Comprehensive Physiology Wiley Online Library

Integration in Spinal Neuronal Systems

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Methodological Considerations
1.1 Selective Stimulation of Primary Afferents
1.2 Stimulation of Central Motor Systems
1.3 Methods for Investigation of Convergence at Interneuronal Level
2 Spinal Neuronal Circuits Used in Common by Segmental Afferents and Supraspinal Motor Centers
2.1 Recurrent Inhibition
2.2 Pathways From Ia‐Afferents and Their Control by γ‐Motoneurons
2.3 Reflex Pathways From Group Ib Tendon Organ Afferents
2.4 Reflex Pathways From Cutaneous and Joint Afferents and From Groups II and III Muscle Afferents
2.5 Propriospinal Neurons
2.6 Presynaptic Inhibition of Transmission From Primary Afferents
3 Reticulospinal Inhibition of Segmental Reflex Transmission
3.1 Dorsal Reticulospinal System
3.2 Ventral Reticulospinal Pathways
3.3 Monoaminergic Reticulospinal Pathways
3.4 Decerebrate Preparation
4 Direct Projections of Descending Pathways to α‐Motoneurons
5 Ascending Pathways that Monitor Segmental Interneuronal Activity
5.1 Evidence That Ascending FRA Pathways Monitor Activity in Interneurons of Reflex Pathways
5.2 Information Via Ascending Collaterals of Interneurons
5.3 Ventral Flexor Reflex Tracts
5.4 Ventral Spinocerebellar Tract
6 General Summary and Epilogue
6.1 General Summary
6.2 Epilogue
Figure 1. Figure 1.

Indirect technique used to investigate convergence in interneurons of reflex pathways. Diagram exemplifies convergence from primary afferents (I, prim. aff.) and descending pathways (II, desc.) on common excitatory interneurons projecting to motoneurons. Neuronal circuits are illustrated on left. A: example of excitatory convergence from both sources. B: example of convergence of excitation from prim. aff. and inhibition from desc. pathways. Interneurons represent a population of interneurons with identical convergence. Traces on right give idealized intracellular records from a single motoneuron after the test stimulus (I, prim. aff.), the conditioning (cond.) stimulus (II, desc.), and combined stimulation (I+II). This example refers to convergence from descending fibers and primary afferents, but the technique can be used for convergence from any source. See text for explanation.

Adapted from Lundberg 412 in The Nervous System. The Basic Neurosciences. ©1975, Raven Press, N. Y
Figure 2. Figure 2.

Relations between individual motoneurons and individual Renshaw cells. A: illustration of experimental arrangement and neuronal circuit. Excitatory effect on Renshaw cells (RC) by single impulses in motoneurons (Mn) and inhibitory effect of individual Renshaw cells onto individual motoneurons were investigated with the aid of 2 microelectrodes: 1 used for intracellular stimulation of and recording from motoneurons, and 1 used for extracellular recording from Renshaw cells. The 2nd electrode was filled with 3 M sodium glutamate, which allowed electrophoretic application of glutamate to induce repetitive firing of the recorded Renshaw cells. B: 2 samples of burst firing in a Renshaw cell (upper traces, large spikes) following single action potentials in the motoneuron (timing shown by lower trace). Stimulation frequency of motoneuron was 8.5 impulses/s. Time calibration in D applies also to records in B. C: graph showing rapid decrease of number of spikes in Renshaw cell burst firing following each motoneuronal spike with increasing stimulation frequency. Dots represent mean value of several trials (cf. records in B) at different stimulation frequencies. This example shows the most powerful excitatory coupling of 21 tested cases. D: slow repetitive firing of Renshaw cell (upper trace, large spike) with simultaneous recording of membrane potential of motoneuron (lower trace; voltage calibration applies to intracellular recording). Same motoneuron and Renshaw cell that was illustrated in B, C, E: averaged record of membrane potential in motoneuron following single discharges of the Renshaw cell (the Renshaw cell firing with “doublets” or “triplets” are excluded). Same motoneuron and Renshaw cell that was illustrated in B–D. Note remarkably long time course of this unitary IPSP. F: a more typical time course of a presumed unitary IPSP from a Renshaw cell in a different motoneuron. Pretriggered averaging was used in both E and F; zero on abscissa refers to a time 1 ms before occurrence of spike of Renshaw cell. (From L. Van Keulen, unpublished material. see ref. 592.)

Figure 3. Figure 3.

Comparison of patterns of monosynaptic excitation from muscle spindle Ia‐afferents and recurrent inhibition via motor axon collaterals in motor nuclei supplying various hindlimb muscles in cat. Experimental arrangement and neuronal circuits as illustrated in the scheme. Motor nuclei recorded from are listed to left of diagram and nerves whose orthodromic and antidromic effects were compared are indicated above. Homonymous connections, always with both Ia excitation and recurrent inhibition, are indicated by dotted areas. Heteronymous Ia excitation and recurrent inhibition are indicated by hatched and shaded areas, respectively. Diagram is constructed from data obtained in the cat. It indicates a complete overlap of Ia excitation and recurrent inhibition, but with recurrent inhibition more widely distributed than Ia excitation. Sart, sartorius; Q, quadriceps; Sm, semimembranosus; AB, anterior biceps; St, semitendinosus; PB, posterior biceps; Per, peroneus; G‐S, gastrocnemius‐soleus; FDL, flexor digitorum longus; Pl, plantaris; Add, adductor femoris and longus; Grac, gracilis.

Data for Ia pathways from Eccles et al. 135 and Eccles and Lundberg 147; data for recurrent inhibition from Hultborn et al. 292, Eccles et al. 131, and Wilson et al. 613
Figure 4. Figure 4.

Convergence of recurrent inhibition in a primate (baboon) gracilis (Grac) motoneuron with correlation to location of motor nuclei whose efferent axons have been stimulated. A: upper traces are intracellular averaged records (calibration pulse 1 mV, 4 ms) from the gracilis motoneuron. Lower traces are simultaneous records from the cord dorsum. Dorsal roots have been cut and nerves indicated have been stimulated at supramaximal strength for α‐fibers (arrival of volleys at spinal cord indicated by arrows). B: segmental location of motor nuclei tested in A (at the end of each experiment the rostrocaudal distribution of motor nuclei were tested by stimulation of individual ventral roots with simultaneous recording from peripheral muscle nerves). Position of recorded gracilis motoneuron is indicated by arrow. Results suggest that recurrent inhibition is not organized according to proximity principle. Note large recurrent IPSPs from semitendinosus and posterior biceps whose motor nuclei are among the most caudal ones, but no effect at all from quadriceps despite a similar rostrocaudal location of quadriceps and gracilis motor nuclei. Q, quadriceps; Sm, semimembranosus; Pl, plantaris; AB, anterior biceps; Sart, sartorius; TA, tibialis anterior; EDL, extensor digitorum longus; St, semitendinosus; PB, posterior biceps. (From H. Hultborn, E. Jankowska, and S. Lindström, unpublished material. Cited in ref. 292.)

Figure 5. Figure 5.

Diagrams illustrating the hypothesis that Renshaw system serves as a variable gain regulator at motoneuronal level. A: input and output connections of α‐motoneurons and Renshaw cells. B: simplified diagram of input‐output relations of a motoneuronal pool during inhibition and facilitation of transmission in recurrent pathway. C: concept of motor output stage. Neurons constituting output are framed by thick lines. Hatched lines indicate parallel connections to α‐ and γ‐motoneurons and corresponding Ia inhibitory interneurons. α, α‐Motoneurons; γ, γ‐motoneurons; RC, Renshaw cells; Ia IN, Ia inhibitory interneurons.

Adapted from Hultborn et al. 293
Figure 6. Figure 6.

Distribution of heteronymous monosynaptic excitation evoked by impulses in muscle spindle Ia‐afferents from proximal hindlimb muscles in cat. Ia‐afferents of an individual muscle (indicated by coils) evoke excitatory effects via monosynaptic pathways in α‐motoneurons of muscles (marked by open triangles). TFL, musculus tensor fasciae latae.

Adapted from Eccles and Lundberg 147
Figure 7. Figure 7.

Long‐lasting excitability increase of α‐motoneurons induced via polysynaptic Ia pathways. B, C: simultaneous intracellular records at different gains from a triceps surae motoneuron in a decerebrate, decerebellate cat (experimental arrangement in A). Ventral roots were intact, but γ‐loop was opened by relaxation with gallamine triethiodide (Flaxedil). Short stimulus trains to medial gastrocnemius nerve (on: MG) caused a depolarizing shift of membrane potential and a firing of motoneuron. This effect was reversed by a short stimulus train to superficial peroneal nerve (off: SP). D, E: monosynaptic reflexes (MSR) evoked by stimulation of nerve from medial gastrocnemius and recorded from central end of cut S1–L7 ventral roots (γ‐loop open) in a spinal unanesthetized cat after injection of 5‐hydroxytryptophan (50 mg/kg; experimental arrangement in D). E: integrated amplitude of MSR (arbitrary units, horizontal lines give mean ± SEM) is plotted against time. Stimulation of lateral gastrocnemius‐soleus nerve (on: LG‐S) increased amplitude of MSR while a short stimulus train to the SP nerve (off: SP) brought it back to control value. Values obtained during LG‐S or SP stimulation are not included in the calculation.

AC adapted from Hultborn and Wigström 301; D, E from H. Hultborn and H. Wigström, unpublished observations
Figure 8. Figure 8.

Distribution of disynaptic inhibition evoked by impulses in muscle spindle Ia‐afferents from proximal hindlimb muscles in cat. Ia‐afferents of an individual muscle (indicated by coils) evoke inhibitory actions via disynaptic pathways in α‐motoneurons of muscles (marked by closed triangles). TFL, musculus tensor fasciae latae.

Adapted from Eccles and Lundberg 147
Figure 9. Figure 9.

Convergence of Ia‐afferents on α‐motoneurons and interneurons that mediate disynaptic Ia inhibition to motoneurons of antagonists (Ia inhibitory interneurons). Scheme in G illustrates experimental arrangement and neuronal circuits. A–C: upper traces are intracellular records from a semitendinosus (St) motoneuron (identified by antidromic invasion in A). Lower traces are records from L7 dorsal root entry zone. Voltage calibrations apply to intracellular records. Activation of Ia‐afferents from St, posterior biceps (PB), and gracilis (Grac) muscles evokes monosynaptic EPSPs. DF: intracellular (averaged) records from a vastocrureus (V–Cr) motoneuron (direct antagonist to St). Calibration pulse: 2 ms, 0.5 mV. Stimulation strength (in times threshold of lowest threshold afferent fibers): D) PB, 1.2; St, 1.1; E) Grac, 1.3; St, 1.05; F) Grac, 1.3; PB, 1.2. Appearance of disynaptic IPSPs in motoneuron (bottom traces in DF) upon conjoint stimulation of PB and St (D), Grac and St (E), and Grac and PB (F) demonstrates convergence of monosynaptic excitation from respective Ia‐afferents on common inhibitory interneurons that thus receive same excitatory convergence as the St motoneurons.

AC adapted from Eccles et al. 135; DF adapted from Hultborn and Udo 299
Figure 10. Figure 10.

Identification of interneurons mediating disynaptic Ia inhibition (Ia inhibitory interneurons) from quadriceps (Q) afferents to posterior biceps and semitendinosus (PBSt) motoneurons. Experimental arrangement and neuronal circuits are illustrated in D (for records AG) and J (for records I, K). Upper traces in AG are intracellular records from a PBSt motoneuron (A–C) and a lamina VII interneuron (E–G). Lower traces are records from L5 dorsal root entry zone. Upper trace in I is an extracellular record from a lamina VII interneuron that is excited by electrophoretic ejection of glutamate from the recording electrode. Lower trace is a simultaneous intracellular record from a PBSt motoneuron. K shows averaged intracellular response from I (lower trace) triggered by extracellular interneuronal spike (upper trace). Voltage calibrations refer to intracellular records. Stimulation strength of Q nerve is given in times threshold of lowest threshold afferent fibers; stimulation of ventral roots (VRs) was supramaximal for α‐fibers. AC: disynaptic Ia IPSPs from Q afferents in a PBSt motoneuron (A, B) and its depression by conditioning stimulation of L5 + L6 VRs (C). E–G: lamina VII interneuron with convergence required for mediation of effects recorded in PBSt motoneuron of AC, i.e., monosynaptic activation from Q Ia‐afferents (E, F) and recurrent inhibition from L6 VR (G). H: summarizing diagram giving location in L6 of a number of interneurons with convergence illustrated in E–G. I–K: synaptic action in a PBSt motoneuron by an interneuron with convergence shown in EG and location shown in H. Monosynaptic unitary IPSP evoked by interneuronal activity proves that interneuron with monosynaptic Ia excitation and disynaptic inhibition from motor axons mediates reciprocal Ia inhibition.

AC from H. Hultborn, E. Jankowska, and S. Lindström, unpublished records; EG adapted from Hultborn et al. 287; H adapted from Hultborn et al. 291; IK adapted from Jankowska and Roberts 336
Figure 11. Figure 11.

Connections to interneuron in reciprocal Ia inhibitory pathway. A: circuit diagram of some connections to interneuron in reciprocal Ia inhibitory pathway, i, Ipsilateral; co, contralateral; Vs, vestibulospinal tract; Cs, corticospinal tract; Rs, rubrospinal tract; Ps, propriospinal tract; cut, cutaneous afferents; FRA, flexor reflex afferents; Mn, motoneurons; R, Renshaw cells. BK: parallel projection from nucleus vestibularis lateralis (ND) onto α‐motoneurons and corresponding Ia inhibitory interneurons. Experimental arrangement and neuronal circuits as illustrated in K. Upper traces are intracellular records from a posterior biceps‐semitendinosus motoneuron (PBSt; B–E), a quadriceps (Q) motoneuron (I, J), and a Ia inhibitory interneuron (F–H) presumably intercalated in the Ia inhibitory pathway from Q to PBSt. Lower traces are records from dorsal root entry zone in L7 (B–E) or L6, (F–J). Voltage calibrations refer to intracellular records. Ipsilateral ND stimulated with 80 μA in CE and 200 μA in G. Q nerve (B–E, F) stimulated with 1.1 times threshold of the lowest threshold afferent fibers. Ventral roots (VR) were stimulated with single stimuli, supramaximal for α‐fibers. BE: facilitation of the Q Ia IPSP in a PBSt motoneuron from vestibulospinal tract (B–D) and depression of facilitated IPSP from L5 + L6 VRs (E). Results indicate convergence of vestibulospinal excitation and inhibition from motor axon collaterals on common Ia inhibitory interneurons projecting to PBSt motoneurons. FH: monosynaptic excitation from the ND (G) of a Q‐activated Ia inhibitory interneuron (F) and its disynaptic inhibition from L5 + L6 VRs (H). Note that these direct recordings from interneuron (F–H) correspond to convergent actions from the ND and the VR on the Ia IPSP recorded in the PBSt motoneuron (B–E). I, J: monosynaptic activation of a Q motoneuron (I, homonymous EPSP) from lateral vestibulospinal tract (J). L: schematic drawing of parallel projections (broken lines) onto α‐ and γ‐motoneurons innervating a muscle and the Ia inhibitory interneuron inhibiting the motoneurons of its antagonist. Notice also the mutual inhibition between opposite Ia inhibitory interneurons. Further description in the text.

BE adapted from Hultborn and Udo 298; FH adapted from Hultborn et al. 289; I, J adapted from Grillner et al. 240; L adapted from Hultborn et al. 287
Figure 12. Figure 12.

Convergence on interneurons in Ib inhibitory pathway to motoneurons. Upper traces, intracellular recordings from motoneurons to gastrocnemius‐soleus (A–C) and to flexor digitorum longus (E–H). Lower traces, incoming volleys recorded from L7 dorsal root entry zone. Voltage calibrations refer to intracellular records. Stimulus strengths of peripheral nerves are given in multiples of thresholds for lowest threshold afferents. AC: facilitatory interaction in inhibitory transmission to motoneurons from Ib muscle afferents from plantaris (Pl) and cutaneous afferents in superficial peroneal nerve (SP). EH: facilitatory interaction between a single descending volley in rubrospinal tract (stimulation of red nucleus, NR) and a Ib volley in quadriceps nerve (Q); G, H compare the lack of facilitation with weaker nerve stimulation that mainly activates Ia‐afferents in E, F. Arrow below records in H indicates time of arrival of the fastest descending volleys in rubrospinal tract to the L6 level. Corresponding graphs in D and I show time course of facilitation obtained by varying conditioning‐testing interval. Zero on abscissa indicates simultaneous arrival at segmental level of both conditioning and testing volleys. Delay of facilitation by a cutaneous volley (D) suggests a disynaptic linkage to Ib inhibitory interneurons. Facilitatory action at simultaneous arrival (and even with the conditioning volley arriving slightly after the test) in the case of rubrospinal volley (I) strongly indicates a monosynaptic linkage from rubrospinal tract. Circuit diagram summarizes most of the neuronal connections to Ib inhibitory interneurons.

AC adapted from Lundberg et al. 417; EH adapted from Hongo et al. 275
Figure 13. Figure 13.

Convergence onto a lamina V and VI interneuron possibly interposed in Ib reflex pathways to motoneurons. Upper traces (A‐F), intracellular records from a lamina V and VI interneuron; lower traces (A‐C) and middle traces (D‐F), records of afferent volleys from L7 dorsal root entry zone; lower traces (D‐F), records of changes in muscle length (increase in length downward). AC: convergence of monosynaptic excitation from group I afferents in nerve from plantaris (Pl), of disynaptic excitation from cutaneous afferents in superficial peroneal nerve (SP) and of monosynaptic excitation from ipsilateral descending fibers (i. desc.). Stimulus strength is given in times threshold for lowest threshold fibers. DF: analysis by graded brief stretches of plantaris muscle of contribution from muscle spindle Ia‐afferents and Golgi tendon organ Ib‐afferents to the group I excitation illustrated in A. Small stretch in D (cf. calibration to right of F) selectively activates muscle spindle Ia‐afferents, the larger muscle stretches in E and F are near maximal for Ia‐afferents and above threshold for Ib‐afferents, respectively. This series of records thus strongly indicates that both Ia‐ and Ib‐afferents contribute monosynaptic excitation. G: reconstruction of soma and part of the axonal projections of interneuron whose excitatory input is illustrated in A‐F. Interneuron was stained by ejecting horseradish peroxidase from the microelectrode after the recording. Notice the axonal projection into region of motoneurons (large neurons, probably motoneurons, are indicated by hatched areas). (From E. Jankowska, T. Johannisson, and J. Lipski, unpublished observations.)

Figure 14. Figure 14.

Facilitation from corticospinal tract of a cutaneous reflex originating from plantar cushion. Scheme in A illustrates experimental arrangement and the involved neuronal circuits. Stippled area symbolizes a pool of interneurons and emphasizes that the actions from both systems are relayed in polysynaptic pathways. Actual site of interaction is not known. Upper traces in BE are records from ventral root; lower traces are simultaneous recordings from dorsal root entry zone. Monosynaptic test reflex of plantar muscles (B) is moderately facilitated by a weak (conditioning) electrical stimulus to the pad (C). Stimulation of cortex (D) had no detectable effect on test reflex, but combined action from the 2 conditioning systems resulted in a large spatial facilitation (E).

Adapted from Engberg 165
Figure 15. Figure 15.

Interaction of cortico‐ and rubrospinal tracts with lumbar cutaneous reflex pathways. Upper traces in AK are intracellular records from different α‐motoneurons and interneurons. AC: posterior biceps‐semitendinosus motoneuron. DF: gastrocnemius‐soleus motoneuron. IK: pretibial flexor motoneuron supplied by deep peroneal nerve. G, H: L7 dorsal horn interneuron. Lower traces are records from dorsal root entry zone at the respective segmental levels. Voltage calibrations refer to intracellular records. AF: facilitation of cutaneous actions from cortex. Liminal synaptic actions evoked from sural nerve (Sur; stimulated alone in A and D) were conditioned by preceding stimulation of cortex (C and F). Cortex was stimulated alone in B and E. The ensuing increase of the polysynaptic potentials indicates excitatory convergence by corticospinal volleys and cutaneous afferent volleys on interneurons in excitatory and inhibitory cutaneous reflex pathways. G, H: monosynaptic activation of a lumbar dorsal horn interneuron from cutaneous (Sur) afferents (G) and short‐latency (probably monosynaptic) activation from the cortex (H). This pattern suggests that, in the action illustrated in AF, the corticospinal tract may project monosynaptically onto first‐order interneurons in lumbar cutaneous pathways. IK: monosynaptic excitation from rubrospinal tract of last order interneurons in a cutaneous reflex pathway. I shows a liminal IPSP evoked from red nucleus (NR) with a single shock of 200 μA (arrows below surface records in I and K mark arrival at segmental level of descending volley in rubrospinal tract). Conditioning stimulation of superficial peroneal nerve (SP) at 1.4 times threshold of the lowest threshold afferents increased the disynaptic rubrospinal IPSP (K; SP alone in J). L: comparison of location in cat lumbar cord of focal potentials evoked from rubrospinal and corticospinal tracts. Transverse section shows areas in which field potentials were evoked from indicated sites with an amplitude above 80% (black and hatched area) and 50% (continuous and interrupted lines) of their maximal amplitudes. Dorsal location of field from cortex and ventral location of NR field would be compatible with hypothesis that corticospinal fibers interact with first–order interneurons (cf. A‐F, G, H) and rubrospinal fibers interact with last‐order interneurons (cf. I‐K) in trisynaptic cutaneous reflex pathways in the lumbar spinal cord.

AF from Lundberg and Voorhoeve 425; G, H adapted from Lundberg et al. 420; IK adapted from Baldissera et al. 35; L adapted from Hongo et al. 277
Figure 16. Figure 16.

EPSPs from group II muscle afferents, joint afferents, and cutaneous afferents in a motoneuron to flexor digitorum longus. Upper traces, intracellular records; lower traces, incoming volleys recorded from dorsal root entry zone in L7. Stimulus strengths are given in multiples of threshold for each nerve. Voltage calibration applies to intracellular records. AD: graded stimulation of nerve from flexor digitorum longus (FDL; maximal homonymous Ia EPSP at 2 times threshold in A). E‐H: graded stimulation of nerves from posterior biceps and semitendinosus (PBSt). IK: graded stimulation of posterior knee joint nerve. L: stimulation of cutaneous afferents in sural nerve (Sur). To estimate the central latency of group II effects, it is necessary to relate them to arrival of the group II volley to the spinal cord. The group II incoming volley was therefore recorded at end of experiment from transected dorsal root. Minimum linkage for group II EPSP in extensor motoneurons seems to be disynaptic. Notice similarity of group II actions from FDL and PBSt as well as from joint and skin afferents. (From A. Lundberg, K. Malmgren, and E. Schomburg, unpublished observations; see ref. 416.)

Figure 17. Figure 17.

EPSPs from group II muscle afferents, cutaneous afferents, and joint afferents in a lamina VII interneuron. Upper traces, intracellular records; lower traces, incoming volleys recorded from dorsal root entry zone in L7. Stimulus strengths are given in multiples of threshold for each nerve. Voltage calibration applies to intracellular records. AD: graded stimulation of nerve from gastrocnemius‐soleus (G‐S). E, F: graded stimulation of nerve from flexor digitorum longus (FDL). G, H: stimulation of quadriceps (Q) and sartorius (Sart) nerves. I, J: stimulation of sural nerve (Sur). K: stimulation of posterior knee joint nerve (joint). Central latency from arrival of group II incoming volley to spinal cord (as judged from recording from transected dorsal roots at end of experiment) of the group II effects from G‐S (C, D) and from FDL (G) indicates a monosynaptic linkage from group II afferents. Note additional convergence of polysynaptic EPSPs from group II afferents (C, D, H) and from cutaneous and joint afferents. This type of interneuron has a pattern of convergence that makes it a likely candidate for transmitting the type of reflex actions illustrated in Fig. 16. (From A. Lundberg, K. Malmgren, and E. Schomburg, unpublished observations.)

Figure 18. Figure 18.

Interaction of descending tracts with flexor reflex afferent (FRA) pathways. Upper traces show intracellular records from a posterior biceps‐semitendinosus motoneuron (A‐C), 2 gastrocnemius‐soleus motoneurons (D‐F and G‐I), and a tibial motoneuron (J‐L). Lower traces are records from dorsal root entry zone. Voltage calibrations apply to intracellular records. Stimulation strengths of gastrocnemius‐soleus (G‐S) in AC and JL and plantaris (Pl) in DF are indicated in times threshold (xT) of lowest threshold afferents. Stimulation of posterior knee joint nerve (joint) in GI activated high‐threshold joint afferents. Stimulation strength of ipsilateral Deiters' nucleus (ND) is given in μA. AI: facilitation of excitatory (A‐C) and inhibitory (D‐I) FRA actions from cortex (postsigmoid gyrus). A, D, G show responses to nerve stimulation alone. B, E, H show responses to stimulation of cortex alone. C, F, I show effects of combined stimulation. Initial negative potential in A and C is a Ia field potential (same shape at an extracellular position), the short‐latency EPSP in D and F is a heteronymous Ia EPSP. The EPSP facilitated in C is due to activation of group II fibers. The IPSP in F is due to activation of relatively high‐threshold group II afferents and group III afferents (stimulation of Pl with 11 xT was without effect). JL: facilitation of a disynaptic vestibulospinal EPSP from contralateral (co) FRA. J illustrates disynaptic vestibulospinal EPSP that is markedly increased by preceding stimulation of the coG‐S (L; G‐S alone in K). M: diagram summarizing neuronal connections revealed by records in A‐I. Stippled area symbolizes a pool of interneurons and emphasizes that the actions from both cortex and peripheral afferents are relayed by polysynaptic pathways. Actual site of convergence is not known. Records in A‐C, D‐F, G‐I do not show whether all the afferent systems are relayed via common interneurons or through separate channels. The first alternative is the most likely, however, since spatial facilitation among all these afferent systems (including cutaneous afferents) have been described in other investigations, and since many interneurons in the intermediate region display the required convergence, including corticospinal excitation 420. As indicated, the same convergence applies for interneurons in excitatory (A‐C) and inhibitory (D‐F, G‐I) pathways to motoneurons. N: diagram summarizing neuronal connections revealed by records in J‐L. Interneurons mediating disynaptic vestibulospinal excitation are also last‐order interneurons in the cross‐extensor reflex.

AI adapted from Lundberg and Voorhoeve 425; JL adapted from Bruggencate and Lundberg 72. T.‐C. Fu, E. Jankowska, and A. Lundberg, unpublished observations, quoted and illustrated in Lundberg 413
Figure 19. Figure 19.

Actions of flexor‐reflex afferents (FRA) in α‐motoneurons before and after activation of reticulospinal noradrenergic pathways by intravenous injection of L‐hydroxyphenylalanine (dopa). Upper traces are intracellular records from 4 different posterior biceps‐semitendinosus motoneurons (A‐E; F‐J; K‐O; P‐S). Lower traces are records from L7 dorsal root entry zone. Voltage calibrations refer to intracellular records. Stimulation strength of afferent nerves is indicated in times threshold of the lowest threshold afferent fibers. Stimulation of L6 ventral roots (VR; Q‐S) was supramaximal for α‐fibers. In AO the left 3 columns show effect of single stimuli at fast sweep speed (calibration below M), the right 2 columns show effect of short trains of stimuli at slow time base (calibration below N). All records were obtained in spinal unanesthetized cats. PBSt, posterior biceps‐semitendinosus; G‐S, gastrocnemius and soleus; Sur, suralis; ABSm, anterior biceps and semimembranosus; joint, posterior knee joint nerve; coH, contralateral hamstring. AE: polysynaptic short‐latency EPSPs from FRA without dopa. FJ: depression of short‐latency effects after injection of dopa and appearance of long‐latency excitation (compare I, J and D, E). Depolarization remaining in F is homonymous monosynaptic Ia EPSP in the motoneuron. KO: reappearance of short‐latency FRA actions about 2 h after dopa and parallel disappearance of long‐latency excitation. PS: reciprocal inhibition of long latency obtained in a flexor motoneuron by stimulation of contralateral (co) FRA after injection of dopa. Depression of IPSP by stimulation of L6 VR (Q, expanded in R) indicates its mediation by Ia inhibitory interneurons (L6 VR alone in S).

AO adapted from Jankowska et al. 324; PS adapted from Fu et al. 193
Figure 20. Figure 20.

Organization of a spinal network giving reciprocal activation of flexor and extensor muscles. Upper traces in AC and GI are intracellular records from 2 α‐motoneurons to posterior biceps‐semitendinosus muscles (PBSt) and gastrocnemius‐soleus muscles (G‐S) in DF and J‐L upper traces are extracellular records from interneurons. Lower traces are from dorsal root entry zone. Voltage calibrations refer to intracellular records. Stimulation strength of afferent nerves is given in times threshold of lowest threshold afferent fibers. NO are simultaneous records from efferent nerves to a flexor and an extensor muscle. All records were obtained in spinal unanesthetized cats after injection of dopa (100 mg/kg). AL: half‐center organization of interneuronal network released after dopa. In flexor motoneuron (A‐C) a train of volleys in high‐threshold muscle afferents evoked the characteristic long‐latency EPSP (B), which was effectively inhibited (C) by a preceding (cond) train of volleys in contralateral high‐threshold muscle afferents (contralateral hamstring nerve, co.H). In the extensor motoneuron (G‐I) long‐latency EPSP was evoked from contralateral high‐threshold muscle afferents (H) and inhibited from the ipsilateral posterior nerve to the knee joint (joint, I). D‐F and JL are corresponding records from interneurons that were located in a region dorsal to motor nuclei. Neuron in DF was excited from high‐threshold afferents in G‐S (E) and inhibited from contralateral high‐threshold cutaneous afferents (co. sural nerve, co. Sur; F). Neuron of JL was excited from co.H (K) and completely inhibited from high‐threshold afferents in the i.G‐S (L). M: tentative diagram showing principle organization of an interneuronal network that could account for results illustrated in A‐L, N,O. Inhibition could be exerted by pre‐ or postsynaptic actions. A single interneuron in the diagram represents a chain of neurons. N,O: alternating discharges in flexor and extensor efferents triggered by short trains of impulses in ipsilateral and contralateral FRA. Acute spinal cat pretreated with nialamide (10 mg/kg) before administration of 100 mg/kg dopa. Records from nerves to a flexor (medial sartorius, Sart) and to an extensor (medial vastus of quadriceps, Vast). N: stimulation of high‐threshold afferents in contralateral quadriceps nerve (co.Q; timing of stimulus train marked by arrow and the vertical interrupted line) during a spontaneous discharge in flexor efferents is followed by a pause in flexor nerve and a discharge in extensor nerve. Activity in flexor nerve is resumed after cessation of extensor discharge, and a 2nd extensor discharge then occurs after this flexor burst. O: stimulation of ipsilateral saphenous nerve (i.saph) evokes a long‐latency discharge in the flexor nerve that, after additional stimulation of co.Q, is followed by a series of alternating bursts in the efferent nerves. P: original hypothesis of organization of spinal “primary half‐centers” projecting to flexor and extensor muscles with reciprocal inhibition acting between them As described in text, Graham Brown 214 later proposed the existence of “interposed half‐centers” to describe interneurons (intercalated between primary afferents and motoneurons) with mutual inhibition as in the diagram in M.

AC, GI, and N from Jankowska et al. 324; DF and JL from Jankowska et al. 325; M adapted from Jankowska et al. 324; O from Jankowska et al. 324 illustrated in Lundberg 413; P from Graham Brown 213.] Graham Brown 209,211
Figure 21. Figure 21.

Inhibitory action from short‐latency flexor‐reflex afferent (FRA) pathways on transmission in long‐latency FRA pathways. Upper traces in AD are intracellular records from a motoneuron innervating posterior biceps‐semitendinosus muscles (PBSt); in EH they are extracellular records from a lumbar interneuron located dorsal to the motor nuclei. Lower traces are records from dorsal root entry zone. Voltage calibration applies to intracellular records. Both neurons were recorded in unanesthetized acute spinal cats after injection of dopa (100 mg/kg). AD: recording from PBSt motoneuron shows EPSPs of long latency and duration following stimulation of ipsilateral FRA (ipsilateral nerve to anterior biceps‐semimembranosus muscle, iABSm; stimulation strength is 37 times threshold of lowest threshold afferent fibers). Bars below AD indicate duration of repetitive stimulation. Notice that prolongation of repetitive stimulation of FRA delays onset of EPSP. EH: stimulation of ipsilateral sural nerve (iSur) elicits long‐latency and long‐lasting activation of interneuron that is typical after dopa. Notice that response is delayed with prolongation of repetitive stimulation (thickening of base lines is due to stimulation artifacts). I: tentative diagram of neuronal connections that could explain illustrated effects. A single interneuron in the diagram represents a chain of interneurons. Inhibition of short‐latency pathway A (A' and A”) from descending noradrenergic systems (NA) releases transmission in long‐latency pathway B (records AH were obtained in this condition). The prolongation in onset of late discharge found in motoneurons and interneurons of B (the interneuron in EH should belong to this population) illustrates inhibitory interaction from short‐latency on long‐latency FRA pathways. The short‐latency pathway A is subdivided into A' and A” to accomodate the finding that the noradrenergic system may block the short‐latency reflex action onto motoneurons when there is still an inhibitory action on transmission through the long‐latency pathway B. Mn, motoneuron.

A‐D, I adapted from Jankowska et al. 324; EH adapted from Jankowska et al. 325
Figure 22. Figure 22.

Diagram showing alternative reflex pathways from flexor reflex afferents (FRA) with descending (desc) excitatory connections to interneurons of these pathways and its inhibitory interactive connections with the other reflex pathways from the FRA. Mn, motoneuron.

Adapted from Lundberg 411
Figure 23. Figure 23.

Definition of propriospinal relay neurons in C3‐C4 segments. Experimental arrangement and neuronal circuits as illustrated in G (for records A‐F) and in N (for records H‐M). Upper traces, intracellular records from 2 biceps (Bi) motoneurons (A‐F; H‐M); lower traces, records from C6 dorsal root entry zone. Voltage calibration refers to intracellular records. Stimulation strength of contralateral pyramid (Pyr) is given in μA. Dots below surface records in B, C, I, J indicate arrival of 3rd pyramidal volley. Dashed lines below records B, E, I mark sections that are shown at an expanded time scale in C, F, J. AF: disynaptic (1.5 ms) Pyr EPSP before (A‐C) and after (D‐F) transection of corticospinal tract (CST) in C5. Comparison of antidromic spike potentials in A and D suggests that recording conditions before and after lesion were similar. Results indicate presence of propriospinal neurons (P) above the C5 lesion that transmit disynaptic corticospinal excitation to forelimb motoneurons (FMn). HM: near disappearance of disynaptic Pyr EPSP after CST transection at rostral (r) C3. Grading of the Pyr stimulation strength in L and M suggests that the very small disynaptic EPSP that remained after lesion was a unitary response. These results indicate that the most rostral location of the propriospinal neurons intercalated in disynaptic corticomotoneuronal pathway is in caudal C2 or rostral C3.

Adapted from Illert et al. 314
Figure 24. Figure 24.

Convergence on C3‐C4 propriospinal neurons (P) from corticospinal (CST) and rubrospinal (RST) tracts and from forelimb cutaneous afferents (superficial radial nerve, SR). Experimental arrangement and neuronal circuits as illustrated in J. Convergence was judged indirectly from intracellular recording from a forelimb motoneuron (FMn; upper traces in A‐H) with CST transected completely in C5 (compare in I records from lateral funiculus in C6 before, upper trace, and after, lower trace, the C5 lesion). The same lesion transected the RST only partially. Lower traces in AH are records from C6 dorsal root entry zone. Voltage calibration applies to intracellular records. Stimulation strengths: SR, 1.5 times threshold of lowest threshold fibers; red nucleus (NR), 65 μA; pyramid (Pyr), 200 μA. Note that disynaptic EPSP appears only in A, B when 2 rubral stimuli and 1 pyramidal stimulus are given together with the cutaneous SR volley that indicates excitation of common propriospinal neurons (P). All other possible combinations were ineffective (C‐H). Dashed line below A indicates the part of the traces that is expanded in B.

Adapted from Illert et al. 314
Figure 25. Figure 25.

Convergence on C3‐C4 propriospinal neurons (P). Scheme in H summarizes monosynaptic excitatory convergence onto them as well as their direct projection to forelimb motoneurons (FMn). Intracellular records in A‐G (upper traces) illustrate a propriospinal neuron (identified from the lateral funiculus in C5 in A) with monosynaptic excitation from forelimb group I muscle afferents (D‐G: deep radial nerve, DR; stimulation strength in times threshold of the lowest threshold fibers), from corticospinal tract (B: stimulation of contralateral pyramid, Pyr, with 100 μA), and from rubrospinal tract (C: stimulation of contralateral red nucleus, NR, with 100 μA). Lower traces are records from dorsal root entry zone in C7 (in G also from C3). Voltage calibration refers to intracellular records.

AG adapted from Illert et al. 311; H from data in refs. 54,226,310,311,314
Figure 26. Figure 26.

Collateral projection of C3‐C4 propriospinal neurons (P) to forelimb segments and to lateral reticular nucleus (LRN). Scheme in G illustrates experimental arrangement and neuronal circuits. A‐F: upper traces, intracellular records from a propriospinal neuron in C3 (antidromic activation from the C7 segment in E); lower traces, records from C3 dorsal root entry zone. Voltage calibration applies to intracellular records. A, B show antidromic activation of propriospinal neuron from LRN (stimulation strength in μA); note monosynaptic IPSP from LRN in A (see Fig. 39). C illustrates monosynaptic EPSP from contralateral pyramid (Pyr, 100 μA). DF: collision of antidromic spike (F) evoked in propriospinal neuron from C7 (E) by preceding LRN stimulation (D).

AF adapted from Illert and Lundberg 309; G, data for descending projection from Illert et al. 311,314; data for direct projection to forelimb motoneurons (FMn) from Illert et al. 314 and Grant et al. 226
Figure 27. Figure 27.

Segmental afferent input depolarizing terminals of muscle spindle Ia‐afferents (A), Golgi tendon organ Ib‐afferents (B), and cutaneous afferents (C). Approximate relative amount of depolarization contributed by each input has been estimated from results quoted in text and is indicated by width of arrows. Mn, motoneurons; Ia, Ib, II, and III are the respective muscle afferents; Cut, myelinated cutaneous fibers. A: ipsilateral inputs depolarizing Ia‐fibers of flexor (left) and extensor (right) muscles. B: ipsi‐ and contralateral inputs depolarizing Ib‐fibers. C: ipsi‐ and contralateral inputs depolarizing cutaneous afferent fibers. Note that this pattern of effects is partly changed after administration of dopa to acute spinal cat (described in the text).

From Schmidt 518
Figure 28. Figure 28.

Inhibitory action from flexor reflex afferents (FRA) on transmission to Ia‐afferents illustrated by recording Ia EPSPs in motoneurons (A‐D), by the antidromic discharge of Ia‐afferents following stimulation of their terminals in the spinal cord (E‐H), and by recording dorsal root potentials (I‐L). A‐D: upper traces, intracellular records from a gastrocnemius‐soleus motoneuron (G‐S); lower traces, from L7 dorsal root entry zone. Voltage calibration refers to intracellular records. Stimulus strengths are given in multiples of threshold for lowest threshold fibers. Homonymous test Ia EPSP is shown in A. In B there is no effect by a conditioning volley in the sural (Sur) nerve. Depression in C is evoked by a short train of maximal Ia volleys in nerve from posterior biceps‐semitendinosus (PBSt). With combined conditioning with sural and PBSt nerves there is a removal of the depression (D). E‐H: intraspinal excitability measurements of Ia‐afferent terminals in gastrocnemius‐soleus (G‐S) motor nucleus. Test response (E) was recorded in G‐S nerve. Conditioning stimulation of superficial peroneal nerve (SP; cutaneous afferents only) does not change excitability (F). The facilitation (G, reflecting an excitability increase) is evoked by a train of PBSt Ia volleys. Record in H shows that a volley in SP can remove facilitatory action from PBSt Ia volleys. I‐L: upper traces are dorsal root potentials (DRPs) recorded from the most caudal dorsal rootlet in L6. Lower traces are recorded from L7 dorsal root entry zone. I shows test DRP evoked from a train of maximal Ia volleys from PBSt. J shows response to stimulation of the sural nerve. In K and L there is combined stimulation of sural and PBSt nerves (cond. + test). Note pronounced depression of DRP from PBSt Ia‐afferents (K) as well as the long duration of this effect (L). M: tentative diagram showing connections through which volleys in FRA depress transmission from Ia to Ia‐terminals. Terminals with 2 branches are excitatory. Circles indicate presynaptic terminals making synaptic contacts with presynaptic terminals. In the diagram a single interneuron may represent a chain of interneurons.

AL from Lund et al. 402; diagram in M from Lundberg 406
Figure 29. Figure 29.

Long‐latency depolarization evoked from flexor reflex afferents (FRA) in Ia‐terminals after dopa administration. Unanesthetized decorticate acute spinal cat. Graphs in AC give excitability measurements of Ia‐afferent terminals from gastrocnemius‐soleus nerve (G‐S) before (A) and after (B, C) injection of dopa. Intraspinal stimulation of G‐S afferent terminals with a microelectrode (test) was conditioned (cond) with a short train of stimuli to anterior biceps–semimembranosus (ABSm) nerve (stimulation strength is indicated in times threshold, XT, of lowest threshold afferent fibers). Abscissa gives interval between 1st conditioning volley and test stimulus to G‐S terminals. Ordinate gives amplitude of conditioned response in percent of unconditioned test response. Comparison of A and C shows that after dopa FRA stimulation causes an increased excitability of Ia‐fibers of long latency (conditioning group I stimulation is without effect, B). DG illustrate simultaneous recordings of dorsal root potentials (DRPs) evoked from ABSm nerve. Comparison of E and G shows that parallel to excitability increase in Ia‐terminals following stimulation of FRA there is also a long‐latency DRP. H: tentative diagram of connections from FRA to Ia‐afferents. Filled circles, connections exerting either pre‐ or postsynaptic inhibition; open circles, termination on presynaptic terminals. Excitatory synaptic terminals are indicated by two branches. Each symbol represents a chain of interneurons. It is assumed that activity in short‐latency FRA pathway (left) inhibits long‐latency pathway to Ia‐afferent terminals. Inhibition from a descending noradrenergic (NA) pathway of short‐latency FRA system releases transmission in long‐latency path from FRA to Ia‐terminals.

AG adapted from Andén et al. 16; H from Lundberg 406
Figure 30. Figure 30.

Rubrospinal facilitation of segmental transmission to primary afferents. A–G: upper traces, dorsal root potentials (DRPs) recorded from the most caudal filament of the L6 dorsal root; lower traces, records from cord dorsum at L7. Voltage calibration beside F applies to DRPs of A–F. A–F: facilitation from red nucleus (NR) of DRPs evoked from cutaneous (B, C: sural nerve, Sur, stimulation at 3 times threshold of lowest threshold afferent fibers) and joint afferents (E, F: high‐threshold afferents in posterior knee joint nerve). Increased amplitude of DRP in C is due to facilitation of 2nd component of cutaneous DRP that is mediated by an FRA pathway. G, H: origin of DRPs evoked by tegmental stimulation. The DRPs of G were evoked with 100‐μA stimulation at sites marked in H in a drawing of a correspnding transverse section of brain stem. The DRPs are evoked from a rather restricted area within the NR. CA, central aqueduct; NIII, oculomotor nucleus; L, lateral and H, horizontal Horsley‐Clarke coordinates.

Adapted from Hongo et al. 276
Figure 31. Figure 31.

Depression by dorsal reticulospinal system of flexor reflex afferent (FRA) inhibition in gastrocnemius‐soleus (G‐S) motor nucleus. A: schematic drawing of experimental arrangements. Right, 3 transverse sections: from brain stem (about 6 mm rostral of obex) showing the stimulation; from lower thoracic cord, hatched area indicating transection of whole cord except the right dorsolateral fascicle; from lumbar cord with microelectrode recording in left ventral horn. Left ventral quadrant and dorsolateral fascicle are mounted on electrodes for recording ascending discharges. Ventral root (VR) discharges are recorded from S1 and L7 segments, and dorsal root potentials are recorded in a caudal filament of the L6 dorsal root (DR fil.), both on left. B–E: upper traces, monosynaptic reflexes recorded from ventral root; lower traces, incoming volleys at the dorsal root entry zone. Recording was done with 2 sweep speeds simultaneously and records therefore consist of double sets (left traces at slow and right traces at fast sweep speeds; see calibrations below E). B: monosynaptic reflex from group Ia‐afferents in G‐S nerve. C: same reflex inhibited by a preceding volley in high‐threshold afferents (20 times threshold for lowest threshold fibers) from anterior biceps and semimembranosus (ABSm). D, E: same as B and C respectively but conditioned by a train of 5 stimuli in brain stem that removes most of FRA inhibition (compare C and E), without any effect on monosynaptic test reflex itself (compare B and D). F: diagram of a transverse brain stem section about 6 mm above obex. Filled circles of different size indicate points of stimulation and magnitude of effect on FRA inhibitory actions; X indicates points stimulated without effect on transmission from FRA. G, H: dorsal root records (upper traces) showing that no dorsal root potential (DRP) is evoked by the same brain stem (BS) stimulation that was used in D, E, whereas there is a large DRP with the same amplification from a single volley in the sural nerve (Sur) at 20 times threshold for the lowest threshold fibers. I: time course of brain stem action illustrated in C and E. Inhibition of monosynaptic reflex by the ABSm volley (100% refers to inhibition seen without preceding brain stem stimulation) is plotted against time interval between onset of brain stem stimulation and arrival of ABSm volley at the cord.

Adapted from Engberg et al. 167
Figure 32. Figure 32.

Action of dorsal reticulospinal system in lumbar interneurons. Upper traces in AJ are records from 3 different interneurons (A–C; D–F; G–J) located at 1.8 mm–2.0 mm depth from cord dorsum. Lower traces are records from dorsal root entry zone (not present in C, F, I). Voltage calibrations refer to intracellular records. Stimulation strength for nerves is indicated in times threshold of lowest threshold afferent fibers. Experimental arrangement is shown in Fig. 31 A. A–F: excitation from high‐threshold fibers of posterior knee joint nerve (A: joint) and inhibition from high‐threshold muscle afferents of gastrocnemius‐soleus nerve (D; G‐S) were both depressed by conditioning stimulation of brain stem (B, E; BS). BS stimulation alone evoked neither postsynaptic potentials in interneurons (C, F) nor dorsal root potentials (not illustrated), which indicates inhibition from BS in FRA pathways to interneurons from which recording was done. GJ: interneuron with synaptic inhibition from BS. G: monosynaptic activation from superficial peroneal nerve (SP; stimulation at 6 times threshold for lowest threshold fibers). H: inhibitory postsynaptic potential (IPSP) evoked by 3 stimuli in BS. I: onset of IPSP evoked by a single shock (beginning of 1st deflection is marked by arrow). J: IPSP evoked by a single stimulus of contralateral dorsolateral funiculus (DLF) just rostral to spinal cord lesion. A comparison of the latencies to onset of inhibition in I and J suggests that responsible descending fibers have a conduction velocity of 25 m/s.

Adapted from Engberg et al. 168
Figure 33. Figure 33.

Effect of dopa on transmission from high‐threshold muscle afferents in motoneurons and reversal of effects by the α‐receptor blocker phenoxybenzamine. Graphs show effect of single conditioning volleys on monosynaptic test reflex (MSR) from flexor posterior biceps‐semitendinosus nerve (PBSt). Ordinate gives amplitude of MSR in percent of unconditioned MSR. Abscissa gives interval between arrival at spinal cord of conditioning and testing group I volleys. Strength of conditioning stimulation of plantaris nerve (PI) is indicated in times threshold (xT) of lowest threshold afferent fibers. A was evoked before injection of dopa; B evoked 10 min after injection of dopa; and C evoked 25 min after dopa and 15 min after injection of phenoxybenzamine. The FRA facilitation of the flexor MSR (A) is abolished after dopa (B), which indicates inhibition of transmission in short‐latency flexor reflex afferent (FRA) pathways. Reversal of inhibition by phenoxybenzamine indicates that effect of injected dopa is mediated by α‐receptors.

Adapted from Andén et al. 15
Figure 34. Figure 34.

Release of Ib‐ and flexor reflex afferent (FRA) actions after bilateral lesions of dorsolateral funiculi. Graphs show effect of single conditioning volleys in nerve to flexor digitorum longus (FDL) on monosynaptic test reflexes evoked from gastrocnemius‐soleus (G‐S). Ordinate gives amplitude of reflexes in percent of unconditioned test reflex. Abscissa gives interval between arrival at spinal cord of conditioning and testing group I volleys. Strength of conditioning stimulation (cond) is indicated in times threshold (xT) of lowest threshold afferent fibers. It was chosen to activate group I fibers in A and D, groups I and II fibers in B, and groups I, II, and III fibers in C and E. At each conditioning strength the curves were obtained after the ipsilateral (i) and contralateral (co) lesions indicated schematically in A (for AC) and in D (for D, E). Transverse spinal cord drawings in A show symbols used in AC to illustrate effects of sequential lesions; the initial lesion of the dorsal column and dorsolateral on the contralateral side, and ventral funiculi (O), the additional transection of ipsilateral ventral funiculus (x, cf. histological reconstruction to left), and the complete spinal transection (•). Drawings in D similarly show initial lesion of dorsal column and dorsolateral on the ipsilateral side, and ventral quadrants (O), the additional lesion of the contralateral dorsal funiculus (x, cf. histological reconstruction to left), and complete spinal transection (•).

Adapted from Holmqvist and Lundberg 267
Figure 35. Figure 35.

Differential release of transmission in excitatory and inhibitory flexor reflex afferent pathways by brain stem lesions. Graphs show effect of single conditioning volleys in nerve to flexor digitorum longus (FDL) on monosynaptic test reflexes evoked from gastrocnemius‐soleus (G‐S, AC) or from posterior biceps‐semitendinosus (PBSt, D, E). Ordinate gives amplitude of reflexes in percent of unconditioned test reflexes. Abscissa gives interval between arrival at spinal cord of conditioning and testing group I volleys. Strength of conditioning stimulation (cond) is indicated in times threshold (xT) of lowest threshold afferent fibers. It was chosen to activate group I fibers in A and D, groups I and II fibers in B, and groups I, II, and III fibers in C and E. At each conditoning strength the curves were obtained after brain stem lesions indicated in F.

Adapted from Holmqvist and Lundberg 268
Figure 36. Figure 36.

Monosynaptic connections to α‐motoneurons from corticospinal and rubrospinal tracts in primate (rhesus monkey). Lower traces in AF are intracellular records from 2 motoneurons. Upper traces are cord dorsum potentials monitoring afferent incoming volleys and corticospinal volleys. All traces in G, H are intracellular records. AF: frequency potentiation of monosynaptic excitatory postsynaptic potential (EPSP) evoked from motor cortex (MC) in a motoneuron to flexor digitorum longus (FDL, in AC) and to gastrocnemius‐soleus (G‐S, in DF). Antidromic invasion from FDL and G‐S nerves in A and D, respectively. Monosynaptic EPSPs following a single shock stimulation of motor cortex are shown in B, E. Growth of monosynaptic EPSPs with repetitive stimuli shown in C, F; same stimulus strengths as in B, E, respectively. Notice the greater potentiation in F than in C. Calibration pulse 1 ms, 5 m V in A, D; 1 ms, 2 m V in B, C, E, F. G, H: convergence of monosynaptic EPSPs from motor cortex (MC) and red nucleus (NR) in an FDL motoneuron. Superimposed traces of monosynaptic EPSPs evoked by stimulation of MC (G) and NR (H) of different intensities given to right of each trace. Note differences in amplitude and time course from MC and NR.

AF from Tamarova et al. 574; G, H from Shapovalov 539
Figure 37. Figure 37.

Diagram illustrating 2 different ways in which ascending neurons may monitor interneuronal activity. A: by ascending collaterals from interneurons interposed in segmental reflex pathways. B: by separate ascending neurons that receive collateral projections from interneurons interposed in segmental reflex pathways. Mn, motoneuron; asc., ascending neuron; prim. aff., primary afferent.

Figure 38. Figure 38.

Decerebrate control of transmission to ascending ventral pathways activated from flexor reflex afferents (FRA). Diagram in J illustrates experimental arrangement with a bilateral section of ventral quadrants (VQ) and recording from right VQ of ascending discharge evoked from highthreshold afferents in left hamstring nerve (1. Ham; A, D, G) and left sural nerve (1. Sur; B, E, H). C, F, I show background activity in VQ without stimulation. Comparison of records obtained before (AC), during (DF), and after (GI) cooling of dorsal part of spinal cord (the dorsal column removed) with a thermode illustrates effective blockade of transmission from FRA to ascending tracts in decerebrate state with conducting dorsolateral funiculi (A–C; G–I), and release in the spinal state, i.e., during blockade of impulse transmission in dorsal part of the spinal cord (DF).

Adapted from Holmqvist et al. 269
Figure 39. Figure 39.

Wiring diagrams illustrating 2 examples of ascending information from collaterals of neurons interposed in spinal neuronal circuits. A: ascending collaterals from propriospinal neurons in C3–C4 (P) to lateral reticular nucleus (LRN). Mn, motoneuron. This connection is illustrated in Fig. 26. B: ascending collaterals to LRN from inhibitory interneurons (with monosynaptic excitation from forelimb afferents) projecting to propriospinal neurons in C3‐C4.

A from data of Illert and Lundberg 309; B from unpublished results of B. Alstermark, M. Illert, and A. Lundberg (see ref. 413
Figure 40. Figure 40.

Wiring diagrams illustrating hypothesis that information transferred by ventral spinocerebellar tract (VSCT) neurons reflects relation between input to and output from inhibitory segmental interneurons. A: diagram illustrating convergence on a VSCT neuron of monosynaptic excitation from primary afferents (prim. aff.) and disynaptic inhibition evoked from the same primary afferents. Disynaptic inhibition of VSCT neuron is due to a collateral projection from interneurons transmitting disynaptic inhibitory postsynaptic potentials to motoneurons (Mn). In this case the VSCT neuron would compare excitatory input to interneurons from primary afferents and the resulting activity of the inhibitory interneuron. B: in addition to connections in A there is also a descending motor pathway (desc) giving monosynaptic excitation and disynaptic inhibition of VSCT neuron. The descending monosynaptic excitation is collateral to excitation of interneurons, and the disynaptic inhibition is again due to collateral projection from interneuron. Such a VSCT neuron thus compares excitatory input to interneurons from primary afferents and the descending pathway with the resulting activation of the interneuron.

Figure 41. Figure 41.

Recurrent depression of a Ia inhibitory postsynaptic potential (IPSP) in a ventral spinocerebellar tract (VSCT) neuron. Experimental arrangement and relevant neuronal circuits are illustrated in A. Mn, motoneuron; RC, Renshaw cell; Q, quadriceps. Upper traces in BE are intracellular records from a VSCT neuron (identification from ipsilateral anterior lobe of cerebellum in E). Lower traces are records from L5 dorsal root entry zone. Voltage calibrations refer to intracellular records. BD: disynaptic Ia IPSPs from Q nerve. Stimulation strength is indicated in times threshold of lowest threshold afferent fibers. F: averaged records of submaximal Ia IPSPs from Q (upper trace) combined with conditioning stimulation of L5–S1 ventral roots (VRs; lower trace). Arrow below lower trace indicates arrival of the VR volley to the spinal cord (calibration pulse 1 mV, 2 ms).

Adapted from Gustafsson and Lindström 251
Figure 42. Figure 42.

Facilitation of Ia inhibitory postsynaptic potentials (IPSPs) in ventral spinocerebellar tract (VSCT) neurons from vestibulospinal and rubrospinal tracts. A shows experimental arrangement and relevant neuronal circuits. Mn, motoneurons; ND, Deiters' nucleus; NR, red nucleus. Upper traces in BD and FH are intracellular records from 2 different VSCT neurons. Lower traces are from dorsal root entry zone. Voltage calibrations refer to intracellular records. Strength of stimulation in ND and NR is indicated in μA, stimulation of nerves in times threshold of lowest threshold afferent fibers. BE: facilitation of a quadriceps‐evoked (Q) Ia IPSP from the ND. Graph in E gives amplitude of IPSP in percent of unconditioned test (ordinate) in relation to time interval between arrival of descending volley from ND (zero ms on abscissa) and incoming Q volley (interrupted line indicates synchronous arrival). Facilitation started with Q volley arriving slightly before ND volley, which indicates monosynaptic coupling of descending fibers with interneuron mediating the Ia IPSP. FI: facilitation from NR of IPSPs evoked from posterior biceps‐semitendinosus nerve (PBSt). Records in FH show a facilitation of a Ia IPSP from PBSt. Graph in I shows time course of facilitation of a Ia IPSP from Q in another VSCT neuron. Amplitude of IPSP is given in percent of unconditioned test (ordinate) in relation to time interval in arrival of 1st rubrospinal volley (zero ms on abscissa) and the incoming Q volley (interrupted line indicates synchronous arrival).

BE adapted from Baldissera and Roberts 40; FI adapted from Baldissera and Bruggencate 34


Figure 1.

Indirect technique used to investigate convergence in interneurons of reflex pathways. Diagram exemplifies convergence from primary afferents (I, prim. aff.) and descending pathways (II, desc.) on common excitatory interneurons projecting to motoneurons. Neuronal circuits are illustrated on left. A: example of excitatory convergence from both sources. B: example of convergence of excitation from prim. aff. and inhibition from desc. pathways. Interneurons represent a population of interneurons with identical convergence. Traces on right give idealized intracellular records from a single motoneuron after the test stimulus (I, prim. aff.), the conditioning (cond.) stimulus (II, desc.), and combined stimulation (I+II). This example refers to convergence from descending fibers and primary afferents, but the technique can be used for convergence from any source. See text for explanation.

Adapted from Lundberg 412 in The Nervous System. The Basic Neurosciences. ©1975, Raven Press, N. Y


Figure 2.

Relations between individual motoneurons and individual Renshaw cells. A: illustration of experimental arrangement and neuronal circuit. Excitatory effect on Renshaw cells (RC) by single impulses in motoneurons (Mn) and inhibitory effect of individual Renshaw cells onto individual motoneurons were investigated with the aid of 2 microelectrodes: 1 used for intracellular stimulation of and recording from motoneurons, and 1 used for extracellular recording from Renshaw cells. The 2nd electrode was filled with 3 M sodium glutamate, which allowed electrophoretic application of glutamate to induce repetitive firing of the recorded Renshaw cells. B: 2 samples of burst firing in a Renshaw cell (upper traces, large spikes) following single action potentials in the motoneuron (timing shown by lower trace). Stimulation frequency of motoneuron was 8.5 impulses/s. Time calibration in D applies also to records in B. C: graph showing rapid decrease of number of spikes in Renshaw cell burst firing following each motoneuronal spike with increasing stimulation frequency. Dots represent mean value of several trials (cf. records in B) at different stimulation frequencies. This example shows the most powerful excitatory coupling of 21 tested cases. D: slow repetitive firing of Renshaw cell (upper trace, large spike) with simultaneous recording of membrane potential of motoneuron (lower trace; voltage calibration applies to intracellular recording). Same motoneuron and Renshaw cell that was illustrated in B, C, E: averaged record of membrane potential in motoneuron following single discharges of the Renshaw cell (the Renshaw cell firing with “doublets” or “triplets” are excluded). Same motoneuron and Renshaw cell that was illustrated in B–D. Note remarkably long time course of this unitary IPSP. F: a more typical time course of a presumed unitary IPSP from a Renshaw cell in a different motoneuron. Pretriggered averaging was used in both E and F; zero on abscissa refers to a time 1 ms before occurrence of spike of Renshaw cell. (From L. Van Keulen, unpublished material. see ref. 592.)



Figure 3.

Comparison of patterns of monosynaptic excitation from muscle spindle Ia‐afferents and recurrent inhibition via motor axon collaterals in motor nuclei supplying various hindlimb muscles in cat. Experimental arrangement and neuronal circuits as illustrated in the scheme. Motor nuclei recorded from are listed to left of diagram and nerves whose orthodromic and antidromic effects were compared are indicated above. Homonymous connections, always with both Ia excitation and recurrent inhibition, are indicated by dotted areas. Heteronymous Ia excitation and recurrent inhibition are indicated by hatched and shaded areas, respectively. Diagram is constructed from data obtained in the cat. It indicates a complete overlap of Ia excitation and recurrent inhibition, but with recurrent inhibition more widely distributed than Ia excitation. Sart, sartorius; Q, quadriceps; Sm, semimembranosus; AB, anterior biceps; St, semitendinosus; PB, posterior biceps; Per, peroneus; G‐S, gastrocnemius‐soleus; FDL, flexor digitorum longus; Pl, plantaris; Add, adductor femoris and longus; Grac, gracilis.

Data for Ia pathways from Eccles et al. 135 and Eccles and Lundberg 147; data for recurrent inhibition from Hultborn et al. 292, Eccles et al. 131, and Wilson et al. 613


Figure 4.

Convergence of recurrent inhibition in a primate (baboon) gracilis (Grac) motoneuron with correlation to location of motor nuclei whose efferent axons have been stimulated. A: upper traces are intracellular averaged records (calibration pulse 1 mV, 4 ms) from the gracilis motoneuron. Lower traces are simultaneous records from the cord dorsum. Dorsal roots have been cut and nerves indicated have been stimulated at supramaximal strength for α‐fibers (arrival of volleys at spinal cord indicated by arrows). B: segmental location of motor nuclei tested in A (at the end of each experiment the rostrocaudal distribution of motor nuclei were tested by stimulation of individual ventral roots with simultaneous recording from peripheral muscle nerves). Position of recorded gracilis motoneuron is indicated by arrow. Results suggest that recurrent inhibition is not organized according to proximity principle. Note large recurrent IPSPs from semitendinosus and posterior biceps whose motor nuclei are among the most caudal ones, but no effect at all from quadriceps despite a similar rostrocaudal location of quadriceps and gracilis motor nuclei. Q, quadriceps; Sm, semimembranosus; Pl, plantaris; AB, anterior biceps; Sart, sartorius; TA, tibialis anterior; EDL, extensor digitorum longus; St, semitendinosus; PB, posterior biceps. (From H. Hultborn, E. Jankowska, and S. Lindström, unpublished material. Cited in ref. 292.)



Figure 5.

Diagrams illustrating the hypothesis that Renshaw system serves as a variable gain regulator at motoneuronal level. A: input and output connections of α‐motoneurons and Renshaw cells. B: simplified diagram of input‐output relations of a motoneuronal pool during inhibition and facilitation of transmission in recurrent pathway. C: concept of motor output stage. Neurons constituting output are framed by thick lines. Hatched lines indicate parallel connections to α‐ and γ‐motoneurons and corresponding Ia inhibitory interneurons. α, α‐Motoneurons; γ, γ‐motoneurons; RC, Renshaw cells; Ia IN, Ia inhibitory interneurons.

Adapted from Hultborn et al. 293


Figure 6.

Distribution of heteronymous monosynaptic excitation evoked by impulses in muscle spindle Ia‐afferents from proximal hindlimb muscles in cat. Ia‐afferents of an individual muscle (indicated by coils) evoke excitatory effects via monosynaptic pathways in α‐motoneurons of muscles (marked by open triangles). TFL, musculus tensor fasciae latae.

Adapted from Eccles and Lundberg 147


Figure 7.

Long‐lasting excitability increase of α‐motoneurons induced via polysynaptic Ia pathways. B, C: simultaneous intracellular records at different gains from a triceps surae motoneuron in a decerebrate, decerebellate cat (experimental arrangement in A). Ventral roots were intact, but γ‐loop was opened by relaxation with gallamine triethiodide (Flaxedil). Short stimulus trains to medial gastrocnemius nerve (on: MG) caused a depolarizing shift of membrane potential and a firing of motoneuron. This effect was reversed by a short stimulus train to superficial peroneal nerve (off: SP). D, E: monosynaptic reflexes (MSR) evoked by stimulation of nerve from medial gastrocnemius and recorded from central end of cut S1–L7 ventral roots (γ‐loop open) in a spinal unanesthetized cat after injection of 5‐hydroxytryptophan (50 mg/kg; experimental arrangement in D). E: integrated amplitude of MSR (arbitrary units, horizontal lines give mean ± SEM) is plotted against time. Stimulation of lateral gastrocnemius‐soleus nerve (on: LG‐S) increased amplitude of MSR while a short stimulus train to the SP nerve (off: SP) brought it back to control value. Values obtained during LG‐S or SP stimulation are not included in the calculation.

AC adapted from Hultborn and Wigström 301; D, E from H. Hultborn and H. Wigström, unpublished observations


Figure 8.

Distribution of disynaptic inhibition evoked by impulses in muscle spindle Ia‐afferents from proximal hindlimb muscles in cat. Ia‐afferents of an individual muscle (indicated by coils) evoke inhibitory actions via disynaptic pathways in α‐motoneurons of muscles (marked by closed triangles). TFL, musculus tensor fasciae latae.

Adapted from Eccles and Lundberg 147


Figure 9.

Convergence of Ia‐afferents on α‐motoneurons and interneurons that mediate disynaptic Ia inhibition to motoneurons of antagonists (Ia inhibitory interneurons). Scheme in G illustrates experimental arrangement and neuronal circuits. A–C: upper traces are intracellular records from a semitendinosus (St) motoneuron (identified by antidromic invasion in A). Lower traces are records from L7 dorsal root entry zone. Voltage calibrations apply to intracellular records. Activation of Ia‐afferents from St, posterior biceps (PB), and gracilis (Grac) muscles evokes monosynaptic EPSPs. DF: intracellular (averaged) records from a vastocrureus (V–Cr) motoneuron (direct antagonist to St). Calibration pulse: 2 ms, 0.5 mV. Stimulation strength (in times threshold of lowest threshold afferent fibers): D) PB, 1.2; St, 1.1; E) Grac, 1.3; St, 1.05; F) Grac, 1.3; PB, 1.2. Appearance of disynaptic IPSPs in motoneuron (bottom traces in DF) upon conjoint stimulation of PB and St (D), Grac and St (E), and Grac and PB (F) demonstrates convergence of monosynaptic excitation from respective Ia‐afferents on common inhibitory interneurons that thus receive same excitatory convergence as the St motoneurons.

AC adapted from Eccles et al. 135; DF adapted from Hultborn and Udo 299


Figure 10.

Identification of interneurons mediating disynaptic Ia inhibition (Ia inhibitory interneurons) from quadriceps (Q) afferents to posterior biceps and semitendinosus (PBSt) motoneurons. Experimental arrangement and neuronal circuits are illustrated in D (for records AG) and J (for records I, K). Upper traces in AG are intracellular records from a PBSt motoneuron (A–C) and a lamina VII interneuron (E–G). Lower traces are records from L5 dorsal root entry zone. Upper trace in I is an extracellular record from a lamina VII interneuron that is excited by electrophoretic ejection of glutamate from the recording electrode. Lower trace is a simultaneous intracellular record from a PBSt motoneuron. K shows averaged intracellular response from I (lower trace) triggered by extracellular interneuronal spike (upper trace). Voltage calibrations refer to intracellular records. Stimulation strength of Q nerve is given in times threshold of lowest threshold afferent fibers; stimulation of ventral roots (VRs) was supramaximal for α‐fibers. AC: disynaptic Ia IPSPs from Q afferents in a PBSt motoneuron (A, B) and its depression by conditioning stimulation of L5 + L6 VRs (C). E–G: lamina VII interneuron with convergence required for mediation of effects recorded in PBSt motoneuron of AC, i.e., monosynaptic activation from Q Ia‐afferents (E, F) and recurrent inhibition from L6 VR (G). H: summarizing diagram giving location in L6 of a number of interneurons with convergence illustrated in E–G. I–K: synaptic action in a PBSt motoneuron by an interneuron with convergence shown in EG and location shown in H. Monosynaptic unitary IPSP evoked by interneuronal activity proves that interneuron with monosynaptic Ia excitation and disynaptic inhibition from motor axons mediates reciprocal Ia inhibition.

AC from H. Hultborn, E. Jankowska, and S. Lindström, unpublished records; EG adapted from Hultborn et al. 287; H adapted from Hultborn et al. 291; IK adapted from Jankowska and Roberts 336


Figure 11.

Connections to interneuron in reciprocal Ia inhibitory pathway. A: circuit diagram of some connections to interneuron in reciprocal Ia inhibitory pathway, i, Ipsilateral; co, contralateral; Vs, vestibulospinal tract; Cs, corticospinal tract; Rs, rubrospinal tract; Ps, propriospinal tract; cut, cutaneous afferents; FRA, flexor reflex afferents; Mn, motoneurons; R, Renshaw cells. BK: parallel projection from nucleus vestibularis lateralis (ND) onto α‐motoneurons and corresponding Ia inhibitory interneurons. Experimental arrangement and neuronal circuits as illustrated in K. Upper traces are intracellular records from a posterior biceps‐semitendinosus motoneuron (PBSt; B–E), a quadriceps (Q) motoneuron (I, J), and a Ia inhibitory interneuron (F–H) presumably intercalated in the Ia inhibitory pathway from Q to PBSt. Lower traces are records from dorsal root entry zone in L7 (B–E) or L6, (F–J). Voltage calibrations refer to intracellular records. Ipsilateral ND stimulated with 80 μA in CE and 200 μA in G. Q nerve (B–E, F) stimulated with 1.1 times threshold of the lowest threshold afferent fibers. Ventral roots (VR) were stimulated with single stimuli, supramaximal for α‐fibers. BE: facilitation of the Q Ia IPSP in a PBSt motoneuron from vestibulospinal tract (B–D) and depression of facilitated IPSP from L5 + L6 VRs (E). Results indicate convergence of vestibulospinal excitation and inhibition from motor axon collaterals on common Ia inhibitory interneurons projecting to PBSt motoneurons. FH: monosynaptic excitation from the ND (G) of a Q‐activated Ia inhibitory interneuron (F) and its disynaptic inhibition from L5 + L6 VRs (H). Note that these direct recordings from interneuron (F–H) correspond to convergent actions from the ND and the VR on the Ia IPSP recorded in the PBSt motoneuron (B–E). I, J: monosynaptic activation of a Q motoneuron (I, homonymous EPSP) from lateral vestibulospinal tract (J). L: schematic drawing of parallel projections (broken lines) onto α‐ and γ‐motoneurons innervating a muscle and the Ia inhibitory interneuron inhibiting the motoneurons of its antagonist. Notice also the mutual inhibition between opposite Ia inhibitory interneurons. Further description in the text.

BE adapted from Hultborn and Udo 298; FH adapted from Hultborn et al. 289; I, J adapted from Grillner et al. 240; L adapted from Hultborn et al. 287


Figure 12.

Convergence on interneurons in Ib inhibitory pathway to motoneurons. Upper traces, intracellular recordings from motoneurons to gastrocnemius‐soleus (A–C) and to flexor digitorum longus (E–H). Lower traces, incoming volleys recorded from L7 dorsal root entry zone. Voltage calibrations refer to intracellular records. Stimulus strengths of peripheral nerves are given in multiples of thresholds for lowest threshold afferents. AC: facilitatory interaction in inhibitory transmission to motoneurons from Ib muscle afferents from plantaris (Pl) and cutaneous afferents in superficial peroneal nerve (SP). EH: facilitatory interaction between a single descending volley in rubrospinal tract (stimulation of red nucleus, NR) and a Ib volley in quadriceps nerve (Q); G, H compare the lack of facilitation with weaker nerve stimulation that mainly activates Ia‐afferents in E, F. Arrow below records in H indicates time of arrival of the fastest descending volleys in rubrospinal tract to the L6 level. Corresponding graphs in D and I show time course of facilitation obtained by varying conditioning‐testing interval. Zero on abscissa indicates simultaneous arrival at segmental level of both conditioning and testing volleys. Delay of facilitation by a cutaneous volley (D) suggests a disynaptic linkage to Ib inhibitory interneurons. Facilitatory action at simultaneous arrival (and even with the conditioning volley arriving slightly after the test) in the case of rubrospinal volley (I) strongly indicates a monosynaptic linkage from rubrospinal tract. Circuit diagram summarizes most of the neuronal connections to Ib inhibitory interneurons.

AC adapted from Lundberg et al. 417; EH adapted from Hongo et al. 275


Figure 13.

Convergence onto a lamina V and VI interneuron possibly interposed in Ib reflex pathways to motoneurons. Upper traces (A‐F), intracellular records from a lamina V and VI interneuron; lower traces (A‐C) and middle traces (D‐F), records of afferent volleys from L7 dorsal root entry zone; lower traces (D‐F), records of changes in muscle length (increase in length downward). AC: convergence of monosynaptic excitation from group I afferents in nerve from plantaris (Pl), of disynaptic excitation from cutaneous afferents in superficial peroneal nerve (SP) and of monosynaptic excitation from ipsilateral descending fibers (i. desc.). Stimulus strength is given in times threshold for lowest threshold fibers. DF: analysis by graded brief stretches of plantaris muscle of contribution from muscle spindle Ia‐afferents and Golgi tendon organ Ib‐afferents to the group I excitation illustrated in A. Small stretch in D (cf. calibration to right of F) selectively activates muscle spindle Ia‐afferents, the larger muscle stretches in E and F are near maximal for Ia‐afferents and above threshold for Ib‐afferents, respectively. This series of records thus strongly indicates that both Ia‐ and Ib‐afferents contribute monosynaptic excitation. G: reconstruction of soma and part of the axonal projections of interneuron whose excitatory input is illustrated in A‐F. Interneuron was stained by ejecting horseradish peroxidase from the microelectrode after the recording. Notice the axonal projection into region of motoneurons (large neurons, probably motoneurons, are indicated by hatched areas). (From E. Jankowska, T. Johannisson, and J. Lipski, unpublished observations.)



Figure 14.

Facilitation from corticospinal tract of a cutaneous reflex originating from plantar cushion. Scheme in A illustrates experimental arrangement and the involved neuronal circuits. Stippled area symbolizes a pool of interneurons and emphasizes that the actions from both systems are relayed in polysynaptic pathways. Actual site of interaction is not known. Upper traces in BE are records from ventral root; lower traces are simultaneous recordings from dorsal root entry zone. Monosynaptic test reflex of plantar muscles (B) is moderately facilitated by a weak (conditioning) electrical stimulus to the pad (C). Stimulation of cortex (D) had no detectable effect on test reflex, but combined action from the 2 conditioning systems resulted in a large spatial facilitation (E).

Adapted from Engberg 165


Figure 15.

Interaction of cortico‐ and rubrospinal tracts with lumbar cutaneous reflex pathways. Upper traces in AK are intracellular records from different α‐motoneurons and interneurons. AC: posterior biceps‐semitendinosus motoneuron. DF: gastrocnemius‐soleus motoneuron. IK: pretibial flexor motoneuron supplied by deep peroneal nerve. G, H: L7 dorsal horn interneuron. Lower traces are records from dorsal root entry zone at the respective segmental levels. Voltage calibrations refer to intracellular records. AF: facilitation of cutaneous actions from cortex. Liminal synaptic actions evoked from sural nerve (Sur; stimulated alone in A and D) were conditioned by preceding stimulation of cortex (C and F). Cortex was stimulated alone in B and E. The ensuing increase of the polysynaptic potentials indicates excitatory convergence by corticospinal volleys and cutaneous afferent volleys on interneurons in excitatory and inhibitory cutaneous reflex pathways. G, H: monosynaptic activation of a lumbar dorsal horn interneuron from cutaneous (Sur) afferents (G) and short‐latency (probably monosynaptic) activation from the cortex (H). This pattern suggests that, in the action illustrated in AF, the corticospinal tract may project monosynaptically onto first‐order interneurons in lumbar cutaneous pathways. IK: monosynaptic excitation from rubrospinal tract of last order interneurons in a cutaneous reflex pathway. I shows a liminal IPSP evoked from red nucleus (NR) with a single shock of 200 μA (arrows below surface records in I and K mark arrival at segmental level of descending volley in rubrospinal tract). Conditioning stimulation of superficial peroneal nerve (SP) at 1.4 times threshold of the lowest threshold afferents increased the disynaptic rubrospinal IPSP (K; SP alone in J). L: comparison of location in cat lumbar cord of focal potentials evoked from rubrospinal and corticospinal tracts. Transverse section shows areas in which field potentials were evoked from indicated sites with an amplitude above 80% (black and hatched area) and 50% (continuous and interrupted lines) of their maximal amplitudes. Dorsal location of field from cortex and ventral location of NR field would be compatible with hypothesis that corticospinal fibers interact with first–order interneurons (cf. A‐F, G, H) and rubrospinal fibers interact with last‐order interneurons (cf. I‐K) in trisynaptic cutaneous reflex pathways in the lumbar spinal cord.

AF from Lundberg and Voorhoeve 425; G, H adapted from Lundberg et al. 420; IK adapted from Baldissera et al. 35; L adapted from Hongo et al. 277


Figure 16.

EPSPs from group II muscle afferents, joint afferents, and cutaneous afferents in a motoneuron to flexor digitorum longus. Upper traces, intracellular records; lower traces, incoming volleys recorded from dorsal root entry zone in L7. Stimulus strengths are given in multiples of threshold for each nerve. Voltage calibration applies to intracellular records. AD: graded stimulation of nerve from flexor digitorum longus (FDL; maximal homonymous Ia EPSP at 2 times threshold in A). E‐H: graded stimulation of nerves from posterior biceps and semitendinosus (PBSt). IK: graded stimulation of posterior knee joint nerve. L: stimulation of cutaneous afferents in sural nerve (Sur). To estimate the central latency of group II effects, it is necessary to relate them to arrival of the group II volley to the spinal cord. The group II incoming volley was therefore recorded at end of experiment from transected dorsal root. Minimum linkage for group II EPSP in extensor motoneurons seems to be disynaptic. Notice similarity of group II actions from FDL and PBSt as well as from joint and skin afferents. (From A. Lundberg, K. Malmgren, and E. Schomburg, unpublished observations; see ref. 416.)



Figure 17.

EPSPs from group II muscle afferents, cutaneous afferents, and joint afferents in a lamina VII interneuron. Upper traces, intracellular records; lower traces, incoming volleys recorded from dorsal root entry zone in L7. Stimulus strengths are given in multiples of threshold for each nerve. Voltage calibration applies to intracellular records. AD: graded stimulation of nerve from gastrocnemius‐soleus (G‐S). E, F: graded stimulation of nerve from flexor digitorum longus (FDL). G, H: stimulation of quadriceps (Q) and sartorius (Sart) nerves. I, J: stimulation of sural nerve (Sur). K: stimulation of posterior knee joint nerve (joint). Central latency from arrival of group II incoming volley to spinal cord (as judged from recording from transected dorsal roots at end of experiment) of the group II effects from G‐S (C, D) and from FDL (G) indicates a monosynaptic linkage from group II afferents. Note additional convergence of polysynaptic EPSPs from group II afferents (C, D, H) and from cutaneous and joint afferents. This type of interneuron has a pattern of convergence that makes it a likely candidate for transmitting the type of reflex actions illustrated in Fig. 16. (From A. Lundberg, K. Malmgren, and E. Schomburg, unpublished observations.)



Figure 18.

Interaction of descending tracts with flexor reflex afferent (FRA) pathways. Upper traces show intracellular records from a posterior biceps‐semitendinosus motoneuron (A‐C), 2 gastrocnemius‐soleus motoneurons (D‐F and G‐I), and a tibial motoneuron (J‐L). Lower traces are records from dorsal root entry zone. Voltage calibrations apply to intracellular records. Stimulation strengths of gastrocnemius‐soleus (G‐S) in AC and JL and plantaris (Pl) in DF are indicated in times threshold (xT) of lowest threshold afferents. Stimulation of posterior knee joint nerve (joint) in GI activated high‐threshold joint afferents. Stimulation strength of ipsilateral Deiters' nucleus (ND) is given in μA. AI: facilitation of excitatory (A‐C) and inhibitory (D‐I) FRA actions from cortex (postsigmoid gyrus). A, D, G show responses to nerve stimulation alone. B, E, H show responses to stimulation of cortex alone. C, F, I show effects of combined stimulation. Initial negative potential in A and C is a Ia field potential (same shape at an extracellular position), the short‐latency EPSP in D and F is a heteronymous Ia EPSP. The EPSP facilitated in C is due to activation of group II fibers. The IPSP in F is due to activation of relatively high‐threshold group II afferents and group III afferents (stimulation of Pl with 11 xT was without effect). JL: facilitation of a disynaptic vestibulospinal EPSP from contralateral (co) FRA. J illustrates disynaptic vestibulospinal EPSP that is markedly increased by preceding stimulation of the coG‐S (L; G‐S alone in K). M: diagram summarizing neuronal connections revealed by records in A‐I. Stippled area symbolizes a pool of interneurons and emphasizes that the actions from both cortex and peripheral afferents are relayed by polysynaptic pathways. Actual site of convergence is not known. Records in A‐C, D‐F, G‐I do not show whether all the afferent systems are relayed via common interneurons or through separate channels. The first alternative is the most likely, however, since spatial facilitation among all these afferent systems (including cutaneous afferents) have been described in other investigations, and since many interneurons in the intermediate region display the required convergence, including corticospinal excitation 420. As indicated, the same convergence applies for interneurons in excitatory (A‐C) and inhibitory (D‐F, G‐I) pathways to motoneurons. N: diagram summarizing neuronal connections revealed by records in J‐L. Interneurons mediating disynaptic vestibulospinal excitation are also last‐order interneurons in the cross‐extensor reflex.

AI adapted from Lundberg and Voorhoeve 425; JL adapted from Bruggencate and Lundberg 72. T.‐C. Fu, E. Jankowska, and A. Lundberg, unpublished observations, quoted and illustrated in Lundberg 413


Figure 19.

Actions of flexor‐reflex afferents (FRA) in α‐motoneurons before and after activation of reticulospinal noradrenergic pathways by intravenous injection of L‐hydroxyphenylalanine (dopa). Upper traces are intracellular records from 4 different posterior biceps‐semitendinosus motoneurons (A‐E; F‐J; K‐O; P‐S). Lower traces are records from L7 dorsal root entry zone. Voltage calibrations refer to intracellular records. Stimulation strength of afferent nerves is indicated in times threshold of the lowest threshold afferent fibers. Stimulation of L6 ventral roots (VR; Q‐S) was supramaximal for α‐fibers. In AO the left 3 columns show effect of single stimuli at fast sweep speed (calibration below M), the right 2 columns show effect of short trains of stimuli at slow time base (calibration below N). All records were obtained in spinal unanesthetized cats. PBSt, posterior biceps‐semitendinosus; G‐S, gastrocnemius and soleus; Sur, suralis; ABSm, anterior biceps and semimembranosus; joint, posterior knee joint nerve; coH, contralateral hamstring. AE: polysynaptic short‐latency EPSPs from FRA without dopa. FJ: depression of short‐latency effects after injection of dopa and appearance of long‐latency excitation (compare I, J and D, E). Depolarization remaining in F is homonymous monosynaptic Ia EPSP in the motoneuron. KO: reappearance of short‐latency FRA actions about 2 h after dopa and parallel disappearance of long‐latency excitation. PS: reciprocal inhibition of long latency obtained in a flexor motoneuron by stimulation of contralateral (co) FRA after injection of dopa. Depression of IPSP by stimulation of L6 VR (Q, expanded in R) indicates its mediation by Ia inhibitory interneurons (L6 VR alone in S).

AO adapted from Jankowska et al. 324; PS adapted from Fu et al. 193


Figure 20.

Organization of a spinal network giving reciprocal activation of flexor and extensor muscles. Upper traces in AC and GI are intracellular records from 2 α‐motoneurons to posterior biceps‐semitendinosus muscles (PBSt) and gastrocnemius‐soleus muscles (G‐S) in DF and J‐L upper traces are extracellular records from interneurons. Lower traces are from dorsal root entry zone. Voltage calibrations refer to intracellular records. Stimulation strength of afferent nerves is given in times threshold of lowest threshold afferent fibers. NO are simultaneous records from efferent nerves to a flexor and an extensor muscle. All records were obtained in spinal unanesthetized cats after injection of dopa (100 mg/kg). AL: half‐center organization of interneuronal network released after dopa. In flexor motoneuron (A‐C) a train of volleys in high‐threshold muscle afferents evoked the characteristic long‐latency EPSP (B), which was effectively inhibited (C) by a preceding (cond) train of volleys in contralateral high‐threshold muscle afferents (contralateral hamstring nerve, co.H). In the extensor motoneuron (G‐I) long‐latency EPSP was evoked from contralateral high‐threshold muscle afferents (H) and inhibited from the ipsilateral posterior nerve to the knee joint (joint, I). D‐F and JL are corresponding records from interneurons that were located in a region dorsal to motor nuclei. Neuron in DF was excited from high‐threshold afferents in G‐S (E) and inhibited from contralateral high‐threshold cutaneous afferents (co. sural nerve, co. Sur; F). Neuron of JL was excited from co.H (K) and completely inhibited from high‐threshold afferents in the i.G‐S (L). M: tentative diagram showing principle organization of an interneuronal network that could account for results illustrated in A‐L, N,O. Inhibition could be exerted by pre‐ or postsynaptic actions. A single interneuron in the diagram represents a chain of neurons. N,O: alternating discharges in flexor and extensor efferents triggered by short trains of impulses in ipsilateral and contralateral FRA. Acute spinal cat pretreated with nialamide (10 mg/kg) before administration of 100 mg/kg dopa. Records from nerves to a flexor (medial sartorius, Sart) and to an extensor (medial vastus of quadriceps, Vast). N: stimulation of high‐threshold afferents in contralateral quadriceps nerve (co.Q; timing of stimulus train marked by arrow and the vertical interrupted line) during a spontaneous discharge in flexor efferents is followed by a pause in flexor nerve and a discharge in extensor nerve. Activity in flexor nerve is resumed after cessation of extensor discharge, and a 2nd extensor discharge then occurs after this flexor burst. O: stimulation of ipsilateral saphenous nerve (i.saph) evokes a long‐latency discharge in the flexor nerve that, after additional stimulation of co.Q, is followed by a series of alternating bursts in the efferent nerves. P: original hypothesis of organization of spinal “primary half‐centers” projecting to flexor and extensor muscles with reciprocal inhibition acting between them As described in text, Graham Brown 214 later proposed the existence of “interposed half‐centers” to describe interneurons (intercalated between primary afferents and motoneurons) with mutual inhibition as in the diagram in M.

AC, GI, and N from Jankowska et al. 324; DF and JL from Jankowska et al. 325; M adapted from Jankowska et al. 324; O from Jankowska et al. 324 illustrated in Lundberg 413; P from Graham Brown 213.] Graham Brown 209,211


Figure 21.

Inhibitory action from short‐latency flexor‐reflex afferent (FRA) pathways on transmission in long‐latency FRA pathways. Upper traces in AD are intracellular records from a motoneuron innervating posterior biceps‐semitendinosus muscles (PBSt); in EH they are extracellular records from a lumbar interneuron located dorsal to the motor nuclei. Lower traces are records from dorsal root entry zone. Voltage calibration applies to intracellular records. Both neurons were recorded in unanesthetized acute spinal cats after injection of dopa (100 mg/kg). AD: recording from PBSt motoneuron shows EPSPs of long latency and duration following stimulation of ipsilateral FRA (ipsilateral nerve to anterior biceps‐semimembranosus muscle, iABSm; stimulation strength is 37 times threshold of lowest threshold afferent fibers). Bars below AD indicate duration of repetitive stimulation. Notice that prolongation of repetitive stimulation of FRA delays onset of EPSP. EH: stimulation of ipsilateral sural nerve (iSur) elicits long‐latency and long‐lasting activation of interneuron that is typical after dopa. Notice that response is delayed with prolongation of repetitive stimulation (thickening of base lines is due to stimulation artifacts). I: tentative diagram of neuronal connections that could explain illustrated effects. A single interneuron in the diagram represents a chain of interneurons. Inhibition of short‐latency pathway A (A' and A”) from descending noradrenergic systems (NA) releases transmission in long‐latency pathway B (records AH were obtained in this condition). The prolongation in onset of late discharge found in motoneurons and interneurons of B (the interneuron in EH should belong to this population) illustrates inhibitory interaction from short‐latency on long‐latency FRA pathways. The short‐latency pathway A is subdivided into A' and A” to accomodate the finding that the noradrenergic system may block the short‐latency reflex action onto motoneurons when there is still an inhibitory action on transmission through the long‐latency pathway B. Mn, motoneuron.

A‐D, I adapted from Jankowska et al. 324; EH adapted from Jankowska et al. 325


Figure 22.

Diagram showing alternative reflex pathways from flexor reflex afferents (FRA) with descending (desc) excitatory connections to interneurons of these pathways and its inhibitory interactive connections with the other reflex pathways from the FRA. Mn, motoneuron.

Adapted from Lundberg 411


Figure 23.

Definition of propriospinal relay neurons in C3‐C4 segments. Experimental arrangement and neuronal circuits as illustrated in G (for records A‐F) and in N (for records H‐M). Upper traces, intracellular records from 2 biceps (Bi) motoneurons (A‐F; H‐M); lower traces, records from C6 dorsal root entry zone. Voltage calibration refers to intracellular records. Stimulation strength of contralateral pyramid (Pyr) is given in μA. Dots below surface records in B, C, I, J indicate arrival of 3rd pyramidal volley. Dashed lines below records B, E, I mark sections that are shown at an expanded time scale in C, F, J. AF: disynaptic (1.5 ms) Pyr EPSP before (A‐C) and after (D‐F) transection of corticospinal tract (CST) in C5. Comparison of antidromic spike potentials in A and D suggests that recording conditions before and after lesion were similar. Results indicate presence of propriospinal neurons (P) above the C5 lesion that transmit disynaptic corticospinal excitation to forelimb motoneurons (FMn). HM: near disappearance of disynaptic Pyr EPSP after CST transection at rostral (r) C3. Grading of the Pyr stimulation strength in L and M suggests that the very small disynaptic EPSP that remained after lesion was a unitary response. These results indicate that the most rostral location of the propriospinal neurons intercalated in disynaptic corticomotoneuronal pathway is in caudal C2 or rostral C3.

Adapted from Illert et al. 314


Figure 24.

Convergence on C3‐C4 propriospinal neurons (P) from corticospinal (CST) and rubrospinal (RST) tracts and from forelimb cutaneous afferents (superficial radial nerve, SR). Experimental arrangement and neuronal circuits as illustrated in J. Convergence was judged indirectly from intracellular recording from a forelimb motoneuron (FMn; upper traces in A‐H) with CST transected completely in C5 (compare in I records from lateral funiculus in C6 before, upper trace, and after, lower trace, the C5 lesion). The same lesion transected the RST only partially. Lower traces in AH are records from C6 dorsal root entry zone. Voltage calibration applies to intracellular records. Stimulation strengths: SR, 1.5 times threshold of lowest threshold fibers; red nucleus (NR), 65 μA; pyramid (Pyr), 200 μA. Note that disynaptic EPSP appears only in A, B when 2 rubral stimuli and 1 pyramidal stimulus are given together with the cutaneous SR volley that indicates excitation of common propriospinal neurons (P). All other possible combinations were ineffective (C‐H). Dashed line below A indicates the part of the traces that is expanded in B.

Adapted from Illert et al. 314


Figure 25.

Convergence on C3‐C4 propriospinal neurons (P). Scheme in H summarizes monosynaptic excitatory convergence onto them as well as their direct projection to forelimb motoneurons (FMn). Intracellular records in A‐G (upper traces) illustrate a propriospinal neuron (identified from the lateral funiculus in C5 in A) with monosynaptic excitation from forelimb group I muscle afferents (D‐G: deep radial nerve, DR; stimulation strength in times threshold of the lowest threshold fibers), from corticospinal tract (B: stimulation of contralateral pyramid, Pyr, with 100 μA), and from rubrospinal tract (C: stimulation of contralateral red nucleus, NR, with 100 μA). Lower traces are records from dorsal root entry zone in C7 (in G also from C3). Voltage calibration refers to intracellular records.

AG adapted from Illert et al. 311; H from data in refs. 54,226,310,311,314


Figure 26.

Collateral projection of C3‐C4 propriospinal neurons (P) to forelimb segments and to lateral reticular nucleus (LRN). Scheme in G illustrates experimental arrangement and neuronal circuits. A‐F: upper traces, intracellular records from a propriospinal neuron in C3 (antidromic activation from the C7 segment in E); lower traces, records from C3 dorsal root entry zone. Voltage calibration applies to intracellular records. A, B show antidromic activation of propriospinal neuron from LRN (stimulation strength in μA); note monosynaptic IPSP from LRN in A (see Fig. 39). C illustrates monosynaptic EPSP from contralateral pyramid (Pyr, 100 μA). DF: collision of antidromic spike (F) evoked in propriospinal neuron from C7 (E) by preceding LRN stimulation (D).

AF adapted from Illert and Lundberg 309; G, data for descending projection from Illert et al. 311,314; data for direct projection to forelimb motoneurons (FMn) from Illert et al. 314 and Grant et al. 226


Figure 27.

Segmental afferent input depolarizing terminals of muscle spindle Ia‐afferents (A), Golgi tendon organ Ib‐afferents (B), and cutaneous afferents (C). Approximate relative amount of depolarization contributed by each input has been estimated from results quoted in text and is indicated by width of arrows. Mn, motoneurons; Ia, Ib, II, and III are the respective muscle afferents; Cut, myelinated cutaneous fibers. A: ipsilateral inputs depolarizing Ia‐fibers of flexor (left) and extensor (right) muscles. B: ipsi‐ and contralateral inputs depolarizing Ib‐fibers. C: ipsi‐ and contralateral inputs depolarizing cutaneous afferent fibers. Note that this pattern of effects is partly changed after administration of dopa to acute spinal cat (described in the text).

From Schmidt 518


Figure 28.

Inhibitory action from flexor reflex afferents (FRA) on transmission to Ia‐afferents illustrated by recording Ia EPSPs in motoneurons (A‐D), by the antidromic discharge of Ia‐afferents following stimulation of their terminals in the spinal cord (E‐H), and by recording dorsal root potentials (I‐L). A‐D: upper traces, intracellular records from a gastrocnemius‐soleus motoneuron (G‐S); lower traces, from L7 dorsal root entry zone. Voltage calibration refers to intracellular records. Stimulus strengths are given in multiples of threshold for lowest threshold fibers. Homonymous test Ia EPSP is shown in A. In B there is no effect by a conditioning volley in the sural (Sur) nerve. Depression in C is evoked by a short train of maximal Ia volleys in nerve from posterior biceps‐semitendinosus (PBSt). With combined conditioning with sural and PBSt nerves there is a removal of the depression (D). E‐H: intraspinal excitability measurements of Ia‐afferent terminals in gastrocnemius‐soleus (G‐S) motor nucleus. Test response (E) was recorded in G‐S nerve. Conditioning stimulation of superficial peroneal nerve (SP; cutaneous afferents only) does not change excitability (F). The facilitation (G, reflecting an excitability increase) is evoked by a train of PBSt Ia volleys. Record in H shows that a volley in SP can remove facilitatory action from PBSt Ia volleys. I‐L: upper traces are dorsal root potentials (DRPs) recorded from the most caudal dorsal rootlet in L6. Lower traces are recorded from L7 dorsal root entry zone. I shows test DRP evoked from a train of maximal Ia volleys from PBSt. J shows response to stimulation of the sural nerve. In K and L there is combined stimulation of sural and PBSt nerves (cond. + test). Note pronounced depression of DRP from PBSt Ia‐afferents (K) as well as the long duration of this effect (L). M: tentative diagram showing connections through which volleys in FRA depress transmission from Ia to Ia‐terminals. Terminals with 2 branches are excitatory. Circles indicate presynaptic terminals making synaptic contacts with presynaptic terminals. In the diagram a single interneuron may represent a chain of interneurons.

AL from Lund et al. 402; diagram in M from Lundberg 406


Figure 29.

Long‐latency depolarization evoked from flexor reflex afferents (FRA) in Ia‐terminals after dopa administration. Unanesthetized decorticate acute spinal cat. Graphs in AC give excitability measurements of Ia‐afferent terminals from gastrocnemius‐soleus nerve (G‐S) before (A) and after (B, C) injection of dopa. Intraspinal stimulation of G‐S afferent terminals with a microelectrode (test) was conditioned (cond) with a short train of stimuli to anterior biceps–semimembranosus (ABSm) nerve (stimulation strength is indicated in times threshold, XT, of lowest threshold afferent fibers). Abscissa gives interval between 1st conditioning volley and test stimulus to G‐S terminals. Ordinate gives amplitude of conditioned response in percent of unconditioned test response. Comparison of A and C shows that after dopa FRA stimulation causes an increased excitability of Ia‐fibers of long latency (conditioning group I stimulation is without effect, B). DG illustrate simultaneous recordings of dorsal root potentials (DRPs) evoked from ABSm nerve. Comparison of E and G shows that parallel to excitability increase in Ia‐terminals following stimulation of FRA there is also a long‐latency DRP. H: tentative diagram of connections from FRA to Ia‐afferents. Filled circles, connections exerting either pre‐ or postsynaptic inhibition; open circles, termination on presynaptic terminals. Excitatory synaptic terminals are indicated by two branches. Each symbol represents a chain of interneurons. It is assumed that activity in short‐latency FRA pathway (left) inhibits long‐latency pathway to Ia‐afferent terminals. Inhibition from a descending noradrenergic (NA) pathway of short‐latency FRA system releases transmission in long‐latency path from FRA to Ia‐terminals.

AG adapted from Andén et al. 16; H from Lundberg 406


Figure 30.

Rubrospinal facilitation of segmental transmission to primary afferents. A–G: upper traces, dorsal root potentials (DRPs) recorded from the most caudal filament of the L6 dorsal root; lower traces, records from cord dorsum at L7. Voltage calibration beside F applies to DRPs of A–F. A–F: facilitation from red nucleus (NR) of DRPs evoked from cutaneous (B, C: sural nerve, Sur, stimulation at 3 times threshold of lowest threshold afferent fibers) and joint afferents (E, F: high‐threshold afferents in posterior knee joint nerve). Increased amplitude of DRP in C is due to facilitation of 2nd component of cutaneous DRP that is mediated by an FRA pathway. G, H: origin of DRPs evoked by tegmental stimulation. The DRPs of G were evoked with 100‐μA stimulation at sites marked in H in a drawing of a correspnding transverse section of brain stem. The DRPs are evoked from a rather restricted area within the NR. CA, central aqueduct; NIII, oculomotor nucleus; L, lateral and H, horizontal Horsley‐Clarke coordinates.

Adapted from Hongo et al. 276


Figure 31.

Depression by dorsal reticulospinal system of flexor reflex afferent (FRA) inhibition in gastrocnemius‐soleus (G‐S) motor nucleus. A: schematic drawing of experimental arrangements. Right, 3 transverse sections: from brain stem (about 6 mm rostral of obex) showing the stimulation; from lower thoracic cord, hatched area indicating transection of whole cord except the right dorsolateral fascicle; from lumbar cord with microelectrode recording in left ventral horn. Left ventral quadrant and dorsolateral fascicle are mounted on electrodes for recording ascending discharges. Ventral root (VR) discharges are recorded from S1 and L7 segments, and dorsal root potentials are recorded in a caudal filament of the L6 dorsal root (DR fil.), both on left. B–E: upper traces, monosynaptic reflexes recorded from ventral root; lower traces, incoming volleys at the dorsal root entry zone. Recording was done with 2 sweep speeds simultaneously and records therefore consist of double sets (left traces at slow and right traces at fast sweep speeds; see calibrations below E). B: monosynaptic reflex from group Ia‐afferents in G‐S nerve. C: same reflex inhibited by a preceding volley in high‐threshold afferents (20 times threshold for lowest threshold fibers) from anterior biceps and semimembranosus (ABSm). D, E: same as B and C respectively but conditioned by a train of 5 stimuli in brain stem that removes most of FRA inhibition (compare C and E), without any effect on monosynaptic test reflex itself (compare B and D). F: diagram of a transverse brain stem section about 6 mm above obex. Filled circles of different size indicate points of stimulation and magnitude of effect on FRA inhibitory actions; X indicates points stimulated without effect on transmission from FRA. G, H: dorsal root records (upper traces) showing that no dorsal root potential (DRP) is evoked by the same brain stem (BS) stimulation that was used in D, E, whereas there is a large DRP with the same amplification from a single volley in the sural nerve (Sur) at 20 times threshold for the lowest threshold fibers. I: time course of brain stem action illustrated in C and E. Inhibition of monosynaptic reflex by the ABSm volley (100% refers to inhibition seen without preceding brain stem stimulation) is plotted against time interval between onset of brain stem stimulation and arrival of ABSm volley at the cord.

Adapted from Engberg et al. 167


Figure 32.

Action of dorsal reticulospinal system in lumbar interneurons. Upper traces in AJ are records from 3 different interneurons (A–C; D–F; G–J) located at 1.8 mm–2.0 mm depth from cord dorsum. Lower traces are records from dorsal root entry zone (not present in C, F, I). Voltage calibrations refer to intracellular records. Stimulation strength for nerves is indicated in times threshold of lowest threshold afferent fibers. Experimental arrangement is shown in Fig. 31 A. A–F: excitation from high‐threshold fibers of posterior knee joint nerve (A: joint) and inhibition from high‐threshold muscle afferents of gastrocnemius‐soleus nerve (D; G‐S) were both depressed by conditioning stimulation of brain stem (B, E; BS). BS stimulation alone evoked neither postsynaptic potentials in interneurons (C, F) nor dorsal root potentials (not illustrated), which indicates inhibition from BS in FRA pathways to interneurons from which recording was done. GJ: interneuron with synaptic inhibition from BS. G: monosynaptic activation from superficial peroneal nerve (SP; stimulation at 6 times threshold for lowest threshold fibers). H: inhibitory postsynaptic potential (IPSP) evoked by 3 stimuli in BS. I: onset of IPSP evoked by a single shock (beginning of 1st deflection is marked by arrow). J: IPSP evoked by a single stimulus of contralateral dorsolateral funiculus (DLF) just rostral to spinal cord lesion. A comparison of the latencies to onset of inhibition in I and J suggests that responsible descending fibers have a conduction velocity of 25 m/s.

Adapted from Engberg et al. 168


Figure 33.

Effect of dopa on transmission from high‐threshold muscle afferents in motoneurons and reversal of effects by the α‐receptor blocker phenoxybenzamine. Graphs show effect of single conditioning volleys on monosynaptic test reflex (MSR) from flexor posterior biceps‐semitendinosus nerve (PBSt). Ordinate gives amplitude of MSR in percent of unconditioned MSR. Abscissa gives interval between arrival at spinal cord of conditioning and testing group I volleys. Strength of conditioning stimulation of plantaris nerve (PI) is indicated in times threshold (xT) of lowest threshold afferent fibers. A was evoked before injection of dopa; B evoked 10 min after injection of dopa; and C evoked 25 min after dopa and 15 min after injection of phenoxybenzamine. The FRA facilitation of the flexor MSR (A) is abolished after dopa (B), which indicates inhibition of transmission in short‐latency flexor reflex afferent (FRA) pathways. Reversal of inhibition by phenoxybenzamine indicates that effect of injected dopa is mediated by α‐receptors.

Adapted from Andén et al. 15


Figure 34.

Release of Ib‐ and flexor reflex afferent (FRA) actions after bilateral lesions of dorsolateral funiculi. Graphs show effect of single conditioning volleys in nerve to flexor digitorum longus (FDL) on monosynaptic test reflexes evoked from gastrocnemius‐soleus (G‐S). Ordinate gives amplitude of reflexes in percent of unconditioned test reflex. Abscissa gives interval between arrival at spinal cord of conditioning and testing group I volleys. Strength of conditioning stimulation (cond) is indicated in times threshold (xT) of lowest threshold afferent fibers. It was chosen to activate group I fibers in A and D, groups I and II fibers in B, and groups I, II, and III fibers in C and E. At each conditioning strength the curves were obtained after the ipsilateral (i) and contralateral (co) lesions indicated schematically in A (for AC) and in D (for D, E). Transverse spinal cord drawings in A show symbols used in AC to illustrate effects of sequential lesions; the initial lesion of the dorsal column and dorsolateral on the contralateral side, and ventral funiculi (O), the additional transection of ipsilateral ventral funiculus (x, cf. histological reconstruction to left), and the complete spinal transection (•). Drawings in D similarly show initial lesion of dorsal column and dorsolateral on the ipsilateral side, and ventral quadrants (O), the additional lesion of the contralateral dorsal funiculus (x, cf. histological reconstruction to left), and complete spinal transection (•).

Adapted from Holmqvist and Lundberg 267


Figure 35.

Differential release of transmission in excitatory and inhibitory flexor reflex afferent pathways by brain stem lesions. Graphs show effect of single conditioning volleys in nerve to flexor digitorum longus (FDL) on monosynaptic test reflexes evoked from gastrocnemius‐soleus (G‐S, AC) or from posterior biceps‐semitendinosus (PBSt, D, E). Ordinate gives amplitude of reflexes in percent of unconditioned test reflexes. Abscissa gives interval between arrival at spinal cord of conditioning and testing group I volleys. Strength of conditioning stimulation (cond) is indicated in times threshold (xT) of lowest threshold afferent fibers. It was chosen to activate group I fibers in A and D, groups I and II fibers in B, and groups I, II, and III fibers in C and E. At each conditoning strength the curves were obtained after brain stem lesions indicated in F.

Adapted from Holmqvist and Lundberg 268


Figure 36.

Monosynaptic connections to α‐motoneurons from corticospinal and rubrospinal tracts in primate (rhesus monkey). Lower traces in AF are intracellular records from 2 motoneurons. Upper traces are cord dorsum potentials monitoring afferent incoming volleys and corticospinal volleys. All traces in G, H are intracellular records. AF: frequency potentiation of monosynaptic excitatory postsynaptic potential (EPSP) evoked from motor cortex (MC) in a motoneuron to flexor digitorum longus (FDL, in AC) and to gastrocnemius‐soleus (G‐S, in DF). Antidromic invasion from FDL and G‐S nerves in A and D, respectively. Monosynaptic EPSPs following a single shock stimulation of motor cortex are shown in B, E. Growth of monosynaptic EPSPs with repetitive stimuli shown in C, F; same stimulus strengths as in B, E, respectively. Notice the greater potentiation in F than in C. Calibration pulse 1 ms, 5 m V in A, D; 1 ms, 2 m V in B, C, E, F. G, H: convergence of monosynaptic EPSPs from motor cortex (MC) and red nucleus (NR) in an FDL motoneuron. Superimposed traces of monosynaptic EPSPs evoked by stimulation of MC (G) and NR (H) of different intensities given to right of each trace. Note differences in amplitude and time course from MC and NR.

AF from Tamarova et al. 574; G, H from Shapovalov 539


Figure 37.

Diagram illustrating 2 different ways in which ascending neurons may monitor interneuronal activity. A: by ascending collaterals from interneurons interposed in segmental reflex pathways. B: by separate ascending neurons that receive collateral projections from interneurons interposed in segmental reflex pathways. Mn, motoneuron; asc., ascending neuron; prim. aff., primary afferent.



Figure 38.

Decerebrate control of transmission to ascending ventral pathways activated from flexor reflex afferents (FRA). Diagram in J illustrates experimental arrangement with a bilateral section of ventral quadrants (VQ) and recording from right VQ of ascending discharge evoked from highthreshold afferents in left hamstring nerve (1. Ham; A, D, G) and left sural nerve (1. Sur; B, E, H). C, F, I show background activity in VQ without stimulation. Comparison of records obtained before (AC), during (DF), and after (GI) cooling of dorsal part of spinal cord (the dorsal column removed) with a thermode illustrates effective blockade of transmission from FRA to ascending tracts in decerebrate state with conducting dorsolateral funiculi (A–C; G–I), and release in the spinal state, i.e., during blockade of impulse transmission in dorsal part of the spinal cord (DF).

Adapted from Holmqvist et al. 269


Figure 39.

Wiring diagrams illustrating 2 examples of ascending information from collaterals of neurons interposed in spinal neuronal circuits. A: ascending collaterals from propriospinal neurons in C3–C4 (P) to lateral reticular nucleus (LRN). Mn, motoneuron. This connection is illustrated in Fig. 26. B: ascending collaterals to LRN from inhibitory interneurons (with monosynaptic excitation from forelimb afferents) projecting to propriospinal neurons in C3‐C4.

A from data of Illert and Lundberg 309; B from unpublished results of B. Alstermark, M. Illert, and A. Lundberg (see ref. 413


Figure 40.

Wiring diagrams illustrating hypothesis that information transferred by ventral spinocerebellar tract (VSCT) neurons reflects relation between input to and output from inhibitory segmental interneurons. A: diagram illustrating convergence on a VSCT neuron of monosynaptic excitation from primary afferents (prim. aff.) and disynaptic inhibition evoked from the same primary afferents. Disynaptic inhibition of VSCT neuron is due to a collateral projection from interneurons transmitting disynaptic inhibitory postsynaptic potentials to motoneurons (Mn). In this case the VSCT neuron would compare excitatory input to interneurons from primary afferents and the resulting activity of the inhibitory interneuron. B: in addition to connections in A there is also a descending motor pathway (desc) giving monosynaptic excitation and disynaptic inhibition of VSCT neuron. The descending monosynaptic excitation is collateral to excitation of interneurons, and the disynaptic inhibition is again due to collateral projection from interneuron. Such a VSCT neuron thus compares excitatory input to interneurons from primary afferents and the descending pathway with the resulting activation of the interneuron.



Figure 41.

Recurrent depression of a Ia inhibitory postsynaptic potential (IPSP) in a ventral spinocerebellar tract (VSCT) neuron. Experimental arrangement and relevant neuronal circuits are illustrated in A. Mn, motoneuron; RC, Renshaw cell; Q, quadriceps. Upper traces in BE are intracellular records from a VSCT neuron (identification from ipsilateral anterior lobe of cerebellum in E). Lower traces are records from L5 dorsal root entry zone. Voltage calibrations refer to intracellular records. BD: disynaptic Ia IPSPs from Q nerve. Stimulation strength is indicated in times threshold of lowest threshold afferent fibers. F: averaged records of submaximal Ia IPSPs from Q (upper trace) combined with conditioning stimulation of L5–S1 ventral roots (VRs; lower trace). Arrow below lower trace indicates arrival of the VR volley to the spinal cord (calibration pulse 1 mV, 2 ms).

Adapted from Gustafsson and Lindström 251


Figure 42.

Facilitation of Ia inhibitory postsynaptic potentials (IPSPs) in ventral spinocerebellar tract (VSCT) neurons from vestibulospinal and rubrospinal tracts. A shows experimental arrangement and relevant neuronal circuits. Mn, motoneurons; ND, Deiters' nucleus; NR, red nucleus. Upper traces in BD and FH are intracellular records from 2 different VSCT neurons. Lower traces are from dorsal root entry zone. Voltage calibrations refer to intracellular records. Strength of stimulation in ND and NR is indicated in μA, stimulation of nerves in times threshold of lowest threshold afferent fibers. BE: facilitation of a quadriceps‐evoked (Q) Ia IPSP from the ND. Graph in E gives amplitude of IPSP in percent of unconditioned test (ordinate) in relation to time interval between arrival of descending volley from ND (zero ms on abscissa) and incoming Q volley (interrupted line indicates synchronous arrival). Facilitation started with Q volley arriving slightly before ND volley, which indicates monosynaptic coupling of descending fibers with interneuron mediating the Ia IPSP. FI: facilitation from NR of IPSPs evoked from posterior biceps‐semitendinosus nerve (PBSt). Records in FH show a facilitation of a Ia IPSP from PBSt. Graph in I shows time course of facilitation of a Ia IPSP from Q in another VSCT neuron. Amplitude of IPSP is given in percent of unconditioned test (ordinate) in relation to time interval in arrival of 1st rubrospinal volley (zero ms on abscissa) and the incoming Q volley (interrupted line indicates synchronous arrival).

BE adapted from Baldissera and Roberts 40; FI adapted from Baldissera and Bruggencate 34
References
 1. Abrahams, V. C. Cervico‐lumbar reflex interactions involving a proprioceptive receiving area of the cerebral cortex. J. Physiol. London 209: 45–56, 1970.
 2. Abrahams, V. C., and S. Falchetto. Hind leg ataxia of cervical origin and cervico‐lumbar spinal interactions with a supratentorial pathway. J. Physiol. London 203: 435–447, 1969.
 3. Adam, D., U. Windhorst, and G. F. Inbar. The effects of recurrent inhibition on the cross‐correlated firing patterns of motoneurones (and their relation to signal transmission in the spinal cord‐muscle channel). Biol. Cybernetics 29: 229–235, 1978.
 4. Ahlman, H., S. Grillner, and M. Udo. The effect of 5‐HTP on the static fusimotor activity and the tonic stretch reflex of an extensor muscle. Brain Res. 27: 393–396, 1971.
 5. Alnaes, E., J. K. S. Jansen, and T. Rudjord. Fusimotor activity in the spinal cat. Acta Physiol. Scand. 63: 197–212, 1965.
 6. Alstermark, B., A. Lundberg, U. Norrsell, and E. Sybirska. Role of C3‐C4 propriospinal neurones in forelimb movements in the cat. Acta Physiol. Scand. 105: 24A–25A, 1979.
 7. Alvord, E. C. Jr., and M. G. F. Fuortes. Reflex activity of extensor motor units following muscular afferent excitation. J. Physiol. London 122: 302–321, 1953.
 8. Amassian, V. E., H. Weiner, and M. Rosenblum. Neural systems subserving the tactile placing reaction: a model for the study of higher level control of movement. Brain Res. 40: 171–178, 1972.
 9. Anastasievic, R., D. A. Vasilenko, A. I. Kostyukov, and N. N. Preobrazhenskii. Reticulofugal influences of interneurones in the lateral region of gray matter of the cat spinal cord. Neurophysiology USSR 5: 525–536, 1973.
 10. Andén, N.‐E., H. Corrodi, K. Fuxe, B. Hökfelt, T. Hökfelt, C. Rydin, and T. Svensson. Evidence for a central noradrenaline receptor stimulation by clonidine. Life Sci. 9: 513–523, 1970.
 11. Andén, N.‐E., J. Engel, and A. Rubenson. Mode of action of l‐DOPA on central noradrenaline mechanisms. Naunyn‐Schmiedebergs Arch. Pharmacol. 273: 1–10, 1972.
 12. Andén, N.‐E., M. G. M. Jukes, and A. Lundberg. Spinal reflexes and monoamine liberation. Nature London 202: 1222–1223, 1964.
 13. Andén, N.‐E., M. G. M. Jukes, and A. Lundberg. The effect of DOPA on the spinal cord. 2. A pharmacological analysis. Acta Physiol. Scand. 67: 387–397, 1966.
 14. Andén, N.‐E., M. G. M. Jukes, A. Lundberg, and L. Vyklický. A new spinal flexor reflex. Nature London 202: 1344–1345, 1964.
 15. Andén, N.‐E., M. G. M. Jukes, A. Lundberg, and L. Vyklický. The effect of DOPA on the spinal cord. 1. Influence on transmission from primary afferents. Acta Physiol. Scand. 67: 373–386, 1966.
 16. Andén, N.‐E., M. G. M. Jukes, A. Lundberg, and L. Vyklický. The effect of DOPA on the spinal cord. 3. Depolarization evoked in the central terminals of ipsilateral Ia afferents by volleys in the flexor reflex afferents. Acta Physiol. Scand. 68: 322–336, 1966.
 17. Andersen, P., J. C. Eccles, R. F. Schmidt, and T. Yokota. Slow potential waves produced in the cuneate nucleus by cutaneous volleys and by cortical stimulation. J. Neurophysiol. 27: 78–91, 1964.
 18. Andersen, P., J. C. Eccles, R. F. Schmidt, and T. Yokota. Depolarization of presynaptic fibers in the cuneate nucleus. J. Neurophysiol. 27: 92–106, 1964.
 19. Andersen, P., J. C. Eccles, and T. A. Sears. Cortically evoked depolarization of primary afferent fibers in the spinal cord. J. Neurophysiol. 27: 63–77, 1964.
 20. Anderson, E. G. Bulbospinal serotonin‐containing neurons and motor control. Federation Proc. 31: 107–112, 1972.
 21. Anderson, G., and B. Sjölund. The ventral spino‐olivocerebellar system in the cat. IV. Spinal transmission after administration of clonidine and l‐Dopa. Exp. Brain Res. 33: 227–240, 1978.
 22. Andersson, O., H. Forssberg, S. Grillner, and M. Lindquist. Phasic gain control of the transmission in cutaneous reflex pathways to motoneurones during “Active” locomotion. Brain Res. 149: 503–507, 1978.
 23. Andrew, B. L., and E. Dodt. The deployment of sensory nerve endings at the knee joint of the cat. Acta Physiol. Scand. 28: 287–296, 1953.
 24. Aoki, M., and A. K. McIntyre. Cortical and long spinal actions on lumbosacral motoneurones in the cat. J. Physiol. London 251: 569–587, 1975.
 25. Appelberg, B. A rubro‐olivary pathway. II. Simultaneous action on dynamic fusimotor neurones and the activity of the posterior lobe of the cerebellar cortex. Exp. Brain Res. 3: 382–390, 1967.
 26. Appelberg, B., and T. Jeneskog. Mesencephalic fusimotor control. Exp. Brain Res. 15: 97–112, 1972.
 27. Appelberg, B., T. Jeneskog, and H. Johansson. Rubrospinal control of static and dynamic fusimotor neurones. Acta Physiol. Scand. 95: 431–440, 1975.
 28. Appelberg, B., H. Johansson, and G. Kalistratov. The influence of group II muscle afferents and low threshold skin afferents on dynamic fusimotor neurones to the triceps surae of the cat. Brain Res. 132: 153–158, 1977.
 29. Appelberg, B., and C. Molander. A rubro‐olivary pathway. I. Identification of a descending system for control of the dynamic sensitivity of muscle spindles. Exp. Brain Res. 3: 372–381, 1967.
 30. Araki, T., J. C. Eccles, and M. Ito. Correlation of the inhibitory postsynaptic potential of motoneurones with the latency and time course of inhibition of monosynaptic reflexes. J. Physiol. London 154: 354–377, 1960.
 31. Arshavsky, Yu., M. B. Berkinblit, O. I. Fukson, I. M. Gelfand, and G. N. Orlovsky. Origin of modulation in neurones of the ventral spinocerebellar tract during locomotion. Brain Res. 43: 276–279, 1972.
 32. Asanuma, H., and I. Rosén. Spread of mono‐ and polysynaptic connections within cat's motor cortex. Exp. Brain Res. 16: 507–520, 1973.
 33. Baldissera, F., and G. Broggi. An analysis of potential changes in the spinal cord during desynchronized sleep. Brain Res. 6: 706–715, 1967.
 34. Baldissera, F., and G. ten eBruggencate. Rubrospinal effects on ventral spinocerebellar tract neurones. Acta Physiol. Scand. 96: 233–249, 1976.
 35. Baldissera, F., G. ten eBruggencate, and A. Lundberg. Rubrospinal monosynaptic connexion with last‐order interneurones of polysynaptic reflex paths. Brain Res. 27: 390–392, 1971.
 36. Baldissera, F., M. G. Cesa‐Bianchi, and M. Mancia. Phasic events indicating presynaptic inhibition of primary afferents to the spinal cord during desynchronized sleep. J. Neurophysiol. 29: 871–887, 1966.
 37. Baldissera, F., A. Lundberg, and M. Udo. Activity evoked from the mesencephalic tegmentum in descending pathways other than the rubrospinal tract. Exp. Brain Res. 15: 133–150, 1972.
 38. Baldissera, F., A. Lundberg, and M. Udo. Stimulation of pre‐ and postsynaptic elements in the red nucleus. Exp. Brain Res. 15: 151–167, 1972.
 39. Baldissera, F., and W. J. Roberts. Effects on the ventral spinocerebellar tract neurones from Deiters' nucleus and the medial longitudinal fascicle in the cat. Acta Physiol. Scand. 93: 228–249, 1975.
 40. Baldissera, F., and W. J. Roberts. Effects from the vestibulospinal tract on transmission from primary afferents to ventral spino‐cerebellar tract neurones. Acta Physiol. Scand. 96: 217–232, 1976.
 41. Ballif, L., J. F. Fulton, and E. G. T. Liddell. Observations on spinal and decerebrate knee‐jerks with special reference to their inhibition by single break‐shocks. Proc. R. Soc. London Ser. B 98: 586–607, 1925.
 42. Bard, P. Studies on the cerebral cortex. 1. Localized control of placing and hopping reactions in the cat and their normal management by small cortical remnants. Arch. Neurol. Psychiatry 30: 40–74, 1933.
 43. Barker, D. The morphology of muscle receptors. In: Handbook of Sensory Physiology. Muscle Receptors, edited by C. C. Hunt. Berlin: Springer‐Verlag, 1974, vol. 3, pt. 2, p. 1–190.
 44. Barnes, C. D., and O. Pompeiano. Inhibition of monosynaptic extensor reflex attributable to presynaptic depolarization of the group Ia afferent fibers produced by vibration of flexor muscle. Arch. Ital. Biol. 108: 233–258, 1970.
 45. Barnes, C. D., and O. Pompeiano. Presynaptic and postsynaptic effects in the monosynaptic reflex pathway to extensor motoneurons following vibration of synergic muscles. Arch. Ital. Biol. 108: 259–294, 1970.
 46. Barron, D. H., and B. H. C. Matthews. The interpretation of potential changes in the spinal cord. J. Physiol. London 92: 276–321, 1938.
 47. Baxendale, R. H., and J. R. Rosenberg. Crossed reflexes evoked by selective activation of tendon organ afferent axons in the decerebrate cat. Brain Res. 127: 323–326, 1977.
 48. Bazett, H. C., and W. G. Penfield. A study of the Sherrington decerebrate animal in the chronic as well as the acute condition. Original articles and clinical cases. Brain 65: 185–265, 1922.
 49. Benecke, R., J. Meyer‐Lohmann, and J. Guntau. Inverse changes in the excitability of Renshaw cells and α‐motoneurones induced by interpositus stimulation. Pfluegers Arch. 365: R40, 1976.
 50. Bergmans, J., R. Burke, L. Fedina, and A. Lundberg. The effect of DOPA on the spinal cord. 8. Presynaptic and “remote” inhibition of transmission from Ia afferents to alpha motoneurones. Acta Physiol. Scand. 90: 618–639, 1974.
 51. Bergmans, J., R. Burke, and A. Lundberg. Inhibition of transmission in the recurrent inhibitory pathway to motoneurones. Brain Res. 13: 600–602, 1969.
 52. Bergmans, J., and S. Grillner. Changes in dynamic sensitivity of primary endings of muscle spindle afferents induced by DOPA. Acta Physiol. Scand. 74: 629–636, 1968.
 53. Bergmans, J., and S. Grillner. Reciprocal control of spontaneous activity and reflex effects in static and dynamic flexor γ‐motoneurones revealed by an injection of DOPA. Acta Physiol. Scand. 77: 106–124, 1969.
 54. Bergmans, J., M. Illert, E. Jankowska, and A. Lundberg. Reticulospinal control of propriospinal neurones mediating disynaptic corticomotoneurones excitation in the cat (Abstract). Acta Physiol. Scand. 96: 5A, 1976.
 55. Bergmans, J., S. Miller, and D. J. Reitsma. Influence of l‐DOPA on transmission in long ascending propriospinal pathways in the cat. Brain Res. 62: 155–167, 1973.
 56. Berkinblit, M. B., T. G. Deliagina, A. G. Feldman, I. M. Gelfand, and G. N. Orlovsky. Generation of scratching. I. Activity of spinal interneurons during scratching. J. Neurophysiol. 41: 1040–1057, 1978.
 57. Berkinblit, M. B., T. G. Deliagina, A. G. Feldman, I. M. Gelfand, and G. N. Orlovsky. Generation of scratching. II. Nonregular regimes of generation. J. Neurophysiol. 41: 1058–1069, 1978.
 58. Berkinblit, M. B., T. G. Deliagina, G. N. Orlovksy, and A. G. Feldman. Activity of propriospinal neurons during scratch reflex in the cat. Neirofiziologiia 9: 504–511, 1977.
 59. Bernhard, C. G., and E. Bohm. Monosynaptic corticospinal activation of fore limb motoneurones in monkeys (Macaca mulatta). Acta Physiol. Scand. 31: 104–112, 1954.
 60. Bernhard, C. G., E. Bohm, and I. Petersén. Investigations on the organization of the corticospinal system in monkeys (Macaca mulatta). Acta Physiol. Scand. 29: (Suppl. 106) 79–105, 1953.
 61. Bessou, P., F. Emonet‐Dénand, and Y. Laporte. Motor fibres innervating extrafusal and intrafusal muscle fibres in the cat. J. Physiol. London 180: 649–672, 1965.
 62. Björklund, A., and G. Skagerberg. Evidence for a major spinal cord projection from the diencephalic A11 dopamine cell group in the rat using transmitter‐specific fluorescent retrograde tracing. Brain Res. 177: 170–175, 1979.
 63. Blessing, W. W., and J. P. Chalmers. Direct projection of catecholamine (presumably dopamine)‐containing neurons from hypothalamus to spinal cord. Neurosci. Lett. 11: 35–40, 1979.
 64. Bok, S. T. Das Rückenmark. In: Handbuch der Mikroskopischen Anatomie des Menschen. Vierter Band: Nervensystem, edited by W. von Möllendorff. Berlin: Springer‐Verlag, 1928, p. 478–578.
 65. Bosemark, B. Some aspects on the crossed extensor reflex in relation to mononeurones supplying fast and slow contracting muscles. In: Muscular Afferents and Motor Control, Nobel Symposium I, edited by R. Granit. Stockholm: Almqvist & Wiksell, 1966, p. 261–268.
 66. Boyd, I. A., and M. R. Davey. Composition of Peripheral Nerves. Edinburgh: Livingstone, 1968.
 67. Bradley, K., D. M. Easton, and J. C. Eccles. An investigation of primary or direct inhibition. J. Physiol. London 122: 474–488, 1953.
 68. Bradley, K., and J. C. Eccles. Analysis of the fast afferent impulses from thigh muscles. J. Physiol. London 122: 462–473, 1953.
 69. Brooks, V. B., and V. J. Wilson. Recurrent inhibition in the cat's spinal cord. J. Physiol. London 146: 380–391, 1959.
 70. Brown, M. C., I. Engberg, and P. B. C. Matthews. The relative sensitivity to vibration of muscle receptors of the cat. J. Physiol. London 192: 773–800, 1967.
 71. Brown, M. C., D. G. Lawrence, and P. B. C. Matthews. Antidromic inhibition of presumed fusimotor neurones by repetitive stimulation of the ventral root in the decerebrate cat. Experientia 24: 1210–1211, 1968.
 72. Bruggencate, G. ten, and A. Lundberg. Facilitatory interaction in transmission to motoneurones from vestibulospinal fibres and contralateral primary afferents. Exp. Brain Res. 19: 248–270, 1974.
 73. Budakova, N. N. Stepping movements evoked by a rhythmic stimulation of a dorsal root in mesencephalic cat [in Russian] Fiziol. Zh. SSSR. im I. M. Sechenova 57: 1632–1640, 1971. [Engl. transl. Neurosci. Behav. Physiol. 5: 355–363, 1972.].
 74. Budakova, N. N. Stepping movements in the spinal cat due to DOPA administration [in Russian]. Fiziol. Zh. SSSR im I. M. Sechenova 59: 1190–1198, 1973.
 75. Burgess, P. R., and F. J. Clark. Characteristics of knee joint receptors in the cat. J. Physiol. London 203: 317–335, 1969.
 76. Burgess, P. R., and E. R. Perl. Cutaneous mechanoreceptors and nociceptors. In: Handbook of Sensory Physiology. Somatosensory System, edited by A. Iggo. Berlin: Springer‐Verlag, 1973, vol. II, p. 29–78.
 77. Burke, D., K.‐E. Hagbarth, and L. Löfstedt. Muscle spindle responses in man to changes in load during accurate position maintenance. J. Physiol. London 276: 159–164, 1978.
 78. Burke, D., K.‐E. Hagbarth, and N. F. Skuse. Voluntary activation of spindle endings in human muscles temporarily paralysed by nerve pressure. J. Physiol. London 287: 329–336, 1979.
 79. Burke, D., L. Knowles, C. Andrews, and P. Ashby. Spasticity, decerebrate rigidity and the clasp‐knife phenomenon: an experimental study in the cat. Brain 95: 31–48, 1972.
 80. Burke, R. E., E. Jankowska, and G. ten Bruggencate. A comparison of peripheral and rubrospinal synaptic input to slow and fast twitch motor units of triceps surae. J. Physiol. London 207: 709–732, 1970.
 81. Burke, R., A. Lundberg, and F. Weight. Spinal border cell origin of the ventral spinocerebellar tract. Exp. Brain Res. 12: 283–294, 1971.
 82. Burke, R. E., and P. Rudomin. Spinal neurons and synapses. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc., 1977, sect. 1, vol. I, pt. 2, chapt. 24, p. 877–944.
 83. Burke, R. E., P. Rudomin, L. Vyklický, and F. E. Zajac. III. Primary afferent depolarization and flexion reflexes produced by radiant heat stimulation of the skin. J. Physiol. London 213: 185–214, 1971.
 84. Burke, R. E., P. L. Strick, K. Kanda, C. C. Kim, and B. Walmsley. Anatomy of medial gastrocnemius and soleus motor nuclei in cat spinal cord. J. Neurophysiol. 40: 667–680, 1977.
 85. Burke, R. E., and P. Tsairis. Histochemical and physiological profile of a skeletofusimotor (β) unit in cat soleus muscle. Brain Res. 129: 341–345, 1977.
 86. Bussel, B., and E. Pierrot‐Deseilligny. Inhibition of human motoneurones, probably of Renshaw origin, elicited by an orthodromic motor discharge. J. Physiol. London 269: 319–339, 1977.
 87. Cajal, S. Ramón y. Histologie du Système Nerveux de L'Homme et des Vertébrés. Paris: Maloine, 1909, vol. 1.
 88. Cangiano, A., W. A. Cook, Jr., and O. Pompeiano. Primary afferent depolarization in the lumbar cord evoked from the fastigial nucleus. Arch. Ital. Biol. 107: 321–340, 1969.
 89. Cangiano, A., W. A. Cook, Jr., and O. Pompeiano. Cerebellar inhibitory control of the vestibular reflex pathways to primary afferents. Arch. Ital. Biol. 107: 341–364, 1969.
 90. Carli, G., K. Diete‐Spiff, and O. Pompeiano. Responses of the muscle spindles and of the extrafusal fibres in an extensor muscle to stimulation of the lateral vestibular nucleus in the cat. Arch. Ital. Biol. 105: 209–242, 1967.
 91. Carlsson, A., B. Falck, K. Fuxe, and N.‐Å. Hillarp. Cellular localization of monoamines in the spinal cord. Acta Physiol. Scand. 60: 112–119, 1964.
 92. Carlsson, A., T. Magnusson, and E. Rosengren. 5‐Hydroxytryptamine of the spinal cord normally and after transection. Experientia 19: 359, 1963.
 93. Carpenter, D., I. Engberg, H. Funkenstein, and A. Lundberg. Decerebrate control of reflexes to primary afferents. Acta Physiol. Scand. 59: 424–437, 1963.
 94. Carpenter, D., I. Engberg, and A. Lundberg. Differential supraspinal control of inhibitory and excitatory actions from the FRA to ascending spinal pathways. Acta Physiol. Scand. 63: 103–110, 1965.
 95. Carpenter, D., I. Engberg, and A. Lundberg. Primary afferent depolarization evoked from the brain stem and the cerebellum. Arch. Ital. Biol. 104: 73–85, 1966.
 96. Carpenter, D., A. Lundberg, and U. Norrsell. Primary afferent depolarization evoked from the sensorimotor cortex. Acta Physiol. Scand. 59: 126–142, 1963.
 97. Clark, F. J., and P. R. Burgess. Slowly adapting receptors in cat knee joint: can they signal joint angle? J. Neurophysiol. 38: 1448–1463, 1975.
 98. Clendenin, M., C.‐F. Ekerot, O. Oscarsson, and I. Rosén. The lateral reticular nucleus in the cat. II. Organization of component activated from bilateral ventral flexor reflex tract (bVFRT). Exp. Brain Res. 21: 487–500, 1974.
 99. Clough, J. F. M., D. Kernell, and C. G. Phillips. The distribution of monosynaptic excitation from the pyramidal tract and from primary spindle afferents to motoneurones of the baboon's hand and forearm. J. Physiol. London 198: 145–166, 1968.
 100. Clough, J. F. M., C. G. Phillips, and J. D. Sheridan. The short‐latency projection from the baboon's motor cortex to fusimotor neurones of the forearm and hand. J. Physiol. London 216: 257–279, 1971.
 101. Cohen, L. A. Localization of stretch reflex. J. Neurophysiol. 16: 272–285, 1953.
 102. Cohen, L. A. Organization of stretch reflex into two types of direct spinal arcs. J. Neurophysiol. 17: 443–453, 1954.
 103. Cook, W. A. Jr., A. Cangiano, and O. Pompeiano. Vestibular control of transmission in primary afferents to the lumbar spinal cord. Arch. Ital. Biol. 107: 296–320, 1969.
 104. Cooper, S., and C. S. Sherrington. Gower's tract and spinal border cells. Brain 63: 123–134, 1940.
 105. Coppin, C. M. L., J. J. B. Jack, and C. R. MacLennan. A method for the selective electrical activation of tendon organ afferent fibres from the cat soleus muscle. J. Physiol. London 210: 18P–20P, 1970.
 106. Coppin, C. M. L., J. J. B. Jack, and A. K. McIntyre. Properties of group I afferent fibres from semitendinosus muscle in the cat. J. Physiol. London 203: 45P–46P, 1969.
 107. Corda, M., C. von Euler, and G. Lennerstrand. Reflex and cerebellar influences on α and on “rhythmic” and “tonic” γ activity in the intercostal muscle. J. Physiol. London 184: 898–923, 1966.
 108. Crago, P. E., J. C. Houk, and Z. Hasan. Regulatory actions of human stretch reflex. J. Neurophysiol. 39: 925–935, 1976.
 109. Creed, R. S., D. Denny‐Brown, J. C. Eccles, E. G. T. Liddell, and C. S. Sherrington. Reflex Activity of the Spinal Cord. London: Oxford Univ. Press, 1932.
 110. Cullheim, S., and J.‐O. Kellerth. A morphological study of the axons and recurrent axon collaterals of cat α‐motoneurones supplying different hind‐limb muscles. J. Physiol. London 281: 285–299, 1978.
 111. Cullheim, S., and J.‐O. Kellerth. A morphological study of the axons and recurrent axon collaterals of cat α‐motoneurones supplying different functional types of muscle unit. J. Physiol. London 281: 301–313, 1978.
 112. Cullheim, S., J.‐O. Kellerth, and S. Conradi. Evidence for direct synaptic interconnections between cat spinal α‐motoneurons via the recurrent axon collaterals: a morphological study using intracellular injection of horseradish peroxidase. Brain Res. 132: 1–10, 1977.
 113. Curtis, D. R., J. W. Phillis, and J. C. Watkins. Cholinergic and non‐cholinergic transmission in the mammalian spinal cord. J. Physiol. London. 158: 296–323, 1961.
 114. Curtis, D. R., and R. W. Ryall. The synaptic excitation of Renshaw cells. Exp. Brain Res. 2: 81–96, 1966.
 115. Czarkowska, J., E. Jankowska, and E. Sybirska. Axonal projections of spinal interneurones excited by group I afferents in the cat, revealed by intracellular staining with horseradish peroxidase. Brain Res. 118: 115–118, 1976.
 116. Czarkowska, J., E. Jankowska, and E. Sybirska. Diameter and internodal length of axons of spinal interneurones excited by group I afferents in the cat. Brain Res. 118: 119–122, 1976.
 117. Czarkowska, J., E. Jankowska, and E. Sybirska. Common interneurones in reflex pathways from group Ia and Ib afferents of knee flexors and extensors. J. Physiol. London. In press.
 118. Dahlström, A., and K. Fuxe. Evidence for the existence of monoamine neurons in the central nervous system. II. Experimentally induced changes in the intraneuronal amine levels of bulbospinal neuron systems. Acta Physiol. Scand. 64: (Suppl. 247) 1–36, 1965.
 119. Decandia, M., L. Provini, and H. Táboříková. Presynaptic inhibition of the monosynaptic reflex following the stimulation of nerves to extensor muscles of the ankle. Exp. Brain Res. 4: 34–42, 1967.
 120. DeGail, P., J. W. Lance, and P. D. Neilson. Differential effects on tonic and phasic reflex mechanisms produced by vibration of muscles in man. J. Neurol. Neurosurg. Psychiatry 29: 1–11, 1966.
 121. Deliagina, T. G., A. G. Feldman, I. M. Gelfand, and G. N. Orlovsky. On the role of central program and afferent inflow in the control of scratching movements in the cat. Brain Res. 100: 297–313, 1975.
 122. Delwaide, P. J. Human monosynaptic reflexes and presynaptic inhibition. In: New Developments in Electromyography and Clinical Neurophysiology, edited by J. E. Desmedt. Basel: Karger, 1973, vol. 3, p. 508–522.
 123. Denny‐Brown, D. Motor mechanisms—introduction: the general principles of motor integration. In: Handbook of Physiology. Neurophysiology, edited by J. Field, H. W. Magoun and V. E. Hall. Washington, DC: Am. Physiol. Soc., 1960, sect. 1, vol. II, p. 781–796.
 124. Devanandan, M. S., R. M. Eccles, and T. Yokota. Depolarization of afferent terminals evoked by muscle stretch. J. Physiol. London 179: 417–429, 1965.
 125. Devanandan, M. S., R. M. Eccles, and T. Yokota. Muscle stretch and the presynaptic inhibition of the group Ia pathway to motoneurones. J. Physiol. London. 179: 430–441, 1965.
 126. Devanandan, M. S., B. Holmqvist, and T. Yokota. Presynaptic depolarization of group I muscle afferents by contralateral afferent volleys. Acta Physiol. Scand. 63: 46–54, 1965.
 127. Downman, C. B. B., and A. Hussain. Spinal tracts and supraspinal centres influencing visceromotor and allied reflexes in cats. J. Physiol. London 141: 489–499, 1958.
 128. Eccles, J. C. The Neurophysiological Basis of Mind: The Principles of Neurophysiology. Oxford: Clarendon, 1953.
 129. Eccles, J. C. Presynaptic inhibition in the spinal cord. In: Progress in Brain Research. Physiology of Spinal Neurons, edited by J. C. Eccles and J. P. Schadé. Amsterdam: Elsevier, 1964, vol. 12, p. 65–91.
 130. Eccles, J. C. The Inhibitory Pathways of the Central Nervous System. The Sherrington Lectures IX. Liverpool: Liverpool Univ. Press, 1969.
 131. Eccles, J. C., R. M. Eccles, A. Iggo, and M. Ito. Distribution of recurrent inhibition among motoneurones. J. Physiol. London 159: 479–499, 1961.
 132. Eccles, J. C., R. M. Eccles, A. Iggo, and A. Lundberg. Electrophysiological studies on gamma motoneurones. Acta Physiol. Scand. 50: 32–40, 1960.
 133. Eccles, J. C., R. M. Eccles, A. Iggo, and A. Lundberg. Electrophysiological investigations of Renshaw cells. J. Physiol. London 159: 461–478, 1961.
 134. Eccles, J. C., R. M. Eccles, and A. Lundberg. Synaptic actions on motoneurones in relation to the two components of the group I muscle afferent volley. J. Physiol. London 136: 527–546, 1957.
 135. Eccles, J. C., R. M. Eccles, and A. Lundberg. The convergence of monosynaptic excitatory afferents on to many different species of alpha motoneurones. J. Physiol. London 137: 22–50, 1957.
 136. Eccles, J. C., R. M. Eccles, and A. Lundberg. Synaptic actions on motoneurones caused by impulses in Golgi tendon organ afferents. J. Physiol. London 138: 227–252, 1957.
 137. Eccles, J. C., R. M. Eccles, and A. Lundberg. Types of neurone in and around the intermediate nucleus of the lumbosacral cord. J. Physiol. London 154: 89–114, 1960.
 138. Eccles, J. C., R. M. Eccles, and F. Magni. Central inhibitory action attributable to presynaptic depolarization produced by muscle afferent volleys. J. Physiol. London 159: 147–166, 1961.
 139. Eccles, J. C., P. Fatt, and K. Koketsu. Cholinergic and inhibitory synapses in a pathway from motor‐axon collaterals to motoneurones. J. Physiol. London 126: 524–562, 1954.
 140. Eccles, J. C., P. Fatt, and S. Landgren. Central pathway for direct inhibitory action of impulses in largest afferent nerve fibres to muscle. J. Neurophysiol. 19: 75–98, 1956.
 141. Eccles, J. C., P. Fatt, S. Landgren, and G. J. Winsbury. Spinal cord potentials generated by volleys in the large muscle afferents. J. Physiol. London 125: 590–606, 1954.
 142. Eccles, J. C., J. I. Hubbard, and O. Oscarsson. Intracellular recording from cells of the ventral spinocerebellar tract. J. Physiol. London 158: 486–516, 1961.
 143. Eccles, J. C., P. G. Kostyuk, and R. F. Schmidt. Presynaptic inhibition of the central actions of flexor reflex afferents. J. Physiol. London 161: 258–281, 1962.
 144. Eccles, J. C., F. Magni, and W. D. Willis. Depolarization of central terminals of group I afferent fibres from muscle. J. Physiol. London 160: 62–93, 1962.
 145. Eccles, J. C., R. F. Schmidt, and W. D. Willis. Depolarization of central terminals of group Ib afferent fibers of muscle. J. Neurophysiol. 26: 1–27, 1963.
 146. Eccles, J. C., R. F. Schmidt, and W. D. Willis. Depolarization of the central terminals of cutaneous afferent fibers. J. Neurophysiol. 26: 646–661, 1963.
 147. Eccles, R. M., and A. Lundberg. Integrative pattern of Ia synaptic actions on motoneurones of hip and knee muscles. J. Physiol. London 144: 271–298, 1958.
 148. Eccles, R. M., and A. Lundberg. Synaptic actions in motoneurones by afferents which may evoke the flexion reflex. Arch. Ital. Biol. 97: 199–221, 1959.
 149. Eccles, R. M., and A. Lundberg. Supraspinal control of interneurones mediating spinal reflexes. J. Physiol. London 147: 565–584, 1959.
 150. Edgerton, V. R., S. Grillner, A. Sjöström, and P. Zangger. Central generation of locomotion in vertebrates. In: Neural Control of Locomotion, Advances in Behavioral Biology, edited by R. M. Herman, S. Grillner, P. S. G. Stein and D. G. Stuart. New York: Plenum, 1976, vol. 18, p. 439–464.
 151. Ekerot, C.‐F., and O. Oscarsson. Inhibitory spinal paths to the lateral reticular nucleus. Brain Res. 99: 157–161, 1975.
 152. Ekholm, J., G. Eklund, and S. Skoglund. On the reflex effects from the knee joint of the cat. Acta Physiol. Scand. 50: 167–174, 1960.
 153. Eklund, G., C. von Euler, and S. Rutkowski. Spontaneous and reflex activity of intercostal gamma motoneurones. J. Physiol. London 171: 139–163, 1964.
 154. Eklund, G., and K.‐E. Hagbarth. Normal variability of tonic vibration reflexes in man. Exp. Neurol. 16: 80–92, 1966.
 155. Eldred, E. Posture and locomotion. In: Handbook of Physiology. Neurophysiology, edited by J. Field, H. W. Magoun, and V. E. Hall. Washington, DC: Am. Physiol. Soc., 1960, sect. 1, vol. II, p. 1067–1088.
 156. Eldred, E., and B. Fujimori. Relations of the reticular formation to muscle spindle activation. In: Reticular Formation of the Brain, edited by H. H. Jasper, L. D. Proctor, R. S. Knighton, W. C. Noshay and R. T. Costello. Boston: Little, Brown, 1958, p. 275–283.
 157. Eldred, E., R. Granit, and P. A. Merton. Supraspinal control of the muscle spindles and its significance. J. Physiol. London 122: 498–523, 1953.
 158. Elger, C. E., E.‐J. Speckmann, H. Caspers, and R. W. C. Janzen. Cortico‐spinal connections in the rat. I. Monosynaptic and polysynaptic responses of cervical motoneurons to epicortical stimulation. Exp. Brain Res. 28: 385–404, 1977.
 159. Ellaway, P. H. Antidromic inhibition of fusimotor neurones. J. Physiol. London 198: 39P–40P, 1968.
 160. Ellaway, P. H. Recurrent inhibition of fusimotor neurones exhibiting background discharges in the decerebrate and the spinal cat. J. Physiol. London 216: 419–439, 1971.
 161. Ellaway, P. H., and J. R. Trott. The mode of action of 5‐hydroxytryptophan in facilitating a stretch reflex in the spinal cat. Exp. Brain Res. 22: 145–162, 1975.
 162. Ellaway, P. H., and J. R. Trott. Reflex connections from muscle stretch receptors to their own fusimotor neurones. In: Progress in Brain Research. Understanding the Stretch Reflex, edited by S. Homma. Amsterdam: Elsevier, 1976, vol. 44, p. 113–122.
 163. Ellaway, P. H., and J. R. Trott. Autogenetic reflex action on to gamma motoneurones by stretch of triceps surae in the decerebrated cat. J. Physiol. London 276: 49–66, 1978.
 164. Emonet‐Dénand, F., L. Jami, and Y. Laporte. Skeleto‐fusimotor axons in hind‐limb muscles of the cat. J. Physiol. London 249: 153–166, 1975.
 165. Engberg, I. Reflexes to foot muscles in the cat. Acta Physiol. Scand. 62: (Suppl. 235) 1–64, 1964.
 166. Engberg, I., and A. Lundberg. An electromyographic analysis of muscular activity in the hindlimb of the cat during unrestrained locomotion. Acta Physiol. Scand. 75: 614–630, 1969.
 167. Engberg, I., A. Lundberg, and R. W. Ryall. Reticulospinal inhibition of transmission in reflex pathways. J. Physiol. London 194: 201–223, 1968.
 168. Engberg, I., A. Lundberg, and R. W. Ryall. Reticulospinal inhibition of interneurones. J. Physiol. London 194: 225–236, 1968.
 169. Engberg, I., A. Lundberg, and R. W. Ryall. The effect of reserpine on transmission in the spinal cord. Acta Physiol. Scand. 72: 115–122, 1968.
 170. Engberg, I., A. Lundberg, and R. W. Ryall. Is the tonic decerebrate inhibition of reflex paths mediated by monoaminergic pathways. Acta Physiol. Scand. 72: 123–133, 1968.
 171. Euler, C. von. The control of respiratory movement. In: Breathlessness, edited by J. B. L. Howell and E. J. M. Campbell. Oxford: Blackwell, 1966, p. 19–32.
 172. Fedina, L., and H. Hultborn. Facilitation from ipsilateral primary afferents of interneuronal transmission in the Ia inhibitory pathway to motoneurones. Acta Physiol. Scand. 86: 59–81, 1972.
 173. Fedina, L., H. Hultborn, and M. Illert. Facilitation from contralateral primary afferents of interneuronal transmission in the Ia inhibitory pathway to motoneurones. Acta Physiol. Scand. 94: 198–221, 1975.
 174. Fedina, L., A. Lundberg, and L. Vyklický. The effect of a noradrenaline liberator (4,alpha‐dimethyl‐meta‐tyramine) on reflex transmission in spinal cats. Acta Physiol. Scand. 83: 495–504, 1971.
 175. Feldman, A. G., and G. N. Orlovsky. Activity of interneurons mediating reciprocal Ia inhibition during locomotion. Brain Res. 84: 181–194, 1975.
 176. Ferrell, W. R. Slowly Adapting Receptors in the Cat Knee Joint and Their Role in Position Sense (Dissertation). Glasgow: Univ. of Glasgow Library, 1977.
 177. Ferrell, W. R. The discharge of mechanoreceptors in the cat knee joint at intermediate angles. J. Physiol. London 268: 23P–24P, 1977.
 178. Fetz, E. E. Pyramidal tract effects on interneurons in the cat lumbar dorsal horn. J. Neurophysiol. 31: 69–80, 1968.
 179. Fetz, E. E., E. Jankowska, T. Johannisson, and J. Lipski. Autogenetic inhibition of motoneurones by impulses in group Ia muscle spindle afferents. J. Physiol. London 293: 173–195, 1979.
 180. Fidone, S. J., and J. B. Preston. Patterns of motor cortex control of flexor and extensor cat fusimotor neurons. J. Neurophysiol. 32: 103–115, 1969.
 181. Forssberg, H. Stumbling corrective reaction: A phase‐dependent compensatory reaction during locomotion. J. Neurophysiol. 42: 936–953, 1979.
 182. Forssberg, H., and S. Grillner. The locomotion of the acute spinal cat injected with clonidine i.v. Brain Res. 50: 184–186, 1973.
 183. Forssberg, H., S. Grillner, and S. Rossignol. Phasic gain control of reflexes from the dorsum of the paw during spinal locomotion. Brain Res. 132: 121–139, 1977.
 184. Forssberg, H., S. Grillner, S. Rossignol, and P. Wallén. Phasic control of reflexes during locomotion in vertebrates. In: Neural Control of Locomotion. Advances in Behavioral Biology, edited by R. M. Herman, S. Grillner, P. S. G. Stein and D. G. Stuart. New York: Plenum, 1976, vol. 18, p. 647–674.
 185. Forssberg, H., S. Grillner, and A. Sjöström. Tactile placing reactions in chronic spinal kittens. Acta Physiol. Scand. 92: 114–120, 1974.
 186. Foster, M. A Textbook of Physiology. London, 1879. [Cited by E. G. T. Liddell: The Discovery of Reflexes. Oxford: Clarendon, 1960, p. 101.].
 187. Frank, K., and M. G. F. Fuortes. Presynaptic and postsynaptic inhibition of monosynaptic reflexes (Abstract). Federation Proc. 16: 39–40, 1957.
 188. Fritz, N., M. Illert, and P. Saggau. Monosynaptic convergence of group I muscle afferents from the forelimb onto dorsal interosseus motoneurones. Neurosci. Lett.: (Suppl. 1) S95, 1978.
 189. Fromm, C., J. Haase, and E. Wolf. Depression of the recurrent inhibition of extensor motoneurons by the action of group II afferents. Brain Res. 120: 459–468, 1977.
 190. Fromm, C., and J. Noth. Autogenetic inhibition of γ‐motoneurones in the spinal cat uncovered by Dopa injection. Pfluegers Arch. 349: 247–256, 1974.
 191. Fromm, C., and J. Noth. Reflex responses of gamma motoneurones to vibration of the muscle they innervate. J. Physiol. London 256: 117–136, 1976.
 192. Fu, T.‐C., H. Hultborn, R. Larsson, and A. Lundberg. Reciprocal inhibition during the tonic stretch reflex in the decerebrate cat. J. Physiol. London 284: 345–369, 1978.
 193. Fu, T.‐C., E. Jankowska, and A. Lundberg. Reciprocal Ia inhibition during the late reflexes evoked from the flexor reflex afferents after DOPA. Brain Res. 85: 99–102, 1975.
 194. Fu, T.‐C., E. Jankowska, and R. Tanaka. Effects of volleys in cortico‐spinal tract fibres on ventral spino‐cerebellar tract cells in the cat. Acta Physiol. Scand 100: 1–13, 1977.
 195. Fu, T.‐C., M. Santini, and E. D. Schomburg. Characteristics and distribution of spinal focal synaptic potentials generated by group II muscle afferents. Acta Physiol. Scand. 91: 298–313, 1974.
 196. Fu, T.‐C., and E. D. Schomburg. Electrophysiological investigation of the projection of secondary muscle spindle afferents in the cat spinal cord. Acta Physiol. Scand. 91: 314–329, 1974.
 197. Fulton, J. F. Muscular Contraction and the Reflex Control of Movement. Baltimore: Williams & Wilkins, 1926, p. 1–644.
 198. Fuxe, K. Evidence for the existence of monoamine neurons in the central nervous system. III. The monoamine nerve terminal. Z. Zellforsch. Mikrosk. Anat. 65: 573–596, 1965.
 199. Fuxe, K. Evidence for the existence of monoamine neurons in the central nervous system. IV. The distribution of monoamine terminals in the central nervous system. Acta Physiol. Scand. 64: (Suppl. 247) 39–85, 1965.
 200. Fuxe, K., and G. Jonsson. Further mapping of central 5‐hydroxytryptamine neurons: studies with the neurotoxic dihydroxytryptamines. In: Advances in Biochemical Psychopharmacology. Serotonin: New Vistas. Histochemistry and Pharmacology, edited by E. Costa, G. L. Gessa, and M. Sandler. New York: Raven, 1974, vol. 10, p. 1–12.
 201. Gelfand, I. M., V. S. Gurfinkel, Y. M. Kots, V. I. Krinskii, M. L. Tsetlin, and M. L. Shik. Investigations of postural activity. Biofizika 9: 710–717, 1964.
 202. Gelfand, I. M., V. S. Gurfinkel, Y. M. Kots, M. L. Tsetlin, and M. L. Shik. Synchronization of motor units and associated model concepts. Biofizika 8: 475–486, 1963.
 203. Gernandt, B. E., and D. Megirian. Ascending propriospinal mechanisms. J. Neurophysiol. 24: 364–376, 1961.
 204. Gernandt, B. E., and M. Shimamura. Mechanisms of interlimb reflexes in cat. J. Neurophysiol. 24: 665–676, 1961.
 205. Gershon, M. D. Biochemistry and physiology of serotonergic transmission. In: Handbook of Physiology. The Nervous System, edited by J. M. Brookhart and V. B. Mountcastle. Bethesda, MD: Am. Physiol. Soc. 1977, sect. 1, vol. I, pt. 1, chapt. 16, p. 573–623.
 206. Giovanelli Barilari, M., and H. G. J. M. Kuypers. Propriospinal fibers interconnecting the spinal enlargements in the cat. Brain Res. 14: 321–330, 1969.
 207. Goodwin, G. M., and E. S. Luschei. Discharge of spindle afferents from jaw‐closing muscles during chewing in alert monkeys. J. Neurophysiol. 38: 560–571, 1975.
 208. Górska, T., E. Jankowska, and M. Mossakowski. Effects of pyramidotomy on instrumental conditioned reflexes in cats. II. Reflexes derived from unconditioned reactions. Acta Biol. Exp. Warsaw 26: 451–462, 1966.
 209. Graham Brown, T. The intrinsic factors in the act of progression in the mammal. Proc. R. Soc. London Ser. B 84: 308–319, 1911.
 210. Graham Brown, T. Studies in the physiology of the nervous system. IX. Reflex terminal phenomena‐rebound‐rhythmic rebound and movements of progression. Q. J. Exp. Physiol. 4: 331–397, 1911.
 211. Graham Brown, T. The factors in rhythmic activity of the nervous system. Proc. R. Soc. London Ser. B 85: 278–289, 1912.
 212. Graham Brown, T. Studies in the physiology of the nervous system. XI. Immediate reflex phenomena in the simple reflex. Q. J. Exp. Physiol. 5: 237–307, 1912.
 213. Graham Brown, T. Die Reflexfunktionen des Zentralnerven‐systems mit besonderer Berücksichtigung der rhythmischen Tätigkeiten beim Säugetier. Wiesbaden: Bergmann, 1916, vol. II, p. 480–790.
 214. Graham Brown, T. Studies in the physiology of the nervous system. XXVIII. Absence of algebraic equality between the magnitudes of central excitation and effective central inhibition given in the reflex centre of a single limb by the same reflex stimulus. Q. J. Exp. Physiol. 14: 1–23, 1924.
 215. Granit, R. Reflex self‐regulation of muscle contraction and autogenetic inhibition. J. Neurophysiol. 13: 351–372, 1950.
 216. Granit, R. Reflexes to stretch and contraction of antagonists around ankle joint. J. Neurophysiol. 15: 269–279, 1952.
 217. Granit, R. Receptors and Sensory Perception. New Haven, CT: Yale Univ. Press, 1955.
 218. Granit, R. The Basis of Motor Control. London: Academic, 1970.
 219. Granit, R. The functional role of the muscle spindles—facts and hypotheses. Brain 98: 531–556, 1975.
 220. Granit, R., J. Haase, and L. T. Rutledge. Recurrent inhibition in relation to frequency of firing and limitation of discharge rate of extensor motoneurones. J. Physiol. London 154: 308–328, 1960.
 221. Granit, R., B. Holmgren, and P. A. Merton. The two routes for excitation of muscle and their subservience to the cerebellum. J. Physiol. London 130: 213–224, 1955.
 222. Granit, R., C. Job, and B. R. Kaada. Activation of muscle spindles in pinna reflex. Acta Physiol. Scand. 27: 161–168, 1952.
 223. Granit, R., and B. R. Kaada. Influence of stimulation of central nervous structures on muscle spindles in cat. Acta Physiol. Scand. 27: 130–160, 1952.
 224. Granit, R., J. E. Pascoe, and G. Steg. The behaviour of tonic α and γ motoneurones during stimulation of recurrent collaterals. J. Physiol. London 138: 381–400, 1957.
 225. Granit, R., C. G. Phillips, S. Skoglund, and G. Steg. Differentiation of tonic from phasic alpha ventral horn cells by stretch, pinna and crossed extensor reflexes. J. Neurophysiol. 20: 470–481, 1957.
 226. Grant, G., M. Illert, and R. Tanaka. Integration in descending motor pathways controlling the forelimb in the cat. 6. Anatomical evidence consistent with the existence of C3–C4 propriospinal neurones projecting to forelimb motornuclei. Exp. Brain Res. 38: 87–93, 1980.
 227. Grigg, P. Mechanical factors influencing response of joint afferent neurons from cat knee. J. Neurophysiol. 38: 1473–1484, 1975.
 228. Grigg, P., and B. J. Greenspan. Response of primate joint afferent neurons to mechanical stimulation of knee joint. J. Neurophysiol. 40: 1–8, 1977.
 229. Grigg, P., E. P. Harrigan, and K. E. Fogarty. Segmental reflexes mediated by joint afferent neurons in cat knee. J. Neurophysiol. 41: 9–14, 1978.
 230. Grigg, P., and J. B. Preston. Baboon flexor and extensor fusimotor neurons and their modulation by motor cortex. J. Neurophysiol. 34: 428–436, 1971.
 231. Grillner, S. The influence of DOPA on the static and the dynamic fusimotor activity to the triceps surae of the spinal cat. Acta Physiol. Scand. 77: 490–509, 1969.
 232. Grillner, S. Supraspinal and segmental control of static and dynamic γ‐motoneurones in the cat. Acta Physiol. Scand.: (Suppl. 327) 1–34, 1969.
 233. Grillner, S. Locomotion in the spinal cat. In: Advances in Behavioral Biology. Control of Posture and Locomotion, edited by R. B. Stein, K. B. Pearson, R. S. Smith and J. B. Redford. New York: Plenum, 1973, vol. 7, p. 515–535.
 234. Grillner, S. Locomotion in vertebrates: central mechanisms and reflex interaction. Physiol. Rev. 55: 247–304, 1975.
 235. Grillner, S. On the neural control of movement—a comparison of different basic rhythmic behaviors. In: Function and Formation of Neural Systems, edited by G. S. Stent. Berlin: Dahlem Konferenzen, 1977, p. 197–224.
 236. Grillner, S., and T. Hongo. Vestibulospinal effects on motoneurones and interneurones in the lumbosacral cord. In: Progress in Brain Research. Basic Aspects of Central Vestibular Mechanisms, edited by A. Brodal and O. Pompeiano. Amsterdam: Elsevier, 1972, vol. 37, p. 243–262.
 237. Grillner, S., T. Hongo, and S. Lund. Interaction between the inhibitory pathways from the Deiters' nucleus and Ia afferents to flexor motoneurones. Acta Physiol. Scand. 68: (Suppl. 277) 61, 1966.
 238. Grillner, S., T. Hongo, and S. Lund. The origin of descending fibres monosynaptically activating spinoreticular neurones. Brain Res. 10: 259–262, 1968.
 239. Grillner, S., T. Hongo, and S. Lund. Descending monosynaptic and reflex control of γ‐motoneurones. Acta Physiol. Scand. 75: 592–613, 1969.
 240. Grillner, S., T. Hongo, and S. Lund. The vestibulospinal tract. Effects on alpha‐motoneurones in the lumbosacral spinal cord in the cat. Exp. Brain Res. 10: 94–120, 1970.
 241. Grillner, S., T. Hongo, and S. Lund. Convergent effects on alpha motoneurones from the vestibulospinal tract and a pathway descending in the medial longitudinal fasciculus. Exp. Brain Res. 12: 457–479, 1971.
 242. Grillner, S., T. Hongo, and A. Lundberg. The effect of DOPA on the spinal cord. 7. Reflex activation of static γ‐motoneurones from the flexor reflex afferents. Acta Physiol. Scand. 70: 403–411, 1967.
 243. Grillner, S., and S. Lund. A descending pathway with monosynaptic action on flexor motoneurones. Experientia 22: 390, 1966.
 244. Grillner, S., and S. Lund. The origin of a descending pathway with monosynaptic action on flexor motoneurones. Acta Physiol. Scand. 74: 274–284, 1968.
 245. Grillner, S., and S. Rossignol. Contralateral reflex reversal controlled by limb position in the acute spinal cat injected with clonidine i.v. Brain Res. 144: 411–414, 1978.
 246. Grillner, S., and S. Rossignol. On the initiation of the swing phase of locomotion in chronic spinal cats. Brain Res. 146: 269–277, 1978.
 247. Grillner, S., and M. L. Shik. On the descending control of the lumbosacral spinal cord from the “mesencephalic locomotor region.” Acta Physiol. Scand. 87: 320–330, 1973.
 248. Grillner, S., and M. Udo. Is the tonic stretch reflex dependent on suppression of autogenetic inhibitory reflexes? Acta Physiol. Scand. 79: 13A–14A, 1970.
 249. Grillner, S., and P. Zangger. Locomotor movements generated by the deafferented spinal cord. Acta Physiol. Scand. 91: 38A–39A, 1974.
 250. Grillner, S., and P. Zangger. On the central generation of locomotion in the low spinal cat. Exp. Brain Res. 34: 241–261, 1979.
 251. Gustafsson, B., and S. Lindström. Recurrent control from motor axon collaterals of Ia inhibitory pathways to ventral spinocerebellar tract neurones. Acta Physiol. Scand. 89: 457–481, 1973.
 252. Ha, H., and C.‐N. Liu. Cell origin of the ventral spinocerebellar tract. J. Comp. Neurol. 133: 185–206, 1968.
 253. Haase, J., S. Cleveland, and H.‐G. Ross. Problems of postsynaptic autogenous and recurrent inhibition in the mammalian spinal cord. Rev. Physiol. Biochem. Pharmacol. 73: 74–129, 1975.
 254. Haase, J., and J. P. Van Der Meulen. Effects of supraspinal stimulation on Renshaw cells belonging to extensor motoneurones. J. Neurophysiol. 24: 510–520, 1961.
 255. Haase, J., and B. Vogel. Die Erregung der Renshaw‐Zellen durch reflektorische Entladungen der α‐Motoneurone. Pfluegers Arch. 325: 14–27, 1971.
 256. Hagbarth, K.‐E. Excitatory and inhibitory skin areas for flexor and extensor motoneurones. Acta Physiol. Scand. 26: (Suppl. 94) 1–58, 1952.
 257. Hagbarth, K.‐E., and G. Eklund. Motor effects of vibratory stimuli in man. In: Muscular Afferents and Motor Control. Nobel Symposium 1, edited by R. Granit. Stockholm: Almqvist & Wiksell, 1966, p. 177–186.
 258. Hammond, P. H. An experimental study of servo action in human muscular control. IEE Conf. Publ. London 3: 190–199, 1960.
 259. Hammond, P. H., P. A. Merton, and G. G. Sutton. Nervous gradation of muscular contraction. Brit. Med. Bull. 12: 214–218, 1956.
 260. Harris, A. S. Mammalian plantar reflexes in terms of afferent fibers. Am. J. Physiol. 124: 117–125, 1938.
 261. Henatsch, H.‐D., H.‐J. Kaese, D. Langrehr, and J. Meyer‐Lohmann. Einflüsse des motorischen Cortex der Katze auf die Renshaw‐Rückkoppelungshemmung der Motoneurone. Pfluegers Arch. 274: 51–52, 1961.
 262. Hikosaka, O., Y. Igusa, S. Nakao, and H. Shimazu. Direct inhibitory synaptic linkage of pontomedullary reticular burst neurons with abducens motoneurons in the cat. Exp. Brain Res. 33: 337–352, 1978.
 263. Hirai, N., T. Hongo, N. Kudo, and T. Yamaguchi. Heterogenous composition of the spinocerebellar tract originating from the cervical enlargement in the cat. Brain Res. 109: 387–391, 1976.
 264. Hirai, N., T. Hongo, and T. Yamaguchi. Spinocerebellar tract neurones with long descending axon collaterals. Brain Res. 142: 147–151, 1978.
 265. Hodge, C. J., Jr. Potential changes inside central afferent terminals secondary to stimulation of large‐ and small‐diameter peripheral nerve fibers. J. Neurophysiol. 35: 30–43, 1972.
 266. Holmqvist, B. Crossed spinal reflex actions evoked by volleys in somatic afferents. Acta Physiol. Scand. 52: (Suppl. 181) 1–67, 1961.
 267. Holmqvist, B., and A. Lundberg. On the organization of the supraspinal inhibitory control of interneurones of various spinal reflex arcs. Arch. Ital. Biol. 97: 340–356, 1959.
 268. Holmqvist, B., and A. Lundberg. Differential supraspinal control of synaptic actions evoked by volleys in the flexion reflex afferents in alpha motoneurons. Acta Physiol. Scand. 54: (Suppl. 186) 1–51, 1961.
 269. Holmqvist, B., A. Lundberg, and O. Oscarsson. Supraspinal inhibitory control of transmission to three ascending spinal pathways influenced by the flexion reflex afferents. Arch. Ital. Biol. 98: 60–80, 1960.
 270. Holmqvist, B., A. Lundberg, and O. Oscarsson. A supraspinal control system monosynaptically connected with an ascending spinal pathway. Arch. Ital. Biol. 98: 402–422, 1960.
 271. Homma, S., M. Mizote, and S. Watanabe. Participation of mono‐ and polysynaptic transmission during tonic activation of the stretch reflex arcs. Jpn. J. Physiol. 25: 135–146, 1975.
 272. Hongo, T., and E. Jankowska. Effects from the sensorimotor cortex on the spinal cord in cats with transected pyramids. Exp. Brain Res. 3: 117–134, 1967.
 273. Hongo, T., E. Jankowska, and A. Lundberg. Convergence of excitatory and inhibitory action on interneurones in the lumbosacral cord. Exp. Brain Res. 1: 338–358, 1966.
 274. Hongo, T., E. Jankowska, and A. Lundberg. The rubrospinal tract. I. Effects on alpha‐motoneurones innervating hindlimb muscles in cats. Exp. Brain Res. 7: 344–364, 1969.
 275. Hongo, T., E. Jankowska, and A. Lundberg. The rubrospinal tract. II. Facilitation of interneuronal transmission in reflex paths to motoneurones. Exp. Brain Res. 7: 365–391, 1969.
 276. Hongo, T., E. Jankowska, and A. Lundberg. The rubrospinal tract. III. Effects on primary afferent terminals. Exp. Brain Res. 15: 39–53, 1972.
 277. Hongo, T., E. Jankowska, and A. Lundberg. The rubrospinal tract. IV. Effects on interneurones. Exp. Brain Res. 15: 54–78, 1972.
 278. Hongo, T., N. Kudo, and R. Tanaka. The vestibulospinal tract: crossed and uncrossed effects on hindlimb motoneurones in the cat. Exp. Brain Res. 24: 37–55, 1975.
 279. Hongo, T., and Y. Okada. Cortically evoked pre‐ and postsynaptic inhibition of impulse transmission to the dorsal spinocerebellar tract. Exp. Brain Res. 3: 163–177, 1967.
 280. Houk, J., and E. Henneman. Feedback control of skeletal muscles. Brain Res. 5: 433–451, 1967.
 281. Houk, J., and E. Henneman. Responses of Golgi tendon organs to active contractions of the soleus muscle of the cat. J. Neurophysiol. 30: 466–481, 1967.
 282. Houk, J. C., J. J. Singer, and M. R. Goldman. An evaluation of length and force feedback to soleus muscles of decerebrate cats. J. Neurophysiol. 33: 784–811, 1970.
 283. Hubbard, J. I., and O. Oscarsson. Localization of the cell bodies of the ventral spino‐cerebellar tract in lumbar segments of the cat. J. Comp. Neurol. 118: 199–204, 1962.
 284. Hulliger, M., E. Nordh, A.‐E. Thelin, and Å. B. Vallbo. The responses of afferent fibres from glabrous skin of the hand during voluntary finger movements in man. J. Physiol. London 291: 233–249, 1979.
 285. Hultborn, H. Convergence on interneurones in the reciprocal Ia inhibitory pathway to motoneurones. Acta Physiol. Scand. 85: (Suppl. 375) 1–42, 1972.
 286. Hultborn, H. Transmission in the pathway of reciprocal Ia inhibition to motoneurones and its control during the tonic stretch reflex. In: Progress in Brain Research. Understanding the Stretch Reflex, edited by S. Homma. Amsterdam: Elsevier, 1976, vol. 44, p. 235–255.
 287. Hultborn, H., M. Illert, and M. Santini. Convergence on interneurones mediating the reciprocal Ia inhibition of motoneurones. I. Disynaptic Ia inhibition of Ia inhibitory interneurones. Acta Physiol. Scand. 96: 193–201, 1976.
 288. Hultborn, H., M. Illert, and M. Santini. Convergence on interneurones mediating the reciprocal Ia inhibition of motoneurones. II. Effects from segmental flexor reflex pathways. Acta Physiol. Scand. 96: 351–367, 1976.
 289. Hultborn, H., M. Illert, and M. Santini. Convergence on interneurones mediating the reciprocal Ia inhibition of motoneurones. III. Effects from supraspinal pathways. Acta Physiol. Scand. 96: 368–391, 1976.
 290. Hultborn, H., E. Jankowska, and S. Lindström. Recurrent inhibition from motor axon collaterals of transmission in the Ia inhibitory pathway to motoneurones. J. Physiol. London 215: 591–612, 1971.
 291. Hultborn, H., E. Jankowska, and S. Lindström. Recurrent inhibition of interneurones monosynaptically activated from group Ia afferents. J. Physiol. London 215: 613–636, 1971.
 292. Hultborn, H., E. Jankowska, and S. Lindström. Relative contribution from different nerves to recurrent depression of Ia IPSPs in motoneurones. J. Physiol. London 215: 637–664, 1971.
 293. Hultborn, H., S. Lindström, and H. Wigström. On the function of recurrent inhibition in the spinal cord. Exp. Brain Res. 37: 399–403, 1979.
 294. Hultborn, H., and E. Pierrot‐Deseilligny. Changes in recurrent inhibition during voluntary soleus contractions in man studied by an H‐reflex technique. J. Physiol. London 297: 229–251, 1979.
 295. Hultborn, H., and E. Pierrot‐Deseilligny. Input‐output relations in the pathway of recurrent inhibition to motoneurones in the cat. J. Physiol. London 297: 267–287, 1979.
 296. Hultborn, H., E. Pierrot‐Deseilligny, and H. Wigström. Distribution of recurrent inhibition within a motor nucleus. Neurosci. Lett.: (Suppl. 1) S97, 1978.
 297. Hultborn, H., E. Pierrot‐Deseilligny, and H. Wigström. Recurrent inhibition and afterhyperpolarization following motoneuronal discharge in the cat. J. Physiol. London 297: 253–266, 1979.
 298. Hultborn, H., and M. Udo. Convergence in the reciprocal Ia inhibitory pathway of excitation from descending pathways and inhibition from motor axon collaterals. Acta Physiol. Scand. 84: 95–108, 1972.
 299. Hultborn, H., and M. Udo. Convergence of large muscle spindle (Ia) afferents at interneuronal level in the reciprocal Ia inhibitory pathway to motoneurones. Acta Physiol. Scand. 84: 493–499, 1972.
 300. Hultborn, H., and M. Udo. Recurrent depression from motor axon collaterals of supraspinal inhibition in motoneurones. Acta Physiol. Scand. 85: 44–57, 1972.
 301. Hultborn, H., and H. Wigström. Motor response with long latency and maintained duration evoked by activity in Ia afferents. In: Progress in Clinical Neurophysiology, Spinal and Supraspinal Mechanisms of Voluntary Motor Control and Locomotion, edited by J. E. Desmedt. Basel: Karger, 1980, vol. 8, p. 99–115.
 302. Hultborn, H., H. Wigström, and B. Wängberg. Prolonged activation of soleus motoneurones following a conditioning train in soleus Ia afferents — a case for a reverberating loop? Neurosci. Lett. 1: 147–152, 1975.
 303. Hunt, C. C. The reflex activity of mammalian small‐nerve fibres. J. Physiol. London 115: 456–469, 1951.
 304. Hunt, C. C. The effect of stretch receptors from muscle on the discharge of motoneurones. J. Physiol. London 117: 359–379, 1952.
 305. Hunt, C. C. Relation of function to diameter in afferent fibers of muscle nerves. J. Gen. Physiol. 38: 117–131, 1954.
 306. Hunt, C. C., and S. W. Kuffler. Stretch receptor discharges during muscle contraction. J. Physiol. London 113: 298–315, 1951.
 307. Hunt, C. C., and A. K. McIntyre. Characteristics of responses from receptors from the flexor longus digitorum muscle and the adjoining interosseous region of the cat. J. Physiol. London 153: 74–87, 1960.
 308. Hunt, C. C., and A. S. Paintal. Spinal reflex regulation of fusimotor neurones. J. Physiol. London 143: 195–212, 1958.
 309. Illert, M., and A. Lundberg. Collateral connections to the lateral reticular nucleus from cervical propriospinal neurones projecting to forelimb motoneurones in the cat. Neurosci. Lett. 7: 167–172, 1978.
 310. Illert, M., A. Lundberg, Y. Padel, and R. Tanaka. Convergence on propriospinal neurones which may mediate disynaptic corticospinal excitation to forelimb motoneurones in the cat. Brain Res. 93: 530–534, 1975.
 311. Illert, M., A. Lundberg, Y. Padel, and R. Tanaka. Integration in descending motor pathways controlling the forelimb in the cat. 5. Properties of and monosynaptic excitatory convergence on C3–C4 propriospinal neurones. Exp. Brain Res. 33: 101–130, 1978.
 312. Illert, M., A. Lundberg, and R. Tanaka. Integration in descending motor pathways controlling the forelimb in the cat. 1. Pyramidal effects on motoneurones. Exp. Brain Res. 26: 509–519, 1976.
 313. Illert, M., A. Lundberg, and R. Tanaka. Integration in descending motor pathways controlling the forelimb in the cat. 2. Convergence on neurones mediating disynaptic cortico‐motoneuronal excitation. Exp. Brain Res. 26: 521–540, 1976.
 314. Illert, M., A. Lundberg, and R. Tanaka. Integration in descending motor pathways controlling the forelimb in the cat. 3. Convergence on propriospinal neurones transmitting disynaptic excitation from the corticospinal tract and other descending tracts. Exp. Brain Res. 29: 323–346, 1977.
 315. Illert, M., and R. Tanaka. Integration in descending motor pathways controlling the forelimb in the cat. 4. Corticospinal inhibition of forelimb motoneurones mediated by short propriospinal neurones. Exp. Brain Res. 31: 131–141, 1978.
 316. Jack, J. J. B. Some methods for selective activation of muscle afferent fibres. In: Studies in Neurophysiology. Essays in Honour of Professor A. K. McIntyre. Cambridge: Cambridge Univ. Press, 1978.
 317. Jack, J. J. B., and C. R. MacLennan. The lack of an electrical threshold discrimination between group Ia and group Ib fibres in the nerve to the cat peroneus longus muscle. J. Physiol. London 212: 35P–36P, 1971.
 318. Jack, J. J. B., and R. C. Roberts. Selective electrical activation of group II muscle afferent fibres. J. Physiol. London 241: 82P–84P, 1974.
 319. Jack, J. J. B., and R. C. Roberts. The role of muscle spindle afferents in stretch and vibration reflexes of the soleus muscle of the decerebrate cat. Brain Res. 146: 366–372, 1978.
 320. Jänig, W., R. F. Schmidt, and M. Zimmerman. Single unit responses and the total afferent outflow from the cat's foot pad upon mechanical stimulation. Exp. Brain Res. 6: 100–115, 1968.
 321. Jänig, W., R. F. Schmidt, and M. Zimmerman. Two specific feedback pathways to the central afferent terminals of phasic and tonic mechanoreceptors. Exp. Brain Res. 6: 116–129, 1968.
 322. Jankowska, E. New observations on neuronal organization of reflexes from tendon organ afferents and their relation to reflexes evoked from muscle spindle afferents. In: Progress in Brain Research. Reflex Control of Posture and Movement, edited by R. Granit and O. Pompeiano. Amsterdam: Elsevier, 1979, vol. 50, p. 29–36.
 323. Jankowska, E., T. Johannisson, and J. Lipski. Common interneurones in reflex pathways from group Ia and Ib afferents of ankle extensors in the cat. J. Physiol. London. In press.
 324. Jankowska, E., M. G. M. Jukes, S. Lund, and A. Lundberg. 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, 1967.
 325. Jankowska, E., M. G. M. Jukes, S. Lund, and A. Lundberg. 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, 1967.
 326. Jankowska, E., and S. Lindström. Morphological identification of Renshaw cells. Acta Physiol. Scand. 81: 428–430, 1971.
 327. Jankowska, E., and S. Lindström. Morphology of interneurones mediating Ia reciprocal inhibition of motoneurones in the spinal cord of the cat. J. Physiol. London 226: 805–823, 1972.
 328. Jankowska, E., and S. Lindström. Procion yellow staining of functionally identified interneurones in the spinal cord of the cat. In: Intracellular Staining in Neurobiology, edited by S. B. Kater and C. Nicholson. New York: Springer‐Verlag, 1973, p. 199–209.
 329. Jankowska, E., S. Lund, and A. Lundberg. The effect of DOPA on the spinal cord. 4. Depolarization evoked in the central terminals of contralateral Ia afferent terminals by volleys in the flexor reflex afferents. Acta Physiol. Scand. 68: 337–341, 1966.
 330. Jankowska, E., S. Lund, A. Lundberg, and O. Pompeiano. Inhibitory effects evoked through ventral reticulospinal pathways. Arch. Ital. Biol. 106: 124–140, 1968.
 331. Jankowska, E., A. Lundberg, W. J. Roberts, and D. Stuart. A long propriospinal system with direct effect on motoneurones and on interneurones in the cat lumbosacral cord. Exp. Brain Res. 21: 169–194, 1974.
 332. Jankowska, E., A. Lundberg, and D. Stuart. Propriospinal control of last order interneurones of spinal reflex pathways in the cat. Brain Res. 53: 227–231, 1973.
 333. Jankowska, E., Y. Padel, and R. Tanaka. Disynaptic inhibition of spinal motoneurones from the motor cortex in the monkey. J. Physiol. London 258: 467–487, 1976.
 334. Jankowska, E., Y. Padel, and P. Zarzecki. Crossed disynaptic inhibition of sacral motoneurones. J. Physiol. London 285: 425–444, 1978.
 335. Jankowska, E., and W. J. Roberts. An electrophysiological demonstration of the axonal projections of single spinal interneurones in the cat. J. Physiol. London 222: 597–622, 1972.
 336. Jankowska, E., and W. J. Roberts. Synaptic actions of single interneurones mediating reciprocal Ia inhibition of motoneurones. J. Physiol. London 222: 623–642, 1972.
 337. Jankowska, E., and D. O. Smith. Antidromic activation of Renshaw cells and their axonal projections. Acta Physiol. Scand. 88: 198–214, 1973.
 338. Jansen, J. K. S., and T. Rudjord. On the silent period and Golgi tendon organs of the soleus muscle of the cat. Acta Physiol. Scand. 62: 364–379, 1964.
 339. Jeneskog, T. Parallel activation of dynamic fusimotor neurones and a climbing fibre system from the cat brain stem. I. Effects from the rubral region. Acta Physiol. Scand. 91: 223–242, 1974.
 340. Jeneskog, T. Parallel activation of dynamic fusimotor neurones and a climbing fibre system from the cat brain stem. II. Effects from the inferior olivary region. Acta Physiol. Scand. 92: 66–83, 1974.
 341. Job, C. Über autogene Inhibition und Reflexumkehr bei spin‐alisierten und decerebrierten Katzen. Pfluegers Arch. 256: 406–418, 1953.
 342. Jurna, I., and A. Lundberg. The influence of an inhibitor of dopamine‐beta‐hydroxylase on the effect of DOPA on transmission in the spinal cord. In: Structure and Functions of Inhibitory Neuronal Mechanisms. Oxford: Pergamon, 1968, p. 469–472. (Fourth Int. Meeting Neurobiol., Stockholm, September 1966.).
 343. Kanda, K. Contribution of polysynaptic pathways to the tonic vibration reflex. Jpn. J. Physiol. 22: 367–377, 1972.
 344. Kanda, R., and W. Z. Rymer. An estimate of the secondary spindle receptor afferent contribution to the stretch reflex in extensor muscles of the decerebrate cat. J. Physiol. London 264: 63–87, 1977.
 345. Kandel, E. R., W. T. Frazier, R. Waziri, and R. E. Coggeshall. Direct and common connections among identified neurons in Aplysia. J. Neurophysiol. 30: 1352–1376, 1967.
 346. Kato, M., and K. Fukushima. Effect of differential blocking of motor axons on antidromic activation of Renshaw cells in the cat. Exp. Brain Res. 20: 135–143, 1974.
 347. Kenins, P., H. Kikillus, and E. D. Schomburg. Short‐ and long‐latency reflex pathways from neck afferents to hindlimb motoneurones in the cat. Brain Res. 149: 235–238, 1978.
 348. Kibakina, T. V., and A. I. Shapovalov. Functional organization of reticulo‐ and vestibulo‐spinal synaptic projections to rat lumbar motoneurons [in Russian]. Fiziol. Zh. SSSR im I. M. Sechenov 64: 435–442, 1978.
 349. Kirkwood, P. A., and T. A. Sears. Monosynaptic excitation of motoneurones from muscle spindle secondary endings of intercostal and triceps surae muscles in the cat. J. Physiol. London 245: 64P–66P, 1975.
 350. Kirkwood, P. A., and T. A. Sears. The synaptic connexions to intercostal motoneurones as revealed by the average common excitation potential. J. Physiol. London 275: 103–134, 1978.
 351. Koehler, W., U. Windhorst, J. Schmidt, J. Meyer‐Lohmann, and H.‐D. Henatsch. Diverging influences on Renshaw cell responses and monosynaptic reflexes from stimulation of capsula interna. Neurosci. Lett. 8: 35–39, 1978.
 352. Koeze, T. H. The independence of corticomotoneuronal and fusimotor pathways in the production of muscle contraction by motor cortex stimulation. J. Physiol. London 197: 87–105, 1968.
 353. Koeze, T. H., C. G. Phillips, and J. D. Sheridan. Thresholds of cortical activation of muscle spindles and α motoneurones of the baboon's hand. J. Physiol. London 195: 419–449, 1968.
 354. Kostyuk, P. G. Interneuronal mechanisms of interaction between descending and afferent signals in the spinal cord. In: Golgi Centennial Symposium. Proceedings, edited by M. Santini. New York: Raven, 1975, p. 247–259.
 355. Kostyuk, P. G., and D. A. Vasilenko. Transformation of cortical motor signals in spinal cord. Proc. IEEE 56: 1049–1058, 1968.
 356. Kostyuk, P. G., D. A. Vasilenko, and E. Lang. Propriospinal pathways in the dorsolateral funiculus and their effects on lumbosacral motoneuronal pools. Brain Res. 28: 233–249, 1971.
 357. Kostyuk, P. G., D. A. Vasilenko, and A. G. Zadorozhny. Reactions of lumbar motoneurons produced by actions of propriospinal pathways. Neurophysiology USSR 1: 5–14, 1969.
 358. Kozhanov, V. M. The propriospinal monosynaptic effects of the ventral descending pathways on the cat lumbar motoneurones. Fiziol. Zh. SSSR im I. M. Sechenov 60: 171–178, 1974.
 359. Kozhanov, V. M., and A. I. Shapovalov. Synaptic organization of the supraspinal control of propriospinal ventral horn interneurons in cat and monkey spinal cord. Neurophysiology USSR 9: 177–184, 1977.
 360. Kozhanov, V. M., and A. I. Shapovalov. Synaptic actions evoked in motoneurons by stimulation of individual propriospinal neurons. Neurophysiology USSR 9: 300–306, 1977.
 361. Kuffler, S. W., and C. C. Hunt. The mammalian small‐nerve fibers: a system for efferent nervous regulation of muscle spindle discharge. In: Patterns of Organization in the Central Nervous System, edited by P. Bard. Baltimore: Williams & Wilkins, 1952, p. 24–47.
 362. Kuno, M. Excitability following antidromic activation in spinal motoneurones supplying red muscles. J. Physiol. London 149: 374–393, 1959.
 363. Kuno, M., and E. R. Perl. Alternation of spinal reflexes by interaction with suprasegmental and dorsal root activity. J. Physiol. London 151: 103–122, 1960.
 364. Kuypers, H. Pericentral cortical projections to motor and sensory nuclei. Science 128: 662–663, 1958.
 365. Lance, J. W., P. De Gail, and P. D. Neilson. Tonic and phasic spinal cord mechanisms in man. J. Neurol. Neurosurg. Psychiatry 29: 535–544, 1966.
 366. Landgren, S., C. G. Phillips, and R. Porter. Minimal synaptic actions of pyramidal impulses on some alpha motoneurones of the baboon's hand and forearm. J. Physiol. London 161: 91–111, 1962.
 367. Laporte, Y. The motor innervation of the mammalian muscle spindle. In: Studies in Neurophysiology. Essays in Honour of Professor A. K. McIntyre, edited by R. Porter. Cambridge: Cambridge Univ., 1978, p. 45–59.
 368. Laporte, Y., and P. Bessou. Étude des sous‐groupes lent et rapide du groupe I (fibres afférentes d'origine musculaire de grand diamètre) chez le chat. J. Physiol. Paris 49: 1025–1037, 1957.
 369. Laporte, Y., and D. P. C. Lloyd. Nature and significance of the reflex connections established by large afferent fibers of muscular origin. Am. J. Physiol. 169: 609–621, 1952.
 370. Laporte, Y., A. Lundberg, and O. Oscarsson. Functional organization of the dorsal spinocerebellar tract in the cat. II. Single fibre recording in Flechsig's fasciculus on electrical stimulation of various peripheral nerves. Acta Physiol. Scand. 36: 188–203, 1956.
 371. Larsson, B., S. Miller, and O. Oscarsson. A spinocerebellar climbing fibre path activated by the flexor reflex afferents from all four limbs. J. Physiol. London 203: 641–649, 1969.
 372. Lawrence, D. G., and D. A. Hopkins. The development of motor control in the rhesus monkey: evidence concerning the role of corticomotoneuronal connections. Brain 99: 235–254, 1976.
 373. Lawrence, D. G., and H. G. J. M. Kuypers. The functional organization of the motor system in the monkey. II. The effects of lesions of the descending brain‐stem pathways. Brain 91: 15–36, 1968.
 374. Lennerstrand, G., and U. Thoden. Muscle spindle responses to concomitant variations in length and in fusimotor activation. Acta Physiol. Scand. 74: 153–165, 1968.
 375. Levy, R. A., and E. G. Anderson. The role of γ‐aminobutyric acid as a mediator of positive dorsal root potentials. Brain Res. 76: 71–82, 1974.
 376. Liddell, E. G. T. Spinal shock and some features in isolation‐alteration of the spinal cord in cats. Brain 57: 386–400, 1934.
 377. Liddell, E. G. T. The influence of experimental lesions of the spinal cord upon the knee‐jerk. II. Chronic lesions. With an appendix ‘a note on the “spinal” and “decerebrate” type of knee‐jerk in the cat’. Brain 59: 160–174, 1936.
 378. Liddell, E. G. T., and C. S. Sherrington. Reflexes in response to stretch (myotatic reflexes). Proc. R. Soc. London Ser. B 96: 212–242, 1924.
 379. Liddell, E. G. T., and C. S. Sherrington. Further observations on myotatic reflexes. Proc. R. Soc. London Ser. B 97: 267–283, 1925.
 380. Lindström, S. Recurrent control from motor axon collaterals of Ia inhibitory pathways in the spinal cord of the cat. Acta Physiol. Scand. (Suppl. 392): 1–43, 1973.
 381. Lindström, S., and E. D. Schomburg. Recurrent inhibition from motor axon collaterals of ventral spinocerebellar tract neurones. Acta Physiol. Scand. 88: 505–515, 1973.
 382. Lindström, S., and E. D. Schomburg. Group I inhibition in Ib excited ventral spinocerebellar tract neurones. Acta Physiol. Scand. 90: 166–185, 1974.
 383. Liu, C.‐N. Fate of cells of Clarke's nucleus following axon section in cat, with evidence of spinal collaterals from dorsal spinocerebellar tract. Anat. Record. 115: 342–343, 1953.
 384. Llinas, R., and C. A. Terzuolo. Mechanisms of supraspinal actions upon spinal cord activities. Reticular inhibitory mechanisms on alpha‐extensor motoneurones. J. Neurophysiol. 27: 579–591, 1964.
 385. Llinas, R., and C. A. Terzuolo. Mechanisms of supraspinal actions upon spinal cord activities. Reticular inhibitory mechanisms upon flexor motoneurones. J. Neurophysiol. 28: 413–422, 1965.
 386. Lloyd, D. P. C. Activity in neurons of the bulbospinal correlation system. J. Neurophysiol. 4: 115–134, 1941.
 387. Lloyd, D. P. C. A direct central inhibitory action of dromically conducted impulses. J. Neurophysiol. 4: 184–190, 1941.
 388. Lloyd, D. P. C. The spinal mechanism of the pyramidal system in cats. J. Neurophysiol. 4: 525–546, 1941.
 389. Lloyd, D. P. C. Mediation of descending long spinal reflex activity. J. Neurophysiol. 5: 435–458, 1942.
 390. Lloyd, D. P. C. Reflex action in relation to pattern and peripheral source of afferent stimulation. J. Neurophysiol. 6: 111–119, 1943.
 391. Lloyd, D. P. C. Neuron patterns controlling transmission of ipsilateral hind limb reflexes in cat. J. Neurophysiol. 6: 293–315, 1943.
 392. Lloyd, D. P. C. Conduction and synaptic transmission of the reflex response to stretch in spinal cats. J. Neurophysiol. 6: 317–326, 1943.
 393. Lloyd, D. P. C. Facilitation and inhibition of spinal motoneurons. J. Neurophysiol. 9: 421–438, 1946.
 394. Lloyd, D. P. C. Integrative pattern of excitation and inhibition in two‐neuron reflex arcs. J. Neurophysiol. 9: 439–444, 1946.
 395. Lloyd, D. P. C. Spinal mechanisms involved in somatic activities. In: Handbook of Physiology. Neurophysiology, edited by J. Field, H. W. Magoun, and V. E. Hall. Washington, DC: Am. Physiol. Soc., 1960, sect. 1, vol. II, chapt. 36, p. 929–949.
 396. Lloyd, D. P. C., and A. K. McIntyre. Analysis of forelimb‐hindlimb reflex activity in acutely decapitate cats. J. Neurophysiol. 11: 455–470, 1948.
 397. Loeb, G. E., M. J. Bak, and J. Duysens. Long‐term unit recording from somatosensory neurons in the spinal ganglia of the freely walking cat. Science 197: 1192–1194, 1977.
 398. Loeb, G. E., and J. Duysens. Activity patterns in individual hindlimb primary and secondary muscle spindle afferents during normal movements in unrestrained cats. J. Neurophysiol. 42: 420–440, 1979.
 399. Longo, V. G., W. R. Martin, and K. R. Unna. A pharmacological study on the Renshaw cell. J. Pharmacol. Exp. Ther. 129: 61–68, 1960.
 400. Lucas, M. E., and W. D. Willis. Identification of muscle afferents which activate interneurons in the intermediate nucleus. J. Neurophysiol. 37: 282–293, 1974.
 401. Lund, J. P., A. M. Smith, B. J. Sessle, and T. Murakami. Activity of trigeminal α‐ and γ‐motoneurons and muscle afferents during performance of a biting task. J. Neurophysiol. 42: 710–725, 1979.
 402. Lund, S., A. Lundberg, and L. Vyklický. Inhibitory action from the flexor reflex afferents on transmission to Ia afferents. Acta Physiol. Scand. 64: 345–355, 1965.
 403. Lund, S., and O. Pompeiano. Monosynaptic excitation of alpha motoneurones from supraspinal structures in the cat. Acta Physiol. Scand. 73: 1–21, 1968.
 404. Lundberg, A. Integrative significance of patterns of connections made by muscle afferents in the spinal cord. In: Congr. Int. Cienc. Fisiol., XXI, Buenos Aires, 1959, p. 100–105.
 405. Lundberg, A. Monoamines and spinal reflexes. In: Studies in Physiology. Berlin: Springer‐Verlag, 1965, p. 186–190.
 406. Lundberg, A. Integration in the reflex pathway. In: Muscular Afferents and Motor Control. Nobel Symposium I, edited by R. Granit. Stockholm: Almqvist & Wiksell, 1966, p. 275–305.
 407. Lundberg, A. Convergence of excitatory and inhibitory action on interneurones in the spinal cord. In: The Interneuron, edited by M. A. B. Brazier. Los Angeles: University of California Press, 1969, p. 231–265. (UCLA Forum Med. Sci. No. 11.).
 408. Lundberg, A. Reflex control of stepping. The Nansen Memorial Lecture V. Oslo: Universitetsforlaget, 1969, p. 5–42.
 409. Lundberg, A. The excitatory control of the Ia inhibitory pathway. In: Excitatory Synaptic Mechanisms, edited by P. Andersen and J. K. S. Jansen. Oslo: Universitetsforlaget, 1970, p. 333–340.
 410. Lundberg, A. Function of the ventral spinocerebellar tract. A new hypothesis. Exp. Brain Res. 12: 317–330, 1971.
 411. Lundberg, A. The significance of segmental spinal mechanisms in motor control. In: Symp. Int. Biophys. Congr., 4th, Moscow, 1972. Moscow: Pushchino, 1973, p. 1–23.
 412. Lundberg, A. Control of spinal mechanisms from the brain. In: The Nervous System. The Basic Neurosciences, edited by D. B. Tower. New York: Raven, 1975, vol. I, p. 253–265.
 413. Lundberg, A. Multisensory control of spinal reflex pathways. Progress in Brain Research. Reflex Control of Posture and Movement, edited by R. Granit and O. Pompeiano. Amsterdam: Elsevier, 1979, vol. 50, p. 11–28.
 414. Lundberg, A. Integration in a propriospinal motor centre controlling the forelimb in the cat. In: Integration in the Nervous System, edited by H. Asanuma and V. J. Wilson. Tokyo: Igaku shoin, 1979, p. 47–64.
 415. Lundberg, A., K. Malmgren and E. D. Schomburg. Convergence from Ib, cutaneous and joint afferents in reflex pathways to motoneurones. Brain Res. 87: 81–84, 1975.
 416. Lundberg, A., K. Malmgren and E. D. Schomburg. Characteristics of the excitatory pathway from group II muscle afferents to alpha motoneurones. Brain Res. 88: 538–542, 1975.
 417. Lundberg, A., K. Malmgren and E. D. Schomburg. Cutaneous facilitation of transmission in reflex pathways from Ib afferents to motoneurones. J. Physiol. London 265: 763–780, 1977.
 418. Lundberg, A., K. Malmgren and E. D. Schomburg. Comments on reflex actions evoked by electrical stimulation of group II muscle afferents. Brain Res. 122: 551–555, 1977.
 419. Lundberg, A., K. Malmgren, and E. D. Schomburg. Role of joint afferents in motor control exemplified by effects on reflex pathways from Ib afferents. J. Physiol. London 284: 327–343, 1978.
 420. Lundberg, A., U. Norrsell, and P. Voorhoeve. Pyramidal effects on lumbo‐sacral interneurones activated by somatic afferents. Acta Physiol. Scand. 56: 220–229, 1962.
 421. Lundberg, A., U. Norrsell, and P. Voorhoeve. Effects from the sensorimotor cortex on ascending spinal pathways. Acta Physiol. Scand. 59: 462–473, 1963.
 422. Lundberg, A., and O. Oscarsson. Functional organization of the dorsal spino‐cerebellar tract in the cat. VII. Identification of units by antidromic activation from the cerebellar cortex with recognition of five functional subdivisions. Acta Physiol. Scand. 50: 356–374, 1960.
 423. Lundberg, A., and O. Oscarsson. Three ascending spinal pathways in the dorsal part of the lateral funiculus. Acta Physiol. Scand. 51: 1–16, 1961.
 424. Lundberg, A., and O. Oscarsson. Two ascending spinal pathways in the ventral part of the cord. Acta Physiol. Scand. 54: 270–286, 1962.
 425. Lundberg, A., and P. Voorhoeve. Effects from the pyramidal tract on spinal reflex arcs. Acta Physiol. Scand. 56: 201–219, 1962.
 426. Lundberg, A., and L. Vyklický. Inhibitory interaction between spinal reflexes to primary afferents. Experientia 19: 247–248, 1963.
 427. Lundberg, A., and L. Vyklický. Inhibition of transmission to primary afferents by electrical stimulation of the brain stem. Arch. Ital. Biol. 104: 86–97, 1966.
 428. Lundberg, A., and F. Weight. Functional organization of connexions to the ventral spinocerebellar tract. Exp. Brain Res. 12: 295–316, 1971.
 429. Lundberg, A., and G. Winsbury. Selective adequate activation of large afferents from muscle spindles and Golgi tendon organs. Acta Physiol. Scand. 49: 155–164, 1960.
 430. MacLean, J. B., and H. Leffman. Supraspinal control of Renshaw cells. Exp. Neurol. 18: 94–104, 1967.
 431. Magni, F., and O. Oscarsson. Cerebral control of transmission to the ventral spino‐cerebellar tract. Arch. Ital. Biol. 99: 369–396, 1961.
 432. Magoun, H. W., and R. Rhines. An inhibitory mechanism in the bulbar reticular formation. J. Neurophysiol. 9: 165–171, 1946.
 433. Marsden, C. D., P. A. Merton, and H. B. Morton. Servo action in human voluntary movement. Nature London 238: 140–143, 1972.
 434. Marsden, C. D., P. A. Merton, and H. B. Morton. Is the human stretch reflex cortical rather than spinal. Lancet 1: 759–761, 1973.
 435. Marsden, C. D., P. A. Merton, and H. B. Morton. Stretch reflex and servo action in a variety of human muscles. J. Physiol. London 259: 531–560, 1976.
 436. Massion, J. The mammalian red nucleus. Physiol. Rev. 47: 383–436, 1967.
 437. Matsushita, M. Some aspects of the interneuronal connections in cat's spinal gray matter. J. Comp. Neurol. 136: 57–79, 1969.
 438. Matsushita, M., and Y. Hosoya. Cells of origin of the spinocerebellar tract in the rat, studied with the method of retrograde transport of horseradish peroxidase. Brain Res. 173: 185–200, 1979.
 439. Matthews, B. H. C. Nerve endings in mammalian muscle. J. Physiol. London 78: 1–53, 1933.
 440. Matthews, P. B. C. The differentiation of two types of fusimotor fibre by their effects on the dynamic response of muscle spindle primary endings. Q. J. Exp. Physiol. 47: 324–333, 1962.
 441. Matthews, P. B. C. Muscle spindles and their motor control. Physiol. Rev. 44: 219–288, 1964.
 442. Matthews, P. B. C. Evidence that the secondary as well as the primary endings of the muscle spindles may be responsible for the tonic stretch reflex of the decerebrate cat. J. Physiol. London 204: 365–393, 1969.
 443. Matthews, P. B. C. The origin and functional significance of the stretch reflex. In: Excitatory Synaptic Mechanism, edited by P. Andersen and J. K. S. Jansen. Olso: Universitetsforlaget, 1970, p. 301–315.
 444. Matthews, P. B. C. Mammalian Muscle Receptors and Their Central Actions. London: Arnold, 1972.
 445. McCrea, D. A., C. A. Pratt, and L. M. Jordan. Renshaw cell activity and recurrent effects on motoneurons during fictive locomotion. J. Neurophysiol. 44: 475–488, 1980.
 446. Melvill‐Jones, G., and D. G. D. Watt. Observations on the control of stepping and hopping movements in man. J. Physiol. London 219: 709–727, 1971.
 447. Mendell, L. Positive dorsal root potentials produced by stimulation of small diameter muscle afferents. Brain Res. 18: 375–379, 1970.
 448. Mendell, L. Properties and distribution of peripherally evoked presynaptic hyperpolarization in cat lumbar spinal cord. J. Physiol. London 226: 769–792, 1972.
 449. Mendell, L. M., and P. D. Wall. Presynaptic hyperpolarization: a role for fine afferent fibers. J. Physiol. London 172: 274–294, 1964.
 450. Merton, P. A. Speculations on the servo‐control of movement. In: The Spinal Cord, edited by G. E. W. Wolstenholme. London: Churchill, 1953, p. 247–255.
 451. Miller, S., D. J. Reitsma, and F. G. A. van der Meché. Functional organization of long ascending propriospinal pathways linking lumbo‐sacral and cervical segments in the cat. Brain Res. 62: 169–188, 1973.
 452. Molenaar, I. The distribution of propriospinal neurons projecting to different motoneuronal cell groups in the cat's brachial cord. Brain Res. 158: 203–206, 1978.
 453. Molenaar, I., and H. G. J. M. Kuypers. Cells of origin of propriospinal fibers and of fibers ascending to supraspinal levels. A HRP study in cat and rhesus monkey. Brain Res. 152: 429–450, 1978.
 454. Molenaar, I., A. Rustioni, and H. G. J. M. Kuypers. The location of cells of origin of the fibers in the ventral and the lateral funiculus of the cat's lumbo‐sacral cord. Brain Res. 78: 239–254, 1974.
 455. Morrison, A. R., and O. Pompeiano. Central depolarization of group Ia afferent fibers during desynchronized sleep. Arch. Ital. Biol. 103: 517–537, 1965.
 456. Morrison, A. R., and O. Pompeiano. Depolarization of central terminals of group Ia muscle afferent fibres during desynchronized sleep. Nature London 210: 201–202, 1966.
 457. Muir, R. B., and R. Porter. The effect of a preceding stimulus on temporal facilitation at corticomotoneuronal synapses. J. Physiol. London 228: 749–763, 1973.
 458. Murakami, F., N. Tsukahara, and Y. Fujito. Properties of the synaptic transmission of the newly formed cortico‐rubral synapses after lesion of the nucleus interpositus of the cerebellum. Exp. Brain Res. 30: 245–258, 1977.
 459. Newsom Davis, J., and T. A. Sears. The proprioceptive reflex control of the intercostal muscles during their voluntary activation. J. Physiol. London 209: 711–738, 1970.
 460. Nichols, T. R., and J. C. Houk. Improvement in linearity and regulation of stiffness that results from actions of stretch reflex. J. Neurophysiol. 39: 119–142, 1976.
 461. Nygren, L.‐G., K. Fuxe, G. Jonsson, and L. Olson. Functional regeneration of 5‐hydroxytryptamine nerve terminals in the rat spinal cord following 5,6‐dihydroxytryptamine induced degeneration. Brain Res. 78: 377–394, 1974.
 462. Nygren, L.‐G., and L. Olson. On spinal noradrenaline receptor supersensitivity: Correlation between nerve terminal densities and flexor reflexes various times after intracisternal 6‐hydroxydopamine. Brain Res. 116: 455–470, 1976.
 463. Nygren, L.‐G., and L. Olson. A new major projection from locus coeruleus: the main source of noradrenergic nerve terminals in the ventral and dorsal columns of the spinal cord. Brain Res. 132: 85–93, 1977.
 464. Orlovsky, G. N. Activity of vestibulospinal neurons during locomotion. Brain Res. 46: 85–98, 1972.
 465. Orlovsky, G. N. Activity of rubrospinal neurons during locomotion. Brain Res. 46: 99–112, 1972.
 466. Oscarsson, O. Functional organization of the ventral spinocerebellar tract in the cat. I. Electrophysiological identification of the tract. Acta Physiol. Scand. 38: 145–165, 1956.
 467. Oscarsson, O. Functional organization of the ventral spinocerebellar tract in the cat. II. Connections with muscle, joint, and skin nerve afferents and effects on adequate stimulation of various receptors. Acta Physiol. Scand. 42: (Suppl. 146) 1–107, 1957.
 468. Oscarsson, O. Further observations on ascending spinal tracts activated from muscle, joint, and skin nerves. Arch. Ital. Biol. 96: 199–215, 1958.
 469. Oscarsson, O. Functional organization of the ventral spinocerebellar tract in the cat. III. Supraspinal control of VSCT units of I‐type. Acta Physiol. Scand. 49: 171–183, 1960.
 470. Oscarsson, O. Termination and functional organization of the ventral spino‐olivocerebellar path. J. Physiol. London 196: 453–478, 1968.
 471. Oscarsson, O. Termination and functional organization of the dorsal spino‐olivocerebellar path. J. Physiol. London 200: 129–149, 1969.
 472. Oscarsson, O. Functional organization of spinocerebellar paths. In: Handbook of Sensory Physiology. Somatosensory System, edited by A. Iggo. Berlin: Springer‐Verlag, 1973, vol. II, p. 339–380.
 473. Oscarsson, O. Spatial distribution of climbing and mossy fibre inputs into the cerebellar cortex. In: Afferent and Intrinsic Organization of Laminated Structures in the Brain, edited by O. Creutzfeldt. Exp. Brain Res. Suppl. 1. Berlin: Springer‐Verlag, 1976, p. 36–42.
 474. Oscarsson, O., and B. Sjölund. The ventral spino‐olivocerebellar system in the cat. I. Identification of five paths and their termination in the cerebellar anterior lobe. Exp. Brain Res. 28: 469–486, 1977.
 475. Oscarsson, O., and N. Uddenberg. Identification of a spinocerebellar tract activated from forelimb afferents in the cat. Acta Physiol. Scand. 62: 125–136, 1964.
 476. Oscarsson, O., and N. Uddenberg. Somatotopic termination of spino‐olivocerebellar path. Brain Res. 3: 204–207, 1966.
 477. Patton, H. D., and V. E. Amassian. The pyramidal tract: its excitation and functions. In: Handbook of Physiology. Neurophysiology, edited by J. Field, H. W. Magoun and V. E. Hall. Washington, DC: Am. Physiol. Soc., 1960, sect. 1, vol. II, chapt. 34, p. 837–861.
 478. Pearson, K. G., and J. Duysens. Function of segmental reflexes in the control of stepping in cockroaches and cats. In: Neural Control of Locomotion. Advances in Behavioral Biology, edited by R. M. Herman, S. Grillner, P. G. S. Stein and D. G. Stuart. New York: Plenum, 1976, vol. 18, p. 519–537.
 479. Perl, E. R. Crossed reflex effects evoked by activity in myelinated afferent fibers of muscle. J. Neurophysiol. 21: 101–112, 1958.
 480. Perl, E. R. Effects of muscle stretch on excitability of contralateral motoneurones. J. Physiol. London 145: 193–203, 1959.
 481. Perret, C. Neural control of locomotion in the decorticate cat. In: Neural Control of Locomotion. Advances in Behavioral Biology, edited by R. M. Herman, S. Grillner, P. S. G. Stein and D. G. Stuart. New York: Plenum, 1976, vol. 18, p. 587–615.
 482. Perret, C., and A. Berthoz. Evidence of static and dynamic fusimotor actions on the spindle response to sinusoidal stretch during locomotor activity in the cat. Exp. Brain Res. 18: 178–188, 1973.
 483. Peterson, B. W., N. G. Pitts, K. Fukushima, and R. Mackel. Reticulospinal excitation and inhibition of neck motoneurons. Exp. Brain Res. 32: 471–489, 1978.
 484. Philippson, M. L'autonomie et la centralisation dans le système nerveux des animaux. Trav. Lab. Physiol. Inst. Solvay Bruxelles. 7: 1–208, 1905.
 485. Phillips, C. G., and R. Porter. The pyramidal projection to motoneurones of some muscle groups of the baboon's forelimb. In: Progress in Brain Research. Physiology of Spinal Neurons, edited by J. C. Eccles and J. P. Schadé. Amsterdam: Elsevier, 1964, vol. 12, p. 222–245.
 486. Piercey, M. F., and J. Goldfarb. Discharge patterns of Renshaw cells evoked by volleys in ipsilateral cutaneous and high‐threshold muscle afferents and their relationship to reflexes recorded in ventral roots. J. Neurophysiol. 37: 294–302, 1974.
 487. Pierrot‐Deseilligny, E., R. Katz, and C. Morin. Evidence for Ib inhibition in human subjects. Brain Res. 166: 176–179, 1979.
 488. Pierrot‐Deseilligny, E., C. Morin, R. Katz, and B. Bussel. Influence of posture and voluntary movement on recurrent inhibition in human subjects. Brain Res. 124: 427–436, 1977.
 489. Pi‐Suñer, J., and J. F. Fulton. The influence of proprioceptive system upon the crossed extensor reflex. Am. J. Physiol. 88: 453–467, 1929.
 490. Pogorelaya, N. Kh. Experimental‐morphological study of primary afferent terminals in the base of dorsal horn of the cat spinal cord. Neurophysiology USSR 5: 406–414, 1973.
 491. Pompeiano, O. Mechanisms of sensorimotor integration during sleep. In: Progress in Physiological Psychology, edited by E. Stellar and J. M. Sprague. New York: Academic, 1970, vol. 3, p. 1–179.
 492. Porter, R. Early facilitation at corticomotoneuronal synapses. J. Physiol. London 207: 733–745, 1970.
 493. Pratt, C. A., and L. M. Jordan. Recurrent inhibition of motoneurons in decerebrate cats during controlled treadmill locomotion. J. Neurophysiol. 44: 489–500, 1980.
 494. Prestige, M. C. Initial collaterals of motor axons within the spinal cord of the cat. J. Comp. Neurol. 126: 123–136, 1966.
 495. Prochazka, A., J. A. Stephens, and P. Wand. Muscle spindle discharge in normal and obstructed movements. J. Physiol. London 287: 57–66, 1979.
 496. Prochazka, A., R. A. Westerman, and S. P. Ziccone. Discharges of single hindlimb afferents in the freely moving cat. J. Neurophysiol. 39: 1090–1104, 1976.
 497. Prochazka, A., R. A. Westerman, and S. P. Ziccone. Ia afferent activity during a variety of voluntary movements in the cat. J. Physiol. London 268: 423–448, 1977.
 498. Rademaker, G. G. S. Das Stehen. Berlin: Springer‐Verlag, 1931.
 499. Rapoport, S., A. Susswein, Y. Uchino, and V. J. Wilson. Synaptic actions of individual vestibular neurones on cat neck motoneurones. J. Physiol. London 272: 367–382, 1977.
 500. Rastad, J., E. Jankowska, and J. Westman. Arborization of initial axon collaterals of spinocervical tract cells stained intracellularly with horseradish peroxidase. Brain Res. 135: 1–10, 1977.
 501. Ratliff, F., H. H. Hartline, and W. H. Miller. Spatial and temporal aspects of retinal inhibitory interaction. J. Opt. Soc. Am. 53: 110–120, 1963.
 502. Renshaw, B. Influence of discharge of motoneurons upon excitation of neighboring motoneurons. J. Neurophysiol. 4: 167–183, 1941.
 503. Renshaw, B. Central effects of centripetal impulses in axons of spinal ventral roots. J. Neurophysiol. 9: 191–204, 1946.
 504. Roberts, T. D. M. Reflex interaction of synergic extensor muscles of the cat hind limb. J. Physiol. London 117: 5P–6P, 1952.
 505. Romanes, G. J. The motor cell columns of the lumbo‐sacral spinal cord of the cat. J. Comp. Neurol. 94: 313–363, 1951.
 506. Ross, H.‐G., S. Cleveland, and J. Haase. Quantitative relation of Renshaw cell discharges to monosynaptic reflex height. Pfluegers Arch. 332: 73–79, 1972.
 507. Ross, H.‐G., S. Cleveland, and J. Haase. Contribution of single motoneurons to Renshaw cell activity. Neurosci. Lett. 1: 105–108, 1975.
 508. Ross, H.‐G., S. Cleveland, and J. Haase. Quantitative relation between discharge frequencies of a Renshaw cell and an intracellularly depolarized motoneuron. Neurosci. Lett. 3: 129–132, 1976.
 509. Rossignol, S., and L. Gauthier. Reversal of contralateral limb reflexes. Proc. Int. Un. Physiol. Sci. 13: 639, 1977.
 510. Rudomin, P., R. Nuñez, J. Madrid, and R. E. Burke. Primary afferent hyperpolarization and presynaptic facilitation of Ia afferent terminals induced by large cutaneous fibers. J. Neurophysiol. 37: 413–429, 1974.
 511. Ryall, R. W. Renshaw cell mediated inhibition of Renshaw cells: Patterns of excitation and inhibition from impulses in motor axon collaterals. J. Neurophysiol. 33: 257–270, 1970.
 512. Ryall, R. W. Excitatory convergence on Renshaw cells. J. Physiol. London 226: 69P–70P, 1972.
 513. Ryall, R. W., and M. F. Piercey. Excitation and inhibition of Renshaw cells by impulses in peripheral afferent nerve fibers. J. Neurophysiol. 34: 242–251, 1971.
 514. Ryall, R. W., M. F. Piercey, and C. Polosa. Intersegmental and intrasegmental distribution of mutual inhibition of Renshaw cells. J. Neurophysiol. 34: 700–707, 1971.
 515. Ryall, R. W., M. F. Piercey, C. Polosa, and J. Goldfarb. Excitation of Renshaw cells in relation to orthodromic and antidromic excitation of motoneurons. J. Neurophysiol. 35: 137–148, 1972.
 516. Rymer, W. Z., J. C. Houk, and P. E. Crago. Mechanisms of the clasp‐knife reflex studied in an animal model. Exp. Brain Res. 37: 93–113, 1979.
 517. Scheibel, M. E., and A. B. Scheibel. Spinal motoneurons, interneurons, and Renshaw cells. A Golgi study. Arch. Ital. Biol. 104: 328–353, 1966.
 518. Scheibel, M. E., and A. B. Scheibel. Inhibition and the Renshaw cell. A structural critique. Brain Behav. Evol. 4: 53–93, 1971.
 519. Schmidt, R. F. Presynaptic inhibition in the vertebrate central nervous system. Ergeb. Physiol. Biol. Exp. Pharmacol. 63: 20–101, 1971.
 520. Schmidt, R. F., J. Senges, and M. Zimmermann. Presynaptic depolarization of cutaneous mechanoreceptor afferents after mechanical skin stimulation. Exp. Brain Res. 3: 234–247, 1967.
 521. Schmidt, R. F., and W. D. Willis. Intracellular recording from motoneurons of the cervical spinal cord of the cat. J. Neurophysiol. 26: 28–43, 1963.
 522. Schoen, J. H. R. Comparative aspects of the descending fibre systems in the spinal cord. In: Progress in Brain Research. Organization of the Spinal Cord, edited by J. C. Eccles and J. P. Schadé. Amsterdam: Elsevier, 1964, vol. 11, p. 203–222.
 523. Schomburg, E. D. Fusimotorische Förderung und Hemmung bei Pinna‐Reizung in Abhängigkeit vom zentralnervösen Zustand. Pfluegers Arch. 304: 164–182, 1968.
 524. Schomburg, E. D. Supraspinal interactions of anesthetic, neuroleptic and analeptic drugs on the fusimotor effects of pinna stimulation in the cat. Exp. Brain Res. 10: 182–196, 1970.
 525. Schomburg, E. D., and H. B. Behrends. Phasic control of the transmission in the excitatory and inhibitory reflex pathways from cutaneous afferents to α‐motoneurones during fictive locomotion in cats. Neurosci. Lett. 8: 277–282, 1978.
 526. Schomburg, E. D., and H. B. Behrends. The possibility of phase‐dependent monosynaptic and polysynaptic Ia excitation to homonymous motoneurones during fictive locomotion. Brain Res. 143: 533–537, 1978.
 527. Schomburg, E. D., H. B. Behrends, and H. Steffens. Alteration of transmission in segmental pathways from flexor reflex afferents (FRA) to α‐motoneurones during spinal locomotor activity. Neurosci. Lett.: (Suppl. 1), S103, 1978.
 528. Schomburg, E. D., and H. D. Henatsch. Aktivierung lumbaler Extensor‐Fusimotoneurone durch mechanische und elektrische Pinna‐Reizung an decerebrierten Katzen. Pfluegers Arch. 304: 152–163, 1968.
 529. Schomburg, E. D., H.‐M. Meinck, and J. Haustein. A fast propriospinal inhibitory pathway from forelimb afferents to motoneurones of hindlimb flexor digitorum longus. Neurosci. Lett. 1: 311–314, 1975.
 530. Schomburg, E. D., H.‐M. Meinck, J. Haustein, and J. Roesler. Functional organization of the spinal reflex pathways from forelimb afferents to hindlimb motoneurones in the cat. Brain Res. 139: 21–33, 1978.
 531. Schomburg, E. D., J. Roesler, and P. Kenins. On the function of the fast long spinal inhibitory pathway from forelimb afferents to flexor digitorum longus motoneurones in cats and dogs. Neurosci. Lett. 7: 55–59, 1978.
 532. Schomburg, E. D., J. Roesler, and H.‐M. Meinck. Phase‐dependent transmission in the excitatory propriospinal reflex pathway from forelimb afferents to lumbar motoneurones during fictive locomotion. Neurosci. Lett. 4: 249–252, 1977.
 533. Sears, T. A. Investigations on respiratory motoneurones of the thoracic spinal cord. In: Progress in Brain Research. Physiology of Spinal Neurons, edited by J. C. Eccles and J. P. Schadé. Amsterdam: Elsevier, 1964, vol. 12, p. 259–273.
 534. Sears, T. A. Efferent discharges in alpha and fusimotor fibres of intercostal nerves of the cat. J. Physiol. London 174: 295–315, 1964.
 535. Sears, T. A., and D. Stagg. Short‐term synchronization of intercostal motoneurone activity. J. Physiol. London 263: 357–381, 1976.
 536. Severin, F. V. On the role of γ‐motor system for extensor α‐motoneurone activation during controlled locomotion [in Russian]. Biofizika 15: 1096–1102, 1970.
 537. Severin, F. V., G. N. Orlovsky, and M. L. Shik. Work of the muscle receptors during controlled locomotion. Biophysics 12: 575–586, 1967. [Engl. transl. 12: 502–511, 1967.].
 538. Severin, F. V., G. N. Orlovsky, and M. L. Shik. Recurrent inhibitory effects on single motoneurones during an electrically evoked locomotion. Bull Exp. Biol. Med. USSR 66: 3–9, 1968.
 539. Shapovalov, A. I. Excitation and inhibition of spinal neurones during supraspinal stimulation. In: Muscular Afferents and Motor Control. Nobel Symp. I, edited by R. Granit. Stockholm: Almqvist & Wiksell, 1966, p. 331–348.
 540. Shapovalov, A. I. Extrapyramidal monosynaptic and disynaptic control of mammalian alpha motoneurons. Brain Res. 40: 105–115, 1972.
 541. Shapovalov, A. I. Evolution of neuronal systems of supraspinal motor control. Neurophysiology USSR 4: 453–470, 1972.
 542. Shapovalov, A. I. Extrapyramidal control of primate motoneurons. In: New Developments in Electromyography and Clinical Neurophysiology, edited by J. E. Desmedt. Basel: Karger, 1973, vol. 3, p. 145–158.
 543. Shapovalov, A. I. Neuronal organization and synaptic mechanisms of supraspinal motor control in vertebrates. Rev. Physiol. Biochem. Pharmacol. 72: 1–54, 1975.
 544. Shapovalov, A. I., A. A. Grantyn, and G. G. Kurchavyi. Short‐latency reticulospinal projections to alpha‐motoneurons. Bull. Exp. Biol. Med. USSR 64: 3–15, 1967.
 545. Shapovalov, A. I., O. A. Karamjan, G. G. Kurchavyi, and Z. A. Repina. Synaptic actions evoked from the red nucleus on the spinal alpha‐motoneurons in the rhesus monkey. Brain Res. 32: 325–348, 1971.
 546. Shapovalov, A. I., and V. M. Kozhanov. Disynaptic brain‐stem‐propriospinal projections to mammalian motoneurones. Neuroscience 3: 105–108, 1978.
 547. Shapovalov, A. I., G. G. Kurchavyi, O. A. Karamjan, and Z. A. Repina. Extrapyramidal pathways with monosynaptic effects upon primate α‐motoneurons. Experientia 27: 522–524, 1971.
 548. Shapovalov, A. I., G. G. Kurchavyi, and M. P. Strogonova. Synaptic mechanisms of vestibulo‐spinal influences on alpha‐motoneurons. Sechenov J. Physiol. USSR 52: 1401–1409, 1966.
 549. Sherrington, C. S. Decerebrate rigidity, and reflex coordination of movements. J. Physiol. London 22: 319–332, 1898.
 550. Sherrington, C. S. On innervation of antagonistic muscles. Sixth Note. Proc. R. Soc. London Ser. B 66: 66–67, 1900.
 551. Sherrington, C. The Integrative Action of the Nervous System. New Haven: Yale Univ. Press, 1906.
 552. Sherrington, C. S. Observations on the scratch‐reflex in the spinal dog. J. Physiol. London 34: 1–50, 1906.
 553. Sherrington, C. S. On reciprocal innervation of antagonistic muscles. Twelfth note. Proprioceptive reflexes. Proc. R. Soc. London Ser. B 80: 552–564, 1908.
 554. Sherrington, C. S. On plastic tonus and proprioceptive reflexes. Q. J. Exp. Physiol. 2: 109–156, 1909.
 555. Sherrington, C. S. Flexion‐reflex of the limb, crossed extension reflex, and reflex stepping and standing. J. Physiol. London 40: 28–121, 1910.
 556. Sherrington, C. S. Notes on the scratch‐reflex of the cat. Q. J. Exp. Physiol. 3: 213–220, 1910.
 557. Sherrington, C. S., and E. E. Laslett. Observations on some spinal reflexes and the interconnection of spinal segments. J. Physiol. London 29: 58–96, 1903.
 558. Sherrington, C. S., and S. C. M. Sowton. Observations on reflex responses to single break‐shocks. J. Physiol. London 49: 331–348, 1915.
 559. Shibuya, T., and E. G. Anderson. The influence of chronic cord transection on the effects of 5‐hydroxytryptophan, 1‐tryptophan and pargyline on spinal neuronal activity. J. Pharmacol. Exp. Ther. 164: 185–190, 1968.
 560. Shik, M. L., F. V. Severin, and G. N. Orlovsky. Control of walking and running by means of electrical stimulation of the mid‐brain. Biofizika 11: 659–666, 1966. [Engl. transl. 11: 756–765, 1966.].
 561. Shimazu, H., T. Hongo, and K. Kubota. Two types of central influences on gamma motor system. J. Neurophysiol. 25: 309–323, 1962.
 562. Sjölund, B. The ventral spino‐olivocerebellar system in the cat. V. Supraspinal control of spinal transmission. Exp. Brain Res. 33: 509–522, 1978.
 563. Sjöström, A., and P. Zangger. Muscle spindle control during locomotor movements generated by the deafferented spinal cord. Acta Physiol. Scand. 97: 281–291, 1976.
 564. Skoglund, S. Anatomical and physiological studies of knee joint innervation in the cat. Acta Physiol. Scand. 36: (Suppl. 124) 1–101, 1956.
 565. Snow, P. J., P. K. Rose, and A. G. Brown. Tracing axons and axon collaterals of spinal neurons using intracellular injection of horseradish peroxidase. Science 191: 312–313, 1976.
 566. Stauffer, E. K., D. G. D. Watt, A. Taylor, R. M. Reinking, and D. G. Stuart. Analysis of muscle receptor connections by spike‐triggered averaging. 2. Spindle group II afferents. J. Neurophysiol. 39: 1393–1402, 1976.
 567. Stephens, J. A., R. M. Reinking, and D. G. Stuart. Tendon organs of cat medial gastrocnemius: responses to active and passive forces as a function of muscle length. J. Neurophysiol. 38: 1217–1231, 1975.
 568. Sterling, P., and H. G. J. M. Kuypers. Anatomical organization of the brachial spinal cord of the cat. II. The motoneuron plexus. Brain Res. 4: 16–32, 1967.
 569. Stewart, D. H., J. B. Preston, and D. G. Whitlock. Spinal pathways mediating motor cortex evoked excitability changes in segmental motoneurons in pyramidal cats. J. Neurophysiol. 31: 928–937, 1968.
 570. Stuart, D. G., C. G. Mosher, R. L. Gerlach, and R. M. Reinking. Selective activation of Ia afferents by transient muscle stretch. Exp. Brain Res. 10: 477–487, 1970.
 571. Sumner, A. J. Properties of Ia and Ib afferent fibres serving stretch receptors of the cat's medial gastrocnemius muscle. Proc. Univ. Otago Med. School 39: 3–5, 1961.
 572. Szentágothai, J. Short propriospinal neurons and intrinsic connections of the spinal gray matter. Acta Morphol. Acad. Sci. Hung. 1: 81–94, 1951.
 573. Szentágothai, J. Propriospinal pathways and their synapses. In: Progress in Brain Research. Organization of the Spinal Cord, edited by J. C. Eccles and J. P. Schadé. Amsterdam: Elsevier, 1964, vol. 11, p. 155–177.
 574. Taber, C. p‐Chlorophenylalanine blockade of the effect of 5‐hydroxytryptophan on spinal synaptic activity (Abstract). Federation Proc. 30: 317, 1971.
 575. Tamarova, Z. A., A. I. Shapovalov, O. A. Karamyan, and G. G. Kurchavyi. Cortico‐pyramidal and cortico‐extrapyramidal synaptic effects on the monkey lumbar motoneurons. Neirofiziologiia 4: 587–596, 1972.
 576. Tanaka, R. Reciprocal Ia inhibition during voluntary movements in man. Exp. Brain Res. 21: 529–540, 1974.
 577. Tanaka, R. Reciprocal Ia inhibition and voluntary movements in man. In: Progress in Brain Research. Understanding the Stretch Reflex, edited by S. Homma. Amsterdam: Elsevier, 1976, vol. 44, p. 291–302.
 578. Taylor, A., and F. W. J. Cody. Jaw muscle spindle activity in the cat during normal movements of eating and drinking. Brain Res. 71: 523–530, 1974.
 579. Taylor, A., and M. R. Davey. Behaviour of jaw muscle stretch receptors during active and passive movements in the cat. Nature London 220: 301–302, 1968.
 580. Thomas, R. C., and V. J. Wilson. Recurrent interactions between motoneurons of known location in the cervical cord of the cat. J. Neurophysiol. 30: 661–674, 1967.
 581. Thorndike, E. L. Animal intelligence: an experimental study of the associative processes in animals. Psychol. Monogr. 2: 1–109, 1898. No. 8.
 582. Trott, J. R. The effect of low amplitude muscle vibration on the discharge of fusimotor neurones in the decerebrate cat. J. Physiol. London 255: 635–649, 1976.
 583. Tsukahara, N., and C. Ohye. Polysynaptic activation of extensor motoneurones from group Ia fibres in the cat spinal cord. Experientia 20: 628–629, 1964.
 584. Vallbo, Å. B. Slowly adapting muscle receptors in man. Acta Physiol. Scand. 78: 315–333, 1970.
 585. Vallbo, Å. B. Discharge patterns in human muscle spindle afferents during isometric voluntary contractions. Acta Physiol. Scand. 80: 552–566, 1970.
 586. Vallbo, Å. B. Muscle spindle response at the onset of isometric voluntary contractions in man. Time difference between fusimotor and skeletomotor effects. J. Physiol. London 218: 405–431, 1971.
 587. Vallbo, Å. B. Muscle spindle afferent discharge from resting and contracting muscles in normal human subjects. In: New Developments in Electromyography and Clinical Neurophysiology, edited by J. E. Desmedt. Basel: Karger, 1973, vol. 3, p. 251–262.
 588. Vallbo, Å. B. The significance of intramuscular receptors in load compensation during voluntary contractions in man. In: Control of Posture and Locomotion. Advances in Behavioral Biology, edited by R. B. Stein, K. B. Pearson, R. S. Smith and J. B. Redford. New York: Plenum, 1973, vol. 7, p. 211–226.
 589. Vallbo, Å. B. Human muscle spindle discharge during isometric voluntary contractions. Amplitude relations between spindle frequency and torque. Acta Physiol. Scand. 90: 319–336, 1974.
 590. Vallbo, Å. B., K.‐E. Hagbarth, H. E. Torebjörk, and B. G Wallin. Somatosensory, proprioceptive and sympathetic activity in human peripheral nerves. Physiol. Rev. 59: 919–957, 1979.
 591. van Keulen, L. C. M. Identification de cellules de Renshaw par l'injection intracellulaire, de la substance fluorescente “procion yellow.” J. Physiol. Paris 63: 131A, 1971.
 592. van Keulen, L. C. M. Axon trajectories of Renshaw cells in the lumbar spinal cord of the cat, as reconstructed after intracellular staining with horseradish peroxidase. Brain Res. 167: 157–162, 1979.
 593. van Keulen, L. Relations between individual motoneurones and individual Renshaw cells. Neurosci. Lett.: (Suppl. 3), S 313, 1979.
 594. Vasilenko, D. A. Propriospinal pathways in the ventral funicles of the cat spinal cord: their effects on lumbosacral motoneurones. Brain Res. 93: 502–506, 1975.
 595. Vasilenko, D. A., and P. G. Kostyuk. Functional properties of interneurons activated monosynaptically by the pyramidal tract. Zh. Vyssh. Nervn. Deyat. im I. P. Pavlova 16: 1046, 1966. [Neurosci. Transl. 1: 66–72, 1967/68.].
 596. Vasilenko, D. A., and A. I. Kostyukov. Brain stem and primary afferent projections to the ventromedial group of propriospinal neurones in the cat. Brain Res. 117: 141–146, 1976.
 597. Vasilenko, D. A., and A. I. Kostyukov. Transmission of reticulofugal activity via the ventromedial group of propriospinal neurons in cat. Neurophysiology USSR 9: 205–209, 1977.
 598. Vasilenko, D. A., A. I. Kostyukov, and A. I. Pilyavsky. Cortico‐ and rubrofugal activation of propriospinal interneurons sending axons into the dorsolateral funiculus of the cat spinal cord. Neurophysiology USSR 4: 489–500, 1972.
 599. Vedel, J. P., and J. Mouillac‐Baudevin. Contrôle de l'activité des fibres fusimotrices dynamiques et statiques par la formation réticulée mésencéphalique chez le chat. Exp. Brain Res. 9: 307–324, 1969.
 600. Vedel, J. P., and J. Mouillac‐Baudevin. Contrôle pyramidal de l'activité des fibres fusimotrices dynamiques et statiques chez le chat. Exp. Brain Res. 10: 39–63, 1970.
 601. Viala, D., and P. Buser. Modalités d'obtention de rythmes locomoteurs chez le lapin spinal par traitements pharmacologiques (DOPA, 5‐HTP, D‐amphétamine). Brain Res. 35: 151–165, 1971.
 602. Viala, G., D. Orsal, and P. Buser. Cutaneous fiber groups involved in the inhibition of fictive locomotion in the rabbit. Exp. Brain Res. 33: 257–267, 1978.
 603. Voorhoeve, P. E., and R. W. van Kanten. Reflex behaviour of fusimotor neurones of the cat upon electrical stimulation of various afferent fibers. Acta Physiol. Pharmacol. Neerl. 10: 391–407, 1962.
 604. Wall, P. D. Excitability changes in afferent fibre terminations and their relation to slow potentials. J. Physiol. London 142: 1–21, 1958.
 605. Wall, P. D. Presynaptic control of impulses at the first central synapse in the cutaneous pathway. In: Progress in Brain Research. Physiology of Spinal Neurons, edited by J. C. Eccles and J. P. Schadé. Amsterdam: Elsevier, 1964, vol. 12, p. 92–118.
 606. Westbury, D. R. A study of stretch and vibration reflexes of the cat by intracellular recording from motoneurones. J. Physiol. London 226: 37–56, 1972.
 607. Whitehorn, D., and P. R. Burgess. Changes in polarization of central branches of myelinated mechanoreceptor and nociceptor fibers during noxious and innocuous stimulation of the skin. J. Neurophysiol. 36: 226–237, 1973.
 608. Willis, W. D. The case for the Renshaw cell. Brain Behav. Evol. 4: 5–52, 1971.
 609. Willis, W. D., G. W. Tate, R. D. Ashworth, and J. C. Willis. Monosynaptic excitation of motoneurons of individual forelimb muscles. J. Neurophysiol. 29: 410–424, 1966.
 610. Willis, W. D., and J. C. Willis. Properties of interneurons in the ventral spinal cord. Arch. Ital. Biol. 104: 354–386, 1966.
 611. Wilson, V. J. Regulation and function of Renshaw cell discharge. In: Muscular Afferents and Motor Control, Nobel Symposium I, edited by R. Granit. Stockholm: Almqvist & Wiksell 1966, p. 317–329.
 612. Wilson, V. J., and M. Kato. Excitation of extensor motoneurons by group II afferent fibers in ipsilateral muscle nerves. J. Neurophysiol. 28: 545–554, 1965.
 613. Wilson, V. J., and W. H. Talbot. Integration at an inhibitory interneurone: Inhibition of Renshaw cells. Nature London 200: 1325–1327, 1963.
 614. Wilson, V. J., W. H. Talbot, and F. P. J. Diecke. Distribution of recurrent facilitation and inhibition in cat spinal cord. J. Neurophysiol. 23: 144–153, 1960.
 615. Wilson, V. J., W. H. Talbot, and M. Kato. Inhibitory convergence upon Renshaw cells. J. Neurophysiol. 27: 1063–1079, 1964.
 616. Wilson, V. J., Y. Uchino, R. A. Maunz, A. Susswein, and K. Fukushima. Properties and connections of cat fastigiospinal neurons. Exp. Brain. Res. 32: 1–17, 1978.
 617. Wilson, V. J., and M. Yoshida. Comparison of effects of stimulation of Deiters' nucleus and medial longitudinal fasciculus on neck, forelimb, and hindlimb motoneurons. J. Neurophysiol. 32: 743–758, 1969.
 618. Wilson, V. J., and M. Yoshida. Monosynaptic inhibition of neck motoneurons by the medial vestibular nucleus. Exp. Brain Res. 9: 365–380, 1969.
 619. Wilson, V. J., M. Yoshida, and R. H. Schor. Supraspinal monosynaptic excitation and inhibition of thoracic back motoneurons. Exp. Brain Res. 11: 282–295, 1970.
 620. Windhorst, U., D. Adam, and G. F. Inbar. The effects of recurrent inhibitory feedback in shaping discharge patterns of motoneurones excited by phasic muscle stretches. Biol. Cybernetics 29: 221–227, 1978.
 621. Windhorst, U., M. Ptok, J. Meyer‐Lohmann, and J. Schmidt. Effects of conditioning stimulation of the contralateral n. ruber on antidromic Renshaw cell responses and monosynaptic reflexes. Pfluegers Arch. 373: R 70, 1978.
 622. Wirth, F. P., J. L. O'Leary, J. M. Smith, and A. B. Jenny. Monosynaptic corticospinal‐motoneuron path in the raccoon. Brain Res. 77: 344–348, 1974.
 623. Yokota, T., and P. E. Voorhoeve. Pyramidal control of fusimotor neurons supplying extensor muscles in the cat's forelimb. Exp. Brain Res. 9: 96–115, 1969.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Fausto Baldissera, Hans Hultborn, Michael Illert. Integration in Spinal Neuronal Systems. Compr Physiol 2011, Supplement 2: Handbook of Physiology, The Nervous System, Motor Control: 509-595. First published in print 1981. doi: 10.1002/cphy.cp010212