Next Article in Journal
Carbon Dots in Treatment of Pediatric Brain Tumors: Past, Present, and Future Directions
Previous Article in Journal
The Exosome-like Vesicles of Giardia Assemblages A, B, and E Are Involved in the Delivering of Distinct Small RNA from Parasite to Parasite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ganglioside GM1 and the Central Nervous System

Department of Chemistry, University of Florida, Gainesville, FL 32611, USA
Int. J. Mol. Sci. 2023, 24(11), 9558; https://doi.org/10.3390/ijms24119558
Submission received: 14 March 2023 / Revised: 18 April 2023 / Accepted: 4 May 2023 / Published: 31 May 2023
(This article belongs to the Special Issue Glycome and Nervous System 2.0)

Abstract

:
GM1 is one of the major glycosphingolipids (GSLs) on the cell surface in the central nervous system (CNS). Its expression level, distribution pattern, and lipid composition are dependent upon cell and tissue type, developmental stage, and disease state, which suggests a potentially broad spectrum of functions of GM1 in various neurological and neuropathological processes. The major focus of this review is the roles that GM1 plays in the development and activities of brains, such as cell differentiation, neuritogenesis, neuroregeneration, signal transducing, memory, and cognition, as well as the molecular basis and mechanisms for these functions. Overall, GM1 is protective for the CNS. Additionally, this review has also examined the relationships between GM1 and neurological disorders, such as Alzheimer’s disease, Parkinson’s disease, GM1 gangliosidosis, Huntington’s disease, epilepsy and seizure, amyotrophic lateral sclerosis, depression, alcohol dependence, etc., and the functional roles and therapeutic applications of GM1 in these disorders. Finally, current obstacles that hinder more in-depth investigations and understanding of GM1 and the future directions in this field are discussed.

1. Introduction

GM1 (or GM1a) is a glycosphingolipid (GSL) consisting of a pentasaccharide and a ceramide (Cer), which are coupled together by a β-glycosidic linkage (Figure 1). GM1 is one of the gangliosides, i.e., sialylated ganglio-series GSLs, since its glycan contains an N-acetylneuraminic acid (Neu5Ac) residue, which is the most common member of the family of sialic acids and thus is frequently referred to as the “sialic acid”. GM1 is usually anchored onto the cell surface through embedding its lipid chains into the plasma membrane bilayer, leaving the glycan as the outer leaflet (Figure 1). GM1 is one of the main GSLs in the plasma membrane of the vertebrate central nervous system (CNS) to play a key role in various functions of the CNS [1]. GM1 is also abundant in peripheral tissues closely associated with many other physiological and pathological processes [2,3,4,5]. However, this article is focused on the research progress related to the distribution and function of GM1 in the CNS, as well as its relationships with various CNS-related diseases.

2. GM1 Metabolism

Like other gangliosides, GM1 is biosynthesized in the endoplasmic reticulum (ER) and Golgi [6]. After Cer is generated in the ER [7], it is transported to the Golgi, where sugar residues are added sequentially by a set of membrane-associated glycosyltransferases (GTs) to furnish GM1 (Figure 2). GM3 is the common intermediate for the synthesis of all gangliosides, whereas GM1 is the immediate precursor of GD1a—another important ganglioside. Besides de novo biosynthesis, GSLs are also generated by salvage pathways [8,9]. For example, GSLs are endocytosed and degraded to monosaccharides, sphingoids, fatty acids, and simpler GSLs such as glucosylceramide (GlcCer) and galactosylceramide (GalCer), which are salvaged by the de novo biosynthetic pathway. Alternatively, after endocytosis, GSLs are directly recycled or remodeled in the Golgi before re-presentation onto the cell surface [10,11]. It has been shown that 80% of the sphingosine derived from 3H-labeld GM3 was metabolically incorporated by cells [12], and fluorescent N-[6-(7-nitrobnez-2-oxa-1,3-diazol-4-yl)aminohexanoyl]-GM1 is directly recycled by hippocampal neurons [13]. In addition, in some cells, GSLs are also remodeled by GTs [14], sialidases [15,16], and other enzymes [17] on the plasma membrane.
Ganglioside degradation is achieved by stepwise enzymatic hydrolysis of their glycans starting from the non-reducing end (Figure 3). Sialidase or neuraminidase (Neu-ase)-catalyzed desialylation is the first step of ganglioside degradation, occurring both on the cell surface and in the endosomes and lysosomes [6,18], where GM1 is the common intermediate. Cell surface GM1 can be endocytosed for further degradation in lysosomes [19]. The enzymes engaged in GM1 hydrolysis include Neu-ases, β-galactosidase (βGal-ase), β-hexosaminidase (βHex-ase), glucocerebrosidase (GCase), and ceramidase [20], which are typically water-soluble hydrolases. To prompt the interaction of these enzymes with membrane-associated GSLs, sphingolipid activator proteins (SAPs) are used as enzyme cofactors (Figure 3) to help extract GSLs from the membrane [6,21].
The expression levels and distribution patterns of GM1 and other gangliosides in the CNS are affected by many factors. First, the activity and location of enzymes engaged in GSL synthesis and degradation have a direct impact on GSL metabolism. Thus, development-associated changes in GSL distribution are typically the results of reprogramed enzyme expression [22]. For example, it has been shown that epigenetic activation of the GalNAc-T gene leads to GM1-induced neural stem cell (NSC) differentiation, whereas nuclear GM1 binds acetylated histones on the promoter of the GalNAc-T gene in differentiated neurons [23]. Second, forming stable clusters of various GTs is also important for the effective biosynthesis of GSLs [24,25]. Third, the levels of enzyme substrates, e.g., sugar nucleotides as glycosyl donors for GTs [26], in cells can affect glycosylation reactions and GSL synthesis. In addition, the cell membrane composition and fluidity, external stimuli, and flow rates of different GSLs through the Golgi are other factors that may also affect GSL metabolism and distribution [6,27,28]. Overall, various GTs and hydrolases form a complex and effective network in the control of GSL expression and distribution and keep them in check during delicate neurological activity. As a result, any changes in these factors, such as enzyme activity, substrate level, or environment, may break the dynamic equilibrium of GSLs and result in functional disorders.

3. GM1 Distribution in the Central Nervous System

It has been well established that GSLs are a major component of the cell membrane and are especially rich in the CNS. For example, >80% of the glycans in the glycocalyx of brain cell are GSLs [29,30]. Moreover, GSL expression levels and patterns are cell-, tissue-, species-, and development-specific [31]. In vertebrate brains, gangliosides are predominant, accounting for >75% of total sialic acids [29], which is 10–30 times higher than that in other tissues [32]. Overall, the expression level and structural complexity of GSLs increase with brain development but start to decrease during old age [33]. In the adult human brain, GM1, GD1a, GD1b, and GT1b are extremely enriched, representing >97% of all gangliosides [34,35]. Studies have shown that they are also the major gangliosides in the brain of many other animals [36,37,38,39,40,41,42,43]. In addition, the Cer structure, i.e., the lipid form, of GSLs is also cell-, tissue-, species-, and development-dependent [29,30].
Specifically, dramatic changes in the expression levels and patterns of gangliosides have been associated with the development of rat and mouse brains. Simple gangliosides (GM3 and GD3, Figure 4) are rich in the early embryonic stages, but more complex gangliosides (GM1, GD1a, GD1b and GT1b, Figure 4) are predominant in the late developmental stages of the brain [38,39,40,41,42,43]. In rats, the decrease in GM3 and GD3 and the increase in GM1, GD1a and GT1a are correlated with the increase of GalNAc-T that transforms GM3 into GM2, and at birth, GM3 and GD3 are already minor components in the rat brain. During brain maturation, the total ganglioside content can increase several fold [30]. However, the total ganglioside content of the mouse brain was almost constant from young adulthood to the beginning of senescence, despite the changes in their composition (e.g., an increase in GM1 and decrease in GD1b, GT1b and GQ1b). During senescence, however, the concentration of total gangliosides may decrease by 20% [44]. These results suggest that GSL expression is age-related.
There are also differences in the ganglioside profile in various brain regions. For example, gangliosides in the cortex, olfactory bulb, corpora quadrigemina, cerebellum, thalamic region, and medulla of the rat brain are distinct, and each region has its own unique development markers [45,46]. GQ1b and GP1 are the markers of neural cell sprouting and arborization; GD1a and GM1 are synaptogenesis and myelination markers, respectively. In the adult mouse brain, GD1b and GT1b are abundant in gray and white matter, and GM1 is dominant in white matter [42]. GD1a is the major ganglioside in the hippocampus, cortex, and caudate nucleus, but GT1b is the main species in the cerebellum, hypothalamus, and pons–medulla [47]. Similarly, significant differences in regional ganglioside distribution were also observed in newborn and adult rats. Newborns have lower ganglioside contents in their forebrain (2.5 times), cerebellum (2.0 times), and brain stem (2.0 times) as compared to the adults [48]. At the cellular level, GM1 was detected on the cell soma, while GD3 was found on non-neuronal cells but not on neuronal cells. Furthermore, a substantial quantity of gangliosides were transformed into more complex products upon neurite outgrowth and morphological changes [49].
Qualitative and quantitative studies of postmortem human brains have also revealed dramatic and steady changes in GSL profiles during neurodevelopment and neuropathogenesis [38,50,51]. For example, fetal human brains showed a stage-specific expression of GSLs [52,53]. At gestational week 10, GT1b, GD1b, and GD3 were predominant, but their ratios were reduced at week 22. GD1a and GM1 expression increased markedly between weeks 12 and 14, slowly between weeks 14 and 18, and rapidly again from week 20. A study on the human frontal lobe over a period from 10 fetal weeks to 80 years of age showed an increase in both the concentration and structural complexity of gangliosides during brain development [54]. The ganglioside content increased by roughly a factor of three from gestational week 10 to 5 years old, whereas GM1 and GD1a increased by a factor of 12–15 during the same period. The most rapid increases of GM1 and GD1a occurred around term. Other gangliosides have exhibited different developmental curves. Another study of adult human brains has demonstrated a continuous change in ganglioside composition between 20 and 70 years of age, namely a decrease in GM1 and GD1a and an increase in GD1b, GM3, and GD3 [55]. A composition of gangliosides in cortical regions further revealed region-specific changes associated with aging. In the frontal cortex, GD1a and GM1 declined with aging but no significant change was observed in the visual cortex. GD1b and GT1b in the cerebellar cortex were reduced with aging, but it was GD1a that showed a moderate decrease in hippocampus [56,57].
In vitro studies using cultured hippocampal neurons revealed that complex gangliosides such as GM1, GD1a, GD1b and GT1b were drastically increased during axonogenesis, accompanied by a decrease in GD3 and GlcCer [58]. Investigation of different types of brain cells has demonstrated the great diversity of gangliosides in their expression and distribution patterns [59]. For example, while GM1 and GM4 were found on all oligodendrocytes and most astrocytes (80%) and Schwann cells (50–70%), only GM1, not GM4, was observed on most neurons (80%) and a fraction of fibroblasts (5–10%).
The structure and distribution of Cer in GSLs are also tissue-, species-, and development-specific, and this is reflected in both the total lipids and individual GSLs [60]. While ceramides consisting of sphingosine (d18:1) and (d20:1) are predominant in mammalian brains, the constitution of sphingosine (d20:1) increases with brain development and aging [60,61]. For example, the fetal human brain contains <3% sphingosine (d20:1) [62,63,64], but it increases to roughly 50% by 30 years of age. While all (d20:1) gangliosides remarkably and progressively increase with aging, this trend is more obvious for GD1b, GT1b, GQ1b, GM1, and GD1a [64,65,66,67]. Mass spectrometry (MS) imaging of 11 regions of rat brains across their lifespan revealed an age-dependent increase in GM1 and GD1 (d20:1), especially in the early developmental stage, and a decrease in the ratio of sphingosine (d20:1)/(d18:1) for GM2 and GM3 [68]. Similar results were also observed in rats [69,70]. Interestingly, GM1 was the only ganglioside that has a higher sphingosine (d20:1) ratio in the synaptosomal fraction than in the myelin fraction. Furthermore, exogenous GM1 (d18:1) can be inserted into the cell membrane 2.43 times more than GM1 (d20:1) [71]. In regard to the fatty acyl component of Cer, palmitic (C16:0) and stearic (C18:0) groups are the most common, but the fatty acyl chain can vary from C14 to C30 [62,72]. In humans, it was found that the stearic component in GSLs decreased from 93% at birth to 78% at 98 years old, whereas C20 acyl chains increased from 3 to 9% with the most obvious changes found after 60 years of age [73]. In addition to the length of the fatty acyl group, its structure also changes [74], namely in different degrees of saturation, which can have a significant impact on the behavior and function of GSLs. As a results, brain development and neurological disorders are often accompanied by structural changes in the Cer of GSLs [75,76,77].
In summary, in the early stages of brain development, simple gangliosides such as GM3 and GD3 are expressed abundantly [66,78,79,80]. With brain maturation, a broader spectrum of more complex gangliosides is expressed, which is correlated with the increasingly sophisticated functions of the brain. In the brains of all adult mammals and birds, GM1, GD1a, GD1b, and GT1b are predominant (Figure 5) [30,35]. These four major gangliosides also exhibit tissue specificity, with GM1 and GD1a as the most abundant gangliosides in the white matter and brain nuclei and tracts, respectively [42]. The ganglioside pattern evolves continuously, and the overall trend is a gradual decrease in GM1 and GD1a and increase in GD1b, GM3, and GD3 alongside aging [81]. In the meantime, the lipid structure of GSLs is also diverse and dependent on tissues and development, which may have an impact on the organization of GSLs on cells and the fluidity or other biophysical properties of the cell membrane, thereby affecting the biological activities of GSLs. For example, it has been shown that an increase in the ratio of C20 fatty acid-containing GM1 and GD1 in the hippocampus may be associated with altered membrane fluidity and other properties and increased susceptibility of cells to degeneration during aging [60]. Similarly, lipid rafts isolated from neurons aged in vitro exhibited an increase in Cer content and a decreased level of gangliosides [82].

4. Biological Functions of GM1 in the Central Nervous System

GSLs are abundant, structurally diverse, and tissue-/development-dependent in the CNS, indicating that they may play a role in the regulation of neurological processes. Specific sets of gangliosides are associated with particular development stages of the brain, suggesting that they may mediate neuronal differentiation and growth [83,84]. For example, GM1 can interact with its receptors on the cell surface to activate processes such as neuronal migration, dendrite emission, neonatal cortical development, and axon growth [85,86]. Thus, GM1 is a differential marker of neurons [87]. GM1 and GQlb, but not other gangliosides, have been shown to boost neuritogenesis in neuroblastoma cells [88,89], and GM1 can accelerate neurite outgrowth from primary peripheral and central neurons [90]. GM1, GM3, and ganglioside mixtures were found to reduce the baseline rate of proliferation of Schwann cells and their response to mitogens [91].
The mechanism by which GM1 mediates neuron differentiation and growth is still not clear. However, it is generally accepted that GSLs on the cell surface may interact with specific receptors to generate signals that initiate appropriate cellular responses [83]. The interaction of GM1 with Ca2+ or membrane proteins is one of the potential mechanisms of GM1 to activate neuronal differentiation and development [92], which is supported by the anionic structure of the Neu5Ac residue that binds Ca2+ and GM1 co-localization with the Ca2+ pump. It was also revealed that, in the rat brain, GM1, GT1b, and GD1a could stimulate Ca2+/calmodulin-dependent protein kinase II (CaM-kinase II), which is independent of Ca2+ [93]. However, at high concentrations, GT1b, GD1a, and GM1 inhibited CaM-kinase II activity in a substrate-dependent manner. Moreover, the ability of GM1 to enhance PC12 cell fiber outgrowth involves nerve growth factor (NGF) [94], which is a key neurotrophic factor regulating neuronal proliferation, growth, maintenance, and survival [95]. Among all gangliosides that can affect the activity of NGF, GM1 is the most important. It can potentiate NGF both in vitro and in vivo [96,97] to prevent vinblastine-induced sympathectomy in newborn rats [98] and to improve the performance of cortically lesioned rats [99]. GM1 activates NGF through interacting with an NGF-associated tropomyosin receptor kinase (Trk) to promote tyrosine phosphorylation [100,101].
Many studies have revealed that GSLs can regulate signal transduction across the cell membrane via interacting with membrane proteins [102,103]. GSLs can bind proteins on the same or opposite cells and in the extracellular matrix to affect protein conformation, receptor aggregation, and receptor interaction with partners in or on the cell membrane, thereby generating binding signals. For example, GM1 interaction with TrkA was shown to lead to the dimerization and activation of this kinase and then initiation of downstream activities [100,104]. GM1 is a co-receptor of fibroblast growth factor 2 (FGF2), required for FGF2 activation [105]. GM1 also possesses neuroprotective and neurorestorative activities for several neurotransmitter systems [106] through binding neurotransmitters released at the synapse to help their presentation to high-affinity receptors [107,108,109]. GM1-mediated signaling is associated with various kinases and phosphorylation processes. For example, GM1 has been shown to trigger the phosphorylation and activation of Trk and Erk receptors [110,111] and phosphatidylinositol 3-kinase [104] and to promote neurotrophin-3 release through stimulating TrkC autophosphorylation. Changes in the endogenous GM1 density around TrkB have been shown to affect the activity of TrkB [112]. The molecular basis for the interaction between GM1 and kinases is not clear, but its glycan is important as GM1 oligosaccharide alone had a similar activity to GM1 [85,113].
One of the fundamental functions of GM1 in the CNS is to protect the nervous tissues against various toxic agents or conditions and maintain neuronal activity [114]. For example, GM1 has been shown to prevent neuronal apoptosis [115], neurodegeneration [116], and neuronal functional decay [117], as well as glutamate- and ethanol-caused neurotoxicity [118,119,120], anoxia-caused neuronal death [121], and cerebral or retinal ischemia- and trauma-induced brain and neural injuries [122,123,124,125]. GM1 has been shown to provide neural protection by a variety of mechanisms. For example, GM1 prevents apoptotic neuronal death through Trk neurotrophin receptors [126]; ketamine-induced hippocampus and cortex apoptosis and cognitive impairment through the PI3K/AKT/GSK3β pathway [127,128]; lead-caused cognitive deficit and brain damages by averting changes in superoxide dismutase and intracellular Ca2+ levels [129] and activating the SIRT1/CREB/BDNF pathway [130]; high altitude-caused cerebral edema by reducing oxidative stress and PI3K/AKT-Nrf2-linked inflammatory response [131]; and cognitive deficits induced by trauma [132] and MK801 [133]. GM1 has also been shown to protect the CNS via eliciting anti-inflammatory activity [134] and brain spectrin–aminophospholipid interaction [135], stabilizing plasma membrane Ca2+ ATPase-neuroplastin complexes [136], interacting with acetylated histones to regulate gene expressions [137], and modulating neuronal plasticity [138]. In addition, it has been found that GM1 can protect brains from neurotoxin-induced inhibition of choline uptake [139], diminish cell death caused by excessive excitatory neurotransmission [140], and reduce amyloid beta (Aβ)-related cognitive deficits and ischemia-induced brain damages by modulating Na+/K+- and Mg2+-ATPases [141,142,143].
GSLs play a critical role in the processes of neural cell differentiation, neurite sprouting and growth, and nerve regeneration [83,84]. However, the abilities of various GSLs to promote neuritogenesis and nerve regeneration vary dramatically. For example, GT1b and GD1b are stronger than GM1. These results suggest the potential involvement of GSL glycans in these processes [144]. The mechanisms by which GM1 influence neuritogenesis and nerve regeneration are not clear but are believed to be related to specific neurotrophic and growth factors. In rats, GM1 can associate with Trk to induce tyrosine autophosphorylation and enhance PC12 cell neurite outgrowth and neurofilament expression elicited by NGF, while NGF alone is insufficient to induce such activity [100]. It was also shown that the neurotrophic activity of GM1 was due to its activation of RET tyrosine kinase receptors via interaction with GFRα1 co-receptors in the glia cell-derived neurotrophic factor (GDNF) receptor complex [145]. These findings suggest the possibility of using GSLs to promote nerve regeneration and treat nervous injuries and neurodegenerative disorders. Indeed, GM1 has shown a positive effect on cerebrovascular [146] and Parkinson’s disease (PD) patients [147] and brain trauma in mice [132]. Recently, it was also reported that intracerebroventricular (ICV) infusion of exogenous gangliosides, such as GD3 and GM1, can promote adult neurogenesis in damaged brains [148].
GSLs also play a critical role in memory and cognition through modulating synaptic activities [32] such as neurotransmission [149], functional synapse and neuronal circuit formation and stabilization [150], etc. These actions are supported by the adoptive change of GSLs in response to brain development, sensory stimulation, CNS activity, and learning processes [151]. Exogenous GM1 was shown to induce neurotransmitter release in the cortical synaptosomes of mouse brains [152], and intraperitoneal injection of GM1 was found to improve spatial learning and memory in cognitively impaired geriatric rats [153], thereby helping improve age- and lesion-caused memory deficits [154]. GM1 was discovered to affect neurotransmission through regulating ion channel activities and related signaling, namely enhancing depolarization-induced Ca2+ influx into the synaptosome [152], which is also the underlying mechanism for the enhanced synaptosome release of acetylcholine (ACh) upon stimulation [155]. GM1 was also shown to affect synaptic plasticity [156], referring to the persistent changes in the strength of connections between neuronal cells [150]. Therefore, enhancing the GM1 content in the synaptic membrane of rat hippocampal slices has been proved to increase the synaptic plasticity and potentiation ability of nerves [157]. It has been further demonstrated that intraperitoneal injection and dietary supplementation of gangliosides can have a positive impact on neuroplasticity and the cognitive functions of the brain [150,158].
GM1 can influence the CNS by several other mechanisms as well. For example, it has been shown that GM1 can affect dopamine-mediated signaling through enhancing dopaminergic neuron regeneration [159,160] and high-affinity dopamine uptake in the striatal synaptosome [161], up-regulating dopamine receptors in rat brains [162], and reviving electrically stimulated dopamine release in the kainic acid-lesioned rat striata [163], thereby restoring cholinergic and dopaminergic deficits in aged rats [164], protect dopaminergic neurons against 6-OHDA (6-hydroxydopamine)-caused damages [165], and improve behavioral changes caused by irreversible blockage of dopaminergic sites [166]. GM1 is also a modulator of Ca2+ ion flux and equilibrium in the membrane [167,168], e.g., as a negative regulator of Ca2+-ATPase [169,170] or a potentiator of the nuclear Na+-Ca2+ exchanger [171,172]. GM1 can also affect the CNS through interacting with specific receptors or enzymes. For instance, GM1 was shown to enhance the efficacy and sensitivity of sensory neurons to opioid [173,174] via modulating the activity of excitatory opioid receptor [175,176] and promote leptin receptor (LepR) signaling via activating the JAK2/STAT3 pathway [177]. In addition, GM1 has been found to function by reducing the regional cerebral metabolic rate of glucose in conscious rats [178] and altering the membrane property [136].
In summary, GM1 is one of the most abundant gangliosides in the vertebrate brain and is closely related to the various functions of the CNS [179] (Table 1). GM1 expression in the CNS is dependent on development, and a decrease in GM1 content has been associated with aging, degenerated neurotransmitter activity, and age-related symptoms such as cognitive impairment [180]. In general, GM1 plays a protective role and is directly involved in cell differentiation, neuritogenesis, neuroregeneration, signaling, and neuroplasticity in the CNS. GM1 carries out its functions by an array of mechanisms. Like other GSLs, GM1 tends to cluster in the cell membrane lipid rafts, which are rich in GSLs and various receptors and signaling molecules [76,181]. However, typically, the lipid chains of GSLs are not long enough to span the entire plasma membrane. To transduce signals through the membrane, GSLs need to interact with other membrane components. As a result, lipid rafts are useful platforms for GM1 to form specific microdomains that enable GM1 interaction with molecules both inside and outside of the cell membrane. Consequently, in addition to their glycans, the Cer moieties in GM1 and other GSLs also play an important role in their biological functions [182,183,184].

5. GM1 in Central Nervous System-Related Diseases

5.1. Alzheimer’s Disease

Alzheimer’s disease (AD) is a neurodegenerative disorder that accounts for roughly 60–70% of all cases of dementia [185]. The pathogenesis of AD is not fully understood yet, but AD brains are typically accompanied by the accumulation of misfolded proteins in the form of neurofibrillary tangles. According to the amyloid cascade hypothesis, enzymatic cleavage of the transmembrane amyloid precursor protein (APP) can lead to the release of extracellular Aβ peptides that aggregate into insoluble and neurotoxic oligomers, fibrils, and plaques [186]. Furthermore, Aβ oligomers can promote tau protein hyperphosphorylation and aggregation, which is another cause of neurotoxicity. In this process, GSLs [187,188,189,190,191,192], especially GM1 [193,194,195,196,197,198,199,200,201,202,203,204], play an important role.
Many studies have shown that in AD brains, GM1 can bind monomeric Aβ to act as the seed to promote Aβ aggregation and fibrillogenesis [193,194,195,196,197,198,199,200,201,202,203,204]. After binding to GM1, Aβ performs a series of structural changes, from random coil to α-helix-rich and then β-sheet-rich conformations, depending on the concentration and the aggregation form of Aβ [205]. The β-sheet-rich aggregates are more likely to form neurotoxic oligomers and fibrils [192,206]. Other factors may also affect the process. For example, both sphingomyelin accumulation-[207] and age-dependent high-density GM1 clustering at the presynaptic neuritic terminal [208] could promote Aβ fibrillogenesis. In addition, both the glycan and the Cer of GSLs are important for their interactions with Aβ. For example, Aβ forms longer fibrils in GM1-, GD1a-, and GT1b-containing membranes than in GD1b- and GM2-containing membranes, and GM1 has exhibited the strongest activity as amyloid seeds, indicating the importance of their glycans [209,210]. Additonally, an increase in the GM1 (d18:1) to (d20:1) ratio was observed in AD patients [211]. To explain the influences of GM1 on AD and its progression, it has been suggested that a decrease in GM1 on neuronal cells activates the PI3K/Akt/mTOR pathway to inhibit Aβ fibrillogenesis [212] and, in addition, GM1 can alleviate oxidative stresses to improve memory deficits in an AD rat model [213]. In addition, the mixture of GM1 and soluble Aβ can enhance NSC proliferation [214].
However, GM1 plays a generally protective role in the CNS and is reported to diminish Aβ-associated neurotoxicity [213,215,216]. For example, GM1 can reduce Aβ-induced neurodegeneration through regulating APP proteolysis [217] and inhibiting Aβ-caused Na+/K+-ATPase inactivation [142,216] or GSK3β activation [215]. The latter signaling pathway is involved in the apoptosis and death of neural cells. In addition, GM1 can prevent sphingomyelin-triggered Aβ oligomerization [203]. The apparently contrary roles of GM1 in AD pathogenesis may be attributed to the unique properties of GM1-Aβ interaction [186]. It has been shown that Aβ deposition in the brain is region-, lipid composition-, and environment-dependent [204,205], and the density and organization of GM1 in the membrane are pivotal in its ability to promote Aβ aggregation [218]. Aβ can selectively bind clustered GM1 in lipid rafts [210,219,220,221,222], whereas the physiological concentrations of GM1 are not enough to provoke Aβ oligomerization [203]. Conversely, at low densities, GM1 has been demonstrated to disrupt Aβ dimers in vivo [206] and inhibit sphingomyelin-triggered Aβ oligomerization in vitro [203].
The involvement of GM1 in AD pathogenesis suggests the potential of GM1-based therapy for AD. It has been shown that midazolam, a benzodiazepine that does not target GSLs directly but decreases GM1 expression and GM1-rich microdomains in presynaptic membranes, inhibits amyloid fibril formation in synaptosomes derived from geriatric mice [223]. In 5XFAD transgenic mice that overexpress mutant human amyloid proteins, gene manipulations to deplete GM1 and other major brain gangliosides and treatment with Neu5Ac-specific Limax flavus agglutinin result in substantial improvement to AD pathology [224]. A chimeric peptide with a strong affinity for GM1-containing raft-like membranes has shown therapeutic potential for AD [225].

5.2. Parkinson’s Disease

PD is another common neurodegenerative disorder, which mainly affects movements at the initial stage but, at advanced stages, other symptoms such as dementia and loss of cognition also develop [226]. PD is associated with the loss of dopaminergic neurons, but its exact cause is not clear. GSLs have been found as being involved in PD pathogenesis [227,228]. In the brain of a mouse PD model, the levels of simple GSLs were elevated while the levels of complex gangliosides were decreased, and these changes occurred in parallel with reduced GCase and increased Neu-ase activities [229]. In particular, GM1 is closely related to PD [230,231,232,233]. For example, GM1 levels were lower in residual dopaminergic neurons of PD patients than in those of the normal control, and a GM1 deficit was correlated with PD in both mice and humans [234,235,236]. In gene knockout mice, GM1 depletion led to motor disability, neurodegeneration, and other PD symptoms [237,238], which were attenuated by LIGA-20, a GM1 analogue [237]. Similarly, increasing the levels of GM1 was shown to provide protection in both chemically and α-synuclein induced PD animal models [239,240,241,242]. Thus, GM1 plays a mainly protective role against PD-related damages such as dysfunction of excitatory amino acid neurotransmitters, disorder in calcium homeostasis, abnormal metabolism of oxygen free radicals, abnormal trace elements distribution and/or deposition, and excessive apoptosis [243].
GM1 performs its protective role against PD by regulating α-synuclein folding and aggregation [239] and stimulating autophagy and α-synuclein clearance [240]. GM1 has been shown to specifically interact with α-synuclein, recruit it to the caveolae and lipid rafts, and inhibit its fibrillation [244]. GM1-containing vesicles can interact with α-synuclein to decrease its fibril formation [245]. In addition, GM1 has been shown to inhibit neuroinflammatory responses by decreasing IL-1β expression and increasing IL-1Ra expression. Due to its protective effect, GM1 may be used to treat PD. Indeed, GM1 has been shown to relieve PD-related symptoms and promote patient recovery [246,247,248]. A 5-year open study has revealed that the long-term use of GM1 is safe and may provide clinical benefits in PD patients [249].

5.3. GM1 Gangliosidosis

GM1 gangliosidosis is a lysosomal storage disease (LSD) caused by mutations of the GLB1 gene, which encodes βGal-ase—the enzyme hydrolyzing GM1 into GM2 (Figure 3). The depletion or attenuation of βGal-ase activity leads to a buildup of GM1 and related molecular species [250,251]. GM1 accumulation in the CNS causes progressive neurodegeneration and neurological dysfunction. Based on onset time and severity of symptoms, GM1 gangliosidosis is divided into three clinical types: infantile (type I), juvenile (type II), and adult or chronic (type III) [252]. The symptom onset time also reflects the degree of residual βGal-ase activity remaining in the patient. An increase of GM1 in the patients’ cerebrospinal fluid was also observed; thus GM1 can be a biomarker for clinical diagnosis of the disease and assessment of therapies [253,254].
How GM1 accumulation induces neurodegeneration is not fully understood [251]. However, it has been observed that excess GM1 accumulation can cause CNS myelin deficiency [255,256,257], lysosomal swelling, and cellular dysfunction [258]. In addition, GM1 accumulation results in immune activation and inflammation [259] in the CNS and deplete Ca2+ in the ER to cause neuronal apoptosis [260] and reduce Ca2+ flux in synaptosomes, thereby impairing Ca2+ homeostasis [168,261,262], neurotransmitter uptake, and membrane fluidity [263]. In a mouse model of GM1 gangliosidosis, enhanced autophagy and mitochondrial dysfunction have been correlated with disease progression [264]. GM1 may also act by enhancing NGF-induced Trk receptor autophosphorylation and dimerization to elevate neurotrophic activity [113,265].
Currently, there has been no cure for GM1 gangliosidosis, and thus all treatments are symptomatic. However, significant progress has been made in several directions. For example, enzyme replacement therapies (ERT) based on ICV injection or efficient enzyme delivery systems such as “Trojan horse” proteins and nanoparticles to address the blood brain barrier (BBB) problem have been explored [266,267,268,269,270]. Chaperon-mediated therapies (CMT) employing small molecule chaperons, such as Gal [271], iminosugars and valienamine derivatives [272,273,274,275,276,277,278,279,280], have been utilized to treat diseases caused by unstable or misfolded mutant enzymes [281,282,283,284]. Some of the therapies result in significant increases in βGal-ase activity [284] and decreases in the GM1 content of neonatal brains [279]. Substrate reduction therapy (SRT) using Miglustat, an inhibitor of glucosylceramide synthase, was shown to reduce GM1 and inflammation and improve function of the CNS in animals [279,285] and reverse disease progression in patients [286]. However, SRT relies on residual βGal-ase activities to degrade the GM1 already accumulated in the CNS. Gene therapy is a promising method for GM1 gangliosidosis despite the challenge of achieving global distribution of genomes or active enzymes in the CNS [287]. In a mouse model, intravenous (IV) or adeno-associated virus (AAV)-assisted delivery of genes was shown to completely repair enzyme deficiency and lysosomal storage in the brain [288,289] and improve animal survival [290,291]. The distribution of intraparenchymally injected AAV vectors in the brain and system is believed to be achieved through diffusion, axonal transportation from neurons at the injection site, and cerebrospinal fluid (CSF) flow in the Virchow-Robin perivascular space [292]. AAV-mediated delivery of GLB1 in a feline model has provided additional support for its therapeutic efficacy in humans [293]. However, intracranial injection suffers from the drawback of invasiveness. Recently, it has been shown in an adult mouse model that IV injection of an AAV9-GLB1 vector, which could penetrate the BBB and target the brain [294], resulted in CNS expression of βGal-ase to significantly ameliorate the disease symptoms and extend the animals’ lifespan [291]. IV injection of the AAV9-GLB1 vector in a cat model of GM1 gangliosidosis gave similar results. These encouraging results support the potential clinical benefits of gene therapies to treat GM1 gangliosidosis [295]. Consequently, gene therapies based on cisterna magna delivery of LYS-GM101 and PBGM01 and IV delivery of AAV9-GLB1 have been approved for clinical trials for type I and type II GM1 gangliosidosis [250,251].

5.4. Huntington’s Disease

Huntington’s disease (HD) is a neurodegenerative disorder characterized by progressive motor, cognitive, and psychiatric dysfunctions [296]. It is caused by mutations in the HTT gene encoding the Huntington protein (HTT). These mutations result in HTT misfolding and aggregation, and subsequently neuronal dysfunction and death in the striatum and cortex [297,298,299,300]. It has been revealed that HD is associated with abnormal GSL expression. For example, a decrease in striatal gangliosides, most significantly GM1, was observed in both rat models and HD patient tissues [301,302,303], but in the cerebellum, a region largely unaffected by HD, the GM1 level was elevated [304]. As discussed earlier, GM1 plays an overall protective role in the CNS, and a decrease in GM1 renders neuronal cells more susceptible to injuries and growth factor deprivation involved in the activation of the neuronal prosurvival PI3K/AKT pathway [303]. Meanwhile, GM1 was able to reduce protein aggregation and toxicity by promoting mutant HTT phosphorylation [305,306] and accelerate mutant HTT clearance from the brain [307]. Furthermore, it has been revealed that mutant HTT shows decreased insertion and oligomerization into GM1-modified membranes made from brain lipid extracts [308].
Currently, there is no cure for HD; thus all mainstream treatments are symptomatic [298,299]. GM1 is a promising prospect for HD therapy. In HD mice, intraventricular infusion of exogenous GM1 has shown many therapeutic benefits, including reversal of motor, cognitive, and psychiatric symptoms; slowdown of myelin atrophy and neurodegeneration; improvement of neurotransmitter levels [305,307]; a decrease in mutant HTT; and recoveries at the molecular, cellular, and behavioral levels.

5.5. Epilepsy and Seizure

Epilepsy is a neurological disorder characterized by seizures caused by aberrant neuronal activities in the brain. Severe epilepsy is often accompanied by abnormal GSL expression [29]. For example, decreased levels of the major gangliosides, including GM1, in the hippocampus and particularly GM1 and GD1a in the CSF were observed in a pilocarpine-induced mouse model of epilepsy [309] and the infantile epilepsy disease West syndrome [310]. Defects in the GM3 synthase gene in some Amish families, which result in the loss of all gangliosides, including GM1, cause serious infantile epileptic disorders [311]. Anti-GM1 antibodies were found to cause seizures and were utilized to generate epilepsy models [312]. Conversely, GM1 administration was shown to protect animals from glutaric acid-induced seizures [313].
The protective function of GM1 against seizures is believed to be associated with its ability to modulate membrane receptors and ion channels. GluR2-AMPAR is a receptor by which gangliosides affect neuronal cells [314]. It has been demonstrated that AMPARs containing GM1-bound GluR2 are functionally segregated from GT1b-bound AMPAR-trafficking complexes and disrupting the biosynthesis and balance of gangliosides results in a reduced synaptic expression of GluR2-contianing AMPARs and, thereby, intellectual deficits and seizure susceptibility in mice and humans.

5.6. Amyotrophic Lateral Sclerosis

Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by progressive loss of motor neurons in the brain and spinal cord [315]. It is associated with significantly elevated levels of GM1 and several other simple GSLs in the spinal cord of ALS patients [316] and ALS mouse models [316,317]. In addition, GM1 levels in the CSF are correlated with disease severity [318]. GSLs may contribute to ALS via interaction with NGF and protein kinases [319,320]. Another theory is that ALS may involve autoimmunity because antibodies against GM1 [321,322] and other gangliosides [323] have been detected in the sera and CSF of most ALS patients. However, the autoimmune theory is controversial, as some studies did not observe clear association of the ALS phenotype with elevated anti-ganglioside antibodies [324].

5.7. Ischemic Stroke

Stroke is one of the leading causes of human death and is characterized by severe neurological deficits in long-term survivors [325]. A substantial reduction in total gangliosides, including GM1, is associated with ischemic stroke in neonatal patients who died of hypoxia [326]. Similar results were observed in animals subjected to hypoxia–ischemia [327] or middle cerebral artery occlusion (MCAO) [328]. Meanwhile, the decrease in brain gangliosides correlated with the severity of brain damage. In addition, the ratio of GM1 (d20:1) lipid form, which is neurotoxic, was also significantly elevated [328,329]. However, changes in GM1 levels vary in different regions of the brain. An increase of GM1 was observed in adult rats’ cerebral cortices that were damaged by ischemic infarct, which may be an autologous mechanism developed by adults to protect the brain from ischemic damages [330].
In a rat model of cerebral ischemia, GM1 was revealed to downregulate the N-methyl-D-aspartate (NMDA) receptor that regulates excitotoxicity, thereby ameliorating aberrant excitatory amino acid neurotransmitter and mitochondrial Ca2+ release and reuptake and reducing reactive oxygen species and associated oxidative stress in the cerebral cortex and hippocampus of the injured regions [331,332]. For example, GM1 has been found to protect the brain against ischemic injuries via preventing hypoxia-induced changes in neurotransmitters, such as choline, dopamine, and glutamate [333,334]. In a rat model of hypoxia induced by high altitudes, GM1 was found to activate the PI3K/AKT-Nrf2 signaling pathway, resulting in the suppression of oxidative stress and inflammation [131]. GM1 can also ameliorate cerebral ischemic injury via inhibition of lipid peroxidation, which results in up-regulated expression of tyrosine hydroxylase and, hence, restoration of the tyrosine/dopa pathway [122]. In a rat model of stroke with diabetes, GM1 was proved to inhibit proline-directed kinase ERK1/2 phosphorylation [335] and ER stress-induced neuronal cell apoptosis [336]. In addition, GM1 was able to reverse ischemia-induced changes in Na+/K+- and Mg2+-ATPase and plasma membrane disorganization, thereby leading to the reduction of ischemic injury and cell death [143,337].
The above results suggest the potential benefit of increasing brain GM1 levels for managing stroke- and hypoxia-induced tissue damages and post-stroke patients. Indeed, GM1 administration has been shown to reduce neuronal damage, functional deficit, and autophagy caused by ischemia or stroke [338,339,340,341,342,343], protect the brain from ischemic injury [123,124], and improve therapeutic outcomes and recovery. However, the protective effect of GM1 was time-dependent; thus early administration provided better protection [331]. Oral administration of LIGA20, a GM1 analogue could reduce infarct volume and behavioral impairment after focal cerebral ischemia [344]. Finally, gangliosides were found to reduce overall mortality after global ischemia [345].

5.8. Depression and Anxiety

Depression and anxiety are correlated with the dysregulation and/or malfunction of neurotransmitters and their receptors [346]. Gangliosides are co-receptors for some neurotransmitters that regulate their interaction with membrane receptors and their functions and, hence, are associated with depression and anxiety. It has been shown that serotonin can bind to gangliosides in the cell membrane and, in the process, the interaction of negatively charged Neu5Ac in GM1 with divalent cations, such as Ca2+ and Mg2+, plays a pivotal role [107,347]. Exogenous GM1 was also found to increase the affinity of serotonin for its receptor by a factor of ten to markedly elevate the activity of serotonin and the level of cyclic AMP [348]. Molecular dynamic simulation further suggested that the glycan of GM1 could interact with the extracellular loop of serotonin receptors to drive a conformational change that favors ligand binding [349]. GM1 showed a similar impact on dopamine and other membrane-binding neurotransmitters. GM1 was found to increase the affinity of neuronal dopamine transporters in the membranes of rat striatal synaptosomes [161]. Treatment of animals with GM1 reduced ethanol-induced locomotion decrement, anxiety, and other intoxicating effects [350,351]. In addition, GM1 treatment was proved to reverse defeat stress-induced reduction of social interaction in mice through activating the neurotrophic factor signaling cascade [352]. GM1 can form a charge-based vestibule in front of the postsynaptic membrane to attract neurotransmitters, such as histamine and dopamine, to foster their membrane association [353].

5.9. Autism

Autism is a neurodevelopmental disorder characterized by a wide spectrum of deficiencies in social skills and communication. It is correlated with an increase in all major brain gangliosides, including GM1, in the patients’ CSF, which may affect synaptic transmission and activities [354]. Elevated serum levels of GM1 [355] and anti-GM1 antibodies [356] are also observed in autistic children, and the antibody levels have a positive correlation with autistic severity. However, the relationship of anti-GM1 antibodies to autism is ambiguous. Some studies suggest at least partial involvement of anti-GM1 antibodies and autoimmune neuropsychiatry in autism [356], whereas other studies suggest a lack of direct relationship between them [357].

5.10. Alcohol Dependence

Alcohol dependence is characterized by patients’ physical and psychological reliance on ethanol, a neurotoxin that affects embryogenesis, cell migration and differentiation, and synaptogenesis [358]. Alcohol dependence is related to GSL because alcohol has an impact on the composition, orientation, and conformation of cell surface GSLs and thus the structure and function of cell membrane [359]. Oral (chronic) consumption of alcohol was found to induce a significant change in ganglioside patterns and content in the rat brain. The most significant impacts were observed with hypothalamus GD1b and GT1b, thalamus GM1 and GD1a, and hippocampus GM1, GD1b, and GT1b [360,361]. Both intraperitoneal injection and oral administration of alcohol could lower the level of GM1 in the serum of mice [362].
Accordingly, GM1 is expected to exhibit a protective effect against alcohol intoxication. Indeed, it has been discovered that daily treatment of animals with GM1 can prevent the development of physical dependence on ethanol and reduce ethanol withdrawal syndrome [363]. Pre-incubation of rat cerebellar granule neurons with GM1 or LIGA20 could reduce alcohol-induced cell apoptosis and activation of the DNA cleavage enzyme cytosolic caspase-3 [364]. Similar results were observed in developing mouse brains [365]. In particular, GM1 can attenuate ethanol-induced neurotoxicity in septal and hippocampal neurons from fetal rats and dorsal root ganglion neurons from embryonic chicks [120]. GM1 treatment could remarkably reduce the retardation of cortical cell development and arborization due to in utero ethanol exposure [358]. Pretreatment of pregnant rats with GM1 could reduce fetal alcohol effects in rat pups [366] and the development of alcohol tolerance [367]. Intraperitoneal injection of GM1 was proved to reduce the extent and duration of alcohol-induced ataxia and prevent tremors and anxiety-like behaviors associated with alcohol withdrawal [351].
The protective effect of GM1 against alcohol-induced neurotoxicity is mainly due to its ability to inhibit changes in the cell membrane induced by alcohol, thereby minimizing its impact on brain maturation and behavioral alteration [366,368]. GM1 was shown to reduce alcohol-induced activation of phospholipase A2 (PLA2), an enzyme that regulates phospholipid metabolism, in synaptosomes [369]. GM1 was also found to mitigate alcohol-induced changes in unsaturated fatty acid incorporation into cerebral phospholipids in mice, which is a mechanism to repair alcohol-caused damages [370].

5.11. Bacterial Infections

GSLs also play a key role in the neurotoxicity of some bacterial infections. Typically, these bacteria need to interact with GSLs on the host cell surface, either directly or through their toxins, to cause neurological damage. Vibrio cholerae and Escherichia coli are such bacteria and use GM1 as the binding receptor [371]. Cholera toxin from V. cholerae consists of an A subunit and five identical B subunits, which bind to GM1 on the host cell membrane to help the entry into cells [372,373]. The heat-labile enterotoxin of E. coli is similar to cholera toxin both structurally and functionally [371]. It was found that treating cholera toxin-resistant cells with GM1 can sensitize the cell to this toxin [371]. Studies on different lipid forms of GM1 have proved the importance of its lipid structure in toxin binding and trafficking [374]. These toxins bind only to GM1 carrying unsaturated lipids for their sorting to the Golgi and ER. Moreover, it is mainly the B subunit that interacts with GM1 in the lipid rafts, leading to the modulation of the architecture and dynamics of the membrane and then toxin uptake in the cell via endocytosis [375]. Thus, GM1 on the host cell not only functions as a binding site for the toxin but also facilitates its transmembrane movement [376]. It has been further shown that the cholera toxin B subunit can stimulate Ca2+ influx gated by interaction with GM1 and neuritogenesis [377], although there are also contradictory reports [378]. Tetanus toxin (TeNT) secreted by Clostridium tetani is another toxin that has a high affinity for GM1 and other gangliosides [371,379]. However, gangliosides with a disialyl moiety attached to the inner Gal residue, such as GD1b, appear to have the highest affinity for this toxin [380].
Based on the above findings, it is expected that reagents that can interrupt GSL–toxin interactions may be useful for the treatment of related infectious diseases. Indeed, it has been demonstrated that synthetic sialooligosaccharides mimicking the glycan of GM1 can efficiently inhibit cholera toxin binding to treat cholera-related gastrointestinal symptoms [381]. Moreover, the ability of these toxins to specifically bind GSLs and target the ER-associated protein degradation (ERAD) mechanism has provided the opportunity to develop new therapeutics for cancer and genetic diseases involving misfolded proteins. To this end, GM1-binding cholera toxin should be especially useful [382].
In summary, many neurological diseases are accompanied by changes in the expression level and lipid form of GM1 (Table 2). The correlation of such changes with and their impacts on various disorders are different. The functions and pathological mechanisms of GM1 are relatively well defined in some diseases but ambiguous in others. Nevertheless, the observation that GM1 is associated with various neurological disorders has provided the opportunity to develop new diagnostic and therapeutic strategies.

6. Concluding Remarks

GM1 is one of the most abundant GSLs in the vertebrate animal brain, and its expression is cell-, tissue-, species-, and development-dependent. Like other complex GSLs, the expression level of GM1 in the brain increases from the embryonic stage to adulthood and starts to decline in old age, which coincides with the development of the brain, its functions, and then progressive neurodegeneration during senescence. Therefore, GM1 is associated with a broad spectrum of biological processes and activities of the CNS, such as neuronal differentiation and growth, neuritogenesis, neuroregeneration, cell signal transduction, memory, and cognition. GM1 performs its functions via a variety of mechanisms, from cis and trans interactions with specific receptors and modulation of ion channels and pumps on the cell surface to affecting neurotransmitters and cell membrane properties, namely, the formation and stabilization of lipid rafts. In this process, both the glycan and lipid of GM1 are important. Due to the diverse functions of GM1 in the CNS, changes in its expression level are often accompanied with neurological disorders [383].
Overall, GM1 is protective for the CNS and promotes functional recovery of the brain from various diseases and pathological damages [125,384]. Therefore, elevating the CNS level of GM1 should be beneficial to most neurological disorders discussed above. In addition, GM1 has also been found to protect the brain from damages induced by neurotoxic chemicals [385,386,387,388,389,390,391,392,393,394], X-ray radiation [395,396], stimuli [397], and brain injuries [398,399,400,401], and to improve the therapeutic outcome of experimental allergic neuritis [402]. The functions of GM1 in protecting the brain from neurological disorders are generally related to its ability to interact with the receptors involved in specific signaling pathways, restore the activity of neurotransmitters, promote neuronal cell regeneration, and sustain the structural integrity and functions of the cell membrane [403]. For example, GM1 was found to enhance choline acetyltransferase activity and repair cholinergic deficits in aged and vincristine-treated rats [404,405,406,407], restore the dopamine pathway and reverse dopamine receptor supersensitivity in mitochondrial permeability transition pore (MPTP)-treated mice and monkeys [408,409,410], and ease NMDA receptor-modulated release of excitatory amino acids [411]. Similarly, the GM1 analogue LIGA20 was proved to increase the striatal dopamine level in MPTP-treated mice [412]. Furthermore, GM1 was discovered to stimulate the collateral sprouting of dopaminergic axons in unilateral hemitransection-damaged rat brain [413], prevent GM2-caused inhibition on axonal regeneration [132,414], promote the recovery of dopaminergic neurons in MPTP-treated mice and monkeys [415], rescue 1-methyl-4-phenylpyridinium (MPP)-damaged dopaminergic neurons [416], and eliminate neuroleptic-induced sensorimotor deficits [417].
The correlation of changes in the brain level of GM1 with various neurological disorders renders GM1 a promising target for new therapy development. Some of the strategies include modulating the CNS level of GM1 through direct supplement of GM1, managing endogenous GM1 by genetic methods or manipulation of related enzymes or biosynthetic precursors, and the combination of different strategies. As discussed above, some of these therapies have demonstrated significant progress. However, for all the strategies, a major issue is how to deliver the therapeutic agents to the CNS through the BBB.
Despite the great progress presented above, there are still unanswered questions concerning the functions of GM1 and its roles in neurological diseases. Our understanding of the functions of GM1 is far from comprehensive, and there is still a lack of large-scale and systematic studies of GM1 in relation to neurological disorders that employ human brain samples, but this type of study is necessary to obtain more reliable data. To gain more insights into the functions and functional mechanisms of GM1 in the CNS and its related diseases, there are several issues that need to be addressed. First, it is necessary to have access to human brain tissues at different developmental and pathological stages on a large scale. Second, GM1 is present in the lipid rafts of cell membrane, which has a complex and dynamic environment, and interacts with diverse biomolecules. To distinguish and examine the interactions of GM1 with various molecules on the cell is a significant challenge. This may be addressed by developing differently labeled GM1 probes. Third, probes available for the study of GM1 and related biology are limited. To address this problem, more effective synthetic methods are desired to provide functionalized GM1 derivatives that are enabled by modern analytical technologies. Fourth, a high-throughput method is required for large-scale and comprehensive analysis and profiling of brain GM1, including its complex lipid forms. Nevertheless, we believe that with the development of new analytical technologies and synthetic methods for GM1 and its probes, our understanding of the functions and functional mechanisms of GM1 and its relationships with neurological disorders will improve, which will help the design of new strategies for the diagnosis, prevention, and therapy of various neurological diseases.

Funding

Our work is partially supported by an NIH grant (R35 GM131686).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Z.G. thanks Steven and Rebecca Scott for the Endowment of our research.

Conflicts of Interest

The author declares no conflict of interest.

Acronyms

AAV: adeno-associated virus; Aβ: amyloid beta; Ach: acetylcholine; AD: Alzheimer’s disease; ALS: amyotrophic lateral sclerosis; APP: amyloid precursor protein; βGal-ase: β-galactosidase; βHex-ase: β-hexosaminidase; BBB: blood brain barrier; CaM-kinase: Ca2+/calmodulin-dependent protein kinase; Cer: ceramide; CMT: chaperon-mediated therapies; CNS: central nervous system; CSF: cerebrospinal fluid; ER: endoplasmic reticulum; ERAD: ER-associated protein degradation; ERT: enzyme replacement therapies; GDNF: glia cell-derived neurotrophic factor; FGF2: fibroblast growth factor 2; GalCer: galactosylceramide; GCase: glucocerebrosidase; GlcCer: glucosylceramide; GSL: glycosphingolipid; GT: glycosyltransferase; HD: Huntington’s disease; HTT: huntingtin protein; IV: intravenous; ICV: intracerebroventricular; LepR: leptin receptor; LSD: lysosomal storage disease; MCAO: middle cerebral artery occlusion; MPP: 1-methyl-4-phenylpyridinium; MPTP: mitochondrial permeability transition pore; MS: mass spectrometry; Neu5Ac: N-acetylneuraminic acid (or sialic acid); NGF: nerve growth factor; NMDA: N-methyl-D-aspartate; NSC: neural stem cell; 6-OHDA: 6-hydroxydopamine; PD: Parkinson’s disease; PLA2: phospholipase A2; SAP: sphingolipid activator protein; SRT: substrate reduction therapy; TeNT: Tetanus toxin; Trk: tropomyosin-receptor kinase.

References

  1. Schnaar, R.L.; Suzuki, A.; Stanley, P. Chapter 10—Glycosphingolipids. In Essentials of Glycobiology; Varki, A., Cummings, R.D., Esko, J.D., Freeze, H.H., Stanley, P., Bertozzi, C.R., Hart, G.W., Etzler, M.E., Eds.; Cold Spring Harbor Laboratory Press: Cold Spring Harbor, NY, USA, 2009. [Google Scholar]
  2. Latov, N. Neuropathy and Anti-Gm1 Antibodies. Ann. Neurol. 1990, 27, S41–S43. [Google Scholar] [CrossRef] [PubMed]
  3. Beadon, K.; Guimarães-Costa, R.; Léger, J.M. Multifocal Motor Neuropathy. Curr. Opin. Neurol. 2018, 31, 559–564. [Google Scholar] [CrossRef] [PubMed]
  4. Sántha, P.; Dobos, I.; Kis, G.; Jancsó, G. Role of Gangliosides in Peripheral Pain Mechanisms. Int. J. Mol. Sci. 2020, 21, 1005. [Google Scholar] [CrossRef]
  5. Goodfellow, J.A.; Willison, H.J. Gangliosides and Autoimmune Peripheral Nerve Diseases. Prog. Mol. Biol. Transl. Sci. 2018, 156, 355–382. [Google Scholar] [PubMed]
  6. Sandhoff, R.; Sandhoff, K. Emerging Concepts of Ganglioside Metabolism. FEBS Lett. 2018, 592, 3835–3864. [Google Scholar] [CrossRef] [PubMed]
  7. van Echten-Deckert, G.; Herget, T. Sphingolipid Metabolism in Neural Cells. Biochim. Biophys. Acta 2006, 1758, 1978–1994. [Google Scholar] [CrossRef] [PubMed]
  8. Tettamanti, G.; Bassi, R.; Viani, P.; Riboni, L. Salvage Pathways in Glycosphingolipid Metabolism. Biochimie 2003, 85, 423–437. [Google Scholar] [CrossRef]
  9. Tettamanti, G. Ganglioside/Glycosphingolipid Turnover: New Concepts. Glycoconj. J. 2004, 20, 301–317. [Google Scholar] [CrossRef]
  10. Sonderfeld, S.; Conzelmann, E.; Schwarzmann, G.; Burg, J.; Hinrichs, U.; Sandhoff, K. Incorporation and Metabolism of Ganglioside Gm2 in Skin Fibroblasts from Normal and Gm2 Gangliosidosis Subjects. Eur. J. Biochem. 1985, 149, 247–255. [Google Scholar] [CrossRef]
  11. Riboni, L.; Viani, P.; Tettamanti, G. Estimating Sphingolipid Metabolism and Trafficking in Cultured Cells Using Radiolabeled Compounds. Methods Enzym. 2000, 311, 656–682. [Google Scholar]
  12. Chigorno, V.; Riva, C.; Valsecchi, M.; Nicolini, M.; Brocca, P.; Sonnino, S. Metabolic Processing of Gangliosides by Human Fibroblasts in Culture—Formation and Recycling of Separate Pools of Sphingosine. Eur. J. Biochem. 1997, 250, 661–669. [Google Scholar] [CrossRef]
  13. Sofer, A.; Schwarzmann, G.; Futerman, A.H. The Internalization of a Short Acyl Chain Analogue of Ganglioside Gm1 in Polarized Neurons. J. Cell Sci. 1996, 109, 2111–2119. [Google Scholar] [CrossRef] [PubMed]
  14. Tulsiani, D.R.; Chayko, C.A.; Orgebin-Crist, M.C.; Araki, Y. Temporal Surge of Glycosyltransferase Activities in the Genital Tract of the Hamster during the Estrous Cycle. Biol. Reprod. 1996, 54, 1032–1037. [Google Scholar] [CrossRef] [PubMed]
  15. Saito, M.; Yu, R.K. Role of Myelin-Associated Neuraminidase in the Ganglioside Metabolism of Rat Brain Myelin. J. Neurochem. 1992, 58, 83–87. [Google Scholar] [CrossRef]
  16. Goi, G.; Bairati, C.; Massaccesi, L.; Lovagnini, A.; Lombardo, A.; Tettamanti, G. Membrane Anchoring and Surface Distribution of Glycohydrolases of Human Erythrocyte Membranes. FEBS Lett. 2000, 473, 89–94. [Google Scholar] [CrossRef]
  17. Bassi, R.; Riboni, L.; Tettamanti, G. Cultured Cerebellar Granule Cells, but Not Astrocytes, Produce an Ester of Ganglioside Gd1b, Presumably Gd1b Monolactone, from Exogenous Gd1b. Biochem. J. 1994, 302, 937–942. [Google Scholar] [CrossRef] [PubMed]
  18. Monti, E.; Bonten, E.; D’Azzo, A.; Bresciani, R.; Venerando, B.; Borsani, G.; Schauer, R.; Tettamanti, G. Sialidases in Vertebrates: A Family of Enzymes Tailored for Several Cell Functions. Adv. Carbohydr. Chem. Biochem. 2010, 64, 403–479. [Google Scholar] [PubMed]
  19. Sandhoff, K.; Harzer, K. Gangliosides and Gangliosidoses: Principles of Molecular and Metabolic Pathogenesis. J. Neurosci. 2013, 33, 10195–10208. [Google Scholar] [CrossRef]
  20. Ryckman, A.E.; Brockhausen, I.; Walia, J.S. Metabolism of Glycosphingolipids and Their Role in the Pathophysiology of Lysosomal Storage Disorders. Int. J. Mol. Sci. 2020, 21, 6881. [Google Scholar] [CrossRef]
  21. Kolter, T.; Sandhoff, K. Principles of Lysosomal Membrane Digestion: Stimulation of Sphingolipid Degradation by Sphingolipid Activator Proteins and Anionic Lysosomal Lipids. Annu. Rev. Cell Dev. Biol. 2005, 21, 81–103. [Google Scholar] [CrossRef]
  22. Yu, R.K.; Bieberich, E.; Xia, T.; Zeng, G. Regulation of Ganglioside Biosynthesis in the Nervous System. J. Lipid Res. 2004, 45, 783–793. [Google Scholar] [CrossRef] [PubMed]
  23. Itokazu, Y.; Tsai, Y.T.; Yu, R.K. Epigenetic Regulation of Ganglioside Expression in Neural Stem Cells and Neuronal Cells. Glycoconj. J. 2017, 34, 749–756. [Google Scholar] [CrossRef]
  24. Roseman, S. The Synthesis of Complex Carbohydrates by Multiglycosyltransferase Systems and Their Potential Function in Intercellular Adhesion. Chem. Phys. Lipids 1970, 5, 270–297. [Google Scholar] [CrossRef] [PubMed]
  25. Giraudo, C.G.; Daniotti, J.L.; Maccioni, H.J. Physical and Functional Association of Glycolipid N-Acetyl-Galactosaminyl and Galactosyl Transferases in the Golgi Apparatus. Proc. Natl. Acad. Sci. USA 2001, 98, 1625–1630. [Google Scholar] [CrossRef] [PubMed]
  26. Yusuf, H.K.; Pohlentz, G.; Sandhoff, K. Tunicamycin Inhibits Ganglioside Biosynthesis in Rat Liver Golgi Apparatus by Blocking Sugar Nucleotide Transport across the Membrane Vesicles. Proc. Natl. Acad. Sci. USA 1983, 80, 7075–7079. [Google Scholar] [CrossRef]
  27. Scheel, G.; Acevedo, E.; Conzelmann, E.; Nehrkorn, H.; Sandhoff, K. Model for the Interaction of Membrane-Bound Substrates and Enzymes. Hydrolysis of Ganglioside Gd1a by Sialidase of Neuronal Membranes Isolated from Calf Brain. Eur. J. Biochem. 1982, 127, 245–253. [Google Scholar] [CrossRef]
  28. Kolter, T.; Proia, R.L.; Sandhoff, K. Combinatorial Ganglioside Biosynthesis. J. Biol. Chem. 2002, 277, 25859–25862. [Google Scholar] [CrossRef]
  29. Sipione, S.; Monyror, J.; Galleguillos, D.; Steinberg, N.; Kadam, V. Gangliosides in the Brain: Physiology, Pathophysiology and Therapeutic Applications. Front. Neurosci. 2020, 14, 572965. [Google Scholar] [CrossRef]
  30. Schnaar, R.L.; Gerardy-Schahn, R.; Hildebrandt, H. Sialic Acids in the Brain: Gangliosides and Polysialic Acid in Nervous System Development, Stability, Disease, and Regeneration. Physiol. Rev. 2014, 94, 461–518. [Google Scholar] [CrossRef]
  31. Wiegandt, H. The Chemical Constitution of Gangliosides of the Vertebrate Nervous System. Behav. Brain Res. 1995, 66, 85–97. [Google Scholar] [CrossRef]
  32. Svennerholm, L. Gangliosides and Synaptic Transmission. Adv. Exp. Med. Biol. 1980, 125, 533–544. [Google Scholar]
  33. Ohsawa, T.; Shumiya, S. Age-Related Alteration of Brain Gangliosides in Senescence-Accelerated Mouse (Sam)-P/8. Mech. Ageing Dev. 1991, 59, 263–274. [Google Scholar] [CrossRef] [PubMed]
  34. Ando, S.; Yu, R.K. Isolation and Characterization of a Novel Trisialoganglioside, Gt1a, from Human Brain. J. Biol. Chem. 1977, 252, 6247–6250. [Google Scholar] [CrossRef] [PubMed]
  35. Tettamanti, G.; Bonali, F.; Marchesini, S.; Zambotti, V. A New Procedure for the Extraction, Purification and Fractionation of Brain Gangliosides. Biochim. Biophys. Acta 1973, 296, 160–170. [Google Scholar] [CrossRef] [PubMed]
  36. Iwamori, M.; Nagai, Y. Comparative Study on Ganglioside Compositions of Various Rabbit Tissues. Tissue-Specificity in Ganglioside Molecular Species of Rabbit Thymus. Biochim. Biophys. Acta 1981, 665, 214–220. [Google Scholar] [CrossRef] [PubMed]
  37. Ledeen, R.W.; Haley, J.E.; Skrivanek, J.A. Study of Ganglioside Patterns with Two-Dimensional and Thin-Layer Chromatography and Radioautography; Detection of New Fucogangliosides and Other Minor Species in Rabbit Brain. Anal. Biochem. 1981, 112, 135–142. [Google Scholar] [CrossRef] [PubMed]
  38. Yu, R.K.; Tsai, Y.T.; Ariga, T. Functional Roles of Gangliosides in Neurodevelopment: An Overview of Recent Advances. Neurochem. Res. 2012, 37, 1230–1244. [Google Scholar] [CrossRef]
  39. Yu, R.K.; Macala, L.J.; Taki, T.; Weinfield, H.M.; Yu, F.S. Developmental Changes in Ganglioside Composition and Synthesis in Embryonic Rat Brain. J. Neurochem. 1988, 50, 1825–1829. [Google Scholar] [CrossRef]
  40. Ngamukote, S.; Yanagisawa, M.; Ariga, T.; Ando, S.; Yu, R.K. Developmental Changes of Glycosphingolipids and Expression of Glycogenes in Mouse Brains. J. Neurochem. 2007, 103, 2327–2341. [Google Scholar] [CrossRef]
  41. Irwin, L.N.; Hunter, G.D.; Crandall, J.E.; McCluer, R.H. Ganglioside Patterns During Cerebral Development in the Normal and Reeler Mouse. J. Neurosci. Res. 1985, 13, 591–597. [Google Scholar] [CrossRef]
  42. Vajn, K.; Viljetić, B.; Degmečić, I.V.; Schnaar, R.L.; Heffer, M. Differential Distribution of Major Brain Gangliosides in the Adult Mouse Central Nervous System. PLoS ONE 2013, 8, e75720. [Google Scholar] [CrossRef] [PubMed]
  43. Bouvier, J.D.; Seyfried, T.N. Ganglioside Composition of Normal and Mutant Mouse Embryos. J. Neurochem. 1989, 52, 460–466. [Google Scholar] [CrossRef]
  44. Ohsawa, T. Changes of Mouse Brain Gangliosides during Aging from Young Adult until Senescence. Mech. Ageing Dev. 1989, 50, 169–177. [Google Scholar] [CrossRef]
  45. Hilbig, R.; Lauke, G.; Rahmann, H. Brain Gangliosides during the Life Span (Embryogenesis to Senescence) of the Rat. Dev. Neurosci. 1983, 6, 260–270. [Google Scholar] [CrossRef]
  46. Bhargava, P.; Rao, P.S.; Vajreshwari, A.; Shankar, R. Total Gangliosides, Ganglioside Species and the Activity of Neuraminidase in Different Brain Regions and Spinal Cord of Normal and Undernourished Rats. Lipids 1984, 19, 179–186. [Google Scholar] [CrossRef] [PubMed]
  47. Dreyfus, H.; Harth, S.; Giuliani-Debernardi, A.; Roos, M.; Mack, G.; Mandel, P. Gangliosides in Various Brain Areas of Three Inbred Strains of Mice. Neurochem. Res. 1982, 7, 477–488. [Google Scholar] [CrossRef]
  48. Cabezas, J.A.; Andrés, R.; Hueso, P.; Llanillo, M.; Martínez-Zorzano, V.S.; Rodrigo, M.; Sánchez-Yagüe, J. Ganglioside and Phospholipid Composition of Forebrain, Cerebellum, and Brain Stem from Adult and Newborn Rats. Neurochem. Res. 1991, 16, 781–785. [Google Scholar] [CrossRef] [PubMed]
  49. Ogiso, M.; Ohta, M.; Harada, Y.; Kubo, H.; Hirano, S. Developmental Change in Ganglioside Expression in Primary Culture of Rat Neurons. Neuroscience 1991, 41, 167–176. [Google Scholar] [CrossRef]
  50. Yu, R.K.; Nakatani, Y.; Yanagisawa, M. The Role of Glycosphingolipid Metabolism in the Developing Brain. J. Lipid Res. 2009, 50, S440–S445. [Google Scholar] [CrossRef]
  51. Sandhoff, R.; Schulze, H.; Sandhoff, K. Ganglioside Metabolism in Health and Disease. Prog. Mol. Biol. Transl. Sci. 2018, 156, 1–62. [Google Scholar]
  52. Ishikawa, Y.; Gasa, S.; Minami, R.; Makita, A. Characterization of Neutral Glycosphingolipids from Fetal Human Brain: Evidence for Stage-Specific Expression of the Globo, Ganglio, and Neolacto Series in the Central Nervous System. J. Biochem. 1987, 101, 1369–1375. [Google Scholar] [CrossRef]
  53. Svennerholm, L.; Rynmark, B.M.; Vilbergsson, G.; Fredman, P.; Gottfries, J.; Månsson, J.E.; Percy, A. Gangliosides in Human Fetal Brain. J. Neurochem. 1991, 56, 1763–1768. [Google Scholar] [CrossRef]
  54. Svennerholm, L.; Boström, K.; Fredman, P.; Månsson, J.E.; Rosengren, B.; Rynmark, B.M. Human Brain Gangliosides: Developmental Changes from Early Fetal Stage to Advanced Age. Biochim. Biophys. Acta 1989, 1005, 109–117. [Google Scholar] [CrossRef] [PubMed]
  55. Svennerholm, L.; Boström, K.; Jungbjer, B.; Olsson, L. Membrane Lipids of Adult Human Brain: Lipid Composition of Frontal and Temporal Lobe in Subjects of Age 20 to 100 Years. J. Neurochem. 1994, 63, 1802–1811. [Google Scholar] [CrossRef] [PubMed]
  56. Kracun, I.; Rosner, H.; Drnovsek, V.; Heffer-Lauc, M.; Cosović, C.; Lauc, G. Human Brain Gangliosides in Development, Aging and Disease. Int. J. Dev. Biol. 1991, 35, 289–295. [Google Scholar] [PubMed]
  57. Kracun, I.; Rosner, H.; Drnovsek, V.; Vukelic, Z.; Cosovic, C.; Trbojevic-Cepe, M.; Kubat, M. Gangliosides in the Human Brain Development and Aging. Neurochem. Int. 1992, 20, 421–431. [Google Scholar] [CrossRef]
  58. Hirschberg, K.; Zisling, R.; van Echten-Deckert, G.; Futerman, A.H. Ganglioside Synthesis During the Development of Neuronal Polarity. Major Changes Occur during Axonogenesis and Axon Elongation, but Not during Dendrite Growth or Synaptogenesis. J. Biol. Chem. 1996, 271, 14876–14882. [Google Scholar] [CrossRef]
  59. Kim, S.U.; Moretto, G.; Lee, V.; Yu, R.K. Neuroimmunology of Gangliosides in Human Neurons and Glial Cells in Culture. J. Neurosci. Res. 1986, 15, 303–321. [Google Scholar] [CrossRef]
  60. Sonnino, S.; Chigorno, V. Ganglioside Molecular Species Containing C18- and C20-Sphingosine in Mammalian Nervous Tissues and Neuronal Cell Cultures. Biochim. Biophys. Acta 2000, 1469, 63–77. [Google Scholar] [CrossRef]
  61. Palestini, P.; Masserini, M.; Sonnino, S.; Giuliani, A.; Tettamanti, G. Changes in the Ceramide Composition of Rat Forebrain Gangliosides with Age. J. Neurochem. 1990, 54, 230–235. [Google Scholar] [CrossRef]
  62. Rosenberg, A.; Stern, N. Changes in Sphingosine and Fatty Acid Components of the Gangliosides in Developing Rat and Human Brain. J. Lipid Res. 1966, 7, 122–131. [Google Scholar] [CrossRef] [PubMed]
  63. Kawamura, N.; Taketomi, T. Variation in Sphingosine and Fatty Acid Compositions of Sphingolipids in Rabbit Brain during Development. Jpn. J. Exp. Med. 1969, 39, 383–392. [Google Scholar] [PubMed]
  64. Palestini, P.; Sonnino, S.; Tettamanti, G. Lack of the Ganglioside Molecular Species Containing the C20-Long-Chain Bases in Human, Rat, Mouse, Rabbit, Cat, Dog, and Chicken Brains During Prenatal Life. J. Neurochem. 1991, 56, 2048–2050. [Google Scholar] [CrossRef] [PubMed]
  65. Mansson, J.E.; Vanier, M.T.; Svennerholm, L. Changes in the Fatty Acid and Sphingosine Composition of the Major Gangliosides of Human Brain with Age. J. Neurochem. 1978, 30, 273–275. [Google Scholar] [CrossRef] [PubMed]
  66. Schengrund, C.L.; Garrigan, O.W. A Comparative Study of Gangliosides from Brains of Various Species. Lipids 1969, 4, 488–495. [Google Scholar] [CrossRef]
  67. Valsecchi, M.; Palestini, P.; Chigorno, V.; Sonnino, S.; Tettamanti, G. Changes in the Ganglioside Long-Chain Base Composition of Rat Cerebellar Granule Cells during Differentiation and Aging in Culture. J. Neurochem. 1993, 60, 193–196. [Google Scholar] [CrossRef]
  68. Caughlin, S.; Maheshwari, S.; Weishaupt, N.; Yeung, K.K.; Cechetto, D.F.; Whitehead, S.N. Age-Dependent and Regional Heterogeneity in the Long-Chain Base of a-Series Gangliosides Observed in the Rat Brain Using Maldi Imaging. Sci. Rep. 2017, 7, e16135. [Google Scholar] [CrossRef]
  69. Valsecchi, M.; Palestini, P.; Chigorno, V.; Sonnino, S. Age-Related Changes of the Ganglioside Long-Chain Base Composition in Rat Cerebellum. Neurochem. Int. 1996, 28, 183–187. [Google Scholar] [CrossRef]
  70. Palestini, P.; Masserini, M.; Fiorilli, A.; Calappi, E.; Tettamanti, G. Age-Related Changes in the Ceramide Composition of the Major Gangliosides Present in Rat Brain Subcellular Fractions Enriched in Plasma Membranes of Neuronal and Myelin Origin. J. Neurochem. 1993, 61, 955–960. [Google Scholar] [CrossRef]
  71. Valsecchi, M.; Chigorno, V.; Sonnino, S.; Tettamanti, G. Rat Cerebellar Granule Cells in Culture Associate and Metabolize Differently Exogenous Gm1 Ganglioside Molecular Species Containing a C18 or C20 Long Chain Base. Chem. Phys. Lipids 1992, 60, 247–252. [Google Scholar] [CrossRef]
  72. Kolesnick, R. The Therapeutic Potential of Modulating the Ceramide/Sphingomyelin Pathway. J. Clin. Investig. 2002, 110, 3–8. [Google Scholar] [CrossRef] [PubMed]
  73. Rouser, G.; Yamamoto, A. The Fatty Acid Composition of Total Gangliosides in Normal Human Whole Brain at Different Ages. J. Neurochem. 1972, 19, 2697–2698. [Google Scholar] [CrossRef] [PubMed]
  74. Bielawska, A.; Crane, H.M.; Liotta, D.; Obeid, L.M.; Hannun, Y.A. Selectivity of Ceramide-Mediated Biology. Lack of Activity of Erythro-Dihydroceramide. J. Biol. Chem. 1993, 268, 26226–26232. [Google Scholar] [CrossRef]
  75. Mencarelli, C.; Martinez-Martinez, P. Ceramide Function in the Brain: When a Slight Tilt Is Enough. Cell. Mol. Life Sci. 2013, 70, 181–203. [Google Scholar] [CrossRef] [PubMed]
  76. Aureli, M.; Grassi, S.; Prioni, S.; Sonnino, S.; Prinetti, A. Lipid Membrane Domains in the Brain. Biochim. Biophys. Acta 2015, 1851, 1006–1016. [Google Scholar] [CrossRef] [PubMed]
  77. Olsen, A.S.B.; Færgeman, N.J. Sphingolipids: Membrane Microdomains in Brain Development, Function and Neurological Diseases. Open Biol. 2017, 7, e170069. [Google Scholar] [CrossRef]
  78. Witting, L.A.; Krishnan, R.S.; Sakr, A.H.; Horwitt, M.K. Brain Gangliosides of Several Species. Anal. Biochem. 1968, 22, 295–303. [Google Scholar] [CrossRef]
  79. Svennerholm, L.; Raal, A. Composition of Brain Gangliosides. Biochim. Biophys. Acta 1961, 53, 422–424. [Google Scholar] [CrossRef]
  80. Irwin, L.N.; Michael, D.B.; Irwin, C.C. Ganglioside Patterns of Fetal Rat and Mouse Brain. J. Neurochem. 1980, 34, 1527–1530. [Google Scholar] [CrossRef]
  81. Posse de Chaves, E.; Sipione, S. Sphingolipids and Gangliosides of the Nervous System in Membrane Function and Dysfunction. FEBS Lett. 2010, 584, 1748–1759. [Google Scholar] [CrossRef]
  82. Prinetti, A.; Chigorno, V.; Prioni, S.; Loberto, N.; Marano, N.; Tettamanti, G.; Sonnino, S. Changes in the Lipid Turnover, Composition, and Organization, as Sphingolipid-Enriched Membrane Domains, in Rat Cerebellar Granule Cells Developing In Vitro. J. Biol. Chem. 2001, 276, 21136–21145. [Google Scholar] [CrossRef]
  83. Schengrund, C.L. The Role(S) of Gangliosides in Neural Differentiation and Repair: A Perspective. Brain Res. Bull. 1990, 24, 131–141. [Google Scholar] [CrossRef] [PubMed]
  84. Rösner, H.; al-Aqtum, M.; Rahmann, H. Gangliosides and Neuronal Differentiation. Neurochem. Int. 1992, 20, 339–351. [Google Scholar] [CrossRef] [PubMed]
  85. Di Biase, E.; Lunghi, G.; Fazzari, M.; Maggioni, M.; Pomè, D.Y.; Valsecchi, M.; Samarani, M.; Fato, P.; Ciampa, M.G.; Prioni, S.; et al. Gangliosides in the Differentiation Process of Primary Neurons: The Specific Role of Gm1-Oligosaccharide. Glycoconj. J. 2020, 37, 329–343. [Google Scholar] [CrossRef] [PubMed]
  86. Mahadik, S.P.; Karpiak, S.E. Gm1 Ganglioside Enhances Neonatal Cortical Development. Neurotoxicology 1986, 7, 161–168. [Google Scholar]
  87. Willinger, M.; Schachner, M. Gm1 Ganglioside as a Marker for Neuronal Differentiation in Mouse Cerebellum. Dev. Biol. 1980, 74, 101–117. [Google Scholar] [CrossRef] [PubMed]
  88. Carine, K.; Schengrund, C.L. Effects of Exogenous Gm1 and Gd1a on S20y Neuroblastoma Cells. J. Neurosci. Res. 1984, 12, 59–69. [Google Scholar] [CrossRef]
  89. Nakajima, J.; Tsuji, S.; Nagai, Y. Bioactive Gangliosides: Analysis of Functional Structures of the Tetrasialoganglioside Gq1b Which Promotes Neurite Outgrowth. Biochim. Biophys. Acta 1986, 876, 65–71. [Google Scholar]
  90. Skaper, S.D.; Katoh-Semba, R.; Varon, S. Gm1 Ganglioside Accelerates Neurite Outgrowth from Primary Peripheral and Central Neurons under Selected Culture Conditions. Brain Res. 1985, 355, 19–26. [Google Scholar] [CrossRef]
  91. Sobue, G.; Taki, T.; Yasuda, T.; Mitsuma, T. Gangliosides Modulate Schwann Cell Proliferation and Morphology. Brain Res. 1988, 474, 287–295. [Google Scholar] [CrossRef]
  92. Rahmann, H.; Rösner, H.; Körtje, K.H.; Beitinger, H.; Seybold, V. Ca(2+)-Ganglioside-Interaction in Neuronal Differentiation and Development. Prog. Brain Res. 1994, 101, 127–145. [Google Scholar] [PubMed]
  93. Fukunaga, K.; Miyamoto, E.; Soderling, T.R. Regulation of Ca2+/Calmodulin-Dependent Protein Kinase Ii by Brain Gangliosides. J. Neurochem. 1990, 54, 103–109. [Google Scholar] [CrossRef] [PubMed]
  94. Ferrari, G.; Fabris, M.; Gorio, A. Gangliosides Enhance Neurite Outgrowth in Pc12 Cells. Brain Res. 1983, 284, 215–221. [Google Scholar] [CrossRef] [PubMed]
  95. Yates, A.J.; Rampersaud, A. Sphingolipids as Receptor Modulators. An Overview. Ann. N. Y. Acad. Sci. 1998, 845, 57–71. [Google Scholar] [CrossRef]
  96. Cuello, A.C.; Garofalo, L.; Kenigsberg, R.L.; Maysinger, D. Gangliosides Potentiate in Vivo and in Vitro Effects of Nerve Growth Factor on Central Cholinergic Neurons. Proc. Natl. Acad. Sci. USA 1989, 86, 2056–2060. [Google Scholar] [CrossRef] [PubMed]
  97. Panni, M.K.; Cooper, J.D.; Sofroniew, M.V. Ganglioside Gm1 Potentiates Ngf Action on Axotomised Medial Septal Cholinergic Neurons. Brain Res. 1998, 812, 76–80. [Google Scholar] [CrossRef]
  98. Vantini, G.; Fusco, M.; Bigon, E.; Leon, A. Gm1 Ganglioside Potentiates the Effect of Nerve Growth Factor in Preventing Vinblastine-Induced Sympathectomy in Newborn Rats. Brain Res. 1988, 448, 252–258. [Google Scholar] [CrossRef] [PubMed]
  99. Garofalo, L.; Cuello, A.C. Nerve Growth Factor and the Monosialoganglioside Gm1: Analogous and Different In Vivo Effects on Biochemical, Morphological, and Behavioral Parameters of Adult Cortically Lesioned Rats. Exp. Neurol. 1994, 125, 195–217. [Google Scholar] [CrossRef]
  100. Mutoh, T.; Tokuda, A.; Miyadai, T.; Hamaguchi, M.; Fujiki, N. Ganglioside Gm1 Binds to the Trk Protein and Regulates Receptor Function. Proc. Natl. Acad. Sci. USA 1995, 92, 5087–5091. [Google Scholar] [CrossRef]
  101. Fukuda, Y.; Fukui, T.; Hikichi, C.; Ishikawa, T.; Murate, K.; Adachi, T.; Imai, H.; Fukuhara, K.; Ueda, A.; Kaplan, A.P.; et al. Neurotropin Promotes Ngf Signaling through Interaction of Gm1 Ganglioside with Trk Neurotrophin Receptor in Pc12 Cells. Brain Res. 2015, 1596, 13–21. [Google Scholar] [CrossRef]
  102. Nagai, Y. Functional Roles of Gangliosides in Bio-Signaling. Behav. Brain Res. 1995, 66, 99–104. [Google Scholar] [CrossRef] [PubMed]
  103. Sonnino, S.; Mauri, L.; Ciampa, M.G.; Prinetti, A. Gangliosides as Regulators of Cell Signaling: Ganglioside-Protein Interactions or Ganglioside-Driven Membrane Organization? J. Neurochem. 2013, 124, 432–435. [Google Scholar] [PubMed]
  104. Duchemin, A.M.; Ren, Q.; Neff, N.H.; Hadjiconstantinou, M. Gm1-Induced Activation of Phosphatidylinositol 3-Kinase: Involvement of Trk Receptors. J. Neurochem. 2008, 104, 1466–1477. [Google Scholar] [CrossRef] [PubMed]
  105. Rusnati, M.; Urbinati, C.; Tanghetti, E.; Dell’Era, P.; Lortat-Jacob, H.; Presta, M. Cell Membrane Gm1 Ganglioside Is a Functional Coreceptor for Fibroblast Growth Factor 2. Proc. Natl. Acad. Sci. USA 2002, 99, 4367–4372. [Google Scholar] [CrossRef]
  106. Hadjiconstantinou, M.; Neff, N.H. Gm1 Ganglioside: In Vivo and in Vitro Trophic Actions on Central Neurotransmitter Systems. J. Neurochem. 1998, 70, 1335–1345. [Google Scholar]
  107. Matinyan, N.S.; Melikyan, G.B.; Arakelyan, V.B.; Kocharov, S.L.; Prokazova, N.V.; Avakian, T.M. Interaction of Ganglioside-Containing Planar Bilayers with Serotonin and Inorganic Cations. Biochim. Biophys. Acta 1989, 984, 313–318. [Google Scholar] [CrossRef]
  108. Postila, P.A.; Vattulainen, I.; Róg, T. Selective Effect of Cell Membrane on Synaptic Neurotransmission. Sci. Rep. 2016, 6, e19345. [Google Scholar]
  109. Fantini, J.; Barrantes, F.J. Sphingolipid/Cholesterol Regulation of Neurotransmitter Receptor Conformation and Function. Biochim. Biophys. Acta 2009, 1788, 2345–2361. [Google Scholar]
  110. Duchemin, A.M.; Ren, Q.; Mo, L.; Neff, N.H.; Hadjiconstantinou, M. Gm1 Ganglioside Induces Phosphorylation and Activation of Trk and Erk in Brain. J. Neurochem. 2002, 81, 696–707. [Google Scholar] [CrossRef]
  111. Duchemin, A.M.; Neff, N.H.; Hadjiconstantinou, M. Induction of Trk Phosphorylation in Rat Brain by Gm1 Ganglioside. Ann. N. Y. Acad. Sci. 1998, 845, 406. [Google Scholar] [CrossRef]
  112. Pitto, M.; Mutoh, T.; Kuriyama, M.; Ferraretto, A.; Palestini, P.; Masserini, M. Influence of Endogenous Gm1 Ganglioside on Trkb Activity, in Cultured Neurons. FEBS Lett. 1998, 439, 93–96. [Google Scholar] [CrossRef]
  113. Chiricozzi, E.; Pomè, D.Y.; Maggioni, M.; Di Biase, E.; Parravicini, C.; Palazzolo, L.; Loberto, N.; Eberini, I.; Sonnino, S. Role of the Gm1 Ganglioside Oligosaccharide Portion in the Trka-Dependent Neurite Sprouting in Neuroblastoma Cells. J. Neurochem. 2017, 143, 645–659. [Google Scholar] [CrossRef]
  114. Arboleda, G.; Morales, L.C.; Benítez, B.; Arboleda, H. Regulation of Ceramide-Induced Neuronal Death: Cell Metabolism Meets Neurodegeneration. Brain Res. Rev. 2009, 59, 333–346. [Google Scholar] [CrossRef]
  115. Ferrari, G.; Minozzi, M.C.; Zanellato, A.M.; Silvestrini, B. Gm1, Like Igf-I and Gdnf, Prevents Neuronal Apoptosis. Ann. N. Y. Acad. Sci. 1998, 845, 408. [Google Scholar] [CrossRef]
  116. Chiricozzi, E.; Lunghi, G.; Di Biase, E.; Fazzari, M.; Sonnino, S.; Mauri, L. Gm1 Ganglioside Is a Key Factor in Maintaining the Mammalian Neuronal Functions Avoiding Neurodegeneration. Int. J. Mol. Sci. 2020, 21, 868. [Google Scholar] [CrossRef]
  117. Gorio, A.; Vitadello, M. Ganglioside Prevention of Neuronal Functional Decay. Prog. Brain Res. 1987, 71, 203–208. [Google Scholar] [PubMed]
  118. Manev, H.; Favaron, M.; Vicini, S.; Guidotti, A. Ganglioside-Mediated Protection from Glutamate-Induced Neuronal Death. Acta Neurobiol. Exp. 1990, 50, 475–488. [Google Scholar]
  119. Costa, E.; Armstrong, D.M.; Guidotti, A.; Kharlamov, A.; Kiedrowski, L.; Manev, H.; Polo, A.; Wroblewski, J.T. Gangliosides in the Protection against Glutamate Excitotoxicity. Prog. Brain Res. 1994, 101, 357–373. [Google Scholar] [PubMed]
  120. Heaton, M.B.; Paiva, M.; Swanson, D.J.; Walker, D.W. Ethanol Neurotoxicity in Vitro: Effects of Gm1 Ganglioside and Protein Synthesis Inhibition. Brain Res. 1994, 654, 336–342. [Google Scholar] [CrossRef] [PubMed]
  121. Skaper, S.D.; Facci, L.; Milani, D.; Leon, A. Monosialoganglioside Gm1 Protects against Anoxia-Induced Neuronal Death In Vitro. Exp. Neurol. 1989, 106, 297–305. [Google Scholar] [CrossRef]
  122. Li, M.; Liao, Y.J.; Hou, G.H.; Yang, Z.B.; Zuo, M.L. Monosialotetrahexosylganglioside Protect Cerebral Ischemia/Reperfusion Injury through Upregulating the Expression of Tyrosine Hydroxylase by Inhibiting Lipid Peroxidation. Biomed. Pharmacother. 2016, 84, 1923–1929. [Google Scholar]
  123. Weber, M.; Mohand-Said, S.; Hicks, D.; Dreyfus, H.; Sahel, J.A. Monosialoganglioside Gm1 Reduces Ischemia—Reperfusion-Induced Injury in the Rat Retina. Investig. Opthalmol. Vis. Sci. 1996, 37, 267–273. [Google Scholar] [CrossRef]
  124. Tan, W.K.; Williams, C.E.; Mallard, C.E.; Gluckman, P.D. Monosialoganglioside Gm1 Treatment after a Hypoxic-Ischemic Episode Reduces the Vulnerability of the Fetal Sheep Brain to Subsequent Injuries. Am. J. Obstet. Gynecol. 1994, 170, 663–669. [Google Scholar] [CrossRef] [PubMed]
  125. Janigro, D.; Di Gregorio, F.; Vyskocil, F.; Gorio, A. Gangliosides’ Dual Mode of Action: A Working Hypothesis. J. Neurosci. Res. 1984, 12, 499–509. [Google Scholar] [CrossRef]
  126. Ferrari, G.; Anderson, B.L.; Stephens, R.M.; Kaplan, D.R.; Greene, L.A. Prevention of Apoptotic Neuronal Death by Gm1 Ganglioside. Involvement of Trk Neurotrophin Receptors. J. Biol. Chem. 1995, 270, 3074–3080. [Google Scholar] [CrossRef]
  127. Zhang, Z.; Liu, W.; Shen, M.; Ma, X.; Li, R.; Jin, X.; Bai, H.; Gao, L. Protective Effect of Gm1 Attenuates Hippocampus and Cortex Apoptosis after Ketamine Exposure in Neonatal Rat Via Pi3k/Akt/Gsk3β Pathway. Mol. Neurobiol. 2021, 58, 3471–3483. [Google Scholar] [CrossRef]
  128. Meng, C.; Yao, X.Q.; Chang, R.J.; Wang, S.L.; Wang, X.; Ma, D.Q.; Li, Q.; Wang, X.Y. Exogenous Gm1 Ganglioside Attenuates Ketamine-Induced Neurocognitive Impairment in the Developing Rat Brain. Anesth. Analg. 2020, 130, 505–517. [Google Scholar] [CrossRef]
  129. She, J.Q.; Wang, M.; Zhu, D.M.; Tang, M.; Chen, J.T.; Wang, L.; Ruan, D.Y. Monosialoanglioside (Gm1) Prevents Lead-Induced Neurotoxicity on Long-Term Potentiation, Sod Activity, Mda Levels, and Intracellular Calcium Levels of Hippocampus in Rats. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2009, 379, 517–524. [Google Scholar] [CrossRef]
  130. Chen, F.; Zhou, C.C.; Yang, Y.; Liu, J.W.; Yan, C.H. Gm1 Ameliorates Lead-Induced Cognitive Deficits and Brain Damage through Activating the Sirt1/Creb/Bdnf Pathway in the Developing Male Rat Hippocampus. Biol. Trace Elem. Res. 2019, 190, 425–436. [Google Scholar] [CrossRef]
  131. Gong, G.; Yin, L.; Yuan, L.; Sui, D.; Sun, Y.; Fu, H.; Chen, L.; Wang, X. Ganglioside Gm1 Protects against High Altitude Cerebral Edema in Rats by Suppressing the Oxidative Stress and Inflammatory Response via the Pi3k/Akt-Nrf2 Pathway. Mol. Immunol. 2018, 95, 91–98. [Google Scholar] [CrossRef]
  132. Benady, A.; Freidin, D.; Pick, C.G.; Rubovitch, V. Gm1 Ganglioside Prevents Axonal Regeneration Inhibition and Cognitive Deficits in a Mouse Model of Traumatic Brain Injury. Sci. Rep. 2018, 8, e13340. [Google Scholar] [CrossRef]
  133. Ni, Y.F.; Zhang, W.; Bao, X.F.; Wang, W.; Song, L.; Jiang, B. Gm1 Ganglioside Reverses the Cognitive Deficits Induced by Mk801 in Mice. Behav. Pharmacol. 2016, 27, 451–459. [Google Scholar] [CrossRef]
  134. Galleguillos, D.; Wang, Q.; Steinberg, N.; Zaidi, A.; Shrivastava, G.; Dhami, K.; Daskhan, G.C.; Schmidt, E.N.; Dworsky-Fried, Z.; Giuliani, F.; et al. Anti-Inflammatory Role of Gm1 and Other Gangliosides on Microglia. J. Neuroinflamm. 2022, 19, e9. [Google Scholar] [CrossRef]
  135. Sarkar, S.; Bose, D.; Giri, R.P.; Mukhopadhyay, M.K.; Chakrabarti, A. Effects of Gm1 on Brain Spectrin-Aminophospholipid Interactions. Biochim. Biophys. Acta Biomembr. 2019, 1861, 298–305. [Google Scholar] [CrossRef]
  136. Ilic, K.; Lin, X.; Malci, A.; Stojanović, M.; Puljko, B.; Rožman, M.; Vukelić, Ž.; Heffer, M.; Montag, D.; Schnaar, R.L.; et al. Plasma Membrane Calcium Atpase-Neuroplastin Complexes Are Selectively Stabilized in Gm1-Containing Lipid Rafts. Int. J. Mol. Sci. 2021, 22, 13590. [Google Scholar] [CrossRef] [PubMed]
  137. Tsai, Y.T.; Itokazu, Y.; Yu, R.K. Gm1 Ganglioside Is Involved in Epigenetic Activation Loci of Neuronal Cells. Neurochem. Res. 2016, 41, 107–115. [Google Scholar] [CrossRef] [PubMed]
  138. Toffano, G.; Agnati, L.F.; Fuxe, K.G. The Effect of the Ganglioside Gm1 on Neuronal Plasticity. Int. J. Dev. Neurosci. 1986, 4, 97–100. [Google Scholar] [CrossRef] [PubMed]
  139. Maysinger, D.; Leavitt, B.R.; Zorc, B.; Butula, I.; Fernandes, L.G.; Ribeiro-da-Silva, A. Inhibition of High Affinity Choline Uptake in the Rat Brain by Neurotoxins: Effect of Monosialoganglioside Gm1. Neurochem. Int. 1992, 20, 289–297. [Google Scholar] [CrossRef] [PubMed]
  140. Skaper, S.D.; Leon, A.; Facci, L. Ganglioside Gm1 Prevents Death Induced by Excessive Excitatory Neurotransmission in Cultured Hippocampal Pyramidal Neurons. Neurosci. Lett. 1991, 126, 98–101. [Google Scholar] [CrossRef]
  141. Leon, A.; Facci, L.; Toffano, G.; Sonnino, S.; Tettamanti, G. Activation of (Na+, K+)-Atpase by Nanomolar Concentrations of Gm1 Ganglioside. J. Neurochem. 1981, 37, 350–357. [Google Scholar] [CrossRef]
  142. Kreutz, F.; Scherer, E.B.; Ferreira, A.G.; Petry Fdos, S.; Pereira, C.L.; Santana, F.; de Souza Wyse, A.T.; Salbego, C.G.; Trindade, V.M. Alterations on Na+,K+-Atpase and Acetylcholinesterase Activities Induced by Amyloid-Β Peptide in Rat Brain and Gm1 Ganglioside Neuroprotective Action. Neurochem. Res. 2013, 38, 2342–2350. [Google Scholar] [CrossRef] [PubMed]
  143. Mahadik, S.P.; Hawver, D.B.; Hungund, B.L.; Li, Y.S.; Karpiak, S.E. Gm1 Ganglioside Treatment after Global Ischemia Protects Changes in Membrane Fatty Acids and Properties of Na+, K+-Atpase and Mg2+-Atpase. J. Neurosci. Res. 1989, 24, 402–412. [Google Scholar] [CrossRef]
  144. Kappagantula, S.; Andrews, M.R.; Cheah, M.; Abad-Rodriguez, J.; Dotti, C.G.; Fawcett, J.W. Neu3 Sialidase-Mediated Ganglioside Conversion Is Necessary for Axon Regeneration and Is Blocked in Cns Axons. J. Neurosci. 2014, 34, 2477–2492. [Google Scholar] [CrossRef] [PubMed]
  145. Newburn, E.N.; Duchemin, A.M.; Neff, N.H.; Hadjiconstantinou, M. Gm1 Ganglioside Enhances Ret Signaling in Striatum. J. Neurochem. 2014, 130, 541–554. [Google Scholar] [CrossRef] [PubMed]
  146. Battistin, L.; Cesari, A.; Galligioni, F.; Marin, G.; Massarotti, M.; Paccagnella, D.; Pellegrini, A.; Testa, G.; Tonin, P. Effects of Gm1 Ganglioside in Cerebrovascular Diseases: A Double-Blind Trial in 40 Cases. Eur. Neurol. 1985, 24, 343–351. [Google Scholar] [CrossRef]
  147. Schneider, J.S.; Gollomp, S.M.; Sendek, S.; Colcher, A.; Cambi, F.; Du, W. A Randomized, Controlled, Delayed Start Trial of Gm1 Ganglioside in Treated Parkinson’s Disease Patients. J. Neurol. Sci. 2013, 324, 140–148. [Google Scholar] [CrossRef]
  148. Itokazu, Y.; Li, D.; Yu, R.K. Intracerebroventricular Infusion of Gangliosides Augments the Adult Neural Stem Cell Pool in Mouse Brain. ASN Neuro 2019, 11, e1759091419884859. [Google Scholar] [CrossRef]
  149. Ledeen, R.W.; Wu, G. Ganglioside Function in Calcium Homeostasis and Signaling. Neurochem. Res. 2002, 27, 637–647. [Google Scholar] [CrossRef]
  150. Palmano, K.; Rowan, A.; Guillermo, R.; Guan, J.; McJarrow, P. The Role of Gangliosides in Neurodevelopment. Nutrients 2015, 7, 3891–3913. [Google Scholar] [CrossRef]
  151. Rahmann, H.; Probst, W.; Mühleisen, M. Gangliosides and Synaptic Transmission. Jpn. J. Exp. Med. 1982, 52, 275–286. [Google Scholar]
  152. Tanaka, Y.; Waki, H.; Kon, K.; Ando, S. Gangliosides Enhance Kcl-Induced Ca2+ Influx and Acetylcholine Release in Brain Synaptosomes. Neuroreport 1997, 8, 2203–2207. [Google Scholar] [CrossRef]
  153. Fong, T.G.; Neff, N.H.; Hadjiconstantinou, M. Gm1 Ganglioside Improves Spatial Learning and Memory of Aged Rats. Behav. Brain Res. 1997, 85, 203–211. [Google Scholar] [CrossRef] [PubMed]
  154. Wang, B. Molecular Mechanism Underlying Sialic Acid as an Essential Nutrient for Brain Development and Cognition. Adv. Nutr. 2012, 3, 465s–472s. [Google Scholar] [CrossRef] [PubMed]
  155. Ando, S.; Tanaka, Y.; Waki, H.; Kon, K.; Iwamoto, M.; Fukui, F. Gangliosides and Sialylcholesterol as Modulators of Synaptic Functions. Ann. N. Y. Acad. Sci. 1998, 845, 232–239. [Google Scholar] [CrossRef]
  156. Egorushkina, N.V.; Ratushnyak, A.S.; Egorushkin, I.V. The Influence of Exogenous Gangliosides on the Dynamics of the Development of Prolonged Posttetanic Potentiation. Neurosci. Behav. Physiol. 1993, 23, 435–438. [Google Scholar] [CrossRef]
  157. Wieraszko, A.; Seifert, W. The Role of Monosialoganglioside Gm1 in the Synaptic Plasticity: In Vitro Study on Rat Hippocampal Slices. Brain Res. 1985, 345, 159–164. [Google Scholar] [CrossRef] [PubMed]
  158. She, J.Q.; Wang, M.; Zhu, D.M.; Sun, L.G.; Ruan, D.Y. Effect of Ganglioside on Synaptic Plasticity of Hippocampus in Lead-Exposed Rats In Vivo. Brain Res. 2005, 1060, 162–169. [Google Scholar] [CrossRef]
  159. Toffano, G.; Savoini, G.; Moroni, F.; Lombardi, G.; Calza, L.; Agnati, L.F. Gm1 Ganglioside Stimulates the Regeneration of Dopaminergic Neurons in the Central Nervous System. Brain Res. 1983, 261, 163–166. [Google Scholar] [CrossRef] [PubMed]
  160. Agnati, L.F.; Fuxe, K.; Calza, L.; Goldstein, M.; Toffano, G.; Giardino, L.; Zoli, M. Further Studies on the Effects of the Gm1 Ganglioside on the Degenerative and Regenerative Features of Mesostriatal Dopamine Neurons. Acta Physiol. Scand. Suppl. 1984, 532, 37–44. [Google Scholar]
  161. Barrier, L.; Page, G.; Barc, S.; Piriou, A.; Portoukalian, J. Sulfatide and Gm1 Ganglioside Modulate the High-Affinity Dopamine Uptake in Rat Striatal Synaptosomes: Evidence for the Involvement of Their Ionic Charges. Neurochem. Int. 2003, 42, 305–313. [Google Scholar] [CrossRef]
  162. Tilson, H.A.; Harry, G.J.; Nanry, K.; Hudson, P.M.; Hong, J.S. Ganglioside Interactions with the Dopaminergic System of Rats. J. Neurosci. Res. 1988, 19, 88–93. [Google Scholar] [CrossRef] [PubMed]
  163. Raiteri, M.; Bonanno, G.; Pende, M.; Versace, P. Gm1 Ganglioside Treatment Promotes Recovery of Electrically-Stimulated [3h]Dopamine Release in Striatal Slices from Rats Lesioned with Kainic Acid. Neurosci. Lett. 1992, 136, 127–130. [Google Scholar] [CrossRef]
  164. Goettl, V.M.; Wemlinger, T.A.; Fong, T.G.; Neff, N.H.; Hadjiconstantinou, M. Retinal Cholinergic and Dopaminergic Deficits of Aged Rats Are Improved Following Treatment with Gm1 Ganglioside. Brain Res. 2000, 877, 1–6. [Google Scholar]
  165. Fadda, E.; Negro, A.; Facci, L.; Skaper, S.D. Ganglioside Gm1 Cooperates with Brain-Derived Neurotrophic Factor to Protect Dopaminergic Neurons from 6-Hydroxydopamine-Induced Degeneration. Neurosci. Lett. 1993, 159, 147–150. [Google Scholar] [CrossRef]
  166. Keller, E.A.; Molina, V.A. Gangliosides Facilitate the Recovery of Behavioral Response Mediated by Dopaminergic Sites following Their Irreversible Blockade. Brain Res. Bull. 1993, 31, 161–163. [Google Scholar] [CrossRef]
  167. Domańska-Janik, K.; Noremberg, K.; Lazarewicz, J. Gangliosides and Synaptosomal Calcium Homeostasis. Int. J. Tissue React. 1986, 8, 373–382. [Google Scholar] [PubMed]
  168. Spoerri, P.E.; Dozier, A.K.; Roisen, F.J. Calcium Regulation of Neuronal Differentiation: The Role of Calcium in Gm1-Mediated Neuritogenesis. Brain Res. Dev. Brain Res. 1990, 56, 177–188. [Google Scholar] [CrossRef] [PubMed]
  169. Wang, Y.; Tsui, Z.; Yang, F. Antagonistic Effect of Ganglioside Gm1 and Gm3 on the Activity and Conformation of Sarcoplasmic Reticulum Ca(2+)-Atpase. FEBS Lett. 1999, 457, 144–148. [Google Scholar] [CrossRef]
  170. Zhao, Y.; Fan, X.; Yang, F.; Zhang, X. Gangliosides Modulate the Activity of the Plasma Membrane Ca(2+)-Atpase from Porcine Brain Synaptosomes. Arch. Biochem. Biophys. 2004, 427, 204–212. [Google Scholar] [CrossRef]
  171. Ledeen, R.W.; Wu, G. Gm1 Ganglioside: Another Nuclear Lipid That Modulates Nuclear Calcium. Gm1 Potentiates the Nuclear Sodium-Calcium Exchanger. Can. J. Physiol. Pharmacol. 2006, 84, 393–402. [Google Scholar] [CrossRef]
  172. Xie, X.; Wu, G.; Lu, Z.H.; Ledeen, R.W. Potentiation of a Sodium-Calcium Exchanger in the Nuclear Envelope by Nuclear Gm1 Ganglioside. J. Neurochem. 2002, 81, 1185–1195. [Google Scholar] [CrossRef] [PubMed]
  173. Shen, K.F.; Crain, S.M.; Ledeen, R.W. Brief Treatment of Sensory Ganglion Neurons with Gm1 Ganglioside Enhances the Efficacy of Opioid Excitatory Effects on the Action Potential. Brain Res. 1991, 559, 130–138. [Google Scholar] [CrossRef]
  174. Crain, S.M.; Shen, K.F. After Chronic Opioid Exposure Sensory Neurons Become Supersensitive to the Excitatory Effects of Opioid Agonists and Antagonists as Occurs after Acute Elevation of Gm1 Ganglioside. Brain Res. 1992, 575, 13–24. [Google Scholar] [CrossRef]
  175. Wu, G.; Lu, Z.H.; Ledeen, R.W. Interaction of the Delta-Opioid Receptor with Gm1 Ganglioside: Conversion from Inhibitory to Excitatory Mode. Brain Res. Mol. Brain Res. 1997, 44, 341–346. [Google Scholar] [CrossRef]
  176. Shen, K.F.; Crain, S.M. Cholera Toxin-B Subunit Blocks Excitatory Effects of Opioids on Sensory Neuron Action Potentials Indicating That Gm1 Ganglioside May Regulate Gs-Linked Opioid Receptor Functions. Brain Res. 1990, 531, 1–7. [Google Scholar] [CrossRef] [PubMed]
  177. Inamori, K.I.; Inokuchi, J.I. Roles of Gangliosides in Hypothalamic Control of Energy Balance: New Insights. Int. J. Mol. Sci. 2020, 21, 5349. [Google Scholar] [CrossRef]
  178. Freo, U.; Dam, M.; Pizzolato, G.; Pietrini, P.; Soncrant, T.T.; Battistin, L. The Monosialoganglioside Gm1 Dose-Dependently Reduces Regional Cerebral Metabolic Rates for Glucose in Awake Rats. Brain Res. 1993, 621, 175–179. [Google Scholar] [CrossRef]
  179. Cochran, F.B., Jr.; Yu, R.K.; Ledeen, R.W. Myelin Gangliosides in Vertebrates. J. Neurochem. 1982, 39, 773–779. [Google Scholar] [CrossRef]
  180. Hadjiconstantinou, M.; Neff, N.H. Gm1 and the Aged Brain. Ann. N. Y. Acad. Sci. 1998, 845, 225–231. [Google Scholar] [CrossRef]
  181. Schengrund, C.L. Lipid Rafts: Keys to Neurodegeneration. Brain Res. Bull. 2010, 82, 7–17. [Google Scholar] [CrossRef] [PubMed]
  182. Masserini, M.; Palestini, P.; Freire, E. Influence of Glycolipid Oligosaccharide and Long-Chain Base Composition on the Thermotropic Properties of Dipalmitoylphosphatidylcholine Large Unilamellar Vesicles Containing Gangliosides. Biochemistry 1989, 28, 5029–5034. [Google Scholar] [CrossRef] [PubMed]
  183. Morrow, M.R.; Singh, D.; Lu, D.; Grant, C.W. Glycosphingolipid Fatty Acid Arrangement in Phospholipid Bilayers: Cholesterol Effects. Biophys. J. 1995, 68, 179–186. [Google Scholar] [CrossRef]
  184. Brockman, H.L.; Momsen, M.M.; Brown, R.E.; He, L.; Chun, J.; Byun, H.S.; Bittman, R. The 4,5-Double Bond of Ceramide Regulates Its Dipole Potential, Elastic Properties, and Packing Behavior. Biophys. J. 2004, 87, 1722–1731. [Google Scholar] [CrossRef] [PubMed]
  185. Leng, F.; Edison, P. Neuroinflammation and Microglial Activation in Alzheimer Disease: Where Do We Go from Here? Nat. Rev. Neurol. 2021, 17, 157–172. [Google Scholar] [CrossRef] [PubMed]
  186. Cebecauer, M.; Hof, M.; Amaro, M. Impact of Gm(1) on Membrane-Mediated Aggregation/Oligomerization of Β-Amyloid: Unifying View. Biophys. J. 2017, 113, 1194–1199. [Google Scholar] [CrossRef]
  187. Yanagisawa, K. Pathological Significance of Ganglioside Clusters in Alzheimer’s Disease. J. Neurochem. 2011, 116, 806–812. [Google Scholar] [CrossRef]
  188. Yanagisawa, K. Role of Gangliosides in Alzheimer’s Disease. Biochim. Biophys. Acta 2007, 1768, 1943–1951. [Google Scholar] [CrossRef]
  189. Ariga, T.; McDonald, M.P.; Yu, R.K. Role of Ganglioside Metabolism in the Pathogenesis of Alzheimer’s Disease—A Review. J. Lipid Res. 2008, 49, 1157–1175. [Google Scholar] [CrossRef]
  190. Fusco, M.; Vantini, G.; Schiavo, N.; Zanotti, A.; Zanoni, R.; Facci, L.; Skaper, S.D. Gangliosides and Neurotrophic Factors in Neurodegenerative Diseases: From Experimental Findings to Clinical Perspectives. Ann. N. Y. Acad. Sci. 1993, 695, 314–317. [Google Scholar] [CrossRef]
  191. He, X.; Huang, Y.; Li, B.; Gong, C.X.; Schuchman, E.H. Deregulation of Sphingolipid Metabolism in Alzheimer’s Disease. Neurobiol. Aging 2010, 31, 398–408. [Google Scholar] [CrossRef]
  192. Matsuzaki, K. How Do Membranes Initiate Alzheimer’s Disease? Formation of Toxic Amyloid Fibrils by the Amyloid Β-Protein on Ganglioside Clusters. Acc. Chem. Res. 2014, 47, 2397–2404. [Google Scholar] [CrossRef] [PubMed]
  193. Yanagisawa, K.; Odaka, A.; Suzuki, N.; Ihara, Y. Gm1 Ganglioside-Bound Amyloid Beta-Protein (a Beta): A Possible Form of Preamyloid in Alzheimer’s Disease. Nat. Med. 1995, 1, 1062–1066. [Google Scholar] [CrossRef] [PubMed]
  194. Yanagisawa, K.; Ihara, Y. Gm1 Ganglioside-Bound Amyloid Beta-Protein in Alzheimer’s Disease Brain. Neurobiol. Aging 1998, 19, S65–S67. [Google Scholar] [CrossRef]
  195. Hayashi, H.; Kimura, N.; Yamaguchi, H.; Hasegawa, K.; Yokoseki, T.; Shibata, M.; Yamamoto, N.; Michikawa, M.; Yoshikawa, Y.; Terao, K.; et al. A Seed for Alzheimer Amyloid in the Brain. J. Neurosci. 2004, 24, 4894–4902. [Google Scholar] [CrossRef]
  196. Yanagisawa, K. Gm1 Ganglioside and the Seeding of Amyloid in Alzheimer’s Disease: Endogenous Seed for Alzheimer Amyloid. Neuroscientist 2005, 11, 250–260. [Google Scholar] [CrossRef]
  197. Yamamoto, N.; Van Nostrand, W.E.; Yanagisawa, K. Further Evidence of Local Ganglioside-Dependent Amyloid Beta-Protein Assembly in Brain. Neuroreport 2006, 17, 1735–1737. [Google Scholar] [CrossRef] [PubMed]
  198. Yamamoto, N.; Matsubara, E.; Maeda, S.; Minagawa, H.; Takashima, A.; Maruyama, W.; Michikawa, M.; Yanagisawa, K. A Ganglioside-Induced Toxic Soluble Abeta Assembly. Its Enhanced Formation from Abeta Bearing the Arctic Mutation. J. Biol. Chem. 2007, 282, 2646–2655. [Google Scholar] [CrossRef]
  199. Yamamoto, N.; Fukata, Y.; Fukata, M.; Yanagisawa, K. Gm1-Ganglioside-Induced Abeta Assembly on Synaptic Membranes of Cultured Neurons. Biochim. Biophys. Acta 2007, 1768, 1128–1137. [Google Scholar] [CrossRef]
  200. Oikawa, N.; Yamaguchi, H.; Ogino, K.; Taki, T.; Yuyama, K.; Yamamoto, N.; Shin, R.W.; Furukawa, K.; Yanagisawa, K. Gangliosides Determine the Amyloid Pathology of Alzheimer’s Disease. Neuroreport 2009, 20, 1043–1046. [Google Scholar] [CrossRef]
  201. Yanagisawa, M.; Ariga, T.; Yu, R.K. Cytotoxic Effects of G(M1) Ganglioside and Amyloid Β-Peptide on Mouse Embryonic Neural Stem Cells. ASN Neuro 2010, 2, e00029. [Google Scholar] [CrossRef]
  202. Yanagisawa, K. Gm1 Ganglioside and Alzheimer’s Disease. Glycoconj. J. 2015, 32, 87–91. [Google Scholar] [CrossRef] [PubMed]
  203. Amaro, M.; Šachl, R.; Aydogan, G.; Mikhalyov, I.I.; Vácha, R.; Hof, M. Gm1 Ganglioside Inhibits Β-Amyloid Oligomerization Induced by Sphingomyelin. Angew. Chem. Int. Ed. 2016, 55, 9411–9415. [Google Scholar] [CrossRef] [PubMed]
  204. Fernández-Pérez, E.J.; Sepúlveda, F.J.; Peoples, R.; Aguayo, L.G. Role of Membrane Gm1 on Early Neuronal Membrane Actions of Aβ during Onset of Alzheimer’s Disease. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 3105–3116. [Google Scholar] [CrossRef]
  205. Matsuzaki, K.; Kato, K.; Yanagisawa, K. Ganglioside-Mediated Assembly of Amyloid Β-Protein: Roles in Alzheimer’s Disease. Prog. Mol. Biol. Transl. Sci. 2018, 156, 413–434. [Google Scholar] [PubMed]
  206. Hong, S.; Ostaszewski, B.L.; Yang, T.; O’Malley, T.T.; Jin, M.; Yanagisawa, K.; Li, S.; Bartels, T.; Selkoe, D.J. Soluble Aβ Oligomers Are Rapidly Sequestered from Brain Isf in Vivo and Bind Gm1 Ganglioside on Cellular Membranes. Neuron 2014, 82, 308–319. [Google Scholar] [CrossRef]
  207. Yuyama, K.; Yanagisawa, K. Sphingomyelin Accumulation Provides a Favorable Milieu for Gm1 Ganglioside-Induced Assembly of Amyloid Beta-Protein. Neurosci. Lett. 2010, 481, 168–172. [Google Scholar] [CrossRef] [PubMed]
  208. Yamamoto, N.; Matsubara, T.; Sato, T.; Yanagisawa, K. Age-Dependent High-Density Clustering of Gm1 Ganglioside at Presynaptic Neuritic Terminals Promotes Amyloid Beta-Protein Fibrillogenesis. Biochim. Biophys. Acta 2008, 1778, 2717–2726. [Google Scholar] [CrossRef]
  209. Matsubara, T.; Nishihara, M.; Yasumori, H.; Nakai, M.; Yanagisawa, K.; Sato, T. Size and Shape of Amyloid Fibrils Induced by Ganglioside Nanoclusters: Role of Sialyl Oligosaccharide in Fibril Formation. Langmuir 2017, 33, 13874–13881. [Google Scholar] [CrossRef]
  210. Kakio, A.; Nishimoto, S.; Yanagisawa, K.; Kozutsumi, Y.; Matsuzaki, K. Interactions of Amyloid Beta-Protein with Various Gangliosides in Raft-Like Membranes: Importance of Gm1 Ganglioside-Bound Form as an Endogenous Seed for Alzheimer Amyloid. Biochemistry 2002, 41, 7385–7390. [Google Scholar] [CrossRef]
  211. Hirano-Sakamaki, W.; Sugiyama, E.; Hayasaka, T.; Ravid, R.; Setou, M.; Taki, T. Alzheimer’s Disease Is Associated with Disordered Localization of Ganglioside Gm1 Molecular Species in the Human Dentate Gyrus. FEBS Lett. 2015, 589, 3611–3616. [Google Scholar] [CrossRef]
  212. Yamamoto, N.; Tanida, M.; Kasahara, R.; Sobue, K.; Suzuki, K. Leptin Inhibits Amyloid Β-Protein Fibrillogenesis by Decreasing Gm1 Gangliosides on the Neuronal Cell Surface through Pi3k/Akt/Mtor Pathway. J. Neurochem. 2014, 131, 323–332. [Google Scholar] [CrossRef] [PubMed]
  213. Yang, R.; Wang, Q.; Min, L.; Sui, R.; Li, J.; Liu, X. Monosialoanglioside Improves Memory Deficits and Relieves Oxidative Stress in the Hippocampus of Rat Model of Alzheimer’s Disease. Neurol. Sci. 2013, 34, 1447–1451. [Google Scholar] [CrossRef] [PubMed]
  214. Itokazu, Y.; Kato-Negishi, M.; Nakatani, Y.; Ariga, T.; Yu, R.K. Effects of Amyloid Β-Peptides and Gangliosides on Mouse Neural Stem Cells. Neurochem. Res. 2013, 38, 2019–2027. [Google Scholar] [CrossRef]
  215. Kreutz, F.; Frozza, R.L.; Breier, A.C.; de Oliveira, V.A.; Horn, A.P.; Pettenuzzo, L.F.; Netto, C.A.; Salbego, C.G.; Trindade, V.M. Amyloid-Β Induced Toxicity Involves Ganglioside Expression and Is Sensitive to Gm1 Neuroprotective Action. Neurochem. Int. 2011, 59, 648–655. [Google Scholar] [CrossRef] [PubMed]
  216. Sokolova, T.V.; Zakharova, I.O.; Furaev, V.V.; Rychkova, M.P.; Avrova, N.F. Neuroprotective Effect of Ganglioside Gm1 on the Cytotoxic Action of Hydrogen Peroxide and Amyloid Beta-Peptide in Pc12 Cells. Neurochem. Res. 2007, 32, 1302–1313. [Google Scholar] [CrossRef] [PubMed]
  217. Zha, Q.; Ruan, Y.; Hartmann, T.; Beyreuther, K.; Zhang, D. Gm1 Ganglioside Regulates the Proteolysis of Amyloid Precursor Protein. Mol. Psychiatry 2004, 9, 946–952. [Google Scholar] [CrossRef]
  218. Matsubara, T.; Iijima, K.; Yamamoto, N.; Yanagisawa, K.; Sato, T. Density of Gm1 in Nanoclusters Is a Critical Factor in the Formation of a Spherical Assembly of Amyloid Β-Protein on Synaptic Plasma Membranes. Langmuir 2013, 29, 2258–2264. [Google Scholar] [CrossRef]
  219. Kakio, A.; Nishimoto, S.I.; Yanagisawa, K.; Kozutsumi, Y.; Matsuzaki, K. Cholesterol-Dependent Formation of Gm1 Ganglioside-Bound Amyloid Beta-Protein, an Endogenous Seed for Alzheimer Amyloid. J. Biol. Chem. 2001, 276, 24985–24990. [Google Scholar] [CrossRef]
  220. Dai, Y.; Zhang, M.; Shi, X.; Wang, K.; Gao, G.; Shen, L.; Sun, T. Kinetic Study of Aβ(1-42) Amyloidosis in the Presence of Ganglioside-Containing Vesicles. Colloids Surf. B Biointerfaces 2020, 185, e110615. [Google Scholar] [CrossRef]
  221. Ikeda, K.; Yamaguchi, T.; Fukunaga, S.; Hoshino, M.; Matsuzaki, K. Mechanism of Amyloid Β-Protein Aggregation Mediated by Gm1 Ganglioside Clusters. Biochemistry 2011, 50, 6433–6440. [Google Scholar] [CrossRef]
  222. Kim, S.I.; Yi, J.S.; Ko, Y.G. Amyloid Beta Oligomerization Is Induced by Brain Lipid Rafts. J. Cell. Biochem. 2006, 99, 878–889. [Google Scholar] [CrossRef] [PubMed]
  223. Yamamoto, N.; Arima, H.; Sugiura, T.; Hirate, H.; Kusama, N.; Suzuki, K.; Sobue, K. Midazolam Inhibits the Formation of Amyloid Fibrils and Gm1 Ganglioside-Rich Microdomains in Presynaptic Membranes through the Gamma-Aminobutyric Acid a Receptor. Biochem. Biophys. Res. Commun. 2015, 457, 547–553. [Google Scholar] [CrossRef] [PubMed]
  224. Dukhinova, M.; Veremeyko, T.; Yung, A.W.Y.; Kuznetsova, I.S.; Lau, T.Y.B.; Kopeikina, E.; Chan, A.M.L.; Ponomarev, E.D. Fresh Evidence for Major Brain Gangliosides as a Target for the Treatment of Alzheimer’s Disease. Neurobiol. Aging 2019, 77, 128–143. [Google Scholar] [CrossRef]
  225. Yahi, N.; Fantini, J. Deciphering the Glycolipid Code of Alzheimer’s and Parkinson’s Amyloid Proteins Allowed the Creation of a Universal Ganglioside-Binding Peptide. PLoS ONE 2014, 9, e104751. [Google Scholar] [CrossRef] [PubMed]
  226. Beitz, J.M. Parkinson’s Disease: A Review. Front. Biosci. 2014, 6, 65–74. [Google Scholar] [CrossRef]
  227. Herrero, M.T.; Kastner, A.; Perez-Otaño, I.; Hirsch, E.C.; Luquin, M.R.; Javoy-Agid, F.; Del Río, J.; Obeso, J.A.; Agid, Y. Gangliosides and Parkinsonism. Neurology 1993, 43, 2132–2134. [Google Scholar] [CrossRef]
  228. Wallom, K.L.; Fernández-Suárez, M.E.; Priestman, D.A.; Te Vruchte, D.; Huebecker, M.; Hallett, P.J.; Isacson, O.; Platt, F.M. Glycosphingolipid Metabolism and Its Role in Ageing and Parkinson’s Disease. Glycoconj. J. 2021, 39, 39–53. [Google Scholar] [CrossRef]
  229. Huebecker, M.; Moloney, E.B.; van der Spoel, A.C.; Priestman, D.A.; Isacson, O.; Hallett, P.J.; Platt, F.M. Reduced Sphingolipid Hydrolase Activities, Substrate Accumulation and Ganglioside Decline in Parkinson’s Disease. Mol. Neurodegener. 2019, 14, e40. [Google Scholar] [CrossRef]
  230. Ledeen, R.; Chowdhury, S.; Lu, Z.H.; Chakraborty, M.; Wu, G. Systemic Deficiency of Gm1 Ganglioside in Parkinson’s Disease Tissues and Its Relation to the Disease Etiology. Glycoconj. J. 2022, 39, 75–82. [Google Scholar] [CrossRef]
  231. Chowdhury, S.; Ledeen, R. The Key Role of Gm1 Ganglioside in Parkinson’s Disease. Biomolecules 2022, 12, 173. [Google Scholar] [CrossRef]
  232. Fazzari, M.; Di Biase, E.; Lunghi, G.; Mauri, L.; Chiricozzi, E.; Sonnino, S. Novel Insights on Gm1 and Parkinson’s Disease: A Critical Review. Glycoconj. J. 2022, 39, 27–38. [Google Scholar] [CrossRef]
  233. Sonnino, S. The Relationship between Depletion of Brain Gm1 Ganglioside and Parkinson’s Disease. FEBS Open Bio 2023. [Google Scholar] [CrossRef] [PubMed]
  234. Wu, G.; Lu, Z.H.; Kulkarni, N.; Ledeen, R.W. Deficiency of Ganglioside Gm1 Correlates with Parkinson’s Disease in Mice and Humans. J. Neurosci. Res. 2012, 90, 1997–2008. [Google Scholar] [CrossRef] [PubMed]
  235. Wu, G.; Lu, Z.H.; Seo, J.H.; Alselehdar, S.K.; DeFrees, S.; Ledeen, R.W. Mice Deficient in Gm1 Manifest Both Motor and Non-Motor Symptoms of Parkinson’s Disease; Successful Treatment with Synthetic Gm1 Ganglioside. Exp. Neurol. 2020, 329, e113284. [Google Scholar] [CrossRef]
  236. Schneider, J.S. Altered Expression of Genes Involved in Ganglioside Biosynthesis in Substantia Nigra Neurons in Parkinson’s Disease. PLoS ONE 2018, 13, e0199189. [Google Scholar] [CrossRef]
  237. Wu, G.; Lu, Z.H.; Kulkarni, N.; Amin, R.; Ledeen, R.W. Mice Lacking Major Brain Gangliosides Develop Parkinsonism. Neurochem. Res. 2011, 36, 1706–1714. [Google Scholar] [CrossRef]
  238. Verma, M.; Schneider, J.S. Sirna-Mediated Knockdown of B3galt4 Decreases Gm1 Ganglioside Expression and Enhances Vulnerability for Neurodegeneration. Mol. Cell. Neurosci. 2019, 95, 25–30. [Google Scholar] [CrossRef] [PubMed]
  239. Schneider, J.S.; Aras, R.; Williams, C.K.; Koprich, J.B.; Brotchie, J.M.; Singh, V. Gm1 Ganglioside Modifies A-Synuclein Toxicity and Is Neuroprotective in a Rat A-Synuclein Model of Parkinson’s Disease. Sci. Rep. 2019, 9, e8362. [Google Scholar] [CrossRef]
  240. Guo, Y.L.; Duan, W.J.; Lu, D.H.; Ma, X.H.; Li, X.X.; Li, Z.; Bi, W.; Kurihara, H.; Liu, H.Z.; Li, Y.F.; et al. Autophagy-Dependent Removal of A-Synuclein: A Novel Mechanism of Gm1 Ganglioside Neuroprotection against Parkinson’s Disease. Acta Pharmacol. Sin. 2021, 42, 518–528. [Google Scholar] [CrossRef]
  241. Akkhawattanangkul, Y.; Maiti, P.; Xue, Y.; Aryal, D.; Wetsel, W.C.; Hamilton, D.; Fowler, S.C.; McDonald, M.P. Targeted Deletion of Gd3 Synthase Protects against Mptp-Induced Neurodegeneration. Genes Brain Behav. 2017, 16, 522–536. [Google Scholar] [CrossRef]
  242. Dhanushkodi, A.; Xue, Y.; Roguski, E.E.; Ding, Y.; Matta, S.G.; Heck, D.; Fan, G.H.; McDonald, M.P. Lentiviral-Mediated Knock-Down of Gd3 Synthase Protects against Mptp-Induced Motor Deficits and Neurodegeneration. Neurosci. Lett. 2019, 692, 53–63. [Google Scholar] [CrossRef]
  243. Xu, R.; Zhou, Y.; Fang, X.; Lu, Y.; Li, J.; Zhang, J.; Deng, X.; Li, S. The Possible Mechanism of Parkinson’s Disease Progressive Damage and the Preventive Effect of Gm1 in the Rat Model Induced by 6-Hydroxydopamine. Brain Res. 2014, 1592, 73–81. [Google Scholar] [CrossRef]
  244. Martinez, Z.; Zhu, M.; Han, S.; Fink, A.L. Gm1 Specifically Interacts with Alpha-Synuclein and Inhibits Fibrillation. Biochemistry 2007, 46, 1868–1877. [Google Scholar] [CrossRef]
  245. Bartels, T.; Kim, N.C.; Luth, E.S.; Selkoe, D.J. N-Alpha-Acetylation of A-Synuclein Increases Its Helical Folding Propensity, Gm1 Binding Specificity and Resistance to Aggregation. PLoS ONE 2014, 9, e103727. [Google Scholar] [CrossRef] [PubMed]
  246. Ba, X.H. Therapeutic Effects of Gm1 on Parkinson’s Disease in Rats and Its Mechanism. Int. J. Neurosci. 2016, 126, 163–167. [Google Scholar] [CrossRef]
  247. Schneider, J.S.; Pope, A.; Simpson, K.; Taggart, J.; Smith, M.G.; DiStefano, L. Recovery from Experimental Parkinsonism in Primates with Gm1 Ganglioside Treatment. Science 1992, 256, 843–846. [Google Scholar] [CrossRef] [PubMed]
  248. Schneider, J.S.; Yuwiler, A. Gm1 Ganglioside Treatment Promotes Recovery of Striatal Dopamine Concentrations in the Mouse Model of Mptp-Induced Parkinsonism. Exp. Neurol. 1989, 105, 177–183. [Google Scholar] [CrossRef] [PubMed]
  249. Schneider, J.S.; Sendek, S.; Daskalakis, C.; Cambi, F. Gm1 Ganglioside in Parkinson’s Disease: Results of a Five Year Open Study. J. Neurol. Sci. 2010, 292, 45–51. [Google Scholar] [CrossRef]
  250. Nicoli, E.R.; Annunziata, I.; d’Azzo, A.; Platt, F.M.; Tifft, C.J.; Stepien, K.M. Gm1 Gangliosidosis—A Mini-Review. Front. Genet. 2021, 12, 734878. [Google Scholar] [CrossRef]
  251. Rha, A.K.; Maguire, A.S.; Martin, D.R. Gm1 Gangliosidosis: Mechanisms and Management. Appl. Clin. Genet. 2021, 14, 209–233. [Google Scholar] [CrossRef]
  252. Brunetti-Pierri, N.; Scaglia, F. Gm1 Gangliosidosis: Review of Clinical, Molecular, and Therapeutic Aspects. Mol. Genet. Metab. 2008, 94, 391–396. [Google Scholar] [CrossRef] [PubMed]
  253. Satoh, H.; Yamato, O.; Asano, T.; Yonemura, M.; Yamauchi, T.; Hasegawa, D.; Orima, H.; Arai, T.; Yamasaki, M.; Maede, Y. Cerebrospinal Fluid Biomarkers Showing Neurodegeneration in Dogs with Gm1 Gangliosidosis: Possible Use for Assessment of a Therapeutic Regimen. Brain Res. 2007, 1133, 200–208. [Google Scholar] [CrossRef]
  254. Satoh, H.; Yamato, O.; Asano, T.; Yamasaki, M.; Maede, Y. Increased Concentration of Gm1-Ganglioside in Cerebrospinal Fluid in Dogs with Gm1- and Gm2-Gangliosidoses and Its Clinical Application for Diagnosis. J. Vet. Diagn. Investig. 2004, 16, 223–226. [Google Scholar] [CrossRef] [PubMed]
  255. Kaye, E.M.; Alroy, J.; Raghavan, S.S.; Schwarting, G.A.; Adelman, L.S.; Runge, V.; Gelblum, D.; Thalhammer, J.G.; Zuniga, G. Dysmyelinogenesis in Animal Model of Gm1 Gangliosidosis. Pediatr. Neurol. 1992, 8, 255–261. [Google Scholar] [CrossRef]
  256. Folkerth, R.D.; Alroy, J.; Bhan, I.; Kaye, E.M. Infantile G(M1) Gangliosidosis: Complete Morphology and Histochemistry of Two Autopsy Cases, with Particular Reference to Delayed Central Nervous System Myelination. Pediatr. Dev. Pathol. 2000, 3, 73–86. [Google Scholar] [CrossRef]
  257. Heinecke, K.A.; Luoma, A.; d’Azzo, A.; Kirschner, D.A.; Seyfried, T.N. Myelin Abnormalities in the Optic and Sciatic Nerves in Mice with Gm1-Gangliosidosis. ASN Neuro 2015, 7, 1759091415568913. [Google Scholar] [CrossRef]
  258. van der Voorn, J.P.; Kamphorst, W.; van der Knaap, M.S.; Powers, J.M. The Leukoencephalopathy of Infantile Gm1 Gangliosidosis: Oligodendrocytic Loss and Axonal Dysfunction. Acta Neuropathol. 2004, 107, 539–545. [Google Scholar] [CrossRef]
  259. Jeyakumar, M.; Thomas, R.; Elliot-Smith, E.; Smith, D.A.; van der Spoel, A.C.; d’Azzo, A.; Perry, V.H.; Butters, T.D.; Dwek, R.A.; Platt, F.M. Central Nervous System Inflammation Is a Hallmark of Pathogenesis in Mouse Models of Gm1 and Gm2 Gangliosidosis. Brain 2003, 126, 974–987. [Google Scholar] [CrossRef] [PubMed]
  260. Tessitore, A.; Martin, M.D.P.; Sano, R.; Ma, Y.; Mann, L.; Ingrassia, A.; Laywell, E.D.; Steindler, D.A.; Hendershot, L.M.; d’Azzo, A. Gm1-Ganglioside-Mediated Activation of the Unfolded Protein Response Causes Neuronal Death in a Neurodegenerative Gangliosidosis. Mol. Cell 2004, 15, 753–766. [Google Scholar] [CrossRef]
  261. Koenig, M.L.; Jope, R.S.; Baker, H.J.; Lally, K.M. Reduced Ca2+ Flux in Synaptosomes from Cats with Gm1 Gangliosidosis. Brain Res. 1987, 424, 169–176. [Google Scholar] [CrossRef]
  262. Wu, G.; Lu, Z.H.; Ledeen, R.W. Gm1 Ganglioside in the Nuclear Membrane Modulates Nuclear Calcium Homeostasis during Neurite Outgrowth. J. Neurochem. 1995, 65, 1419–1422. [Google Scholar] [CrossRef]
  263. Singer, H.S.; Coyle, J.T.; Weaver, D.L.; Kawamura, N.; Baker, H.J. Neurotransmitter Chemistry in Feline Gm1 Gangliosidosis: A Model for Human Ganglioside Storage Disease. Ann. Neurol. 1982, 12, 37–41. [Google Scholar] [CrossRef]
  264. Takamura, A.; Higaki, K.; Kajimaki, K.; Otsuka, S.; Ninomiya, H.; Matsuda, J.; Ohno, K.; Suzuki, Y.; Nanba, E. Enhanced Autophagy and Mitochondrial Aberrations in Murine G(M1)-Gangliosidosis. Biochem. Biophys. Res. Commun. 2008, 367, 616–622. [Google Scholar] [CrossRef]
  265. Takamura, A.; Higaki, K.; Ninomiya, H.; Takai, T.; Matsuda, J.; Iida, M.; Ohno, K.; Suzuki, Y.; Nanba, E. Lysosomal Accumulation of Trk Protein in Brain of GM1-Gangliosidosis Mouse and Its Restoration by Chemical Chaperone. J. Neurochem. 2011, 118, 399–406. [Google Scholar] [CrossRef]
  266. Condori, J.; Acosta, W.; Ayala, J.; Katta, V.; Flory, A.; Martin, R.; Radin, J.; Cramer, C.L.; Radin, D.N. Enzyme Replacement for GM1-Gangliosidosis: Uptake, Lysosomal Activation, and Cellular Disease Correction Using a Novel β-Galactosidase:Rtb Lectin Fusion. Mol. Genet. Metab. 2016, 117, 199–209. [Google Scholar] [CrossRef] [PubMed]
  267. Przybilla, M.J.; Stewart, C.; Carlson, T.W.; Ou, L.; Koniar, B.L.; Sidhu, R.; Kell, P.J.; Jiang, X.; Jarnes, J.R.; O’Sullivan, M.G.; et al. Examination of a Blood-Brain Barrier Targeting Β-Galactosidase-Monoclonal Antibody Fusion Protein in a Murine Model of Gm1-Gangliosidosis. Mol. Genet. Metab. Rep. 2021, 27, e100748. [Google Scholar] [CrossRef]
  268. Gupta, M.; Pandey, H.; Sivakumar, S. Intracellular Delivery of Β-Galactosidase Enzyme Using Arginase-Responsive Dextran Sulfate/Poly-L-Arginine Capsule for Lysosomal Storage Disorder. ACS Omega 2017, 2, 9002–9012. [Google Scholar] [CrossRef] [PubMed]
  269. Kelly, J.M.; Gross, A.L.; Martin, D.R.; Byrne, M.E. Polyethylene Glycol-B-Poly(Lactic Acid) Polymersomes as Vehicles for Enzyme Replacement Therapy. Nanomedicine 2017, 12, 2591–2606. [Google Scholar] [CrossRef]
  270. Chen, J.C.; Luu, A.R.; Wise, N.; Angelis, R.; Agrawal, V.; Mangini, L.; Vincelette, J.; Handyside, B.; Sterling, H.; Lo, M.J.; et al. Intracerebroventricular Enzyme Replacement Therapy with Β-Galactosidase Reverses Brain Pathologies due to Gm1 Gangliosidosis in Mice. J. Biol. Chem. 2020, 295, 13532–13555. [Google Scholar] [CrossRef]
  271. Caciotti, A.; Donati, M.A.; d’Azzo, A.; Salvioli, R.; Guerrini, R.; Zammarchi, E.; Morrone, A. The Potential Action of Galactose as a “Chemical Chaperone”: Increase of Beta Galactosidase Activity in Fibroblasts from an Adult Gm1-Gangliosidosis Patient. Eur. J. Paediatr. Neurol. 2009, 13, 160–164. [Google Scholar] [CrossRef] [PubMed]
  272. Higaki, K.; Ninomiya, H.; Suzuki, Y.; Nanba, E. Candidate Molecules for Chemical Chaperone Therapy of Gm1-Gangliosidosis. Futur. Med. Chem. 2013, 5, 1551–1558. [Google Scholar] [CrossRef]
  273. Rigat, B.A.; Tropak, M.B.; Buttner, J.; Crushell, E.; Benedict, D.; Callahan, J.W.; Martin, D.R.; Mahuran, D.J. Evaluation of N-Nonyl-Deoxygalactonojirimycin as a Pharmacological Chaperone for Human Gm1 Gangliosidosis Leads to Identification of a Feline Model Suitable for Testing Enzyme Enhancement Therapy. Mol. Genet. Metab. 2012, 107, 203–212. [Google Scholar] [CrossRef] [PubMed]
  274. Front, S.; Biela-Banaś, A.; Burda, P.; Ballhausen, D.; Higaki, K.; Caciotti, A.; Morrone, A.; Charollais-Thoenig, J.; Gallienne, E.; Demotz, S.; et al. (5ar)-5a-C-Pentyl-4-Epi-Isofagomine: A Powerful Inhibitor of Lysosomal Β-Galactosidase and a Remarkable Chaperone for Mutations Associated with Gm1-Gangliosidosis and Morquio Disease Type B. Eur. J. Med. Chem. 2017, 126, 160–170. [Google Scholar] [CrossRef] [PubMed]
  275. Takai, T.; Higaki, K.; Aguilar-Moncayo, M.; Mena-Barragán, T.; Hirano, Y.; Yura, K.; Yu, L.; Ninomiya, H.; García-Moreno, M.I.; Sakakibara, Y.; et al. A Bicyclic 1-Deoxygalactonojirimycin Derivative as a Novel Pharmacological Chaperone for Gm1 Gangliosidosis. Mol. Ther. 2013, 21, 526–532. [Google Scholar] [CrossRef] [PubMed]
  276. Suzuki, Y.; Ichinomiya, S.; Kurosawa, M.; Matsuda, J.; Ogawa, S.; Iida, M.; Kubo, T.; Tabe, M.; Itoh, M.; Higaki, K.; et al. Therapeutic Chaperone Effect of N-Octyl 4-Epi-Β-Valienamine on Murine G(M1)-Gangliosidosis. Mol. Genet. Metab. 2012, 106, 92–98. [Google Scholar] [CrossRef] [PubMed]
  277. Thonhofer, M.; Gonzalez Santana, A.; Fischer, R.; Torvisco Gomez, A.; Saf, R.; Schalli, M.; Stütz, A.E.; Withers, S.G. 5-Fluoro Derivatives of 4-Epi-Isofagomine as D-Galactosidase Inhibitors and Potential Pharmacological Chaperones for Gm1-Gangliosidosis as Well as Fabry’s Disease. Carbohydr. Res. 2016, 420, 6–12. [Google Scholar] [CrossRef] [PubMed]
  278. Fantur, K.; Hofer, D.; Schitter, G.; Steiner, A.J.; Pabst, B.M.; Wrodnigg, T.M.; Stütz, A.E.; Paschke, E. Dlhex-Dgj, a Novel Derivative of 1-Deoxygalactonojirimycin with Pharmacological Chaperone Activity in Human G(M1)-Gangliosidosis Fibroblasts. Mol. Genet. Metab. 2010, 100, 262–268. [Google Scholar] [CrossRef] [PubMed]
  279. Kasperzyk, J.L.; El-Abbadi, M.M.; Hauser, E.C.; D’Azzo, A.; Platt, F.M.; Seyfried, T.N. N-Butyldeoxygalactonojirimycin Reduces Neonatal Brain Ganglioside Content in a Mouse Model of Gm1 Gangliosidosis. J. Neurochem. 2004, 89, 645–653. [Google Scholar] [CrossRef]
  280. Kasperzyk, J.L.; d’Azzo, A.; Platt, F.M.; Alroy, J.; Seyfried, T.N. Substrate Reduction Reduces Gangliosides in Postnatal Cerebrum-Brainstem and Cerebellum in Gm1 Gangliosidosis Mice. J. Lipid Res. 2005, 46, 744–751. [Google Scholar] [CrossRef]
  281. Suzuki, Y.; Ichinomiya, S.; Kurosawa, M.; Ohkubo, M.; Watanabe, H.; Iwasaki, H.; Matsuda, J.; Noguchi, Y.; Takimoto, K.; Itoh, M.; et al. Chemical Chaperone Therapy: Clinical Effect in Murine G(M1)-Gangliosidosis. Ann. Neurol. 2007, 62, 671–675. [Google Scholar] [CrossRef]
  282. Suzuki, Y. Chemical Chaperone Therapy for Gm1-Gangliosidosis. Cell. Mol. Life Sci. 2008, 65, 351–353. [Google Scholar] [CrossRef]
  283. Higaki, K.; Li, L.; Bahrudin, U.; Okuzawa, S.; Takamuram, A.; Yamamoto, K.; Adachi, K.; Paraguison, R.C.; Takai, T.; Ikehata, H.; et al. Chemical Chaperone Therapy: Chaperone Effect on Mutant Enzyme and Cellular Pathophysiology in Β-Galactosidase Deficiency. Hum. Mutat. 2011, 32, 843–852. [Google Scholar] [CrossRef]
  284. Matsuda, J.; Suzuki, O.; Oshima, A.; Yamamoto, Y.; Noguchi, A.; Takimoto, K.; Itoh, M.; Matsuzaki, Y.; Yasuda, Y.; Ogawa, S.; et al. Chemical Chaperone Therapy for Brain Pathology in G(M1)-Gangliosidosis. Proc. Natl. Acad. Sci. USA 2003, 100, 15912–15917. [Google Scholar] [CrossRef]
  285. Elliot-Smith, E.; Speak, A.O.; Lloyd-Evans, E.; Smith, D.A.; van der Spoel, A.C.; Jeyakumar, M.; Butters, T.D.; Dwek, R.A.; d’Azzo, A.; Platt, F.M. Beneficial Effects of Substrate Reduction Therapy in a Mouse Model of Gm1 Gangliosidosis. Mol. Genet. Metab. 2008, 94, 204–211. [Google Scholar] [CrossRef]
  286. Deodato, F.; Procopio, E.; Rampazzo, A.; Taurisano, R.; Donati, M.A.; Dionisi-Vici, C.; Caciotti, A.; Morrone, A.; Scarpa, M. The Treatment of Juvenile/Adult Gm1-Gangliosidosis with Miglustat May Reverse Disease Progression. Metab. Brain Dis. 2017, 32, 1529–1536. [Google Scholar]
  287. Hayward, C.; Patel, H.C.; Manohar, S.G.; Lyon, A.R. Gene Therapy for Gm1 Gangliosidosis: Challenges of Translational Medicine. Ann. Transl. Med. 2015, 3, S28. [Google Scholar]
  288. Takaura, N.; Yagi, T.; Maeda, M.; Nanba, E.; Oshima, A.; Suzuki, Y.; Yamano, T.; Tanaka, A. Attenuation of Ganglioside Gm1 Accumulation in the Brain of Gm1 Gangliosidosis Mice by Neonatal Intravenous Gene Transfer. Gene Ther. 2003, 10, 1487–1493. [Google Scholar] [CrossRef] [PubMed]
  289. Broekman, M.L.; Baek, R.C.; Comer, L.A.; Fernandez, J.L.; Seyfried, T.N.; Sena-Esteves, M. Complete Correction of Enzymatic Deficiency and Neurochemistry in the Gm1-Gangliosidosis Mouse Brain by Neonatal Adeno-Associated Virus-Mediated Gene Delivery. Mol. Ther. 2007, 15, 30–37. [Google Scholar] [CrossRef] [PubMed]
  290. Baek, R.C.; Broekman, M.L.; Leroy, S.G.; Tierney, L.A.; Sandberg, M.A.; d’Azzo, A.; Seyfried, T.N.; Sena-Esteves, M. Aav-Mediated Gene Delivery in Adult Gm1-Gangliosidosis Mice Corrects Lysosomal Storage in Cns and Improves Survival. PLoS ONE 2010, 5, e13468. [Google Scholar] [CrossRef] [PubMed]
  291. Weismann, C.M.; Ferreira, J.; Keeler, A.M.; Su, Q.; Qui, L.; Shaffer, S.A.; Xu, Z.; Gao, G.; Sena-Esteves, M. Systemic Aav9 Gene Transfer in Adult Gm1 Gangliosidosis Mice Reduces Lysosomal Storage in Cns and Extends Lifespan. Hum. Mol. Genet. 2015, 24, 4353–4364. [Google Scholar] [CrossRef]
  292. Broekman, M.L.; Tierney, L.A.; Benn, C.; Chawla, P.; Cha, J.H.; Sena-Esteves, M. Mechanisms of Distribution of Mouse Beta-Galactosidase in the Adult Gm1-Gangliosidosis Brain. Gene Ther. 2009, 16, 303–308. [Google Scholar] [CrossRef]
  293. McCurdy, V.J.; Johnson, A.K.; Gray-Edwards, H.L.; Randle, A.N.; Brunson, B.L.; Morrison, N.E.; Salibi, N.; Johnson, J.A.; Hwang, M.; Beyers, R.J.; et al. Sustained Normalization of Neurological Disease after Intracranial Gene Therapy in a Feline Model. Sci. Transl. Med. 2014, 6, 231ra248. [Google Scholar] [CrossRef] [PubMed]
  294. Foust, K.D.; Nurre, E.; Montgomery, C.L.; Hernandez, A.; Chan, C.M.; Kaspar, B.K. Intravascular Aav9 Preferentially Targets Neonatal Neurons and Adult Astrocytes. Nat. Biotechnol. 2009, 27, 59–65. [Google Scholar] [CrossRef]
  295. Gray-Edwards, H.L.; Regier, D.S.; Shirley, J.L.; Randle, A.N.; Salibi, N.; Thomas, S.E.; Latour, Y.L.; Johnston, J.; Golas, G.; Maguire, A.S.; et al. Novel Biomarkers of Human Gm1 Gangliosidosis Reflect the Clinical Efficacy of Gene Therapy in a Feline Model. Mol. Ther. 2017, 25, 892–903. [Google Scholar] [CrossRef] [PubMed]
  296. Walker, F.O. Huntington’s Disease. Lancet 2007, 369, 218–228. [Google Scholar] [CrossRef] [PubMed]
  297. Bates, G.P.; Dorsey, R.; Gusella, J.F.; Hayden, M.R.; Kay, C.; Leavitt, B.R.; Nance, M.; Ross, C.A.; Scahill, R.I.; Wetzel, R.; et al. Huntington Disease. Nat. Rev. Dis. Primers 2015, 1, e15005. [Google Scholar] [CrossRef]
  298. Jimenez-Sanchez, M.; Licitra, F.; Underwood, B.R.; Rubinsztein, D.C. Huntington’s Disease: Mechanisms of Pathogenesis and Therapeutic Strategies. Cold Spring Harb. Perspect. Med. 2017, 7, e024240. [Google Scholar] [CrossRef]
  299. Ross, C.A.; Tabrizi, S.J. Huntington’s Disease: From Molecular Pathogenesis to Clinical Treatment. Lancet Neurol. 2011, 10, 83–98. [Google Scholar] [CrossRef]
  300. McColgan, P.; Tabrizi, S.J. Huntington’s Disease: A Clinical Review. Eur. J. Neurol. 2018, 25, 24–34. [Google Scholar] [CrossRef]
  301. Higatsberger, M.R.; Sperk, G.; Bernheimer, H.; Shannak, K.S.; Hornykiewicz, O. Striatal Ganglioside Levels in the Rat Following Kainic Acid Lesions: Comparison with Huntington’s Disease. Exp. Brain Res. 1981, 44, 93–96. [Google Scholar] [CrossRef]
  302. Desplats, P.A.; Denny, C.A.; Kass, K.E.; Gilmartin, T.; Head, S.R.; Sutcliffe, J.G.; Seyfried, T.N.; Thomas, E.A. Glycolipid and Ganglioside Metabolism Imbalances in Huntington’s Disease. Neurobiol. Dis. 2007, 27, 265–277. [Google Scholar] [CrossRef] [PubMed]
  303. Maglione, V.; Marchi, P.; Di Pardo, A.; Lingrell, S.; Horkey, M.; Tidmarsh, E.; Sipione, S. Impaired Ganglioside Metabolism in Huntington’s Disease and Neuroprotective Role of Gm1. J. Neurosci. 2010, 30, 4072–4080. [Google Scholar] [CrossRef] [PubMed]
  304. Denny, C.A.; Desplats, P.A.; Thomas, E.A.; Seyfried, T.N. Cerebellar Lipid Differences between R6/1 Transgenic Mice and Humans with Huntington’s Disease. J. Neurochem. 2010, 115, 748–758. [Google Scholar] [CrossRef]
  305. Di Pardo, A.; Maglione, V.; Alpaugh, M.; Horkey, M.; Atwal, R.S.; Sassone, J.; Ciammola, A.; Steffan, J.S.; Fouad, K.; Truant, R.; et al. Ganglioside Gm1 Induces Phosphorylation of Mutant Huntingtin and Restores Normal Motor Behavior in Huntington Disease Mice. Proc. Natl. Acad. Sci. USA 2012, 109, 3528–3533. [Google Scholar] [CrossRef]
  306. Gu, X.; Greiner, E.R.; Mishra, R.; Kodali, R.; Osmand, A.; Finkbeiner, S.; Steffan, J.S.; Thompson, L.M.; Wetzel, R.; Yang, X.W. Serines 13 and 16 Are Critical Determinants of Full-Length Human Mutant Huntingtin Induced Disease Pathogenesis in Hd Mice. Neuron 2009, 64, 828–840. [Google Scholar] [CrossRef] [PubMed]
  307. Alpaugh, M.; Galleguillos, D.; Forero, J.; Morales, L.C.; Lackey, S.W.; Kar, P.; Di Pardo, A.; Holt, A.; Kerr, B.J.; Todd, K.G.; et al. Disease-Modifying Effects of Ganglioside Gm1 in Huntington’s Disease Models. EMBO Mol. Med. 2017, 9, 1537–1557. [Google Scholar] [CrossRef]
  308. Chaibva, M.; Gao, X.; Jain, P.; Campbell, W.A., 4th; Frey, S.L.; Legleiter, J. Sphingomyelin and Gm1 Influence Huntingtin Binding to, Disruption of, and Aggregation on Lipid Membranes. ACS Omega 2018, 3, 273–285. [Google Scholar] [CrossRef]
  309. de Freitas, R.M.; do Nascimento, K.G.; Ferreira, P.M.; Jordán, J. Neurochemical Changes on Oxidative Stress in Rat Hippocampus During Acute Phase of Pilocarpine-Induced Seizures. Pharmacol. Biochem. Behav. 2010, 94, 341–345. [Google Scholar] [CrossRef]
  310. Izumi, T.; Ogawa, T.; Koizumi, H.; Fukuyama, Y. Low Levels of Csf Gangliotetraose-Series Gangliosides in West Syndrome: Implication of Brain Maturation Disturbance. Pediatr. Neurol. 1993, 9, 293–296. [Google Scholar] [CrossRef]
  311. Simpson, M.A.; Cross, H.; Proukakis, C.; Priestman, D.A.; Neville, D.C.; Reinkensmeier, G.; Wang, H.; Wiznitzer, M.; Gurtz, K.; Verganelaki, A.; et al. Infantile-Onset Symptomatic Epilepsy Syndrome Caused by a Homozygous Loss-of-Function Mutation of Gm3 Synthase. Nat. Genet. 2004, 36, 1225–1229. [Google Scholar] [CrossRef]
  312. Karpiak, S.E.; Mahadik, S.P.; Graf, L.; Rapport, M.M. An Immunological Model of Epilepsy: Seizures Induced by Antibodies to Gm1 Ganglioside. Epilepsia 1981, 22, 189–196. [Google Scholar] [CrossRef]
  313. Fighera, M.R.; Royes, L.F.; Furian, A.F.; Oliveira, M.S.; Fiorenza, N.G.; Frussa-Filho, R.; Petry, J.C.; Coelho, R.C.; Mello, C.F. Gm1 Ganglioside Prevents Seizures, Na+, K+-Atpase Activity Inhibition and Oxidative Stress Induced by Glutaric Acid and Pentylenetetrazole. Neurobiol. Dis. 2006, 22, 611–623. [Google Scholar] [CrossRef]
  314. Prendergast, J.; Umanah, G.K.; Yoo, S.W.; Lagerlöf, O.; Motari, M.G.; Cole, R.N.; Huganir, R.L.; Dawson, T.M.; Dawson, V.L.; Schnaar, R.L. Ganglioside Regulation of Ampa Receptor Trafficking. J. Neurosci. 2014, 34, 13246–13258. [Google Scholar] [CrossRef]
  315. van Es, M.A.; Hardiman, O.; Chio, A.; Al-Chalabi, A.; Pasterkamp, R.J.; Veldink, J.H.; van den Berg, L.H. Amyotrophic Lateral Sclerosis. Lancet 2017, 390, 2084–2098. [Google Scholar] [CrossRef] [PubMed]
  316. Dodge, J.C.; Treleaven, C.M.; Pacheco, J.; Cooper, S.; Bao, C.; Abraham, M.; Cromwell, M.; Sardi, S.P.; Chuang, W.L.; Sidman, R.L.; et al. Glycosphingolipids Are Modulators of Disease Pathogenesis in Amyotrophic Lateral Sclerosis. Proc. Natl. Acad. Sci. USA 2015, 112, 8100–8105. [Google Scholar] [CrossRef]
  317. Henriques, A.; Croixmarie, V.; Priestman, D.A.; Rosenbohm, A.; Dirrig-Grosch, S.; D’Ambra, E.; Huebecker, M.; Hussain, G.; Boursier-Neyret, C.; Echaniz-Laguna, A.; et al. Amyotrophic Lateral Sclerosis and Denervation Alter Sphingolipids and up-Regulate Glucosylceramide Synthase. Hum. Mol. Genet. 2015, 24, 7390–7405. [Google Scholar] [CrossRef] [PubMed]
  318. Henriques, A.; Huebecker, M.; Blasco, H.; Keime, C.; Andres, C.R.; Corcia, P.; Priestman, D.A.; Platt, F.M.; Spedding, M.; Loeffler, J.P. Inhibition of Β-Glucocerebrosidase Activity Preserves Motor Unit Integrity in a Mouse Model of Amyotrophic Lateral Sclerosis. Sci. Rep. 2017, 7, e5235. [Google Scholar] [CrossRef]
  319. Rapport, M.M. Implications of Altered Brain Ganglioside Profiles in Amyotrophic Lateral Sclerosis (Als). Acta Neurobiol. Exp. 1990, 50, 505–513. [Google Scholar]
  320. Palomo, V.; Nozal, V.; Rojas-Prats, E.; Gil, C.; Martinez, A. Protein Kinase Inhibitors for Amyotrophic Lateral Sclerosis Therapy. Br. J. Pharmacol. 2021, 178, 1316–1335. [Google Scholar] [CrossRef]
  321. Haggiag, S.; Steiner-Birmanns, B.; Wirguin, I.; Sicsic, C.; Brenner, T.; Steiner, I. Seroconversion of Anti-Gm1 Antibodies in Patients with Amyotrophic Lateral Sclerosis. Neurology 2004, 63, 755–756. [Google Scholar] [CrossRef]
  322. Adams, D.; Kuntzer, T.; Burger, D.; Chofflon, M.; Magistris, M.R.; Regli, F.; Steck, A.J. Predictive Value of Anti-Gm1 Ganglioside Antibodies in Neuromuscular Diseases: A Study of 180 Sera. J. Neuroimmunol. 1991, 32, 223–230. [Google Scholar] [CrossRef]
  323. Stevens, A.; Weller, M.; Wiethölter, H. A Characteristic Ganglioside Antibody Pattern in the Csf of Patients with Amyotrophic Lateral Sclerosis. J. Neurol. Neurosurg. Psychiatry 1993, 56, 361–364. [Google Scholar] [CrossRef] [PubMed]
  324. Kollewe, K.; Wurster, U.; Sinzenich, T.; Körner, S.; Dengler, R.; Mohammadi, B.; Petri, S. Anti-Ganglioside Antibodies in Amyotrophic Lateral Sclerosis Revisited. PLoS ONE 2015, 10, e0125339. [Google Scholar] [CrossRef] [PubMed]
  325. Hankey, G.J. Stroke. Lancet 2017, 389, 641–654. [Google Scholar] [CrossRef] [PubMed]
  326. Qi, Y.; Xue, Q.M. Ganglioside Levels in Hypoxic Brains from Neonatal and Premature Infants. Mol. Chem. Neuropathol. 1991, 14, 87–97. [Google Scholar] [CrossRef] [PubMed]
  327. Ramirez, M.R.; Muraro, F.; Zylbersztejn, D.S.; Abel, C.R.; Arteni, N.S.; Lavinsky, D.; Netto, C.A.; Trindade, V.M. Neonatal Hypoxia-Ischemia Reduces Ganglioside, Phospholipid and Cholesterol Contents in the Rat Hippocampus. Neurosci. Res. 2003, 46, 339–347. [Google Scholar] [CrossRef]
  328. Whitehead, S.N.; Chan, K.H.; Gangaraju, S.; Slinn, J.; Li, J.; Hou, S.T. Imaging Mass Spectrometry Detection of Gangliosides Species in the Mouse Brain Following Transient Focal Cerebral Ischemia and Long-Term Recovery. PLoS ONE 2011, 6, e20808. [Google Scholar] [CrossRef]
  329. Caughlin, S.; Hepburn, J.D.; Park, D.H.; Jurcic, K.; Yeung, K.K.; Cechetto, D.F.; Whitehead, S.N. Increased Expression of Simple Ganglioside Species Gm2 and Gm3 Detected by Maldi Imaging Mass Spectrometry in a Combined Rat Model of Aβ Toxicity and Stroke. PLoS ONE 2015, 10, e0130364. [Google Scholar] [CrossRef]
  330. Kwak, D.H.; Kim, S.M.; Lee, D.H.; Kim, J.S.; Kim, S.M.; Lee, S.U.; Jung, K.Y.; Seo, B.B.; Choo, Y.K. Differential Expression Patterns of Gangliosides in the Ischemic Cerebral Cortex Produced by Middle Cerebral Artery Occlusion. Mol. Cells 2005, 20, 354–360. [Google Scholar]
  331. Liu, J.R.; Ding, M.P.; Wei, E.Q.; Luo, J.H.; Song, Y.; Huang, J.Z.; Ge, Q.F.; Hu, H.; Zhu, L.J. Gm1 Stabilizes Expression of Nmda Receptor Subunit 1 in the Ischemic Hemisphere of Mcao/Reperfusion Rat. J. Zhejiang Univ. Sci. B 2005, 6, 254–258. [Google Scholar] [CrossRef]
  332. Zhang, J.; Fang, X.; Zhou, Y.; Deng, X.; Lu, Y.; Li, J.; Li, S.; Wang, B.; Xu, R. The Possible Damaged Mechanism and the Preventive Effect of Monosialotetrahexosylganglioside in a Rat Model of Cerebral Ischemia-Reperfusion Injury. J. Stroke Cerebrovasc. Dis. 2015, 24, 1471–1478. [Google Scholar] [CrossRef]
  333. Tan, W.K.; Williams, C.E.; Gunn, A.J.; Mallard, E.C.; Gluckman, P.D. Pretreatment with Monosialoganglioside Gm1 Protects the Brain of Fetal Sheep against Hypoxic-Ischemic Injury without Causing Systemic Compromise. Pediatr. Res. 1993, 34, 18–22. [Google Scholar] [CrossRef] [PubMed]
  334. Krajnc, D.; Wemlinger, T.A.; Neff, N.H.; Hadjiconstantinou, M. Neonatal Hypoxia: Early Neurotransmitter Responses and the Consequences of Treatment with Gm1 Ganglioside. J. Pharmacol. Exp. Ther. 1994, 271, 1299–1305. [Google Scholar]
  335. Zhang, J.Z.; Jing, L.; Ma, Y.; Guo, F.Y.; Chang, Y.; Li, P.A. Monosialotetrahexosy-1 Ganglioside Attenuates Diabetes-Enhanced Brain Damage after Transient Forebrain Ischemia and Suppresses Phosphorylation of Erk1/2 in the Rat Brain. Brain Res. 2010, 1344, 200–208. [Google Scholar] [CrossRef] [PubMed]
  336. Su, D.; Ma, J.; Yang, J.; Kang, Y.; Lv, M.; Li, Y. Monosialotetrahexosy-1 Ganglioside Attenuates Diabetes-Associated Cerebral Ischemia/Reperfusion Injury through Suppression of the Endoplasmic Reticulum Stress-Induced Apoptosis. J. Clin. Neurosci. 2017, 41, 54–59. [Google Scholar] [CrossRef] [PubMed]
  337. Mahadik, S.P.; Hungund, B.L.; Gokhale, V.S.; Ortiz, A.; Makar, T.K.; Karpiak, S.E. Monosialoganglioside (Gm1) Restores Membrane Fatty Acid Levels in Ischemic Tissue after Cortical Focal Ischemia in Rat. Neurochem. Int. 1993, 23, 163–172. [Google Scholar] [CrossRef]
  338. Zhang, W.; Krafft, P.R.; Wang, T.; Zhang, J.H.; Li, L.; Tang, J. Pathophysiology of Ganglioside Gm1 in Ischemic Stroke: Ganglioside Gm1: A Critical Review. Cell Transplant. 2019, 28, 657–661. [Google Scholar] [CrossRef]
  339. Bharucha, V.A.; Wakade, C.G.; Mahadik, S.P.; Karpiak, S.E. Gm1 Ganglioside Treatment Reduces Functional Deficits Associated with Cortical Focal Ischemia. Exp. Neurol. 1991, 114, 136–139. [Google Scholar] [CrossRef]
  340. Ortiz, A.; MacDonall, J.S.; Wakade, C.G.; Karpiak, S.E. Gm1 Ganglioside Reduces Cognitive Dysfunction after Focal Cortical Ischemia. Pharmacol. Biochem. Behav. 1990, 37, 679–684. [Google Scholar] [CrossRef]
  341. Hicks, D.; Heidinger, V.; Mohand-Said, S.; Sahel, J.; Dreyfus, H. Growth Factors and Gangliosides as Neuroprotective Agents in Excitotoxicity and Ischemia. Gen. Pharmacol. 1998, 30, 265–273. [Google Scholar] [CrossRef]
  342. Rotondo, G.; Maniero, G.; Toffano, G. New Perspectives in the Treatment of Hypoxic and Ischemic Brain Damage: Effect of Gangliosides. Aviat. Space Environ. Med. 1990, 61, 162–164. [Google Scholar] [PubMed]
  343. Li, L.; Tian, J.; Long, M.K.; Chen, Y.; Lu, J.; Zhou, C.; Wang, T. Protection against Experimental Stroke by Ganglioside Gm1 Is Associated with the Inhibition of Autophagy. PLoS ONE 2016, 11, e0144219. [Google Scholar] [CrossRef] [PubMed]
  344. Seren, M.S.; Lazzaro, A.; Yang, C.L.; Canella, R.; Bassan, M.; Zanoni, R.; Manev, H. Orally Administered Glycolipid Derivative Liga20 Reduces Infarct Volume and Behavioral Impairment after Focal Cerebral Ischemia. J. Pharmacol. Exp. Ther. 1994, 268, 460–465. [Google Scholar]
  345. Karpiak, S.E.; Li, Y.S.; Mahadik, S.P. Gangliosides Reduce Mortality due to Global Ischemia: Membrane Protection. Clin. Neuropharmacol. 1986, 9 (Suppl. S4), 338–340. [Google Scholar] [PubMed]
  346. Jesulola, E.; Micalos, P.; Baguley, I.J. Understanding the Pathophysiology of Depression: From Monoamines to the Neurogenesis Hypothesis Model—Are We There Yet? Behav. Brain Res. 2018, 341, 79–90. [Google Scholar] [CrossRef]
  347. Price, H.C.; Byard, C.; Sims, W.; Wilson, R. Gangliosides and Other Lipid Micelles. A Study of Amine Binding by a Dialysis/Fluorescence Method. Neurochem. Res. 1979, 4, 63–72. [Google Scholar] [CrossRef] [PubMed]
  348. Berry-Kravis, E.; Dawson, G. Possible Role of Gangliosides in Regulating an Adenylate Cyclase-Linked 5-Hydroxytryptamine (5-Ht1) Receptor. J. Neurochem. 1985, 45, 1739–1747. [Google Scholar] [CrossRef]
  349. Chattopadhyay, A.; Paila, Y.D.; Shrivastava, S.; Tiwari, S.; Singh, P.; Fantini, J. Sphingolipid-Binding Domain in the Serotonin(1a) Receptor. Adv. Exp. Med. Biol. 2012, 749, 279–293. [Google Scholar]
  350. Gilmore, N.; Cherian, L.; Klemm, W.R. Ganglioside or Sialic Acid Attenuates Ethanol-Induced Decrements in Locomotion, Nose-Poke Exploration, and Anxiety, but Not Body Temperature. Prog. Neuro-Psychopharmacol. Biol. Psychiatry 1991, 15, 91–104. [Google Scholar] [CrossRef]
  351. Wallis, C.J.; Rezazadeh, S.M.; Lal, H. Gm1 Ganglioside Reduces Ethanol Intoxication and the Development of Ethanol Dependence. Alcohol 1995, 12, 573–580. [Google Scholar] [CrossRef]
  352. Jiang, B.; Song, L.; Wang, C.N.; Zhang, W.; Huang, C.; Tong, L.J. Antidepressant-Like Effects of Gm1 Ganglioside Involving the Bdnf Signaling Cascade in Mice. Int. J. Neuropsychopharmacol. 2016, 19, pyw046. [Google Scholar] [CrossRef]
  353. Juhola, H.; Postila, P.A.; Rissanen, S.; Lolicato, F.; Vattulainen, I.; Róg, T. Negatively Charged Gangliosides Promote Membrane Association of Amphipathic Neurotransmitters. Neuroscience 2018, 384, 214–223. [Google Scholar] [CrossRef] [PubMed]
  354. Lekman, A.; Skjeldal, O.; Sponheim, E.; Svennerholm, L. Gangliosides in Children with Autism. Acta Paediatr. 1995, 84, 787–790. [Google Scholar] [CrossRef] [PubMed]
  355. Nordin, V.; Lekman, A.; Johansson, M.; Fredman, P.; Gillberg, C. Gangliosides in Cerebrospinal Fluid in Children with Autism Spectrum Disorders. Dev. Med. Child Neurol. 1998, 40, 587–594. [Google Scholar] [CrossRef]
  356. Mostafa, G.A.; Al-Ayadhi, L.Y. Increased Serum Levels of Anti-Ganglioside M1 Auto-Antibodies in Autistic Children: Relation to the Disease Severity. J. Neuroinflamm. 2011, 8, e39. [Google Scholar] [CrossRef] [PubMed]
  357. Moeller, S.; Lau, N.M.; Green, P.H.; Hellberg, D.; Higgins, J.J.; Rajadhyaksha, A.M.; Alaedini, A. Lack of Association between Autism and Anti-Gm1 Ganglioside Antibody. Neurology 2013, 81, 1640–1641. [Google Scholar] [CrossRef] [PubMed]
  358. Laev, H.; Karpiak, S.E.; Gokhale, V.S.; Hungund, B.L. In Utero Ethanol Exposure Retards Growth and Alters Morphology of Cortical Cultures: Gm1 Reverses Effects. Alcohol. Clin. Exp. Res. 1995, 19, 1226–1233. [Google Scholar] [CrossRef] [PubMed]
  359. Harris, R.A.; Groh, G.I.; Baxter, D.M.; Hitzemann, R.J. Gangliosides Enhance the Membrane Actions of Ethanol and Pentobarbital. Mol. Pharmacol. 1984, 25, 410–417. [Google Scholar]
  360. Vrbaski, S.R.; Grujić-Injac, B.; Ristić, M. Phospholipid and Ganglioside Composition in Rat Brain after Chronic Intake of Ethanol. J. Neurochem. 1984, 42, 1235–1239. [Google Scholar] [CrossRef]
  361. Vrbaski, S.R. Region Distribution of the Gangliosides in Rat Brain after Chronic Ethanol Treatment. Mol. Chem. Neuropathol. 1995, 25, 273–281. [Google Scholar] [CrossRef]
  362. Saito, M.; Saito, M.; Cooper, T.B.; Vadasz, C. Alcohol Reduces Gm1 Ganglioside Content in the Serum of Inbred Mouse Strains. Alcohol. Clin. Exp. Res. 2004, 28, 1107–1113. [Google Scholar] [CrossRef]
  363. Snell, L.D.; Szabó, G.; Tabakoff, B.; Hoffman, P.L. Gangliosides Reduce the Development of Ethanol Dependence without Affecting Ethanol Tolerance. J. Pharmacol. Exp. Ther. 1996, 279, 128–136. [Google Scholar]
  364. Saito, M.; Saito, M.; Berg, M.J.; Guidotti, A.; Marks, N. Gangliosides Attenuate Ethanol-Induced Apoptosis in Rat Cerebellar Granule Neurons. Neurochem. Res. 1999, 24, 1107–1115. [Google Scholar] [CrossRef] [PubMed]
  365. Saito, M.; Mao, R.F.; Wang, R.; Vadasz, C.; Saito, M. Effects of Gangliosides on Ethanol-Induced Neurodegeneration in the Developing Mouse Brain. Alcohol. Clin. Exp. Res. 2007, 31, 665–674. [Google Scholar] [CrossRef]
  366. Hungund, B.L.; Ross, D.C.; Gokhale, V.S. Ganglioside Gm1 Reduces Fetal Alcohol Effects in Rat Pups Exposed to Ethanol in Utero. Alcohol. Clin. Exp. Res. 1994, 18, 1248–1251. [Google Scholar] [CrossRef]
  367. Hungund, B.L.; Gokhale, V.S. Reduction of Fatty Acid Ethyl Ester Accumulation by Ganglioside Gm1 in Rat Fetus Exposed to Ethanol. Biochem. Pharmacol. 1994, 48, 2103–2108. [Google Scholar] [CrossRef] [PubMed]
  368. Laev, H.; Hungund, B.L.; Karpiak, S.E. Cortical Cell Plasma Membrane Alterations after In Vitro Alcohol Exposure: Prevention by Gm1 Ganglioside. Alcohol 1996, 13, 187–194. [Google Scholar] [CrossRef]
  369. Hungund, B.L.; Zheng, Z.; Lin, L.; Barkai, A.I. Ganglioside Gm1 Reduces Ethanol Induced Phospholipase A2 Activity in Synaptosomal Preparations from Mice. Neurochem. Int. 1994, 25, 321–325. [Google Scholar] [CrossRef]
  370. Zheng, Z.; Barkai, A.I.; Hungund, B.L. Effect of Ganglioside Gm1 on Ethanol-Induced Changes in the Incorporation of Free Fatty Acids into Membrane Phospholipids in Mouse Brain. Alcohol Alcohol. 1997, 32, 693–701. [Google Scholar] [CrossRef] [PubMed]
  371. Fishman, P.H. Role of Membrane Gangliosides in the Binding and Action of Bacterial Toxins. J. Membr. Biol. 1982, 69, 85–97. [Google Scholar] [CrossRef]
  372. Wernick, N.L.; Chinnapen, D.J.; Cho, J.A.; Lencer, W.I. Cholera Toxin: An Intracellular Journey into the Cytosol by Way of the Endoplasmic Reticulum. Toxins 2010, 2, 310–325. [Google Scholar] [CrossRef]
  373. Lubran, M.M. Bacterial Toxins. Ann. Clin. Lab. Sci. 1988, 18, 58–71. [Google Scholar] [PubMed]
  374. Chinnapen, D.J.; Hsieh, W.T.; te Welscher, Y.M.; Saslowsky, D.E.; Kaoutzani, L.; Brandsma, E.; D’Auria, L.; Park, H.; Wagner, J.S.; Drake, K.R.; et al. Lipid Sorting by Ceramide Structure from Plasma Membrane to Er for the Cholera Toxin Receptor Ganglioside Gm1. Dev. Cell 2012, 23, 573–586. [Google Scholar] [CrossRef] [PubMed]
  375. Day, C.A.; Kenworthy, A.K. Functions of Cholera Toxin B-Subunit as a Raft Cross-Linker. Essays Biochem. 2015, 57, 135–145. [Google Scholar] [PubMed]
  376. Zuverink, M.; Barbieri, J.T. Protein Toxins That Utilize Gangliosides as Host Receptors. Prog. Mol. Biol. Transl. Sci. 2018, 156, 325–354. [Google Scholar]
  377. Fang, Y.; Xie, X.; Ledeen, R.W.; Wu, G. Characterization of Cholera Toxin B Subunit-Induced Ca(2+) Influx in Neuroblastoma Cells: Evidence for a Voltage-Independent Gm1 Ganglioside-Associated Ca(2+) Channel. J. Neurosci. Res. 2002, 69, 669–680. [Google Scholar] [CrossRef]
  378. Nedelkoska, L.; Benjamins, J.A. Binding of Cholera Toxin B Subunit: A Surface Marker for Murine Microglia but Not Oligodendrocytes or Astrocytes. J. Neurosci. Res. 1998, 53, 605–612. [Google Scholar] [CrossRef]
  379. Chen, C.; Fu, Z.; Kim, J.J.; Barbieri, J.T.; Baldwin, M.R. Gangliosides as High Affinity Receptors for Tetanus Neurotoxin. J. Biol. Chem. 2009, 284, 26569–26577. [Google Scholar] [CrossRef]
  380. Critchley, D.R.; Habig, W.H.; Fishman, P.H. Reevaluation of the Role of Gangliosides as Receptors for Tetanus Toxin. J. Neurochem. 1986, 47, 213–222. [Google Scholar] [CrossRef]
  381. Sinclair, H.R.; Smejkal, C.W.; Glister, C.; Kemp, F.; van den Heuvel, E.; de Slegte, J.; Gibson, G.R.; Rastall, R.A. Sialyloligosaccharides Inhibit Cholera Toxin Binding to the Gm1 Receptor. Carbohydr. Res. 2008, 343, 2589–2594. [Google Scholar] [CrossRef]
  382. Lingwood, C. Therapeutic Uses of Bacterial Subunit Toxins. Toxins 2021, 13, e378. [Google Scholar] [CrossRef]
  383. Schneider, J.S. Gangliosides and Glycolipids in Neurodegenerative Disorders. Adv. Neurobiol. 2014, 9, 449–461. [Google Scholar]
  384. Aureli, M.; Mauri, L.; Ciampa, M.G.; Prinetti, A.; Toffano, G.; Secchieri, C.; Sonnino, S. Gm1 Ganglioside: Past Studies and Future Potential. Mol. Neurobiol. 2016, 53, 1824–1842. [Google Scholar] [CrossRef]
  385. Lipartiti, M.; Lazzaro, A.; Zanoni, R.; Mazzari, S.; Toffano, G.; Leon, A. Monosialoganglioside Gm1 Reduces Nmda Neurotoxicity in Neonatal Rat Brain. Exp. Neurol. 1991, 113, 301–305. [Google Scholar] [CrossRef]
  386. Lipartiti, M.; Lazzaro, A.; Manev, H. Ganglioside Derivative Liga20 Reduces Nmda Neurotoxicity in Neonatal Rat Brain. Neuroreport 1992, 3, 919–921. [Google Scholar] [CrossRef] [PubMed]
  387. Lazarewicz, J.W.; Salińska, E.; Matyja, E. Ganglioside Gm1 Prevents N-Methyl-D-Aspartate Neurotoxicity in Rabbit Hippocampus in Vivo. Effects on Calcium Homeostasis. Mol. Chem. Neuropathol. 1995, 24, 165–177. [Google Scholar] [CrossRef]
  388. Fazzini, E.; Durso, R.; Davoudi, H.; Szabo, G.K.; Albert, M.L. Gm1 Gangliosides Alter Acute Mptp-Induced Behavioral and Neurochemical Toxicity in Mice. J. Neurol. Sci. 1990, 99, 59–68. [Google Scholar] [CrossRef]
  389. Bellot, R.G.; Vital, M.A.; Palermo-Neto, J.; Frussa-Filho, R. Repeated Monosialoganglioside Administration Attenuates Behavioral Sensitization to Amphetamine. Brain Res. 1997, 747, 169–172. [Google Scholar] [CrossRef] [PubMed]
  390. Fighera, M.R.; Bonini, J.S.; de Oliveira, T.G.; Frussa-Filho, R.; Rocha, J.B.; Dutra-Filho, C.S.; Rubin, M.A.; Mello, C.F. Gm1 Ganglioside Attenuates Convulsions and Thiobarbituric Acid Reactive Substances Production Induced by the Intrastriatal Injection of Methylmalonic Acid. Int. J. Biochem. Cell Biol. 2003, 35, 465–473. [Google Scholar] [CrossRef]
  391. Lombardi, G.; Zanoni, R.; Moroni, F. Systemic Treatments with Gm1 Ganglioside Reduce Quinolinic Acid-Induced Striatal Lesions in the Rat. Eur. J. Pharmacol. 1989, 174, 123–125. [Google Scholar] [CrossRef] [PubMed]
  392. Kojima, H.; Yamada, S.; Yokoo, H.; Tsutsumi, T.; Nishi, S.; Inanaga, K.; Nagatsu, I.; Jonsson, G.; Toffano, G. The Effects of Exogenous Gm1 Ganglioside on Neurotoxin Induced Damage of Cerebral Serotonin Nerve Terminals in Adult Rats. Kurume Med. J. 1988, 35, 49–61. [Google Scholar] [CrossRef]
  393. Skaper, S.D.; Facci, L.; Leon, A. Gangliosides Attenuate the Delayed Neurotoxicity of Aspartic Acid In Vitro. Neurosci. Lett. 1990, 117, 154–159. [Google Scholar] [CrossRef]
  394. Sabel, B.A.; DelMastro, R.; Dunbar, G.L.; Stein, D.G. Reduction of Anterograde Degeneration in Brain Damaged Rats by Gm1-Gangliosides. Neurosci. Lett. 1987, 77, 360–366. [Google Scholar] [CrossRef]
  395. Guelman, L.R.; Zorrilla Zubilete, M.A.; Ríos, H.; Dopico, A.M.; Zieher, L.M. Gm1 Ganglioside Treatment Protects against Long-Term Neurotoxic Effects of Neonatal X-Irradiation on Cerebellar Cortex Cytoarchitecture and Motor Function. Brain Res. 2000, 858, 303–311. [Google Scholar] [CrossRef]
  396. Guelman, L.R.; Zieher, L.M.; Fiszman, M.L. The Effect of X-Radiation on Cerebellar Granule Cells Grown in Culture. Ganglioside Gm1 Neuroprotective Activity. Neurochem. Int. 1996, 29, 521–527. [Google Scholar] [CrossRef] [PubMed]
  397. Goettl, V.M.; Lindsey, A.E.; Neff, N.H.; Hadjiconstantinou, M. Gm1 Ganglioside Restores Abnormal Responses to Acute Thermal and Mechanical Stimuli in Aged Rats. Brain Res. 2000, 858, 380–385. [Google Scholar] [CrossRef] [PubMed]
  398. Sautter, J.; Schwartz, M.; Duvdevani, R.; Sabel, B.A. Gm1 Ganglioside Treatment Reduces Visual Deficits after Graded Crush of the Rat Optic Nerve. Brain Res. 1991, 565, 23–33. [Google Scholar] [CrossRef]
  399. Glasier, M.M.; Janis, L.S.; Goncalves, M.I.; Stein, D.G. Gm1 Produces Attenuation of Short-Term Memory Deficits in Hebb-Williams Maze Performance after Unilateral Entorhinal Cortex Lesions. Physiol. Behav. 1999, 66, 441–446. [Google Scholar] [CrossRef] [PubMed]
  400. Tseng, E.E.; Brock, M.V.; Lange, M.S.; Troncoso, J.C.; Blue, M.E.; Lowenstein, C.J.; Johnston, M.V.; Baumgartner, W.A. Monosialoganglioside Gm1 Inhibits Neurotoxicity after Hypothermic Circulatory Arrest. Surgery 1998, 124, 298–306. [Google Scholar] [CrossRef] [PubMed]
  401. Baumgartner, W.A.; Redmond, J.M.; Zehr, K.J.; Brock, M.V.; Tseng, E.E.; Blue, M.E.; Troncoso, J.C.; Johnston, M.V. The Role of the Monosialoganglioside, Gm1 as a Neuroprotectant in an Experimental Model of Cardiopulmonary Bypass and Hypothermic Circulatory Arrest. Ann. N. Y. Acad. Sci. 1998, 845, 382–390. [Google Scholar] [CrossRef]
  402. Oderfeld-Nowak, B.; Brosnan, C.; Cervone, A.; Oderfeld, J.; Ledeen, R.W. Gangliosides Improve the Outcome of Experimental Allergic Neuritis (Ean). Acta Neurobiol. Exp. 1990, 50, 495–504. [Google Scholar]
  403. Laev, H.; Mahadik, S.P.; Bonheur, J.L.; Hernandez, N.; Karpiak, S.E. Gm1 Ganglioside Reduces Glutamate Toxicity to Cortical Cells. Lowered Ldh Release and Preserved Membrane Integrity. Mol. Chem. Neuropathol. 1993, 20, 229–243. [Google Scholar] [CrossRef] [PubMed]
  404. Fong, T.G.; Neff, N.H.; Hadjiconstantinou, M. Systemic Administration of Gm1 Ganglioside Increases Choline Acetyltransferase Activity in the Brain of Aged Rats. Exp. Neurol. 1995, 132, 157–161. [Google Scholar] [CrossRef] [PubMed]
  405. Hadjiconstantinou, M.; Karadsheh, N.S.; Rattan, A.K.; Tejwani, G.A.; Fitkin, J.G.; Neff, N.H. Gm1 Ganglioside Enhances Cholinergic Parameters in the Brain of Senescent Rats. Neuroscience 1992, 46, 681–686. [Google Scholar] [CrossRef]
  406. Vogelsberg, V.; Fong, T.G.; Neff, N.H.; Hadjiconstantinou, M. Cholinergic Deficits in Aged Rat Spinal Cord: Restoration by Gm1 Ganglioside. Brain Res. 1997, 761, 250–256. [Google Scholar] [CrossRef] [PubMed]
  407. Di Patre, P.L.; Abbamondi, A.; Bartolini, L.; Pepeu, G. Gm1 Ganglioside Counteracts Cholinergic and Behavioral Deficits Induced in the Rat by Intracerebral Injection of Vincristine. Eur. J. Pharmacol. 1989, 162, 43–50. [Google Scholar] [CrossRef] [PubMed]
  408. Hadjiconstantinou, M.; Neff, N.H. Treatment with Gm1 Ganglioside Restores Striatal Dopamine in the 1-Methyl-4-Phenyl-1,2,3,6-Tetrahydropyridine-Treated Mouse. J. Neurochem. 1988, 51, 1190–1196. [Google Scholar] [CrossRef]
  409. Hadjiconstantinou, M.; Mariani, A.P.; Neff, N.H. Gm1 Ganglioside-Induced Recovery of Nigrostriatal Dopaminergic Neurons after Mptp: An Immunohistochemical Study. Brain Res. 1989, 484, 297–303. [Google Scholar] [CrossRef]
  410. Hadjiconstantinou, M.; Weihmuller, F.; Neff, N.H. Treatment with Gm1 Ganglioside Reverses Dopamine D-2 Receptor Supersensitivity Induced by the Neurotoxin Mptp. Eur. J. Pharmacol. 1989, 168, 261–264. [Google Scholar] [CrossRef]
  411. Nicoletti, F.; Cavallaro, S.; Bruno, V.; Virgili, M.; Catania, M.V.; Contestabile, A.; Canonico, P.L. Gangliosides Attenuate Nmda Receptor-Mediated Excitatory Amino Acid Release in Cultured Cerebellar Neurons. Neuropharmacology 1989, 28, 1283–1286. [Google Scholar] [CrossRef]
  412. Schneider, J.S.; DiStefano, L. Liga 20 Increases Striatal Dopamine Levels in Aged Mptp-Treated Mice Refractory to Gm1 Ganglioside Treatment. Neuroreport 1993, 5, 103–104. [Google Scholar] [CrossRef]
  413. Toffano, G.; Savoini, G.; Aldinio, C.; Valenti, G.; Dal Toso, R.; Leon, A.; Calza, L.; Zini, I.; Agnati, L.F.; Fuxe, K. Effects of Gangliosides on the Functional Recovery of Damaged Brain. Adv. Exp. Med. Biol. 1984, 174, 475–488. [Google Scholar] [PubMed]
  414. Rubovitch, V.; Zilberstein, Y.; Chapman, J.; Schreiber, S.; Pick, C.G. Restoring Gm1 Ganglioside Expression Ameliorates Axonal Outgrowth Inhibition and Cognitive Impairments Induced by Blast Traumatic Brain Injury. Sci. Rep. 2017, 7, e41269. [Google Scholar] [CrossRef]
  415. Herrero, M.T.; Perez-Otaño, I.; Oset, C.; Kastner, A.; Hirsch, E.C.; Agid, Y.; Luquin, M.R.; Obeso, J.A.; Del Rio, J. Gm-1 Ganglioside Promotes the Recovery of Surviving Midbrain Dopaminergic Neurons in Mptp-Treated Monkeys. Neuroscience 1993, 56, 965–972. [Google Scholar] [CrossRef] [PubMed]
  416. Stull, N.D.; Schneider, J.S.; Iacovitti, L. Gm1 Ganglioside Partially Rescues Cultured Dopaminergic Neurons from Mpp(+)-Induced Damage: Dependence on Initial Damage and Time of Treatment. Brain Res. 1994, 640, 308–315. [Google Scholar] [CrossRef] [PubMed]
  417. Weihmuller, F.B.; Hadjiconstantinou, M.; Bruno, J.P.; Neff, N.H. Administration of Gm1 Ganglioside Eliminates Neuroleptic-Induced Sensorimotor Deficits in Mptp-Treated Mice. Neurosci. Lett. 1988, 92, 207–212. [Google Scholar] [CrossRef]
Figure 1. The structure of GM1 and its association with the cell membrane through insertion of the lipid tails (blue) into the membrane.
Figure 1. The structure of GM1 and its association with the cell membrane through insertion of the lipid tails (blue) into the membrane.
Ijms 24 09558 g001
Figure 2. Biosynthesis of ganglio-series GSLs. The numbers in superscript indicate the glycosylation sites of GTs. GM1, GD1a, GD1b, and GT1b are the major gangliosides in the human brain.
Figure 2. Biosynthesis of ganglio-series GSLs. The numbers in superscript indicate the glycosylation sites of GTs. GM1, GD1a, GD1b, and GT1b are the major gangliosides in the human brain.
Ijms 24 09558 g002
Figure 3. The degradation of some important gangliosides. Complex gangliosides are degraded to form GM1 on the cell surface and then translocated into the cell through endocytosis to be further hydrolyzed to free monosaccharides and Cer.
Figure 3. The degradation of some important gangliosides. Complex gangliosides are degraded to form GM1 on the cell surface and then translocated into the cell through endocytosis to be further hydrolyzed to free monosaccharides and Cer.
Ijms 24 09558 g003
Figure 4. Chemical structures of the several dominant gangliosides in the vertebrate brain, (GM3, GD3, GM1, GD1a, GD1b, GT1b, GM1b) which have different numbers of sialic acid residues (in red) or have sialic acid at different locations.
Figure 4. Chemical structures of the several dominant gangliosides in the vertebrate brain, (GM3, GD3, GM1, GD1a, GD1b, GT1b, GM1b) which have different numbers of sialic acid residues (in red) or have sialic acid at different locations.
Ijms 24 09558 g004
Figure 5. Major ganglioside species and their proportions in the adult human brain. This figure is generated using the data presented in the literature [29,35].
Figure 5. Major ganglioside species and their proportions in the adult human brain. This figure is generated using the data presented in the literature [29,35].
Ijms 24 09558 g005
Table 1. Major functions of GM1 in the central nervous system.
Table 1. Major functions of GM1 in the central nervous system.
CNS Functions That Involve GM1References
  • GM1 is a differential marker of neurons
  • GM1 is involved in neuronal differentiation and development and nerve regeneration
  • GM1 promotes neuritogenesis
  • GM1 accelerates neurite outgrowth
  • GM1 regulates signal transduction
  • GM1 inhibits neuronal apoptosis
  • GM1 prevents neuronal functional decay
  • GM1 is protective against neurodegeneration, neurotoxicity, and neuronal injury and death caused by various factors
  • GM1 is involved in memory and cognition via modulating the synaptic activities
[87]
[83,84,92]
[88,89]
[90]
[102,103]
[115]
[117]
[116,118,119,120,121,122,123,124,125]

[32,149,150]
Table 2. GM1 in CNS-related neurological disorders.
Table 2. GM1 in CNS-related neurological disorders.
DiseasesCorrelation of GM1 with the DiseaseReferences
Alzheimer’s disease
  • Age-dependent clustering of GM1 at the presynaptic neuritic terminal
  • Increased GM1 (d18:1) to (d20:1) ratio
  • At low density, GM1 disrupts Aβ dimers and inhibits sphingomyelin-triggered Aβ oligomerization
  • Clustered GM1 in the lipid rafts binds to Aβ selectively to promote Aβ aggregation and fibrillogenesis
[208]
[211]
[203,206]


[210,219,220,221,222]
Parkinson’s disease
  • GM1 deficit is observed in PD patients and animal models
  • GM1 can stimulate autophagy and α-synuclein clearance and modulate α-synuclein folding and aggregation
[234,235,236]

[239,240]
GM1 gangliosidosis
  • GM1 accumulation is observed in the central nervous system, including the cerebrospinal fluid
[250,251,253,254]
Huntington’s disease
  • Decreased GM1 in HD patients and animal models
  • GM1 accelerates mutant HTT clearance and reduces mutant HTT aggregation and toxicity
[301,302,303]

[305,306,307]
Epilepsy, seizure
  • Decreased GM1
  • Anti-GM1 antibody can cause seizure
[310]
[312]
Amyotrophic lateral sclerosis
  • Elevated GM1 in the spinal cord
[316]
Ischemic stroke
  • Decreased GM1
  • Elevated level of GM1 (d20:1) lipid form
[326]
[328,329]
Depression, anxiety
  • GM1 is a co-receptor of serotonin, dopamine and other membrane-binding neurotransmitters
  • GM1 interacts with serotonin receptor to enhance ligand binding
[107,161,347]

[349]
Autism
  • Elevated GM1
  • Anti-GM1 antibodies are observed
[354,355,356]
Alcohol dependence
  • Decreased GM1 in animal models
  • GM1 inhibits alcohol-induced changes in the cell membrane
[362]
[366,368]
Bacterial infections
  • GM1 is the binding receptor for some bacterial toxins
  • The lipid form of GM1 is important for toxin binding and trafficking
[371,372,373]
[374]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, Z. Ganglioside GM1 and the Central Nervous System. Int. J. Mol. Sci. 2023, 24, 9558. https://doi.org/10.3390/ijms24119558

AMA Style

Guo Z. Ganglioside GM1 and the Central Nervous System. International Journal of Molecular Sciences. 2023; 24(11):9558. https://doi.org/10.3390/ijms24119558

Chicago/Turabian Style

Guo, Zhongwu. 2023. "Ganglioside GM1 and the Central Nervous System" International Journal of Molecular Sciences 24, no. 11: 9558. https://doi.org/10.3390/ijms24119558

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop