pan MK, Tseng SH, Wang TR, Kuo CC. Hyperpolarization of the subthalamic nucleus alleviates hyperkinetic movement disorders. Sci Rep. 2020;10(1):8278.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Tai CH, Pan MK, Lin JJ, Huang CS, Yang YC, Kuo CC. Subthalamic discharges as a causal determinant of parkinsonian motor deficits. Ann Neurol. 2012;72(3):464–76.

    Article  PubMed  Google Scholar 

  63. Beurrier C, Congar P, Bioulac B, Hammond C. Subthalamic nucleus neurons switch from single-spike activity to burst-firing mode. J Neurosci. 1999;19(2):599–609.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Gajendiran M, Cepeda C, Kha HT, Sison JD, Levine MS. Electrophysiological alterations in subthalamic neurons after unilateral dopamine depletion in the rat. J Neurosci Res. 2005;80(2):203–10.

    Article  CAS  PubMed  Google Scholar 

  65. Ammari R, Bioulac B, Garcia L, Hammond C. The subthalamic nucleus becomes a generator of bursts in the dopamine-depleted state. Its high frequency stimulation dramatically weakens transmission to the globus pallidus. Front Syst Neurosci. 2011;5:43.

    Article  PubMed  PubMed Central  Google Scholar 

  66. Pan MK, Tai CH, Liu WC, Pei JC, Lai WS, Kuo CC. Deranged NMDAergic cortico-subthalamic transmission underlies parkinsonian motor deficits. J Clin Invest. 2014;124(10):4629–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Sharott A, Gulberti A, Zittel S, Tudor Jones AA, Fickel U, Munchau A, et al. Activity parameters of subthalamic nucleus neurons selectively predict motor symptom severity in Parkinson’s disease. J Neurosci. 2014;34(18):6273–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Georgiades MJ, Shine JM, Gilat M, McMaster J, Owler B, Mahant N, Lewis SJG. Hitting the brakes: pathological subthalamic nucleus activity in Parkinson’s disease gait freezing. Brain. 2019;142(12):3906–16.

    Article  PubMed  Google Scholar 

  69. Tai CH. Subthalamic burst firing: a pathophysiological target in Parkinson’s disease. Neurosci Biobehav Rev. 2022;132:410–9.

    Article  PubMed  Google Scholar 

  70. Scherer M, Steiner LA, Kalia SK, Hodaie M, Kühn AA, Lozano AM, et al. Single-neuron bursts encode pathological oscillations in subcortical nuclei of patients with Parkinson’s disease and essential tremor. Proc Natl Acad Sci U S A. 2022;119(35):e2205881119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Milosevic L, Kalia SK, Hodaie M, Lozano A, Popovic MR, Hutchison W. Subthalamic suppression defines therapeutic threshold of deep brain stimulation in Parkinson’s disease. J Neurol Neurosurg Psychiatry. 2019;90(10):1105–8.

    Article  PubMed  Google Scholar 

  72. Tai CH, Boraud T, Bezard E, Bioulac B, Gross C, Benazzouz A. Electrophysiological and metabolic evidence that high-frequency stimulation of the subthalamic nucleus bridles neuronal activity in the subthalamic nucleus and the substantia nigra reticulata. FASEB J. 2003;17(13):1820–30.

    Article  CAS  PubMed  Google Scholar 

  73. Tai CH, Yang YC, Pan MK, Huang CS, Kuo CC. Modulation of subthalamic T-type Ca(2+) channels remedies locomotor deficits in a rat model of Parkinson disease. J Clin Invest. 2011;121(8):3289–305.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Zhuang QX, Li GY, Li B, Zhang CZ, Zhang XY, Xi K, et al. Regularizing firing patterns of rat subthalamic neurons ameliorates parkinsonian motor deficits. J Clin Invest. 2018;128(12):5413–27.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Pan MK, Kuo SH, Tai CH, Liou JY, Pei JC, Chang CY, et al. Neuronal firing patterns outweigh circuitry oscillations in parkinsonian motor control. J Clin Invest. 2016;126(12):4516–26.

    Article  PubMed  PubMed Central  Google Scholar 

  76. Huang CS, Wang GH, Tai CH, Hu CC, Yang YC. Antiarrhythmics cure brain arrhythmia: the imperativeness of subthalamic ERG K(+) channels in parkinsonian discharges. Sci Adv. 2017;3(5):e1602272.

    Article  PubMed  PubMed Central  Google Scholar 

  77. Cleary DR, Raslan AM, Rubin JE, Bahgat D, Viswanathan A, Heinricher MM, Burchiel KJ. Deep brain stimulation entrains local neuronal firing in human globus pallidus internus. J Neurophysiol. 2013;109(4):978–87.

    Article  PubMed  Google Scholar 

  78. Bar-Gad I, Elias S, Vaadia E, Bergman H. Complex locking rather than complete cessation of neuronal activity in the globus pallidus of a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated primate in response to pallidal microstimulation. J Neurosci. 2004;24(33):7410–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Debanne D, Poo MM. Spike-timing dependent plasticity beyond synapse - pre- and post-synaptic plasticity of intrinsic neuronal excitability. Front Synaptic Neurosci. 2010;2:21.

    PubMed  PubMed Central  Google Scholar 

  80. Mozzachiodi R, Byrne JH. More than synaptic plasticity: role of nonsynaptic plasticity in learning and memory. Trends Neurosci. 2010;33(1):17–26.

    Article  CAS  PubMed  Google Scholar 

  81. Hashimoto T, Elder CM, Okun MS, Patrick SK, Vitek JL. Stimulation of the subthalamic nucleus changes the firing pattern of pallidal neurons. J Neurosci. 2003;23(5):1916–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Valverde S, Vandecasteele M, Piette C, Derousseaux W, Gangarossa G, Aristieta Arbelaiz A, et al. Deep brain stimulation-guided optogenetic rescue of parkinsonian symptoms. Nat Commun. 2020;11(1):2388.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. McIntyre CC, Savasta M, Kerkerian-Le Goff L, Vitek JL. Uncovering the mechanism(s) of action of deep brain stimulation: activation, inhibition, or both. Clin Neurophysiol. 2004;115(6):1239–48.

    Article  PubMed  Google Scholar 

  84. McIntyre CC, Grill WM, Sherman DL, Thakor NV. Cellular effects of deep brain stimulation: model-based analysis of activation and inhibition. J Neurophysiol. 2004;91(4):1457–69.

    Article  PubMed  Google Scholar 

  85. Brocker DT, Grill WM. Principles of electrical stimulation of neural tissue. Handb Clin Neurol. 2013;116:3–18.

    Article  PubMed  Google Scholar 

  86. Histed MH, Bonin V, Reid RC. Direct activation of sparse, distributed populations of cortical neurons by electrical microstimulation. Neuron. 2009;63(4):508–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Anderson RW, Farokhniaee A, Gunalan K, Howell B, McIntyre CC. Action potential initiation, propagation, and cortical invasion in the hyperdirect pathway during subthalamic deep brain stimulation. Brain Stimul. 2018;11(5):1140–50.

    Article  PubMed  PubMed Central  Google Scholar 

  88. Li Q, Ke Y, Chan DC, Qian ZM, Yung KK, Ko H, et al. Therapeutic deep brain stimulation in Parkinsonian rats directly influences motor cortex. Neuron. 2012;76(5):1030–41.

    Article  CAS  PubMed  Google Scholar 

  89. Finkelstein DI, Walker RH, Moore C, Davies G, Dirling LB, Koch RJ, Meshul CK. Effects of subthalamic nucleus lesions and stimulation upon corticostriatal afferents in the 6-hydroxydopamine-lesioned rat. PLoS One. 2012;7(3):e32919.

    Article  Google Scholar 

  90. Kahan J, Urner M, Moran R, Flandin G, Marreiros A, Mancini L, et al. Resting state functional MRI in Parkinson’s disease: the impact of deep brain stimulation on ‘effective’ connectivity. Brain. 2014;137(4):1130–44.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Liu LD, Prescott IA, Dostrovsky JO, Hodaie M, Lozano AM, Hutchison WD. Frequency-dependent effects of electrical stimulation in the globus pallidus of dystonia patients. J Neurophysiol. 2012;108(1):5–17.

    Article  PubMed  Google Scholar 

  92. Pierantozzi M, Palmieri MG, Mazzone P, Marciani MG, Rossini PM, Stefani A, et al. Deep brain stimulation of both subthalamic nucleus and internal globus pallidus restores intracortical inhibition in Parkinson’s disease paralleling apomorphine effects: a paired magnetic stimulation study. Clin Neurophysiol. 2002;113(1):108–13.

    Article  CAS  PubMed  Google Scholar 

  93. Xiao C, Ji YW, Luan YW, Jia T, Yin C, Zhou CY. Differential modulation of subthalamic projection neurons by short-term and long-term electrical stimulation in physiological and parkinsonian conditions. Acta Pharmacol Sin. 2022;43(8):1928–39.

    Article  CAS  PubMed  Google Scholar 

  94. Benazzouz A, Piallat B, Pollak P, Benabid AL. Responses of substantia nigra pars reticulata and globus pallidus complex to high frequency stimulation of the subthalamic nucleus in rats: electrophysiological data. Neurosci Lett. 1995;189(2):77–80.

    Article  CAS  PubMed  Google Scholar 

  95. Chiken S, Nambu A. Disrupting neuronal transmission: mechanism of DBS? Front Syst Neurosci. 2014;8:33.

    Article  PubMed  PubMed Central  Google Scholar 

  96. Bosch C, Degos B, Deniau JM, Venance L. Subthalamic nucleus high-frequency stimulation generates a concomitant synaptic excitation-inhibition in substantia nigra pars reticulata. J Physiol. 2011;589(17):4189–207.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Beiske AG, Loge JH, Rønningen A, Svensson E. Pain in Parkinson’s disease: prevalence and characteristics. Pain. 2009;141(1–2):173–7.

    Article  CAS  PubMed  Google Scholar 

  98. Luan Y, Tang D, Wu H, Gu W, Wu Y, Cao JL, et al. Reversal of hyperactive subthalamic circuits differentially mitigates pain hypersensitivity phenotypes in parkinsonian mice. Proc Natl Acad Sci U S A. 2020;117(18):10045–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Bahners BH, Waterstraat G, Kannenberg S, Curio G, Schnitzler A, Nikulin V, Florin E. Electrophysiological characterization of the hyperdirect pathway and its functional relevance for subthalamic deep brain stimulation. Exp Neurol. 2022;352:114031.

    Article  CAS  PubMed  Google Scholar 

  100. Jorge A, Lipski WJ, Wang D, Crammond DJ, Turner RS, Richardson RM. Hyperdirect connectivity of opercular speech network to the subthalamic nucleus. Cell Rep. 2022;38(10):110477.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Sitti I, Acar G, Zisakis AK, Ozdemir M, Acar F, Burchiel KJ. Effect of subthalamic nucleus stimulation on pedunculopontine nucleus neural activity. Stereotact Funct Neurosurg. 2016;94(1):54–9.

    Article  PubMed  Google Scholar 

  102. Wilson CL, Puntis M, Lacey MG. Overwhelmingly asynchronous firing of rat subthalamic nucleus neurones in brain slices provides little evidence for intrinsic interconnectivity. Neuroscience. 2004;123(1):187–200.

    Article  CAS  PubMed  Google Scholar 

  103. Atherton JF, Kitano K, Baufreton J, Fan K, Wokosin D, Tkatch T, et al. Selective participation of somatodendritic HCN channels in inhibitory but not excitatory synaptic integration in neurons of the subthalamic nucleus. J Neurosci. 2010;30(47):16025–40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Bevan MD, Hallworth NE, Baufreton J. GABAergic control of the subthalamic nucleus. Prog Brain Res. 2007;160:173–88.

    Article  CAS  PubMed  Google Scholar 

  105. Wang L, Kitai ST, Xiang Z. Activity-dependent bidirectional modification of inhibitory synaptic transmission in rat subthalamic neurons. J Neurosci. 2006;26(28):7321–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Li S, Arbuthnott GW, Jutras MJ, Goldberg JA, Jaeger D. Resonant antidromic cortical circuit activation as a consequence of high-frequency subthalamic deep-brain stimulation. J Neurophysiol. 2007;98(6):3525–37.

    Article  CAS  PubMed  Google Scholar 

  107. Gradinaru V, Mogri M, Thompson KR, Henderson JM, Deisseroth K. Optical deconstruction of parkinsonian neural circuitry. Science. 2009;324(5925):354–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Sanders TH, Jaeger D. Optogenetic stimulation of cortico-subthalamic projections is sufficient to ameliorate bradykinesia in 6-ohda lesioned mice. Neurobiol Dis. 2016;95:225–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Kita T, Kita H. The subthalamic nucleus is one of multiple innervation sites for long-range corticofugal axons: a single-axon tracing study in the rat. J Neurosci. 2012;32(17):5990–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Johnson LA, Wang J, Nebeck SD, Zhang J, Johnson MD, Vitek JL. Direct activation of primary motor cortex during subthalamic but not pallidal deep brain stimulation. J Neurosci. 2020;40(10):2166–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Chu HY, Atherton JF, Wokosin D, Surmeier DJ, Bevan MD. Heterosynaptic regulation of external globus pallidus inputs to the subthalamic nucleus by the motor cortex. Neuron. 2015;85(2):364–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Xu W, Russo GS, Hashimoto T, Zhang J, Vitek JL. Subthalamic nucleus stimulation modulates thalamic neuronal activity. J Neurosci. 2008;28(46):11916–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Anderson ME, Postupna N, Ruffo M. Effects of high-frequency stimulation in the internal globus pallidus on the activity of thalamic neurons in the awake monkey. J Neurophysiol. 2003;89(2):1150–60.

    Article  PubMed  Google Scholar 

  114. Hunnicutt BJ, Jongbloets BC, Birdsong WT, Gertz KJ, Zhong H, Mao T. A comprehensive excitatory input map of the striatum reveals novel functional organization. Elife. 2016;5:e19103.

    Article  PubMed  PubMed Central  Google Scholar 

  115. Smith Y, Galvan A, Ellender TJ, Doig N, Villalba RM, Huerta-Ocampo I, et al. The thalamostriatal system in normal and diseased states. Front Syst Neurosci. 2014;8:5.

    Article  PubMed  PubMed Central  Google Scholar 

  116. Brown HD, Baker PM, Ragozzino ME. The parafascicular thalamic nucleus concomitantly influences behavioral flexibility and dorsomedial striatal acetylcholine output in rats. J Neurosci. 2010;30(43):14390–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Tanimura A, Du Y, Kondapalli J, Wokosin DL, Surmeier DJ. Cholinergic interneurons amplify thalamostriatal excitation of striatal indirect pathway neurons in Parkinson’s disease models. Neuron. 2019;101(3):444-458.e446.

    Article  CAS  PubMed  Google Scholar 

  118. Kita T, Shigematsu N, Kita H. Intralaminar and tectal projections to the subthalamus in the rat. Eur J Neurosci. 2016;44(11):2899–908.

    Article  PubMed  PubMed Central  Google Scholar 

  119. Watson GDR, Hughes RN, Petter EA, Fallon IP, Kim N, Severino FPU, Yin HH. Thalamic projections to the subthalamic nucleus contribute to movement initiation and rescue of parkinsonian symptoms. Sci Adv. 2021;7(6):eabe9192.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Zhang Y, Roy DS, Zhu Y, Chen Y, Aida T, Hou Y, et al. Targeting thalamic circuits rescues motor and mood deficits in PD mice. Nature. 2022;607(7918):321–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Johnson KA, Cagle JN, Lopes JL, Wong JK, Okun MS, Gunduz A, et al. Globus pallidus internus deep brain stimulation evokes resonant neural activity in Parkinson’s disease. Brain Commun. 2023;5(2):fcad025.

    Article  PubMed  PubMed Central  Google Scholar 

  122. Steiner LA, Milosevic L. A convergent subcortical signature to explain the common efficacy of subthalamic and pallidal deep brain stimulation. Brain Commun. 2023;5(2):fcad033.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Wiest C, Tinkhauser G, Pogosyan A, Bange M, Muthuraman M, Groppa S, et al. Local field potential activity dynamics in response to deep brain stimulation of the subthalamic nucleus in Parkinson’s disease. Neurobiol Dis. 2020;143:105019.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Wiest C, He S, Duchet B, Pogosyan A, Benjaber M, Denison T, et al. Evoked resonant neural activity in subthalamic local field potentials reflects basal ganglia network dynamics. Neurobiol Dis. 2023;178:106019.

    Article  PubMed  PubMed Central  Google Scholar 

  125. Schmidt SL, Brocker DT, Swan BD, Turner DA, Grill WM. Evoked potentials reveal neural circuits engaged by human deep brain stimulation. Brain Stimul. 2020;13(6):1706–18.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Ozturk M, Viswanathan A, Sheth SA, Ince NF. Electroceutically induced subthalamic high-frequency oscillations and evoked compound activity may explain the mechanism of therapeutic stimulation in Parkinson’s disease. Commun Biol. 2021;4(1):393.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Sinclair NC, McDermott HJ, Fallon JB, Perera T, Brown P, Bulluss KJ, Thevathasan W. Deep brain stimulation for Parkinson’s disease modulates high-frequency evoked and spontaneous neural activity. Neurobiol Dis. 2019;130:104522.

    Article  PubMed  PubMed Central  Google Scholar 

  128. Xu SS, Lee WL, Perera T, Sinclair NC, Bulluss KJ, McDermott HJ, Thevathasan W. Can brain signals and anatomy refine contact choice for deep brain stimulation in Parkinson’s disease? J Neurol Neurosurg Psychiatry. 2022. https://doi.org/10.1136/jnnp-2021-327708.

    Article  PubMed  Google Scholar 

  129. Sobesky L, Goede L, Odekerken VJJ, Wang Q, Li N, Neudorfer C, et al. Subthalamic and pallidal deep brain stimulation: are we modulating the same network? Brain. 2022;145(1):251–62.

    Article  PubMed  Google Scholar 

  130. Wallace ML, Saunders A, Huang KW, Philson AC, Goldman M, Macosko EZ, et al. Genetically distinct parallel pathways in the entopeduncular nucleus for limbic and sensorimotor output of the basal ganglia. Neuron. 2017;94(1):138-152 e135.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Calabresi P, Picconi B, Tozzi A, Ghiglieri V, Di Filippo M. Direct and indirect pathways of basal ganglia: a critical reappraisal. Nat Neurosci. 2014;17(8):1022–30.

    Article  CAS  PubMed  Google Scholar 

  132. Wichmann T, Dostrovsky JO. Pathological basal ganglia activity in movement disorders. Neuroscience. 2011;198:232–44.

    Article  CAS  PubMed  Google Scholar 

  133. Heimer G, Bar-Gad I, Goldberg JA, Bergman H. Dopamine replacement therapy reverses abnormal synchronization of pallidal neurons in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine primate model of parkinsonism. J Neurosci. 2002;22(18):7850–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Stamatakis AM, Stuber GD. Activation of lateral habenula inputs to the ventral midbrain promotes behavioral avoidance. Nat Neurosci. 2012;15(8):1105–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Hong S, Hikosaka O. The globus pallidus sends reward-related signals to the lateral habenula. Neuron. 2008;60(4):720–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Plantinga BR, Temel Y, Duchin Y, Uludağ K, Patriat R, Roebroeck A, et al. Individualized parcellation of the subthalamic nucleus in patients with Parkinson’s disease with 7T MRI. Neuroimage. 2018;168:403–11.

    Article  PubMed  Google Scholar 

  137. Zhang F, Wang F, Li W, Wang N, Han C, Fan S, et al. Relationship between electrode position of deep brain stimulation and motor symptoms of Parkinson’s disease. BMC Neurol. 2021;21(1):122.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Herzog J, Fietzek U, Hamel W, Morsnowski A, Steigerwald F, Schrader B, et al. Most effective stimulation site in subthalamic deep brain stimulation for Parkinson’s disease. Mov Disord. 2004;19(9):1050–4.

    Article  PubMed  Google Scholar 

  139. Schor JS, Nelson AB. Multiple stimulation parameters influence efficacy of deep brain stimulation in parkinsonian mice. J Clin Invest. 2019;129(9):3833–8.

    Article  PubMed  PubMed Central  Google Scholar 

  140. Muller J, Alizadeh M, Mohamed FB, Riley J, Pearce JJ, Trieu B, et al. Clinically applicable delineation of the pallidal sensorimotor region in patients with advanced Parkinson’s disease: study of probabilistic and deterministic tractography. J Neurosurg. 2019;131:1520–31. https://thejns.org/view/journals/j-neurosurg/131/5/article-p1520.xml

  141. Nowacki A, Fiechter M, Fichtner J, Debove I, Lachenmayer L, SchĂźpbach M, et al. Using MDEFT MRI sequences to target the GPi in DBS surgery. PLoS One. 2015;10(9):e0137868.

    Article  PubMed  PubMed Central  Google Scholar 

  142. Gut NK, Winn P. Deep brain stimulation of different pedunculopontine targets in a novel rodent model of parkinsonism. J Neurosci. 2015;35(12):4792–803.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Thevathasan W, Debu B, Aziz T, Bloem BR, Blahak C, Butson C, et al. Pedunculopontine nucleus deep brain stimulation in Parkinson’s disease: a clinical review. Mov Disord. 2018;33(1):10–20.

    Article  PubMed  Google Scholar 

  144. Thevathasan W, Coyne TJ, Hyam JA, Kerr G, Jenkinson N, Aziz TZ, Silburn PA. Pedunculopontine nucleus stimulation improves gait freezing in Parkinson disease. Neurosurgery. 2011;69(6):1248–54.

    Article  PubMed  Google Scholar 

  145. Takakusaki K, Chiba R, Nozu T, Okumura T. Brainstem control of locomotion and muscle tone with special reference to the role of the mesopontine tegmentum and medullary reticulospinal systems. J Neural Transm (Vienna). 2016;123(7):695–729.

    Article  PubMed  Google Scholar 

  146. Karachi C, Francois C, Parain K, Bardinet E, Tande D, Hirsch E, Yelnik J. Three-dimensional cartography of functional territories in the human striatopallidal complex by using calbindin immunoreactivity. J Comp Neurol. 2002;450(2):122–34.

    Article  PubMed  Google Scholar 

  147. Lambert C, Zrinzo L, Nagy Z, Lutti A, Hariz M, Foltynie T, et al. Confirmation of functional zones within the human subthalamic nucleus: patterns of connectivity and sub-parcellation using diffusion weighted imaging. Neuroimage. 2012;60(1):83–94.

    Article  PubMed  Google Scholar 

  148. Haynes WI, Haber SN. The organization of prefrontal-subthalamic inputs in primates provides an anatomical substrate for both functional specificity and integration: implications for Basal Ganglia models and deep brain stimulation. J Neurosci. 2013;33(11):4804–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Rodriguez-Oroz MC, Rodriguez M, Guridi J, Mewes K, Chockkman V, Vitek J, et al. The subthalamic nucleus in Parkinson’s disease: somatotopic organization and physiological characteristics. Brain. 2001;124(Pt 9):1777–90.

    Article  CAS  PubMed  Google Scholar 

  150. Jeon H, Lee H, Kwon DH, Kim J, Tanaka-Yamamoto K, Yook JS, et al. Topographic connectivity and cellular profiling reveal detailed input pathways and functionally distinct cell types in the subthalamic nucleus. Cell Rep. 2022;38(9):110439.

    Article  CAS  PubMed  Google Scholar 

  151. Kita T, Osten P, Kita H. Rat subthalamic nucleus and zona incerta share extensively overlapped representations of cortical functional territories. J Comp Neurol. 2014;522(18):4043–56.

    Article  PubMed  PubMed Central  Google Scholar 

  152. Bertino S, Basile GA, Bramanti A, Anastasi GP, Quartarone A, Milardi D, Cacciola A. Spatially coherent and topographically organized pathways of the human globus pallidus. Hum Brain Mapp. 2020;41(16):4641–61.

    Article  PubMed  PubMed Central  Google Scholar 

  153. Williams NR, Foote KD, Okun MS. STN vs. GPi deep brain stimulation: translating the rematch into clinical practice. Mov Disord Clin Pract. 2014;1(1):24–35.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Kaku H, Ozturk M, Viswanathan A, Shahed J, Sheth SA, Kumar S, Ince NF. Unsupervised clustering reveals spatially varying single neuronal firing patterns in the subthalamic nucleus of patients with Parkinson’s disease. Clin Park Relat Disord. 2020;3:100032.

    PubMed  Google Scholar 

  155. Roseberry TK, Lee AM, Lalive AL, Wilbrecht L, Bonci A, Kreitzer AC. Cell-type-specific control of brainstem locomotor circuits by basal ganglia. Cell. 2016;164(3):526–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Ferreira-Pinto MJ, Kanodia H, Falasconi A, Sigrist M, Esposito MS, Arber S. Functional diversity for body actions in the mesencephalic locomotor region. Cell. 2021;184(17):4564-4578 e4518.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. French IT, Muthusamy KA. A review of the pedunculopontine nucleus in Parkinson’s disease. Front Aging Neurosci. 2018;10:99.

    Article  PubMed  PubMed Central  Google Scholar 

  158. Vitale F, Capozzo A, Mazzone P, Scarnati E. Neurophysiology of the pedunculopontine tegmental nucleus. Neurobiol Dis. 2019;128:19–30.

    Article  CAS  PubMed  Google Scholar 

  159. Winn P. How best to consider the structure and function of the pedunculopontine tegmental nucleus: evidence from animal studies. J Neurol Sci. 2006;248(1–2):234–50.

    Article  PubMed  Google Scholar 

  160. Masini D, Kiehn O. Targeted activation of midbrain neurons restores locomotor function in mouse models of parkinsonism. Nat Commun. 2022;13(1):504.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Wang HL, Morales M. Pedunculopontine and laterodorsal tegmental nuclei contain distinct populations of cholinergic, glutamatergic and GABAergic neurons in the rat. Eur J Neurosci. 2009;29(2):340–58.

    Article  PubMed  Google Scholar 

  162. Carvalho MM, Tanke N, Kropff E, Witter MP, Moser MB, Moser EI. A brainstem locomotor circuit drives the activity of speed cells in the medial entorhinal cortex. Cell Rep. 2020;32(10):108123.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Caggiano V, Leiras R, Goñi-Erro H, Masini D, Bellardita C, Bouvier J, et al. Midbrain circuits that set locomotor speed and gait selection. Nature. 2018;553(7689):455–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Picillo M, Lozano AM, Kou N, Puppi Munhoz R, Fasano A. Programming deep brain stimulation for Parkinson’s disease: the Toronto Western Hospital algorithms. Brain Stimul. 2016;9(3):425–37.

    Article  PubMed  Google Scholar 

  165. Moro E, Esselink RJ, Xie J, Hommel M, Benabid AL, Pollak P. The impact on Parkinson’s disease of electrical parameter settings in STN stimulation. Neurology. 2002;59(5):706–13.

    Article  CAS  PubMed  Google Scholar 

  166. Moon HC, Won SY, Kim EG, Kim HK, Cho CB, Park YS. Effect of optogenetic modulation on entopeduncular input affects thalamic discharge and behavior in an AAV2-α-synuclein-induced hemiparkinson rat model. Neurosci Lett. 2018;662:129–35.

    Article  CAS  PubMed  Google Scholar 

  167. Yoon HH, Nam M-H, Choi I, Min J, Jeon SR. Optogenetic inactivation of the entopeduncular nucleus improves forelimb akinesia in a Parkinson’s disease model. Behav Brain Res. 2020;386:112551.

    Article  CAS  PubMed  Google Scholar 

  168. Yoon HH, Park JH, Kim YH, Min J, Hwang E, Lee CJ, et al. Optogenetic inactivation of the subthalamic nucleus improves forelimb akinesia in a rat model of Parkinson disease. Neurosurgery. 2014;74(5):533–41.

    Article  PubMed  Google Scholar 

  169. Yoon HH, Min J, Hwang E, Lee CJ, Suh JK, Hwang O, Jeon SR. Optogenetic inhibition of the subthalamic nucleus reduces levodopa-induced dyskinesias in a rat model of Parkinson’s disease. Stereotact Funct Neurosurg. 2016;94(1):41–53.

    Article  PubMed  Google Scholar 

  170. Parolari L, Schneeberger M, Heintz N, Friedman JM. Functional analysis of distinct populations of subthalamic nucleus neurons on Parkinson’s disease and OCD-like behaviors in mice. Mol Psychiatry. 2021;26(11):7029–46.

    Article  CAS  PubMed  Google Scholar 

  171. Jun NY, Cardin JA. Activation of distinct channelrhodopsin variants engages different patterns of network activity. eNeuro. 2020;7(1):ENEURO.0222-0218.2019.

    Article  Google Scholar 

  172. Yizhar O, Fenno LE, Davidson TJ, Mogri M, Deisseroth K. Optogenetics in neural systems. Neuron. 2011;71(1):9–34.

    Article  CAS  PubMed  Google Scholar 

  173. Yu C, Cassar IR, Sambangi J, Grill WM. Frequency-specific optogenetic deep brain stimulation of subthalamic nucleus improves parkinsonian motor behaviors. J Neurosci. 2020;40(22):4323–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Klapoetke NC, Murata Y, Kim SS, Pulver SR, Birdsey-Benson A, Cho YK, et al. Independent optical excitation of distinct neural populations. Nat Methods. 2014;11(3):338–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Llewellyn ME, Thompson KR, Deisseroth K, Delp SL. Orderly recruitment of motor units under optical control in vivo. Nat Med. 2010;16(10):1161–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Denker A, Rizzoli SO. Synaptic vesicle pools: an update. Front Synaptic Neurosci. 2010;2:135.

    PubMed  PubMed Central  Google Scholar 

  177. Dutar P, Nicoll RA. A physiological role for GABAB receptors in the central nervous system. Nature. 1988;332(6160):156–8.

    Article  CAS  PubMed  Google Scholar 

  178. Mintz IM, Bean BP. GABAB receptor inhibition of P-type Ca2+ channels in central neurons. Neuron. 1993;10(5):889–98.

    Article  CAS  PubMed  Google Scholar 

  179. Neher E, Sakaba T. Multiple roles of calcium ions in the regulation of neurotransmitter release. Neuron. 2008;59(6):861–72.

    Article  CAS  PubMed  Google Scholar 

  180. McIntyre CC, Mori S, Sherman DL, Thakor NV, Vitek JL. Electric field and stimulating influence generated by deep brain stimulation of the subthalamic nucleus. Clin Neurophysiol. 2004;115(3):589–95.

    Article  PubMed  Google Scholar 

  181. Butson CR, Cooper SE, Henderson JM, McIntyre CC. Patient-specific analysis of the volume of tissue activated during deep brain stimulation. Neuroimage. 2007;34(2):661–70.

    Article  PubMed  Google Scholar 

  182. Aström M, Tripoliti E, Hariz MI, Zrinzo LU, Martinez-Torres I, Limousin P, Wårdell K. Patient-specific model-based investigation of speech intelligibility and movement during deep brain stimulation. Stereotact Funct Neurosurg. 2010;88(4):224–33.

    Article  PubMed  PubMed Central  Google Scholar 

  183. Lai Y, Song Y, Huang P, Wang T, Wang L, Pan Y, et al. Subthalamic stimulation for camptocormia in Parkinson’s disease: association of volume of tissue activated and structural connectivity with clinical effectiveness. J Parkinsons Dis. 2021;11(1):199–210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Agharazi H, Hardin EC, Flannery K, Beylergil SB, Noecker A, Kilbane C, et al. Physiological measures and anatomical correlates of subthalamic deep brain stimulation effect on gait in Parkinson’s disease. J Neurol Sci. 2023;449:120647.

    Article  PubMed  Google Scholar 

  185. Rizzone M, Lanotte M, Bergamasco B, Tavella A, Torre E, Faccani G, et al. Deep brain stimulation of the subthalamic nucleus in Parkinson’s disease: effects of variation in stimulation parameters. J Neurol Neurosurg Psychiatry. 2001;71(2):215–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Zheng Z, Li Y, Li J, Zhang Y, Zhang X, Zhuang P. Stimulation-induced dyskinesia in the early stage after subthalamic deep brain stimulation. Stereotact Funct Neurosurg. 2010;88(1):29–34.

    Article  PubMed  Google Scholar 

  187. Dayal V, Limousin P, Foltynie T. Subthalamic nucleus deep brain stimulation in Parkinson’s disease: the effect of varying stimulation parameters. J Parkinsons Dis. 2017;7(2):235–45.

    Article  PubMed  PubMed Central  Google Scholar 

  188. Milosevic L, Gramer R, Kim TH, Algarni M, Fasano A, Kalia SK, et al. Modulation of inhibitory plasticity in basal ganglia output nuclei of patients with Parkinson’s disease. Neurobiol Dis. 2019;124:46–56.

    Article  PubMed  Google Scholar 

  189. Temperli P, Ghika J, Villemure JG, Burkhard PR, Bogousslavsky J, Vingerhoets FJ. How do parkinsonian signs return after discontinuation of subthalamic DBS? Neurology. 2003;60(1):78–81.

    Article  CAS  PubMed  Google Scholar 

  190. Tass PA. A model of desynchronizing deep brain stimulation with a demand-controlled coordinated reset of neural subpopulations. Biol Cybern. 2003;89(2):81–8.

    Article  PubMed  Google Scholar 

  191. Tass PA, Qin L, Hauptmann C, Dovero S, Bezard E, Boraud T, Meissner WG. Coordinated reset has sustained aftereffects in Parkinsonian monkeys. Ann Neurol. 2012;72(5):816–20.

    Article  PubMed  Google Scholar 

  192. Adamchic I, Hauptmann C, Barnikol UB, Pawelczyk N, Popovych O, Barnikol TT, et al. Coordinated reset neuromodulation for Parkinson’s disease: proof-of-concept study. Mov Disord. 2014;29(13):1679–84.

    Article  PubMed  PubMed Central  Google Scholar 

  193. Wang J, Nebeck S, Muralidharan A, Johnson MD, Vitek JL, Baker KB. Coordinated reset deep brain stimulation of subthalamic nucleus produces long-lasting, dose-dependent motor improvements in the 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine non-human primate model of parkinsonism. Brain Stimul. 2016;9(4):609–17.

    Article  PubMed  PubMed Central  Google Scholar 

  194. Bosley KM, Luo Z, Amoozegar S, Acedillo K, Nakajima K, Johnson LA, et al. Effect of subthalamic coordinated reset deep brain stimulation on Parkinsonian gait. Front Neuroinform. 2023;17:1185723.

    Article  PubMed  PubMed Central  Google Scholar 

  195. Gerstner W, Kempter R, van Hemmen JL, Wagner H. A neuronal learning rule for sub-millisecond temporal coding. Nature. 1996;383(6595):76–81.

    Article  CAS  PubMed  Google Scholar 

  196. Markram H, Lübke J, Frotscher M, Sakmann B. Regulation of synaptic efficacy by coincidence of postsynaptic APs and EPSPs. Science. 1997;275(5297):213–5.

    Article  CAS  PubMed  Google Scholar 

  197. Ebert M, Hauptmann C, Tass PA. Coordinated reset stimulation in a large-scale model of the STN-GPe circuit. Front Comput Neurosci. 2014;8:154.

    Article  PubMed  PubMed Central  Google Scholar 

  198. Tass PA, Majtanik M. Long-term anti-kindling effects of desynchronizing brain stimulation: a theoretical study. Biol Cybern. 2006;94(1):58–66.

    Article  PubMed  Google Scholar 

  199. Spix TA, Nanivadekar S, Toong N, Kaplow IM, Isett BR, Goksen Y, et al. Population-specific neuromodulation prolongs therapeutic benefits of deep brain stimulation. Science. 2021;374(6564):201–6.

    Article  CAS  PubMed  Google Scholar 

  200. Kita H. Globus pallidus external segment. In: Gaba and the Basal Ganglia - from molecules to systems. 2007; 111–133.

  201. Mastro KJ, Bouchard RS, Holt HA, Gittis AH. Transgenic mouse lines subdivide external segment of the globus pallidus (GPe) neurons and reveal distinct GPe output pathways. J Neurosci. 2014;34(6):2087–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. de la Crompe B, Aristieta A, Leblois A, Elsherbiny S, Boraud T, Mallet NP. The globus pallidus orchestrates abnormal network dynamics in a model of Parkinsonism. Nat Commun. 2020;11(1):1570.

    Article  PubMed  PubMed Central  Google Scholar 

  203. Mastro KJ, Zitelli KT, Willard AM, Leblanc KH, Kravitz AV, Gittis AH. Cell-specific pallidal intervention induces long-lasting motor recovery in dopamine-depleted mice. Nat Neurosci. 2017;20(6):815–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  204. Fasano A, Romito LM, Daniele A, Piano C, Zinno M, Bentivoglio AR, Albanese A. Motor and cognitive outcome in patients with Parkinson’s disease 8 years after subthalamic implants. Brain. 2010;133(9):2664–76.

    Article  PubMed  Google Scholar 

  205. Krauss JK, Lipsman N, Aziz T, Boutet A, Brown P, Chang JW, et al. Technology of deep brain stimulation: current status and future directions. Nat Rev Neurol. 2021;17(2):75–87.

    Article  PubMed  Google Scholar 

  206. Bouthour W, Mégevand P, Donoghue J, Lüscher C, Birbaumer N, Krack P. Biomarkers for closed-loop deep brain stimulation in Parkinson disease and beyond. Nat Rev Neurol. 2019;15(6):343–52.

    Article  PubMed  Google Scholar 

  207. Neumann WJ, Gilron R, Little S, Tinkhauser G. Adaptive deep brain stimulation: from experimental evidence toward practical implementation. Mov Disord. 2023;38(6):937–48.

    Article  PubMed  Google Scholar 

  208. Rosa M, Arlotti M, Ardolino G, Cogiamanian F, Marceglia S, Di Fonzo A, et al. Adaptive deep brain stimulation in a freely moving Parkinsonian patient. Mov Disord. 2015;30(7):1003–5.

    Article  PubMed  PubMed Central  Google Scholar 

  209. Arlotti M, Marceglia S, Foffani G, Volkmann J, Lozano AM, Moro E, et al. Eight-hours adaptive deep brain stimulation in patients with Parkinson disease. Neurology. 2018;90(11):e971–6.

    Article  PubMed  PubMed Central  Google Scholar 

  210. Rosa M, Arlotti M, Marceglia S, Cogiamanian F, Ardolino G, Fonzo AD, et al. Adaptive deep brain stimulation controls levodopa-induced side effects in Parkinsonian patients. Mov Disord. 2017;32(4):628–9.

    Article  PubMed  PubMed Central  Google Scholar 

  211. Little S, Pogosyan A, Neal S, Zavala B, Zrinzo L, Hariz M, et al. Adaptive deep brain stimulation in advanced Parkinson disease. Ann Neurol. 2013;74(3):449–57.

    Article  PubMed  PubMed Central  Google Scholar 

  212. Sasaki F, Oyama G, Sekimoto S, Nuermaimaiti M, Iwamuro H, Shimo Y, et al. Closed-loop programming using external responses for deep brain stimulation in Parkinson’s disease. Parkinsonism Relat Disord. 2021;84:47–51.

    Article  PubMed  Google Scholar 

  213. Kühn AA, Tsui A, Aziz T, Ray N, Brücke C, Kupsch A, et al. Pathological synchronisation in the subthalamic nucleus of patients with Parkinson’s disease relates to both bradykinesia and rigidity. Exp Neurol. 2009;215(2):380–7.

    Article  PubMed  Google Scholar 

  214. Neumann WJ, Staub-Bartelt F, Horn A, Schanda J, Schneider GH, Brown P, Kühn AA. Long term correlation of subthalamic beta band activity with motor impairment in patients with Parkinson’s disease. Clin Neurophysiol. 2017;128(11):2286–91.

    Article  PubMed  PubMed Central  Google Scholar 

  215. Ray NJ, Jenkinson N, Wang S, Holland P, Brittain JS, Joint C, et al. Local field potential beta activity in the subthalamic nucleus of patients with Parkinson’s disease is associated with improvements in bradykinesia after dopamine and deep brain stimulation. Exp Neurol. 2008;213(1):108–13.

    Article  CAS  PubMed  Google Scholar 

  216. Qasim SE, de Hemptinne C, Swann NC, Miocinovic S, Ostrem JL, Starr PA. Electrocorticography reveals beta desynchronization in the basal ganglia-cortical loop during rest tremor in Parkinson’s disease. Neurobiol Dis. 2016;86:177–86.

    Article  PubMed  Google Scholar 

  217. Swann NC, de Hemptinne C, Miocinovic S, Qasim S, Wang SS, Ziman N, et al. Gamma oscillations in the hyperkinetic state detected with chronic human brain recordings in Parkinson’s disease. J Neurosci. 2016;36(24):6445–58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Swann NC, de Hemptinne C, Thompson MC, Miocinovic S, Miller AM, Gilron R, et al. Adaptive deep brain stimulation for Parkinson’s disease using motor cortex sensing. J Neural Eng. 2018;15(4):046006.

    Article  PubMed  PubMed Central  Google Scholar 

  219. Little S, Brown P. What brain signals are suitable for feedback control of deep brain stimulation in Parkinson’s disease? Ann N Y Acad Sci. 2012;1265(1):9–24.

    Article  PubMed  PubMed Central  Google Scholar 

  220. Steigerwald F, Müller L, Johannes S, Matthies C, Volkmann J. Directional deep brain stimulation of the subthalamic nucleus: a pilot study using a novel neurostimulation device. Mov Disord. 2016;31(8):1240–3.

    Article  PubMed  PubMed Central  Google Scholar 

  221. Rammo RA, Ozinga SJ, White A, Nagel SJ, Machado AG, Pallavaram S, et al. Directional stimulation in Parkinson’s disease and essential tremor: the Cleveland Clinic experience. Neuromodulation. 2022;25(6):829–35.

    Article  PubMed  Google Scholar 

  222. Rebelo P, Green AL, Aziz TZ, Kent A, Schafer D, Venkatesan L, Cheeran B. Thalamic directional deep brain stimulation for tremor: spend less, get more. Brain Stimul. 2018;11(3):600–6.

    Article  CAS  PubMed  Google Scholar 

  223. Pollo C, Kaelin-Lang A, Oertel MF, Stieglitz L, Taub E, Fuhr P, et al. Directional deep brain stimulation: an intraoperative double-blind pilot study. Brain. 2014;137(Pt 7):2015–26.

    Article  PubMed  Google Scholar 

  224. Dembek TA, Reker P, Visser-Vandewalle V, Wirths J, Treuer H, Klehr M, et al. Directional DBS increases side-effect thresholds-a prospective, double-blind trial. Mov Disord. 2017;32(10):1380–8.

    Article  PubMed  Google Scholar 

  225. Klug JR, Engelhardt MD, Cadman CN, Li H, Smith JB, Ayala S, et al. Differential inputs to striatal cholinergic and parvalbumin interneurons imply functional distinctions. Elife. 2018;7:e35657.

    Article  PubMed  PubMed Central  Google Scholar 

  226. Mena-Segovia J, Bolam JP, Martinez-Gonzalez C. Topographical organization of the pedunculopontine nucleus. Front Neuroanat. 2011;5:22.

    PubMed  PubMed Central  Google Scholar 

  227. Lafreniere-Roula M, Kim E, Hutchison WD, Lozano AM, Hodaie M, Dostrovsky JO. High-frequency microstimulation in human globus pallidus and substantia nigra. Exp Brain Res. 2010;205(2):251–61.

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank Mr. Ze-Jie Lin for critical comments on the manuscript.

Funding

This work was supported by grants from the STI2030-Major Projects (2022ZD0208605), the National Natural Science Foundation of China (32071023, 82271274, 82171242), the Science and Technology Commission of Shanghai Municipality (22XD1420700), the Shanghai Municipal Health Commission (2022XD046), and innovative research team of high-level local universities in Shanghai (SHSMU-ZDCX20211901).

Author information

Authors and Affiliations

Authors

Contributions

WX and JW contributed equally to literature search, writing, and editing of the manuscript. XNL, LS, and JL edited the manuscript. YW, ZL and BS read and commented on the final version of the manuscript. WGL supervised the literature search, discussion, and writing of the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Zhenguo Liu, Bomin Sun or Wei-Guang Li.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xu, W., Wang, J., Li, XN. et al. Neuronal and synaptic adaptations underlying the benefits of deep brain stimulation for Parkinson's disease. Transl Neurodegener 12, 55 (2023). https://doi.org/10.1186/s40035-023-00390-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40035-023-00390-w

Keywords