Skip to main content
Canna~Fangled Abstracts

Therapeutic neuroprotective agents for amyotrophic lateral sclerosis

By July 18, 2013No Comments
pub med big
Logo of nihpa

Cell Mol Life Sci. Author manuscript; available in PMC Sep 23, 2014.
Published in final edited form as:
PMCID: PMC4172456
NIHMSID: NIHMS628427

Therapeutic neuroprotective agents for amyotrophic lateral sclerosis

Abstract

Amyotrophic lateral sclerosis (ALS) is a fatal chronic neurodegenerative disease whose hallmark is proteinaceous, ubiquitinated, cytoplasmic inclusions in motor neurons and surrounding cells. Multiple mechanisms proposed as responsible for ALS pathogenesis include dysfunction of protein degradation, glutamate excitotoxicity, mitochondrial dysfunction, apoptosis, oxidative stress, and inflammation. It is therefore essential to gain a better understanding of the underlying disease etiology and search for neuroprotective agents that might delay disease onset, slow progression, prolong survival, and ultimately reduce the burden of disease. Because riluzole, the only Food and Drug Administration (FDA)-approved treatment, prolongs the ALS patient’s life by only 3 months, new therapeutic agents are urgently needed. In this review, we focus on studies of various small pharmacological compounds targeting the proposed pathogenic mechanisms of ALS and discuss their impact on disease progression.

Keywords: Neuroprotective agents, Motor neurons, Glial cells, Muscle, Amyotrophic lateral sclerosis pathogenesis

Amyotrophic lateral sclerosis

Amyotrophic lateral sclerosis (ALS), a.k.a. Lou Gehrig’s disease, is a chronic neurodegenerative disease with a prevalence of 1–2 per 100,000 [1] and an incidence of 6–8 per 100,000 people/year. More than 300,000 Americans will die from ALS over the next 10 years unless an effective cure or means of prevention is found. Approximately 90–95 % of ALS cases are sporadic ALS (sALS), with familial ALS (fALS) or inherited cases comprising the remaining 5–10 % of cases. The degeneration of motor neurons in primary motor cortex, corticospinal tracts, brain stem, and spinal cord is responsible for the muscle weakness that typifies ALS.

Skeletal muscle is considered a key target in the development of ALS. Mice expressing mutant G93A SOD1 (mSOD1G93A) selectively in skeletal muscle demonstrate progressive muscle atrophy and spinal motor neuron degeneration [23]. Because muscle cells can clear misfolded mSOD1 proteins more efficiently than motor neuron-like cells [4], several treatments have targeted skeletal muscle. Unfortunately, though these treatments help sustain muscle function, they do not significantly extend survival in ALS mice, which may imply that motor neuron-targeted treatment alone or combined with skeletal muscle-targeted treatment would be more effective in ALS treatment [58].

In ALS pathogenesis, it has been assumed that damage to a selective population of motor neurons leads to disease onset, duration, disease progression, and length of survival. Emerging evidence supports the involvement of neighboring non-motor neuron cell types including microglia, astrocytes, and other glial-type Schwann cells, oligodendrocytes, and NG2 cells, as well as targeted muscle cells in the process of ALS-related pathological development; astrogliosis and microgliosis are notable hallmarks of ALS disease. We will focus on therapeutic agents targeting motor neurons and summarize the effects of agents targeting astrocytes and microglia.

The neuroprotective agent riluzole, the only treatment currently approved by the US Food and Drug Administration (FDA), demonstrates only marginal efficacy, prolonging the patient’s life by only a few months [9]; its proposed mechanism of action is thought to be anti-excitotoxicity. Hence there is hope that neuroprotective agents counteracting excitotoxicity or other pathogenic mechanisms might ameliorate the clinical symptoms of ALS. Below we review the current knowledge regarding mechanisms underlying the disease pathology with the goal of identifying neuroprotective agents for ALS. We highlight the single small-chemical compounds that affect one or more proposed pathogenic ALS pathways and show neuroprotective efficacy in ALS mouse models. Although one agent may target multiple action mechanisms, we have divided neuroprotective agents into categories based on whether they mainly affect protein-degradation-clearing pathways, excitotoxicity, mitochondrial dysfunction, apoptosis, oxidative stress, or inflammation (Table 1Fig. 1).

Fig. 1

Neuroprotective agents target the pathogenic pathways in ALS
Table 1

Therapeutic neuroprotective agents and their effects on models of ALS

Genetic factors in the ALS and mutant SOD1G93A mouse model

Amyotrophic lateral sclerosis is a heterogeneous disease. Though mutations in copper–zinc superoxide dismutase 1 (SOD1) have been linked to ~20 % of fALS cases [10], other genes have been implicated, including ubiquilin 2 [11] (a recent finding), TDP-43 [12], FUS/TLS [1314], TAF-15 [15], and the hexanucleotide-repeat expansion in the uncharacterized gene C9ORF72 [1617].

Transgenic mice overexpressing mutant G93A SOD1 constituted the first and most commonly used animal model to evaluate the efficacy of potential treatments including small chemical compounds for ALS [1819] although other animal models are available. The mSOD1 mice models (including G37R, G85R, G127X, D90A, and H46R) share similar pathological hallmarks but differ in disease onset and progression related to gender, generation, and copy number [20]. Like humans with ALS, mSOD1 mice exhibit symptoms such as muscle loss [3], respiratory distress, upper and lower motor neuron involvement, immune system activation, blood–brain barrier (BBB) disruption, inclusion bodies and protein aggregation, and neuromuscular junction disruption [18192122]. Recent mouse studies began intervention at symptom onset [12325] or after onset [2627] to mimic the clinical treatment of ALS patients. However, most early studies treated mice before onset (preventively), which may have made that animal work less applicable to clinical studies, in that human ALS patients almost invariably receive treatment after diagnosis. Here we include mSOD1 mouse studies involving treatment before, at, or after disease onset.

Protein degradation pathways and SOD1-clearing agents

A significant amount of ALS research has been focused on understanding protein aggregation pathology. Protein degradation pathways play a crucial role in removing misfolded proteins and preventing protein aggregation; the ubiquitin proteasome system (UPS) and autophagy are the two major known protein degradation pathways [28]. Cytoplasmic inclusions in ALS are usually ubiquitin-positive [2934], and alterations in ubiquitin and proteasome are found not only in motor neurons but also in astrocytes from ALS patients. That evidence supports the involvement of the ubiquitin–proteasome system in both motor neurons and astrocytes in the pathogenesis of ALS [3233]. However, because accumulation of ubiquitinated proteins in patient samples could reflect the failure of the UPS, autophagy, or both, the role of the UPS in ALS is ambiguous. Most studies have reported mild-to-severe proteasomal dysfunction in ALS [35], although upregulation has also been reported [36]. A recent study using an in vivo UPS activity reporter mouse model demonstrated a mild decrease in UPS activity in ALS mice [37]. Mutations in ubiquilin 2, a protein that regulates the degradation of ubiquitinated proteins, have recently been reported to cause dominantly inherited, chromosome-X-linked ALS and ALS/dementia [11]. Although the details of ubiquilin 2’s role in ALS disease etiology remain to be determined, this new finding supports a significant relationship among proteasomal dysfunction, abnormal protein aggregation, and neurodegeneration.

It is critical to maintain autophagy activity in the central nervous system (CNS). Deletion of the key autophagy genes atg5 or atg7 in neurons in transgenic mice resulted in neurodegeneration and the accumulation of polyubiquitinated proteins and ubiquitin-immunoreactive inclusions [38]. The involvement of autophagy in ALS has been reported, but its functional role in disease etiology is less clear. Analysis of post-mortem spinal cord samples from sALS and fALS patients revealed elevated levels of poly-ubiquitin and autophagy markers Beclin-1 and LC3-II, suggesting autophagy [39]. The protein inclusions in ALS are often immunopositive for ubiquitin and p62, both of which are found in protein inclusions in atg5 and atg7 knockout mice. Moreover, alterations in autophagy caused by different gene mutations were reported in multiple fALS models.

p62 is an adaptor protein for autophagy substrates found in pathological inclusions in ALS. Ubiquitin is commonly found in p62-positive inclusions [40]. P62-immunopositive inclusions were reported in fALS caused by CHMP2B mutation [41], ANG/angiogenin mutation [42], FIG 4 mutation [43], and more recently in TDP-43 [44] and FUS inclusions [45]. p62 co-localized with SOD1- and ubiquitin-positive inclusions in G93A SOD1 mice and was co-immunoprecipitated with fALS mutants but not with wild-type SOD1 [30]. Moreover, p62 can function as an adaptor between mSOD1 and the autophagy machinery, potentially by a ubiquitin-independent mechanism [46]. P62 provides a potential target for novel therapeutic strategies that focus on clearing misfolded and aggregated proteins.

Several chemical compounds are capable of reducing SOD1 levels/aggregates in models of ALS. Arimoclomol, an amplifier of heat shock protein expression, delays disease progression and extends the lifespan of pre- or early symptomatic stages of mSOD1G93A mice [47]. Late-stage treatment improves muscle function [26]. Arimoclomol reduces ubiquitin aggregates in the spinal cord of G93A mice. Because the accumulation of ubiquitinated proteins reflects UPS failure and/or autophagy, these results may suggest a direct role of arimoclomol in protein aggregation [26] and effects on UPS and/or autophagy. Ongoing phase II/III clinical trials indicated arimoclomol’s good safety and tolerability [4849].

Another SOD1-clearing agent is edaravone. As an antioxidant and a free-radical scavenger, it effectively slows symptom progression, body weight loss, and motor neuron degeneration, and decreases the mean area with SOD1 aggregates in mSOD1G93A mice; these effects may be attributed to an enhancement of proteasomal activity yet to be characterized [25]. Pyrimethamine (Daraprim), an FDA-approved medication for the treatment of malaria and toxoplasmosis, reduces SOD1 levels in cultured cells, mice, and ALS patients [50], although it cannot decrease SOD1 expression in PC12 cells [51]. Lithium also reduces ubiquitin and SOD1 aggregates in motor neurons [52], inhibits excitotoxic motor neuron death in organotypic spinal cord cultures [53], and provides neuroprotection in cerebellar granule cells [54]. Moreover, lithium significantly delayed disease onset and duration, augmented the lifespan of the G93A mouse, and reduced reactive astrogliosis [52]. The same report suggests that lithium combined with riluzole delayed disease progression in ALS patients [52]. However, another report on G93A mice failed to show neuroprotection [55]. An Italian study suggested that lithium combined with riluzole delayed disease progression in ALS patients [52]. However, two larger multicenter trials failed to confirm these findings and were halted due to serious safety and efficacy concerns [56].

Excitotoxic mechanisms and anti-excitotoxic agents

Compared with other neurons, motor neurons are particularly vulnerable to excitotoxicity. In addition, astrocytes facilitate the removal of excessive glutamate and affect the calcium permeability of AMPA receptors of motor neurons [57]. Glutamate-induced excitotoxicity leading to motor neuron death is one pathogenic mechanism of ALS. NMDA and AMPA receptors are responsible for calcium entry. Under the pathological conditions of ALS, mSOD1 increases the sensitivity of the AMPA receptor to glutamatergic stimulation, disrupts mitochondrial function, and affects the surrounding astrocytes by interrupting their maintenance of extracellular glutamate levels. Astrocytes expressing mSOD1 are highly vulnerable to glutamate-induced excitotoxicity mediated by metabotropic glutamate receptor 5 (mGluR5) [3258].

As stated earlier, riluzole is the only FDA-approved treatment for ALS, but has only a modest effect on survival. Riluzole protects against motor neuron degeneration through excitotoxicity by interrupting glutamatergic transmission and lowering glutamate concentration, with effects on NMDA or AMPA receptors. This explains the strong interest in developing drug candidates that reduce glutamate-induced excitotoxicity. Riluzole significantly extended survival in ALS transgenic mice although it had no effect on disease onset. Riluzole also blocks Ca2+ and Na+ channels and modulates the GABAergic system, which suggests that it targets multiple pathogenic pathways [5960]. Combining riluzole with other drugs showing protective efficacy in ALS mice or patients may yield greater therapeutic potential for ALS. Ceftriaxone, a β-lactam antibiotic approved for treatment of bacterial infections, protected neurons against apoptosis and was identified from among 1,040 FDA-approved drugs as capable of increasing glutamate transporter gene expression [61]. Ceftriaxone prevents glutamate neurotoxicity, and when initiated at disease onset delays loss of muscle strength and body weight and prolongs survival [61]. Talampanel, a noncompetitive AMPA antagonist, reduces motorneuronal calcium levels if applied presymptomatically [62]. Memantine is an FDA-approved drug used for the treatment of Parkinsonism, vascular dementia, and Alzheimer’s disease. It is a N-methyl-D-aspartate (NMDA) receptor blocker that attenuates excitotoxicity [63]. If administered orally to mSOD1G93A mice either at symptom onset or pre-symptomatically [64], it prolongs survival [23]. However, memantine did not elicit any reversal of motor deterioration [23].

The histone deacetylases (HDACs) play an important role in protein acetylation in histones and in regulation of transcription [65]. Along with anti-excitotoxic effects, HDAC inhibitors have anti-inflammatory and neurotrophic properties. The HDAC inhibitor valproic acid (VPA) protects against glutamate or kainite-induced excitotoxicity in cultured neurons [6667] and achieves neuroprotection by reducing the number of apoptotic cells [68] and upregulating Bcl-2 [69]. Combined valproate and lithium treatment delays disease onset, reduces neurological deficits, and prolongs survival in ALS mice [70]. Sodium phenylbutyrate, another HDAC inhibitor, protects against glutathione-induced oxidative stress in cultured cortical neurons [71]. Sodium phenylbutyrate, administered by injection to mSOD1G93A mice either before or after symptom onset, prolongs survival and reduces the severity of pathological phenotypes [72].

Mitochondrial dysfunction and mitochondrial protectants

Mitochondria, cellular organelles that generate energy, play a key role in the intrinsic apoptotic death pathway, cellular calcium homeostasis, and ALS pathogenesis. A tendency to accumulate and aggregate in mitochondria is common to all mSOD1 [73], which may disrupt the activity of complex IV and impair the association of cytochrome c with the inner membrane [74]. mSOD1 aggregation in the mitochondria is linked with spinal cord–specific dysfunction of mitochondria, suggesting an association between defective mitochondria due to the toxic function of mSOD1 and the pathogenesis of ALS, prompting the search for neuroprotective agents targeting mitochondria [75]. Below is a brief review of candidate compounds.

KNS-760704 (dexpramipexole) is the optical enantiomer of pramipexole. Dexpramipexole was developed to maintain the neuroprotective properties while reducing the dopaminergic side effects of pramipexole. Both drugs include the benzothiazole core also present in riluzole [76]. Dexpramipexole functions at the level of mitochondria to enhance ATP output, reduce the generation of reactive oxygen species, and suppress apoptosis. It shows neuroprotection in vitro and in vivo, including the G93A mouse model [76], and hence is a candidate agent against ALS. However, dexpramipexole recently failed in a phase III trail.

Olesoxime (TRO-19622 or mitotarget), a small molecule with a cholesterol-like structure, is neuroprotective in animal and cellular models of ALS. It acts on the mitochondrial permeability transition pore (mPTP) [77]. Mitochondrial swelling and vacuolization are early pathological features of ALS [18], and genetic deletion of a major regulator of the mPTP in ALS mice delays disease onset and extends survival [78]. Olesoxime protects against motor neuron death, increasing the lifespan of mSOD1G93A mice via delaying the onset of motor dysfunction and weight loss rather than slowing disease progression [79]. Furthermore, olesoxime delays muscle denervation, astrogliosis, microglial activation, and motoneuron death in mSOD1G93A mice [80]. Olesoxime directly binds TSPO and VDAC, two proteins of the outer mitochondrial membrane [79].

Creatine helps prevent bioenergetic dysfunction and mitochondrial impairment, producing a dose-dependent improvement in motor performance and extending survival in G93A mice [81]. Creatine supplementation also reduces oxidative damage and neuron loss in G93A mice [81]. The addition of other agents, including minocycline or CELE-BREX, produces additive effects in ALS mice [8283].

Induction of mPTP results in the release of mitochondrial proteins such as cytochrome c into the cytoplasm, which eventually induces cell-death pathways and leads to cell demise [84]. Inhibitors of mPT have been suggested to be neuroprotective [85]. Among these inhibitors, nortriptyline is an FDA-approved tricyclic antidepressant [85] that significantly delays disease onset and reduces motor neuron loss in ALS mice [86]. Caspase-3 activation and cytochrome c release in the lumbar spinal cord of ALS mice were inhibited by nortriptyline [86]. Another mPT inhibitor, the immunosuppressant cyclosporine, has mitochondrial protective properties. It prevents the assembly of mPT and stabilizes mitochondrial membranes, thus preventing apoptosis [87]. Intrathecal injections of cyclosporine starting at disease onset extended the survival of ALS mice [87]. Taken together, these studies support the potential of mPT inhibitors in therapeutic applications for ALS.

P7C3 and its analog P7C3A20 are aminopropyl carbazoles shown to be neuroprotective and to encourage neurogenesis. P7C3 exhibits protection of mitochondrial member integrity against calcium in vitro [88]. Besides its neuroprotective effect in a mouse model of Parkinson disease [89], when administrated at onset, P7C3A20 was also neuroprotective in a G93A SOD1 mouse model of ALS, protecting lumbar spinal motor neurons and improving motor performance in accelerating rotarod test and walking gait [90].

Apoptotic pathways and anti-apoptotic agents

Though debatable, there is evidence that apoptosis, particularly the mitochondrial apoptotic pathway, has some involvement in ALS. An early event in the mechanism of toxicity of ALS is the activation of initiator caspase-1 [349192]. The deletion of the Bax/Bak pathways of mitochondrial apoptosis in G93A mice provides strong support for the role of mitochondrially regulated apoptosis in ALS [93]. Understanding these apoptotic pathways is an active area for future ALS therapeutic studies.

Melatonin has an indole core skeleton, easily penetrates the BBB, and acts as an anti-apoptotic agent, an antioxidant, and a free radical scavenger [94]. Our own recent studies indicate that melatonin administrated by intraperitoneal injection not only delays disease onset and increases mortality in both SOD1G93A ALS mice and R6/2 Huntington’s disease transgenic mice but also inhibits mutant huntingtin-induced cell death and caspase activation and preserves melatonin receptor 1A in mutant huntingtin ST14A cells [9597]. In addition, melatonin reduces superoxide-induced cell death and attenuates glutamate excitotoxicity in NSC-34 cells [98]. Astrocyte activation with increased levels of GFAP (a marker for astrocytes) and microglial activation with increased levels of RCA-1 (a marker for microglia/macrophages) in the spinal cord are both correlated with motor-neuron degeneration. Interestingly, we found that melatonin treatment reduced the expression of GFAP and RCA-1 in mSOD1G93A ALS mice [98]. High oral doses of melatonin delayed disease progression and increased longevity in SOD1G93A mice [98]. In a clinical safety study, high-dose rectally administered melatonin was well tolerated in patients with sALS [98]. However, melatonin has two weaknesses: a relatively short half-life (t½ = 20–50 min in adults) and multiple binding sites. Future research should be directed at identifying analogs and further drug design to overcome melatonin’s weaknesses and strengthen its neuroprotective effects in slowing disease progression in ALS.

The hematopoietic growth factor erythropoietin (EPO) inhibits apoptotic neuronal changes [99]. Non-hematopoietic erythropoietin derivatives, including asialo and carbamylated erythropoietin, improve motor behavior and reduce motor neuron loss. They reduce the activation of microglia and astrocytes in vivo and in vitro [100]. EPO prevented neuronal injury and early motor neuron degeneration and delayed the onset of motor deterioration in female animals without prolonging survival [101]. EPO levels in the CSF have been reported to be decreased in ALS patients [102].

Minocycline is an antibiotic approved by the FDA for the treatment of bacterial infection. As an anti-apoptotic agent, it inhibits the release of cytochrome c, the activation of caspase-1 and -3, reactive microgliosis, as well as p38 mitogen-activated protein kinase [103105]. Minocycline improved muscle strength, increased longevity, and delayed the onset of motor neuron degeneration in G93A mice [103] and SOD1G37R mice [106]. It also reduced microglial activation in G93A mice [107]. Moreover, a cocktail of minocycline and creatine offered additive neuroprotection, improving motor performance and extending survival in mSOD1G93A mice [82]. TCH346 treatment at disease onset slowed disease progression, increased lifespan, and preserved both body weight and motor performance in a mouse model of mutant with progressive motor neuronopathy [108]. Additionally, TCH346 prevents p53-related neuronal apoptosis. However, chronic subcutaneous treatment with TCH346 offers no benefit in high-copy SOD1G93A mice [109].

zVAD-fmk is a broad caspase inhibitor that delays disease onset and mortality in mSOD1G93A mice. zVAD-fmk inhibits caspase-1 activity as well as caspase-1 and caspase-3 mRNA upregulation [92110]. However, because zVAD-fmk is too toxic for human use, development of safer versions is under way.

Oxidative damage and anti-oxidative agents

Several neuroprotective agents with antioxidant capabilities have been studied in relation to ALS. Here we summarize their neuroprotective effects and the success of some antioxidants that target the oxidative stress pathogenic pathway in ALS animals.

Manganese porphyrin (AEOL10150), an antioxidant and a free-radical scavenger, markedly extends survival of mSOD1G93A mice when administered at symptom onset [111] as well as improving motor neuron architecture and reducing astrogliosis. Manganese porphyrin also shows anti-inflammatory properties [112]. Rasagiline (AZI-LECT), a monoamine oxidase inhibitor used in the treatment of Parkinson’s disease [113], has neuroprotective properties via its antioxidant activity and has demonstrated mitochondrial protection [114]. Rasagiline, either alone or in combination with riluzole, improves both motor performance and survival in mSOD1G93A mice [114].

Various metals are cytotoxic to motor neurons, leading to neurodegeneration. The lipophilic metal chelators DP-109 and DP-460 chelate calcium, copper, and zinc; both extended survival, improved motor performance, and reduced spinal cord cell loss and oxidative damage markers, as well as decreased reactive astrogliosis and microgliosis in ALS mice [115]. Iron dysregulation promotes oxidative damage, and altered iron homeostasis has been found in ALS patients [116]. The HFE gene is involved in iron regulation, and HFE polymorphism is increased in ALS patients [117]. The multifunctional iron-chelating drugs M30 and HLA20 reduce neurotoxicity induced by the peroxynitrite ion generator SIN-1 and H2O2, augment the expression of iron metabolism—related protein Tfr in NSC-34 cells [118], and prevent G93A SOD1-induced toxicity. M30 significantly delays disease onset and extends the survival of mSOD1G93Amice [119]. Thus, metal chelators may become a source for screening drugs against ALS.

H2O2 and UV light produce oxidative stress and cytotoxicity in SH-SY5Y cells. These effects are reversed by treatment with N-acetyl-L-cysteine (NAC), an antioxidant [120]. NAC improves survival and delays the onset of motor impairment in G93A mice [121], and the NAC precursor glutathione significantly reduced lower motor neuron degeneration in the wobbler mouse [122]. The antioxidant vitamin E (α-tocopherol) slows the onset and progression of paralysis in ALS mice, implying a possible future role in ALS prevention [123].

Inflammatory pathways and anti-inflammatory agents

Inflammation plays a major role in the pathogenesis of motor neuron death in ALS, and neuroinflammation accelerates disease progression.

Selective mutant expression of SOD1 in motor neurons did not progress to ALS disease [124125], while reduction of mSOD1 in motor neurons delayed disease onset and slowed early progression but had no benefit in terms of later disease progression and survival [126127]. A mixture of wild-type and mSOD1 in motor neurons was not sufficient to trigger disease onset [128129], and total ablation of mSOD1 expression (in either astrocytes or microglia) slowed disease progression and extended survival [126127130]. Together, these findings indicate that ALS is not motor–neuron autonomous and that glial cells play important roles in motor neuron degeneration.

Evidence further supports the contention that expression of mSOD1 in motor neurons dominantly and primarily initiates ALS pathogenesis (in other words, affects disease onset and early stages of the disease), whereas neighboring cell types other than motor neurons (including microglia, astrocytes, and Schwann cells) as well as other cells interact both with damaged motor neurons and each other to mediate and affect ALS disease progression and duration [131132].

Rather than playing simply a supportive role, astrocytes surrounding motor neurons provide nutrients, maintain the homeostatic environment, and carry out multiple functions. Impaired astrocytic functions including extracellular glutamate clearance and neurotrophic factors release have been implicated in ALS disease. Astrogliosis is present at symptomatic stages of ALS mice [97133] and may precede and drive the deterioration of motor activities in animal models of ALS [134135], while astrogliosis is detectable in post-mortem spinal cord tissue from fALS and sALS patients [136]. Accumulating evidence shows that astrocytes may modulate microglial activation and infiltration, speeding disease progression [127137138]. We and other researchers have reported that astrocytes carrying ALS-causing genes, especially mutated SOD1, as the non-cell autonomous components in ALS pathogenesis, may play a critical role in stimulating damage and degeneration of neighboring motor neurons [58137143]. The selective toxicity induced by mSOD1 astrocytes is involved in multiple mechanisms including activation of oxidative stress, secretion of toxic factors, disruption of Ca2+ oscillations, Wnt signaling dysfunction, and glutamate- and mGluR5-mediated excitotoxicity [3258143145]. Astrocytes derived from familiar ALS and sporadic ALS patients are similarly toxic to motor neurons [146]. On the other hand, wild-type glial cells extend the survival of mSOD1 motor neurons [144].

Microglial cells immigrate from the periphery and enter the brain and spinal cord through the BBB in response to damaged motor neurons. Indeed, activated microglia are implicated in the pathogenesis of ALS, and the intensity of microglial activation is correlated with the severity of motor neuron damage [131147148]. The fact that progression can be slowed by selectively decreasing mSOD1 in microglia or the addition of microglia expressing normal SOD1 suggests that mSOD1 affects disease progression more than disease initiation [126130]. In addition, by wrapping full-length axons of lower motor neurons in myelin, neighboring Schwann cells protect signal transduction. However, there are controversial reports that other glial-type oligodendrocytes and NG2 cells are involved in ALS pathogenesis [131]. Additionally, T lymphocytes also play a role in mSOD1-caused ALS [149150].

Below we summarize a number of protective agents targeting inflammatory pathways shown in animal models to demonstrate beneficial effects against ALS.

Cyclooxygenase-2 (COX-2) is a key molecule in the inflammatory pathogenic pathway in ALS. Expressed in spinal astrocytes and neurons, COX-2 catalyzes prostaglandin 2, which stimulates glutamate release from astrocytes via a calcium-dependent pathway [151155].

Moreover, along with cytokines, reactive oxygen species, and free radicals, COX-2 is involved in the inflammatory process in the central nervous system [156]. The COX-2 inhibitor and anti-inflammatory agent CELEBREX (celecoxib) significantly delays the onset of weakness and weight loss and prolongs survival of ALS mice; the administration of CELEBREX in mice provides significant preservation of spinal neurons and reduced astrogliosis and microglial activation [157].

TNF-α activates microglia and introduces neuronal apoptosis, and elevated levels of TNF-α have been found in the spinal cords of G93A mice and serum of humans with ALS [158160]. TNF-α in the lumbar spinal cord of G93A mice is increased before symptom onset and loss of motor neurons and is correlated with pathological progression, indicating that activation of inflammation (including TNF-α) plays a critical role in ALS pathogenesis [161162].

Both thalidomide and its analog lenalidomide are immunomodulatory agents that inhibit TNF-α production, attenuate weight loss, enhance motor performance, decrease motor neuron death, and significantly increase survival in mSOD1G93A mice when administered prior to onset of disease [133]. When administered after symptom onset, lenalidomide also provides neuroprotection in mSOD1G93A mice [1]. The anti-inflammatory and immunomodulatory properties of both agents play a role in their neuroprotection; both reduce the expression of proinflammatory cytokines [1133]. Microglial activation was ameliorated and neuronal loss was reduced in lenalidomide-treated mice [1].

Cannabinoids produce anti-inflammatory actions via cannabinoid receptor 2 (CB2) and cannabinoid receptor 1 (CB1). There is significant temporal elevation of CB2 mRNA and CB2-related activation of microglia/macrophages in the spinal cord of G93A mice [24] and in post-mortem human ALS samples [163]. Cannabinoids ameliorate disease progression in animal models of ALS. Δ9-tetrahydrocannabinol (THC), the major effective active ingredient of cannabis, is a CB2 antagonist. It delays motor impairment and prolongs survival when administered either before or after disease onset in ALS mice [164]. Δ9-THC is also anti-excitotoxic, reducing oxidative damage in vitro [164]. AM-1241 is a CB2 selective agonist that prolongs survival in mSOD1G93A mice when administered either after disease onset [165] or at symptom onset [24]. Cannabinoids may prove useful in the treatment of ALS.

Conclusions

Except for riluzole, promising results from animal models of ALS have not translated well into humans. For instance, talampanel, CELEBREX, and thalidomide did not succeed in phase II trials in ALS patients [166168], while minocycline, olesoxime, and TCH346 failed in large phase III or phase II/III clinical trials among ALS patients [169170]. These results highlight the challenges in predicting human clinical trial success from animal models of disease [171]. Applying rational design to produce significant improvements over current ALS preclinical studies should help produce more reliable and translatable results.

The translational failures may be explained in part by the following: (i) though SOD1G93A transgenic mice have been the most common model for therapeutic agent studies, there is genetic variation/mutation among ALS patients [172]. To what extent the SOD1 mouse model faithfully represents human ALS disease remains debatable; it may more accurately represent a small fraction of fALS rather than sALS patients [173174]. Unanswered questions include (1) how generalizable the drugs targeting pathogenic mechanisms of mSOD1G93A mice are to other animal models of ALS, and whether similar events occur in human ALS. Given the difficulty of treating a progressive disease with symptoms appearing in late adulthood, besides the timing of drug treatment before or on onset, mice trials could be more carefully designed, controlled, and elaborated to make them more translatable to the clinic, (2) methodological flaws such as randomization, blinded outcome assessment, and sample size calculation should be corrected [175], and more accurate controls such as sibling matched controls and monitoring the copy number of transgenes would improve the design of pre-clinical trials, and (3) Clinical trials have been marked by the failure to determine the correct dosage and ideal duration, inappropriate sample sizes, and poor selection of timing factors to identify a person as eligible. It may be necessary to identify a primary outcome for measurement other than survival, with biomarkers providing information on target engagement and effects on disease progression [173176178].

The neurobiology of ALS is complicated by multiple pathogenic mechanisms; which one(s) will predominate in any particular patient remains unpredictable. Drug candidates that ameliorate symptoms (including damage to muscle function) but provide no benefits in terms of survival or delayed onset are still valuable and informative for further investigation. A pleiotropic neuroprotective agent targeting multiple mechanisms may be most promising.

Of the current clinical trials, both melatonin [9798] and Δ9-THC [179], though in the early stages of clinical study, have multiple modes of action, including anti-oxidant, anti-excitotoxity, and anti-inflammatory effects. Along with their protection against apoptosis and mitochondrial damage (in the case of melatonin), those qualities may predict strong showings in further investigations. In addition, ceftriaxone’s effects against excitotoxity and apoptosis imply great promise. Melatonin’s indole core skeleton may facilitate further rational drug design and analog search, while the CB2 receptor’s non-selective or selective agonist/antagonists may provide a novel source of candidate drugs for screening against ALS. On the other hand, finding a single “magic bullet” for this heterogeneous disease may be difficult, and it is likely that the most effective treatment will turn out to be a combination therapy involving drugs that target pathogenic pathways.

Further direction of basic research should focus on deepening our understanding of the mechanisms underlying ALS pathogenesis, highlighting first/early triggers of this disease. Given the common appearance of reactive microglial and astroglial cells in ALS, Alzheimer’s disease, Parkinson’s disease, and Huntington’s disease, the notion of “non-cell autonomous” neurodegeneration has been applied to these neurodegenerative diseases. Further research on the biology, function, activation mechanisms, and regulation of astrocyte and microglial cell function is urgently needed and should make significant contributions to the discovery of therapeutic agents that will slow ALS disease progression, benefiting sALS and fALS patients, as well as those suffering from other neurodegenerative diseases. We anticipate that the accumulating knowledge of pathogenic mechanisms, drug discovery, and drug bioavailability, along with the evidence gathered in experimental preclinical studies and human trials, will soon yield more effective therapies for ALS.

Acknowledgments

We thank Drs. Chunyan Li and Xiao-Yun Liu for helpful discussions of this article. This work was supported by grants from the Muscular Dystrophy Association (157511 and 254530 to X.W.), the ALS Therapy Alliance (to X.W.), and the National Institutes of Health/National Institute of Neurological Disorders and Stroke (NS55072 to X.W.).

Contributor Information

Rachna S. Pandya, Department of Neurosurgery, Brigham and Women’s Hospital, Harvard Medical School, Boston, MA 02115, USA.
Haining Zhu, Department of Molecular and Cellular Biochemistry, College of Medicine, University of Kentucky, Lexington, KY 40536, USA.
Wei Li, Department of Neurosurgery, Brigham and Women’s Hospital, Harvard Medical School, Boston, MA 02115, USA.
Robert Bowser, Division of Neurobiology, Barrow Neurological Institute, Phoenix, AZ 85013, USA.
Robert M. Friedlander, Department of Neurosurgery, University of Pittsburgh School of Medicine, Pittsburgh, PA 15213, USA.
Xin Wang, Department of Neurosurgery, Brigham and Women’s Hospital, Harvard Medical School, Boston, MA 02115, USA.

References

1. Neymotin A, Petri S, Calingasan NY, Wille E, Schafer P, Stewart C, Hensley K, Beal MF, Kiaei M. Lenalidomide (Revlimid) administration at symptom onset is neuroprotective in a mouse model of amyotrophic lateral sclerosis. Exp Neurol. 2009;220:191–197. [PMC free article]  [PubMed]
2. Wong M, Martin LJ. Skeletal muscle-restricted expression of human SOD1 causes motor neuron degeneration in transgenic mice. Hum Mol Genet. 2010;19:2284–2302. [PMC free article]  [PubMed]
3. Dobrowolny G, Aucello M, Rizzuto E, Beccafico S, Mammucari C, Boncompagni S, Belia S, Wannenes F, Nicoletti C, Del Prete Z, et al. Skeletal muscle is a primary target of SOD1G93A-mediated toxicity. Cell Metab. 2008;8:425–436.  [PubMed]
4. Onesto E, Rusmini P, Crippa V, Ferri N, Zito A, Galbiati M, Poletti A. Muscle cells and motoneurons differentially remove mutant SOD1 causing familial amyotrophic lateral sclerosis. J Neurochem. 2011;118:266–280. [PMC free article]  [PubMed]
5. Liang H, Ward WF, Jang YC, Bhattacharya A, Bokov AF, Li Y, Jernigan A, Richardson A, Van Remmen H. PGC-1alpha protects neurons and alters disease progression in an amyotrophic lateral sclerosis mouse model. Muscle Nerve. 2011;44:947–956.  [PubMed]
6. Da Cruz S, Parone PA, Lopes VS, Lillo C, McAlonis-Downes M, Lee SK, Vetto AP, Petrosyan S, Marsala M, Murphy AN, et al. Elevated PGC-1alpha activity sustains mitochondrial biogenesis and muscle function without extending survival in a mouse model of inherited ALS. Cell Metab. 2012;15:778–786.[PMC free article]  [PubMed]
7. Holzbaur EL, Howland DS, Weber N, Wallace K, She Y, Kwak S, Tchistiakova LA, Murphy E, Hinson J, Karim R, et al. Myostatin inhibition slows muscle atrophy in rodent models of amyotrophic lateral sclerosis. Neurobiol Dis. 2006;23:697–707.  [PubMed]
8. Morrison BM, Lachey JL, Warsing LC, Ting BL, Pullen AE, Underwood KW, Kumar R, Sako D, Grinberg A, Wong V, et al. A soluble activin type IIB receptor improves function in a mouse model of amyotrophic lateral sclerosis. Exp Neurol. 2009;217:258–268.  [PubMed]
9. Miller RG, Mitchell JD, Lyon M, Moore DH. Riluzole for amyotrophic lateral sclerosis (ALS)/motor neuron disease (MND) Cochrane Database Syst Rev. 2007;1:CD001447.  [PubMed]
10. Deng HX, Hentati A, Tainer JA, Iqbal Z, Cayabyab A, Hung WY, Getzoff ED, Hu P, Herzfeldt B, Roos RP, et al. Amyotrophic lateral sclerosis and structural defects in Cu, Zn superoxide dismutase. Science. 1993;261:1047–1051.  [PubMed]
11. Deng HX, Chen W, Hong ST, Boycott KM, Gorrie GH, Siddique N, Yang Y, Fecto F, Shi Y, Zhai H, et al. Mutations in UBQLN2 cause dominant X-linked juvenile and adult-onset ALS and ALS/dementia. Nature. 2011;477:211–215. [PMC free article]  [PubMed]
12. Neumann M, Sampathu DM, Kwong LK, Truax AC, Micsenyi MC, Chou TT, Bruce J, Schuck T, Grossman M, Clark CM, et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science. 2006;314:130–133.  [PubMed]
13. Kwiatkowski TJ, Jr, Bosco DA, Leclerc AL, Tamrazian E, Vanderburg CR, Russ C, Davis A, Gilchrist J, Kasarskis EJ, Munsat T, et al. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science. 2009;323:1205–1208.  [PubMed]
14. Vance C, Rogelj B, Hortobagyi T, De Vos KJ, Nishimura AL, Sreedharan J, Hu X, Smith B, Ruddy D, Wright P, et al. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science. 2009;323:1208–1211.  [PubMed]
15. Couthouis J, Hart MP, Shorter J, Dejesus-Hernandez M, Erion R, Oristano R, Liu AX, Ramos D, Jethava N, Hosangadi D, et al. Feature Article: a yeast functional screen predicts new candidate ALS disease genes. Proc Natl Acad Sci USA. 2011;108:20881–20890. [PMC free article]  [PubMed]
16. DeJesus-Hernandez M, Mackenzie IR, Boeve BF, Boxer AL, Baker M, Rutherford NJ, Nicholson AM, Finch NA, Flynn H, Adamson J, et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron. 2011;72:245–256. [PMC free article][PubMed]
17. Renton AE, Majounie E, Waite A, Simon-Sanchez J, Rollinson S, Gibbs JR, Schymick JC, Laaksovirta H, van Swieten JC, Myllykangas L, et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron. 2011;72:257–268. [PMC free article]  [PubMed]
18. Gurney ME, Pu H, Chiu AY, Dal Canto MC, Polchow CY, Alexander DD, Caliendo J, Hentati A, Kwon YW, Deng HX, et al. Motor neuron degeneration in mice that express a human Cu, Zn superoxide dismutase mutation. Science. 1994;264:1772–1775.  [PubMed]
19. Gurney ME. The use of transgenic mouse models of amyotrophic lateral sclerosis in preclinical drug studies. J Neurol Sci. 1997;152(Suppl 1):S67–S73.  [PubMed]
20. Leitner M, Menzies S, Lutz C.  Prize4Life. Cambridge, MA: The Jackson Laboratory, Bar Harbor, ME; 2009. Working with ALS mice. Guidelines for preclinical testing and colony management.
21. Turner BJ, Talbot K. Transgenics, toxicity and therapeutics in rodent models of mutant SOD1-mediated familial ALS. Prog Neurobiol. 2008;85:94–134.  [PubMed]
22. Peviani M, Caron I, Pizzasegola C, Gensano F, Tortarolo M, Bendotti C. Unraveling the complexity of amyotrophic lateral sclerosis: recent advances from the transgenic mutant SOD1 mice. CNS Neurol Disord: Drug Targets. 2010;9:491–503.  [PubMed]
23. Joo IS, Hwang DH, Seok JI, Shin SK, Kim SU. Oral administration of memantine prolongs survival in a transgenic mouse model of amyotrophic lateral sclerosis. J Clin Neurol. 2007;3:181–186.[PMC free article]  [PubMed]
24. Shoemaker JL, Seely KA, Reed RL, Crow JP, Prather PL. The CB2 cannabinoid agonist AM-1241 prolongs survival in a transgenic mouse model of amyotrophic lateral sclerosis when initiated at symptom onset. J Neurochem. 2007;101:87–98. [PMC free article]  [PubMed]
25. Ito H, Wate R, Zhang J, Ohnishi S, Kaneko S, Nakano S, Kusaka H. Treatment with edaravone, initiated at symptom onset, slows motor decline and decreases SOD1 deposition in ALS mice. Exp Neurol. 2008;213:448–455.  [PubMed]
26. Kalmar B, Novoselov S, Gray A, Cheetham ME, Margulis B, Greensmith L. Late-stage treatment with arimoclomol delays disease progression and prevents protein aggregation in the SOD1 mouse model of ALS. J Neurochem. 2008;107:339–350.  [PubMed]
27. Yoo YE, Ko CP. Treatment with trichostatin A initiated after disease onset delays disease progression and increases survival in a mouse model of amyotrophic lateral sclerosis. Exp Neurol. 2011;231:147–159.[PubMed]
28. Bennett EJ, Bence NF, Jayakumar R, Kopito RR. Global impairment of the ubiquitin-proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation. Mol Cell. 2005;17:351–365.  [PubMed]
29. Boillee S, Vande Velde C, Cleveland DW. ALS: a disease of motor neurons and their nonneuronal neighbors. Neuron. 2006;52:39–59.  [PubMed]
30. Gal J, Strom AL, Kilty R, Zhang F, Zhu H. p62 accumulates and enhances aggregate formation in model systems of familial amyotrophic lateral sclerosis. J Biol Chem. 2007;282:11068–11077.  [PubMed]
31. Maekawa S, Leigh PN, King A, Jones E, Steele JC, Bodi I, Shaw CE, Hortobagyi T, Al-Sarraj S. TDP-43 is consistently co-localized with ubiquitinated inclusions in sporadic and Guam amyotrophic lateral sclerosis but not in familial amyotrophic lateral sclerosis with and without SOD1 mutations. Neuropathology. 2009;29:672–683.  [PubMed]
32. Rossi D, Brambilla L, Valori CF, Roncoroni C, Crugnola A, Yokota T, Bredesen DE, Volterra A. Focal degeneration of astrocytes in amyotrophic lateral sclerosis. Cell Death Differ. 2008;15:1691–1700.[PubMed]
33. Mendonca DM, Chimelli L, Martinez AM. Expression of ubiquitin and proteasome in motorneurons and astrocytes of spinal cords from patients with amyotrophic lateral sclerosis. Neurosci Lett. 2006;404:315–319.  [PubMed]
34. Pasinelli P, Houseweart MK, Brown RH, Jr, Cleveland DW. Caspase-1 and -3 are sequentially activated in motor neuron death in Cu, Zn superoxide dismutase-mediated familial amyotrophic lateral sclerosis. Proc Natl Acad Sci USA. 2000;97:13901–13906. [PMC free article]  [PubMed]
35. Urushitani M, Kurisu J, Tsukita K, Takahashi R. Proteasomal inhibition by misfolded mutant superoxide dismutase 1 induces selective motor neuron death in familial amyotrophic lateral sclerosis. J Neurochem. 2002;83:1030–1042.  [PubMed]
36. Aquilano K, Rotilio G, Ciriolo MR. Proteasome activation and nNOS down-regulation in neuroblastoma cells expressing a Cu, Zn superoxide dismutase mutant involved in familial ALS. J Neurochem. 2003;85:1324–1335.  [PubMed]
37. Cheroni C, Marino M, Tortarolo M, Veglianese P, De Biasi S, Fontana E, Zuccarello LV, Maynard CJ, Dantuma NP, Bendotti C. Functional alterations of the ubiquitin–proteasome system in motor neurons of a mouse model of familial amyotrophic lateral sclerosis. Hum Mol Genet. 2009;18:82–96. [PMC free article][PubMed]
38. Komatsu M, Waguri S, Chiba T, Murata S, Iwata J, Tanida I, Ueno T, Koike M, Uchiyama Y, Kominami E, et al. Loss of autophagy in the central nervous system causes neurodegeneration in mice. Nature. 2006;441:880–884.  [PubMed]
39. Hetz C, Thielen P, Matus S, Nassif M, Court F, Kiffin R, Martinez G, Cuervo AM, Brown RH, Glimcher LH. XBP-1 deficiency in the nervous system protects against amyotrophic lateral sclerosis by increasing autophagy. Genes Dev. 2009;23:2294–2306. [PMC free article]  [PubMed]
40. Nakano T, Nakaso K, Nakashima K, Ohama E. Expression of ubiquitin-binding protein p62 in ubiquitin-immunoreactive intraneuronal inclusions in amyotrophic lateral sclerosis with dementia: analysis of five autopsy cases with broad clinicopathological spectrum. Acta Neuropathol. 2004;107:359–364.[PubMed]
41. Parkinson N, Ince PG, Smith MO, Highley R, Skibinski G, Andersen PM, Morrison KE, Pall HS, Hardiman O, Collinge J, et al. ALS phenotypes with mutations in CHMP2B (charged multivesicular body protein 2B) Neurology. 2006;67:1074–1077.  [PubMed]
42. Seilhean D, Cazeneuve C, Thuries V, Russaouen O, Millecamps S, Salachas F, Meininger V, Leguern E, Duyckaerts C. Accumulation of TDP-43 and alpha-actin in an amyotrophic lateral sclerosis patient with the K17I ANG mutation. Acta Neuropathol. 2009;118:561–573.  [PubMed]
43. Ferguson CJ, Lenk GM, Meisler MH. Defective autophagy in neurons and astrocytes from mice deficient in PI(3,5)P2. Hum Mol Genet. 2009;18:4868–4878. [PMC free article]  [PubMed]
44. Braak H, Ludolph A, Thal DR, Del Tredici K. Amyotrophic lateral sclerosis: dash-like accumulation of phosphorylated TDP-43 in somatodendritic and axonal compartments of somatomotor neurons of the lower brainstem and spinal cord. Acta Neuropathol. 2010;120:67–74.  [PubMed]
45. Deng HX, Zhai H, Bigio EH, Yan J, Fecto F, Ajroud K, Mishra M, Ajroud-Driss S, Heller S, Sufit R, et al. FUS-immunoreactive inclusions are a common feature in sporadic and non-SOD1 familial amyotrophic lateral sclerosis. Ann Neurol. 2010;67:739–748.  [PubMed]
46. Gal J, Strom AL, Kwinter DM, Kilty R, Zhang J, Shi P, Fu W, Wooten MW, Zhu H. Sequestosome 1/p62 links familial ALS mutant SOD1 to LC3 via an ubiquitin-independent mechanism. J Neurochem. 2009;111:1062–1073. [PMC free article]  [PubMed]
47. Kieran D, Kalmar B, Dick JR, Riddoch-Contreras J, Burnstock G, Greensmith L. Treatment with arimoclomol, a coinducer of heat shock proteins, delays disease progression in ALS mice. Nat Med. 2004;10:402–405.  [PubMed]
48. Cudkowicz ME, Shefner JM, Simpson E, Grasso D, Yu H, Zhang H, Shui A, Schoenfeld D, Brown RH, Wieland S, et al. Arimoclomol at dosages up to 300 mg/day is well tolerated and safe in amyotrophic lateral sclerosis. Muscle Nerve. 2008;38:837–844.  [PubMed]
49. Lanka V, Wieland S, Barber J, Cudkowicz M. Arimoclomol: a potential therapy under development for ALS. Expert Opin Investig Drugs. 2009;18:1907–1918.  [PubMed]
50. Lange D. Abstract C46: pyrimethamine as a therapy for SOD1 associated FALS: early findings. Amyotroph Lateral Scler. 2008;9(Suppl 1):45–47.
51. Wright PD, Huang M, Weiss A, Matthews J, Wightman N, Glicksman M, Brown RH., Jr Screening for inhibitors of the SOD1 gene promoter: pyrimethamine does not reduce SOD1 levels in cell and animal models. Neurosci Lett. 2010;482:188–192. [PMC free article]  [PubMed]
52. Fornai F, Longone P, Cafaro L, Kastsiuchenka O, Ferrucci M, Manca ML, Lazzeri G, Spalloni A, Bellio N, Lenzi P, et al. Lithium delays progression of amyotrophic lateral sclerosis. Proc Natl Acad Sci USA. 2008;105:2052–2057. [PMC free article]  [PubMed]
53. Caldero J, Brunet N, Tarabal O, Piedrafita L, Hereu M, Ayala V, Esquerda JE. Lithium prevents excitotoxic cell death of motoneurons in organotypic slice cultures of spinal cord. Neuroscience. 2010;165:1353–1369.  [PubMed]
54. Chen RW, Qin ZH, Ren M, Kanai H, Chalecka-Franaszek E, Leeds P, Chuang DM. Regulation of c-Jun N-terminal kinase, p38 kinase and AP-1 DNA binding in cultured brain neurons: roles in glutamate excitotoxicity and lithium neuroprotection. J Neurochem. 2003;84:566–575.  [PubMed]
55. Gill A, Kidd J, Vieira F, Thompson K, Perrin S. No benefit from chronic lithium dosing in a sibling-matched, gender-balanced, investigator-blinded trial using a standard mouse model of familial ALS. Plos One. 2009;4:e6489. [PMC free article]  [PubMed]
56. Aggarwal SP, Zinman L, Simpson E, McKinley J, Jackson KE, Pinto H, Kaufman P, Conwit RA, Schoenfeld D, Shefner J, et al. Safety and efficacy of lithium in combination with riluzole for treatment of amyotrophic lateral sclerosis: a randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2010;9:481–488. [PMC free article]  [PubMed]
57. Staats KA, Van Den Bosch L. Astrocytes in amyotrophic lateral sclerosis: direct effects on motor neuron survival. J Biol Phys. 2009;35:337–346. [PMC free article]  [PubMed]
58. Martorana F, Brambilla L, Valori CF, Bergamaschi C, Roncoroni C, Aronica E, Volterra A, Bezzi P, Rossi D. The BH4 domain of Bcl-X(L) rescues astrocyte degeneration in amyotrophic lateral sclerosis by modulating intracellular calcium signals. Hum Mol Genet. 2012;21:826–840.  [PubMed]
59. Bryson HM, Fulton B, Benfield P. Riluzole. A review of its pharmacodynamic and pharmacokinetic properties and therapeutic potential in amyotrophic lateral sclerosis. Drugs. 1996;52:549–563.  [PubMed]
60. Bellingham MC. A review of the neural mechanisms of action and clinical efficiency of riluzole in treating amyotrophic lateral sclerosis: what have we learned in the last decade? CNS Neurosci Ther. 2011;17:4–31.  [PubMed]
61. Rothstein JD, Patel S, Regan MR, Haenggeli C, Huang YH, Bergles DE, Jin L, Dykes Hoberg M, Vidensky S, Chung DS, et al. Beta-lactam antibiotics offer neuroprotection by increasing glutamate transporter expression. Nature. 2005;433:73–77.  [PubMed]
62. Paizs M, Tortarolo M, Bendotti C, Engelhardt JI, Siklos L. Talampanel reduces the level of motoneuronal calcium in transgenic mutant SOD1 mice only if applied presymptomatically. Amyotroph Lateral Scler. 2011;12:340–344. [PMC free article]  [PubMed]
63. Chen HS, Pellegrini JW, Aggarwal SK, Lei SZ, Warach S, Jensen FE, Lipton SA. Open-channel block of N-methyl-D-aspartate (NMDA) responses by memantine: therapeutic advantage against NMDA receptor-mediated neurotoxicity. J Neurosci. 1992;12:4427–4436.  [PubMed]
64. Wang R, Zhang D. Memantine prolongs survival in an amyotrophic lateral sclerosis mouse model. Eur J Neurosci. 2005;22:2376–2380.  [PubMed]
65. Lv L, Tang YP, Han X, Wang X, Dong Q. Therapeutic application of histone deacetylase inhibitors for stroke. Cent Nerv Syst Agents Med Chem. 2011;11:138–149.  [PubMed]
66. Leng Y, Chuang DM. Endogenous alpha-synuclein is induced by valproic acid through histone deacetylase inhibition and participates in neuroprotection against glutamate-induced excitotoxicity. J Neurosci. 2006;26:7502–7512.  [PubMed]
67. Ragancokova D, Song Y, Nau H, Dengler R, Krampfl K, Petri S. Modulation of synaptic transmission and analysis of neuroprotective effects of valproic acid and derivates in rat embryonic motoneurons. Cell Mol Neurobiol. 2010;30:891–900.  [PubMed]
68. Lv L, Han X, Sun Y, Wang X, Dong Q. Valproic acid improves locomotion in vivo after SCI and axonal growth of neurons in vitro. Exp Neurol. 2012;233:783–790. [PMC free article]  [PubMed]
69. Sugai F, Yamamoto Y, Miyaguchi K, Zhou Z, Sumi H, Hamasaki T, Goto M, Sakoda S. Benefit of valproic acid in suppressing disease progression of ALS model mice. Eur J Neurosci. 2004;20:3179–3183.[PubMed]
70. Feng HL, Leng Y, Ma CH, Zhang J, Ren M, Chuang DM. Combined lithium and valproate treatment delays disease onset, reduces neurological deficits and prolongs survival in an amyotrophic lateral sclerosis mouse model. Neuroscience. 2008;155:567–572. [PMC free article]  [PubMed]
71. Ryu H, Lee J, Olofsson BA, Mwidau A, Dedeoglu A, Escudero M, Flemington E, Azizkhan-Clifford J, Ferrante RJ, Ratan RR. Histone deacetylase inhibitors prevent oxidative neuronal death independent of expanded polyglutamine repeats via an Sp1-dependent pathway. Proc Natl Acad Sci USA. 2003;100:4281–4286. [PMC free article]  [PubMed]
72. Chuang DM, Leng Y, Marinova Z, Kim HJ, Chiu CT. Multiple roles of HDAC inhibition in neurodegenerative conditions. Trends Neurosci. 2009;32:591–601. [PMC free article]  [PubMed]
73. Ferri A, Cozzolino M, Crosio C, Nencini M, Casciati A, Gralla EB, Rotilio G, Valentine JS, Carri MT. Familial ALS-superoxide dismutases associate with mitochondria and shift their redox potentials. Proc Natl Acad Sci USA. 2006;103:13860–13865. [PMC free article]  [PubMed]
74. Kirkinezos IG, Bacman SR, Hernandez D, Oca-Cossio J, Arias LJ, Perez-Pinzon MA, Bradley WG, Moraes CT. Cytochrome c association with the inner mitochondrial membrane is impaired in the CNS of G93A-SOD1 mice. J Neurosci. 2005;25:164–172.  [PubMed]
75. Liu J, Lillo C, Jonsson PA, Vande Velde C, Ward CM, Miller TM, Subramaniam JR, Rothstein JD, Marklund S, Andersen PM, et al. Toxicity of familial ALS-linked SOD1 mutants from selective recruitment to spinal mitochondria. Neuron. 2004;43:5–17.  [PubMed]
76. Gribkoff VK, Bozik ME. KNS-760704 [(6R)-4,5,6,7-tet-rahydro-N6-propyl-2, 6-benzothiazole-diamine dihydrochloride monohydrate] for the treatment of amyotrophic lateral sclerosis. CNS Neurosci Ther. 2008;14:215–226.  [PubMed]
77. Martin LJ. Olesoxime, a cholesterol-like neuroprotectant for the potential treatment of amyotrophic lateral sclerosis. IDrugs. 2010;13:568–580. [PMC free article]  [PubMed]
78. Carri MT, Ferri A, Battistoni A, Famhy L, Gabbianelli R, Poccia F, Rotilio G. Expression of a Cu, Zn superoxide dismutase typical of familial amyotrophic lateral sclerosis induces mitochondrial alteration and increase of cytosolic Ca2 + concentration in transfected neuroblastoma SH-SY5Y cells. FEBS Lett. 1997;414:365–368.  [PubMed]
79. Bordet T, Buisson B, Michaud M, Drouot C, Galea P, Delaage P, Akentieva NP, Evers AS, Covey DF, Ostuni MA, et al. Identification and characterization of cholest-4-en-3-one, oxime (TRO19622), a novel drug candidate for amyotrophic lateral sclerosis. J Pharmacol Exp Ther. 2007;322:709–720.  [PubMed]
80. Sunyach C, Michaud M, Arnoux T, Bernard-Marissal N, Aebischer J, Latyszenok V, Gouarne C, Raoul C, Pruss RM, Bordet T, et al. Olesoxime delays muscle denervation, astrogliosis, microglial activation and motoneuron death in an ALS mouse model. Neuropharmacology. 2012;62:2346–2352.  [PubMed]
81. Klivenyi P, Ferrante RJ, Matthews RT, Bogdanov MB, Klein AM, Andreassen OA, Mueller G, Wermer M, Kaddurah-Daouk R, Beal MF. Neuroprotective effects of creatine in a transgenic animal model of amyotrophic lateral sclerosis. Nat Med. 1999;5:347–350.  [PubMed]
82. Zhang W, Narayanan M, Friedlander RM. Additive neuroprotective effects of minocycline with creatine in a mouse model of ALS. Ann Neurol. 2003;53:267–270.  [PubMed]
83. Klivenyi P, Kiaei M, Gardian G, Calingasan NY, Beal MF. Additive neuroprotective effects of creatine and cyclooxygenase 2 inhibitors in a transgenic mouse model of amyotrophic lateral sclerosis. J Neurochem. 2004;88:576–582.  [PubMed]
84. Li P, Nijhawan D, Budihardjo I, Srinivasula SM, Ahmad M, Alnemri ES, Wang X. Cytochrome c and dATP-dependent formation of Apaf-1/caspase-9 complex initiates an apoptotic protease cascade. Cell. 1997;91:479–489.  [PubMed]
85. Stavrovskaya IG, Narayanan MV, Zhang W, Krasnikov BF, Heemskerk J, Young SS, Blass JP, Brown AM, Beal MF, Friedlander RM, et al. Clinically approved heterocyclics act on a mitochondrial target and reduce stroke-induced pathology. J Exp Med. 2004;200:211–222. [PMC free article]  [PubMed]
86. Wang H, Guan Y, Wang X, Smith K, Cormier K, Zhu S, Stavrovskaya IG, Huo C, Ferrante RJ, Kristal BS, et al. Nortriptyline delays disease onset in models of chronic neurodegeneration. Eur J Neurosci. 2007;26:633–641.  [PubMed]
87. Keep M, Elmer E, Fong KS, Csiszar K. Intrathecal cyclosporin prolongs survival of late-stage ALS mice. Brain Res. 2001;894:327–331.  [PubMed]
88. Pieper AA, Xie S, Capota E, Estill SJ, Zhong J, Long JM, Becker GL, Huntington P, Goldman SE, Shen CH, et al. Discovery of a proneurogenic, neuroprotective chemical. Cell. 2010;142:39–51.[PMC free article]  [PubMed]
89. De Jesus-Cortes H, Xu P, Drawbridge J, Estill SJ, Huntington P, Tran S, Britt J, Tesla R, Morlock L, Naidoo J, et al. Neuroprotective efficacy of aminopropyl carbazoles in a mouse model of Parkinson disease. Proc Natl Acad Sci USA. 2012;109:17010–17015. [PMC free article]  [PubMed]
90. Tesla R, Wolf HP, Xu P, Drawbridge J, Estill SJ, Huntington P, McDaniel L, Knobbe W, Burket A, Tran S, et al. Neuroprotective efficacy of aminopropyl carbazoles in a mouse model of amyotrophic lateral sclerosis. Proc Natl Acad Sci USA. 2012;109:17016–17021. [PMC free article]  [PubMed]
91. Friedlander RM. Apoptosis and caspases in neurodegenerative diseases. N Engl J Med. 2003;348:1365–1375.  [PubMed]
92. Friedlander RM, Brown RH, Gagliardini V, Wang J, Yuan J. Inhibition of ICE slows ALS in mice. Nature. 1997;388:31.  [PubMed]
93. Reyes NA, Fisher JK, Austgen K, VandenBerg S, Huang EJ, Oakes SA. Blocking the mitochondrial apoptotic pathway preserves motor neuron viability and function in a mouse model of amyotrophic lateral sclerosis. J Clin Investig. 2010;120:3673–3679. [PMC free article]  [PubMed]
94. Wang X. The antiapoptotic activity of melatonin in neurodegenerative diseases. CNS Neurosci Ther. 2009;15:345–357. [PMC free article]  [PubMed]
95. Wang X, Zhu S, Pei Z, Drozda M, Stavrovskaya IG, Del Signore SJ, Cormier K, Shimony EM, Wang H, Ferrante RJ, et al. Inhibitors of cytochrome c release with therapeutic potential for Huntington’s disease. J Neurosci. 2008;28:9473–9485. [PMC free article]  [PubMed]
96. Wang X, Sirianni A, Pei Z, Cormier K, Smith K, Jiang J, Zhou S, Wang H, Zhao R, Yano H, et al. The melatonin MT1 receptor axis modulates mutant huntingtin-mediated toxicity. J Neurosci. 2011;31:14496–14507. [PMC free article]  [PubMed]
97. Zhang Y, Cook A, Kim J, Baranov SV, Jiang J, Smith K, Cormier K, Bennett E, Browser RP, Day AL, et al. Melatonin inhibits the caspase-1/cytochrome c/caspase-3 cell death pathway, inhibits MT1 receptor loss and delays disease progression in a mouse model of amyotrophic lateral sclerosis. Neurobiol Dis. 2013;55:26–35. [PMC free article]  [PubMed]
98. Weishaupt JH, Bartels C, Polking E, Dietrich J, Rohde G, Poeggeler B, Mertens N, Sperling S, Bohn M, Huther G, et al. Reduced oxidative damage in ALS by high-dose enteral melatonin treatment. J Pineal Res. 2006;41:313–323.  [PubMed]
99. Naganska E, Taraszewska A, Matyja E, Grieb P, Rafalowska J. Neuroprotective effect of erythropoietin in amyotrophic lateral sclerosis (ALS) model in vitro. Ultrastructural study. Folia Neuropathol. 2010;48:35–44.  [PubMed]
100. Mennini T, De Paola M, Bigini P, Mastrotto C, Fumagalli E, Barbera S, Mengozzi M, Viviani B, Corsini E, Marinovich M, et al. Nonhematopoietic erythropoietin derivatives prevent motoneuron degeneration in vitro and in vivo. Mol Med. 2006;12:153–160. [PMC free article]  [PubMed]
101. Grunfeld JF, Barhum Y, Blondheim N, Rabey JM, Melamed E, Offen D. Erythropoietin delays disease onset in an amyotrophic lateral sclerosis model. Exp Neurol. 2007;204:260–263.  [PubMed]
102. Brettschneider J, Widl K, Schattauer D, Ludolph AC, Tumani H. Cerebrospinal fluid erythropoietin (EPO) in amyotrophic lateral sclerosis. Neurosci Lett. 2007;416:257–260.  [PubMed]
103. Zhu S, Stavrovskaya IG, Drozda M, Kim BY, Ona V, Li M, Sarang S, Liu AS, Hartley DM, du Wu C, et al. Minocycline inhibits cytochrome c release and delays progression of amyotrophic lateral sclerosis in mice. Nature. 2002;417:74–78.  [PubMed]
104. Wang X, Zhu S, Drozda M, Zhang W, Stavrovskaya IG, Cattaneo E, Ferrante RJ, Kristal BS, Friedlander RM. Minocycline inhibits caspase-independent and -dependent mitochondrial cell death pathways in models of Huntington’s disease. Proc Natl Acad Sci USA. 2003;100:10483–10487.[PMC free article]  [PubMed]
105. Tikka T, Fiebich BL, Goldsteins G, Keinanen R, Koistinaho J. Minocycline, a tetracycline derivative, is neuroprotective against excitotoxicity by inhibiting activation and proliferation of microglia. J Neurosci. 2001;21:2580–2588.  [PubMed]
106. Kriz J, Nguyen MD, Julien JP. Minocycline slows disease progression in a mouse model of amyotrophic lateral sclerosis. Neurobiol Dis. 2002;10:268–278.  [PubMed]
107. Van Den Bosch L, Tilkin P, Lemmens G, Robberecht W. Minocycline delays disease onset and mortality in a transgenic model of ALS. NeuroReport. 2002;13:1067–1070.  [PubMed]
108. Sagot Y, Toni N, Perrelet D, Lurot S, King B, Rixner H, Mattenberger L, Waldmeier PC, Kato AC. An orally active anti-apoptotic molecule (CGP 3466B) preserves mitochondria and enhances survival in an animal model of motoneuron disease. Br J Pharmacol. 2000;131:721–728. [PMC free article]  [PubMed]
109. Groeneveld GJ, van Muiswinkel FL, de Leeuw van Weenen J, Blauw H, Veldink JH, Wokke JH, van den Berg LH, Bar PR. CGP 3466B has no effect on disease course of (G93A) mSOD1 transgenic mice. Amyotroph Lateral Scler Other Motor Neuron Disord. 2004;5:220–225.  [PubMed]
110. Li M, Ona VO, Guegan C, Chen M, Jackson-Lewis V, Andrews LJ, Olszewski AJ, Stieg PE, Lee JP, Przedborski S, et al. Functional role of caspase-1 and caspase-3 in an ALS transgenic mouse model. Science. 2000;288:335–339.  [PubMed]
111. Crow JP, Calingasan NY, Chen J, Hill JL, Beal MF. Manganese porphyrin given at symptom onset markedly extends survival of ALS mice. Ann Neurol. 2005;58:258–265.  [PubMed]
112. Bowler RP, Sheng H, Enghild JJ, Pearlstein RD, Warner DS, Crapo JD. A catalytic antioxidant (AEOL 10150) attenuates expression of inflammatory genes in stroke. Free Radic Biol Med. 2002;33:1141–1152.[PubMed]
113. Oldfield V, Keating GM, Perry CM. Rasagiline: a review of its use in the management of Parkinson’s disease. Drugs. 2007;67:1725–1747.  [PubMed]
114. Waibel S, Reuter A, Malessa S, Blaugrund E, Ludolph AC. Rasagiline alone and in combination with riluzole prolongs survival in an ALS mouse model. J Neurol. 2004;251:1080–1084.  [PubMed]
115. Petri S, Calingasan NY, Alsaied OA, Wille E, Kiaei M, Friedman JE, Baranova O, Chavez JC, Beal MF. The lipophilic metal chelators DP-109 and DP-460 are neuroprotective in a transgenic mouse model of amyotrophic lateral sclerosis. J Neurochem. 2007;102:991–1000.  [PubMed]
116. Mitchell HM, White DM, Domowicz MS, Kraig RP. Cold pre-conditioning neuroprotection depends on TNF-alpha and is enhanced by blockade of interleukin-11. J Neurochem. 2010;117:187–196.[PMC free article]  [PubMed]
117. Schymick JC, Talbot K, Traynor BJ. Genetics of sporadic amyotrophic lateral sclerosis. Hum Mol Genet. 2007;16(Spec No. 2):R233–R242.  [PubMed]
118. Kupershmidt L, Weinreb O, Amit T, Mandel S, Carri MT, Youdim MB. Neuroprotective and neuritogenic activities of novel multimodal iron-chelating drugs in motor-neuron-like NSC-34 cells and transgenic mouse model of amyotrophic lateral sclerosis. Faseb J. 2009;23:3766–3779.  [PubMed]
119. Wang Q, Zhang X, Chen S, Zhang S, Youdium M, Le W. Prevention of motor neuron degeneration by novel iron chelators in SOD1(G93A) transgenic mice of amyotrophic lateral sclerosis. Neurodegener Dis. 2011;8:310–321.  [PubMed]
120. Olivieri G, Baysang G, Meier F, Muller-Spahn F, Stahelin HB, Brockhaus M, Brack C. N-acetyl-L-cysteine protects SHSY5Y neuroblastoma cells from oxidative stress and cell cytotoxicity: effects on beta-amyloid secretion and tau phosphorylation. J Neurochem. 2001;76:224–233.  [PubMed]
121. Andreassen OA, Dedeoglu A, Klivenyi P, Beal MF, Bush AI. N-acetyl-L-cysteine improves survival and preserves motor performance in an animal model of familial amyotrophic lateral sclerosis. NeuroReport. 2000;11:2491–2493.  [PubMed]
122. Henderson JT, Javaheri M, Kopko S, Roder JC. Reduction of lower motor neuron degeneration in wobbler mice by N-acetyl-L-cysteine. J Neurosci. 1996;16:7574–7582.  [PubMed]
123. Wang H, O’Reilly EJ, Weisskopf MG, Logroscino G, McCullough ML, Schatzkin A, Kolonel LN, Ascherio A. Vitamin E intake and risk of amyotrophic lateral sclerosis: a pooled analysis of data from 5 prospective cohort studies. Am J Epidemiol. 2011;173:595–602. [PMC free article]  [PubMed]
124. Pramatarova A, Laganiere J, Roussel J, Brisebois K, Rouleau GA. Neuron-specific expression of mutant superoxide dismutase 1 in transgenic mice does not lead to motor impairment. J Neurosci. 2001;21:3369–3374.  [PubMed]
125. Lino MM, Schneider C, Caroni P. Accumulation of SOD1 mutants in postnatal motoneurons does not cause motoneuron pathology or motoneuron disease. J Neurosci. 2002;22:4825–4832.  [PubMed]
126. Boillee S, Yamanaka K, Lobsiger CS, Copeland NG, Jenkins NA, Kassiotis G, Kollias G, Cleveland DW. Onset and progression in inherited ALS determined by motor neurons and microglia. Science. 2006;312:1389–1392.  [PubMed]
127. Yamanaka K, Chun SJ, Boillee S, Fujimori-Tonou N, Yamashita H, Gutmann DH, Takahashi R, Misawa H, Cleveland DW. Astrocytes as determinants of disease progression in inherited amyotrophic lateral sclerosis. Nat Neurosci. 2008;11:251–253. [PMC free article]  [PubMed]
128. Clement AM, Nguyen MD, Roberts EA, Garcia ML, Boillee S, Rule M, McMahon AP, Doucette W, Siwek D, Ferrante RJ, et al. Wild-type nonneuronal cells extend survival of SOD1 mutant motor neurons in ALS mice. Science. 2003;302:113–117.  [PubMed]
129. Yamanaka K, Boillee S, Roberts EA, Garcia ML, McAlonis-Downes M, Mikse OR, Cleveland DW, Goldstein LS. Mutant SOD1 in cell types other than motor neurons and oligodendrocytes accelerates onset of disease in ALS mice. Proc Natl Acad Sci USA. 2008;105:7594–7599. [PMC free article]  [PubMed]
130. Wang L, Sharma K, Grisotti G, Roos RP. The effect of mutant SOD1 dismutase activity on non-cell autonomous degeneration in familial amyotrophic lateral sclerosis. Neurobiol Dis. 2009;35:234–240.[PMC free article]  [PubMed]
131. Lasiene J, Yamanaka K. Glial cells in amyotrophic lateral sclerosis. Neurol Res Int. 2011;2011:718987. [PMC free article]  [PubMed]
132. Philips T, Robberecht W. Neuroinflammation in amyotrophic lateral sclerosis: role of glial activation in motor neuron disease. Lancet Neurol. 2011;10:253–263.  [PubMed]
133. Kiaei M, Petri S, Kipiani K, Gardian G, Choi DK, Chen J, Calingasan NY, Schafer P, Muller GW, Stewart C, et al. Thalidomide and lenalidomide extend survival in a transgenic mouse model of amyotrophic lateral sclerosis. J Neurosci. 2006;26:2467–2473.  [PubMed]
134. Fujita K, Yamauchi M, Matsui T, Titani K, Takahashi H, Kato T, Isomura G, Ando M, Nagata Y. Increase of glial fibrillary acidic protein fragments in the spinal cord of motor neuron degeneration mutant mouse. Brain Res. 1998;785:31–40.  [PubMed]
135. Keller AF, Gravel M, Kriz J. Live imaging of amyotrophic lateral sclerosis pathogenesis: disease onset is characterized by marked induction of GFAP in Schwann cells. Glia. 2009;57:1130–1142.  [PubMed]
136. Schiffer D, Cordera S, Cavalla P, Migheli A. Reactive astrogliosis of the spinal cord in amyotrophic lateral sclerosis. J Neurol Sci. 1996;139(Suppl):27–33.  [PubMed]
137. Chen Y, Guan Y, Zhang Z, Liu H, Wang S, Yu L, Wu X, Wang X. Wnt signaling pathway is involved in the pathogenesis of amyotrophic lateral sclerosis in adult transgenic mice. Neurol Res. 2012;34:390–399.  [PubMed]
138. Chen Y, Guan Y, Liu H, Wu X, Yu L, Wang S, Zhao C, Du H, Wang X. Activation of the Wnt/beta-catenin signaling pathway is associated with glial proliferation in the adult spinal cord of ALS transgenic mice. Biochem Biophys Res Commun. 2012;420:397–403.  [PubMed]
139. Papadeas ST, Kraig SE, O’Banion C, Lepore AC, Maragakis NJ. Astrocytes carrying the superoxide dismutase 1 (SOD1G93A) mutation induce wild-type motor neuron degeneration in vivo. Proc Natl Acad Sci USA. 2011;108:17803–17808. [PMC free article]  [PubMed]
140. Lobsiger CS, Cleveland DW. Glial cells as intrinsic components of non-cell-autonomous neurodegenerative disease. Nat Neurosci. 2007;10:1355–1360. [PMC free article]  [PubMed]
141. Wang S, Guan Y, Chen Y, Li X, Zhang C, Yu L, Zhou F, Wang X. Role of Wnt1 and Fzd1 in the spinal cord pathogenesis of amyotrophic lateral sclerosis-transgenic mice. Biotechnol Lett. 2013;35:1199–1207.[PubMed]
142. Van Damme P, Bogaert E, Dewil M, Hersmus N, Kiraly D, Scheveneels W, Bockx I, Braeken D, Verpoorten N, Verhoeven K, et al. Astrocytes regulate GluR2 expression in motor neurons and their vulnerability to excitotoxicity. Proc Natl Acad Sci USA. 2007;104:14825–14830. [PMC free article][PubMed]
143. Li X, Guan Y, Chen Y, Zhang C, Shi C, Zhou F, Yu L, Juan J, Wang X. Expression of Wnt5a and its receptor Fzd2 is changed in the spinal cord of adult amyotrophic lateral sclerosis transgenic mice. Int J Clin Exp Pathol. 2013;6:1245–1260. [PMC free article]  [PubMed]
144. Marchetto MC, Muotri AR, Mu Y, Smith AM, Cezar GG, Gage FH. Non-cell-autonomous effect of human SOD1 G37R astrocytes on motor neurons derived from human embryonic stem cells. Cell Stem Cell. 2008;3:649–657.  [PubMed]
145. Nagai M, Re DB, Nagata T, Chalazonitis A, Jessell TM, Wichterle H, Przedborski S. Astrocytes expressing ALS-linked mutated SOD1 release factors selectively toxic to motor neurons. Nat Neurosci. 2007;10:615–622. [PMC free article]  [PubMed]
146. Haidet-Phillips AM, Hester ME, Miranda CJ, Meyer K, Braun L, Frakes A, Song S, Likhite S, Murtha MJ, Foust KD, et al. Astrocytes from familial and sporadic ALS patients are toxic to motor neurons. Nat Biotechnol. 2011;29:824–828. [PMC free article]  [PubMed]
147. Henkel JS, Beers DR, Zhao W, Appel SH. Microglia in ALS: the good, the bad, and the resting. J Neuroimmune Pharmacol. 2009;4:389–398.  [PubMed]
148. McGeer PL, McGeer EG. Inflammatory processes in amyotrophic lateral sclerosis. Muscle Nerve. 2002;26:459–470.  [PubMed]
149. Beers DR, Henkel JS, Zhao W, Wang J, Appel SH. CD4 + T cells support glial neuroprotection, slow disease progression, and modify glial morphology in an animal model of inherited ALS. Proc Natl Acad Sci USA. 2008;105:15558–15563. [PMC free article]  [PubMed]
150. Chiu IM, Chen A, Zheng Y, Kosaras B, Tsiftsoglou SA, Vartanian TK, Brown RH, Jr, Carroll MC. T lymphocytes potentiate endogenous neuroprotective inflammation in a mouse model of ALS. Proc Natl Acad Sci USA. 2008;105:17913–17918. [PMC free article]  [PubMed]
151. Drachman DB, Rothstein JD. Inhibition of cyclooxy-genase-2 protects motor neurons in an organotypic model of amyotrophic lateral sclerosis. Ann Neurol. 2000;48:792–795.  [PubMed]
152. Almer G, Guegan C, Teismann P, Naini A, Rosoklija G, Hays AP, Chen C, Przedborski S. Increased expression of the pro-inflammatory enzyme cyclooxygenase-2 in amyotrophic lateral sclerosis. Ann Neurol. 2001;49:176–185.  [PubMed]
153. Yasojima K, Tourtellotte WW, McGeer EG, McGeer PL. Marked increase in cyclooxygenase-2 in ALS spinal cord: implications for therapy. Neurology. 2001;57:952–956.  [PubMed]
154. Bezzi P, Carmignoto G, Pasti L, Vesce S, Rossi D, Rizzini BL, Pozzan T, Volterra A. Prostaglandins stimulate calcium-dependent glutamate release in astrocytes. Nature. 1998;391:281–285.  [PubMed]
155. Sanzgiri RP, Araque A, Haydon PG. Prostaglandin E(2) stimulates glutamate receptor-dependent astrocyte neuromodulation in cultured hippocampal cells. J Neurobiol. 1999;41:221–229.  [PubMed]
156. McGeer PL. COX-2 and ALS. Amyotroph Lateral Scler Other Motor Neuron Disord. 2001;2:121–122.  [PubMed]
157. Drachman DB, Frank K, Dykes-Hoberg M, Teismann P, Almer G, Przedborski S, Rothstein JD. Cyclooxygenase 2 inhibition protects motor neurons and prolongs survival in a transgenic mouse model of ALS. Ann Neurol. 2002;52:771–778.  [PubMed]
158. Robertson J, Beaulieu JM, Doroudchi MM, Durham HD, Julien JP, Mushynski WE. Apoptotic death of neurons exhibiting peripherin aggregates is mediated by the proinflammatory cytokine tumor necrosis factor-alpha. J Cell Biol. 2001;155:217–226. [PMC free article]  [PubMed]
159. Hensley K, Floyd RA, Gordon B, Mou S, Pye QN, Stewart C, West M, Williamson K. Temporal patterns of cytokine and apoptosis-related gene expression in spinal cords of the G93A-SOD1 mouse model of amyotrophic lateral sclerosis. J Neurochem. 2002;82:365–374.  [PubMed]
160. Poloni M, Facchetti D, Mai R, Micheli A, Agnoletti L, Francolini G, Mora G, Camana C, Mazzini L, Bachetti T. Circulating levels of tumour necrosis factor-alpha and its soluble receptors are increased in the blood of patients with amyotrophic lateral sclerosis. Neurosci Lett. 2000;287:211–214.  [PubMed]
161. Yoshihara T, Ishigaki S, Yamamoto M, Liang Y, Niwa J, Takeuchi H, Doyu M, Sobue G. Differential expression of inflammation- and apoptosis-related genes in spinal cords of a mutant SOD1 transgenic mouse model of familial amyotrophic lateral sclerosis. J Neurochem. 2002;80:158–167.  [PubMed]
162. Hensley LL, Ranganathan G, Wagner EM, Wells BD, Daniel JC, Vu D, Semenkovich CF, Zechner R, Kern PA. Transgenic mice expressing lipoprotein lipase in adipose tissue. Absence of the proximal 3′-untranslated region causes translational upregulation. J Biol Chem. 2003;278:32702–32709.  [PubMed]
163. Yiangou Y, Facer P, Durrenberger P, Chessell IP, Naylor A, Bountra C, Banati RR, Anand P. COX-2, CB2 and P2X7-immunoreactivities are increased in activated microglial cells/macrophages of multiple sclerosis and amyotrophic lateral sclerosis spinal cord. BMC Neurol. 2006;6:12. [PMC free article][PubMed]
164. Raman C, McAllister SD, Rizvi G, Patel SG, Moore DH, Abood ME. Amyotrophic lateral sclerosis: delayed disease progression in mice by treatment with a cannabinoid. Amyotroph Lateral Scler Other Motor Neuron Disord. 2004;5:33–39.  [PubMed]
165. Kim K, Moore DH, Makriyannis A, Abood ME. AM1241, a cannabinoid CB2 receptor selective compound, delays disease progression in a mouse model of amyotrophic lateral sclerosis. Eur J Pharmacol. 2006;542:100–105.  [PubMed]
166. Pascuzzi RM, Shefner J, Chappell AS, Bjerke JS, Tamura R, Chaudhry V, Clawson L, Haas L, Rothstein JD. A phase II trial of talampanel in subjects with amyotrophic lateral sclerosis. Amyotroph Lateral Scler. 2010;11:266–271.  [PubMed]
167. Cudkowicz ME, Shefner JM, Schoenfeld DA, Zhang H, Andreasson KI, Rothstein JD, Drachman DB. Trial of celecoxib in amyotrophic lateral sclerosis. Ann Neurol. 2006;60:22–31.  [PubMed]
168. Stommel EW, Cohen JA, Fadul CE, Cogbill CH, Graber DJ, Kingman L, Mackenzie T, Channon Smith JY, Harris BT. Efficacy of thalidomide for the treatment of amyotrophic lateral sclerosis: a phase II open label clinical trial. Amyotroph Lateral Scler. 2009;10:393–404. [PMC free article]  [PubMed]
169. Gordon PH, Moore DH, Miller RG, Florence JM, Verheijde JL, Doorish C, Hilton JF, Spitalny GM, MacArthur RB, Mitsumoto H, et al. Efficacy of minocycline in patients with amyotrophic lateral sclerosis: a phase III randomised trial. Lancet Neurol. 2007;6:1045–1053.  [PubMed]
170. Miller R, Bradley W, Cudkowicz M, Hubble J, Meininger V, Mitsumoto H, Moore D, Pohlmann H, Sauer D, Silani V, et al. Phase II/III randomized trial of TCH346 in patients with ALS. Neurology. 2007;69:776–784.  [PubMed]
171. Ludolph AC, Bendotti C, Blaugrund E, Chio A, Greensmith L, Loeffler JP, Mead R, Niessen HG, Petri S, Pradat PF, et al. Guidelines for preclinical animal research in ALS/MND: a consensus meeting. Amyotroph Lateral Scler. 2010;11:38–45.  [PubMed]
172. Pandya RS, Mao LL, Zhou EW, Bowser R, Zhu Z, Zhu Y, Wang X. Neuroprotection for amyotrophic lateral sclerosis: role of stem cells, growth factors, and gene therapy. Cent Nerv Syst Agents Med Chem. 2012;12:15–27. [PMC free article]  [PubMed]
173. Gamez J. Minocycline for the treatment of amyotrophic lateral sclerosis: neuroprotector or neurotoxin? Reflections on another failure of translational medicine. Neurologia. 2008;23:484–493.[PubMed]
174. Benatar M. Lost in translation: treatment trials in the SOD1 mouse and in human ALS. Neurobiol Dis. 2007;26:1–13.  [PubMed]
175. van der Worp HB, Howells DW, Sena ES, Porritt MJ, Rewell S, O’Collins V, Macleod MR. Can animal models of disease reliably inform human studies? Plos Med. 2010;7:e1000245. [PMC free article][PubMed]
176. Ganesalingam J, Bowser R. The application of biomarkers in clinical trials for motor neuron disease. Biomark Med. 2010;4:281–297. [PMC free article]  [PubMed]
177. Wilson ME, Boumaza I, Lacomis D, Bowser R. Cystatin C: a candidate biomarker for amyotrophic lateral sclerosis. PLoS ONE. 2010;5:e15133. [PMC free article]  [PubMed]
178. Collins M, Riascos D, Kovalik T, An J, Krupa K, Hood BL, Conrads TP, Renton AE, Traynor BJ, Bowser R. The RNA-binding motif 45 (RBM45) protein accumulates in inclusion bodies in amyotrophic lateral sclerosis (ALS) and frontotemporal lobar degeneration with TDP-43 inclusions (FTLD-TDP) patients. Acta Neuropathol. 2012;124:717–732. [PMC free article]  [PubMed]
179. Weber M, Goldman B, Truniger S. Tetrahydrocannabinol (THC) for cramps in amyotrophic lateral sclerosis: a randomised, double-blind crossover trial. J Neurol Neurosurg Psychiatry. 2010;81:1135–1140.[PubMed]
pot logo blog2