Skip to main content
Log in

Quantifying harmony between direct and indirect pathways in the basal ganglia: healthy and Parkinsonian states

  • Research Article
  • Published:
Cognitive Neurodynamics Aims and scope Submit manuscript

Abstract

The basal ganglia (BG) show a variety of functions for motor and cognition. There are two competitive pathways in the BG; direct pathway (DP) which facilitates movement and indirect pathway (IP) which suppresses movement. It is well known that diverse functions of the BG may be made through “balance” between DP and IP. But, to the best of our knowledge, so far no quantitative analysis for such balance was done. In this paper, as a first time, we introduce the competition degree \({{\mathcal {C}}}_d\) between DP and IP. Then, by employing \({{\mathcal {C}}}_d\), we quantify their competitive harmony (i.e., competition and cooperative interplay), which could lead to improving our understanding of the traditional “balance” so clearly and quantitatively. We first consider the case of normal dopamine (DA) level of \(\phi ^*=0.3\). In the case of phasic cortical input (10 Hz), a healthy state with \({{\mathcal {C}}}_d^* = 2.82\) (i.e., DP is 2.82 times stronger than IP) appears. In this case, normal movement occurs via harmony between DP and IP. Next, we consider the case of decreased DA level, \(\phi = \phi ^*(=0.3)~x_{DA}\) (\(1 > x_{DA} \ge 0\)). With decreasing \(x_{DA}\) from 1, the competition degree \({{\mathcal {C}}}_d\) between DP and IP decreases monotonically from \({{\mathcal {C}}}_d^*\), which results in appearance of a pathological Parkinsonian state with reduced \({{\mathcal {C}}}_d\). In this Parkinsonian state, strength of IP is much increased than that in the case of normal healthy state, leading to disharmony between DP and IP. Due to such break-up of harmony between DP and IP, impaired movement occurs. Finally, we also study treatment of the pathological Parkinsonian state via recovery of harmony between DP and IP.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Albin RL, Young AB, Penne JB (1989) The functional anatomy of basal ganglia disorders. Trends Neurosci 12:366–375

    Article  CAS  PubMed  Google Scholar 

  • Alexander GE, Crutcher MD (1990) Functional architecture of basal ganglia circuits: neural substrates of parallel processing. Trends Neurosci 13:266–272

    Article  CAS  PubMed  Google Scholar 

  • Ammari R, Bioulac B, Garcia L, Hammond C (2011) 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 5:43

    Article  PubMed  PubMed Central  Google Scholar 

  • Andres DS, Darbin O (2018) Complex dynamics in the basal ganglia: health and disease beyond the motor system. J Neuropsychiatry Clin Neurosci 30:101–114

    Article  PubMed  Google Scholar 

  • Armstrong MJ, Okun MS (2020) Diagnosis and treatment of Parkinson disease: a review. JAMA 323:548–560

    Article  PubMed  Google Scholar 

  • Bahuguna J, Aertsen A, Kumar A (2015) Existence and control of Go/No-Go decision transition threshold in the striatum. PLoS Comput Biol 11:e1004233

    Article  PubMed  PubMed Central  Google Scholar 

  • Baladron J, Hamker FH (2015) A spiking neural network based on the basal ganglia functional anatomy. Neural Netw 67:1–13

    Article  PubMed  Google Scholar 

  • Bar-Gad I, Morris G, Bergman H (2003) Information processing, dimensionality reduction and reinforcement learning in the basal ganglia. Prog Neurobiol 71:439–473

    Article  PubMed  Google Scholar 

  • Bariselli S, Fobbs WC, Creed MC, Kravitz AV (2019) A competitive model for striatal action selection. Brain Res 1713:70–79

    Article  CAS  PubMed  Google Scholar 

  • Baufreton J, Atherton JF, Surmeier DJ, Bevan MD (2005) Enhancement of excitatory synaptic integration by GABAergic inhibition in the subthalamic nucleus. J Neurosci 25:8505–8517

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Bauswein E, Fromm C, Preuss A (1989) Corticostriatal cells in comparison with pyramidal tract neurons: contrasting properties in the behaving monkey. Brain Res 493:198–203

    Article  CAS  PubMed  Google Scholar 

  • Bear MF, Connors BM, Paradiso MA (2007) Neuroscience: Exploring the Brain. Lippincott Williams & Wikins, Philadelphia

  • Belforte JE, Zsiros V, Sklar ER, Yu Jiang Z, Li Y, Quinlan EM, Nakazawa K (2010) Postnatal NMDA receptor ablation in corticolimbic interneurons confers schizophrenia-like phenotypes. Nat Neurosci 13:76–83

    Article  CAS  PubMed  Google Scholar 

  • Bevan MD, Wilson CJ (1999) Mechanisms underlying spontaneous oscillation and rhtymic firing in rat subthalamic neurons. J Neurosci 19:7617–7628

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Bevan MD, Wilson CJ, Bolam JP, Magill PJ (2000) Equilibrium potential of GABA-A current and implications for rebound burst firing in rat subthalamic neurons in vitro. J Neurophysiol 83:3169–3172

    Article  CAS  PubMed  Google Scholar 

  • Bevan MD, Magill PJ, Hallworth NE, Bolam JP, Wilson CJ (2002) Regulation of the timing and pattern of action potential generation in rat subthalamic neurons in vitro by GABA-A IPSPs. J Neurophysiol 87:1348–1362

    Article  CAS  PubMed  Google Scholar 

  • Bolam JP, Bergman H, Graybiel AM, Kimura M, Plenz D, Seung HS, Surmeier DJ, Wickens JR (2006) Microcircuits in the striatum. In: Grillner S, Graybiel AM (eds) Microcircuits: The Interface Between Neurons and Global Brain Function. MIT Press, Cambridge, pp 165–190

    Google Scholar 

  • Brette R, Gerstner W (2005) Adaptive exponential integrate-and-fire model as an effective description of neuronal activity. J Neurophysiol 94:3637–3642

    Article  PubMed  Google Scholar 

  • Brunel N, Hakim V (2008) Sparsely synchronized neuronal oscillations. Chaos 18:015113

    Article  PubMed  Google Scholar 

  • Brunel N, Wang XJ (2003) What determines the frequency of fast network oscillations with irregular neural discharges? I. Synaptic dynamics and excitation-inhibition balance. J Neurophysiol 90:415–430

    Article  PubMed  Google Scholar 

  • Bugaysen J, Bronfeld M, Tischler H, Bar-Gad I, Korngreen A (2010) Electrophysiological characteristics of globus pallidus neurons. PLoS ONE 5:e12001

    Article  PubMed  PubMed Central  Google Scholar 

  • Cakir Y (2019) The synchronization behavior of basal ganglia. J Cogn Syst 4:38–45

    Google Scholar 

  • Celikok U, Sengör NS (2016) Realizing medium spiny neurons with a simple neuron model. Poster session presentation at the meeting of the International Conference on Artificial Neural Networks, Barcelona, Spain

  • Celikok U, Navarro-López EM, Sengör NS (2016) A computational model describing the interplay of basal ganglia and subcortical background oscillations during working memory processes. arXiv https://doi.org/10.48550/arXiv.1601.07740

  • Connelly WM, Schulz JM, Lees G, Reynolds JN (2010) Differential short-term plasticity at convergent inhibitory synapses to the substantia nigra pars reticulata. J Neurosci 30:14854–14861

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Dayan P, Abbott LF (2001) Theoretical Neuroscience. MIT Press, Cambridge

    Google Scholar 

  • DeLong MR (1990) Primate models of movement disorders of basal ganglia origin. Trends Neurosci 13:281–285

    Article  CAS  PubMed  Google Scholar 

  • Fountas Z, Shanahan M (2014) Phase offset between slow oscillatory cortical inputs influences competition in a model of basal ganglia. International joint conference on neural networks (IJCNN) pp 2407–2414

  • Fountas Z, Shanahan M (2017) The role of cortical oscillations in a spiking neural network model of the basal ganglia. PLoS ONE 12:e0189109

    Article  PubMed  PubMed Central  Google Scholar 

  • Frank MJ (2005) Dynamic dopamine modulation in the basal ganglia: a neurocomputational account of cognitive deficits in medicated and non-medicated Parkinsonism. J Cogn Neurosci 17:51–72

    Article  PubMed  Google Scholar 

  • Frank MJ, Loughry B, O’Reilly RC (2001) Interactions between frontal cortex and basal ganglia in working memory: a computational model. Cogn Affect Behav Neurosci 1:137–160

    Article  CAS  PubMed  Google Scholar 

  • Frank MJ, Seeberger LC, O’Reilly RC (2004) By carrot or by stick: cognitive reinforcement learning in Parkinsonism. Science 306:1940–1943

    Article  CAS  PubMed  Google Scholar 

  • Fujimoto K, Kita H (1993) Response characteristics of subthalamic neurons to the stimulation of the sensorimotor cortex in the rat. Brain Res 609:185–192

    Article  CAS  PubMed  Google Scholar 

  • Geisler C, Brunel N, Wang XJ (2005) Contributions of intrinsic membrane dynamics to fast network oscillations with irregular neuronal discharges. J Neurophysiol 94:4344–4361

    Article  PubMed  Google Scholar 

  • Gerstner W, Kistler W (2002) Spiking Neuron Models. Cambridge University Press, New York

    Book  Google Scholar 

  • Gertler TS, Chan CS, Surmeier DJ (2008) Dichotomous anatomical properties of adult striatal medium spiny neurons. J Neurosci 28:10814–10824

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Gittis AH, Yttri EA (2018) Translating insights from optogenetics into therapies for Parkinson’s disease. Curr Opin Biomed Eng 8:14–19

    Article  PubMed  PubMed Central  Google Scholar 

  • Góngora-Alfaro JL, Hernández-López S, Flores-Hernández J, Galarraga E (1997) Firing frequency modulation of substantia nigra reticulata neurons by 5-hydroxytryptamine. Neurosci Res 29:225–231

    Article  PubMed  Google Scholar 

  • Götz T, Kraushaar U, Geiger J, Lübke J, Berger T, Jonas P (1997) Functional properties of AMPA and NMDA receptors expressed in identified types of basal ganglia neurons. J Neurosci 17:204–215

    Article  PubMed  PubMed Central  Google Scholar 

  • Guridi J, González-Redondo R, Obeso JA (2012) Clinical features, pathophysiology, and treatment of levodopa-induced dyskinesias in Parkinson’s disease. Parkinsons Dis 943159

  • Gurney K, Prescott TJ, Redgrave P (2001a) A computational model of action selection in the basal ganglia. I A new functional anatomy. Biol Cybern 84:401–410

    Article  CAS  PubMed  Google Scholar 

  • Gurney K, Prescott TJ, Redgrave P (2001b) A computational model of action selection in the basal ganglia. II. Analysis and simulation of behavior. Biol Cybern 84:411–423

    Article  CAS  PubMed  Google Scholar 

  • Hallworth NE, Wilson CJ, Bevan MD (2003) Apamin-sensitive small conductance calcium-activated potassium channels, through their selective coupling to voltage-gated calcium channels, are critical determinants of the precision, pace, and pattern of action potential generation in rat subthalamic nucleus neurons in vitro. J Neurosci 23:7525–7542

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Humphries MD (2014) Basal ganglia: Mechanisms for action selection. In: Encyclopedia of Computation Neuroscience. Springer, New York

  • Humphries MD, Gurney K (2021) Making decisions in the dark basement of the brain: a look back at the GPR model of action selection and the basal ganglia. Biol Cybern 115:323–329

    Article  PubMed  Google Scholar 

  • Humphries MD, Stewart RD, Gurney KN (2006) A physiologically plausible model of action selection and oscillatory activity in the basal ganglia. J Neurosci 26:12921–12942

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Humphries MD, Lepora N, Wood R, Gurney K (2009a) Capturing dopaminergic modulation and bimodal membrane behaviour of striatal medium spiny neurons in accurate, reduced models. Front Comput Neurosci 3:26

    Article  PubMed  PubMed Central  Google Scholar 

  • Humphries MD, Wood R, Gurney K (2009b) Dopamine-modulated dynamic cell assemblies generated by the GABAergic striatal microcircuit. Neural Netw 22:1174–1188

    Article  PubMed  Google Scholar 

  • Humphries MD, Obeso JA, Dreyer JA (2018) Insights into Parkinson’s disease from computational model of the basal ganglia. J Neurol Neurosurg Psychiatry 89:1181–1188

    Article  PubMed  Google Scholar 

  • Izhikevich EM (2003) Simple model of spiking neurons. IEEE Trans Neural Netw 14:1569–1572

    Article  CAS  PubMed  Google Scholar 

  • Izhikevich EM (2004) Which model to use for cortical spiking neurons? IEEE Trans Neural Netw 15:1063–1070

    Article  PubMed  Google Scholar 

  • Izhikevich EM (2007a) Solving the distal reward problem through linkage of STDP and dopamine signaling. Cereb Cortex 17:2443–2452

    Article  PubMed  Google Scholar 

  • Izhikevich EM (2007b) Dynamical Systems in Neuroscience: The Geometry of Excitability and Bursting. MIT Press, Cambridge

    Google Scholar 

  • Jahr CE, Stevens CF (1990) Voltage dependence of NMDA-activated macroscopic conductances predicted by single-channel kinetics. J Neurosci 10:3178–3182

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Kandel ER, Schwartz JH, Jessell TM (1991) Principles of Neural Science. McGraw-Hill, New York

    Google Scholar 

  • Kim SY, Lim W (2014) Realistic thermodynamic and statistical-mechanical measures for neural synchronization. J Neurosci Meth 226:161–170

    Article  Google Scholar 

  • Kim SY, Lim W (2018a) Stochastic spike synchronization in a small-world neural network with spike-timing-dependent plasticity. Neural Netw 97:92–106

    Article  PubMed  Google Scholar 

  • Kim SY, Lim W (2018b) Effect of inhibitory spike-timing-dependent plasticity on fast sparsely synchronized rhythms in a small-world neuronal network. Neural Netw 106:50–66

    Article  PubMed  Google Scholar 

  • Kim SY, Lim W (2020) Effect of interpopulation spike-timing-dependent plasticity on synchronized rhythms in neuronal networks with inhibitory and excitatory populations. Cogn Neurodyn 14:535–567

    Article  PubMed  PubMed Central  Google Scholar 

  • Kim SY, Lim W (2021a) Effect of diverse recoding of granule cells on optokinetic response in a cerebellar ring network with synaptic plasticity. Neural Netw 134:173–204

    Article  PubMed  Google Scholar 

  • Kim SY, Lim W (2021b) Influence of various temporal recoding on Pavlovian eyeblink conditioning in the cerebellum. Cogn Neurodyn 15:1067–1099

    Article  PubMed  PubMed Central  Google Scholar 

  • Kim SY, Lim W (2022a) Dynamical origin for winner-take-all competition in a biological network of the hippocampal dentate gyrus. Phys Rev E 105:014418

    Article  CAS  PubMed  Google Scholar 

  • Kim SY, Lim W (2022b) Population and individual firing behaviors in sparsely synchronized rhythms in the hippocampal dentate gyrus. Cogn Neurodyn 16:643–665

    Article  PubMed  Google Scholar 

  • Kim SY, Lim W (2022c) Disynaptic effect of hilar cells on pattern separation in a spiking neural network of hippocampal dentate gyrus. Cogn Neurodyn 16:1427–1447

    Article  PubMed  PubMed Central  Google Scholar 

  • Kim SY, Lim W (2023) Effect of adult-born immature granule cells on pattern separation in the hippocampal dentate gyrus. Cogn Neurodyn. https://doi.org/10.1007/s11571-023-09985-5

    Article  PubMed  Google Scholar 

  • Koós T, Tepper JM (1999) Inhibitory control of nestriatal projection neurons by GABAergic interneurons. Nature Neurosci 2:467–472

    Article  PubMed  Google Scholar 

  • Kravitz AV, Freeze BS, Parker PRL, Kay K, Thwin MT, Deisseroth K, Kreitzer AC (2010) Regulation of Parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry. Nature 466:622–626

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Kumaravelu K, Brocker DT, Grill WM (2016) A biophysical model of the cortex-basal ganglia-thalamus network in the 6-OHDA lesioned rat model of Parkinson’s disease. J Comput Neurosci 40:207–229

    Article  PubMed  PubMed Central  Google Scholar 

  • Lindahl M, Kotaleski JH (2016) Untangling basal ganglia network dynamics and function: Role of dopamine depletion and inhibition investigated in a spiking network model. eNeuro 3:e0156

    Article  Google Scholar 

  • Lindahl M, Sarvestani IK, Ekeberg O, Kotaleski JH (2013) Signal enhancement in the output stage of the basal ganglia by synaptic short-term plasticity in the direct, indirect, and hyperdirect pathways. Front Comput Neurosci 7:76

    Article  PubMed  PubMed Central  Google Scholar 

  • Liu C, Wang J, Yu H, Deng B, Wei X, Li H, Loparo KA, Fietkiewicz C (2015) Dynamical analysis of Parkinsonian state emulated by hybrid Izhikevich neuron models. Commun Nonlinear Sci Numer Simul 28:10–26

    Article  CAS  Google Scholar 

  • Liu X, Zhang Q, Wang Y, Chen F (2022) Electrophysiological characterization of substantia nigra pars reticulata an anesthetized rats. J Shanghai Jiaotong Univ (Sci) 27:505–511

    Article  Google Scholar 

  • Luo L (2016) Principles of Neurobiology. Garland Science, New York

    Google Scholar 

  • Mailly P, Charpier S, Menetrey A, Deniau JM (2003) Three-dimensional organization of the recurrent axon collateral network of the substantia nigra pars reticulata neurons in the rat. J Neurosci 23:5247–5257

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Maith O, Escudero FV, Dinkelbach HÜ, Baladron J, Horn A, Irmen F, Kühn AA, Hamker FH (2021) A computational model-based analysis of basal ganglia pathway changes in Parkinson’s disease inferred from resting-state fMRI. Eur J Neurosci 53:2278–2295

    Article  PubMed  Google Scholar 

  • Mandali A, Rengaswamy M, Chakravarthy VS, Moustafa AA (2015) A spiking basal ganglia model;synchrony, exploration and decision making. Front Neurosci 9:191

    Article  PubMed  PubMed Central  Google Scholar 

  • Marino BLB, de Souza LR, Sousa KPA, Ferreira JV, Padilha EC, da Silva CHTP, Taft CA, Hage-Melim LIS (2020) Parkinson’s disease: a review from pathophysiology to treatment. Mini Rev Med Chem 20:754–767

    Article  CAS  PubMed  Google Scholar 

  • Mark MD, Wood R, Gurney K (2010) Reconstructing the three-dimensional GABAergic microcircuit of the striatum. PLoS Comput Biol 6:e1001011

    Article  Google Scholar 

  • Michmizos KP, Nikita KS (2011) Local field potential driven Izhikevich model predicts a subthalamic nucleus neuron activity. In: Engineering in Medicine and Biology Society, EMBC, 2011 annual international conference of the IEEE, IEEE, pp. 5900–5903

  • Moyer J, Wolf JA, Finkel LH (2007) Effects of dopaminergic modulation on the integrative properties of the ventral striatal medium spiny neuron. J Neurophysiol 98:3731–3748

    Article  CAS  PubMed  Google Scholar 

  • Nakanishi H, Kita H, Kitai ST (1990) Intracellular study of rat entopeduncular nucleus neurons in an in vitro slice preparation: electrical membrane properties. Brain Res 527:81–88

    Article  CAS  PubMed  Google Scholar 

  • Narayanan S (2003) The role ofcortico-basal-thalamic loops in cognition: a computational model and preliminary results. Neurocomput 52–54:605–614

    Article  Google Scholar 

  • Navarro-López EM, Celikok U, Sengör NS (2016) Chapter 9 - Hybrid systems neuroscience. In: Hady AE (ed) Closed Loop Neuroscience. Elsevier, London, pp 113–129

    Chapter  Google Scholar 

  • Navarro-López EM, Celikok U, Sengör NS (2021) A dynamical model for the basal ganglia-thalamo-cortical oscillatory activity and its implications in Parkinson’s disease. Cogn Neurodyn 15:693–720

    Article  PubMed  Google Scholar 

  • Obeso JA, Rodriguez-Oroz M, Marın C, Alonso P, Zamarbide I, Lanciego JL, Rodriguez-Diaz M (2004) The origin of motor fluctuations in Parkinson’s disease: importance of dopaminergic innervation and basal ganglia circuits. Neurology 62:S17–S30

    Article  CAS  PubMed  Google Scholar 

  • Obeso JA, Marin C, Rodriguez-Oroz C, Blesa J, Benitez-Temiño B, Mena-Segovia J, Rodríguez M, Olanow CW (2008) The basal ganglia in Parkinson’s disease: current concepts and unexplained observations. Ann Neurol 64:S30–S46

    Article  PubMed  Google Scholar 

  • Oorschot DE (1996) Total number of neurons in the neostriatal, pallidal, subthalamic, and substantia nigral nuclei of the rat basal ganglia: a stereological study using the cavalieri and optical disector methods. J Comp Neurol 366:580–599

    Article  CAS  PubMed  Google Scholar 

  • Park MR, Falls WM, Kitai ST (1982) An intracellular HRP study of the rat globus pallidus. I. Responses and light microscopic analysis. J Comp Neurol 211:284–294

    Article  CAS  PubMed  Google Scholar 

  • Reed JL, Qi HZ, Zhou Z, Bernard MR, Burish MJ, Bonds A, Kaas JH (2010) Response properties of neurons in primary somatosensory cortex of owl monkeys reflect widespread spatiotemporal integration. J Neurophysiol 103:2139–2157

    Article  PubMed  PubMed Central  Google Scholar 

  • Richards C, Shiroyama T, Kitai S (1997) Electrophysiological and immunocytochemical characterization of GABA and dopamine neurons in the substantia nigra of the rat. Neurosci 80:545–557

    Article  CAS  Google Scholar 

  • Rubin JE (2017) Computational models of basal ganglia dysfunction: the dynamics is in the details. Curr Opin Neurobiol 46:127–135

    Article  CAS  PubMed  Google Scholar 

  • Sabatini BL, Regehr WG (1998) Optical measurement of presynaptic calcium currents. Biophys J 74:1549–1563

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Sadek AR, Magill PJ, Bolam JP (2007) A single-cell analysis of intrinsic connectivity in the rat globus pallidus. J Neurosci 27:6352–6362

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Sen-Bhattacharya B, James S, Rhodes O, Sugiarto I, Rowley A, Stokes AB, Gurney K, Furber SB (2018) Building a spiking neural network model of the basal ganglia on SpiNNaker. IEEE Trans Cogn Develop Syst 10:823–836

    Article  Google Scholar 

  • Shen KZ, Johnson SW (2006) Subthalamic stimulation evokes complex EPSCs in the rat substantia nigra pars reticulata in vitro. J Physiol 573:697–709

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Shen Y, Campbell RE, Côté DC, Paquet ME (2020) Challenges for therapeutic applications of opsin-based optogenetic tools in humans. Front Neural Circuits 14:41

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Shimazaki H, Shinomoto S (2010) Kernel bandwidth optimization in spike rate estimation. J Comput Neurosci 29:171–182

    Article  PubMed  Google Scholar 

  • Squire LR, Bloom FE, McConnell SK, Roberts JL, Spitzer NC, Zigmond MJ (2003) Fundamental Neuroscience. Academic Press, New York

    Google Scholar 

  • Tecuapetla F, Matias S, Dugue GP, Mainen ZF, Costa RM (2014) Balanced activity in basal ganglia projection pathways is critical for contraversive movements. Nat Commun 5:4315

    Article  CAS  PubMed  Google Scholar 

  • Thibeault CM, Srinivasa N (2013) Using a hybrid neuron in physiologically inspired models of the basal ganglia. Front Comput Neurosci 7:88

    Article  PubMed  PubMed Central  Google Scholar 

  • Tomkins A, Vasilaki E, Beste C, Gurney K, Humphries MD (2014) Transient and steady-state selection in the striatal microcircuit. Front Comput Neurosci 7:192

    Article  PubMed  PubMed Central  Google Scholar 

  • Turner RS, DeLong MR (2000) Corticostriatal activity in primary motor cortex of the macaque. J Neurosci 20:7096–7108

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Wang XJ (2010) Neurophysiological and computational principles of cortical rhythms in cognition. Physiol Rev 90:1195–1268

    Article  PubMed  Google Scholar 

  • Wang X, Yu Y, Han F, Wang Q (2022) Beta-band bursting activity in computational model of heterogeneous external globus pallidus circuits. Commun Nonlinear Sci Numer Simul 110:106388

    Article  Google Scholar 

  • Wolf JA, Moyer JT, Lazarewicz MT, Contreras D, Benoit-Marand M, O’Donnell P, Finkel LH (2005) NMDA/AMPA ratio impacts state transitions and entrainment to oscillations in a computational model of the nucleus accumbens medium spiny projection neuron. J Neurosci 25:9080–9095

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Yin L, Han F, Yu Y, Wang Q (2023) A computational network dynamical modeling for abnormal oscillation and deep brain stimulation control of obsessive-compulsive disorder. Cogn Neurodyn 17:1167–1184

    Article  PubMed  Google Scholar 

  • Yu Y, Wang X, Wang Q, Wang Q (2020) A review of computational modeling and deep brain stimulation: applications to Parkinson’s disease. Appl Math Mech 41:1747–1768

    Article  PubMed  PubMed Central  Google Scholar 

  • Yu Y, Han F, Wang Q, Wang Q (2022) Model-based optogenetic stimulation to regulate beta oscillations in Parkinsonian neural networks. Cogn Neurodyn 16:667–681

    Article  PubMed  Google Scholar 

  • Yzejian B, DiGregorio DA, Vergara JL, Poage RE, Meriney SD, Grinnell AD (1997) Direct measurements of presynaptic calcium and calcium-activated potassium currents regulating neurotransmitter release at cultured Xenopus nerve-muscle synapses. J Neurosci 17:2009–3001

    Google Scholar 

  • Zheng T, Wilson CJ (2002) Corticostriatal combinatorics: the implications of corticostriatal axonal arborizations. J Neurophysiol 87:1007–1017

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (Grant No. 20162007688).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Woochang Lim.

Ethics declarations

Conflict of interest

The authors declare that they have no Conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Single neuron models and DA effects

Table 3 9 single-cell parameter values in the X (= D1 SPN, D2 SPN, STN, GP, SNr) population

The Izhikevich neuron models are considered as single neuron models in the BG SNN (Izhikevich 2003, 2004, 2007a, b). Evolution of dynamical states of individual cells in the X population [\(X=\) D1 (SPN), D2 (SPN), STN, GP, and SNr] is governed by the following equations:

$$\begin{aligned} C_X \frac{dv_i^{(X)}}{dt}= & {} k_X (v_i^{(X)} - v_r^{(X)}) (v_i^{(X)} - v_t^{(X)}) - u_i^{(X)} +I_i^{(X)}, \end{aligned}$$
(8)
$$\begin{aligned} \frac{du_i^{(X)}}{dt}= & {} a_X \left\{ b_X (v_i^{(X)} - v_r^{(X)}) - u_i^{(X)} \right\} ; i = 1,..., N_X, \end{aligned}$$
(9)

with the auxiliary after-spike resetting:

$$\begin{aligned} \text{if}~ v_i^{(X)} \ge v_{peak}^{(X)}, ~\text{then}~ v_i^{(X)} \leftarrow c_X ~\text{and}~ u_i^{(X)} \leftarrow u_i^{(X)}+d_X, \end{aligned}$$
(10)

where \(N_X\) and \(I_i^{(X)}(t)\) are the total number of cells and the current into the ith cell in the X population, respectively. In Eqs. (8) and (9), the dynamical state of the ith cell in the X population at a time t (msec) is characterized by its membrane potential \(v_i^{(X)}(t)\) (mV) and the slow recovery variable \(u_i^{(X)}(t)\) (pA). When \(v_i^{(X)}(t)\) reaches a threshold \(v_{peak}^{(X)}\) (i.e., spike cutoff value), firing a spike occurs, and then \(v_i^{(X)}\) and \(u_i^{(X)}\) are reset in accord with the rules of Eq. (10).

There are 9 intrinsic parameters in each X population; \(C_X\) (pF): membrane capacitance, \(v_r^{(X)}\) (mV): resting membrane potential, \(v_t^{(X)}\) (mV): instantaneous threshold potential, \(k_X\) (nS/mV): parameter associated with the cell’s rheobase, \(a_X\) (\(\text{msec}^{-1}\)): recovery time constant, \(b_X\) (nS): parameter associated with the input resistance, \(c_X\) (mV): after-spike reset value of \(v_i^{(X)}\), \(d_X\) (pA): after-spike jump value of \(u_i^{(X)}\), and \(v_{peak}^{(X)}\) (mV): spike cutoff value. Table 3 shows the 9 intrinsic parameter values of the BG cells. Along with the parameter values of the D1/D2 SPNs provided in Humphries et al. (2009a); Tomkins et al. (2014), we get the parameter values of the other cells (STN, GP, SNr), founded on the work in Fountas and Shanahan (2017). In the case of GP and STN, we consider the major subpopulations of high frequency pauser (85 \(\%\)) and short rebound bursts (60 \(\%\)), respectively. Also, we use the standard 2-variable Izhikevich neuron model for the STN, instead of the 3-variable Izhikevich neuron model in Fountas and Shanahan (2017); these two models give nearly the same results for the STN.

We also consider influences of DA modulation on the D1 and D2 SPNs (Humphries et al. 2009a; Tomkins et al. 2014; Fountas and Shanahan 2017). D1 receptors activation has two opposing influences on intrinsic ion channels. It enhances the inward-rectifying potassium current (KIR), leading to hyperpolarization of the D1 SPN. In contrast, it lowers the activation threshold of the L type \(\text{Ca}^{2+}\) current, resulting in depolarization of the D1 SPN. These two hyperpolarization and depolarization influences are modelled via variations in intrinsic parameters of the D1 SPN:

$$\begin{aligned} v_r\leftarrow & {} v_r (1+\beta _1^{(\text{D1})} \phi _1), \end{aligned}$$
(11)
$$\begin{aligned} d\leftarrow & {} d(1-\beta _2^{(\text{D1})} \phi _1). \end{aligned}$$
(12)

Here, Eq. (11) models the hyperpolarizing effect of the increasing KIR by upscaling \(v_r\), while Eq. (12) models enhanced depolarizing effect of the L type \(\text{Ca}^{2+}\) current by downscaling d. The parameters \(\beta _1^{\rm (D1)}\) and \(\beta _2^{\rm (D1)}\) represent the amplitudes of their respective influences, and \(\phi _1\) is the DA level (i.e., fraction of active DA receptors) for the D1 SPNs.

Table 4 Effects of DA modulation on intrinsic parameters of the D1/D2 SPNs
Table 5 In-vivo firing activities of BG cells in awake resting state with tonic cortical input (3 Hz) in the case of the normal DA level of \(\phi =0.3\). Spontaneous current \(I_{vivo}^{(X)}\), firing rates \(f_{vivo}^{(X)}\), and random background input \(D_X^*\) (\(X=\) D1 SPN, D2 SPN, STN, GP, and SNr)

Next, D2 receptors activation has small inhibitory influence on the slow A-type potassium current, leading to decrease in the cell’s rheobase current. This depolarizing effect is well modelled by downscaling the parameter, k:

$$\begin{aligned} k \leftarrow k (1-\beta ^{(\text{D2})} \phi _2), \end{aligned}$$
(13)

where \(\phi _2\) is the DA level for the D2 SPNs, and the parameter \(\beta ^{\rm (D2)}\) represents the downscaling degree in k. Table 4 shows DA modulation on the intrinsic parameters of the D1/D2 SPNs where the parameter values of \(\beta _1^{\rm (D1)}\), \(\beta _2^{\rm (D1)}\), and \(\beta ^{\rm (D2)}\) are given (Humphries et al. 2009a; Tomkins et al. 2014; Fountas and Shanahan 2017). In this paper, we consider the case of \(\phi _1 = \phi _2 = \phi\).

Time-evolution of \(v_i^{(X)}(t)\) and \(u_i^{(X)}(t)\) in Eqs. (8) and (9) is governed by the current \(I_i^{(X)}(t)\) into the ith cell in the X population:

$$\begin{aligned} I_i^{(X)}(t) = I_{ext,i}^{(X)}(t) - I_{syn,i}^{(X)}(t) + I_{stim}^{(X)}(t). \end{aligned}$$
(14)

Here, \(I_{ext,i}^{(X)}\), \(I_{syn,i}^{(X)}(t)\), and \(I_{stim}^{(X)}(t)\) denote the external current from the external background region (which is not considered in the modeling), the synaptic current, and the injected stimulation current, respectively. In the BG SNN, we consider the case of \(I_{stim}=0\) (i.e., no injected stimulation DC current).

The external current \(I_{ext,i}^{(X)}(t)\) may be modeled in terms of \(I_{spon,i}^{(X)}\) [spontaneous current for spontaneous firing activity, corresponding to time average of \(I_{ext,i}^{(X)}(t)\)] and \(I_{back,i}^{(X)}(t)\) [random background input, corresponding to fluctuation from time average of \(I_{ext,i}^{(X)}(t)\)]. In the BG population, \(I_{spon}^{(X)}\) (independent of i) is just the spontaneous in-vivo current, \(I_{vivo}^{(X)}\), to get the spontaneous in-vivo firing rate \(f_{vivo}^{(X)}\) in the presence of synaptic inputs in the resting state (in-vivo recording in awake resting state with tonic cortical input). The random background current \(I_{back,i}^{(X)}(t)\) is given by:

$$\begin{aligned} I_{back,i}^{(X)}(t) = D_X \cdot \xi _i^{(X)}(t). \end{aligned}$$
(15)

Here, \(D_X\) is the parameter controlling the noise intensity and \(\xi _i^{(X)}\) is the Gaussian white noise, satisfying the zero mean and the unit variance (Kim and Lim 2018a, b, 2020):

$$\begin{aligned} \langle \xi _i^{(X)}(t) \rangle = 0 ~\text{and}~ \langle \xi _i^{(X)}(t) \xi _j^{(X)}(t') \rangle = \delta _{ij}\delta (t-t'). \end{aligned}$$
(16)

Table 5 shows in-vivo firing activities of BG cells in awake resting state with tonic cortical input for the normal DA level of \(\phi =0.3\); spontaneous in-vivo currents \(I_{vivo}^{(X)}\), in-vivo firing rates \(f_{vivo}^{(X)}\), and random background inputs \(D_X^*\) for Humphries et al. (2006); Lindahl et al. (2013); Fountas and Shanahan (2017) are given.

Synaptic currents and DA effects

We explain the synaptic current \(I_{syn,i}^{(X)}(t)\) in Eq. (14). There are two kinds of excitatory synaptic currents, \(I_{\text{AMPA},i}^{(X,Y)}(t)\) and \(I_{\text{NMDA},i}^{(X,Y)}(t)\), which are are the AMPA (\(\alpha\)-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid) receptor-mediated and NMDA (N-methyl-D-aspartate) receptor-mediated currents from the presynaptic source Y population to the postsynaptic ith cell in the target X population, respectively. In addition to these excitatory synaptic currents, there exists another inhibitory synaptic current, \(I_{\text{GABA},i}^{(X,Z)}(t)\), which is the \(\mathrm GABA_A\) (\(\gamma\)-aminobutyric acid type A) receptor-mediated current from the presynaptic source Z population to the postsynaptic ith cell in the target X population.

Here, we follow the “canonical” formalism for the synaptic currents, as in our previous works in the cerebellum (Kim and Lim 2021a, b) and the hippocampus (Kim and Lim 2022a, b, c, 2023). The synaptic current \(I_{R,i}^{(T,S)}(t)\) R (= AMPA, NMDA, or GABA) obeys the following equation:

$$\begin{aligned} I_{R,i}^{(T,S)}(t) = g_{R,i}^{(T,S)}(t)~(v_{i}^{(T)}(t) - V_{R}^{(S)}), \end{aligned}$$
(17)

where \(g_{(R,i)}^{(T,S)}(t)\) and \(V_R^{(S)}\) are synaptic conductance and synaptic reversal potential, respectively.

The synaptic conductance \(g_{R,i}^{(T,S)}(t)\) is given by:

$$\begin{aligned} g_{R,i}^{(T,S)}(t) = {{\widetilde{g}}}_{max,R}^{(T,S)} \sum _{j=1}^{N_S} w_{ij}^{(T,S)} ~ s_{j}^{(T,S)}(t), \end{aligned}$$
(18)

where \({{\widetilde{g}}}_{max,R}^{(T,S)}\) and \(N_S\) are the maximum synaptic conductance and the number of cells in the source population S, respectively. Here, the connection weight \(w_{ij}^{(T,S)}\) is 1 when the jth presynaptic cell is connected to the ith postsynaptic cell; otherwise (i.e., in the absence of such synaptic connection), \(w_{ij}^{(T,S)}=0\).

We note that, \(s^{(T,S)}(t)\) in Eq. (18) denotes fraction of open postsynaptic ion channels which are opened through binding of neurotransmitters (emitted from the source population S). A sum of exponential-decay functions \(E_{R}^{(T,S)} (t - t_{f}^{(j)}-\tau _{R,l}^{(T,S)})\) provides time evolution of \(s_j^{(T,S)}(t)\) of the jth cell in the source S population:

$$\begin{aligned} s_{j}^{(T,S)}(t) = \sum _{f=1}^{F_{j}^{(S)}} E_{R}^{(T,S)} (t - t_{f}^{(j)}-\tau _{R,l}^{(T,S)}), \end{aligned}$$
(19)

where \(F_j^{(S)},\) \(t_f^{(j)},\) and \(\tau _{R,l}^{(T,S)}\) are the total number of spikes and the fth spike time of the jth cell, and the synaptic latency time constant, respectively.

Similar to our previous works in the cerebellum (Kim and Lim 2021a, b), we use the exponential-decay function \(E_{R}^{(T,S)} (t)\):

$$\begin{aligned} E_{R}^{(T,S)}(t) = e^{-t/\tau _{R,d}^{(T,S)}} \cdot \varTheta (t), \end{aligned}$$
(20)

where \(\tau _{R,d}^{(T,S)}\) is the synaptic decay time constant and the Heaviside step function satisfies \(\varTheta (t)=1\) for \(t \ge 0\) and 0 for \(t <0\).

Table 6 Synaptic parameter values. S: source population, T: target population, R: receptor, \({\tilde{g}}_{max,R}^{(T,S)}\): maximum synaptic conductances, \(\tau _{R,d}^{(T,S)}\): synaptic decay times, \(\tau _{R,l}^{(T,S)}\): synaptic delay times, and \(V_R^{(S)}\): synaptic reversal potential
Table 7 Effects of DA modulation on synaptic currents into the target population (T); T: D1 SPN, D2 SPN, STN, and GP

We also note that, in the case of NMDA-receptor, the positive magnesium ions Mg\(^{2+}\) block some of the postsynaptic NMDA channels. For this case, fraction of non-blocked NMDA channels is given by a sigmoidal function \(f(v^{(T)})\) (Jahr and Stevens 1990; Humphries et al. 2009a; Fountas and Shanahan 2017),

$$\begin{aligned} f(v^{(T)}(t)) = \frac{1}{1+0.28 \cdot [\text{Mg}^{2+}] \cdot e^{-0.062 v^{(T)}(t)}}, \end{aligned}$$
(21)

where \(v^{(T)}\) is the membrane potential of a cell in the target population T and \([\text{Mg}^{2+}]\) is the equilibrium concentration of magnesium ions (\([\text{Mg}^{2+}]\) = 1 mM). Then, the synaptic current into the ith cell in the target X population becomes

$$\begin{aligned} I_{syn,i}^{(X)}(t) = I_{\text{AMPA},i}^{(X,Y)}(t) + f(v_i^{(X)}(t)) \cdot I_{\text{NMDA},i}^{(X,Y)}(t) + I_{\text{GABA},i}^{(X,Z)}(t). \end{aligned}$$
(22)

Table 6 shows the synaptic parameters; synaptic parameter values. S: source population, T: target population, R: receptor, \({\tilde{g}}_{max,R}^{(T,S)}\): maximum synaptic conductances, \(\tau _{R,d}^{(T,S)}\): synaptic decay times, \(\tau _{R,l}^{(T,S)}\): synaptic delay times, and \(V_R^{(S)}\): synaptic reversal potential.

We also take into consideration the influence of DA modulation on the synaptic currents into D1 SPN, D2 SPN, STN, and GP cells in Fig. 1 (Humphries et al. 2009a; Tomkins et al. 2014; Fountas and Shanahan 2017). For the synaptic currents into the D1 SPNs, influence of DA modulation ma be modeled by upscaling the NMDA receptor-mediated current \(I_\text{NMDA}\) with the factor \(\beta ^{(\text{D1})}\):

$$\begin{aligned} I_\text{NMDA} \leftarrow I_\text{NMDA} (1+\beta ^{(\text{D1})} \phi _1). \end{aligned}$$
(23)

Here, \(\phi _1\) is the DA level for the D1 SPNs. (There is no DA influence on \(I_\text{AMPA}\) for the D1 SPNs.) On the other hand, for the synaptic currents into the D2 SPNs, influence of DA modulation could be modeled by downscaling the AMPA receptor-mediated current \(I_\text{AMPA}\) with the factor \(\beta ^{(\text{D2})}\):

$$\begin{aligned} I_\text{AMPA} \leftarrow I_\text{AMPA} (1-\beta ^{(\text{D2})} \phi _2). \end{aligned}$$
(24)

Here, \(\phi _2\) is the DA level for the D2 SPNs. (There is no DA influence on \(I_\text{NMDA}\) for the D2 SPNs.) The scaling factors \(\beta ^{(\text{D1})}\) and \(\beta ^{(\text{D2})}\) are given in Table 7. Also, effects of DA modulation on synaptic currents into STN cells and GP cells are well given in Table 7. In these cases, all excitatory and inhibitory synaptic currents, \(I_\text{AMPA}\), \(I_\text{NMDA}\), and \(I_\text{GABA}\), are downscaled with their scaling factors, depending on \(\phi _2\). Here, \(\phi _1 = \phi _2 = \phi\).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, SY., Lim, W. Quantifying harmony between direct and indirect pathways in the basal ganglia: healthy and Parkinsonian states. Cogn Neurodyn (2024). https://doi.org/10.1007/s11571-024-10119-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11571-024-10119-8

Keywords

Navigation