Skip to main content

MINI REVIEW article

Front. Synaptic Neurosci., 06 November 2017
Volume 9 - 2017 | https://doi.org/10.3389/fnsyn.2017.00015

The Search for an Effective Therapy to Treat Fragile X Syndrome: Dream or Reality?

  • 1Université Côte d’Azur, CNRS, Institut de Pharmacologie Moléculaire et Cellulaire (IPMC), Valbonne, France
  • 2Université Côte d’Azur, INSERM, CNRS, Institut de Pharmacologie Moléculaire et Cellulaire (IPMC), Valbonne, France

Fragile X Syndrome (FXS) is the most common form of intellectual disability and a primary cause of autism. It originates from the lack of the Fragile X Mental Retardation Protein (FMRP), which is an RNA-binding protein encoded by the Fragile X Mental Retardation Gene 1 (FMR1) gene. Multiple roles have been attributed to this protein, ranging from RNA transport (from the nucleus to the cytoplasm, but also along neurites) to translational control of mRNAs. Over the last 20 years many studies have found a large number of FMRP mRNA targets, but it is still not clear which are those playing a critical role in the etiology of FXS. So far, no therapy for FXS has been found, making the quest for novel targets of considerable importance. Several pharmacological approaches have been attempted, but, despite some promising preclinical results, no strategy gave successful outcomes, due either to the induction of major side effects or to the lack of improvement of the phenotypes. However, these studies suggested that, in order to measure the effectiveness of a specific treatment, trials should be redesigned and new endpoints defined in FXS patients. Nevertheless, the search for new therapeutic targets for FXS is very active. In this context, the advances in animal modeling, coupled with better understanding of neurobiology and physiopathology of FXS, are of crucial importance in developing new selected treatments. Here, we discuss the pathways that were recently linked to the physiopathology of FXS (mGluR, GABAR, insulin, Insulin-like Growth Factor 1 (IGF-1), MPP-9, serotonin, oxytocin and endocannabinoid signaling) and that suggest new approaches to find an effective therapy for this disorder. Our goal with this review article is to summarize some recent relevant findings on FXS treatment strategies in order to have a clearer view of the different pathways analyzed to date emphasizing those shared with other synaptic disorders.

Introduction

Fragile X Syndrome

Fragile X syndrome (FXS) is the most common form of inherited intellectual disability (ID) and a primary genetic cause of autism. Due to its X-linked nature, it occurs in around 1:4000 males and 1:7000 females and one in three FXS patients display Autism Spectrum Disorders (ASD). Furthermore, the syndrome is characterized by delay in development and intellectual capacity, which impairs cognitive, executive and language performance. The severity of the symptoms can be very different among patients, and some of the most common features of the disease include repetitive behaviors, hyperactivity, anxiety, mood disorders and epilepsy (Hagerman et al., 2009). At the cellular level, there are abnormalities in neuronal maturation and pruning (Bassell and Warren, 2008; Scotto-Lomassese et al., 2011). In the brain of FXS patients as well as in FXS animal models (Fmr1-KO mouse and dFMR1-KO flies), dendritic spines have an altered morphology, and they appear longer, thinner and more tortuous than normal (Bassell and Warren, 2008). The morphological abnormalities are associated with deregulated synaptic plasticity: in mouse hippocampal mGluR dependent Long Term Depression (LTD) is increased (Huber et al., 2002), while Long Term Potentiation (LTP) is reduced in somatosensory and anterior cingulate cortex (Desai et al., 2006) as well as in amygdala (Suvrathan and Chattarji, 2011) and in the CA1 region (Lauterborn et al., 2007; Seese et al., 2012).

FXS arises from a mutation in a single gene called Fragile X Mental Retardation Gene 1 (FMR1). The 5′ UTR of FMR1 contains a CGG trinucleotide repeat that is polymorphic in the population. Once the repeats exceed 200 in number, methylation of the promoter is triggered, and this in turn causes the lack of expression of the gene and translation of its encoded protein, the Fragile X Mental Retardation Protein (FMRP; Bardoni et al., 2000). FMRP is an RNA-binding protein involved in different steps of mRNA metabolism, such as translational control (in soma and dendritic spines) and RNA transport (Maurin et al., 2014).

Therapeutic Strategies for FXS

Exploration of the physiopathology of FXS as well as the search for targets of FMRP have been very active over the last 20 year. These targets include mRNAs and proteins. Most proteins interacting with FMRP are RNA-binding protein components of FMRP-containing ribonucleoproteic particles. However, interactors of FMRP are also ion channels, molecular motors and proteins involved in cytoskeleton remodeling pathways (Menon et al., 2004; Bardoni et al., 2006; Davidovic et al., 2007; Abekhoukh and Bardoni, 2014; Maurin et al., 2014, 2015; Ferron, 2016; Abekhoukh et al., 2017; Bienkowski et al., 2017). This effort resulted in the identification of multiple putative targets for the treatment of FXS. However, so far no efficient therapy is available for this disorder (reviewed in Maurin et al., 2014). For instance, the imbalance between excitatory and inhibitory systems in FXS has been known for 20 years, and it has been thought to be the key target for therapy. Indeed, the first ever therapeutic strategies to be tested for FXS in clinical trials targeted the glutamatergic system (the excitatory pathway known to be up-regulated in FXS; Bear et al., 2004; Berry-Kravis et al., 2016) and the γ-aminobutyric acid (GABA) system (the inhibitory pathway also dysregulated in the disease; D’Antuono et al., 2003).

The mGluR theory regarding the pathophysiology of FXS states that the lack of FMRP hyperactivates the mGluR5-mediated pathway leading to the most prominent features of FXS (Bear et al., 2004). Even if the molecular reasons of this exaggerated activation are not completely clear, many efforts have been done based on this theory in order to develop an effective strategy for both pharmacological and genetic rescue through the inhibition of the mGluR pathway (Bear et al., 2004). In the mammalian FXS animal model (Fmr1-KO mouse), the treatment with specific antagonists of the mGluR5 resulted in the rescue of cellular (dendritic spine morphology), synaptic (exaggerated LTD) and behavioral defects, but these successes were not translated to the human treatments. Indeed, no beneficial effects of mavoglurant (AFQ056, a previous mGluR5 well-characterized antagonist) were observed in a 12-weeks, double-blind study involving a large cohort of adolescent and adult FXS patients (Bailey et al., 2016; Berry-Kravis et al., 2016). A similar new clinical trial was performed in a cohort of 183 FXS using basimglurant, a potent and selective mGluR5-negative allosteric modulator (NAM), already administrated as a treatment of depression (Quiroz et al., 2016). Again, according to the evaluation of the ADAMS score, no improvement in patients’ behavior was observed (Youssef et al., 2017).

GABA is the main inhibitory neurotransmitter of the Central Nervous System (CNS). Based on an altered expression of GABA receptor subunits in the absence of FMRP (Adusei et al., 2010) and a reduced production of GABA (Davidovic et al., 2011), it has been shown that an impairment of the GABAergic system is involved in FXS (D’Hulst and Kooy, 2007; Braat et al., 2015). Direct administration of GABA to the patients is not possible, partly because of its poor brain penetration. Thus, to restore the normal inhibition rate of the CNS and consequently reverse some major phenotype features of the disease, investigators have tried to treat FXS patients with GABA receptor agonists. So far, compounds like Acamprosate, Ganaxolone, Arbaclofen and Riluzole have been tested, but, despite the very mild side effects and good tolerability, they have caused very limited improvements in preliminary studies (Berry-Kravis et al., 2012; Erickson et al., 2014a,b; Ligsay et al., 2017).

Besides Glutamate and GABA, other deregulated pathways have been identified in FXS. Recent pre-clinical studies revealed that modulating other signaling pathways could ameliorate FXS symptoms in the mouse model of FXS. Our goal with this review is to summarize some recent relevant findings on FXS treatment strategies, focusing on promising molecular pathways that are altered not only in FXS but also in other forms of synaptic disorders that might lead to the discovery of a future treatment for neurodevelopmental diseases. In addition to those described here, other therapeutic targets for future treatment of FXS can be considered, for example, Amyloid Precursor Protein (APP), Brain-Derived Neurotrophic Factor (BDNF), cAMP and N-methyl-D-aspartate (NMDA) receptor as illustrated in several recent reviews (Wei et al., 2012; Castrén and Castrén, 2014; Androschuk et al., 2015; Tian et al., 2015; Westmark et al., 2016).

Insulin and Insulin-Like Growth Factor 1 Pathway

Transcriptomic analysis of cultured hippocampal neurons obtained from WT and Fmr1-KO embryos showed that at the top of the list of enriched gene expression pathways was the “Insulin signaling pathway” (Prilutsky et al., 2015). In dFMR1-KO flies insulin signaling is altered due to an elevated expression of Drosophila insulin-like peptide 2 (Dilp2) gene in the insulin-producing cells (IPCs) of the brain. Administration of metformin, an FDA-approved anti-diabetic drug, to dFMR1-KO flies (Viollet et al., 2012), leads to an amelioration of memory defects (Monyak et al., 2017). In this context it is interesting to notice that recent studies have clearly demonstrated the improvement of different in vitro/in vivo hallmarks in Fmr1-KO mice (Gantois et al., 2017) and in seven FXS patients (Dy et al., 2017) using metformin treatment. The precise mechanism of action of metformin has not been deciphered yet, but it has been shown that this drug inhibits the mitochondrial respiratory-chain, specifically at the complex 1 level, without affecting any other steps of the mitochondrial machinery. This leads to a reduction in proton-driven synthesis of ATP from ADP and inorganic phosphate. In addition, through AMPK-dependent and -independent regulation, metformin can lead to the inhibition of glucose production by disrupting gluconeogenesis gene expression (Viollet et al., 2012). In Fmr1-KO mice, metformin treatment inhibited the mTORC1 and ERK pathways, that resulted in the compensation of up-regulated translation, a hallmark of FXS (Maurin et al., 2014; Gantois et al., 2017). Indeed, in Fmr1-null neurons an increased level of phosphorylation of the serine/treonine kinase S6K1 has been observed. S6K1 is a common target of the mTORC1 and ERK pathways that are deregulated in the absence of FMRP (Sharma et al., 2010; Bhattacharya et al., 2012; Gross and Bassell, 2014; Gross et al., 2015a,b). It is interesting to notice here that two inhibitors of S6K1 (PF-4708671 and FS-115) were used in preclinical studies in mouse improving aberrant social interaction and behavioral inflexibility in Y-maze (for review see Gross and Bhattacharya, 2017). On the other side, also reducing the activation of the ERK pathway with lovastatin (a drug that inhibits the Ras-ERK1/2 activation by interfering with Ras recruitment to the membrane) or rimonabant (see endocannabinoid pathway) resulted in an improvement of Fmr1-KO cognition (Busquets-Garcia et al., 2013; Osterweil et al., 2013).

Obesity is often a co-morbid issue observed in FXS patients (Tounian et al., 1999; Raspa et al., 2010). Importantly, metformin has been used to treat seven obese FXS patients who showed improved cognition, language behavior along with obesity condition (Dy et al., 2017) proving to be an effective treatment for the FXS patients.

Insulin-like Growth Factor 1 (IGF-1) is a hormone primarily secreted by hepatocytes in response to Growth Hormone (GH) and, like insulin, promotes a decrease of glycaemia. IGF-1 promotes anabolic processes and tissue growth throughout life and it is a central factor for pathways involved in cell development and survival, proliferation and renewal. IGF-1 exerts its function by interacting with its receptor IGF receptor 1 (IGF1R; Costales and Kolevzon, 2016). In the CNS, IGF-1 plays a role in growth and development of all major CNS cell types and their synapse maturation. Imbalances in the IGF-1 pathway are associated to neuronal developmental impairment and, in particular, to ASD (Vahdatpour et al., 2016). Indeed, recombinant IGF-1, as well as some related compounds, have emerged as potential therapeutics to treat neurodevelopmental disorders and, indeed, clinical trials for ASD are in progress with these molecules (Wrigley et al., 2017).

In the mouse model of FXS, decreasing the levels of IGF1R corrects a number of phenotypic features (Deacon et al., 2015). Trofinetide is a neurotrophic peptide derived from IGF-1 that shows a long half-life and is well tolerated. A chronic treatment of Fmr1-KO mice with trofinetide corrected learning and memory deficits, hyperactivity and social interaction deficits displayed by these animals. At the microscopic level, abnormal dendritic spine density was rescued (Deacon et al., 2015). Considering this promising premise, a phase II clinical trial for trofinetide was performed and after only 28 days of treatment, improvements in higher sensory tolerance, reduced anxiety, better self-regulation and more social engagement were observed. No serious side effects were reported. Interestingly, Neuren Pharmaceuticals reported that trofinetide had significant clinical benefits in a Phase II clinical trial in 5–15 years old girls affected by Rett syndrome, another form of neurodevelopmental disorder characterized by ASD and intellectual disability (Bedogni et al., 2014)1.

Matrix Metalloproteinases Pathway

Matrix metalloproteinases (MMPs) are endopeptidases implicated in both physiological and pathological remodeling of tissues, and their activity is dependent on the zinc ion of their catalytic site. The MMPs family counts 25 members (22 of them found in humans), which can cleave both extracellular matrix (ECM) components and non-ECM elements. MMP-9 is a 92 kD collagenase involved in a broad spectrum of remodeling events of the ECM and plays a major proteolytic role in many cell types. Indeed, it acts during embryo implantation, cardiac tissue development and immune cell functioning (Yabluchanskiy et al., 2013). In the brain, MMP-9 controls synaptic plasticity, and thus learning and memory formation (Ganguly et al., 2013; Knapska et al., 2013).

Until recently, brain disorders associated to defects in MMP-9 has been linked uniquely to its involvement in inflammatory and immune responses. Nevertheless, a hyperactivation of MMP-9 in neurons may cause a massive degradation of the ECM surrounding them that could have severe consequences on function and maturation of synapses. So far, misregulated activation of this enzyme has been implicated in a number of neurodegenerative disorders, including traumatic brain injury, multiple sclerosis and Alzheimer’s disease but also in neurodevelopmental disorders (Reinhard et al., 2015).

The mRNA coding MMP-9 is a target of FMRP, which negatively modulates its expression. Indeed, in the absence of FMRP the expression levels and the activity of MMP-9 are increased. Interestingly this abnormal activity as well as the aberrant dendritic spines could be rescued by treating neuronal cultures with minocycline (Bilousova et al., 2009). This drug is an FDA-approved broad-spectrum antibiotic that shows two major effects: (1) it increases the phosphorylation of GluR1; (2) it promotes the membrane insertion of AMPA receptors (Imbesi et al., 2008). With its action, minocycline lowers the abnormally elevated levels of MMP-9 in FXS, and, when used in vivo on Fmr1-KO mice, reduces anxiety and reverses the deficit in ultrasonic vocalizations (Rotschafer et al., 2012). Genetic reduction of MMP-9 was obtained by crossing the viable MMP-9 KO mice with Fmr1-KO mice. Double KO mice lacked the typical major symptoms of FXS observed in Fmr1-KOs (Sidhu et al., 2014).

Until now, minocycline has been shown to be successful in two different clinical trials (Paribello et al., 2010; Leigh et al., 2013) and, importantly, treatment with mynocline resulted in the improvement of some event-related potentials compared with placebo (Schneider et al., 2013). However, the presence of side effects reduces the enthusiasm for the utilization of this molecule in clinic. It is worth to note that the relevance of MMP-9 levels for the physiopathology of FXS has been recently underlined by the rescue of several FXS related behaviors after a chronic treatment of Fmr1-KO with metformin that, interestingly, resulted in the reduction of MMP-9 levels in the Fmr1-KO mouse brain (Gantois et al., 2017).

Endocannabinoid Pathway

The endocannabinoid system (eCS) is represented by a group of neuromodulatory lipids and their receptors, notably the cannabinoid receptors 1 (CB1) and 2 (CB2). This system is present in mammalian tissues and, in particular, regulates the cardiovascular, nervous and immune systems. In the brain, the eCS is a key modulator of different neuronal aspects, including synaptic plasticity, cognition, anxiety, nociception and susceptibility to epileptic seizures (Khan et al., 2016). All these features also characterize the FXS patients’ phenotype. Indeed, the lack of FMRP has been associated with impaired functioning of the eCB pathway in glutamatergic synapses, and this identifies the endocannabinoid signaling complex as a possible therapeutic target for FXS (Jung et al., 2012). Rimonabant, a selective antagonist of the CB1 receptor, has been the first drug targeting the eCS used to attenuate the FXS symptoms (Busquets-Garcia et al., 2013). This drug was first developed for the treatment of obesity, but was withdrawn from the market because of its significant psychiatric side effects, such as depression, anxiety and suicidal thoughts. Nevertheless, the adverse effects only appear in patients that were given the highest administered doses. Notably, it was shown that the treatment with very low doses of rimonabant rescues synaptic plasticity in the hippocampus of Fmr1-KO mice and also learning and memory (Gomis-González et al., 2016). A new CB1 receptor neutral antagonist (NESS0327) was recently shown to have the same beneficial effects on Fmr1-KO mice behavior as rimonabant, supporting the CB1 receptor as a target to treat FXS (Gomis-González et al., 2016).

Deregulation of eCS could be involved in the physiopathology of ASD forms other than FXS. Indeed, 450 children born between 2006 and 2014 from mothers who, during their pregnancy, had taken valproic acid (VPA), an anti-epileptic and mood stabilizing drug showed neurobehavioral dysfunctions. This suggested that VPA plays an important role in developing ASD (Inspection Générale des Affaires Sociales (IGAS), 2016) and subsequently led to the generation of an environmental model of ASD by exposing pregnant rodents to VPA (Williams and Hersch, 1997; Williams et al., 2001). Recently, some social deficits displayed by two VPA-induced ASD rat models have been corrected by treatment with anandamide, an endocannabinoid positively stimulating the eCS pathway (Servadio et al., 2016). These findings further confirm that the eCS may be an interesting and important target for ASD therapy by using both agonist or antagonist approaches.

Serotonin Pathway

Hyperserotonemia was the first biomarker identified in patients affected by ASD. More recently, it has been shown that the levels of the amino acid tryptophan, the precursor of serotonin, is lower than normal in autistic brains, and that a diet poor in tryptophan worsens autistic symptoms (Boccuto et al., 2013). It was shown that polymorphisms in a gene encoding a 5-HT reuptake transporter protein cause lower synaptic serotonin availability and correlate with increased aggression and destructive behaviors (Hessl et al., 2008). Interestingly, decreased serotonin production was observed especially in young ASD children between 2 and 5 years of age, when the serotonin level should be at its maximal production (Chugani et al., 1999). Animal models in which genes involved in serotonin signaling have been inactivated display altered social interaction (Muller et al., 2016). Conversely, several mouse models of ASD, such as the 15q11-13 duplication and Smith-Laemli-Opitz syndrome models, to which we can also include the Fmr1-KO mouse, display altered 5-HT signaling (Muller et al., 2016). Indeed, it was shown that the stimulation of 5-HT7 serotonin receptors in post-synaptic compartments reverses mGluR-LTD in hippocampal slices of FXS mouse brains, suggesting that 5-HT7 receptor agonists might be envisaged as novel therapeutic tools for FXS (Costa et al., 2012). These same authors characterized two new molecules with very high binding affinity and selectivity for 5-HT7 receptors and ability to rescue exaggerated mGluR-LTD that might be used as novel pharmacological tools for the therapy of FXS (Costa et al., 2015).

The growing body of evidence linking ASD to abnormalities in serotonin function caused the use of the selective serotonin re-uptake inhibitors (SSRIs: citalopram, escitalopram, fluoxetine, fluvoxamine, and sertraline) to target various symptoms of the disorders. Most studies resulted in significant improvements in global functioning and in symptoms associated with anxiety and repetitive behaviors with mild side effects (Kolevzon et al., 2006). Due to the commonalities between ASD and FXS, low-dose sertraline was used to treat young children (12–50 months) affected by FXS. This drug, considered to be one of the most potent inhibitors of serotonin re-uptake, gave significant benefits in behavioral and cognitive features, especially in language skills (Winarni et al., 2012). More recently, a double-blind control trial was performed in 57 FXS patients aged between 2 years and 5 years using another serotonin re-uptake inhibitor (SSRI named Zoloft), an anti-depressant typically used in the treatment of depression, obsessive-compulsive disorder, panic disorder. Despite disappointing primary endpoint results, this treatment demonstrated a positive effect on cognition, visual reception score improving social interaction and early expressive language development (Greiss Hess et al., 2016), strongly suggesting serotonin re-uptake as a promising target to improve the FXS phenotype. Some selective serotonin (5-HT) re-uptake inhibitors seem to act through oxytocin release. Further, the administration of fenfluramine, a serotonergic agonist, to healthy subjects increases plasma oxytocin levels (Marazziti et al., 2012). It is interesting to note that the oxytocin signaling has been proposed as a target to treat FXS (see below). Thus, a better exploration of this cross-talk could result in new therapeutic approaches for FXS and ASD.

Oxytocin

Oxytocin is a neuropeptide that acts both as a hormone and as a neurotransmitter exerting pleiotropic effects in humans. It is well known to trigger labor, but also induces trust, empathy, and parental-infant relationships. It promotes social behavior and reduces stress and anxiety. In the last decade, it has been shown that intranasal administration of oxytocin is a potential treatment that improves social communication skills in various disorders. Some promising studies in animal models (Meyer-Lindenberg, 2008) paved the way for clinical trials focusing on the treatment of impaired social skills in a variety of conditions, including ASD and schizophrenia. Recent studies suggest that intranasal administration of oxytocin can ameliorate some symptoms of FXS, showing anxiolytic and pro-social qualities (Hall et al., 2012). Ben-Ari et al. (1989) shed light on the molecular mechanisms of the benefits of oxytocin treatment. Oxytocin plays a key role in regulating the effect of GABA on neuron activity. GABA is mostly known as an inhibitory neurotransmitter that acts on a receptor channel complex permeable to chloride anions. These anions flow through the channel according to their electrochemical gradient across the plasma membrane. It is noteworthy that the net value of this gradient changes along the development of the brain, depending on the expression of two major chloride co-transporters (KCC2 and NKCC1) that consequently modify the effects of GABA stimulation (Ben-Ari et al., 1989). A critical period for this shift occurs perinatally and the excitatory-to-inhibitory change of GABA effect is actually mediated by oxytocin receptors (Tyzio et al., 2006).

These authors further reported that the oxytocin-mediated GABA excitatory-inhibitory shift during delivery is abolished in the VPA-treated and Fmr1-KO mice, both rodent models of ASD. Consistently, blocking oxytocin signaling in naïve mothers resulted in the production of offspring with electrophysiological and behavioral autistic-like features (Tyzio et al., 2014). These authors also showed that during delivery, both Fmr1-KO and VPA-treated mice have elevated intracellular chloride levels in hippocampal neurons. These elevated chloride levels may explain the paradoxical effects of benzodiazapines and phenobarbital that occur through GABA modulation in patients with ASD. Interestingly, the maternal pretreatment with bumetanide, a diuretic belonging to the sulfamyl category that blocks the chloride co-transporter NKCC1, restored electrophysiological and behavioral phenotypes in the offspring (Tyzio et al., 2014). Very recently, the same authors performed a clinical trial and remarkably showed that bumetanide improves the core symptoms of ASD and presents a favorable benefit/risk ratio particularly at 1.0 mg twice a day (Lemonnier et al., 2017). Collectively, these studies strongly indicate that the oxytocin pathway is involved in ASD physiopathology and that bumetanide is a promising treatment for various forms of ASD, including FXS.

Concluding Remarks

The intense efforts to unravel the physiopathology of FXS appear to have produced some relevant pharmacological targets to treat this disorder. It is interesting to underline that most of the FXS deregulated pathways have been found to be unbalanced also in other forms of ASD or ID-associated diseases. This supports the relevance of these pathways in the physiopathology of ASD and/or intellectual disability. The success of preclinical treatments in mouse, rat or fly FXS models should be considered an exceptionally important result. However, we have learned from past clinical trials that to evaluate the effectiveness of therapies in humans, the design of clinical trials as well as the definition of disease-specific endpoints are of critical importance. Indeed, even on the basis of excellent pre-clinical results, these same preclinical targets have not been translated to therapies to improve behavior and cognition of FXS patients. It is clear that the trials could benefit from the analysis of previous trials results by performing, for instance, longer treatments (Berry-Kravis et al., 2016; Erickson et al., 2017). By further understanding the molecular deregulations in FXS, it should be possible to combine two (or more) drugs targeting different altered pathways, as proposed by the Willemsen laboratory (Zeidler et al., 2015), or treating patients at different ages with different targeted treatments, as we have recently proposed (Bardoni et al., 2017). The positive impact of metformin on FXS behavior troubles underlines the importance to focus on repositioning existing drugs to find new targeted treatments for FXS, as well as for other neurodevelopmental disorders. Lastly, the fact that some clinical trials worked only in small sets of patients (Berry-Kravis et al., 2016), suggests the importance of stratification of FXS patients on the basis of their multiple phenotypes for “ad hoc” therapies.

Furthermore, according to Budimirovic et al. (2017), 22 double-blind controlled clinical trials in FXS have been finalized between 2008 and 2015. The accurate analysis of these studies led the authors to the conclusion that the readouts employed to evaluate the outcome of treatments were in general of moderate/poor quality (Budimirovic et al., 2017). In this context, the search for specific and easily measurable biomarkers for FXS should be encouraged. Even if efforts are in progress concerning blood-based and neurophysiological measures (Ethridge et al., 2016; Ray et al., 2016; AlOlaby et al., 2017; Pellerin et al., 2017; Wang et al., 2017), it would be interesting to develop cell-based biomarkers improving analysis of FXS iPS cell lines, while parameters analyzed so far have highlighted the great heterogeneity of these cells (for review see Khalfallah et al., 2017). Also in this case, more detailed studies could result into personalized treatments or treatments concerning subsets of patients. We can underline that only a few examples of FXS cell lines exist that can be used to investigate preclinical treatment in cellulo or screening with chemical libraries (Castets et al., 2005; Khalfallah et al., 2017). The availability of such tools could accelerate the definition of new pharmacological approaches identified by the dissection of altered pathways in FMRP-null brains. This dissection can be realized by the analysis of FMRP targets (mRNAs and proteins) and the FXS-translatome in different brain regions and/or neuron subtypes at different ages of neurodevelopment upon different conditions (e.g., stress and learning) or upon various stimuli.

All these considerations indicate that research on FXS has still some stimulating areas to investigate in order to define new treatments. In addition, the recent promising studies presented here suggest the conclusion that treatments for all FXS patients will be available in a near future and will not remain only a dream for patients’ families and researchers in the field.

Author Contributions

SC, BB and TM wrote the manuscript.

Conflict of Interest Statement

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

Acknowledgments

We thank Enzo Lalli and Maria Capovilla for critical reading of the manuscript. This study was supported by: INSERM; CNRS; CNRS LIA “NEOGENEX”; ANR-11-LABX-0028-01, Monaco Against Autism Foundation, ANR-12-BSV4-0020, and Fondation pour la Recherche Médicale (FRM) DEQ20140329490 to BB; FRAXA Foundation to TM. SC is recipient of an international PhD fellowship “Signalife- LabEx” Program.

Footnotes

  1. ^ https://www.fraxa.org/tag/trofinetide/

References

Abekhoukh, S., and Bardoni, B. (2014). CYFIP family proteins between autism and intellectual disability: links with Fragile X syndrome. Front. Cell. Neurosci. 8:81. doi: 10.3389/fncel.2014.00081

PubMed Abstract | CrossRef Full Text | Google Scholar

Abekhoukh, S., Sahin, H. B., Grossi, M., Zongaro, S., Maurin, T., Madrigal, I., et al. (2017). New insights into the regulatory function of CYFIP1 in the context of WAVE- and FMRP-containing complexes. Dis. Model. Mech. 10, 463–474. doi: 10.1242/dmm.025809

PubMed Abstract | CrossRef Full Text | Google Scholar

Adusei, D. C., Pacey, L. K., Chen, D., and Hampson, D. R. (2010). Early developmental alterations in GABAergic protein expression in Fragile X knockout mice. Neuropharmacology 59, 167–171. doi: 10.1016/j.neuropharm.2010.05.002

PubMed Abstract | CrossRef Full Text | Google Scholar

AlOlaby, R. R., Sweha, S. R., Silva, M., Durbin-Johnson, B., Yrigollen, C. M., Pretto, D., et al. (2017). Molecular biomarkers predictive of sertraline treatment response in young children with Fragile X syndrome. Brain Dev. 39, 483–492. doi: 10.1016/j.braindev.2017.01.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Androschuk, A., Al-Jabri, B., and Bolduc, F. V. (2015). From learning to memory: what flies can tell us about intellectual disability treatment. Front. Psychiatry 6:85. doi: 10.3389/fpsyt.2015.00085

PubMed Abstract | CrossRef Full Text | Google Scholar

Bailey, D. B. Jr., Berry-Kravis, E., Wheeler, A., Raspa, M., Merrien, F., Ricart, J., et al. (2016). Mavoglurant in adolescents with Fragile X syndrome: analysis of Clinical Global Impression-Improvement source data from a double-blind therapeutic study followed by an open-label, long-term extension study. J. Neurodev. Disord. 8:1. doi: 10.1186/s11689-015-9134-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Bardoni, B., Capovilla, M., and Lalli, E. (2017). Modeling Fragile X syndrome in neurogenesis: an unexpected phenotype and a novel tool for future therapies. Neurogenesis 4:e1270384. doi: 10.1080/23262133.2016.1270384

PubMed Abstract | CrossRef Full Text | Google Scholar

Bardoni, B., Davidovic, L., Bensaid, M., and Khandjian, E. W. (2006). The fragile X syndrome: exploring its molecular basis and seeking a treatment. Expert Rev. Mol. Med. 8, 1–16. doi: 10.1017/s1462399406010751

PubMed Abstract | CrossRef Full Text | Google Scholar

Bardoni, B., Mandel, J. L., and Fisch, G. S. (2000). FMR1 gene and fragile X syndrome. Am. J. Med. Genet. 97, 153–163. doi: 10.1002/1096-8628(200022)97:2<153::aid-ajmg7>3.0.co;2-m

PubMed Abstract | CrossRef Full Text | Google Scholar

Bassell, G. J., and Warren, S. T. (2008). Fragile X syndrome: loss of local mRNA regulation alters synaptic development and function. Neuron 60, 201–214. doi: 10.1016/j.neuron.2008.10.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Bear, M. F., Huber, K. M., and Warren, S. T. (2004). The mGluR theory of fragile X mental retardation. Trends Neurosci. 27, 370–377. doi: 10.1016/j.tins.2004.04.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Bedogni, F., Rossi, R. L., Galli, F., Cobolli Gigli, C., Gandaglia, A., Kilstrup-Nielsen, C., et al. (2014). Rett syndrome and the urge of novel approaches to study MeCP2 functions and mechanisms of action. Neurosci. Biobehav. Rev. 46, 187–201. doi: 10.1016/j.neubiorev.2014.01.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Ben-Ari, Y., Cherubini, E., Corradetti, R., and Gaïarsa, J. L. (1989). Giant synaptic potentials in immature rat CA3 hippocampal neurones. J. Physiol. 416, 303–325. doi: 10.1113/jphysiol.1989.sp017762

PubMed Abstract | CrossRef Full Text | Google Scholar

Berry-Kravis, E., Des Portes, V., Hagerman, R., Jacquemont, S., Charles, P., Visootsak, J., et al. (2016). Mavoglurant in fragile X syndrome: results of two randomized, double-blind, placebo-controlled trials. Sci. Transl. Med. 8:321ra5. doi: 10.1126/scitranslmed.aab4109

PubMed Abstract | CrossRef Full Text | Google Scholar

Berry-Kravis, E. M., Hessl, D., Rathmell, B., Zarevics, P., Cherubini, M., Walton-Bowen, K., et al. (2012). Effects of STX209 (arbaclofen) on neurobehavioral function in children and adults with fragile X syndrome: a randomized, controlled, phase 2 trial. Sci. Transl. Med. 4:152ra127. doi: 10.1126/scitranslmed.3004214

PubMed Abstract | CrossRef Full Text | Google Scholar

Bhattacharya, A., Kaphzan, H., Alvarez-Dieppa, A. C., Murphy, J. P., Pierre, P., and Klann, E. (2012). Genetic removal of p70 S6 kinase 1 corrects molecular, synaptic, and behavioral phenotypes in fragile X syndrome mice. Neuron 76, 325–337. doi: 10.1016/j.neuron.2012.07.022

PubMed Abstract | CrossRef Full Text | Google Scholar

Bienkowski, R. S., Banerjee, A., Rounds, J. C., Rha, J., Omotade, O. F., Gross, C., et al. (2017). The conserved, disease-associated RNA binding protein dNab2 interacts with the fragile X protein ortholog in drosophila neurons. Cell Rep. 20, 1372–1384. doi: 10.1016/j.celrep.2017.07.038

PubMed Abstract | CrossRef Full Text | Google Scholar

Bilousova, T. V., Dansie, L., Ngo, M., Aye, J., Charles, J. R., Ethell, D. W., et al. (2009). Minocycline promotes dendritic spine maturation and improves behavioral performance in the fragile X mouse model. J. Med. Genet. 46, 94–102. doi: 10.1136/jmg.2008.061796

PubMed Abstract | CrossRef Full Text | Google Scholar

Boccuto, L., Chen, C. F., Pittman, A. R., Skinner, C. D., McCartney, H. J., Jones, K., et al. (2013). Decreased tryptophan metabolism in patients with autism spectrum disorders. Mol. Autism 4:16. doi: 10.1186/2040-2392-4-16

PubMed Abstract | CrossRef Full Text | Google Scholar

Braat, S., D’Hulst, C., Heulens, I., De Rubeis, S., Mientjes, E., Nelson, D. L., et al. (2015). The GABAA receptor is an FMRP target with therapeutic potential in fragile X syndrome. Cell Cycle 14, 2985–2995. doi: 10.4161/15384101.2014.989114

PubMed Abstract | CrossRef Full Text | Google Scholar

Budimirovic, D. B., Berry-Kravis, E., Erickson, C. A., Hall, S. S., Hessl, D., Reiss, A. L., et al. (2017). Updated report on tools to measure outcomes of clinical trials in fragile X syndrome. J. Neurodev. Disord. 9:14. doi: 10.1186/s11689-017-9193-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Busquets-Garcia, A., Gomis-Gonzàlez, M., Guegan, T., Augustìn-Pavòn, C., Pastor, A., Mato, S., et al. (2013). Targeting the endocannabinoid system in the treatment of Fragile X syndrome. Nat. Med. 19, 603–607. doi: 10.1038/nm.3127

PubMed Abstract | CrossRef Full Text | Google Scholar

Castets, M., Schaeffer, C., Bechara, E., Schenck, A., Khandjian, E. W., Luche, S., et al. (2005). FMRP interferes with the Rac1 pathway and controls actin cytoskeleton dynamics in murine fibroblasts. Hum. Mol. Genet. 14, 835–844. doi: 10.1093/hmg/ddi077

PubMed Abstract | CrossRef Full Text | Google Scholar

Castrén, M. L., and Castrén, E. (2014). BDNF in fragile X syndrome. Neuropharmacology 76, 729–736. doi: 10.1016/j.neuropharm.2013.05.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Chugani, D. C., Muzik, O., Behen, M., Rothermel, R., Janisse, J. J., Lee, J., et al. (1999). Developmental changes in brain serotonin synthesis capacity in autistic and nonautistic children. Ann. Neurol. 45, 287–295. doi: 10.1002/1531-8249(199903)45:3<287::aid-ana3>3.0.co;2-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Costa, L., Sardone, L. M., Lacivita, E., Leopoldo, M., and Ciranna, L. (2015). Novel agonists for serotonin 5-HT7 receptors reverse metabotropic glutamate receptor-mediated long-term depression in the hippocampus of wild-type and Fmr1 KO mice, a model of Fragile X Syndrome. Front. Behav. Neurosci. 9:65. doi: 10.3389/fnbeh.2015.00065

PubMed Abstract | CrossRef Full Text | Google Scholar

Costa, L., Spatuzza, M., D’Antoni, S., Bonaccorso, C. M., Trovato, C., Musumeci, S. A., et al. (2012). Activation of 5-HT7 serotonin receptors reverses metabotropic glutamate receptor- mediated synaptic plasticity in wild-type and Fmr1 knockout mice, a model of fragile X syndrome. Biol. Psychiatry 72, 924–933. doi: 10.1016/j.biopsych.2012.06.008

PubMed Abstract | CrossRef Full Text | Google Scholar

Costales, J., and Kolevzon, A. (2016). The therapeutic potential of insulin-like growth factor-1 in central nervous system disorders. Neurosci. Biobehav. Rev. 63, 207–222. doi: 10.1016/j.neubiorev.2016.01.001

PubMed Abstract | CrossRef Full Text | Google Scholar

D’Antuono, M., Merlo, D., and Avoli, M. (2003). Involvement of cholinergic and gabaergic systems in the fragile X knockout mice. Neuroscience 119, 9–13. doi: 10.1016/s0306-4522(03)00103-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Davidovic, L., Jaglin, X. H., Lepagnol-Bestel, A. M., Tremblay, S., Simonneau, M., Bardoni, B., et al. (2007). The fragile X mental retardation protein is a molecular adaptor between the neurospecific KIF3C kinesin and dendritic RNA granules. Hum. Mol. Genet. 16, 3047–3058. doi: 10.1093/hmg/ddm263

PubMed Abstract | CrossRef Full Text | Google Scholar

Davidovic, L., Navratil, V., Bonaccorso, C. M., Catania, M. V., Bardoni, B., and Dumas, M. E. (2011). A metabolomic and systems biology perspective on the brain of the fragile X syndrome mouse model. Genome Res. 21, 2190–2202. doi: 10.1101/gr.116764.110

PubMed Abstract | CrossRef Full Text | Google Scholar

Deacon, R. M., Glass, L., Snape, M., Hurley, M. J., Altimiras, F. J., Biekofsky, R. R., et al. (2015). NNZ-2566, a novel analog of (1–3) IGF-1, as a potential therapeutic agent for fragile X syndrome. Neuromolecular Med. 17, 71–82. doi: 10.1007/s12017-015-8341-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Desai, N. S., Casimiro, T. M., Gruber, S. M., and Vanderklish, P. W. (2006). Early postnatal plasticity in neocortex of Fmr1 knockout mice. J. Neurophysiol. 96, 1734–1745. doi: 10.1152/jn.00221.2006

PubMed Abstract | CrossRef Full Text | Google Scholar

D’Hulst, C., and Kooy, R. F. (2007). The GABAA receptor: a novel target for treatment of fragile X? Trends Neurosci. 30, 425–431. doi: 10.1016/j.tins.2007.06.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Dy, A. B. C., Tassone, F., Eldeeb, M., Salcedo-Arellano, M. J., Tartaglia, N., and Hagerman, R. (2017). Metformin as targeted treatment in fragile X syndrome. Clin. Genet. doi: 10.1111/cge.13039 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Erickson, C. A., Davenport, M. H., Schaefer, T. L., Wink, L. K., Pedapati, E. V., Sweeney, J. A., et al. (2017). Fragile X targeted pharmacotherapy: lessons learned and future directions. J. Neurodev. Dis. 9:7. doi: 10.1186/s11689-017-9186-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Erickson, C. A., Ray, B., Maloney, B., Wink, L. K., Bowers, K., Schaefer, T. L., et al. (2014a). Impact of acamprosate on plasma amyloid-β precursor protein in youth: a pilot analysis in fragile X syndrome-associated and idiopathic autism spectrum disorder suggests a pharmacodynamic protein marker. J. Psychiatr. Res. 59, 220–228. doi: 10.1016/j.jpsychires.2014.07.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Erickson, C. A., Veenstra-VanderWeele, J. M., Melmed, R. D., McCracken, J. T., Ginsberg, L. D., Sikich, L., et al. (2014b). STX209 (arbaclofen) for autism spectrum disorders: an 8-week open-label study. J. Autism Dev. Disord. 44, 958–964. doi: 10.1007/s10803-013-1963-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Ethridge, L. E., White, S. P., Mosconi, M. W., Wang, J., Byerly, M. J., and Sweeney, J. A. (2016). Reduced habituation of auditory evoked potentials indicate cortical hyper-excitability in Fragile X Syndrome. Transl. Psychiatry 6:e787. doi: 10.1038/tp.2016.48

PubMed Abstract | CrossRef Full Text | Google Scholar

Ferron, L. (2016). Fragile X mental retardation protein controls ion channel expression and activity. J. Physiol. 594, 5861–5867. doi: 10.1113/JP270675

PubMed Abstract | CrossRef Full Text | Google Scholar

Ganguly, K., Rejmak, E., Mikosz, M., Nikolaev, E., Knapska, E., and Kaczmarek, L. (2013). Matrix metalloproteinase (MMP) 9 transcription in mouse brain induced by fear learning. J. Biol. Chem. 288, 20978–20991. doi: 10.1074/jbc.M113.457903

PubMed Abstract | CrossRef Full Text | Google Scholar

Gantois, I., Khoutorsky, A., Popic, J., Aguilar-Valles, A., Freemantle, E., Cao, R., et al. (2017). Metformin ameliorates core deficits in a mouse model of fragile X syndrome. Nat. Med. 23, 674–677. doi: 10.1038/nm.4335

PubMed Abstract | CrossRef Full Text | Google Scholar

Gomis-González, M., Busquets-Garcia, A., Matute, C., Maldonado, R., Mato, S., and Ozaita, A. (2016). Possible therapeutic doses of cannabinoid type 1 receptor antagonist reverses key alterations in Fragile X syndrome mouse model. Genes 7:E56. doi: 10.3390/genes7090056

PubMed Abstract | CrossRef Full Text | Google Scholar

Greiss Hess, L., Fitzpatrick, S. E., Nguyen, D. V., Chen, Y., Gaul, K. N., Schneider, A., et al. (2016). A randomized, double-blind, placebo-controlled trial of low-dose sertraline in young children with fragile X syndrome. J. Dev. Behav. Ped. 37, 619–628. doi: 10.1097/DBP.0000000000000334

PubMed Abstract | CrossRef Full Text | Google Scholar

Gross, C., and Bassell, G. J. (2014). Neuron-specific regulation of class I PI3K catalytic subunits and their dysfunction in brain disorders. Front. Mol. Neurosci. 7:12. doi: 10.3389/fnmol.2014.00012

PubMed Abstract | CrossRef Full Text | Google Scholar

Gross, C., and Bhattacharya, A. (2017). “Intracellular signaling network in fragile X syndrome: approaches to drug discovery and therapeutics,” in Fragile X Syndrome—From Genetics to Targeted Treatments, eds R. Willemsen and F. Kooy (Amsterdam: Elsevier), 498.

Gross, C., Chang, C. W., Kelly, S. M., Bhattacharya, A., McBride, S. M., Danielson, S. W., et al. (2015a). Increased expression of the PI3K enhancer PIKE mediates deficits in synaptic plasticity and behavior in fragile X syndrome. Cell Rep. 11, 727–736. doi: 10.1016/j.celrep.2015.03.060

PubMed Abstract | CrossRef Full Text | Google Scholar

Gross, C., Raj, N., Molinaro, G., Allen, A. G., Whyte, A. J., Gibson, J. R., et al. (2015b). Selective role of the catalytic PI3K subunit p110β in impaired higher order cognition in fragile X syndrome. Cell Rep. 11, 681–688. doi: 10.1016/j.celrep.2015.03.065

PubMed Abstract | CrossRef Full Text | Google Scholar

Hagerman, R. J., Berry-Kravis, E., Kaufmann, W. E., Ono, M. Y., Tartaglia, N., Lachiewicz, A., et al. (2009). Advances in the treatment of fragile X syndrome. Pediatrics 123, 378–390. doi: 10.1542/peds.2008-0317

PubMed Abstract | CrossRef Full Text | Google Scholar

Hall, S. S., Lightbody, A. A., McCarty, B. E., Parker, K. J., and Reiss, A. L. (2012). Effects of intranasal oxytocin on social anxiety in males with fragile X syndrome. Psychoneuroendocrinology 37, 509–518. doi: 10.1016/j.psyneuen.2011.07.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Hessl, D., Tassone, F., Cordeiro, L., Koldewyn, K., McCormick, C., Green, C., et al. (2008). Brief report: aggression and stereotypic behavior in males with fragile x syndrome—moderating secondary genes in a “single gene” disorder. J. Autism Dev. Disord. 38, 184–189. doi: 10.1007/s10803-007-0365-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Huber, K. M., Gallagher, S. M., Warren, S. T., and Bear, M. F. (2002). Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proc. Natl. Acad. Sci. U S A 99, 7746–7750. doi: 10.1073/pnas.122205699

PubMed Abstract | CrossRef Full Text | Google Scholar

Imbesi, M., Uz, T., Manev, R., Sharma, R. P., and Manev, H. (2008). Minocycline increases phosphorylation and membrane insertion of neuronal GluR1 receptors. Neurosci. Lett. 447, 134–137. doi: 10.1016/j.neulet.2008.10.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Inspection Générale des Affaires Sociales (IGAS). (2016). Enquête Relative aux Spécialités Pharmaceutiques Contenant du Valproate de Sodium. France: IGAS.

Jung, K. M., Sepers, M., Henstridge, C. M., Lassalle, O., Neuhofer, D., Martin, H., et al. (2012). Uncoupling of the endocannabinoid signalling complex in a mouse model of fragile X syndrome. Nat. Commun. 3:1080. doi: 10.1038/ncomms2045

PubMed Abstract | CrossRef Full Text | Google Scholar

Khalfallah, O., Jarjat, M., Davidovic, L., Nottet, N., Cestèle, S., Mantegazza, M., et al. (2017). Depletion of the fragile X mental retardation protein in embryonic stem cells alters the kinetics of neurogenesis. Stem Cells 35, 374–385. doi: 10.1002/stem.2505

PubMed Abstract | CrossRef Full Text | Google Scholar

Khan, M. I., Sobocińska, A. A., Czarnecka, A. M., Król, M., Botta, B., and Szczylik, C. (2016). The therapeutic aspects of the endocannabinoid system (ECS) for cancer and their development: from nature to laboratory. Curr. Pharm. Des. 22, 1756–1766. doi: 10.2174/1381612822666151211094901

PubMed Abstract | CrossRef Full Text | Google Scholar

Knapska, E., Lioudyno, V., Kiryk, A., Mikosz, M., Górkiewicz, T., Michaluk, P., et al. (2013). Reward learning requires activity of matrix metalloproteinase-9 in the central amygdala. J. Neurosci. 33, 14591–14600. doi: 10.1523/JNEUROSCI.5239-12.2013

PubMed Abstract | CrossRef Full Text | Google Scholar

Kolevzon, A., Mathewson, K. A., and Hollander, E. (2006). Selective serotonin reuptake inhibitors in autism: a review of efficacy and tolerability. J. Clin. Psychiatry 67, 407–414. doi: 10.4088/jcp.v67n0311

PubMed Abstract | CrossRef Full Text | Google Scholar

Lauterborn, J. C., Rex, C. S., Kramár, E., Chen, L. Y., Pandyarajan, V., Lynch, G., et al. (2007). Brain-derived neurotrophic factor rescues synaptic plasticity in a mouse model of fragile X syndrome. J. Neurosci. 27, 10685–10694. doi: 10.1523/JNEUROSCI.2624-07.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Leigh, M. J., Nguyen, D. V., Mu, Y., Winarni, T. I., Schneider, A., Chechi, T., et al. (2013). A randomized double-blind, placebo-controlled trial of minocycline in children and adolescents with fragile X syndrome. J. Dev. Behav. Pediatr. 34, 147–155. doi: 10.1097/DBP.0b013e318287cd17

PubMed Abstract | CrossRef Full Text | Google Scholar

Lemonnier, E., Villeneuve, N., Sonie, S., Serret, S., Rosier, A., Roue, M., et al. (2017). Effects of bumetanide on neurobehavioral function in children and adolescents with autism spectrum disorders. Transl. Psychiatry 7:e1124. doi: 10.1038/tp.2017.101

PubMed Abstract | CrossRef Full Text | Google Scholar

Ligsay, A., Van Dijck, A., Nguyen, D. V., Lozano, R., Chen, Y., Bickel, E. S., et al. (2017). A randomized double-blind, placebo-controlled trial of ganaxolone in children and adolescents with fragile X syndrome. J. Neurodev. Disord. 9:26. doi: 10.1186/s11689-017-9207-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Marazziti, D., Baroni, S., Giannaccini, G., Betti, L., Massimetti, G., Carmassi, C., et al. (2012). A link between oxytocin and serotonin in humans: supporting evidence from peripheral markers. Eur. Neuropsychopharmacol. 22, 578–583. doi: 10.1016/j.euroneuro.2011.12.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Maurin, T., Melko, M., Abekhoukh, S., Khalfallah, O., Davidovic, L., Jarjat, M., et al. (2015). The FMRP/GRK4 mRNA interaction uncovers a new mode of binding of the Fragile X mental retardation protein in cerebellum. Nucleic Acids Res. 43, 8540–8550. doi: 10.1093/nar/gkv801

PubMed Abstract | CrossRef Full Text | Google Scholar

Maurin, T., Zongaro, S., and Bardoni, B. (2014). Fragile X Syndrome: from molecular pathology to therapy. Neurosci. Biobehav. Rev. 46, 242–255. doi: 10.1016/j.neubiorev.2014.01.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Menon, R. P., Gibson, T. J., and Pastore, A. (2004). The C terminus of fragile X mental retardation protein interacts with the multi-domain Ran-binding protein in the microtubule-organising centre. J. Mol. Biol. 343, 43–53. doi: 10.1016/j.jmb.2004.08.024

PubMed Abstract | CrossRef Full Text | Google Scholar

Meyer-Lindenberg, A. (2008). Impact of prosocial neuropeptides on human brain function. Prog. Brain Res. 170, 463–470. doi: 10.1016/S0079-6123(08)00436-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Monyak, R. E., Emerson, D., Schoenfeld, B. P., Zheng, X., Chambers, D. B., Rosenfelt, C., et al. (2017). Insulin signaling misregulation underlies circadian and cognitive deficits in a Drosophila Fragile X model. Mol. Psychiatry 22, 1140–1148. doi: 10.1038/mp.2016.51

PubMed Abstract | CrossRef Full Text | Google Scholar

Muller, C. L., Anacker, A. M., and Veenstra-VanderWeele, J. (2016). The serotonin system in autism spectrum disorder: from biomarker to animal models. Neuroscience 321, 24–41. doi: 10.1016/j.neuroscience.2015.11.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Osterweil, E. K., Chuang, S. C., Chubykin, A. A., Sidorov, M., Bianchi, R., Wong, R. K. S., et al. (2013). Lovastatin corrects excess protein synthesis and prevents epileptogenesis in a mouse model of Fragile X syndrome. Neuron 77, 243–250. doi: 10.1016/j.neuron.2012.01.034

PubMed Abstract | CrossRef Full Text | Google Scholar

Paribello, C., Tao, L., Folino, A., Berry-Kravis, E., Tranfaglia, M., Ethell, I. M., et al. (2010). Open-label add-on treatment trial of minocycline in fragile X syndrome. BMC Neurol. 10:91. doi: 10.1186/1471-2377-10-91

PubMed Abstract | CrossRef Full Text | Google Scholar

Pellerin, D., Lortie, A., and Corbin, F. (2017). Platelets as a surrogate disease model of neurodevelopmental disorders: insights from Fragile X syndrome. Platelets doi: 10.1080/09537104.2017.1317733 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Prilutsky, D., Kho, A. T., Palmer, N. P., Bhakar, A. L., Smedemark-Margulies, N., Kong, S. W., et al. (2015). Gene expression analysis in Fmr1KO mice identifies an immunological signature in brain tissue and mGluR5-related signaling in primary neuronal cultures. Mol. Autism 6:66. doi: 10.1186/s13229-015-0061-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Quiroz, J. A., Tamburri, P., Deptula, D., Banken, L., Beyer, U., Rabbia, M., et al. (2016). Efficacy and safety of basimglurant as adjunctive therapy for major depression: a randomized clinical trial. JAMA Psychiatry. 73, 675–684. doi: 10.1001/jamapsychiatry.2016.0838

PubMed Abstract | CrossRef Full Text | Google Scholar

Raspa, M., Bailey, D. B., Bishop, E., Holiday, D., and Olmsted, M. (2010). Obesity, food selectivity, and physical activity in individuals with fragile X syndrome. Am. J. Intellect. Dev. Disabil. 115, 482–495. doi: 10.1352/1944-7558-115.6.482

PubMed Abstract | CrossRef Full Text | Google Scholar

Ray, B., Sokol, D. K., Maloney, B., and Lahiri, D. K. (2016). Finding novel distinctions between the sAPPα-mediated anabolic biochemical pathways in autism spectrum disorder and Fragile X syndrome plasma and brain tissue. Sci. Rep. 6:26052. doi: 10.1038/srep26052

PubMed Abstract | CrossRef Full Text | Google Scholar

Reinhard, S. M., Razak, K., and Ethell, I. M. (2015). A delicate balance: role of MMP-9 in brain development and pathophysiology of neurodevelopmental disorders. Front. Cell. Neurosci. 9:280. doi: 10.3389/fncel.2015.00280

PubMed Abstract | CrossRef Full Text | Google Scholar

Rotschafer, S. E., Trujillo, M. S., Dansie, L. E., Ethell, I. M., and Razak, K. A. (2012). Minocycline treatment reverses ultrasonic vocalization production deficit in a mouse model of Fragile X Syndrome. Brain Res. 1439, 7–14. doi: 10.1016/j.brainres.2011.12.041

PubMed Abstract | CrossRef Full Text | Google Scholar

Schneider, A., Leigh, M. J., Adams, P., Nanakul, R., Chechi, T., Olichney, J., et al. (2013). Electrocortical changes associated with minocycline treatment in fragile X syndrome. J. Psychopharmacol. 27, 956–963. doi: 10.1177/0269881113494105

PubMed Abstract | CrossRef Full Text | Google Scholar

Scotto-Lomassese, S., Nissant, A., Mota, T., Néant-Féry, M., Oostra, B. A., Greer, C. A., et al. (2011). Fragile X mental retardation protein regulates new neuron differentiation in the adult olfactory bulb. J. Neurosci. 31, 2205–2215. doi: 10.1523/jneurosci.5514-10.2011

PubMed Abstract | CrossRef Full Text | Google Scholar

Seese, R. R., Babayan, A. H., Katz, A. M., Cox, C. D., Lauterborn, J. C., Lynch, G., et al. (2012). LTP induction translocates cortactin at distant synapses in wild type but not Fmr1 knock-out mice. J. Neurosci. 32, 7403–7413. doi: 10.1523/jneurosci.0968-12.2012

PubMed Abstract | CrossRef Full Text | Google Scholar

Servadio, M., Melancia, F., Manduca, A., di Masi, A., Schiavi, S., Cartocci, V., et al. (2016). Targeting anandamide metabolism rescues core and associated autistic-like symptoms in rats prenatally exposed to valproic acid. Transl. Psychiatry 6:e902. doi: 10.1038/tp.2016.182

PubMed Abstract | CrossRef Full Text | Google Scholar

Sharma, N. R., Mani, P., Nandwani, N., Mishra, R., Rana, A., and Sarkar, D. P. (2010). Reciprocal regulation of AKT and MAP kinase dictates virus-host cell fusion. J. Virol. 84, 4366–4382. doi: 10.1128/jvi.01940-09

PubMed Abstract | CrossRef Full Text | Google Scholar

Sidhu, H., Dansie, L. E., Hickmott, P. W., Ethell, D. W., and Ethell, I. M. (2014). Genetic removal of matrix metalloproteinase 9 rescues the symptoms of fragile X syndrome in a mouse model. J. Neurosci. 34, 9867–9879. doi: 10.1523/jneurosci.1162-14.2014

PubMed Abstract | CrossRef Full Text | Google Scholar

Suvrathan, A., and Chattarji, S. (2011). Fragile X syndrome and the amygdala. Curr. Opin. Neurobiol. 21, 509–515. doi: 10.1016/j.conb.2011.04.005

PubMed Abstract | CrossRef Full Text | Google Scholar

Tian, M., Zeng, Y., Hu, Y., Yuan, X., Liu, S., Li, J., et al. (2015). 7, 8-Dihydroxyflavone induces synapse expression of AMPA GluA1 and ameliorates cognitive and spine abnormalities in a mouse model of fragile X syndrome. Neuropharmacology 89, 43–53. doi: 10.1016/j.neuropharm.2014.09.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Tounian, P., Frelut, M. L., Parlier, G., Abounaufal, C., Aymard, N., Veinberg, F., et al. (1999). Weight loss and changes in energy metabolism in massively obese adolescents. Int. J. Obes. Relat. Metab. Disord. 23, 830–837. doi: 10.1038/sj.ijo.0800959

PubMed Abstract | CrossRef Full Text | Google Scholar

Tyzio, R., Cossart, R., Khalilov, I., Minlebaev, M., Hübner, C. A., Represa, A., et al. (2006). Maternal oxytocin triggers a transient inhibitory switch in GABA signaling in the fetal brain during delivery. Science 314, 1788–1792. doi: 10.1126/science.1133212

PubMed Abstract | CrossRef Full Text | Google Scholar

Tyzio, R., Nardou, R., Ferrari, D. C., Tsintsadze, T., Shahrokhi, A., Eftekhari, S., et al. (2014). Oxytocin-mediated GABA inhibition during delivery attenuates autism pathogenesis in rodent offspring. Science 343, 675–679. doi: 10.1126/science.1247190

PubMed Abstract | CrossRef Full Text | Google Scholar

Vahdatpour, C., Dyer, A. H., and Tropea, D. (2016). Insulin-like growth factor 1 and related compounds in the treatment of childhood-onset neurodevelopmental disorders. Front. Neurosci. 10:450. doi: 10.3389/fnins.2016.00450

PubMed Abstract | CrossRef Full Text | Google Scholar

Viollet, B., Guigas, B., Sanz Garcia, N., Leclerc, J., Foretz, M., and Andreelli, F. (2012). Cellular and molecular mechanisms of metformin: an overview. Clin. Sci. 122, 253–270. doi: 10.1042/cs20110386

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, J., Ethridge, L. E., Mosconi, M. W., White, S. P., Binder, D. K., Pedapati, E. V., et al. (2017). A resting EEG study of neocortical hyperexcitability and altered functional connectivity in fragile X syndrome. J. Neurodev. Disord. 9:11. doi: 10.1186/s11689-017-9191-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, H., Dobkin, C., Sheikh, A. M., Malik, M., Brown, W. T., and Li, X. (2012). The therapeutic effect of memantine through the stimulation of synapse formation and dendritic spine maturation in autism and fragile X syndrome. PLoS One 7:e36981. doi: 10.1371/journal.pone.0036981

PubMed Abstract | CrossRef Full Text | Google Scholar

Westmark, C. J., Sokol, D. K., Maloney, B., and Lahiri, D. K. (2016). Novel roles of amyloid-β precursor protein metabolites in fragile X syndrome and autism. Mol. Psychiatry 21, 1333–1341. doi: 10.1038/mp.2016.134

PubMed Abstract | CrossRef Full Text | Google Scholar

Williams, P. G., and Hersch, J. H. (1997). A male with fetal valproate syndrome and autism. Dev. Med. Child Neurol. 39, 632–634. doi: 10.1111/j.1469-8749.1997.tb07500.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Williams, G., King, J., Cunningham, M., Stephan, M., Kerr, B., and Hersch, J. H. (2001). Fetal valproate syndrome and autism: additional evidence and association. Dev. Med. Child Neurol. 43, 202–206. doi: 10.1111/j.1469-8749.2001.tb00188.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Winarni, T. I., Schneider, A., Borodyanskara, M., and Hagerman, R. J. (2012). Early intervention combined with targeted treatment promotes cognitive and behavioral improvements in young children with fragile x syndrome. Case Rep. Genet. 2012:280813. doi: 10.1155/2012/280813

PubMed Abstract | CrossRef Full Text | Google Scholar

Wrigley, S., Arafa, D., and Tropea, D. (2017). Insulin-like growth factor 1: at the crossroads of brain development and aging. Front. Cell. Neurosci. 11:14. doi: 10.3389/fncel.2017.00014

PubMed Abstract | CrossRef Full Text | Google Scholar

Yabluchanskiy, A., Ma, Y., Iyer, R. P., Hall, M. E., and Lindsey, M. (2013). Matrix Metalloproteinase-9: many shades of function in cardiovascular disease. Physiology 28, 391–403. doi: 10.1152/physiol.00029.2013

PubMed Abstract | CrossRef Full Text | Google Scholar

Youssef, E. A., Berry-Kravis, E., Czech, C., Hagerman, R. J., Hessl, D., Wong, C. Y., et al. (2017). Effect of the mGluR5-NAM basimglurant on behavior in adolescents and adults with fragile X syndrome in a randomized, double-blind, placebo-controlled trial: fragxis phase 2 results. Neuropsychopharmacology doi: 10.1038/npp.2017.177 [Epub ahead of print].

PubMed Abstract | CrossRef Full Text | Google Scholar

Zeidler, S., Hukema, R. K., and Willemsen, R. (2015). The quest for targeted therapy in fragile X syndrome. Expert Opin. Ther. Targets 19, 1277–1281. doi: 10.1517/14728222.2015.1079176

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Fragile X Syndrome, FMR1, ASD, MMP-9, IGF-1, endocannabinoid system, oxytocin, serotonin

Citation: Castagnola S, Bardoni B and Maurin T (2017) The Search for an Effective Therapy to Treat Fragile X Syndrome: Dream or Reality? Front. Synaptic Neurosci. 9:15. doi: 10.3389/fnsyn.2017.00015

Received: 02 August 2017; Accepted: 16 October 2017;
Published: 06 November 2017.

Edited by:

Hansen Wang, University of Toronto, Canada

Reviewed by:

Christina Gross, Cincinnati Children’s Hospital Medical Center, United States
Emily Osterweil, University of Edinburgh, United Kingdom
Lucia Ciranna, Università degli Studi di Catania, Italy

Copyright © 2017 Castagnola, Bardoni and Maurin. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) or licensor are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Barbara Bardoni, bardoni@ipmc.cnrs.fr

Download