Skip to main content

Modeling the Organization of Spinal Cord Neural Circuits Controlling Two-Joint Muscles

  • Chapter
  • First Online:
Neuromechanical Modeling of Posture and Locomotion

Abstract

The activity of most motoneurons controlling one-joint muscles during locomotion are locked to either extensor or flexor phase of locomotion. In contrast, bifunctional motoneurons, controlling two-joint muscles such as posterior biceps femoris and semitendinosus (PBSt) or rectus femoris (RF), express a variety of activity patterns including firing bursts during both locomotor phases, which may depend on locomotor conditions. Although afferent feedback and supraspinal inputs significantly contribute to shaping the activity of PBSt and RF motoneurons during real locomotion, these motoneurons show complex firing patterns and variable behaviors under the conditions of fictive locomotion in the immobilized decerebrate cat, i.e., with a lack of patterned supraspinal and afferent inputs. This suggests that firing patterns of PBSt and RF motoneurons are defined by neural interactions inherent to the locomotor central pattern generator (CPG) within the spinal cord. In this study, we use computational modeling to suggest the architecture of spinal circuits representing the locomotor CPG and the connectivity pattern of spinal interneurons defining the behavior of bifunctional PBSt and RF motoneurons. The proposed model reproduces the complex firing patterns of these motoneurons during fictive locomotion under different conditions including spontaneous deletions of flexor and extensor activities and provides insights into the organization of spinal circuits controlling locomotion in mammals.

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

Access this chapter

Institutional subscriptions

Abbreviations

CPG:

Central pattern generator

EMG:

Electromyoagam

ENG:

Electroneurogram

GS:

Gastrocnemius combined with soleus

LGS:

Lateral gastrocnemius combined with soleus

MG:

Medial gastrocnemius

MLR:

Mesencephalic locomotor region

PB:

Posterior biceps femoris

PBSt:

PB combined with semitendinosus

PF:

Pattern formation

Plant:

Plantaris

RF:

Rectus femoris

RG:

Rhythm generator

Sart:

Sartorius

SmAB:

Semimembranosus combined with anterior biceps femoris

St:

Semitendinosus

TA:

Tibialis anterior

UBG:

Unit burst generator

References

  • Anderson FC, Pandy MG (2001) Static and dynamic optimization solutions for gait are practically equivalent. J Biomech 34:153–161

    Article  CAS  PubMed  Google Scholar 

  • Booth V, Rinzel J, Kiehn O (1997) Compartmental model of vertebrate motoneurons for Ca2+ -dependent spiking and plateau potentials under pharmacological treatment. J Neurophysiol 78:3371–3385

    CAS  PubMed  Google Scholar 

  • Brown TG (1914) On the nature of the fundamental activity of the nervous centres; together with an analysis of the conditioning of rhythmic activity in progression, and a theory of the evolution of function in the nervous system. J Physiol 48:18–46

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Butera RJ, Jr., Rinzel J, Smith JC (1999) Models of respiratory rhythm generation in the pre-Botzinger complex. I. Bursting pacemaker neurons. J Neurophysiol 82:382–397

    PubMed  Google Scholar 

  • Carlson-Kuhta P, Trank TV, Smith JL (1998) Forms of forward quadrupedal locomotion. II. A comparison of posture, hindlimb kinematics, and motor patterns for upslope and level walking. J Neurophysiol 79:1687–1701

    CAS  PubMed  Google Scholar 

  • Gregor RJ, Smith DW, Prilutsky BI (2006) Mechanics of slope walking in the cat: quantification of muscle load, length change, and ankle extensor EMG patterns. J Neurophysiol 95:1397–1409

    Article  PubMed  Google Scholar 

  • Grillner S (1981) Control of locomotion in bipeds, tetrapods, and fish. In: Brooks V (ed) Handbook of physiology. section I. The nervous system, Vol II. American Physiological Society, Bethesda, pp 1179–236

    Google Scholar 

  • Grillner S, Zangger P (1979) On the central generation of locomotion in the low spinal cat. Exp Brain Res 34:241–261

    Article  CAS  PubMed  Google Scholar 

  • Guertin P, Angel MJ, Perreault MC, McCrea DA (1995) Ankle extensor group I afferents excite extensors throughout the hindlimb during fictive locomotion in the cat. J Physiol 487(Pt 1):197–209

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Halbertsma JM (1983) The stride cycle of the cat: the modelling of locomotion by computerized analysis of automatic recordings. Acta Physiol Scand Suppl 521:1–75

    CAS  PubMed  Google Scholar 

  • Hamade K, Shevtsova NA, Markin SN, Chakrabarty S, McCrea DA, Rybak IA (2008) How a bipartite CPG can control the activity of bifunctional motoneurons: a modeling study with insights from deletions during fictive locomotion. In 2008 Neuroscience Meeting Planner. Society for Neuroscience. Abstract 925.4. Washington, DC.

    Google Scholar 

  • Huguenard JR, McCormick DA (1991) Vclamp and Cclamp. A computational simulation of single thalamic relay and cortical pyramidal neurons. Neural simulation instruction manual. Stanford University, Stanford

    Google Scholar 

  • Huguenard JR, McCormick DA (1992) Simulation of the currents involved in rhythmic oscillations in thalamic relay neurons. J Neurophysiol 68:1373–1383

    CAS  PubMed  Google Scholar 

  • Jankowska E, Jukes MG, Lund S, Lundberg A (1967a) The effect of DOPA on the spinal cord. 5. Reciprocal organization of pathways transmitting excitatory action to alpha motoneurones of flexors and extensors. Acta Physiol Scand 70:369–388

    Article  CAS  PubMed  Google Scholar 

  • Jankowska E, Jukes MG, Lund S, Lundberg A (1967b) The effect of DOPA on the spinal cord. 6. Half-centre organization of interneurones transmitting effects from the flexor reflex afferents. Acta Physiol Scand 70:389–402

    Article  CAS  PubMed  Google Scholar 

  • Lafreniere-Roula M, McCrea DA (2005) Deletions of rhythmic motoneuron activity during fictive locomotion and scratch provide clues to the organization of the mammalian central pattern generator. J Neurophysiol 94:1120–1132

    Article  PubMed  Google Scholar 

  • Lundberg A (1981) Half-centres revisited. In: Szentagothai J, Palkovits M, Hamori J (eds) Regulatory functions of the CNS. Motion and organization principles. Pergamon Akadem Kiado, Budapest, pp 155–167

    Chapter  Google Scholar 

  • MacGregor RI (1987) Neural and brain modelling. Academic Press, New York

    Google Scholar 

  • Markin SN, Lemay MA, Prilutsky BI, Rybak IA (2012) Motoneuronal and muscle synergies involved in cat hindlimb control during fictive and real locomotion: a comparison study. J Neurophysiol 107:2057–2071

    Article  PubMed  Google Scholar 

  • McCrea DA, Chakrabarty S (2007) Activity patterns in bifunctional PBSt motoneuron pools during fictive locomotion in decerebrate cats: clues to CPG organization. In 2007 Neuroscience Meeting Planner. Society for Neuroscience. Abstract 925.3. Washington, DC.

    Google Scholar 

  • McCrea DA, Rybak IA (2007) Modeling the mammalian locomotor CPG: insights from mistakes and perturbations. Prog Brain Res 165:235–253

    Article  PubMed  PubMed Central  Google Scholar 

  • McCrea DA, Rybak IA (2008) Organization of mammalian locomotor rhythm and pattern generation. Brain Res Rev 57:134–146

    Article  PubMed  Google Scholar 

  • Orsal D, Perret C, Cabelguen JM (1986) Evidence of rhythmic inhibitory synaptic influences in hindlimb motoneurons during fictive locomotion in the thalamic cat. Exp Brain Res 64:217–224

    Article  CAS  PubMed  Google Scholar 

  • Perret C (1983) Centrally generated pattern of motoneuron activity during locomotion in the cat. Symp Soc Exp Biol 37:405–422

    CAS  PubMed  Google Scholar 

  • Perret C, Cabelguen JM (1980) Main characteristics of the hindlimb locomotor cycle in the decorticate cat with special reference to bifunctional muscles. Brain Res 187:333–352

    Article  CAS  PubMed  Google Scholar 

  • Perret C, Cabelguen JM, Orsal D (1988) Analysis of the pattern of activity in “knee flexor” motoneurons during locomotion in cat. In: Gurfinkel VS, Ioffe ME, Massim J (eds) Stance and motion: facts and concepts. Plenum Press, New York, pp. 133–141

    Chapter  Google Scholar 

  • Pratt CA, Buford JA, Smith JL (1996) Adaptive control for backward quadrupedal walking V. Mutable activation of bifunctional thigh muscles. J Neurophysiol 75:832–842

    CAS  PubMed  Google Scholar 

  • Prilutsky BI (2000) Coordination of two- and one-joint muscles: functional consequences and implications for motor control. Motor Control 4:1–44

    Article  CAS  PubMed  Google Scholar 

  • Prilutsky BI, Gregor RJ (2000) Analysis of muscle coordination strategies in cycling. IEEE Trans Rehabil Eng 8:362–370

    Article  CAS  PubMed  Google Scholar 

  • Prilutsky BI, Zatsiorsky VM (2002) Optimization-based models of muscle coordination. Exerc Sport Sci Rev 30:32–38

    Article  PubMed  PubMed Central  Google Scholar 

  • Prilutsky BI, Gregor RJ, Ryan MM (1998a) Coordination of two-joint rectus femoris and hamstrings during the swing phase of human walking and running. Exp Brain Res 120:479–486

    Article  CAS  PubMed  Google Scholar 

  • Prilutsky BI, Isaka T, Albrecht AM, Gregor RJ (1998b) Is coordination of two-joint leg muscles during load lifting consistent with the strategy of minimum fatigue? J Biomech 31:1025–1034

    Article  CAS  PubMed  Google Scholar 

  • Rybak IA, Paton JF, Schwaber JS (1997) Modeling neural mechanisms for genesis of respiratory rhythm and pattern. II. Network models of the central respiratory pattern generator. J Neurophysiol 77:2007–2026

    CAS  PubMed  Google Scholar 

  • Rybak IA, Ptak K, Shevtsova NA, McCrimmon DR (2003) Sodium currents in neurons from the rostroventrolateral medulla of the rat. J Neurophysiol 90:1635–1642

    Article  CAS  PubMed  Google Scholar 

  • Rybak IA, Shevtsova NA, Lafreniere-Roula M, McCrea DA (2006a) Modelling spinal circuitry involved in locomotor pattern generation: insights from deletions during fictive locomotion. J Physiol 577:617–639

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Rybak IA, Stecina K, Shevtsova NA, McCrea DA (2006b) Modelling spinal circuitry involved in locomotor pattern generation: insights from the effects of afferent stimulation. J Physiol 577:641–658

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Safronov BV, Vogel W (1995) Single voltage-activated Na+ and K+ channels in the somata of rat motoneurones. J Physiol 487(Pt 1):91–106

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Shevtsova NA (2015) Two-level model of mammalian locomotor CPG. In: Jaeger D, Jung R (eds) Encyclopedia of Computational Neuroscience. New York, NY, Springer, pp. 2999–3017

    Google Scholar 

  • Shevtsova NA, Chakrabarty S, Hamade K, Markin SN, McCrea DA, Rybak IA (2007) Computational model of mammalian locomotor CPG reproducing firing patterns of flexor, extensor and bifunctional motoneurons during fictive locomotion. In 2014 Neuroscience Meeting Planner. Society for Neuroscience. Abstract 925.4. Washington, DC.

    Google Scholar 

  • Smith JL, Carlson-Kuhta P, Trank TV (1998a) Forms of forward quadrupedal locomotion. III. A comparison of posture, hindlimb kinematics, and motor patterns for downslope and level walking. J Neurophysiol 79:1702–1716

    CAS  PubMed  Google Scholar 

  • Smith JL, Carlson-Kuhta P, Trank TV (1998b) Motor patterns for different forms of walking: cues for the locomotor central pattern generator. Ann NY Acad Sci 860:452–455

    Article  CAS  PubMed  Google Scholar 

  • Stein PSG, Smith JL (1997) Neural and biomechanical control strategies for different forms of vertebrate hindlimb motor tasks. In: Stein P, Grillner S, Selverston AI, Stuart DG (eds) Neurons, networks, and motor behavior. MIT Press, Cambridge, pp 61–73

    Google Scholar 

  • Stuart DG, Hultborn H (2008) Thomas Graham Brown (1882–1965), Anders Lundberg (1920-), and the neural control of stepping. Brain Res Rev 59:74–95

    Article  PubMed  Google Scholar 

  • Wells R, Evans N (1987) Functions and recruitment patterns of one-joint and 2-joint muscles under isometric and walking conditions. Hum Mov Sci 6:349–372

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by the NIH grants R01 NS048844, R01 NS081713, R01 NS090919 , R01 NS095366 and R01 IB012855. The authors would like to thank Dr. David McCrea for useful comments and discussions. We would also like to thank Drs. Angel, Gosgnach, Guertin, Jordan, Lafreniere-Roula, McCrea, Perreault, Stecina, and Quevedo for collection of experimental data used in this modeling study.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Natalia A. Shevtsova PhD .

Editor information

Editors and Affiliations

Appendix

Appendix

All neurons were modelled in the Hodgkin-Huxley style. Motoneurons had two compartments: soma and dendrite and were described based on the previous models (Booth et al. 1997; Rybak et al. 2006a) . The membrane potentials of motoneuron soma (V (S) ) and dendrite (V (D) ) obey the following differential equations:

$$ \begin{aligned}& C\times \frac{d{{V}_{(S)}}}{dt}=-{{I}_{Na(S)}}-{{I}_{K(S)}}-{{I}_{A(S)}}-{{I}_{CaN(S)}}-{{I}_{K,Ca(S)}}-{{I}_{L(S)}}-{{I}_{C(S)}}\; \\& C\times \frac{d{{V}_{(D)}}}{dt}=-{{I}_{NaP(D)}}-{{I}_{CaN(D)}}-{{I}_{CaL(D)}}-{{I}_{K,Ca(D)}}-{{I}_{L(D)}}-{{I}_{C(D)}}-{{I}_{SynE}}-{{I}_{SynI}}\, \end{aligned} $$
(5.A1)

where C is the membrane capacitance and t is time (C = 1 μF/cm2), subscripts S and D indicate the soma or dendrite compartments, respectively.

The dendrite-soma coupling currents (with conductance g C ) for soma I C(S) and dendrite I C(D) are described as follows:

$$ \begin{aligned}& {{I}_{C(S)}}=\frac{{{g}_{C}}}{p}\times ({{V}_{(S)}}-{{V}_{(D)}}); \\& {{I}_{C(D)}}=\frac{{{g}_{C}}}{1-p}\times ({{V}_{(D)}}-{{V}_{(S)}}),\end{aligned} $$
(5.A2)

where p is the parameter defining the ratio of somatic surface area to total surface area (p = 0.1); g C  = 0.1 mS/cm2.

The following currents (with the corresponding maximal channel conductances) are included into motoneuron soma compartments (Booth et al. 1997; Rybak et al. 2006a): fast sodium, I Na (maximal conductance \({{\bar{g}}_{Na}}\) = 120 mS/cm2); persistent sodium, I K ( = \({{\bar{g}}_{NaP}}\) 100 mS/cm2); calcium-N, I CaN (\({{\bar{g}}_{CaN}}\) = 14 mS/cm2); calcium-dependent potassium, I K, Ca (\({{\bar{g}}_{CaL}}\) = 2 mS/cm2), and leakage, I L (g L  = 0.51 mS/cm2) currents. In addition, based on evidence of the presence of the transient (rapidly inactivating) potassium current in the spinal cord interneurons and motoneurons (Safronov and Vogel 1995) this current has been also included in our motoneuron models (I A with maximal conductance \({{\bar{g}}_{A}}\) = 200 ± 40 mS/cm2). The following currents (with the corresponding maximal channel conductances) are included into motoneuron dendritic compartment: persistent sodium, I NaP (\({{\bar{g}}_{NaP}}\) = 0.1 mS/cm2); calcium-N, I CaN (\({{\bar{g}}_{CaN}}\) = 0.3 mS/cm2); calcium-L (I CaL , \({{\bar{g}}_{CaL}}\) = 0.33 mS/cm2), calcium-dependent potassium, I K, Ca (\({{\bar{g}}_{K,Ca}}\) = 0.8 mS/cm2), and leakage, I L (g L  = 0.51 mS/cm2) currents.

Interneurons are simulated as single-compartment models. The neurons within the RG-F, RG-E, PF-F, PF-E, In-eF, and In-eE populations contain fast sodium, I Na ; persistent sodium, I NaP ; delayed-rectifier potassium, I K ; and leakage, I L currents:

$$C \times \frac{{dV}}{{dt}} = - {I_{Na}} - {I_K} - {I_L} - {I_{SynE}} - {I_{SynI}}{\kern 1pt} $$
(5.A3)

The maximal channel conductances for neurons in theses populations are as follows: g L  = 0.51 mS/cm2; \({{\bar{g}}_{Na}}\) = 150 mS/cm2 in RG-F and RG-E neurons and 120 mS/cm2 in the PF-F, PF-E, In-eF, and In-eE populations; \({{\bar{g}}_{NaP}}\) = 1.25 mS/cm2 in RG-F, RG-E, In-eF, and In-eF neurons and 0.1 mS/cm2 in the PF-F and PF-E populations; \({{\bar{g}}_{K}}\) = 5 mS/cm2 in the RG-F and RG-E populations and 10 mS/cm2 in the PF-F, PF-E, In-eE, and In-eF populations.

For simplicity, all other interneurons contain only minimal set of ionic currents:

$$ C\times \frac{dV}{dt}=-{{I}_{Na}}-{{I}_{K}}-{{I}_{L}}-{{I}_{SynE}}-{{I}_{SynI}}$$
(5.A4)

with the following maximal conductances: \({{\bar{g}}_{Na}}\) = 120 mS/cm2; \({{\bar{g}}_{K}}\) = 10 mS/cm2; g L  = 0.51 mS/cm2.

The ionic currents included into the modelled neurons are described as follows:

$$ \begin{aligned}& {{I}_{Na}}={{{\bar{g}}}_{Na}}\times m_{Na}^{3}\times {{h}_{Na}}\times ({{V}_{{}}}-{{E}_{Na}}); \\& {{I}_{NaP}}={{{\bar{g}}}_{NaP}}\times {{m}_{NaP}}\times {{h}_{NaP}}\times (V-{{E}_{Na}}); \\& {{I}_{K}}={{{\bar{g}}}_{K}}\times m_{K}^{4}\times ({{V}_{{}}}-{{E}_{K}}); \\& {{I}_{A}}={{{\bar{g}}}_{A}}\times (0.6\times m_{A1}^{4}\times {{h}_{A1}}+0.4\times m_{A2}^{4}\times {{h}_{A2}})\times (V-{{E}_{K}}); \\& {{I}_{CaN}}={{{\bar{g}}}_{CaN}}\times m_{CaN}^{2}\times {{h}_{CaN}}\times ({{V}_{{}}}-{{E}_{Ca}}); \\& {{I}_{CaL}}={{{\bar{g}}}_{CaL}}\times {{m}_{CaL}}\times (V-{{E}_{Ca}}); \\& {{I}_{K,Ca}}={{{\bar{g}}}_{K,Ca}}\times m_{K,Ca}^{{}}\times ({{V}_{{}}}-{{E}_{K}}); \\& {{I}_{L}}={{g}_{L}}\times ({{V}_{{}}}-{{E}_{L}}),\end{aligned} $$
(5.A5)

where V is the membrane potential of the corresponding neuron compartment (soma, V (S) , or dendrite, V (D) ) in two-compartment models, or the neuron membrane potential V in one-compartment models; E Na , E K , E Ca , and E L are the reversal potentials for sodium, potassium, calcium, and leakage current respectively; variables m and h with indexes indicating ionic currents represent, respectively, the activation and inactivation variables of the corresponding ionic channels.

The reversal potential values in the model are as follows: E Na  = 55 mV; E K= − 80 mV; E Ca= 80 mV; E L  = − 64 ± 0.64 mV in RG-F and RG-E neurons, E L  = − 65 ± 0.325 mV in Inrg-F and Inrg-E interneurons and motoneurons, and E L  = − 68 ± 0.34 mV in all other neurons.

Activation m and inactivation h of voltage-dependent ionic channels (e.g., Na, NaP, K, A, CaN, CaL) are described by the following differential equations:

$$\begin{aligned}& {{\tau }_{mi}}(V)\times \frac{d}{dt}{{m}_{i}}={{m}_{\infty i}}(V)-{{m}_{i}}; \\ & {{\tau }_{hi}}(V)\times \frac{d}{dt}{{h}_{i}}={{h}_{\infty i}}(V)-h, \\ \end{aligned}$$
(5.A6)

where i identifies the name of the channel, m i (V) and h i (V) represent the voltage-dependent steady-state activation and inactivation respectively, and τ mi (V) and τ hi (V) define the corresponding time constants (see their descriptions in Table 5.A1). Activation of the sodium channels is considered to be instantaneous (τ mNa  = τ mNaP  = 0, see (Booth et al. 1997; Butera et al. 1999)) .

Activation of the Ca2+ -dependent potassium channels is also considered instantaneous and described as follows (Booth et al. 1997) :

$$ {{m}_{K,Ca}}=\frac{Ca}{Ca+{{K}_{d}}}, $$
(5.A7)

where Ca is the Ca2+ concentration within the corresponding compartment of motoneuron, and K d defines the half-saturation level of this conductance.

The kinetics of intracellular Ca2+ concentration (Ca, described separately for each compartment) is modelled according to the following equation (Booth et al. 1997) :

$$ \frac{d}{dt}C{{a}_{{}}}=f\times (-\alpha \times {{I}_{Ca}}-{{k}_{Ca}}\times Ca), $$
(5.A8)

where f defines the percent of free to total Ca2+; α converts the total Ca2+ current, I Ca , to Ca2+ concentration; k Ca represents the Ca2+ removal rate.

The synaptic excitatory (I synE with conductance g synE and reversal potential E SynE= −10 mV) and inhibitory (I synI with conductance g synI and reversal potential E SynI= −70 mV) currents are described as follows:

$$ \begin{aligned}& {{I}_{SynE}}={{g}_{SynE}}\times ({{V}_{{}}}-{{E}_{SynE}}); \\& {{I}_{SynI}}={{g}_{SynI}}\times ({{V}_{{}}}-{{E}_{SynI}}). \end{aligned} $$
(5.A9)

The excitatory (g SynE ) and inhibitory synaptic (g SynI ) conductances are equal to zero at rest and may be activated (opened) by the excitatory or inhibitory inputs to neuron i respectively:

$$\begin{array}{*{20}{l}}{{g_{SynEi}}(t) = {{\bar g}_E} \times \sum\limits_{j{\rm{\backslash kern}}1pt} {\sum\limits_{{t_{kj}} < t} {S\{ {w_{ji}}\} \times \exp ( - (t - {t_{kj}})/{\tau _{SynE}})} } + {{\bar g}_{Ed}} \times \sum\limits_m {S\{ {w_{dmi}}\} \times {d_{mi}}} ;}\\{{g_{SynIi}}(t) = {{\bar g}_I} \times \sum\limits_{j{\rm{\backslash kern}}1pt} {\sum\limits_{{t_{kj}} < t} {S\{ - {w_{ji}}\} \times \exp ( - (t - {t_{kj}})/{\tau _{SynI}})} } + {{\bar g}_{Id}} \times \sum\limits_m {S\{ - {w_{dmi}}\} \times {d_{mi}}} ,}\end{array}$$
(5.A10)

where the function S{x} = x, if x ≥ 0, and 0 if x < 0. According to equation (5.A10), the excitatory and inhibitory synaptic conductances have two terms: the first term describes the effects of inputs from other neurons in the network (excitatory and inhibitory respectively), and the second one describes effects of inputs from external drives d mi (see also Rybak et al. 1997) . Each spike arriving to neuron i from neuron j at time t kj increases the excitatory synaptic conductance by \({{\bar{g}}_{E}}\times {{w}_{ji}}\) if the synaptic weight w ji  > 0, or increases the inhibitory synaptic conductance by -\({{\bar{g}}_{I}}\times {{w}_{ji}}\) if the synaptic weight w ji  < 0. \({{\bar{g}}_{E}}\) = 0.05 mS/cm2 and \({{\bar{g}}_{I}}\) = 0.05 mS/cm2 are the parameters defining an increase in the excitatory or inhibitory synaptic conductance, respectively, produced by one arriving spike at |w ji | = 1. τ SynE  = 5 ms and τ SynE  = 15 ms are the decay time constants for the excitatory and inhibitory conductances respectively. In the second terms of equations (5.A10), \({{\bar{g}}_{Ed}}\) = \({{\bar{g}}_{Id}}\) = 1 mS/cm2 is the parameter defining the increase in the excitatory synaptic conductance, produced by external input drive d mi= 1 with a synaptic weight of |w dmi | = 1. All synaptic weights used in the model can be found in Table 5.A2. The values of input drives used in particular simulation are shown in Table 5.A3.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer Science+Business Media New York

About this chapter

Cite this chapter

Shevtsova, N., Hamade, K., Chakrabarty, S., Markin, S., Prilutsky, B., Rybak, I. (2016). Modeling the Organization of Spinal Cord Neural Circuits Controlling Two-Joint Muscles. In: Prilutsky, B., Edwards, D. (eds) Neuromechanical Modeling of Posture and Locomotion. Springer Series in Computational Neuroscience. Springer, New York, NY. https://doi.org/10.1007/978-1-4939-3267-2_5

Download citation

Publish with us

Policies and ethics