Skip to main content

PERSPECTIVE article

Front. Neurosci., 20 March 2018
Sec. Neurodegeneration

Serotonergic Dysfunction in Amyotrophic Lateral Sclerosis and Parkinson's Disease: Similar Mechanisms, Dissimilar Outcomes

  • 1Laboratory of Neurochemistry and Behavior, Department of Biomedical Sciences, Institute Born-Bunge, University of Antwerp, Antwerp, Belgium
  • 2Department of Neurology and Alzheimer Research Center, University of Groningen and University Medical Center Groningen, Groningen, Netherlands
  • 3Department of Neurology, Memory Clinic of Hospital Network Antwerp (ZNA) Middelheim and Hoge Beuken, Antwerp, Belgium

Amyotrophic lateral sclerosis (ALS) and Parkinson's disease (PD) share similar pathophysiological mechanisms. From a neurochemical point of view, the serotonin (5-hydroxytryptamine; 5-HT) dysfunction in both movement disorders—related to probable lesioning of the raphe nuclei—is profound, and, therefore, may be partially responsible for motor as well as non-motor disturbances. More specifically, in ALS, it has been hypothesized that serotonergic denervation leads to loss of its inhibitory control on glutamate release, resulting into glutamate-induced neurotoxicity in lower and/or upper motor neurons, combined with a detrimental decrease of its facilitatory effects on glutamatergic motor neuron excitation. Both events then may eventually give rise to the well-known clinical motor phenotype. Similarly, disruption of the organized serotonergic control on complex mesencephalic dopaminergic connections between basal ganglia (BG) nuclei and across the BG-cortico-thalamic circuits, has shown to be closely involved in the onset of parkinsonian symptoms. Levodopa (L-DOPA) therapy in PD largely seems to confirm the influential role of 5-HT, since serotonergic rather than dopaminergic projections release L-DOPA-derived dopamine, particularly in extrastriatal regions, emphasizing the strongly interwoven interactions between both monoamine systems. Apart from its orchestrating function, the 5-HT system also exerts neuroprotective and anti-inflammatory effects. In line with this observation, emerging therapies have recently focused on boosting the serotonergic system in ALS and PD, which may provide novel rationale for treating these devastating conditions both on the disease-modifying, as well as symptomatic level.

Background

The neurotransmitter serotonin (5-hydroxytryptamine; 5-HT) is produced in the raphe nuclei (RN), a moderately sized cluster of caudal and rostral neurons (B1-B9) found in the brainstem (Dahlstroem and Fuxe, 1964). Axons arising from the caudal group (B1-B4) form a descending system projecting to the spinal cord, cerebellum, pontine, and midbrain structures, whereas ascending fibers emanating from the more rostral clusters (B5-B9) connect with the cerebral cortex, (hypo)thalamus, basal ganglia and hippocampus among others. About roughly 300,000 5-HT-containing neurons in the human brain bear a tremendous number of collateral branches so that the serotonergic system densely innervates nearly all brain regions (Jacobs and Azmitia, 1992). It is, therefore, not surprising that this extensive neuronal network is implicated in the regulation of numerous physiological events, such as hormone secretion, sleep-wake cycle, motor control, immune system functioning, nociception, food intake, energy balance/metabolism, cardiovascular/respiratory functioning, body temperature, affect/aggression, consciousness, learning, and memory (Ciranna, 2006; Sandyk, 2006). Serotonin receptors are subdivided into seven families (5-HT1−7), based on structural, biochemical and pharmacological characteristics, resulting into 14 subtypes (5-HT1A/1B/1D/1e/1F, 5-HT2A/2B/2C, 5-HT3, 5-HT4, 5-HT5A/5b, 5-HT6, and 5-HT7). With the sole exception of 5-HT3, which belongs to the ligand gated ion channels, all 5-HT receptors are G protein-coupled receptors, mediating a variety of physiological and behavioral functions (Filip and Bader, 2009). Regarding amyotrophic lateral sclerosis (ALS) and Parkinson's disease (PD) pathophysiology, especially 5-HT1A/1B and 5-HT2A/2B/2C seem crucial (Cummings et al., 2013; Miyazaki et al., 2013; De Deurwaerdère and Di Giovanni, 2016; El Oussini et al., 2016). In short, the 5-HT1A receptor is expressed in the RN as a presynaptic autoreceptor, while it also functions as a postsynaptic heteroreceptor in areas of the limbic system, such as the prefrontal cortex, hippocampus, lateral septum, and amygdala, as well as in (hypo)thalamus, and basal ganglia (Hoyer et al., 1994). Activation of 5-HT1A autoreceptors on the cell bodies or dendrites of the RN neurons exerts inhibitory feedback in response to local 5-HT release. The 5-HT1B receptors are centered on axonal terminals of (non)serotonergic neurons, mainly found in the basal ganglia and substantia nigra (SN) (Bonaventure et al., 1997). Interestingly, it has been indicated that 5-HT1B receptors are preferentially located on presynaptic terminals of ɤ-amino-butyric acid (GABA)ergic neurons, and, it has also been suggested that thalamostriatal and/or corticostriatal glutamatergic neurons express presynaptic 5-HT1B receptors (Bonaventure et al., 1998). In contrast, the 5-HT2A receptor is found mainly in the periphery and neocortical areas, where they are implicated in the pathogenesis of schizophrenia and hallucinations (Burnet et al., 1995; Hannon and Hoyer, 2008). These receptors are highly expressed in both pyramidal cells and GABAergic interneurons. Moreover, cerebral 5-HT2B receptors are present in the cerebellum, cerebral cortex, hypothalamus, corpus callosum and amygdala, causing anxiolytic effects among others (Duxon et al., 1997). Noteworthy, this receptor subtype is likely to be expressed by RN neurons, where this autoreceptor might play a role in the regulation of the serotonin transporter (SERT) (Diaz et al., 2012). Next, the 5-HT2C subtype is widely distributed throughout the brain and has been proposed as the main mediators of the different actions of 5-HT in the central nervous system (Hannon and Hoyer, 2008). Additionally, 5-HT2C receptors are commonly found in the choroid plexus, where they modulate cerebrospinal fluid (CSF) production (Pasqualetti et al., 1999).

The serotonergic system is organized in such a way that it exerts widespread effects on targeted neurons, such as motor neuron excitability threshold control, and interacts with many other neurotransmitters, including dopamine (DA), noradrenaline (NA), glutamate, GABA, and various peptides (Ciranna, 2006). Remarkably, 5-HT also plays an important part in the development of the embryonic nervous system, which relates to neurite outgrowth and other aspects of neuronal differentiation, including synaptogenesis (Lauder, 1990). Given its complex but critical modulating characteristics, 5-HT can be regarded as one of the principal orchestrators of the central nervous system, with a very significant role in motor activity. In PD and ALS, two invariably fatal neurodegenerative conditions, the motor and non-motor features have been partially attributed to disease-related malfunctioning of this overseeing neurotransmitter system.

Serotonergic Degeneration in PD and ALS

Staging of brain pathology in PD demonstrated an early involvement of Lewy body depositions within the RN. In more detail, Halliday et al. (1990) firstly described a 56% loss of serotonergic neurons in the median RN of PD compared to control brain. Afterwards, Braak et al. (2003) determined six stages in the evolution of PD-related pathology, with lesions being present in the median RN in the caudal brainstem already from stage two onwards. Furthermore, 5-HT depletion was observed in various target areas of the RN, such as in the basal ganglia, hypothalamus, hippocampus, and prefrontal cortex (Fahn et al., 1971; Shannak et al., 1994). This was later confirmed by in vivo imaging studies, revealing new insights. For instance, Politis et al. (2010) applied 11C-DASB-PET to early-stage PD patients, and demonstrated reduced SERT binding in the caudate nucleus, (hypo)thalamus, and anterior cingulate cortex, whereas PD subjects with established disease showed additional 11C-DASB binding reductions in the putamen, insula, posterior cingulate cortex, and, prefrontal cortex. Further binding reductions were noticed in the ventral striatum, RN, and amygdala of advanced PD patients. Interestingly, the loss of SERT binding in the RN occurred in later stages, pointing to an earlier loss of serotonergic projections instead of the neurons themselves.

In ALS, distribution patterns of TAR DNA-binding protein (TDP)-43 intraneuronal inclusions have only recently been investigated, summing up into a total of four discriminative neuropathological stages (Brettschneider et al., 2013). Notably, it has been theorized that sites with projections to the cortex remain intact in ALS, unlike those receiving corticofugal axonal projections, supporting the hypothesis of prion-like propagation of TDP-43, potentially from the motor cortex downwards (dying forward/back hypotheses, Figure 1). In agreement, the upper RN with diffuse cortical projections barely become affected by TDP-43 pathology in ALS, which is in great contrast with PD or Alzheimer's disease (AD) (Braak et al., 2013). Nevertheless, a marked reduction in both cortical and RN 5-HT1A receptor binding (21%) has been observed (Turner et al., 2005), and, several studies previously evidenced decreased levels of 5-HT, 5-hydroxyindoleacetic acid (5-HIAA; main metabolite of 5-HT) or tryptophan (precursor of 5-HT) in CSF, plasma, and/or spinal cord (Monaco et al., 1979; Ohsugi et al., 1987; Bertel et al., 1991; Sofic et al., 1991). Platelet 5-HT levels also positively correlated with survival in ALS subjects (Dupuis et al., 2010). Consequently, it has been postulated that 5-HT1A/2 receptor (anta)agonists, 5-HT precursors [e.g., 5-hydroxytryptophan (5-HTP)] (Turner et al., 2003) or 5-HT2B/C receptor inverse agonists (Dentel et al., 2013) might improve locomotor function and even strategically interfere with ALS disease course. On the whole, the serotonergic theory in ALS has gained renewed interest especially due to several recent publications (Dentel et al., 2013; El Oussini et al., 2016, 2017).

FIGURE 1
www.frontiersin.org

Figure 1. Schematic representation of dysfunctional serotonergic pathway interactions in ALS and PD, mediated by lesioned raphe nuclei (RN) centered in the brainstem. In summary, serotonergic loss in amyotrophic lateral sclerosis (ALS) brain and subsequent loss of its inhibitory control on glutamate release cause glutamate-induced excitotoxicity leading to upper/lower motor neuron damage. In this regard, the dying forward hypothesis proposes that ALS is a disorder primarily of the corticomotoneurons, with anterior horn cell degeneration propagated via an anterograde glutamate-dependent excitotoxic process. In contrast, the dying back hypothesis proposes that ALS begins within the muscle or neuromuscular junction, with pathogens being retrogradely transported from the neuromuscular junction to the cell body where they may exert their deleterious effects. Simultaneously, this figure illustrates the pathophysiological serotonergic-dopaminergic interactions on the striatal level in Parkinson's disease (PD), where lesioning of the RN (red spheres) in addition to dopamine (DA) depletion in the striatum and substantia nigra (black-bolded dashes) give rise to a net decreased activity of the motor cortex. Adapted from Vucic et al. (2013), ©2013 with permission from BMJ Publishing Group Ltd.

5-HT and the Control of Motor Neuron Excitability: Possible Implications

The indolamine 5-HT has facilitatory effects on glutamatergic motor neuron excitation by augmenting weak or polysynaptic inputs, bringing motor neurons to threshold. This effect on spinal motor neurons is exerted through 5-HT1/2 receptors (for review: Sandyk, 2006). In ALS, serotonergic denervation has been hypothesized to lead to significant loss of inhibitory control on glutamate release, via decreased binding on presynaptic 5-HT1B receptors, triggering glutamate-induced neurotoxicity, and, eventually, rapid-onset loss of upper and lower motor neurons (Muramatsu et al., 1998). Upper motor neurons are glutamatergic neurons located in layer V of the motor cortex, project to spinal motor neurons through the corticospinal tract, and are the major source of descending motor commands for voluntary movement (Lemon, 2008). Meanwhile, progressive degeneration of 5-HT neurons in the motor cortex, RN and their projections may lead to a compensatory increase in glutamate excitation (Bertel et al., 1991), adding up to the clinical motor phenotype (Figure 1). Conversely, motor neuron groups such as the oculomotor, trochlear, and abducens nuclei, and the cerebellum, which only receive sparse serotonergic innervation, appear more resistant to the process of neurodegeneration in ALS. Moreover, it is possible that differences between bulbar and spinal ALS in the course of the disease may be related to the degree of cerebral 5-HT depletion, which seems more extensive in the bulbar subtype (Turner et al., 2005). Additionally, 5-HT is a precursor of melatonin, which inhibits glutamate release and glutamate-induced neurotoxicity (Zhang et al., 1999).

One of the possible implications of serotonergic degeneration with regard to motor symptoms in ALS, is spasticity (Dentel et al., 2013; El Oussini et al., 2017). For instance, El Oussini et al. (2017) recently demonstrated that degeneration of brainstem 5-HT neurons in transgenic SOD1 (G37R) mice, more particularly the dorsal and median RN, induced spasticity. This hyperreflexia is able to compensate for motor deficits, allowing the maintenance of motor function after disease onset. Spasticity is a painful symptom which can severely restrict quality of life on a daily basis. Remarkably, SB206553 administration, a 5-HT2B/C receptor inverse agonist, completely abolished spasticity symptoms (Murray et al., 2010; El Oussini et al., 2017). The authors further stress that selective degeneration of the RN might directly lead to motor neuron hyperexcitability and spasticity—rather than degeneration of upper motor neurons in the cerebral motor cortex.

Related Treatment Options

So far, riluzole and edaravone—of which latter drug has recently been FDA-licensed in the US and Japan (Hardiman and van den Bergh, 2017)—only modestly improve motor symptoms and daily functioning in ALS patients, but with a reasonable safety profile of riluzole 100 mg daily (Miller et al., 2007). However, the treatment regimen for edaravone is inconvenient and costly (Hardiman and van den Bergh, 2017). Riluzole acts as an N-methyl-D-aspartate (NMDA)-receptor antagonist, whereas edaravone is a free radical scavenger. Both mechanisms of action thus support the glutamate excitotoxicity-driven hypothesis.

As for clinical trials with serotonergic therapies, Meininger et al. (2004) carried out two randomized, double-blind, placebo-controlled multicenter studies (phase III) with xaliproden, a 5-HT1A receptor agonist which has neurotrophic and neuroprotective effects, to assess its safety and efficacy at two doses. ALS patients were randomly assigned to placebo, 1 or 2 mg xaliproden orally once daily as a monotherapy (867 patients), or, to the same regimen with addition of riluzole 50 mg (1,210 patients). In the end, however, the primary outcome measures (time to death, tracheostomy, or permanent assisted ventilation) did not reach statistical significance. There only was a therapeutic benefit on the second outcome measure, i.e., vital capacity (maximum volume of air exhaled at a steady state after maximum inhalation of a single breath) at the 1 and 2 mg dose without riluzole, which let the authors conclude that xaliproden does not effectively slow down disease progression. In short, strong evidence is currently lacking and insufficient regarding the potential benefits of serotonergic therapies in ALS, despite of its remarkable 5-HT-related pathophysiological characteristics described above.

Finally, monoamine oxidase-B (MAO-B) inhibitors such as deprenyl, rasagiline, or selegiline affect the release and increase the content of not only DA and NA, but also of 5-HT (for review: Finberg, 2014). MAO-B inhibitors also have neuroprotective properties. Following the introduction of rasagiline to the therapeutic armamentarium for PD, various successes have been reported (Rascol et al., 2005, 2011). In ALS, results are less consistent, with selegiline treatment having no significant effect on the rate of clinical progression or outcome in ALS as evidenced by Lange et al. (1998), whereas deprenyl and rasagiline seem more promising, but necessitate further scrutiny (Jossan et al., 1994; Macchi et al., 2015).

Serotonergic Modulation of Basal Ganglia and Mesencephalic Dopaminergic Activity in PD

The basal ganglia (BG) are composed of the striatum (caudate nucleus and putamen), subthalamic nucleus (STN), internal and external globus pallidus (GPi/e) and SN, and are part of the BG-cortico-thalamic circuits. This highly organized network is important for motor control, emotion, and cognition. It has been firmly established that BG nuclei receive vast serotonergic input mainly coming from the rostral RN clusters (B7), with effects on mesencephalic dopaminergic activity depending on the specific nucleus and its receptor distribution (excitatory via 5-HT1A/1B/2A/3/4/7 and inhibitory via 5-HT2C/6 receptors (Paolucci et al., 2003; Miguelez et al., 2014; De Deurwaerdère and Di Giovanni, 2016). In PD, lesioning of the RN in addition to DA depletion in the striatum and SN—particularly of the pars compacta (SNc)—are hallmarks of the disease, leading to overactivation of the output regions of the BG, i.e., GPi and SN pars reticulata (SNr), which contain large GABAergic neurons. This cascade results in a net decreased activity of the supplementary motor areas, premotor, and primary motor cortices, triggering parkinsonian symptoms (Albin et al., 1989; Figure 1). Overall, the loss of 5-HT neurons is not as profound as the loss of DA neurons, and may not be sufficient to cause motor or non-motor symptoms per se, however, both systems closely interact, and combined depletion certainly seems to aggravate the situation, as was shown in a parkinsonian rat model (Delaville et al., 2012). Moreover, 5-HT and 5-HIAA levels, as well as SERT expression, are reduced in various BG nuclei (Scatton et al., 1983; Guttman et al., 2007; Kish et al., 2008), and the serotonergic system is strongly involved in the mechanism of action of antiparkinsonian therapeutics, such as levodopa (L-DOPA), and high frequency stimulation of the STN (Navailles and De Deurwaerdère, 2012).

L-DOPA Actions via Serotonergic Nerve Terminals in PD: The Influential Effect of 5-HT

Levodopa (L-DOPA), the metabolic precursor of DA, is a well-established symptomatic treatment for the motor deficits in PD. Paradoxically, L-DOPA-induced dyskinesia (LID), as well as hallucinations, are unfortunate but more or less inevitable corollaries of its long-term administration (De Deurwaerdère et al., 2017). Despite the traditional assumption that L-DOPA is transformed in residual striatal dopaminergic neurons into DA, interestingly, a more important role for serotonergic than dopaminergic projections has been identified for the increase of extracellular DA, predominantly in prefrontal cortex, nucleus accumbens, STN, hippocampus, and additional extrastriatal regions (De Deurwaerdère et al., 2017). Briefly, 5-HT neurons convert exogenous L-DOPA into DA and store neo-synthesized DA into vesicles for exocytosis via vesicular monoamine transporter 2, as was originally shown in rats (Arai et al., 1995). Since the distribution of 5-HT terminals in the brain is very different from dopaminergic innervation, the magnitude of effect in extrastriatal regions is tremendous compared to physiological conditions, especially at low L-DOPA doses, so that 5-HT in fact controls the dopaminergic output in a state and region-dependent manner (Navailles and De Deurwaerdère, 2012).

Latter phenomenon has even led to the assumption that future envisaged pharmacotherapeutic strategies to treat LID should specifically aim at controlling L-DOPA-stimulated DA release from extrastriatal 5-HT neurons (Miguelez et al., 2014; De Deurwaerdère et al., 2017). Recently, the use of 5-HTP (Tronci et al., 2013) or 5-HT1A/B receptor agonists (e.g., eltoprazine or buspirone; Svenningsson et al., 2015; De Deurwaerdère et al., 2017)—influencing DA release indirectly via action on the overall 5-HT tone—has been proposed. As for exacerbation of psychosis by L-DOPA treatment—attributed to excessive DA release in the mesolimbic areas rather than the motor striatum, mediated by hypersensitive 5-HT signaling—a favorable role for 5-HT2A receptor inverse agonists (e.g., pimavanserin) or 5-HT2A antagonists (e.g., low doses of clozapine) has likewise been demonstrated (Cummings et al., 2013). These findings suggest that the serotonergic system may even adapt to the lack of DA by adopting anatomical and functional transformations in PD.

Alterations in Other Monoamine Neurotransmitter Systems

NA levels have been previously reported to be significantly increased in the cervical, thoracic and lumbar spinal cord of ALS patients compared to controls (Bertel et al., 1991), with highest concentrations measured in ventral and intermediate gray matter. In CSF, a similar increase has been noted (Ziegler et al., 1980). Independently of 5-HT, NA increases the excitability of motor neurons to glutamate (White and Neuman, 1980). Bertel et al. (1991) further discussed that in all probability, it is unlikely that the noradrenergic changes were due to the effect of tissue shrinkage—since concentrations were expressed in ng/g wet weighed tissue—but rather a consequence of denser noradrenergic (neosympathetic) innervation, such as from sprouting of noradrenergic fibers into affected areas. In PD, the noradrenergic dysfunction has been investigated in more detail. In summary, α-synuclein depositions in the locus coeruleus (stage 2), the brain's main source of NA, have been evidenced to precede that of the SN (stage 3) (Braak et al., 2003). Consequently, neuronal loss in this noradrenergic nucleus and the accompanying noradrenergic deficiency both on the central and peripheral level have been related to various motor and non-motor (cognitive) symptoms, including the progression to (prodromal) dementia (Vermeiren and De Deyn, 2017).

A potential dopaminergic deficit in ALS has only been scarcely investigated, with significantly reduced striatal DA transporter expression in patients with bulbar- or limb-onset compared to controls ([I-123]-IPT-SPECT) (Borasio et al., 1998), while in drug-naïve, sporadic ALS patients, decreased striatal D2-receptor binding could be partially reversed by riluzole (Vogels et al., 1999). On the contrary, no differences in spinal DA concentrations were found between ALS and control subjects (Bertel et al., 1991). No research has been performed yet with regard to 5-HT-DA interactions in mesencephalic brain areas or BG nuclei, but a study by Xu et al. (2017) observed an abnormal cortical-BG network in ALS after applying resting state fMRI and voxel-wise network analysis.

The Non-motor Outcomes of Serotonergic Dysfunction in ALS and PD

New findings point at an important link between non-linear progressive degeneration of serotonergic terminals and non-motor disturbances in PD, such as depression, fatigue, weight loss, and anxiety (Politis and Niccolini, 2015; Huot et al., 2017). Similarly, cognitive impairment and dementia are major issues in PD, and might be ascribed to serotonergic dysfunction too (Huot et al., 2017). In this respect, a phase 2 trial is currently evaluating the safety, tolerability and efficacy of SYN120, a dual 5-HT6/5-HT2A antagonist, in 80 PD dementia patients over a 16-week period [SYNAPSE; NCT02258152 (clinicaltrials.gov)]. As for ALS, fatigue and abnormal peripheral glucose metabolism have been suggested (Reyes et al., 1984). Major depressive disorder, in which an ~12% reduction of cortical 5-HT1A binding is seen in non-ALS cases (Sargent et al., 2000), is relatively rare in ALS patients, even in later stages (Goldstein et al., 2006). More recently, Vercruysse et al. (2016) indicated that serotonergic axonal loss in the arcuate nucleus of the hypothalamus in combination with decreased hypothalamic 5-HT levels primarily caused a melanocortin deficit in mutant SOD1 (G86R) mice, which contributed to dysregulated food intake/weight loss.

Furthermore, self-referential thinking (i.e., reflecting one's mental self) is a key cognitive process which seems to be regulated by 5-HT1A receptors within the default mode network, which comprises the precuneus, posterior cingulate cortex, medial prefrontal cortex, and, the temporoparietal junction (Hahn et al., 2012). In this regard, Fomina et al. (2017) observed electroencephalography correlates (bandpower) of self-referential thinking in the medial prefrontal cortex of healthy individuals, but not ALS patients. The authors concluded that these cognitive abnormalities, such as anosognosia, may well be in compliance with the proposed serotonergic theory in ALS.

(DIS)Similarities: 5-HT as a Crucial Disease Modifier

The process of normal, healthy aging has complex effects on central and peripheral serotonergic transmission. Accumulating (pre)clinical evidence suggests a linear and gradual decline of 5-HT connections from the RN, as well as altered SERT and 5-HT1A/2A receptor expressions in multiple brain regions (Rodríguez et al., 2012). However, in ALS and PD, RN neuronal loss and/or loss of serotonergic projections due to marked and early TDP-43 and α-synuclein depositions in target areas might cause major imbalance in monoaminergic neurotransmission across the brain (Turner et al., 2005; Dentel et al., 2013; Politis and Niccolini, 2015), accounting for numerous motor, behavioral and cognitive dysfunctions.

The neurochemical similarity between ALS and PD, is that in both conditions, the supervising but damaged serotonergic system has lost pre- and postsynaptic regulatory functions on neighboring systems, leading to loss of inhibitory control of glutamate release and loss of facilitatory effects of glutamatergic motor neuron excitation in ALS, whereas in PD, this results in alterations of the complex serotonergic modulation of mesencephalic dopaminergic systems. Maybe unexpectedly, this largely and selectively affects upper and lower motor neurons in ALS brain and spinal cord, causing neuromuscular disease, while in PD, the effects rather remain central, i.e., at the level of the SNc/r, BG nuclei and (extra)striatal regions. Serotonergic alterations in ALS brain and RN have been found before (Turner et al., 2005), but the overall outcome of the serotonergic shortage on the corticocerebral level remains to be elucidated. In addition, autonomic and olfactory dysfunction in PD have been ascribed to peripheral noradrenergic alterations, potentially resulting from LC lesioning, and, likely, far preceding the motor deficits (Vermeiren and De Deyn, 2017). In contrast, there is a fairly dissimilar clinical outcome for both neurodegenerative diseases, with more 5-HT-associated non-motor disturbances in PD vs. a very typical motor—but less non-motor—region-dependent degenerative pattern in ALS, causing the well-characterized limb- or bulbar-onset phenotype (Figure 2).

FIGURE 2
www.frontiersin.org

Figure 2. Venn diagram—visualization of the complex interplay of neurochemical and clinical keystones in ALS and PD. This figure depicts the complex interplay between neurochemical characteristics in ALS and PD related to their disease-specific clinical (i.e., motor and non-motor) outcome. Mutually influential and/or synergistic interactions are indicated with arrowheads at both ends. Question marks over the arrows refer to partially-proven or suggestive mechanisms of neurochemical features. The Venn diagram clearly shows that serotonergic dysfunction is the central overlap in the overall pathophysiology of ALS and PD, whereas the neurochemical causes of the clinical non-motor (i.e., behavioral) disturbances, particularly in ALS, necessitate further scrutiny. Similarly, the reciprocal interaction between the noradrenergic and serotonergic/dopaminergic disturbances in ALS and PD remains to be elucidated, even though in PD, NA dysfunction has been previously linked with CD and dementia progression. 5-HT, serotonin (5-hydroxytryptamine); ALS, amyotrophic lateral sclerosis; CD, cognitive deterioration; DA, dopamine; LID, levodopa-induced dyskinesia; MDD, major depressive disorder; NA, noradrenaline; OCD, obsessive-compulsive disorder; PD, Parkinson's disease.

Another very important peculiarity of 5-HT, which underscores its disease-influencing potential in ALS and PD, is its neuroprotective role through controlling energy homeostasis via still incompletely characterized mechanisms (Tecott, 2007). As such, new preclinical studies are emerging, which have already shown that 5-HT1A/2B receptor stimulation on astrocytes and microglia promotes proliferation and upregulation of antioxidative molecules, slowing down or even reversing the disease process in ALS (El Oussini et al., 2016), and protecting dopaminergic neurons in PD (Miyazaki et al., 2013).

Why IT Matters?

So far, there is total absence of easily-accessible biological markers in CSF or blood for ALS or PD, rendering the diagnosis of both disease entities sometimes fairly complex, laborious and challenging. Accordingly, the differential diagnosis among similar syndromes, including progressive supranuclear palsy, multiple system atrophy or corticobasal degeneration, may be quite difficult. Future studies should, therefore, focus on the serotonergic dysfunction in ALS and PD, and reveal if serotonergic markers alone or in combination with other biological factors, such as the LDL/HDL ratio, plasma ApoE, or various neuroinflammatory compounds (Dupuis et al., 2010), could be useful for routine diagnostic work-up of patients in clinics.

Additionally, serotonergic approaches in ALS and PD may alleviate disease burden on both the motor and non-motor level, and may hold great potential to influence the disease course, even though clinical trials with 5-HT modulating agents are currently scarce. Hypothetically, other neurodegenerative disorders, such as AD, dementia with Lewy bodies or PD plus syndromes, could—at least in part—share a fundamentally-alike monoaminergic pathophysiology, promoted by very early protein depositions in strategic brainstem nuclei. One might, therefore, wonder whether the universal quest for efficient symptomatic and disease-modifying therapies might, perhaps, be narrowed down to a monoaminergic-based derivative.

Author Contributions

All authors listed have made a substantial, direct, and intellectual contribution to the work, and approved it for publication.

Funding

The authors gratefully acknowledge the support of the Alzheimer Research Foundation Belgium (SAO-FRA; grant P#16003). This work was further supported by the Research Foundation Flanders (FWO), the Institute Born-Bunge, the Medical Research Foundation Antwerp, the Thomas Riellaerts research fund, Neurosearch Antwerp, and the Alzheimer Research Center of the University Medical Center Groningen (ARCG-UMCG).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

Albin, R. L., Young, A. B., and Penney, J. B. (1989). The functional anatomy of basal ganglia disorders. Trends Neurosci. 12, 366–375. doi: 10.1016/0166-2236(89)90074-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Arai, R., Karasawa, N., Geffard, M., and Nagatsu, I. (1995). L-DOPA is converted to dopamine in serotonergic fibers of the striatum of the rat: a double-labeling immunofluorescence study. Neurosci. Lett. 195, 195–198. doi: 10.1016/0304-3940(95)11817-G

PubMed Abstract | CrossRef Full Text | Google Scholar

Bertel, O., Malessa, S., Sluga, E., and Hornykiewicz, O. (1991). Amyotrophic lateral sclerosis: changes of noradrenergic and serotonergic transmitter systems in the spinal cord. Brain Res. 566, 54–60. doi: 10.1016/0006-8993(91)91680-Y

PubMed Abstract | CrossRef Full Text | Google Scholar

Bonaventure, P., Schotte, A., Cras, P., and Leysen, J. E. (1997). Autoradiographic mapping of 5-HT1B- and 5-HT1D receptors in human brain using [3H]alniditan, a new radioligand. Receptors Channels 5, 225–230.

Google Scholar

Bonaventure, P., Voorn, P., Luyten, W. H., Jurzak, M., Schotte, A., and Leysen, J. E. (1998). Detailed mapping of serotonin 5-HT1B and 5-HT1D receptor messenger RNA and ligand binding sites in guinea-pig brain and trigeminal ganglion: clues for function. Neuroscience 82, 469–484. doi: 10.1016/S0306-4522(97)00302-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Borasio, G. D., Linke, R., Schwarz, J., Schlamp, V., Abel, A., Mozley, P. D., et al. (1998). Dopaminergic deficit in amyotrophic lateral sclerosis assessed with [I-123] IPT single photon emission computed tomography. J. Neurol. Neurosurg. Psychiatry 65, 263–265. doi: 10.1136/jnnp.65.2.263

CrossRef Full Text | Google Scholar

Braak, H., Brettschneider, J., Ludolph, A. C., Lee, V. M., Trojanowski, J. Q., and Del Tredici, K. (2013). Amyotrophic lateral sclerosis–a model of corticofugal axonal spread. Nat. Rev. Neurol. 9, 708–714. doi: 10.1038/nrneurol.2013.221

PubMed Abstract | CrossRef Full Text | Google Scholar

Braak, H., Del Tredici, K., Rüb, U., de Vos, R. A., Jansen Steur, E. N., and Braak, E. (2003). Staging of brain pathology related to sporadic Parkinson's disease. Neurobiol. Aging 24, 197–211. doi: 10.1016/S0197-4580(02)00065-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Brettschneider, J., Del Tredici, K., Toledo, J. B., Robinson, J. L., Irwin, D. J., Grossman, M., et al. (2013). Stages of pTDP-43 pathology in amyotrophic lateral sclerosis. Ann. Neurol. 74, 20–38. doi: 10.1002/ana.23937

PubMed Abstract | CrossRef Full Text | Google Scholar

Burnet, P. W., Eastwood, S. L., Lacey, K., and Harrison, P. J. (1995). The distribution of 5-HT1A and 5-HT2A receptor mRNA in human brain. Brain Res. 676, 157–168. doi: 10.1016/0006-8993(95)00104-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Ciranna, L. (2006). Serotonin as a modulator of glutamate- and GABA-mediated neurotransmission: implications in physiological functions and in pathology. Curr. Neuropharmacol. 4, 101–114. doi: 10.2174/157015906776359540

PubMed Abstract | CrossRef Full Text | Google Scholar

Cummings, J., Isaacson, S., Mills, R., Williams, H., Chi-Burris, K., Corbett, A., et al. (2013). Pimavanserin for patients with Parkinson's disease psychosis: a randomised, placebo-controlled phase 3 trial. Lancet 383, 533–540. doi: 10.1016/S0140-6736(13)62106-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Dahlstroem, A., and Fuxe, K. (1964). Evidence for the existence of monoamine-containing neurons in the central nervous system. I. Demonstration of monoamines in the cell bodies of brainstem neurons. Acta Physiol. Scand. Suppl. 232, 1–55.

Google Scholar

De Deurwaerdère, P., and Di Giovanni, G. (2016). Serotonergic modulation of the activity of mesencephalic dopaminergic systems: therapeutic implications. Prog. Neurobiol. 151, 175–236. doi: 10.1016/j.pneurobio.2016.03.004

PubMed Abstract | CrossRef Full Text | Google Scholar

De Deurwaerdère, P., Di Giovanni, G., and Millan, M. J. (2017). Expanding the repertoire of L-DOPA's actions: a comprehensive review of its functional neurochemistry. Prog. Neurobiol. 151, 57–100. doi: 10.1016/j.pneurobio.2016.07.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Delaville, C., Chetrit, J., Abdallah, K., Morin, S., Cardoit, L., De Deurwaerdère, P., et al. (2012). Emerging dysfunctions consequent to combined monoaminergic depletions in Parkinsonism. Neurobiol. Dis. 45, 763–773. doi: 10.1016/j.nbd.2011.10.023

PubMed Abstract | CrossRef Full Text | Google Scholar

Dentel, C., Palamiuc, L., Henriques, A., Lannes, B., Spreux-Varoquaux, O., Gutknecht, L., et al. (2013). Degeneration of serotonergic neurons in amyotrophic lateral sclerosis: a link to spasticity. Brain 136, 483–493. doi: 10.1093/brain/aws274

PubMed Abstract | CrossRef Full Text | Google Scholar

Diaz, S. L., Doly, S., Narboux-Nême, N., Fernández, S., Mazot, P., Banas, S. M., et al. (2012). 5-HT(2B) receptors are required for serotonin-selective antidepressant actions. Mol. Psychiatry 17, 154–163. doi: 10.1038/mp.2011.159

PubMed Abstract | CrossRef Full Text | Google Scholar

Dupuis, L., Spreux-Varoquaux, O., Bensimon, G., Jullien, P., Lacomblez, L., Salachas, F., et al. (2010). Platelet serotonin level predicts survival in amyotrophic lateral sclerosis. PLoS ONE 5:e13346. doi: 10.1371/journal.pone.0013346

PubMed Abstract | CrossRef Full Text | Google Scholar

Duxon, M. S., Kennett, G. A., Lightowler, S., Blackburn, T. P., and Fone, K. C. (1997). Activation of 5-HT2B receptors in the medial amygdala causes anxiolysis in the social interaction test in the rat. Neuropharmacology 36, 601–608. doi: 10.1016/S0028-3908(97)00042-7

PubMed Abstract | CrossRef Full Text | Google Scholar

El Oussini, H., Bayer, H., Scekic-Zahirovic, J., Vercruysse, P., Sinniger, J., Dirrig-Grosch, S., et al. (2016). Serotonin 2B receptor slows disease progression and prevents degeneration of spinal cord mononuclear phagocytes in amyotrophic lateral sclerosis. Acta Neuropathol. 131, 465–480. doi: 10.1007/s00401-016-1534-4

PubMed Abstract | CrossRef Full Text | Google Scholar

El Oussini, H., Scekic-Zahirovic, J., Vercruysse, P., Marques, C., Dirrig-Grosch, S., Dieterlé, S., et al. (2017). Degeneration of serotonin neurons triggers spasticity in amyotrophic lateral sclerosis. Ann. Neurol. 82, 444–456. doi: 10.1002/ana.25030

PubMed Abstract | CrossRef Full Text | Google Scholar

Fahn, S., Libsch, L. R., and Cutler, R. W. (1971). Monoamines in the human neostriatum: topographic distribution in normal and in Parkinson's disease and their role in akinesia, rigidity, chorea, and tremor. J. Neurol. Sci. 14, 427–455. doi: 10.1016/0022-510X(71)90178-X

CrossRef Full Text | Google Scholar

Filip, M., and Bader, M. (2009). Overview on 5-HT-receptors and their role in physiology and pathology of the central nervous system. Pharmacol. Rep. 61, 761–777. doi: 10.1016/S1734-1140(09)70132-X

PubMed Abstract | CrossRef Full Text | Google Scholar

Finberg, J. P. (2014). Update on the pharmacology of selective inhibitors of MAO-A and MAO-B: focus on modulation of CNS monoamine neurotransmitter release. Pharmacol. Ther. 143, 133–152. doi: 10.1016/j.pharmthera.2014.02.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Fomina, T., Weichwald, S., Synofzik, M., Just, J., Schöls, L., Schölkopf, B., et al. (2017). Absence of EEG correlates of self-referential processing depth in ALS. PLoS ONE 12:e0180136. doi: 10.1371/journal.pone.0180136

PubMed Abstract | CrossRef Full Text | Google Scholar

Goldstein, L. H., Atkins, L., Landau, S., Brown, R. G., and Leigh, P. N. (2006). Longitudinal predictors of psychological distress and self-esteem in people with ALS. Neurology 67, 1652–1658. doi: 10.1212/01.wnl.0000242886.91786.47

PubMed Abstract | CrossRef Full Text | Google Scholar

Guttman, M., Boileau, I., Warsh, J., Saint-Cyr, J. A., Ginovart, N., McCluskey, T., et al. (2007). Brain serotonin transporter binding in non-depressed patients with Parkinson's disease. Eur. J. Neurol. 14, 523–528. doi: 10.1111/j.1468-1331.2007.01727.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hahn, A., Wadsak, W., Windischberger, C., Baldinger, P., Höflich, A. S., Losak, J., et al. (2012). Differential modulation of the default mode network via serotonin-1A receptors. Proc. Natl. Acad. Sci. U.S.A. 109, 2619–2624. doi: 10.1073/pnas.1117104109

PubMed Abstract | CrossRef Full Text | Google Scholar

Halliday, G. M., Blumbergs, P. C., Cotton, R. G., Blessing, W. W., and Geffen, L. B. (1990). Loss of brainstem serotonin- and substance P-containing neurons in Parkinson's disease. Brain Res. 510, 104–107. doi: 10.1016/0006-8993(90)90733-R

PubMed Abstract | CrossRef Full Text | Google Scholar

Hannon, J., and Hoyer, D. (2008). Molecular biology of 5-HT receptors. Behav. Brain Res. 195, 198–213. doi: 10.1016/j.bbr.2008.03.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Hardiman, O., and van den Bergh, L. H. (2017). Edaravone: a new treatment for ALS on the horizon? Lancet Neurol. 16, 490–491. doi: 10.1016/S1474-4422(17)30163-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoyer, D., Clarke, D. E., Fozard, J. R., Hartig, P. R., Martin, G. R., Mylecharane, E. J., et al. (1994). International Union of pharmacology classification of receptors for 5-hydroxytryptamine (Serotonin). Pharmacol. Rev. 46, 157–203.

PubMed Abstract | Google Scholar

Huot, P., Sgambato-Faure, V., Fox, S. H., and McCreary, A. C. (2017). Serotonergic approaches in Parkinson's disease: translational perspectives, an update. ACS Chem. Neurosci. 8, 973–986. doi: 10.1021/acschemneuro.6b00440

PubMed Abstract | CrossRef Full Text | Google Scholar

Jacobs, B. L., and Azmitia, E. C. (1992). Structure and function of the brain serotonin system. Physiol. Rev. 72, 165–229. doi: 10.1152/physrev.1992.72.1.165

PubMed Abstract | CrossRef Full Text | Google Scholar

Jossan, S. S., Ekblom, J., Gudjonsson, O., Hagbarth, K. E., and Aquilonius, S. M. (1994). Double blind cross over trial with deprenyl in amyotrophic lateral sclerosis. J. Neural. Transm. Suppl. 41, 237–241. doi: 10.1007/978-3-7091-9324-2_30

PubMed Abstract | CrossRef Full Text | Google Scholar

Kish, S. J., Tong, J., Hornykiewicz, O., Rajput, A., Chang, L. J., Guttman, M., et al. (2008). Preferential loss of serotonin markers in caudate versus putamen in Parkinson's disease. Brain 131, 120–131. doi: 10.1093/brain/awm239

PubMed Abstract | CrossRef Full Text | Google Scholar

Lange, D. J., Murphy, P. L., Diamond, B., Appel, V., Lai, E. C., Younger, D. S., et al. (1998). Selegiline is ineffective in a collaborative double-blind, placebo-controlled trial for treatment of amyotrophic lateral sclerosis. Arch. Neurol. 55, 93–96. doi: 10.1001/archneur.55.1.93

PubMed Abstract | CrossRef Full Text | Google Scholar

Lauder, J. M. (1990). Ontogeny of the serotonergic system in the rat: serotonin as a developmental signal. Ann. N. Y. Acad. Sci. 600, 297–313. doi: 10.1111/j.1749-6632.1990.tb16891.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemon, R. N. (2008). Descending pathways in motor control. Annu. Rev. Neurosci. 31, 195–218. doi: 10.1146/annurev.neuro.31.060407.125547

PubMed Abstract | CrossRef Full Text | Google Scholar

Macchi, Z., Wang, Y., Moore, D., Katz, J., Saperstein, D., and Walk, D. (2015). A multi-center screening trial of rasagiline in patients with amyotrophic lateral sclerosis: possible mitochondrial biomarker target engagement. Amyotroph. Lateral Scler. Frontotemporal Degener. 16, 345–352. doi: 10.3109/21678421.2015.1026826

PubMed Abstract | CrossRef Full Text | Google Scholar

Meininger, V., Bensimon, G., Bradley, W. R., Brooks, B., Douillet, P., Eisen, A. A., et al. (2004). Efficacy and safety of xaliproden in amyotrophic lateral sclerosis: results of two phase III trials. Amyotroph. Lateral Scler. Other Motor Neuron Disord. 5, 107–117. doi: 10.1080/14660820410019602

PubMed Abstract | CrossRef Full Text | Google Scholar

Miguelez, C., Morera-Herreras, T., Torrecilla, M., Ruiz-Ortega, J. A., and Ugedo, L. (2014). Interaction between the 5-HT system and the basal ganglia: functional implication and therapeutic perspective in Parkinson's disease. Front. Neural Circuits 8:21. doi: 10.3389/fncir.2014.00021

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, R. G., Mitchell, J. D., Lyon, M., and Moore, D. H. (2007). Riluzole for amyotrophic lateral sclerosis (ALS)/motor neuron disease (MND). Cochrane Database Syst. Rev. 1:CD001447. doi: 10.1002/14651858.CD001447.pub2

CrossRef Full Text | Google Scholar

Miyazaki, I., Asanuma, M., Murakami, S., Takeshima, M., Torigoe, N., Kitamura, Y., et al. (2013). Targeting 5-HT(1A) receptors in astrocytes to protect dopaminergic neurons in Parkinsonian models. Neurobiol. Dis. 59, 244–256. doi: 10.1016/j.nbd.2013.08.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Monaco, F., Fumero, S., Mondino, A., and Mutani, R. (1979). Plasma and cerebrospinal fluid tryptophan in multiple sclerosis and degenerative diseases. J. Neurol. Neurosurg. Psychiatry 42, 640–641. doi: 10.1136/jnnp.42.7.640

PubMed Abstract | CrossRef Full Text | Google Scholar

Muramatsu, M., Lapiz, M. D., Tanaka, E., and Grenhoff, J. (1998). Serotonin inhibits synaptic glutamate currents in rat nucleus accumbens neurons via presynaptic 5-HT1B receptors. Eur. J. Neurosci 10, 2371–2379. doi: 10.1046/j.1460-9568.1998.00248.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Murray, K. C., Nakae, A., Stephens, M. J., Rank, M., D'Amico, J., Harvey, P. J., et al. (2010). Recovery of motoneuron and locomotor function after spinal cord injury depends on constitutive activity in 5-HT2C receptors. Nat. Med. 16, 694–700. doi: 10.1038/nm.2160

PubMed Abstract | CrossRef Full Text | Google Scholar

Navailles, S., and De Deurwaerdère, P. (2012). Imbalanced dopaminergic transmission mediated by serotonergic neurons in L-DOPA-induced dyskinesia. Parkinsons Dis. 2012:323686. doi: 10.1155/2012/323686

PubMed Abstract | CrossRef Full Text | Google Scholar

Ohsugi, K., Adachi, K., Mukoyama, M., and Ando, K. (1987). Lack of change in the indoleamine metabolism in spinal cord of patients with amyotrophic lateral sclerosis. Neurosci. Lett. 79, 351–354. doi: 10.1016/0304-3940(87)90458-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Paolucci, E., Berretta, N., Tozzi, A., Bernardi, G., and Mercuri, N. B. (2003). Depression of mGluR-mediated IPSCs by 5-HT in dopamine neurons of the rat substantia nigra pars compacta. Eur. J. Neurosci. 18, 2743–2750. doi: 10.1111/j.1460-9568.2003.03015.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Pasqualetti, M., Ori, M., Castagna, M., Marazziti, D., Cassano, G. B., and Nardi, I. (1999). Distribution and cellular localization of the serotonin type 2C receptor messenger RNA in human brain. Neuroscience 92, 601–611. doi: 10.1016/S0306-4522(99)00011-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Politis, M., and Niccolini, F. (2015). Serotonin in Parkinson's disease. Behav. Brain Res. 277, 136–145. doi: 10.1016/j.bbr.2014.07.037

PubMed Abstract | CrossRef Full Text | Google Scholar

Politis, M., Wu, K., Loane, C., Kiferle, L., Molloy, S., Brooks, D. J., et al. (2010). Staging of serotonergic dysfunction in Parkinson's disease: an in vivo 11C-DASB PET study. Neurobiol. Dis. 40, 216–221. doi: 10.1016/j.nbd.2010.05.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Rascol, O., Brooks, D. J., Melamed, E., Oertel, W., Poewe, W., Stocchi, F., et al. (2005). Rasagiline as an adjunct to levodopa in patients with Parkinson's disease and motor fluctuations (LARGO, Lasting effect in Adjunct therapy with Rasagiline Given Once daily, study): a randomised, double-blind, parallel-group trial. Lancet 365, 947–954. doi: 10.1016/S0140-6736(05)71083-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Rascol, O., Fitzer-Attas, C. J., Hauser, R., Jankovic, J., Lang, A., Langston, J. W., et al. (2011). A double-blind, delayed-start trial of rasagiline in Parkinson's disease (the ADAGIO study): prespecified and post-hoc analyses of the need for additional therapies, changes in UPDRS scores, and non-motor outcomes. Lancet Neurol. 10, 415–423. doi: 10.1016/S1474-4422(11)70073-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Reyes, E. T., Perurena, O. H., Festoff, B. W., Jorgensen, R., and Moore, W. V. (1984). Insulin resistance in amyotrophic lateral sclerosis. J. Neurol. Sci. 63, 317–324. doi: 10.1016/0022-510X(84)90154-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Rodríguez, J. J., Noristani, H. N., and Verkhratsky, A. (2012). The serotonergic system in ageing and Alzheimer's disease. Prog. Neurobiol. 99, 15–41. doi: 10.1016/j.pneurobio.2012.06.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Sandyk, R. (2006). Serotonergic mechanisms in amyotrophic lateral sclerosis. Int. J. Neurosci. 116, 775–826. doi: 10.1080/00207450600754087

PubMed Abstract | CrossRef Full Text | Google Scholar

Sargent, P. A., Kjaer, K. H., Bench, C. J., Rabiner, E. A., Messa, C., Meyer, J., et al. (2000). Brain serotonin1A receptor binding measured by positron emission tomography with [11C]WAY-100635: effects of depression and antidepressant treatment. Arch. Gen. Psychiatry 57, 174–180. doi: 10.1001/archpsyc.57.2.174

CrossRef Full Text | Google Scholar

Scatton, B., Javoy-Agid, F., Rouquier, L., Dubois, B., and Agid, Y. (1983). Reduction of cortical dopamine, noradrenaline, serotonin and their metabolites in Parkinson's disease. Brain Res. 275, 321–328. doi: 10.1016/0006-8993(83)90993-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Shannak, K., Rajput, A., Rozdilsky, B., Kish, S., Gilbert, J., and Hornykiewicz, O. (1994). Noradrenaline, dopamine and serotonin levels and metabolism in the human hypothalamus: observations in Parkinson's disease and normal subjects. Brain Res. 639, 33–41. doi: 10.1016/0006-8993(94)91761-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Sofic, E., Riederer, P., Gsell, W., Gavranovic, M., Schmidtke, A., and Jellinger, K. (1991). Biogenic amines and metabolites in spinal cord of patients with Parkinson's disease and amyotrophic lateral sclerosis. J. Neural Transm. 3, 133–142. doi: 10.1007/BF02260888

PubMed Abstract | CrossRef Full Text | Google Scholar

Svenningsson, P., Rosenblad, C., Af Edholm Arvidsson, K., Wictorin, K., Keywood, C., Shankar, B., et al. (2015). Eltoprazine counteracts L-DOPA-induced dyskinesias in Parkinson's disease: a dose-finding study. Brain 138, 963–973. doi: 10.1093/brain/awu409

PubMed Abstract | CrossRef Full Text | Google Scholar

Tecott, L. H. (2007). Serotonin and the orchestration of energy balance. Cell Metab. 6, 352–361. doi: 10.1016/j.cmet.2007.09.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Tronci, E., Lisci, C., Stancampiano, R., Fidalgo, C., Collu, M., Devoto, P., et al. (2013). 5-Hydroxy-tryptophan for the treatment of L-DOPA-induced dyskinesia in the rat Parkinson's disease model. Neurobiol. Dis. 60, 108–114. doi: 10.1016/j.nbd.2013.08.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Turner, B. J., Lopes, E. C., and Cheema, S. S. (2003). The serotonin precursor 5-hydroxytryptophan delays neuromuscular disease in murine familial amyotrophic lateral sclerosis. Amyotroph. Lateral Scler. Other Motor Neuron Disord. 4, 171–176. doi: 10.1080/14660820310009389

PubMed Abstract | CrossRef Full Text | Google Scholar

Turner, M. R., Rabiner, E. A., Hammers, A., Al-Chalabi, A., Grasby, P. M., Shaw, C. E., et al. (2005). [11C]-WAY100635 PET demonstrates marked 5-HT1A receptor changes in sporadic ALS. Brain 128, 896–905. doi: 10.1093/brain/awh428

PubMed Abstract | CrossRef Full Text | Google Scholar

Vercruysse, P., Sinniger, J., El Oussini, H., Scekic-Zahirovic, J., Dieterlé, S., Dengler, R., et al. (2016). Alterations in the hypothalamic melanocortin pathway in amyotrophic lateral sclerosis. Brain 139, 1106–1122. doi: 10.1093/brain/aww004

PubMed Abstract | CrossRef Full Text | Google Scholar

Vermeiren, Y., and De Deyn, P. P. (2017). Targeting the norepinephrinergic system in Parkinson's disease and related disorders: the locus coeruleus story. Neurochem. Int. 102, 22–32. doi: 10.1016/j.neuint.2016.11.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Vogels, O. J., Oyen, W. J., van Engelen, B. G., Padberg, G. W., and Horstink, M. W. (1999). Decreased striatal dopamine-receptor binding in sporadic ALS: glutamate hyperactivity? Neurology 52, 1275–1277. doi: 10.1212/WNL.52.6.1275

PubMed Abstract | CrossRef Full Text | Google Scholar

Vucic, S., Ziemann, U., Eisen, A., Hallett, M., and Kiernan, M. C. (2013). Transcranial magnetic stimulation and amyotrophic lateral sclerosis: pathophysiological insights. J. Neurol. Neurosurg. Psychiatry 84, 1161–1170. doi: 10.1136/jnnp-2012-304019

PubMed Abstract | CrossRef Full Text | Google Scholar

White, S. R., and Neuman, R. S. (1980). Facilitation of spinal motoneurone excitability by 5-hydroxytryptamine and noradrenaline. Brain Res. 188, 119–127. doi: 10.1016/0006-8993(80)90561-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, J., Li, H., Li, C., Yao, J. C., Hu, J., Wang, J., et al. (2017). Abnormal cortical-basal ganglia network in amyotrophic lateral sclerosis: a voxel-wise network efficiency analysis. Behav. Brain Res. 333, 123–128. doi: 10.1016/j.bbr.2017.06.050

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Q. Z., Gong, Y. S., and Zhang, J. T. (1999). Antagonistic effects of melatonin on glutamate release and neurotoxicity in cerebral cortex. Zhongguo Yao Li Xue Bao 20, 829–834.

PubMed Abstract | Google Scholar

Ziegler, M. G., Brooks, B. R., Lake, C. R., Wood, J. H., and Enna, S. J. (1980). Norepinephrine and gamma-aminobutyric acid in amyotrophic lateral sclerosis. Neurology 30, 98–101. doi: 10.1212/WNL.30.1.98

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: serotonin (5-HT), dopamine, glutamate, amyotrophic lateral sclerosis (ALS), Parkinson's disease (PD), raphe nuclei, basal ganglia

Citation: Vermeiren Y, Janssens J, Van Dam D and De Deyn PP (2018) Serotonergic Dysfunction in Amyotrophic Lateral Sclerosis and Parkinson's Disease: Similar Mechanisms, Dissimilar Outcomes. Front. Neurosci. 12:185. doi: 10.3389/fnins.2018.00185

Received: 21 December 2017; Accepted: 06 March 2018;
Published: 20 March 2018.

Edited by:

Tibor Hortobágyi, University of Debrecen, Hungary

Reviewed by:

Luc Dupuis, INSERM U1118, France
Patrizia Longone, Fondazione Santa Lucia (IRCCS), Italy
Philippe De Deurwaerdere, Université de Bordeaux, France

Copyright © 2018 Vermeiren, Janssens, Van Dam and De Deyn. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Yannick Vermeiren, yannick.vermeiren@uantwerpen.be
Peter P. De Deyn, dedeyn@skynet.be

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.