258

Autism and Fragile X Syndrome Elizabeth Berry-Kravis, MD, PhD2

1 Division of Genetics and Genomics, Boston Children’s Hospital, and

Department of Pediatrics, Harvard Medical School, Boston, Massachusetts 2 Departments of Pediatrics, Neurological Sciences, and Biochemistry, Rush University Medical Center, Chicago, Illinois Semin Neurol 2014;34:258–265.

Abstract

Keywords

► fragile X syndrome ► autism spectrum disorder ► fragile X mental retardation protein ► fragile X mental retardation 1 geneintellectual disability

Address for correspondence Timothy W. Yu, MD, PhD, Division of Genetics and Genomics, Boston Children’s Hospital, and Department of Pediatrics, Harvard Medical School, 300 Longwood Avenue, Boston, MA 02115 (e-mail: [email protected]).Elizabeth Berry-Kravis MD, PhD, Departments of Pediatrics, Neurological Sciences, and Biochemistry, Rush University Medical Center, 1725 West Harrison Street, Suite 718, Chicago, IL 60612 (e-mail: [email protected]).

Autistic spectrum disorders (ASDs) are characterized by impairments in language, social skills, and repetitive behaviors, often accompanied by intellectual disability. Advances in the genetics of ASDs are providing new glimpses into the underlying neurobiological mechanisms disrupted in these conditions. These glimpses on one hand reinforce the idea that synapse development and plasticity are one of the major pathways disrupted in autism, but beyond that are providing fresh molecular support to the idea of mechanistic parallels between idiopathic ASD and specific syndromic neurodevelopmental disorders like fragile X syndrome (FXS). Fragile X syndrome is already recognized as the most common identifiable genetic cause of intellectual disability and ASDs, with many overlapping phenotypic features. Fragile X syndrome is associated with a variety of cognitive, behavioral, physical, and medical problems, which are managed through supportive treatment. Recent major advances in the understanding of the underlying neurobiology in FXS have led to the discovery of agents that rescue phenotypes in the FXS mouse model, and early clinical trials of targeted treatments in humans with FXS. Thus translational strategies in FXS may be poised to serve as models for ASD and other cognitive disorders.

Autism Autistic spectrum disorder (ASD) is a neurodevelopmental disorder characterized by early-onset deficits in language, impaired social skills, and patterns of repetitive behaviors or restricted interests. Once considered rare, recent studies have estimated ASD prevalence worldwide to be approximately 1% to 2%,1,2 with a striking 4:1 male bias.3 Autistic spectrum disorders are associated with serious medical and psychosocial comorbidity, including concomitant intellectual disability (16%–50%), epilepsy (7%–35%), sleep disorders (40%–80%), as well as additional conditions (gastrointestinal [GI] disorders, mitochondrial and metabolic disorders) whose prevalence is incompletely understood.1,4,5 Although some individuals with ASDs have relatively mild impairment compatible with independent living, many affected individuals require

Issue Theme Neurogenetics; Guest Editor, Ali Fatemi, MD

educational and lifelong social support, with significant economic impact on their families and society.6 Historically, the extreme heterogeneity of ASD has posed a major challenge to traditional clinicopathologic approaches to investigating disease mechanisms. With some exceptions (e.g., autism with regression,7 or autism with macrocephaly8), the diversity of symptoms and severities encompassed by ASD have made it difficult to formulate robust endophenotypes based on clinical symptoms alone.9,10 Similarly, neuroanatomical investigations of brain tissue from individuals with ASD have highlighted abnormalities in the cerebellum, amygdala, and frontal lobes, but a consistent and unifying neuropathology has not been forthcoming.11 Nonetheless, in recent years, genetic and molecular approaches have made significant progress in penetrating this heterogeneity, and are beginning to reveal autism’s neurobiological underpinnings.

Copyright © 2014 by Thieme Medical Publishers, Inc., 333 Seventh Avenue, New York, NY 10001, USA. Tel: +1(212) 584-4662.

DOI http://dx.doi.org/ 10.1055/s-0034-1386764. ISSN 0271-8235.

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Timothy W. Yu, MD, PhD1

Twin studies have long implicated a strong genetic component to autism, with formal heritability estimates of approximately 50% to 90%.12–19 Simultaneously, it has also been long recognized that many genetic disorders have higher rates of autism than the general population, including phenylketonuria, fragile X syndrome, Rett syndrome, tuberous sclerosis, Angelman syndrome, and Smith Lemli-Opitz syndrome,20 and over 100 different genetic disorders have been reported that may include autistic features as one of their presenting phenotypes,21 although many research studies have tried to include only cases considered to be clinically idiopathic—a somewhat awkward and shifting standard. Rare, recurrent chromosomal abnormalities were among the first genetic abnormalities reported in idiopathic autism, including contiguous gene syndromes (e. g.,15q11-q13 duplication, 22q13 deletion), but occasionally implicating single genes (e.g., FOXP1, FOXP2, NLGN3, NLGN4X, NRXN1, SHANK2, SHANK3).22–29 Most of these chromosomal abnormalities were hundreds of kilobases to megabases in size, and were discovered in close examination of a few sporadic or rare familial cases. More recently, microarray studies in large autism cohorts have revealed additional recurrent ASD-associated copy number variants (CNVs) and translocations (e.g., 16p11.2 deletion/ duplication, 7q11.23 duplication, and others) as well as an excess of many smaller rare CNVs in ASD.30–40 Many of these occur in a de novo fashion, whereas others are inherited with incomplete penetrance. These studies also allowed estimation of the contribution of CNVs to autism as a whole; in aggregate de novo or inherited CNVs likely account for approximately 5% to 7% of cases, and the number of ASD-associated CNV loci is likely in the several hundreds.35–39 Identification of these CNVs is a critical piece of the puzzle for understanding the genetic architecture of ASD, but notably, does not always point to single causative genes or mechanisms, even when largescale resequencing of candidate genes in the interval is undertaken,41 and arguments for plausible biological candidates can be made. 42 This and other evidence suggests that CNVs may exert their effects through the combinatorial action of more than one included or nearby gene, or in concert with additional genetic or nongenetic factors.43,44 The challenge of identifying specific, impactful ASD genes is now being met with advances in high-throughput sequencing. The declining costs and increasing efficiency of sequencing have enabled novel study designs, a particularly valuable example of which has been whole exome sequencing (WES) in ASD trios (affected children and their parents) or quartets (trios plus their unaffected siblings) to find de novo point mutations. Three pilot reports published in 2012 analyzed whole exome sequences from approximately 600 trios and quartets and demonstrated an excess of disruptive de novo mutations in ASD cases.45–47 This excess was particularly evident in brain-expressed genes, and its magnitude was consistent with the idea that de novo mutations may contribute to up to 5% to 15% of sporadic ASD cases. Because de novo

Yu, Berry-Kravis

mutations are rare (approximately 1 per exome per generation), finding multiple recurrent events in the same gene in different cases constitutes strong statistical evidence of disease association. Combined with an important follow-up study in which the top candidate genes were resequenced in an additional  2400 probands,48 this approach yielded six new specific ASD genes that act via de novo mutations (CHD8, DYRK1A, GRIN2B, TBR1, PTEN, and TBL1XR1), and which together contribute to approximately 1% of sporadic ASD. Large-scale efforts to extend this approach to larger and larger datasets are underway.49 In addition to offering insight into de novo mutations, WES has also begun to provide insights into the contribution of inherited point mutations to ASD, critical to understanding the heritability of this disorder. These insights range from recessive analyses of larger and larger cohorts of multiplex and/or consanguineous families, 50,51 to population analyses of large case-control datasets to provide baseline estimates for the proportion of sporadic ASD cases that might be affected by recessive mutations.52 These analyses have uncovered examples of ASD resulting from recessive mutations in previously unrecognized single gene disorders (e.g., UBE3B50,53), as well as examples in which ASD is due to an unusually mild clinical presentation of a more severe neurodevelopmental or neurometabolic single-gene disorder due to a hypomorphic mutation (e.g., functional missense mutations in AMT causing atypical glycine encephalopathy51). In parallel, sequencing advances are also fueling many smaller-scale (but equally biologically interesting) discoveries of genes associated with ASD and/or intellectual disability with autistic features (e.g., BCKDK,54 ANK2,55 and ADNP56). These advances are shaping and refining our understanding of ASD and its causes. Clinically, the heterogeneity of ASDs remains a critical issue and reinforces the importance of careful consideration of underlying neurometabolic or other syndromic single-gene disorders even in nominally idiopathic cases, especially when they may be potentially intervenable. Nonetheless, there is beginning to emerge evidence of core biological mechanisms. Autistic spectrum disorder genes implicated by copy number variation and/or point mutations, as well as parallel measurements of transcriptional modules altered in postmortem brain tissue of individuals with ASD,55,57,58 implicate several broad biological domains in ASD pathogenesis—synaptic function, neuronal signaling and development, and chromatin regulation32,36,45,55,59–63 —but more importantly, with the increasing resolution afforded by more established loci, evidence is strengthening for more specific molecular pathways, such as glutamatergic signaling61,63 and MAPK signal transduction.61 One of the most intriguing and increasingly replicated results is the finding of significant overlap of ASD loci with the fragile X mental retardation protein (FMRP) signaling pathway.47,60,64 These mechanistic connections raise the possibility of leveraging knowledge about well-studied conditions like fragile X syndrome to further hypotheses about ASD pathophysiology, and are being used to guide the development and evaluation of candidate therapies. Seminars in Neurology

Vol. 34

No. 3/2014

259

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Autism and Fragile X Syndrome

Autism and Fragile X Syndrome

Yu, Berry-Kravis

Fragile X Syndrome Fragile X syndrome (FXS) is the most common identifiable single gene cause of intellectual disability (ID) and autism spectrum disorder (ASD), with an estimated frequency of approximately 1:4000 to 5000.65 The disorder affects all ethnic groups worldwide, although prevalence varies between different populations. Fragile X syndrome is one of the fragile X-associated disorders (FXDs), which arise from a trinucleotide repeat (CGG) expansion mutations in the promoter region of FMR1 (fragile X mental retardation 1 gene). This CGG sequence is transcribed into the FMR1 mRNA, but as the sequence is located in the 5′untranslated region of the mRNA, the number of repeat units does not affect the sequence of the protein product of the FMR1 gene (FMRP).66 Smaller mutations in the gene (55–200 CGG repeats) are termed a “premutation,” now known to have a prevalence of approximately 1:151 to 1:209 females and 1:430 to 1:468 males in populations in the United States.67,68 The premutation is associated with risk for two well-defined adult-onset diseases: fragile X-associated tremor/ataxia syndrome (FXTAS) and fragile X-associated primary ovarian insufficiency (FXPOI). The mechanism for premutation-associated disease is thought to relate to elevated FMR1 mRNA levels and resultant CGG repeat-mediated RNA toxicity, although recent findings of elevated antisense FMR1 transcripts (ASFMR) containing the repeat in premutation carriers69 and polyglycine produced due to aberrant translation of the premutation repeat at the ribosome,70 have suggested multiple mechanisms of toxicity in FMR1 premutation diseases. Indeed, individuals with a large premutation (> 150 repeats) also have low FMRP levels due to repeat-mediated translational stalling and manifest symptoms milder than but overlapping with those seen with FXS.66 Large “full mutation” expansions in FMR1 (> 200 repeats) cause FXS, which results from the methylation and transcriptional silencing of the FMR1 promoter with consequent loss or significant reduction of FMRP.66 FMR1 expansions tend to increase in size as they are passed from generation to generation, so FXD affect families in multiple generations. Adults with FXTAS present with tremor and/or ataxia, and may have executive dysfunction, neuropathy, parkinsonism, vestibular dysfunction, psychiatric symptoms like anxiety, and in some cases autonomic dysfunction and dementia.71 Clinically, patients often have a mixture of neurologic symptoms that make them difficult to categorize diagnostically. Progression is variable, but is typically over many years. Brain atrophy, increased signal in the deep white matter, and middle cerebellar peduncle hyperintensity (MCP sign) are seen on magnetic resonance imaging (MRI) scans, whereas the hallmark pathological finding is that of intranuclear inclusions in neurons.71 Diagnosis is through FMR1 DNA testing by PCR to identify the premutation expansion. Consistent with the RNA toxicity mechanism, the CGG repeat length in the premutation correlates with onset and severity of disease, MRI findings, and number of inclusions.71 Treatment is supportive; β-blockers for tremor, L-DOPA for parkinsonism, antidepressants for anxiety, and donepezil or Seminars in Neurology

Vol. 34

No. 3/2014

memantine for executive and cognitive problems have sometimes been helpful anecdotally.71 Children with FXS typically present with developmental delay. Although motor delays are often seen, these tend to be mild, and affected males most commonly come to medical attention because of language delay. Hypotonia is often present early in life, but improves with development, and by school age evolves into coordination and praxis problems. Females and higher-functioning males with FXS may have fairly normal early language acquisition, and present with behavioral problems, anxiety, attention deficit hyperactivity disorder (ADHD) symptoms, or learning delays in school. Males with FXS typically display ID that can range from mild to severe, with characteristic strengths and weaknesses in specific areas of cognition and language. Their IQs tend to decline with age during childhood, and this is not the result of regression, but rather failure to keep pace with the normal rate of intellectual development. Average IQ in adult males with FXS is 40 to 50, with a mental age of about of 5 to 6 years.72 Physical features include macro-orchidism in most adult males, as well as prominent ears, macrocephaly, long face, prominent jaw and forehead, midfacial hypoplasia, high arched palate, and connective tissue laxity, leading to hyperextensible joints, flat feet, and soft skin. Physical features are sufficiently variable that they cannot be used as indicators of which patients to screen. Medical problems commonly include seizures (11%–15%, usually in childhood), strabismus (10%–30%), and sleep disorders (32%).73 Frequent ear infections, gastroesophageal reflux, sleep apnea and other sleep problems, loose stools, and allergies have been reported to be more prevalent in FXS than the general population, although this awaits confirmation with more rigorous epidemiological studies that are currently underway through the Fragile X Clinical and Research Consortium (FXCRC), a group of 27 fragile X clinics organized to promote multi-institutional research and optimize clinical practices in FXS.73 Behavioral symptoms are very prominent in FXS and include hyperactivity, impulsivity, attention problems, anxiety, mood lability, aggression, self-injury, and autistic features such as poor eye contact, self-talk, hand flapping, hand biting, hyperarousal to sensory stimuli, and perseverative language and behavior.72,74 Females with a full mutation are more variably and typically more mildly affected than males, due to production of FMRP from the normal FMR1 allele in cells expressing the nonmutated X chromosome. Average IQ in females is approximately 80, with a range from severe impairment to normal or even superior ability. The pattern of cognitive strengths and weaknesses is similar to that seen in males. Even when females with FXS have a normal IQ, there is frequently learning disability, including particularly executive and social deficits, as well as anxiety, shyness, and attention problems. Severity of cognitive impairment in females with the full mutation is thought to be related to the activation ratio for the normal FMR1 allele and the amount of FMRP expression.75 Males with mosaicism for a full- and premutation or a partially unmethylated full mutation may also be mildly

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

260

affected, with severity in this case also related to the amount of unmethylated DNA and FMRP levels.75 About a half to two-thirds of males and approximately 20% of females with FXS meet criteria for ASD.76 Behavioral and neurologic phenotypes common to FXS and ASD include social interactional and communication/language deficits, poor eye contact, repetitive motor movements such as stereotypies, unusual reactions to sensory input, including hypersensitivity (e.g., tactile defensiveness) and hyposensitivity (e.g., high tolerance for pain), cognitive strength in visual memory, increased prevalence of seizures relative to the general population (although peak incidence is in childhood in FXS and adolescence in ASD), macrocephaly, difficulties with regulation of attention, activity level, emotional behavior, and mood, often leading to additional diagnoses (e. g., ADHD, anxiety, mood disorder), other problematic behaviors (e.g., aggression, noncompliance, self-injury), and sleep problems. Some behavioral characteristics or symptoms are seen in both ASD and FXS, but for different underlying reasons. In FXS, for example, lack of social initiation and poor eye contact is thought to be mostly due to anxiety77 rather than lack of social awareness or interest commonly seen in ASD. Characteristics that tend to differ between the FXS and ASD phenotypes include a higher rate of ID in FXS than ASD, significantly more severe motor coordination deficits in FXS than ASD, worse expressive than receptive language in FXS with the reverse pattern more likely in ASD, generally higher interest in socialization in FXS (although limited by anxiety), and better imitation skills in FXS than ASD. Individuals with FXS þ ASD, relative to those with FXS alone, have less-developed language skills, lower IQ and adaptive skills, moresevere overall behavioral problems, and reduced social interaction.76 Current treatment of FXS is supportive,72,74 is similar to supportive treatment employed in ASD, and includes speech, occupational, and physical therapy, and educational strategies designed based on the pattern of cognitive and behavioral strengths and weaknesses in FXS. Treatment of medical problems is important, as these problems can impact development or behavior. Such treatment depends on symptoms manifested by the individual patient, and would include aggressive treatment of frequent otitis media to prevent an impact on already compromised language, management of obstructive sleep apnea with tonsillectomy/adenoidectomy to prevent cognitive and behavioral aggravation, orthotics when needed for foot pronation and flat feet to avoid leg pain and reduce gait problems, management of strabismus with eye patching, vision therapy, or surgery to avoid amblyopia and compounding of visual processing problems, use of antacids when needed for GE reflux to prevent pain and resulting behavioral decompensation, and seizure control with anticonvulsants that do not have a tendency to aggravate behavior. Behavior problems in FXS typically require intensive consistent behavioral modification strategies implemented at home, at school, and in other settings. Frequently psychopharmacology is employed for attention deficits, anxiety, or other problematic maladaptive behaviors to improve overall

Yu, Berry-Kravis

functioning.72,74 Psychopharmacologic treatment of ADHD symptoms, anxiety, hyperarousal, and irritable/aggressive behaviors with medications such as stimulants, SSRIs, αagonists, and antipsychotics, respectively, appears to be helpful in a clinical setting in approximately 50% to 70% of patients.74 Response is typically incomplete, and data from a National Survey on FXS showed that approximately 10% to 20% of respondents felt that medication was not helpful for the behavior problems being treated in their child with FXS, while only approximately 40% felt the medication was helping a lot.78 There is clearly an unmet need in FXS for better medications to treat behavior and for development of treatments that target cognitive deficits, and thus treatments that modify the underlying disorder would be a tremendously important advance. In the past decade, study of the neurobiology and synaptic mechanisms resulting from absence of FMRP in FXS has emerged as an important window to future targeted treatments for FXS and for ASD and related neurodevelopmental disorders (NDDs).76,79 The Fmr1 knockout mouse, which makes no functional FMRP, has been a critical resource to understand the role of FMRP in neurons, identify cellular treatment targets, and explore effects of proposed diseasemodifying agents. FMRP is an mRNA binding protein involved in the transport, localization and translational regulation of a subset of dendritic mRNAs.79 FMRP functions in a protein complex at the ribosome, where it regulates dendritic protein translation in response to synaptic activation by Group 1 metabotropic glutamate receptors (mGluR1 and mGluR5),80 muscarinic (M1) acetylcholine receptors and probably multiple Gq-linked receptors, including dopamine D1 receptors. Activation of these receptors results in signaling through ERKand mTOR-dependent signaling pathways that ultimately results in loss of FMRP repressor function at the ribosome, and a pulse of new protein synthesis. Translation of a broad array of proteins involved in synaptic functions is regulated by FMRP, such as STEP and Arc, which are linked to AMPA receptor internalization.81 Fragile X mental retardation protein also regulates activity of some pre- and postsynaptic ion channels such as BK and SLACK channels through direct protein–protein interactions.82 These regulatory functions of FMRP appear to be critical for synaptic maturation and strength, as in the absence of FMRP, there is a constitutive elevation of synaptic proteins usually controlled by FMRP, immature elongated morphology of dendritic spines,83 abnormal spine density, abnormal synaptic plasticity including enhanced mGluR-activated hippocampal and cerebellar LTD, and impaired LTP in hippocampus, cortex and amygdala, and proneness to abnormal epileptiform discharges.79,82 The morphological abnormalities and synaptic plasticity deficits found in the fmr1 knockout mouse are associated with numerous cognitive, behavioral, and electrophysiological phenotypes, including abnormal ocular dominance plasticity, olfactory learning deficits, impaired memory formation, decreased motor learning, increased open-field hyperactivity, abnormal marble burying, abnormal social behaviors, abnormal prepulse inhibition (PPI), prolonged epileptiform bursts, neuronal network hyperexcitability, audiogenic seizures, Seminars in Neurology

Vol. 34

No. 3/2014

261

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Autism and Fragile X Syndrome

Autism and Fragile X Syndrome

Yu, Berry-Kravis

abnormal growth patterns, and increased protein synthesis.79,82 The Drosophila model of FXS, in which there is loss of dfmr1 (homolog of the FMR1 gene in the Drosophila genome), shows defects in circadian rhythms, synaptic branching, courtship behavior, and learning.79,82 The abnormalities observed in the absence of FMRP in the mouse model of FXS has led to identification of treatment targets directed at (1) reduction of excess activity in signal transduction pathways leading from group 1 mGluRs or other Gq-linked receptors to the dendritic translational machinery, (2) reduction of excessive activity of individual proteins normally regulated by FMRP, (3) increasing expression and activation of surface AMPA receptors, (4) modification of activity of GABA and other receptors/proteins that regulate glutamate signaling, and (5) blocking excessive translation of mRNAs normally regulated by FMRP using miRNAs.79,82 Treatments aimed at all of these types of targets have shown success in reversing phenotypes in the Fmr1 knockout mouse84–86 and dfxr fly87 models even in adulthood,88 suggesting that there may not be an absolute developmental requirement for FMRP. Successful preclinical testing in FXS models has led to early proof-of-concept clinical trials and subsequent larger trials for some of the proposed targeted treatments. Lithium, thought to reduce excess mGluR-dependent activation of translation by attenuating GSK3β activity and possibly phosphatidyl inositol (PI) turnover, resulted in significant improvement in behavioral scales, verbal memory, and abnormal ERK phosphorylation rates in lymphocytes in a 2month pilot open-label proof-of-concept trial in children and young adults with FXS.89 A pilot placebo-controlled crossover trial of minocycline, an antibiotic that inhibits overexpressed synaptic MMP9 in FXS models, conducted in children with FXS showed mild global clinical improvement and reduction of blood MMP9 levels in responders.90 GABA-B agonist arbaclofen presumably lowers presynaptic glutamate release with resultant reduction of group 1 mGluR signaling. In a phase II double-blind placebo-controlled crossover trial,91 arbaclofen showed improvement over placebo in the entire per protocol group for parent-nominated problem behaviors and social withdrawal. However, a large-phase III placebo-controlled trial in adolescents and adults with FXS did not show benefits for arbaclofen over placebo in the primary outcome of social withdrawal. An additional phase III trial in children with FXS is pending. Acamprosate, currently approved by the U.S. Food and Drug Administration (FDA) for alcohol withdrawal, with agonist properties at both GABA-A and GABA-B receptors, has shown promise for hyperactivity and social functioning in FXS in an open-label trial.92 Acamprosate and GABA-A agonist ganaxolone are being tested in small placebo-controlled trials in FXS (clinicaltrials.gov). Multiple negative modulators of the mGluR5 receptor (group 1 mGluR expressed throughout the brain) have been in trials in FXS. A single oral dose of fenobam93 resulted in a significant improvement in abnormal PPI compared with untreated control subjects with FXS. A phase II double-blind placebo-controlled, crossover trial of AFQ056 in 30 adult Seminars in Neurology

Vol. 34

No. 3/2014

males with FXS treated for 28 days each with AFQ and placebo,94 suggested improvement in maladaptive behavior in a post hoc analysis in the subgroup with full methylation of FMR1. Larger multinational trials have not supported this behavioral outcome, but have not addressed cognitive outcomes. RO4917523, another mGluR5 negative modulator is also in large multinational trials in FXS (clinicaltrials.gov). Although many neuronal targets for treating the underlying disorder in FXS have emerged, and early translational work has begun, there have been problems with demonstrating disease modification in early trials because there are still many uncertainties about how to optimally demonstrate treatment effects in a clinical trial setting.78,79,95 Major trial design issues in FXS trials potentially include variable but narrow dosing windows; timing (age) and length of treatment necessary; potential need for cognitive or behavioral interventions to see drug effects on learning; large placebo effects; and lack of validated, sensitive biomarkers and functional outcome measures in FXS.95 These trial design issues will need to be resolved to be able to demonstrate disease modification in FXS, but it is hoped that in the future treatment to reverse the underlying disorder will replace or complement supportive treatment. There is significant overlap in molecular and cellular pathways involving FMRP and those that include gene products associated with ASD.75,79 This overlap falls into three broad categories: (1) defects in proteins in the signaling cascade for activation of FMRP-regulated translation, such as SHANK, mTOR, PAK, and PTEN; (2) defects in proteins regulated directly by FMRP, such as PSD95 and Arc; and (3) defects in proteins that regulate the balance of activity in brain glutamate and GABA systems. Indeed, this pathway overlap has been recently supported by the findings that (1) FMRP binds to one-third to one-half of all genes identified as associated with ASD in a meta-analysis of exome screening studies47; (2) FMRP target genes are more likely than other genes with similar expression patterns to contribute to ASD64; and (3) common variants in genes involved in postsynaptic regulation of FMRP (CAMK4, GRM1, CYFIP1) are risk factors for ASD.96 Treatments directed at mechanisms involving all of these overlap areas are becoming available in FXS trials and if successful, progress in development of targeted treatments for FXS is likely to result in targeted treatments to reverse central nervous system defects and clinical manifestations of ASD, other NDDs, and intellectual disability.

References 1 Baio J. Prevalence of autism spectrum disorder among children

aged 8 years - autism and developmental disabilities monitoring network, 11 sites, United States, 2010. MMWR Surveill Summ 2014;63(2):1–21 2 Kim YS, Leventhal BL, Koh Y-J, et al. Prevalence of autism spectrum disorders in a total population sample. Am J Psychiatry 2011; 168(9):904–912 3 Werling DM, Geschwind DH. Sex differences in autism spectrum disorders. Curr Opin Neurol 2013;26(2):146–153 4 Bauman ML. Medical comorbidities in autism: challenges to diagnosis and treatment. Neurotherapeutics 2010;7(3):320–327

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

262

Yu, Berry-Kravis

5 Doshi-Velez F, Ge Y, Kohane I. Comorbidity clusters in autism

27 Durand CM, Betancur C, Boeckers TM, et al. Mutations in the gene

spectrum disorders: an electronic health record time-series analysis. Pediatrics 2014;133(1):e54–e63 Lavelle TA, Weinstein MC, Newhouse JP, Munir K, Kuhlthau KA, Prosser LA. Economic burden of childhood autism spectrum disorders. Pediatrics 2014;133(3):e520–e529 Rossignol DA, Frye RE. Mitochondrial dysfunction in autism spectrum disorders: a systematic review and meta-analysis. Mol Psychiatry 2012;17(3):290–314 Butler MG, Dasouki MJ, Zhou X-P, et al. Subset of individuals with autism spectrum disorders and extreme macrocephaly associated with germline PTEN tumour suppressor gene mutations. J Med Genet 2005;42(4):318–321 Geschwind DH. Advances in autism. Annu Rev Med 2009; 60:367–380 Jeste SS. The neurology of autism spectrum disorders. Curr Opin Neurol 2011;24(2):132–139 Amaral DG, Schumann CM, Nordahl CW. Neuroanatomy of autism. Trends Neurosci 2008;31(3):137–145 Folstein S, Rutter M. Genetic influences and infantile autism. Nature 1977;265(5596):726–728 Steffenburg S, Gillberg C, Hellgren L, et al. A twin study of autism in Denmark, Finland, Iceland, Norway and Sweden. J Child Psychol Psychiatry 1989;30(3):405–416 Bailey A, Le Couteur A, Gottesman I, et al. Autism as a strongly genetic disorder: evidence from a British twin study. Psychol Med 1995;25(1):63–77 Lichtenstein P, Carlström E, Råstam M, Gillberg C, Anckarsäter H. The genetics of autism spectrum disorders and related neuropsychiatric disorders in childhood. Am J Psychiatry 2010;167(11): 1357–1363 Ronald A, Happé F, Plomin R. Genetic research into autism. Science 2006;311(5763):952, author reply 952 Hallmayer J, Cleveland S, Torres A, et al. Genetic heritability and shared environmental factors among twin pairs with autism. Arch Gen Psychiatry 2011;68(11):1095–1102 Constantino JN, Todorov A, Hilton C, et al. Autism recurrence in half siblings: strong support for genetic mechanisms of transmission in ASD. Mol Psychiatry 2013;18(2):137–138 Sandin S, Lichtenstein P, Kuja-Halkola R, Larsson H, Hultman CM, Reichenberg A. The familial risk of autism. JAMA 2014;311(17): 1770–1777 Cohen D, Pichard N, Tordjman S, et al. Specific genetic disorders and autism: clinical contribution towards their identification. J Autism Dev Disord 2005;35(1):103–116 Betancur C. Etiological heterogeneity in autism spectrum disorders: more than 100 genetic and genomic disorders and still counting. Brain Res 2011;1380:42–77 Wandstrat AE, Leana-Cox J, Jenkins L, Schwartz S. Molecular cytogenetic evidence for a common breakpoint in the largest inverted duplications of chromosome 15. Am J Hum Genet 1998;62(4):925–936 Goizet C, Excoffier E, Taine L, et al. Case with autistic syndrome and chromosome 22q13.3 deletion detected by FISH. Am J Med Genet 2000;96(6):839–844 Prasad C, Prasad AN, Chodirker BN, et al. Genetic evaluation of pervasive developmental disorders: the terminal 22q13 deletion syndrome may represent a recognizable phenotype. Clin Genet 2000;57(2):103–109 Jamain S, Quach H, Betancur C, et al; Paris Autism Research International Sibpair Study. Mutations of the X-linked genes encoding neuroligins NLGN3 and NLGN4 are associated with autism. Nat Genet 2003;34(1):27–29 Kim H-G, Kishikawa S, Higgins AW, et al. Disruption of neurexin 1 associated with autism spectrum disorder. Am J Hum Genet 2008; 82(1):199–207

encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders. Nat Genet 2007;39(1):25–27 Berkel S, Marshall CR, Weiss B, et al. Mutations in the SHANK2 synaptic scaffolding gene in autism spectrum disorder and mental retardation. Nat Genet 2010;42(6):489–491 Lai CSL, Fisher SE, Hurst JA, Vargha-Khadem F, Monaco AP. A forkhead-domain gene is mutated in a severe speech and language disorder. Nature 2001;413(6855):519–523 Weiss LA, Shen Y, Korn JM, et al; Autism Consortium. Association between microdeletion and microduplication at 16p11.2 and autism. N Engl J Med 2008;358(7):667–675 Christian SL, Brune CW, Sudi J, et al. Novel submicroscopic chromosomal abnormalities detected in autism spectrum disorder. Biol Psychiatry 2008;63(12):1111–1117 Gilman SR, Iossifov I, Levy D, Ronemus M, Wigler M, Vitkup D. Rare de novo variants associated with autism implicate a large functional network of genes involved in formation and function of synapses. Neuron 2011;70(5):898–907 Marshall CR, Noor A, Vincent JB, et al. Structural variation of chromosomes in autism spectrum disorder. Am J Hum Genet 2008;82(2):477–488 Sebat J, Lakshmi B, Malhotra D, et al. Strong association of de novo copy number mutations with autism. Science 2007;316(5823): 445–449 Shen Y, Dies KA, Holm IA, et al; Autism Consortium Clinical Genetics/DNA Diagnostics Collaboration. Clinical genetic testing for patients with autism spectrum disorders. Pediatrics 2010; 125(4):e727–e735 Sanders SJ, Ercan-Sencicek AG, Hus V, et al. Multiple recurrent de novo CNVs, including duplications of the 7q11.23 Williams syndrome region, are strongly associated with autism. Neuron 2011; 70(5):863–885 Levy D, Ronemus M, Yamrom B, et al. Rare de novo and transmitted copy-number variation in autistic spectrum disorders. Neuron 2011;70(5):886–897 Pinto D, Pagnamenta AT, Klei L, et al. Functional impact of global rare copy number variation in autism spectrum disorders. Nature 2010;466(7304):368–372 Glessner JT, Wang K, Cai G, et al. Autism genome-wide copy number variation reveals ubiquitin and neuronal genes. Nature 2009;459(7246):569–573 Talkowski ME, Rosenfeld JA, Blumenthal I, et al. Sequencing chromosomal abnormalities reveals neurodevelopmental loci that confer risk across diagnostic boundaries. Cell 2012;149(3):525–537 Konyukh M, Delorme R, Chaste P, et al. Variations of the candidate SEZ6L2 gene on Chromosome 16p11.2 in patients with autism spectrum disorders and in human populations. PLoS ONE 2011; 6(3):e17289 Golzio C, Willer J, Talkowski ME, et al. KCTD13 is a major driver of mirrored neuroanatomical phenotypes of the 16p11.2 copy number variant. Nature 2012;485(7398):363–367 Girirajan S, Rosenfeld JA, Cooper GM, et al. A recurrent 16p12.1 microdeletion supports a two-hit model for severe developmental delay. Nat Genet 2010;42(3):203–209 Girirajan S, Rosenfeld JA, Coe BP, et al. Phenotypic heterogeneity of genomic disorders and rare copy-number variants. N Engl J Med 2012;367(14):1321–1331 O’Roak BJ, Vives L, Girirajan S, et al. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 2012;485(7397):246–250 Sanders SJ, Murtha MT, Gupta AR, et al. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 2012;485(7397):237–241 Iossifov I, Ronemus M, Levy D, et al. De novo gene disruptions in children on the autistic spectrum. Neuron 2012;74(2):285–299

6

7

8

9 10 11 12 13

14

15

16 17

18

19

20

21

22

23

24

25

26

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

Seminars in Neurology

Vol. 34

No. 3/2014

263

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

Autism and Fragile X Syndrome

Autism and Fragile X Syndrome

Yu, Berry-Kravis

48 O’Roak BJ, Vives L, Fu W, et al. Multiplex targeted sequencing

68 Tassone F, Iong KP, Tong T-H, et al. FMR1 CGG allele size and

identifies recurrently mutated genes in autism spectrum disorders. Science 2012;338(6114):1619–1622 Buxbaum JD, Daly MJ, Devlin B, Lehner T, Roeder K, State MW; Autism Sequencing Consortium. The autism sequencing consortium: large-scale, high-throughput sequencing in autism spectrum disorders. Neuron 2012;76(6):1052–1056 Chahrour MH, Yu TW, Lim ET, et al; ARRA Autism Sequencing Collaboration. Whole-exome sequencing and homozygosity analysis implicate depolarization-regulated neuronal genes in autism. PLoS Genet 2012;8(4):e1002635 Yu TW, Chahrour MH, Coulter ME, et al. Using whole-exome sequencing to identify inherited causes of autism. Neuron 2013; 77(2):259–273 Lim ET, Raychaudhuri S, Sanders SJ, et al; NHLBI Exome Sequencing Project. Rare complete knockouts in humans: population distribution and significant role in autism spectrum disorders. Neuron 2013;77(2):235–242 Basel-Vanagaite L, Dallapiccola B, Ramirez-Solis R, et al. Deficiency for the ubiquitin ligase UBE3B in a blepharophimosis-ptosisintellectual-disability syndrome. Am J Hum Genet 2012;91(6): 998–1010 Novarino G, El-Fishawy P, Kayserili H, et al. Mutations in BCKDkinase lead to a potentially treatable form of autism with epilepsy. Science 2012;338(6105):394–397 Willsey AJ, Sanders SJ, Li M, et al. Coexpression networks implicate human midfetal deep cortical projection neurons in the pathogenesis of autism. Cell 2013;155(5):997–1007 Helsmoortel C, Vulto-van Silfhout AT, Coe BP, et al. A SWI/SNFrelated autism syndrome caused by de novo mutations in ADNP. Nat Genet 2014;46(4):380–384 Voineagu I, Wang X, Johnston P, et al. Transcriptomic analysis of autistic brain reveals convergent molecular pathology. Nature 2011;474(7351):380–384 Parikshak NN, Luo R, Zhang A, et al. Integrative functional genomic analyses implicate specific molecular pathways and circuits in autism. Cell 2013;155(5):1008–1021 Neale BM, Kou Y, Liu L, et al. Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 2012;485(7397): 242–245 Uddin M, Tammimies K, Pellecchia G, et al. Brain-expressed exons under purifying selection are enriched for de novo mutations in autism spectrum disorder. Nat Genet 2014;46(7):742–747. Pinto D, Delaby E, Merico D, et al. Convergence of genes and cellular pathways dysregulated in autism spectrum disorders. Am J Hum Genet 2014;94(5):677–694 Luo R, Sanders SJ, Tian Y, et al. Genome-wide transcriptome profiling reveals the functional impact of rare de novo and recurrent CNVs in autism spectrum disorders. Am J Hum Genet 2012;91(1):38–55 Kelleher RJ III, Geigenmüller U, Hovhannisyan H, et al. Highthroughput sequencing of mGluR signaling pathway genes reveals enrichment of rare variants in autism. PLoS ONE 2012;7(4): e35003 Steinberg J, Webber C. The roles of FMRP-regulated genes in autism spectrum disorder: single- and multiple-hit genetic etiologies. Am J Hum Genet 2013;93(5):825–839 Coffee B, Keith K, Albizua I, et al. Incidence of fragile X syndrome by newborn screening for methylated FMR1 DNA. Am J Hum Genet 2009;85(4):503–514 Willemsen R, Levenga J, Oostra BA. CGG repeat in the FMR1 gene: size matters. Clin Genet 2011;80(3):214–225 Seltzer MM, Baker MW, Hong J, Maenner M, Greenberg J, Mandel D. Prevalence of CGG expansions of the FMR1 gene in a US population-based sample. Am J Med Genet B Neuropsychiatr Genet 2012;159B(5):589–597

prevalence ascertained through newborn screening in the United States. Genome Med 2012;4(12):100 Ladd PD, Smith LE, Rabaia NA, et al. An antisense transcript spanning the CGG repeat region of FMR1 is upregulated in premutation carriers but silenced in full mutation individuals. Hum Mol Genet 2007;16(24):3174–3187 Iliff AJ, Renoux AJ, Krans A, Usdin K, Sutton MA, Todd PK. Impaired activity-dependent FMRP translation and enhanced mGluR-dependent LTD in Fragile X premutation mice. Hum Mol Genet 2013; 22(6):1180–1192 Berry-Kravis E, Abrams L, Coffey SM, et al. Fragile X-associated tremor/ataxia syndrome: clinical features, genetics, and testing guidelines. Mov Disord 2007;22(14):2018–2030, quiz 2140 Berry-Kravis E, Grossman AW, Crnic LS, et al. Understanding fragile X syndrome. Curr Pediatr 2002;12:316–324 Kidd SL, Lachiewicz A, Barbouth D, et al. Fragile X syndrome: a review of associated medical problems. J Pediatr 2014; In press Berry-Kravis E, Sumis A, Hervey C, Mathur S. Clinic-based retrospective analysis of psychopharmacology for behavior in fragile x syndrome. Int J Pediatr 2012;2012:843016 Loesch DZ, Huggins RM, Hagerman RJ. Phenotypic variation and FMRP levels in fragile X. Ment Retard Dev Disabil Res Rev 2004; 10(1):31–41 Wang LW, Berry-Kravis E, Hagerman RJ. Fragile X: leading the way for targeted treatments in autism. Neurotherapeutics 2010;7(3): 264–274 Talisa VB, Boyle L, Crafa D, Kaufmann WE. Autism and anxiety in males with fragile X syndrome: an exploratory analysis of neurobehavioral profiles from a parent survey. Am J Med Genet A 2014; 164A(5):1198–1203 Bailey DB Jr, Raspa M, Bishop E, Olmsted M, Mallya UG, BerryKravis E. Medication utilization for targeted symptoms in children and adults with fragile X syndrome: US survey. J Dev Behav Pediatr 2012;33(1):62–69 Berry-Kravis E, Knox A, Hervey C. Targeted treatments for fragile X syndrome. J Neurodev Disord 2011;3(3):193–210 Huber KM, Gallagher SM, Warren ST, Bear MF. Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proc Natl Acad Sci U S A 2002;99(11):7746–7750 Deng P-Y, Rotman Z, Blundon JA, et al. FMRP regulates neurotransmitter release and synaptic information transmission by modulating action potential duration via BK channels. Neuron 2013; 77(4):696–711 Berry-Kravis E. Mechanism-based treatments in neurodevelopmental disorders: fragile X syndrome. Pediatr Neurol 2014;50(4): 297–302 Grossman AW, Aldridge GM, Weiler IJ, Greenough WT. Local protein synthesis and spine morphogenesis: Fragile X syndrome and beyond. J Neurosci 2006;26(27):7151–7155 Bilousova TV, Dansie L, Ngo M, et al. Minocycline promotes dendritic spine maturation and improves behavioural performance in the fragile X mouse model. J Med Genet 2009;46(2): 94–102 Henderson C, Wijetunge L, Kinoshita MN, et al. Reversal of diseaserelated pathologies in the fragile X mouse model by selective activation of GABAB receptors with arbaclofen. Sci Transl Med 2012;4(152):152ra128 Yuskaitis CJ, Mines MA, King MK, Sweatt JD, Miller CA, Jope RS. Lithium ameliorates altered glycogen synthase kinase-3 and behavior in a mouse model of fragile X syndrome. Biochem Pharmacol 2010;79(4):632–646 McBride SMJ, Choi CH, Wang Y, et al. Pharmacological rescue of synaptic plasticity, courtship behavior, and mushroom body defects in a Drosophila model of fragile X syndrome. Neuron 2005; 45(5):753–764

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66 67

Seminars in Neurology

Vol. 34

No. 3/2014

69

70

71

72 73 74

75

76

77

78

79 80

81

82

83

84

85

86

87

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

264

Autism and Fragile X Syndrome

89

90

91

92

mGlu5 inhibition corrects fragile X in adult mice. Neuron 2012; 74(1):49–56 Berry-Kravis E, Sumis A, Hervey C, et al. Open-label treatment trial of lithium to target the underlying defect in fragile X syndrome. J Dev Behav Pediatr 2008;29(4):293–302 Leigh MJS, Nguyen DV, Mu Y, et al. A randomized double-blind, placebo-controlled trial of minocycline in children and adolescents with fragile x syndrome. J Dev Behav Pediatr 2013;34(3): 147–155 Berry-Kravis EM, Hessl D, Rathmell B, et al. Effects of STX209 (arbaclofen) on neurobehavioral function in children and adults with fragile X syndrome: a randomized, controlled, phase 2 trial. Sci Transl Med 2012;4 )152):152 ra127 Erickson CA, Wink LK, Ray B, et al. Impact of acamprosate on behavior and brain-derived neurotrophic factor: an open-label

93

94

95

96

265

study in youth with fragile X syndrome. Psychopharmacology (Berl) 2013;228(1):75–84 Berry-Kravis E, Hessl D, Coffey S, et al. A pilot open label, single dose trial of fenobam in adults with fragile X syndrome. J Med Genet 2009;46(4):266–271 Jacquemont S, Curie A, des Portes V, et al. Epigenetic modification of the FMR1 gene in fragile X syndrome is associated with differential response to the mGluR5 antagonist AFQ056. Sci Transl Med 2011;3(64):64ra1 Berry-Kravis E, Hessl D, Abbeduto L, Reiss AL, Beckel-Mitchener A, Urv TK; Outcome Measures Working Groups. Outcome measures for clinical trials in fragile X syndrome. J Dev Behav Pediatr 2013; 34(7):508–522 Waltes R, Duketis E, Knapp M, et al. Common variants in genes of the postsynaptic FMRP signalling pathway are risk factors for autism spectrum disorders. Hum Genet 2014;133(6):781–792

This document was downloaded for personal use only. Unauthorized distribution is strictly prohibited.

88 Michalon A, Sidorov M, Ballard TM, et al. Chronic pharmacological

Yu, Berry-Kravis

Seminars in Neurology

Vol. 34

No. 3/2014

Copyright of Seminars in Neurology is the property of Thieme Medical Publishing Inc. and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.

Autism and fragile X syndrome.

Autistic spectrum disorders (ASDs) are characterized by impairments in language, social skills, and repetitive behaviors, often accompanied by intelle...
160KB Sizes 3 Downloads 19 Views