Open access peer-reviewed chapter

From Motor Neuron Specification to Function: Filling in the Gaps

Written By

Mudassar Nazar Khan and Till Marquardt

Submitted: 28 September 2023 Reviewed: 12 February 2024 Published: 22 May 2024

DOI: 10.5772/intechopen.114298

From the Edited Volume

Motor Neurons - New Insights

Edited by Natalia Szejko and Kamila Saramak

Chapter metrics overview

22 Chapter Downloads

View Full Metrics

Abstract

Motor neurons operate at the interface between nervous system and movement apparatus and play several roles in movement generation. During development, motor neurons emerge from progenitor cells in the ventral neural tube and eventually settle into stereotypic position that predict the identity of their target muscles. The specification of these ‘positional’ identities has been studied in detail and involves a coordinate grid of intersecting extrinsic signals that result in the activation of unique combinations of transcription factors acting as cell-autonomous determinants. Eventually, motor neurons diversify into ‘functional’ (e.g., fast/intermediate/slow alpha, beta, and gamma) subtypes essential for proper movement execution, a process linked to the acquisition of unique sets of functional properties. Recent progress has provided insights into the molecular composition and specification of motor neuron functional identities, but little is known about their relationship to the mechanisms underlying the specification of positional identities. In this chapter, we attempt to provide a framework for consolidating both aspects of motor neuron diversification, in addition to outlining the gaps in our knowledge to guide future research directions aiming at understanding the events on a motor neuron’s journey from specification to function.

Keywords

  • spinal motor neurons
  • motor neuron functional specification
  • motor neuron positional identities
  • motor neuron development
  • neurogenesis
  • movement control
  • neuronal development
  • gamma
  • beta
  • alpha motor neurons
  • fast
  • slow motor neurons

1. Introduction

To paraphrase Sherrington, motor neurons represent the final common pathway in the generation of behaviors by linking the nervous system with the movement apparatus [1]. This role is reflected by two levels of organization, one spatial and one functional, that together allow motor neurons to serve as the interface through which the brain can engage with and act upon the external world. On the one hand, motor neurons are somatotopically organized, with the position of motor neuron somas in the spinal cord predicting the specific muscle it controls (Figure 1) [2]. These ‘positional’ identities of motor neurons are specified early in development, prior to the establishment of neuromuscular connections [2]. On the other hand, motor neuron diversity is defined by different roles in movement generation and by the different muscle fiber types that are innervated (Figure 1) [3]. In tetrapod vertebrates, a skeletal muscle typically contains a mixture of different muscle fiber types and is supplied by a range of motor neuron types involved in different aspects of movement generation or control [2, 4, 5, 6]. To avoid confusion, we will refer to the first instance of motor neuron diversity as ‘positional’ diversity and the second as ‘functional’ diversity. While the mechanisms promoting positional or functional diversity may operate independently form each other, an individual motor neuron can eventually be assigned both by a positional and a functional identity (Figure 1). Several excellent reviews have appeared in recent years covering either positional or functional motor neuron diversification [2, 3, 7]. To avoid redundancy, we will in this chapter just briefly touch upon the fundamentals of both spatial and functional motor neuron diversity and mostly focus on what is known—or rather what is not—about the link between the mechanisms that underlie the development of both types of motor neuron diversity necessary for proper movement execution and control.

Figure 1.

Schematic summarizing the different levels of motor neuron specification. All spinal motor neurons originate from neuron progenitor cells in the ventral neural tube. Motor neuron positional identities (columns, divisions innervating certain muscle groups and motor pools innervating individual muscles) are acquired after cell-cycle exit. A typical motor pool of tetrapod vertebrates contains a mixture of different ‘functional’ motor neuron types: Alpha motor neurons (fast/f-MNs, intermediate/i-MNs and slow/s-MNs) as well as beta and gamma motor neurons, which all possess different biophysical properties such as firing rates (such as high firing rates for f-MNs and gamma motor neurons) and roles in movement execution and control. It remains unclear, when exactly motor neuron functional diversity is generated and to what extent, if at all, this is coordinated with the specification of motor pool identities.

Advertisement

2. Motor neuron positional identities

2.1 Initial specification of motor neuron identity

Positionally distinct classes of motor neurons emerge in the developing spinal cord whose fates are specified under the influence of intersecting gradients of signaling molecules called morphogens along the dorsoventral (d-v) and rostrocaudal (r-c) axes of the neural tube [8, 9]. Morphogens are secreted from several embryonic structures to the neural tube: Sonic hedgehog (Shh) from the notochord and floor plate and bone morphogenetic protein (BMP) and wingless (Wnt) from the dorsal roof plate and surface ectoderm, retinoic acid (RA) from the paraxial mesoderm and fibroblast growth factors (FGFs) from the caudal mesoderm and tail bud, that together contribute to the specification of molecular identities of neural progenitor cells or domains along the d-v and r-c axes of the neural tube [8, 9]. Shh drives the specification of motor neuron progenitor (pMN) and interneuron populations (V0, V1, V2, V3) in the ventral spinal cord, while RA and FGFs also contribute to motor neuron progenitor identities [10, 11, 12]. The graded expression of Shh in the d-v axis regulates Gli (Glioma-associated oncogene) family transcription factors activity, which establishes the neural progenitor domains via the expression of Class I and II homeodomain transcription factors (TFs) [9].

Progenitor domain boundaries develop and are sustained through the cross-repressive actions of class I (Pax7, Pax6, Irx3, Dbx1, Dbx2) and Class II TFs (Nkx6.1 and Nkx2.2) [13]. The co-expression of Pax6 and Nkx6.1 activates the expression of basic helix-basic-helix (bHLH) protein Olig2, which together mark the motor neuron progenitor domain [14, 15, 16]. The expression of Olig2 and its repressive activities (specifically, the repression of a repressor that normally represses Ngn2 expression) leads to the indirect upregulation of Ngn2. This results in the expression of post-mitotic motor neuron genes of Lhx3/Isl1 and motor-neuron specific gene Mnx1 (Hb9) [13, 17, 18]. Moreover, the expression levels of Olig2/Ngn2 must be balanced to ensure timely motor neuron specific gene expression, since high expression of Olig2 sustains the pMN state, while high levels of Ngn2 activates the conversion of pMN to post-mitotic motor neurons [19]. Post-mitotic neuron classes, like the progenitor classes, are specified by the combinatorial expression of transcription factors, and in addition, develop specific patterns of connectivity, a neurotransmitter system and electrophysiological properties [8, 9]. Newly born post-mitotic motor neurons express a set of TFs like Lhx3, Isl1 and Mnx1 and send axons peripherally to muscles, use acetylcholine and glutamate as neurotransmitters [20, 21]. This set of early post-mitotic TFs bind to specific enhancers to specify and maintain motor neuron identity [22, 23].

2.2 Specification of motor neuron positional identities: columns, divisions and pools

A second level of motor neuron organization within the spinal cord is the clustering of somas into columns within the r-c axis (Figure 1). These positional identities are patterned by reciprocally graded concentrations of fibroblast growth factors (FGFs) and retinoic acid (RA) that regulate the temporal and spatial expression of homeobox (Hox) family transcription factors [24, 25, 26]. Hox family transcription factors are sequentially arranged on chromosomes in four clusters (HoxA, HoxB, HoxC, HoxD) that encode 39 genes that are expressed in the spinal cord and hindbrain during both progenitor and post-mitotic phases of motor neuron differentiation [27, 28]. RA regulates the expression of Hox4-Hox6 genes at the cervical/brachial levels of the spinal cord [7, 26]. While FGF induces Hox4-Hox10 in cervical/brachial, thoracic and lumbar levels of the spinal cord and Gdf11/FGF8 induce Hox10 in the lumbar spinal cord [7, 26].

Hox gene expression pattern in post-mitotic motor neurons determines their columnar subtype identity. For example, in the phrenic motor column (PMC), motor neurons at the cervical level of the spinal cord express Hox5 [29]. Loss of Hox5 in motor neurons results in failure of dendritic arborization in the diaphragm muscle and neuronal death leading to respiratory failure and perinatal death in mice [29]. The lumbar motor column (LMC) motor neurons of the brachial and lumbar levels express Hox6 and Hox10, respectively [25, 30, 31, 32, 33, 34]. The specification of LMC depends on the expression of Hox genes and their regulation of Foxp1 expression pattern [30, 32]. The preganglionic motor column (PGC) motor neuron identities depend on Hoxc9 expression [35]. While, motor neurons that innervate axial muscles do not depend on Hox gene programs. For example, the medial motor column (MMC) columnar subtype motor neurons identity depends on the Wnt genes and are marked by the expression of Prdm family transcription factor Mecom [7, 36, 37]. Still, the programs involved in the specification of other motor columns, like hypaxial motor column (HMC) which innervates axial muscles are unknown.

A third level of motor neuron organization is known as divisional identity which is defined by the motor axons and the muscles they target. The genetic programs that are involved in the specification of muscle targets have been studied extensively in the lateral motor column (LMC) motor neurons. While the generation of LMC motor neurons depends on Hox genes and the expression pattern of transcription factors they regulate, the maturation of LMC neurons is dependent on both the limb-derived cues and molecular programs intrinsic to motor neurons. Hox genes promote high Foxp1 expression levels, which is required for the expression of Raldh2, an enzyme that catalyzes retinoic acid (RA) synthesis. Localized synthesis of RA in motor neurons at the brachial and lumbar spinal cord levels establish LMC divisional identities [38, 39, 40]. The LMC motor neurons express specific Lim homeodomain (HD) proteins and their axons target specific muscles groups of the limbs, which confers their divisional identity: the lateral division (LMCl) motor neurons express Lhx1 and send axons to the muscles within the dorsal compartment, while the medial division (LMCm) motor neurons express Isl1 and send axons to the muscles within the dorsal compartment [41]. To ensure LMCl identity, Lhx1 expression is maintained by the repressive interaction between Lhx1 and Isl1, the protein signaling controlled by the sources of RA from Raldh2+ motor neuron and the paraxial mesoderm [7, 38, 39, 40]. Moreover, the routes that LMCl motor neuron axons select within the limbs is dependent upon Lhx1 expression. Lhx1 expression within LMCl regulates the expression of axonal guidance receptor Eph4 (which repels ventral axons that express ephrin), leading to the dorsal projection of axons [7, 42, 43]. Moreover, ventrally projecting LMCm axons use similar signaling strategies: Lim HD proteins regulate the expression of axonal guidance molecules like ephrin and Eph receptors that ultimately determine axonal trajectories [7, 44].

A fourth level of motor neuron organization observed in the spinal cord are motor pools (Figure 1). Motor neurons within a specific motor pool cluster in a specific position in the spinal cord, innervate a single muscle, have specific type (alpha/gamma) ratios and show motor pool-specific molecular marker expression, morphology, central and peripheral connectivity, and electrophysiological patterns. These features are determined by Hox genes, in part, and by molecular cues from peripheral targets. Early on in development, motor pool diversity is regulated by Hox genes and their regulation of transcription factor expression levels. Combinatorial expression of Hoxc8 and Hoxc6 determines the expression of several proteins, including ETS domain transcription factor Pea3, the Pou domain protein Scip (also known as Pou3f1), and Nkx6.1, which ultimately characterize LMC motor pool identities [7]. Thus, the loss of Hoxc8 and Hoxc6 in mice leads to the decrease in Pea3-expressing motor neurons and decrease in axonal arborization of targeted muscles [31, 45, 46]. Moreover, specific muscle innervation in mice is disrupted in motor pools that no longer express Nkx6.1, a downstream target of Hox signaling [47], while Foxp1 mutant mice show reduced expression of motor pool markers Pea3, Scip and Nkx6.1 [32]. Motor pool identities are also regulated by the molecular cues from peripheral muscles. For example, neurotrophic factors like glial-derived neurotrophic factor (GDNF), regulate the expression of Pea3 and thus, determine motor pool soma position and motor pool targeted muscle innervation [48].

Advertisement

3. Motor neuron functional identities

3.1 Motor neuron functional diversity: alpha, beta and gamma motor neurons

The mammalian spinal cord displays a diverse population of motor neurons which is correlated with the heterogeneity of the muscle fiber types they innervate (Figure 1). Motor neurons within the mammalian spinal cord can be categorized into functionally diverse classes and subtypes based on several properties: size and morphology, electrical properties, molecular marker expression and function. Based on these characteristics, motor neurons can be divided into three main types, alpha motor neurons (α-MNs), beta motor neurons (β-MNs) and gamma motor neurons (γ-MNs) and several subtypes. Somatic α-MNs exclusively innervate extrafusal muscle fibers that regulate muscle force and movement. While gamma motor neurons innervate the smaller intrafusal fibers located within the muscle spindle proprioceptive organs, regulate the sensory information reported during muscle stretch from the muscle spindle, and thereby contribute to motor control. And beta motor neuron, innervate both extrafusal and intrafusal fibers, and may contribute to both muscle contraction and sensory information gathered from muscle spindles during muscle stretch, although their exact functional contribution remains unknown.

3.2 Alpha motor neurons

Motor neurons and the muscle fibers their axons innervate display a spectrum of properties that are used to categorize them into subtypes. For instance, there is exquisite matching of muscle contractile properties such as isometric twitch speed, maximum force and endurance and alpha motor neuron properties like size and morphology, excitability and firing pattern, which together allow them to contract synchronously as a “motor unit” to drive muscle contraction [49, 50]. Alpha motor units can be classified into three subtypes: (1) slow-twitch fatigue resistant (S), fast-twitch fatigue-resistant (FR), and fast-twitch fatigable (FF) [51]. Each of these motor units possesses a spectrum of properties: S-type are made of type I muscle fibers that contract slow, develop small quantity of force and are very fatigue-resistant, FR-type are comprised of type IIA fibers that contract faster, develop more force and are less fatigue-resistant than the S-type, while the FF-type are made of type IIB fibers that contract the fastest, develop the highest force and are highly fatigable [50, 51].

Alpha motor neurons, which comprise of one-third of motor neurons in a given motor pool, can be identified based on their size and morphology, excitability, biophysical properties, and the expression of molecular markers. Alpha motor neurons are sequentially activated depending on two linked properties of their cell membrane, that is their soma size and morphology (dendritic arborization) and intrinsic properties (quantity and diversity of ion channels) [52, 53]. For example, as a motor pool receives information from descending inputs for eliciting the contraction of the muscle it innervates, S-type motor neurons, possessing smaller soma sizes and more simple dendritic arborization, are activated first because: (1) they possess a higher input resistance, meaning a larger change in cell membrane voltage over constant current injection, and (2) a lower threshold due to cell membrane Na+ ion channel sensitivity. Thus, S-type motor neurons fire action potentials with lower synaptic current when compared to the larger FF-type motor neurons. Thus, this sequential activation from S-type motor neuron to FF-type motor neuron is thought to follow Henneman’s “size principle” and has implications for motor unit recruitment [54, 55]. Moreover, FR-type motor neurons and motor units have intermediate characteristics between S- and FF-subtypes. Thus, activities that require sustained muscle contraction (standing or walking) results in the recruitment of S-type motor neurons that activate slow motor units, while activities that require potent bursts of muscle contraction (running or jumping) initiates FF-type motor neurons recruitment and subsequent activation of fast motor units, elegantly matching motor neuron morphology and electrical properties with motor unit size required for specific movement tasks [56].

Since motor neurons fire repetitive action potentials, their firing rate can be used to classify them. The firing rate is shaped by persistent inward currents (PICs) generated by voltage-gated Na+ and Ca2+ currents, which are prolonged on dendrites of S-type motor neurons than FF-type motor neurons [57]. PICs amplify and limit the modulation of firing rate, making the S-type motor neurons highly excitable with an initial steep firing rate and subsequent saturation [57]. Another property that determines motor neuron firing rate is the after-hyperpolarization (AHP) phase after the action potential, which is generated by Ca2+-dependent K+ currents. The influx of Ca2+ during the firing of action potentials and the intracellular diffusion, pumping and interaction with proteins regulates AHP-decay times [58, 59, 60]. Thus, FF-type motor neurons have a shorter AHP-decay time, therefore a higher maximum firing frequency than S-type motor neurons, which ultimately matches the contractile frequency of the muscle fiber type these motor neurons innervate [3, 61, 62].

Alpha motor neuron subtypes can also be distinguished by the expression of a subset of genes. For instance, studies have shown that FF-type motor neurons express: Calcitonin gene-related peptide (CGRP)/calca, Chondrolectin, Matrix metallopeptidase 9 (MMP-9), and Delta-like homolog 1 (Dlk1) [63, 64, 65, 66, 67]. While a synaptic vesicle protein, SV2a is expressed postnatally in presynaptic terminals of S-type motor neurons that innervate type I and small type IIA muscle fibers [68]. Moreover, a study in rat showed that Ca2+-activated K+ (SK) channels are expressed in S-type alpha motor neurons and the electrophysiological recordings of these SK3+ motor neurons showed medium size AHP-duration, which seems to be in the range of S-type alpha motor neurons [69]. Other putative markers that are expressed in all three alpha motor neuron subtypes (FF, FR and S) are Hb9::GFP, NeuN, Osteopontin, and Na+/K+ ATPase (Atp1a1 and Atp1a3) (both alpha 1 and 3 isoforms in FF- and FR-subtypes), while UCHL1::eGFP and Na+/K+ ATPase (alpha 1 isoform only) are expressed in S-type [70, 71, 72, 73, 74, 75, 76, 77].

3.3 Beta motor neurons

Another class of motor neuron, the beta motor neurons, were thought to exist only in lower vertebrates such as reptiles, amphibians and birds but are also shown to exist in mammals and comprise one-third of all motor units and innervate three-fourths of all muscle spindles [3, 78, 79]. Despite their abundance in quantity, limited anatomical and functional characterization suggests that beta motor neurons possess intermediate properties between alpha and gamma motor neurons and may play important roles in regulating motor behavior. Unlike alpha and gamma motor neurons which exclusively innervate extrafusal and intrafusal fibers, respectively, beta motor neurons innervate both extrafusal and intrafusal fibers, and thus, regulate both muscle contraction and sensory information from the muscle spindle [78]. Based on their hybrid innervation pattern, beta motor neurons can be subdivided into two subtypes: static and dynamic. Static beta motor neurons innervate type IIa extrafusal fibers and bag2 intrafusal fibers, while dynamic motor neurons innervate type I extrafusal fibers and bag1 intrafusal fibers [3]. Thus, the functional role of beta motor neurons is currently unresolved, however, based on their anatomical properties, they seem to play a role in movement and maintenance of posture [3]. Moreover, studies that aim to identify molecular marker expression and electrophysiological properties of beta motor neurons are especially needed in testing what type of functional role they may play in movement and movement control.

3.4 Gamma motor neurons

Like the alpha motor neurons, gamma motor neuron subtypes can be classified based on their distinct patterns of morphology, connectivity, electrical and functional properties. In mammals, gamma motor neurons represent about one-third of the motor neurons in a given spinal motor pool and are distinguished by their singular role of regulating muscle spindle sensitivity and motor control. Gamma motor rely on GDNF secreted from muscle spindle sensory receptors buried deep within skeletal muscle for their postnatal survival, although it is not known how they survive before maturation [70]. Gamma motor neurons innervate the intrafusal fibers within the muscle spindle and receive presynaptic input from sensory neurons within the muscle spindles. Thus, gamma motor neurons enable continuous flow of sensory information about muscle length and ensure fluid muscular action during muscular contractions [6, 80]. Gamma motor neurons can be subdivided into two types based on their intrafusal fiber innervation patterns and their role in regulating muscle spindle sensory information. Dynamic gamma motor neurons innervate bag1 intrafusal fibers and are thought to regulate muscle length during locomotion, while static gamma motor neurons innervate bag2 and nuclear chain fibers and are thought to be involved in body posture [81].

Gamma motor neurons possess unique biophysical properties which were identified based on intracellular and patch clamp recordings in different animal models. Early on, intracellular recordings in the ventral spinal cord of the cat identified that gamma motor neurons had slower conduction velocity (since they have small axon diameter) compared to alpha motor neurons [82, 83, 84]. Moreover, studies showed that gamma motor neurons possess lower discharge threshold, higher discharge rates and lower membrane input resistance when compared to alpha motor neurons [85]. Furthermore, gamma motor neuron subtypes display unique firing properties: dynamic gamma motor neurons increase the discharge rate of primary sensory afferents when muscle is stretched, while static gamma motor neurons seem to have no effect on primary sensory afferent firing [86]. Studies in mice have also characterized gamma motor neuron electrophysiological properties. Immature gamma motor neurons in young mice (P0-P6) expressing GFP under serotonin receptor 1d (5-ht1d) promoter were patch clamped and revealed that gamma motor neuron electrical properties of rheobase current, input resistance and AHP-decay time ranged between FF-type and S-type alpha motor neurons [87]. In a recent study, more mature gamma motor neurons from mice (P20–22) labeled with high levels of Fluorogold (FG) (retrograde tracer) showed low rheobase current, high firing frequencies and gain when compared to alpha motor neurons [88]. This study matches some of the gamma motor neuron signature properties observed in the cat.

3.5 Mechanisms underlying motor neuron functional diversification

Differences in the connectivity and physiological properties of motor neurons were first reported in the 1940s and 1950s [84, 89, 90, 91], and the significance of these differences have been recognized soon afterwards [6, 84, 92]. Beta and gamma motor neurons were described a few years later [84, 93, 94], and their discovery together significantly broadened our understanding of the neuromuscular bases of movement control and have become canonical topics of neurobiology and neurophysiology textbooks. It therefore seems astonishing that it took over 60 years until the first mechanisms promoting motor neuron functional diversification were discovered [67, 88, 95, 96]. This is likely explained by the difficulties in discovering molecular markers for functional motor neuron types in tetrapod vertebrates due to their general lack of correlation with fixed anatomical features.

Since the first molecular markers for functional motor neuron types were reported, it then took another couple of years until first insights into the mechanisms promoting the diversification of alpha motor neurons into fast and slow types were reported. Herein, the type I transmembrane protein and non-canonical Notch ligand Delta-like homolog 1 (DLK1) was shown to be both necessary and sufficient to promote fast alpha motor neuron gene expression and biophysical signatures required for peak force execution [67]. DLK1 appears to operate in part through the activation of regulatory ion channel subunits including Kcng4/KV6.3/4, apparently reflecting its requirement for both specification as well as maturation of fast alpha motor neurons (Figure 2) [67]. The transition between initial specification and maturation of motor neurons is further reflected by global shifts in gene expression and chromatin accessibility for different transcription factor classes [97].

Figure 2.

Schematic summarizing mechanisms identified so far promoting motor neuron functional diversification. While mechanisms promote the acquisition of motor neuron positional identities (not shown) and the organization into different motor pools, parallel or subsequent mechanisms promote the diversification of motor neurons into different alpha, beta and gamma motor neuron types and subtypes. Cell-autonomous mechanisms have been identified underlying the specification and/or maturation alpha and gamma motor neuron types and subtypes, while the maintenance of some alpha motor neuron properties appear to rely on non-cell-autonomous signals by ventral horn astrocytes. Mechanisms underlying the specification of slow alpha motor neurons and beta motor neurons remain to be identified.

More recent work identified two transcription factors PRDM16 and MECOM as early determinants of primary and fast secondary motor neurons in zebrafish, which are likely homologous to tetrapod fast alpha motor neurons [96]. Both transcription factors broadly, but not completely, regulate fast motor neuron gene expression, raising the question of how these molecules are linked to the actions of DLK1 and Notch signaling (Figure 2) [96]. The incorporation of neurons into functional circuitries entails not only their specification but also their maturation including the acquisition of specific membrane electrical and firing properties [98]. An intriguing non-cell-autonomous mechanism was found to promote and maintain the mature state of fast alpha motor neuron properties [95]. The potassium channel subunit KIR4.1 expressed by ventral horn astrocytes appears to be required for maintaining fast alpha motor neuron soma size and function through the paracrine activation of mTOR signaling (Figure 2) [95]. Again, it remains to be determined how these cell-autonomous (DLK1, PRDM16, MECOM) and non-cell-autonomous mechanisms intersect during alpha motor neuron diversification and maturation (Figure 2).

Advertisement

4. Phylogenetic considerations

The spinal motor systems of fish and tetrapod vertebrates differ in key aspects that reflect adaptations specific to swimming and terrestrial locomotion. For instance, fish generally lack muscle spindles and consequently lack gamma or beta motor neurons regulating muscle spindle proprioception [99]. The appearance of muscle spindle and fusimotor systems providing and regulating muscle proprioception obviously represent adaptations to terrestrial locomotion. Another difference between fish and tetrapods is the spatial arrangement of the force-generating alpha motor neurons in the spinal cord. In zebrafish, these neurons are organized into distinct motor columns presynaptically connected to dedicated interneuron modules and postsynaptically to muscle groups containing either slow, intermediate or fast muscle fiber types [100, 101, 102, 103]. This arrangement allows rapid transitions from slow undulating swimming to fast escape movements and is apparently adapted to pelagic locomotion. Zebrafish larvae initially possess only fast muscle fibers and primary motor neurons facilitating escape swimming movements, while intermediate and slow muscle fibers and matching secondary motor neuron subtypes are generated later during the transition to adulthood.

The modular architecture of motor neuron subtypes in fish contrasts with that found in tetrapods [102], particularly in mammals, in which most motor pools comprise a mosaic of motor neuron types, including the different alpha motor neuron types, gamma motor neurons and small numbers of beta motor neurons [3]. Exceptions can be found in motor pools connecting to specialized muscles, such as predominantly slow antigravity muscles like the m. soleus of the calf or muscles involved in explosive force generation [104], such as the m. rectus femoris of the thigh, which are enriched in slow or fast alpha motor neurons, respectively [67]. Moreover, some muscles in birds and non-avian reptiles display regionalized abundances of slow or fast muscle fibers, which are reflected by similar spatial separation of alpha motor neuron types in the corresponding motor pools. Moreover, particularly in primates, motor pools supplying distal forelimb muscles involved in dexterous movements tend to be enriched in gamma motor neurons. Nevertheless, in mammals, and tetrapod vertebrates in general, motor neuron types within the motor pools typically intermingle and do not spatially segregate, as do the fiber types in the skeletal muscles.

While a degree of conservation of the mechanisms underlying the specification of alpha motor neuron types and the axial motor neurons of fish are expected, it is likely that the transition to mosaic organization of motor pools in tetrapods will be reflected by differences in the mechanisms of specification and functional specializations of alpha motor neuron types. In mouse, for instance, DLK1 promotes a gene expression signature specific for fast alpha motor neurons [67], which, however, shows little overlap with that promoted by PRDM16 and MECOM in zebrafish primary and fast secondary motor neurons [96]. While these discrepancies may in part stem from differences in the transcriptome profiling methodologies used by both studies, it is likely that the spatial and functional reorganization of functional motor neuron types in tetrapods reflects an underlying reorganization of genetic circuitries promoting alpha motor neuron diversification.

A similar difference concerns the molecular profile of slow alpha and gamma motor neurons, with slow secondary motor neurons in zebrafish apparently expressing ERR2. In mouse and chick, high levels of ERR2 (together with its paralogue ERR3) mark gamma motor neurons, while lower levels are initially also expressed by alpha motor neurons [65, 71, 88, 96]. It is therefore tempting to speculate that tetrapod fusimotor (beta and gamma) motor neurons evolved from slow motor neurons with a genetic program that enhanced certain sets of preconfigured properties, such as low firing thresholds and small soma sizes. Intriguingly, while fish trunk muscles lack muscle spindles [99], the jaw muscles of some fishes possess muscle-spindle like structures likely supporting rapid prey capture and manipulation movements [105]. It will be highly interesting to test whether such spindle-like structures would receive innervation from motor neurons and how such hypothetical ‘ancestral’ beta motor neurons would differ in their gene expression profile from axial motor neurons.

Advertisement

5. Conclusions

The two aspects motor neuron diversity discussed in this chapter raise the question of whether both are coordinated and if yes, how? At first glance, both types of motor neuron identity, positional and functional, appear to be established independently form each other. In tetrapod vertebrates, a typical motor pool contains a complement of alpha, beta and gamma motor neuron types and subtypes. Developing motor neurons eventually coalesce into motor pools under the influence of both cell-intrinsic and axon target-induced mechanisms, which in addition to positioning motor neuron somas shape dendritic arbors and presynaptic connectivity. Thus, subsets of motor neurons must somehow acquire distinct sets of properties overlayed on these motor pool-specific features. Moreover, some motor pools are enriched in certain functional motor neuron types mirroring the enrichment in certain fiber types in the muscle they supply. This suggests at least some degree of coordination or crosstalk between motor neuron and muscle fiber type diversification, which is supported by the mutual influence motor neuron and muscle fiber types have on each other. However, this mutual influence only seems to work to some degree, as the conversion of motor neurons from one type to another does not lead to a complete conversion to the corresponding muscle fiber type and vice versa. Moreover, there are several lines of evidence that the initial specification of both motor neuron and muscle fiber types occurs in a cell-autonomous fashion but that they to some degree (yet not completely) remain sculptable throughout later life. This initial cell-autonomy suggests that there may be some intersection between the interpretation of positional signal by motor neurons and the eventual generation of certain ratios of motor neuron types, such as the relative abundance of slow motor neuron and muscle fiber types in the soleus motor pool and muscle, respectively. There are therefore many open questions remaining, not only regarding the functional diversification of motor neurons but also regarding to what degree and how the underlying mechanisms are coordinated with those promoting motor neuron positional identities.

Advertisement

Acknowledgments

We thank Louisa-Carole Neumann and Dr. Daniel Müller for help and suggestions.

References

  1. 1. Levine DN. Sherrington’s “the integrative action of the nervous system”: A centennial appraisal. Journal of the Neurological Sciences. 2007;253(1-2):1-6. DOI: 10.1016/j.jns.2006.12.002. Epub 2007 Jan 12
  2. 2. Catela C, Shin MM, Dasen JS. Assembly and function of spinal circuits for motor control. Annual Review of Cell and Developmental Biology. 2015;31:669-698. DOI: 10.1146/annurev-cellbio-100814-125155. Epub 2015 Sep 21
  3. 3. Manuel M, Zytnicki D. Alpha, beta and gamma motoneurons: Functional diversity in the motor system’s final pathway. Journal of Integrative Neuroscience. 2011;10(3):243-276. DOI: 10.1142/S0219635211002786
  4. 4. Barnard EA, Lyles JM, Pizzey JA. Fibre types in chicken skeletal muscles and their changes in muscular dystrophy. The Journal of Physiology. 1982;331:333-354. DOI: 10.1113/jphysiol.1982.sp014375
  5. 5. Laidlaw DH, Callister RJ, Stuart DG. Fiber-type composition of hindlimb muscles in the turtle, Pseudemys (Trachemys) scripta elegans. Journal of Morphology. 1995;225(2):193-211. DOI: 10.1002/jmor.1052250205
  6. 6. Murthy KS. Vertebrate fusimotor neurones and their influences on motor behavior. Progress in Neurobiology. 1978;11(3-4):249-307. DOI: 10.1016/0301-0082(78)90015-1
  7. 7. Dasen JS. Establishing the molecular and functional diversity of spinal motoneurons. Advances in Neurobiology. 2022;28:3-44. DOI: 10.1007/978-3-031-07167-6_1
  8. 8. Jessell TM. Neuronal specification in the spinal cord: Inductive signals and transcriptional codes. Nature Reviews. Genetics. 2000;1(1):20-29. DOI: 10.1038/35049541
  9. 9. Shirasaki R, Pfaff SL. Transcriptional codes and the control of neuronal identity. Annual Review of Neuroscience. 2002;25:251-281. DOI: 10.1146/annurev.neuro.25.112701.142916. Epub 2002 Mar 27
  10. 10. Ericson J, Rashbass P, Schedl A, Brenner-Morton S, Kawakami A, van Heyningen V, et al. Pax6 controls progenitor cell identity and neuronal fate in response to graded Shh signaling. Cell. 1997;90(1):169-180. DOI: 10.1016/s0092-8674(00)80323-2
  11. 11. Ericson J, Briscoe J, Rashbass P, van Heyningen V, Jessell TM. Graded sonic hedgehog signaling and the specification of cell fate in the ventral neural tube. Cold Spring Harbor Symposia on Quantitative Biology. 1997;62:451-466
  12. 12. Novitch BG, Wichterle H, Jessell TM, Sockanathan S. A requirement for retinoic acid-mediated transcriptional activation in ventral neural patterning and motor neuron specification. Neuron. 2003;40(1):81-95. DOI: 10.1016/j.neuron.2003.08.006
  13. 13. Briscoe J, Pierani A, Jessell TM, Ericson J. A homeodomain protein code specifies progenitor cell identity and neuronal fate in the ventral neural tube. Cell. 2000;101(4):435-445. DOI: 10.1016/s0092-8674(00)80853-3
  14. 14. Novitch BG, Chen AI, Jessell TM. Coordinate regulation of motor neuron subtype identity and pan-neuronal properties by the bHLH repressor Olig2. Neuron. 2001;31(5):773-789. DOI: 10.1016/s0896-6273(01)00407-x
  15. 15. Marquardt T, Pfaff SL. Cracking the transcriptional code for cell specification in the neural tube. Cell. 2001;106(6):651-654. DOI: 10.1016/s0092-8674(01)00499-8
  16. 16. Zhou Q , Anderson DJ. The bHLH transcription factors OLIG2 and OLIG1 couple neuronal and glial subtype specification. Cell. 2002;109(1):61-73. DOI: 10.1016/s0092-8674(02)00677-3
  17. 17. Lee SK, Pfaff SL. Synchronization of neurogenesis and motor neuron specification by direct coupling of bHLH and homeodomain transcription factors. Neuron. 2003;38(5):731-745. DOI: 10.1016/s0896-6273(03)00296-4
  18. 18. Ma YC, Song MR, Park JP, Henry Ho HY, Hu L, Kurtev MV, et al. Regulation of motor neuron specification by phosphorylation of neurogenin 2. Neuron. 2008;58(1):65-77. DOI: 10.1016/j.neuron.2008.01.037
  19. 19. Lee SK, Lee B, Ruiz EC, Pfaff SL. Olig2 and Ngn2 function in opposition to modulate gene expression in motor neuron progenitor cells. Genes & Development. 2005;19(2):282-294. DOI: 10.1101/gad.1257105
  20. 20. Mentis GZ, Alvarez FJ, Bonnot A, Richards DS, Gonzalez-Forero D, Zerda R, et al. Noncholinergic excitatory actions of motoneurons in the neonatal mammalian spinal cord. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(20):7344-7349. DOI: 10.1073/pnas.0502788102. Epub 2005 May 9
  21. 21. Nishimaru H, Restrepo CE, Ryge J, Yanagawa Y, Kiehn O. Mammalian motor neurons corelease glutamate and acetylcholine at central synapses. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(14):5245-5249. DOI: 10.1073/pnas.0501331102. Epub 2005 Mar 21
  22. 22. Rhee HS, Closser M, Guo Y, Bashkirova EV, Tan GC, Gifford DK, et al. Expression of terminal effector genes in mammalian neurons is maintained by a dynamic relay of transient enhancers. Neuron. 2016;92(6):1252-1265. DOI: 10.1016/j.neuron.2016.11.037. Epub 2016 Dec 8
  23. 23. Velasco S, Ibrahim MM, Kakumanu A, Garipler G, Aydin B, Al-Sayegh MA, et al. A multi-step transcriptional and chromatin state cascade underlies motor neuron programming from embryonic stem cells. Cell Stem Cell. 2017;20(2):205-217.e8. DOI: 10.1016/j.stem.2016.11.006. Epub 2016 Dec 8
  24. 24. Bel-Vialar S, Itasaki N, Krumlauf R. Initiating Hox gene expression: In the early chick neural tube differential sensitivity to FGF and RA signaling subdivides the HoxB genes in two distinct groups. Development. 2002;129(22):5103-5115. DOI: 10.1242/dev.129.22.5103
  25. 25. Dasen JS, Liu JP, Jessell TM. Motor neuron columnar fate imposed by sequential phases of Hox-c activity. Nature. 2003;425(6961):926-933. DOI: 10.1038/nature02051
  26. 26. Liu JP, Laufer E, Jessell TM. Assigning the positional identity of spinal motor neurons: Rostrocaudal patterning of Hox-c expression by FGFs, Gdf11, and retinoids. Neuron. 2001;32(6):997-1012. DOI: 10.1016/s0896-6273(01)00544-x
  27. 27. Parker HJ, Krumlauf R. A Hox gene regulatory network for hindbrain segmentation. Current Topics in Developmental Biology. 2020;139:169-203. DOI: 10.1016/bs.ctdb.2020.03.001. Epub 2020 Apr 9
  28. 28. Philippidou P, Dasen JS. Hox genes: Choreographers in neural development, architects of circuit organization. Neuron. 2013;80(1):12-34. DOI: 10.1016/j.neuron.2013.09.020. Epub 2013 Oct 2
  29. 29. Philippidou P, Walsh CM, Aubin J, Jeannotte L, Dasen JS. Sustained Hox5 gene activity is required for respiratory motor neuron development. Nature Neuroscience. 2012;15(12):1636-1644. DOI: 10.1038/nn.3242. Epub 2012 Oct 28
  30. 30. Dasen JS, De Camilli A, Wang B, Tucker PW, Jessell TM. Hox repertoires for motor neuron diversity and connectivity gated by a single accessory factor, FoxP1. Cell. 2008;134(2):304-316. DOI: 10.1016/j.cell.2008.06.019
  31. 31. Lacombe J, Hanley O, Jung H, Philippidou P, Surmeli G, Grinstein J, et al. Genetic and functional modularity of Hox activities in the specification of limb-innervating motor neurons. PLoS Genetics. 2013;9(1):e1003184. DOI: 10.1371/journal.pgen.1003184. Epub 2013 Jan 24
  32. 32. Rousso DL, Gaber ZB, Wellik D, Morrisey EE, Novitch BG. Coordinated actions of the forkhead protein Foxp1 and Hox proteins in the columnar organization of spinal motor neurons. Neuron. 2008;59(2):226-240. DOI: 10.1016/j.neuron.2008.06.025
  33. 33. Shah V, Drill E, Lance-Jones C. Ectopic expression of Hoxd10 in thoracic spinal segments induces motoneurons with a lumbosacral molecular profile and axon projections to the limb. Developmental Dynamics. 2004;231(1):43-56. DOI: 10.1002/dvdy.20103
  34. 34. Wu Y, Wang G, Scott SA, Capecchi MR. Hoxc10 and Hoxd10 regulate mouse columnar, divisional and motor pool identity of lumbar motoneurons. Development. 2008;135(1):171-182. DOI: 10.1242/dev.009225
  35. 35. Jung H, Lacombe J, Mazzoni EO, Liem KF Jr, Grinstein J, Mahony S, et al. Global control of motor neuron topography mediated by the repressive actions of a single hox gene. Neuron. 2010;67(5):781-796. DOI: 10.1016/j.neuron.2010.08.008
  36. 36. Agalliu D, Takada S, Agalliu I, McMahon AP, Jessell TM. Motor neurons with axial muscle projections specified by Wnt4/5 signaling. Neuron. 2009;61(5):708-720. DOI: 10.1016/j.neuron.2008.12.026
  37. 37. Hanley O, Zewdu R, Cohen LJ, Jung H, Lacombe J, Philippidou P, et al. Parallel Pbx-dependent pathways govern the coalescence and fate of motor columns. Neuron. 2016;91(5):1005-1020. DOI: 10.1016/j.neuron.2016.07.043. Epub 2016 Aug 25
  38. 38. Ji SJ, Zhuang B, Falco C, Schneider A, Schuster-Gossler K, Gossler A, et al. Mesodermal and neuronal retinoids regulate the induction and maintenance of limb innervating spinal motor neurons. Developmental Biology. 2006;297(1):249-261. DOI: 10.1016/j.ydbio.2006.05.015. Epub 2006 May 19
  39. 39. Kania A, Johnson RL, Jessell TM. Coordinate roles for LIM homeobox genes in directing the dorsoventral trajectory of motor axons in the vertebrate limb. Cell. 2000;102(2):161-173. DOI: 10.1016/s0092-8674(00)00022-2
  40. 40. Sockanathan S, Jessell TM. Motor neuron-derived retinoid signaling specifies the subtype identity of spinal motor neurons. Cell. 1998;94(4):503-514. DOI: 10.1016/s0092-8674(00)81591-3
  41. 41. Tsuchida T, Ensini M, Morton SB, Baldassare M, Edlund T, Jessell TM, et al. Topographic organization of embryonic motor neurons defined by expression of LIM homeobox genes. Cell. 1994;79(6):957-970. DOI: 10.1016/0092-8674(94)90027-2
  42. 42. Eberhart J, Swartz ME, Koblar SA, Pasquale EB, Krull CE. EphA4 constitutes a population-specific guidance cue for motor neurons. Developmental Biology. 2002;247(1):89-101. DOI: 10.1006/dbio.2002.0695
  43. 43. Kania A, Jessell TM. Topographic motor projections in the limb imposed by LIM homeodomain protein regulation of ephrin-A:EphA interactions. Neuron. 2003;38(4):581-596. DOI: 10.1016/s0896-6273(03)00292-7
  44. 44. Luria V, Krawchuk D, Jessell TM, Laufer E, Kania A. Specification of motor axon trajectory by ephrin-B: EphB signaling: Symmetrical control of axonal patterning in the developing limb. Neuron. 2008;60(6):1039-1053. DOI: 10.1016/j.neuron.2008.11.011
  45. 45. Tiret L, Le Mouellic H, Maury M, Brûlet P. Increased apoptosis of motoneurons and altered somatotopic maps in the brachial spinal cord of Hoxc-8-deficient mice. Development. 1998;125(2):279-291. DOI: 10.1242/dev.125.2.279
  46. 46. Catela C, Shin MM, Lee DH, Liu JP, Dasen JS. Hox proteins coordinate motor neuron differentiation and connectivity programs through ret/Gfrα genes. Cell Reports. 2016;14(8):1901-1915. DOI: 10.1016/j.celrep.2016.01.067. Epub 2016 Feb 18
  47. 47. De Marco Garcia NV, Jessell TM. Early motor neuron pool identity and muscle nerve trajectory defined by postmitotic restrictions in Nkx6.1 activity. Neuron. 2008;57(2):217-231. DOI: 10.1016/j.neuron.2007.11.033
  48. 48. Haase G, Dessaud E, Garcès A, de Bovis B, Birling M, Filippi P, et al. GDNF acts through PEA3 to regulate cell body positioning and muscle innervation of specific motor neuron pools. Neuron. 2002;35(5):893-905. DOI: 10.1016/s0896-6273(02)00864-4
  49. 49. Zengel JE, Reid SA, Sypert GW, Munson JB. Membrane electrical properties and prediction of motor-unit type of medial gastrocnemius motoneurons in the cat. Journal of Neurophysiology. 1985;53(5):1323-1344. DOI: 10.1152/jn.1985.53.5.1323
  50. 50. Kanning KC, Kaplan A, Henderson CE. Motor neuron diversity in development and disease. Annual Review of Neuroscience. 2010;33:409-440. DOI: 10.1146/annurev.neuro.051508.135722
  51. 51. Burke RE, Levine DN, Tsairis P, Zajac FE 3rd. Physiological types and histochemical profiles in motor units of the cat gastrocnemius. The Journal of Physiology. 1973;234(3):723-748. DOI: 10.1113/jphysiol.1973.sp010369
  52. 52. Burke RE, Rudomin P. Spinal nervous and synapses. In: Handbook of Physiology. The Nervous System. Cellular Biology of Neurons. Hoboken, New Jersey, USA: John Wiley & Sons, Inc.; 2011. DOI: 10.1002/cphy.cp010124
  53. 53. Burke RE. Motor units: Anatomy, physiology, and functional organization. In: Handbook of Physiology, the Nervous System, Motor Control. Hoboken, New Jersey, USA: John Wiley & Sons, Inc.; 2011. DOI: 10.1002/cphy.cp010210
  54. 54. Henneman E, Somjen G, Carpenter DO. Functional significance of cell size inspinal motoneurons. Journal of Neurophysiology. 1965;28:560-580
  55. 55. Henneman E, Somjen G, Carpenter D, O. Excitability and inhibitability of motoneurons of different sizes. Journal of Neurophysiology. 1965;28(3):599-620
  56. 56. Burke RE. Motor unit types: Functional specializations in motor control. Trends in Neurosciences. 1980;3:255-258
  57. 57. Heckman CJ, Johnson M, Mottram C, Schuster J. Persistent inward currents in spinal motoneurons and their influence on human motoneuron firing patterns. The Neuroscientist. 2008;14(3):264-275. DOI: 10.1177/1073858408314986. Epub 2008 Apr 1
  58. 58. Kernell D. High frequency repetitive firing of cat lumbosacral motoneurones stimulated by long-lasting injected currents. Acta Physiologica Scandinavica. 1965;65:74-86. DOI: 10.1111/j.1748-1716.1965.tb04251.x
  59. 59. Heckman CJ, Enoka RM. Physiology of the motor neuron and the motor unit. In: Clinical Neurophysiology of Motor Neuron Diseases Handbook of Clinical Neurophysiology. Vol. 4. Amsterdam, The Netherlands: Elsevier; 2004. DOI: 10.1016/S1567-4231(04)04006-7
  60. 60. Binder MD, Heckman CJ, Powers RK. The physiological control of motor neuron activity. In: Comprehensive Physiology. Hoboken, New Jersey, USA: John Wiley & Sons, Inc.; 2011. DOI: 10.1002/cphy.cp120101
  61. 61. Bakels R, Kernell D. Matching between motoneurone and muscle unit properties in rat medial gastrocnemius. The Journal of Physiology. 1993;463:307-324. DOI: 10.1113/jphysiol.1993.sp019596
  62. 62. Gardiner PF. Physiological properties of motoneurons innervating different muscle unit types in rat gastrocnemius. Journal of Neurophysiology. 1993;69(4):1160-1170. DOI: 10.1152/jn.1993.69.4.1160
  63. 63. Piehl F, Arvidsson U, Hökfelt T, Cullheim S. Calcitonin gene-related peptide-like immunoreactivity in motoneuron pools innervating different hind limb muscles in the rat. Experimental Brain Research. 1993;96(2):291-303. DOI: 10.1007/BF00227109
  64. 64. Ringer C, Weihe E, Schütz B. Calcitonin gene-related peptide expression levels predict motor neuron vulnerability in the superoxide dismutase 1-G93A mouse model of amyotrophic lateral sclerosis. Neurobiology of Disease. 2012;45(1):547-554. DOI: 10.1016/j.nbd.2011.09.011. Epub 2011 Sep 21
  65. 65. Enjin A, Rabe N, Nakanishi ST, Vallstedt A, Gezelius H, Memic F, et al. Identification of novel spinal cholinergic genetic subtypes disclose Chodl and Pitx2 as markers for fast motor neurons and partition cells. The Journal of Comparative Neurology. 2010;518(12):2284-2304. DOI: 10.1002/cne.22332
  66. 66. Kaplan A, Spiller KJ, Towne C, Kanning KC, Choe GT, Geber A, et al. Neuronal matrix metalloproteinase-9 is a determinant of selective neurodegeneration. Neuron. 2014;81(2):333-348. DOI: 10.1016/j.neuron.2013.12.009
  67. 67. Mueller D, Cherukuri P, Henningfeld K, Poh CH, Wittler L, Grote P, et al. Dlk1 promotes a fast motor neuron biophysical signature required for peak force execution. Science. 2014;343(6176):1264-1266. DOI: 10.1126/science.1246448
  68. 68. Chakkalakal JV, Nishimune H, Ruas JL, Spiegelman BM, Sanes JR. Retrograde influence of muscle fibers on their innervation revealed by a novel marker for slow motoneurons. Development. 2010;137(20):3489-3499. DOI: 10.1242/dev.053348. Epub 2010 Sep 15
  69. 69. Deardorff AS, Romer SH, Deng Z, Bullinger KL, Nardelli P, Cope TC, et al. Expression of postsynaptic Ca2+-activated K+ (SK) channels at C-bouton synapses in mammalian lumbar-motoneurons. The Journal of Physiology. 2013;591(4):875-897. DOI: 10.1113/jphysiol.2012.240879. Epub 2012 Nov 5
  70. 70. Shneider NA, Brown MN, Smith CA, Pickel J, Alvarez FJ. Gamma motor neurons express distinct genetic markers at birth and require muscle spindle-derived GDNF for postnatal survival. Neural Development. 2009;4:42. DOI: 10.1186/1749-8104-4-42
  71. 71. Friese A, Kaltschmidt JA, Ladle DR, Sigrist M, Jessell TM, Arber S. Gamma and alpha motor neurons distinguished by expression of transcription factor Err3. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(32):13588-13593. DOI: 10.1073/pnas.0906809106. Epub 2009 Jul 27
  72. 72. Misawa H, Hara M, Tanabe S, Niikura M, Moriwaki Y, Okuda T. Osteopontin is an alpha motor neuron marker in the mouse spinal cord. Journal of Neuroscience Research. 2012;90(4):732-742. DOI: 10.1002/jnr.22813
  73. 73. Edwards IJ, Bruce G, Lawrenson C, Howe L, Clapcote SJ, Deuchars SA, et al. Na+/K+ ATPase α1 and α3 isoforms are differentially expressed in α- and γ-motoneurons. The Journal of Neuroscience. 2013;33(24):9913-9919. DOI: 10.1523/JNEUROSCI.5584-12.2013
  74. 74. Yasvoina MV, Genç B, Jara JH, Sheets PL, Quinlan KA, Milosevic A, et al. eGFP expression under UCHL1 promoter genetically labels corticospinal motor neurons and a subpopulation of degeneration-resistant spinal motor neurons in an ALS mouse model. The Journal of Neuroscience. 2013;33(18):7890-7904. DOI: 10.1523/JNEUROSCI.2787-12.2013
  75. 75. Ruegsegger C, Maharjan N, Goswami A, Filézac de L’Etang A, Weis J, Troost D, et al. Aberrant association of misfolded SOD1 with Na(+)/K(+)ATPase-α3 impairs its activity and contributes to motor neuron vulnerability in ALS. Acta Neuropathologica. 2016;131(3):427-451. DOI: 10.1007/s00401-015-1510-4. Epub 2015 Nov 30
  76. 76. Dobretsov M, Stimers JR. Neuronal function and alpha3 isoform of the Na/K-ATPase. Frontiers in Bioscience. 2005;10:2373-2396. DOI: 10.2741/1704
  77. 77. Manuel M, Zytnicki D. Molecular and electrophysiological properties of mouse motoneuron and motor unit subtypes. Current Opinion in Physiology. 2019;8:23-29. DOI: 10.1016/j.cophys.2018.11.008. Epub 2018 Dec 1
  78. 78. Bessou P, Emonet-Dénand F, Laporte Y. Motor fibres innervating extrafusal and intrafusal muscle fibres in the cat. The Journal of Physiology. 1965;180(3):649-672. DOI: 10.1113/jphysiol.1965.sp007722
  79. 79. Hulliger M. The mammalian muscle spindle and its central control. Reviews of Physiology, Biochemistry and Pharmacology. 1984;101:1-110. DOI: 10.1007/BFb0027694
  80. 80. Taylor A. Muscle receptors in the control of voluntary movement. Paraplegia. 1972;9(4):167-172. DOI: 10.1038/sc.1971.28
  81. 81. Fitz-Ritson D. The anatomy and physiology of the muscle spindle, and its role in posture and movement: A review. The Journal of the Canadian Chiropractic Association. 1982;26(4):144-150
  82. 82. Kuffler SW, Hunt CC, Quilliam JP. Function of medullated small-nerve fibers in mammalian ventral roots; efferent muscle spindle innervation. Journal of Neurophysiology. 1951;14(1):29-54. DOI: 10.1152/jn.1951.14.1.29
  83. 83. Hunt CC, Kuffler SW. Further study of efferent small-nerve fibers to mammalian muscle spindles; multiple spindle innervation and activity during contraction. The Journal of Physiology. 1951;113(2-3):283-297. DOI: 10.1113/jphysiol.1951.sp004572
  84. 84. Eccles JC, Eccles RM, Iggo A, Lundberg A. Electrophysiological studies on gamma motoneurones. Acta Physiologica Scandinavica. 1960;50:32-40. DOI: 10.1111/j.1748-1716.1960.tb02070.x
  85. 85. Kemm RE, Westbury DR. Some properties of spinal gamma-motoneurones in the cat, determined by micro-electrode recording. The Journal of Physiology. 1978;282:59-71. DOI: 10.1113/jphysiol.1978.sp012448
  86. 86. Matthews PB. The differentiation of two types of fusimotor fibre by their effects on the dynamic response of muscle spindle primary endings. Quarterly Journal of Experimental Physiology and Cognate Medical Sciences. 1962;47:324-333
  87. 87. Enjin A, Leão KE, Mikulovic S, Le Merre P, Tourtellotte WG, Kullander K. Sensorimotor function is modulated by the serotonin receptor 1d, a novel marker for gamma motor neurons. Molecular and Cellular Neurosciences. 2012;49(3):322-332. DOI: 10.1016/j.mcn.2012.01.003. Epub 2012 Jan 17
  88. 88. Khan MN, Cherukuri P, Negro F, Rajput A, Fabrowski P, Bansal V, et al. ERR2 and ERR3 promote the development of gamma motor neuron functional properties required for proprioceptive movement control. PLoS Biology. 2022;20(12):e3001923. DOI: 10.1371/journal.pbio.3001923
  89. 89. Leksell L. The action potentials and excitatory effects of the small ventral root fibres to skeletal muscle. Acta Physiologica Scandinavica. 1945;10(1(Suppl. 31)):1-84
  90. 90. Katz B. The efferent regulation of the muscle spindle in the frog. The Journal of Experimental Biology. 1949;26(2):201-217. DOI: 10.1242/jeb.26.2.201
  91. 91. Henneman E. Relation between size of neurons and their susceptibility to discharge. Science. 1957;126(3287):1345-1347. DOI: 10.1126/science.126.3287.1345
  92. 92. Ellaway PH, Taylor A, Durbaba R. Muscle spindle and fusimotor activity in locomotion. Journal of Anatomy. 2015;227(2):157-166. DOI: 10.1111/joa.12299. Epub 2015 Jun 5
  93. 93. Bessou P, Emonet-Denand F, Laporte Y. Occurrence of intrafusal muscle fibres innervation by branches of slow α motor fibres in the cat. Nature. 1963;198(4880):594-595
  94. 94. Kernell D. Input resistance, electrical excitability, and size of ventral horn cells in cat spinal cord. Science. 1966;152(3729):1637-1640. DOI: 10.1126/science.152.3729.1637
  95. 95. Kelley KW, Ben Haim L, Schirmer L, Tyzack GE, Tolman M, Miller JG, et al. Kir4.1-dependent astrocyte-fast motor neuron interactions are required for peak strength. Neuron. 2018;98(2):306-319.e7. DOI: 10.1016/j.neuron.2018.03.010. Epub 2018 Apr 5
  96. 96. D’Elia KP, Hameedy H, Goldblatt D, Frazel P, Kriese M, Zhu Y, et al. Determinants of motor neuron functional subtypes important for locomotor speed. Cell Reports. 2023;42(9):113049. DOI: 10.1016/j.celrep.2023.113049. Epub ahead of print
  97. 97. Patel T, Hammelman J, Aziz S, et al. Transcriptional dynamics of murine motor neuron maturation in vivo and in vitro. Nature Communications. 2022;13:5427. DOI: 10.1038/s41467-022-33022-4
  98. 98. Hobert O, Kratsios P. Neuronal identity control by terminal selectors in worms, flies, and chordates. Current Opinion in Neurobiology. 2019;56:97-105. DOI: 10.1016/j.conb.2018.12.006. Epub 2019 Jan 18
  99. 99. Barker D. The morphology of muscle receptors. In: Hunt CC, editor. Handbook of Sensory Physiology Vol. III/2, Muscle Receptors. Berlin-Heidelberg-New York: Springer-Verlag; 1974
  100. 100. Goody MF, Carter EV, Kilroy EA, Maves L, Henry CA. “muscling” throughout life: Integrating studies of muscle development, homeostasis, and disease in zebrafish. Current Topics in Developmental Biology. 2017;124:197-234. DOI: 10.1016/bs.ctdb.2016.11.002. Epub 2016 Dec 23
  101. 101. Ampatzis K, Song J, Ausborn J, El Manira A. Pattern of innervation and recruitment of different classes of motoneurons in adult zebrafish. The Journal of Neuroscience. 2013;33(26):10875-10886. DOI: 10.1523/JNEUROSCI.0896-13.2013
  102. 102. Ampatzis K, Song J, Ausborn J, El Manira A. Separate microcircuit modules of distinct v2a interneurons and motoneurons control the speed of locomotion. Neuron. 2014;83(4):934-943. DOI: 10.1016/j.neuron.2014.07.018. Epub 2014 Aug 7
  103. 103. Gabriel JP, Ausborn J, Ampatzis K, Mahmood R, Eklöf-Ljunggren E, El Manira A. Principles governing recruitment of motoneurons during swimming in zebrafish. Nature Neuroscience. 2011;14(1):93-99. DOI: 10.1038/nn.2704. Epub 2010 Nov 28
  104. 104. Armstrong RB, Phelps RO. Muscle fiber type composition of the rat hindlimb. The American Journal of Anatomy. 1984;171(3):259-272. DOI: 10.1002/aja.1001710303
  105. 105. Maeda N, Miyoshi S, Toh H. First observation of a muscle spindle in fish. Nature. 1983;302(5903):61-62. DOI: 10.1038/302061a0

Written By

Mudassar Nazar Khan and Till Marquardt

Submitted: 28 September 2023 Reviewed: 12 February 2024 Published: 22 May 2024