CHAPTER TWO

Contribution of Long Noncoding RNAs to Autism Spectrum Disorder Risk Brent Wilkinson*,†, Daniel B. Campbell†,{,1

*Program in Biological and Biomedical Sciences, University of Southern California, Los Angeles, California, USA † Zilkha Neurogenetic Institute, University of Southern California, Los Angeles, California, USA { Department of Psychiatry and the Behavioral Sciences, University of Southern California, Los Angeles, California, USA 1 Corresponding author: e-mail address: [email protected]

Contents 1. 2. 3. 4. 5.

Small ncRNA Long ncRNA LncRNA in Fundamental Genetic Mechanisms LncRNAs in Cancer LncRNAs in the Brain 5.1 Brain development and function 5.2 Neurodegenerative diseases 5.3 Neurodevelopmental disorders 6. LncRNAs Contribute to ASD 7. Conclusions References

36 38 38 40 41 41 44 45 47 48 49

Abstract Accumulating evidence indicates that long noncoding RNAs (lncRNAs) contribute to autism spectrum disorder (ASD) risk. Although a few lncRNAs have long been recognized to have important functions, the vast majority of this class of molecules remains uncharacterized. Because lncRNAs are more abundant in human brain than proteincoding RNAs, it is likely that they contribute to brain disorders, including ASD. We review here the known functions of lncRNAs and the potential contributions of lncRNAs to ASD.

Projects aiming to characterize all of the functional elements in the mammalian genome such as ENCODE (Djebali et al., 2012) and FANTOM3 (Carninci et al., 2005) have made significant strides in our understanding of the complexity of transcriptional regulation. It is now clear that while less

International Review of Neurobiology, Volume 113 ISSN 0074-7742 http://dx.doi.org/10.1016/B978-0-12-418700-9.00002-2

#

2013 Elsevier Inc. All rights reserved.

35

36

Brent Wilkinson and Daniel B. Campbell

than 2% of the mammalian genome codes for proteins, a majority of what used to be thought of as “junk DNA” actually undergoes transcription and produces noncoding RNAs (ncRNAs) (Bertone et al., 2004; Birney et al., 2007; Carninci et al., 2005, 2006; Cheng et al., 2005; Core, Waterfall, & Lis, 2008; Djebali et al., 2012; Kapranov et al., 2007; Lander et al., 2001; Seila et al., 2008). Emergence of the functional characteristics of ncRNA has shown that these regulatory RNAs play several critical roles in modulating gene expression and influence the developmental complexity of higher species. Considering the proportion of ncRNA relative to total genome size, there is a significant correlation between increased ncRNA in a species with developmental complexity (Taft, Pheasant, & Mattick, 2007). Because of this role, it can be inferred that ncRNA will take part in many critical functions in the most complex organ of the body, the brain. NcRNA can be broadly categorized into two classes based on their size: short ncRNAs (200 nucleotides in length) (Kapranov et al., 2007), both of which contain a number of diverse subclasses with their own distinct properties. Currently, relatively few ncRNAs have been functionally characterized compared to the large amount shown to undergo transcription. But for those that have, it is clear that many are integral to the processes of development and maintenance of an organism. Being essential components of regulatory networks within our cells also means that when they become aberrant, they can influence disease. Indeed, ncRNAs have been implicated in a number of conditions including cancer, HIV, heart disease, and neurological disorders. This highlights the critical importance of characterizing ncRNAs which may provide a pathway to discover routes of pathogenesis that are currently unknown and improve on those that are. Here, we outline the current progress of characterizing ncRNAs and their relationship to several neurological disorders. Specifically, we show that a number of ncRNAs have been identified in autism spectrum disorder (ASD) and associated neurodevelopmental disorders, may play critical roles in these, and may serve as potential therapeutic targets.

1. SMALL ncRNA The classification, small ncRNA, encompasses a wide variety of subclasses including microRNAs (miRNAs), short interfering RNAs (siRNAs), small nucleolar RNAs (snoRNAs), and PIWI-interacting RNAs (piRNAs). These can be derived from a variety of sources within the

LncRNAs in Autism

37

genome (review in Mattick & Makunin, 2005) and it has been suggested that one potential source is also from the processing of long noncoding RNAs (lncRNAs) (Jalali, Jayaraj, & Scaria, 2012; Keniry et al., 2012). In general, one of the most prominent forms of genetic regulation carried out by small ncRNAs is by binding to their specific target transcripts via complementary nucleotide sequences and subsequently influencing the translation of the target through a diverse set of downstream events. For an in-depth explanation of small ncRNA, including subclasses, biogenesis, and forms of genetic regulation, we refer the reader to these reviews: Carthew and Sontheimer (2009), Kim, Han, and Siomi (2009), Luteijn and Ketting (2013), Matera, Terns, and Terns (2007), Mattick and Makunin (2005), and Winter, Jung, Keller, Gregory, and Diederichs (2009). One of the most thoroughly studied classes of small ncRNAs, miRNAs, was first discovered in Caenorhabditis elegans in 1993 (Lee, Feinbaum, & Ambros, 1993) and since then has been postulated to be relatively conserved among vertebrate species (Altuvia et al., 2005; Berezikov et al., 2005; Iba´n˜ez-Ventoso, Vora, & Driscoll, 2008). MiRNAs can participate in posttranscriptional gene regulation by binding to either the 30 -UTR (Lai, 2002) or the 50 -UTR (Lytle, Yario, & Steitz, 2007) of a target transcript which subsequently inhibits its translation. A single miRNA has the potential to target multiple transcripts and similarly, a given transcript may be regulated by more than one miRNA (Krek et al., 2005). Both the importance and prevalence of miRNAs in the human genome is underscored by the fact that greater than 60% of human protein-coding genes are predicted targets of miRNAs (Friedman, Farh, Burge, & Bartel, 2009). The regulation performed by miRNAs is a critical process in the development of mammals and like many other ncRNAs, their expression can be specific in both particular tissue types and developmental stages as reviewed in AlvarezGarcia and Miska (2005), Sayed and Abdellatif (2011), and Wienholds and Plasterk (2005). Like miRNA, siRNA also regulates gene expression by binding to complementary sequences on target transcripts, but with more stringent requirements as they have to have close to perfect sequence complementation (Elbashir, Martinez, Patkaniowska, Lendeckel, & Tuschl, 2001). In general, this difference results in two separate mechanisms where targets with close to perfect complementation undergo direct cleavage (as with siRNA) and those with relatively loose complementation are destabilized or transcriptionally repressed (as with miRNA) (Guo, Ingolia, Weissman, & Bartel, 2010; Hutva´gner & Zamore, 2002; Martinez, Patkaniowska, Urlaub,

38

Brent Wilkinson and Daniel B. Campbell

Lu¨hrmann, & Tuschl, 2002). Endogenous siRNAs (endo-siRNAs) in Drosophila have been proposed to protect against dsDNA viruses and transposable elements in somatic tissues (Chung, Okamura, Martin, & Lai, 2008; Li, Li, & Ding, 2002; Wang et al., 2006). Endo-siRNAs have also been identified in human cells (Chan et al., 2013; Xia, Joyce, Bowcock, & Zhang, 2013; Yang & Kazazian, 2006) and in mice (Babiarz, Ruby, Wang, Bartel, & Blelloch, 2008; Song et al., 2011; Tam et al., 2008; Watanabe et al., 2008), showing another layer of genetic regulation in mammals. siRNA has been explored in numerous settings as a therapeutic agent used for gene therapy and has potential for correcting the dysregulation of ncRNA in several different diseases (Burnett & Rossi, 2012).

2. LONG ncRNA LncRNA is a broad category which encompasses transcripts of diverse structural features and mechanisms of action. They can be derived from sense or antisense strands overlapping with protein-coding genes, within intergenic regions (lincRNAs), or within pseudogenes. Once transcribed they may undergo splicing and be processed to include a 50 methyl-guanosine cap and 30 -poly (A) tail. Within the cell, lncRNAs can be localized to the nucleus or the cytoplasm (Ponting, Oliver, & Reik, 2009). They have been found to perform a wide variety of functions including participating in the recruitment of chromatin-modifying complexes, acting as competing endogenous RNAs (ceRNAs), providing a scaffold for the assembly of protein complexes, modulating alternative splicing, and employing enhancer-like functions (Nagano & Fraser, 2011; Rinn & Chang, 2012; Wang & Chang, 2011). This diverse set of functions further emphasizes the notion that lncRNAs are key components of numerous cellular processes (Table 2.1).

3. LncRNA IN FUNDAMENTAL GENETIC MECHANISMS X-chromosome inactivation (XCI) is essential to the identity of female mammals and is a highly intricate process involving multiple lncRNAs that work together in order to ultimately downregulate mass quantities of genes for dosage compensation. X-inactive specific transcript (Xist) is a conserved lncRNA specifically expressed in the inactive X-chromosome and required for XCI (Brown et al., 1992; Brown, 1991). This was confirmed in XX murine embryonic stem cells (mESCs)

39

LncRNAs in Autism

Table 2.1 LncRNAs implicated in neurodevelopmental and neurodegenerative disorders lncRNA Function Reference

ASFMR1

Unknown, co-expressed with FMR1 Ladd et al. (2007)

ATXN8OS Unknown

Moseley et al. (2006)

BACE1-AS Regulation of BACE1 and roles in amyloid plaque formation

Faghihi et al. (2008)

BC200

Initiation of protein translation in dendritic processes

Muddashetty et al. (2002), Mus, Hof, and Tiedge (2007)

BDNFOS

Regulation of BDNF

Lipovich et al. (2012)

DISC2

Unknown, antisense to DISC1

Millar et al. (2000)

EVF2

Involvement in GABAergic interneuron function

Bond et al. (2009), Feng et al. (2006)

FMR4

Antiapoptic functions, co-expressed with FMR1

Khalil, Faghihi, Modarresi, Brothers, and Wahlestedt (2008)

Gomafu

Alternative splicing of DISC1 and ERBB4

Barry et al. (2013)

HAR1F

Unknown, expressed in Cajal-Retzius Pollard et al. (2006) neurons of developing human neocortex

MALAT1

Regulation of alternative splicing and Bernard et al. (2010) synaptogenesis

MSNP1AS Regulation of Moesin

Kerin et al. (2012)

SOX2OT

Neurogenesis

Amaral et al. (2009)

UBE3AATS

Genomic imprinting of UBE3A

Meng, Person, and Beaudet (2012)

by introducing a targeted deletion of Xist into a single allele. Following differentiation, the targeted X chromosome would always fail to inactivate and only the X chromosome expressing Xist would undergo XCI (Penny, Kay, Sheardown, Rastan, & Brockdorff, 1996). Xist acts by coating the chromosome to be inactivated and then recruiting polycomb-group proteins for inactivation via epigenetic mechanisms (Plath et al., 2003). By manipulating the expression of Xist, selective silencing of one copy of chromosome 21 was achieved in induced pluripotent stem cells (iPSCs) derived from patients

40

Brent Wilkinson and Daniel B. Campbell

with Down’s syndrome. In the iPSC model, this corrected for the trisomy of chromosome 21 involved in the pathogenesis of Down’s syndrome and improved deficiencies in proliferation and neural rosette formation (Jiang et al., 2013). In addition, it was recently discovered that there is also a human specific lncRNA, XACT, that coats the active X chromosome in pluripotent cells (Vallot et al., 2013). Xist displays a common trait among lncRNAs as they can associate with chromatin and recruit proteins with epigenetic functions in order to regulate the transcription of multiple genes. LncRNAs have also been shown to be involved in genomic imprinting, the process by which genes are expressed in a parent-specific manner. The paternally expressed lncRNA, Antisense Igf 2r (Air), silences three proteincoding genes (Igf2r, Slc22a2, and Slc22a3) located within the Igf2r cluster on the same allele (Sleutels, Zwart, & Barlow, 2002). Like Xist, Air is conserved between mice and humans (Yotova et al., 2008) and is involved in epigenetic regulation as it inactivates the Slc22a3 gene by recruiting G9a (a histone methyltransferase) (Nagano et al., 2008). Factors influencing genomic imprinting are of extreme importance as dysregulation of this process is hallmark of a number of conditions, including the neurodevelopmental disorders, Angelman Syndrome (AS), and Prader-Willi Syndrome (PWS) (Horsthemke & Wagstaff, 2008).

4. LncRNAs IN CANCER Downregulation of the tumor suppressor, PTEN, has been associated with numerous types of cancers (Cairns et al., 1997; Li et al., 1997; Vlietstra, van Alewijk, Hermans, van Steenbrugge, & Trapman, 1998). The lncRNA, PTENpg1 (also known as PTENP1), is selectively lost in cancer, has similar expression levels to that of PTEN, and functions as an miRNA decoy by intercepting miRNA species targeting PTEN (Poliseno et al., 2010). Two PTENpg1 antisense (as) isoforms, a and b, regulate the expression levels of PTEN by diverse mechanisms, illustrating the complex networks of regulation lncRNAs can take part in. PTENpg1 asRNA a is able to decrease PTEN expression levels through epigenetic mechanisms involving DNMT3A and EZH2, while PTENpg1 asRNA b interacts with PTENpg1 in order to positively influence its stability and decoy function ( Johnsson et al., 2013). Interestingly, abnormalities in PTEN have also been implicated in ASD (Butler et al., 2005; O’Roak et al., 2012). The HOX genes are essential for specifying patterning during the development of bilateral animals (Pearson, Lemons, & McGinnis, 2005).

LncRNAs in Autism

41

HOTAIR is an lncRNA transcribed from the HOXC locus that acts in trans to regulate transcription at the HOXD locus. Here, HOTAIR associates with Polycomb Repressive Complex 2 (PRC2) in order to influence Histone H3 Lysine-27 trimethylation which results in transcriptional silencing (Rinn et al., 2007). Increased expression of HOTAIR has been associated with a variety of cancer types including breast (Gupta et al., 2010), colon (Kogo et al., 2011), and liver (Ishibashi et al., 2013) cancers. This may be due to massive epigenetic changes in which increased HOTAIR reverts the epigenetic profile of cancerous cells to something resembling that of embryonic fibroblasts through retargeting of PRC2 (Gupta et al., 2010). Maternally expressed gene 3 (MEG3) is a conserved, imprinted gene encoding an lncRNA. In meningiomas, there is a strong association between loss of MEG3 expression and increase in tumor grade (Zhang et al., 2010). This can be attributed to the activation of the p53 pathway (Zhao, Dahle, Zhou, Zhang, & Klibanski, 2005) and increased methylation of MEG3 regulatory region. The effect of decreased expression of MEG3 has also been implicated in pituitary adenomas (Zhang et al., 2003), gliomas (Wang, Ren, & Sun, 2012), cervical cancer (Qin et al., 2013), and bladder cancer (Ying et al., 2013).

5. LncRNAs IN THE BRAIN LncRNAs have been shown to be highly expressed within the central nervous system and in particular, the brain, where they can exhibit spatiotemporal expression patterns (Lipovich et al., 2013; Ponjavic, Oliver, Lunter, & Ponting, 2009). This highly diverse class of ncRNAs has been shown to be involved in several key roles of brain development and function. Therefore, dysregulation of lncRNAs can be contributing factors to neurological disorders, which we will show examples with respect to neurodevelopmental and neurodegenerative disorders. For many complex neurological disorders such as autism, the pathogenesis is quite unclear. The functional characterization of lncRNAs could unveil another layer of transcriptional regulation involved in their pathogenesis and potentially provide routes of therapeutic intervention.

5.1. Brain development and function LncRNAs are essential to the development, maintenance, and function of the brain. They have been shown to take part in fundamental processes such as synaptogenesis, neurogenesis, and GABAergic interneuron function.

42

Brent Wilkinson and Daniel B. Campbell

Abnormalities in these processes have been implicated in several neurodevelopmental disorders including ASD and schizophrenia (SZ). Dysregulation of lncRNAs involved in these processes may ultimately impact the molecular mechanisms that underlie the observed phenotypes in these disorders. Neurogenesis, the differentiation of neurons from neural stem cells or neural progenitor cells, is a critical process that occurs throughout the life of an individual (Ming & Song, 2005). Studies analyzing the differential expression of lncRNAs upon differentiating human ESCs or iPSCs to neurons have identified several lncRNAs as integral components of neurogenesis (Lin et al., 2011; Ng, Johnson, & Stanton, 2012). The proteincoding gene, SOX2, has been shown to play key roles in both embryonic and adult neurogenesis (Ellis et al., 2004; Favaro et al., 2009; Ferri et al., 2004) and is located within an intron of the lncRNA Sox2 overlapping transcript (SOX2OT) (Fantes et al., 2003). Sox2ot and an alternatively spliced isoform, Sox2dot, are both expressed in the mouse brain and enriched in areas associated with neurogenesis (Amaral et al., 2009; Mercer, Dinger, Sunkin, Mehler, & Mattick, 2008). Defects in neurogenesis during development and adulthood have been linked to a number of neurodevelopmental and neurodegenerative diseases (Amiri et al., 2012; Guidi et al., 2008; Hsieh & Eisch, 2010; Reif et al., 2006; Wegiel et al., 2010). In the human brain, synaptogenesis is characterized by increased proliferation of neuronal cells and an overproduction of synaptic connections from gestation to about 3 years of age, followed by subsequent “trimming” of these into adulthood (Bourgeron, 2009; Huttenlocher & Dabholkar, 1997; Petanjek et al., 2011). This timeline has many important implications in neurodevelopmental disorders, as ASD typically presents itself prior to the age of 3 (Bourgeron, 2009; Investigators & Prevention, 2012). Metastasisassociated lung-adenocarcinoma transcript 1 (MALAT1) was originally identified as being overexpressed in non-small cell lung cancer ( Ji et al., 2003) and is also highly expressed in neurons where it plays key roles in synaptogenesis. This lncRNA was shown to regulate synaptic density and the expression levels of neuroligin1 (NLGN1) and synaptic cell-adhesion molecule (SynCAM1), which are involved in controlling synapse formation (Bernard et al., 2010). As with neurogenesis, aberrant regulation of synaptogenesis is a common theme among many neurological disorders including ASD and SZ (Bourgeron, 2009; Grant, 2012; Zoghbi, 2003). In addition, MALAT1 takes part in the recruitment of SR-type pre-mRNA splicing factors to nuclear speckle domains where they participate in

LncRNAs in Autism

43

regulating alternative splicing (Tripathi et al., 2010). This is particularly relevant to the human brain, which contains one of the highest proportions of alternatively spliced transcripts (de la Grange, Gratadou, Delord, Dutertre, & Auboeuf, 2010; Grosso et al., 2008). Dysregulation of alternative splicing can impact the expression levels of large quantities of transcripts and has been implicated in SZ and ASD (Morikawa & Manabe, 2010; Voineagu et al., 2011). GABA is one of the most abundant neurotransmitters in the brain and has key roles in development (Wonders & Anderson, 2006). Because of this, irregular GABAergic interneuron function has been linked to ASD (Fatemi, Folsom, Kneeland, & Liesch, 2011; Hogart, Nagarajan, Patzel, Yasui, & Lasalle, 2007; Horder et al., 2013) and SZ (Lewis, Hashimoto, & Volk, 2005). Dlx homeobox genes are critical for the differentiation and migration of GABAergic interneurons in the developing brain (Anderson, Eisenstat, Shi, & Rubenstein, 1997; Kuwajima, Nishimura, & Yoshikawa, 2006). Evf2 is an lncRNA transcribed from the Dlx-5/6 ultraconserved region that recruits DLX and MECP2 transcription factors to this same region in order to influence the expression of Dlx5, Dlx6, and Gad1. This occurs through a combination of both cis and trans mechanisms (Bond et al., 2009) and illustrates how lncRNAs are able to regulate specific targets through diverse mechanisms. Evf2 knockout mice show reduced numbers of GABAergic interneurons in the hippocampus and dentate gyrus during infancy. Although the quantity of interneurons returns to normal in adults, defects in synaptic connectivity remain (Bond et al., 2009). The early dysregulation in Evf2 having a long-lasting influence highlights the importance of lncRNA in neurodevelopmental disorders. BDNF, the most abundant neurotrophin in the brain, takes part in several fundamental functions including the regulation of neuron morphology, neuronal cell survival, and neuronal plasticity and memory (Egan et al., 2003; Horch & Katz, 2002; Tanaka et al., 2008). Due to its diverse roles in fundamental processes, BDNF has been linked to a wide range of neurodevelopmental and neurodegenerative disorders including SZ (Krebs et al., 2000; Neves-Pereira et al., 2005; Nieto, Kukuljan, & Silva, 2013), Alzheimer’s disease (AD) (Hock, Heese, Hulette, Rosenberg, & Otten, 2000), Parkinson’s disease (Howells et al., 2000), and ASD (Gadow, Roohi, DeVincent, Kirsch, & Hatchwell, 2009; Katoh-Semba et al., 2007; Ricci et al., 2013). BDNFOS or anti-BDNF is a conserved natural antisense transcript that forms dsRNA duplexes in the brain in order to downregulate BDNF transcript levels (Liu et al., 2006; Pruunsild,

44

Brent Wilkinson and Daniel B. Campbell

Kazantseva, Aid, Palm, & Timmusk, 2007). In neocortical regions of brain tissue removed from patients to treat intractable seizures, an increase in BDNF expression along with a reciprocal decrease in BDNFOS expression has been observed (Lipovich et al., 2012). This same trend was seen in human neuronal cells after repeated depolarization that mimicked the effects of epileptic seizures. In addition, the lncRNAs RPPH1, NEAT1, and MALAT1 were also upregulated in both experimental settings, showing that lncRNAs can have activity-dependent expression (Lipovich et al., 2012).

5.2. Neurodegenerative diseases In addition to contributing to the fundamental processes of development, lncRNAs can also be causative factors in diseases characterized by rapid decline of the brain. BACE1 is involved in cleaving APP in order to generate the toxic Ab peptides that contribute to the pathogenesis of AD (Cai et al., 2001) and decreased expression of BACE1 leads to reduced levels of Ab (Atwal et al., 2011; Singer et al., 2005). BACE1-antisense transcript (BACE1-AS) is upregulated in both a transgenic mouse model and human patients with AD. With increased BACE1-AS, there is a concurrent upregulation of BACE1 which is proposed to be due to stabilization of BACE1 by BACE1-AS (Faghihi et al., 2008). This is in contrast to the mechanisms proposed for other protein-coding genes and their corresponding antisense transcripts such as BDNF/BDNFOS and UBE3A/UBE3-ATS where the antisense transcript imposes downregulation of the protein-coding gene (Lipovich et al., 2012; Meng et al., 2012). Upon exposure to Ab 1–42, expression of both BACE1 and BACE1-AS increases which subsequently acts to further increase the levels of Ab 1–42. This can lead to the formation of amyloid plaques and progression of AD (Faghihi et al., 2008). BC200 RNA is a brain-specific ncRNA that is homologous to the rodent BC1 RNA (Martignetti & Brosius, 1993; Tiedge, Chen, & Brosius, 1993). In the dendritic processes of neurons, BC200 RNA associates with poly(A)-binding protein (PABP1) and plays a role in regulating the initiation of protein translation (Muddashetty et al., 2002). Expression of BC200 RNA is upregulated in selectively vulnerable areas of the brain in AD. This upregulation increases with severity of AD and skews the distribution of BC200 RNA, causing it to be clustered in the cell soma (Mus et al., 2007).

LncRNAs in Autism

45

5.3. Neurodevelopmental disorders Silencing of the Fragile X Mental Retardation Gene (FMR1) resulting from extensive 50 -UTR CGG trinucleotide repeats is a causative factor in the most common form of inherited mental retardation, fragile X syndrome (FXS). Normal FMR1 genes contain 5–54 repeats, 55–200 repeats is categorized as the premutation allele, and above 200 repeats is termed the full mutation allele and has a direct link with FXS. The premutation allele confers increased expression of FMR1 mRNA and is associated with fragile X-associated tremor and ataxia syndrome (FXTAS) while the full mutation allele silences the transcription of FMR1 (Garber, Visootsak, & Warren, 2008). Two lncRNAs, ASFMR1 and FMR4, are transcribed from the FMR1 locus and have been shown to have similar patterns of expression with respect to FMR1 in both the premutation and full mutation alleles. Although the function of ASFMR1 is currently unknown, FMR4 has been shown to have antiapoptotic properties in vitro (Khalil et al., 2008; Ladd et al., 2007). In addition, carriers of the premutation allele have been associated with mitochondrial dysfunction (Ross-Inta et al., 2010), which may predispose individuals to neurodegenerative disorders such as Parkinson’s disease (Loesch et al., 2011). Based on the similar expression patterns of both ASFMR1 and FMR4, there is a significant possibility of their contribution to the pathogenesis of the FXS and FXTAS. Like ASD, SZ is a complex disorder thought to be the result of a variety of genetic and environmental influences (Sullivan, Kendler, & Neale, 2003). The DISC locus was originally identified in a large Scottish family as a candidate gene for SZ and contains DISC1 along with a human-specific lncRNA antisense to DISC1 and DISC2 (Millar et al., 2000; Taylor, Devon, Millar, & Porteous, 2003). Recently, it was shown that the lncRNA, Gomafu, is involved in the alternative splicing of DISC1 and another SZ-associated gene, ERBB4. Gomafu is downregulated in an activity-dependent manner in response to depolarization in both mouse and human neuronal cell lines. Knockdown of Gomafu results in increased expression of alternatively spliced isoforms of DISC1 and ERBB4 but not the unspliced genes; this is the same expression pattern previously observed in postmortem brain tissue of schizophrenic patients (Barry et al., 2013; Law, Kleinman, Weinberger, & Weickert, 2007; Nakata et al., 2009). Importantly, decreased expression of Gomafu was also observed in postmortem brain tissue of individuals with SZ, implicating this lncRNA in the pathogenesis of SZ and as a potential therapeutic target (Barry et al., 2013).

46

Brent Wilkinson and Daniel B. Campbell

UBE3A-ATS is an lncRNA antisense to UBE3A within the 15q11–13 chromosomal region (Rougeulle, Cardoso, Fonte´s, Colleaux, & Lalande, 1998), which is termed the PWS/AS region due to chromosomal abnormalities arising here being causative factors to these disorders (Nicholls & Knepper, 2001). The failure to inherit/express a maternal copy of UBE3A is known to cause AS in a majority of the cases (Kishino, Lalande, & Wagstaff, 1997). UBE3A is imprinted in a neuron-specific manner, where it is only expressed from the maternal allele due to silencing of UBE3A on the paternal allele by the lncRNA, UBE3A-ATS (Meng et al., 2012; Yamasaki et al., 2003). This suggests UBE3-ATS may also be a useful therapeutic target as dysregulation of UBE3A-ATS can cause abnormal expression of paternal UBE3A and contribute to the pathogenesis of AS. Due to the tissue- and cell type-specific expression of many lncRNAs, they have the potential to be utilized as biomarkers in order to monitor developmental changes within the brain. Rhabdomyosarcoma 2-associated transcript (RMST ) was originally identified as differentially expressed in various tumor types (Chan, Thorner, Squire, & Zielenska, 2002). It has subsequently been shown that the expression of this conserved lncRNA is mainly restricted to the CNS (Chodroff et al., 2010) and is a marker for midbrain dopaminergic neurons in mice (Uhde, Vives, Jaeger, & Li, 2010). Although a function in human neurons is yet to be elucidated, a recent genome-wide association study identified RMST as a risk gene for severe obesity in which the other loci having significant association were involved in neuronal regulation of energy homeostasis (Wheeler et al., 2013). While some lncRNAs are conserved among species, a majority of these are not in comparison to the prevalence of conserved protein-coding genes (Mercer, Dinger, & Mattick, 2009; Pang, Frith, & Mattick, 2006). Those that are human specific may be linked to human complexity and elucidate pathways involved in complex disorders. By comparing the human genome against that of the chimpanzee and looking for regions that display accelerated evolutionary change, Pollard et al. identified “human accelerated region” (HAR1). HAR1 is part of a pair of overlapping transcripts termed HAR1F and HAR1R. HAR1F displays both time- and cell type-specific expression in Cajal–Retzius neurons of the developing human neocortex that is not detectable past 24 gestational weeks (Pollard et al., 2006). This developmental specificity, or lncRNAs with expression patterns like this, could be important in neurodevelopmental disorders that have defects present from early on such as ASD or SZ. In fact, within the Cajal–Retzius neurons, HAR1F is coexpressed with reelin (Pollard et al., 2006), which has

LncRNAs in Autism

47

been implicated in SZ (Eastwood & Harrison, 2006; Impagnatiello et al., 1998) and ASD (Ashley-Koch et al., 2007; Persico et al., 2001). The involvement of lncRNA in contributing to the pathogenesis of neurodevelopmental disorders is limited to a handful of examples. But, these examples do show that lncRNAs play critical roles in human development and emphasize the need for further characterization of others. The fact that several of these disorders can have comorbid diagnoses with one another points to the possibility of some common molecular pathways that could involve lncRNA. For example, some patients diagnosed with AS also have mutations in the MECP2 gene responsible for a majority of Rett Syndrome cases (Samaco, Hogart, & LaSalle, 2005; Watson et al., 2001).

6. LncRNAs CONTRIBUTE TO ASD Direct evidence for a contribution of lncRNAs to ASD continues to accumulate. Differential expression of lncRNAs has been observed in both postmortem brain tissue and lymphoblastoid cell lines. Recently, we reported that a genome-wide significant association signal implicated an lncRNA, not the neighboring protein-coding genes. To date there has only been one study characterizing lncRNA expression profiles in postmortem tissues of individuals with ASD. Ziat and colleges detected over 222 differentially expressed lncRNAs between individuals with ASD compared to controls. Within these 222 lncRNAs, 82 were unique to the prefrontal cortex (PFC), while 143 were unique to the cerebellum (Ziats & Rennert, 2013). This observation again underlies the fact that lncRNA expression can be highly tissue specific. They also reported increased transcriptional homogeneity between the PFC and cerebellum of ASD brain tissue when compared to controls in both mRNA and annotated lncRNA (1375 differentially expressed lncRNAs in control samples versus 236 in the ASD samples) (Ziats & Rennert, 2013). Although the conclusions drawn from these results are exciting, the sample size of brain tissue in this study is relatively small (n ¼ 2 ASD patients and n ¼ 2 age, sex-matched controls). More studies of this nature will have to be repeated with a larger sample size in order to assess variability and determine the significance of these conclusions. A scan for differential expression of transcripts in LCLs derived from three subgroups of individuals diagnosed with ASD identified 20 common lncRNAs that were dysregulated in all of the subgroups compared to controls. A majority of the lncRNAs identified were also shown to be

48

Brent Wilkinson and Daniel B. Campbell

androgen-responsive, suggesting a gender-specific component (Hu, 2013; Hu et al., 2009). A genome-wide association study (GWAS) of ASD indicated genomewide significant association (P ¼ 1010) of rs4307059 on chromosome 5p14.1 (Wang et al., 2009). The same rs4307059 allele was also associated with social communication phenotypes in a general population sample (St Pourcain et al., 2010). However, rs4307059 genotype was not correlated with expression of either of the flanking protein-coding genes, CDH9 and CDH10 (Wang et al., 2009). We identified a 3.9 kb noncoding RNA that is transcribed directly at the site of the chromosome 5p14.1 ASD GWAS peak (Kerin et al., 2012). The noncoding RNA is encoded by the opposite (antisense) strand of moesin pseudogene 1 (MSNP1) and is thus designated MSNP1AS (moesin pseudogene 1, antisense). MSNP1AS is 94% identical and antisense to the X chromosome transcript MSN, which encodes a protein (moesin) that regulates neuronal architecture and immune response. Expression of MSNP1AS in postmortem temporal cortex is increased 12.7-fold in individuals with ASD and increased 22-fold in individuals with the rs4307059 risk allele (Kerin et al., 2012). The MSNP1AS noncoding RNA binds MSN and its overexpression in cultured neurons causes significant decreases in MSN transcript, moesin protein, neurite number, and neurite length. Thus, our discovery reveals a functional lncRNA which, based on the GWAS findings, contributes to ASD risk (Fig. 2.1).

7. CONCLUSIONS The Central Dogma of Molecular Biology posits that DNA is transcribed into RNA, which is translated into protein. Genetic information is stored in protein-coding genes, while RNA is merely an intermediary between genes and functional proteins. Prior to the completion of the Human Genome Project, the prevailing hypothesis was that the human genome would produce approximately 100,000 protein-coding genes. This seemed like a reasonable estimate based on the size of the human genome and the complexity of human anatomy. However, the human genome contains only approximately 21,000 protein-coding genes, slightly more than a mouse and slightly less than a grape. Over the last decade, RNA has become increasingly recognized as a functional entity. The whole genome and transcriptome sequencing suggests that the complexity of an organism may be regulated by noncoding portions of the genome rather than by proteins.

49

LncRNAs in Autism

A

B

C

rs7704909

rs4307059

MSNP1: Chr 5p14.1 pseudogene with 94% sequence identity to Chr X gene MSN

+ Strand – Strand MSNP1AS: Chr 5p14.1 transcribed 3.9 kb non-coding anti-sense RNA

Figure 2.1 MSNP1AS maps within the chromosome 5p14.1 GWAS-significant ASDassociation peak and is the only significantly expressed transcript within 500 kb of the GWAS peak. (A) The GWAS results from Wang et al. (2009), indicating ASDassociated markers on chromosome 5p14.1. (B) Genome-wide RNA-Seq data from a variety of tissue sources indicate that a single major transcript of 4 kb is expressed within 500 kb of the GWAS peak. (C) The þ strand of this 4 kb chromosome 5p14.1 region is the pseudogene moesin-like 1 (MSNP1), which has 94% sequence identity to the X chromosome gene-encoding moesin (MSN) but does not appear to be transcribed. Instead, our data indicate that a 3.9 kb RNA is transcribed from the  strand, producing a non-protein-coding RNA that is antisense to the X chromosome geneencoding moesin (MSN). Because the chromosome 5p14.1 transcript represents the antisense of the pseudogene, we designate it MSNP1AS (moesin pseudogene 1, antisense). Data from the UCSC Genome Browser.

Among long RNAs produced in the human brain, the majority do not code for proteins but are instead lncRNAs. Increasing evidence suggests that these lncRNAs may contribute to brain disorders.

REFERENCES Altuvia, Y., Landgraf, P., Lithwick, G., Elefant, N., Pfeffer, S., Aravin, A., et al. (2005). Clustering and conservation patterns of human microRNAs. Nucleic Acids Research, 33, 2697–2706. Alvarez-Garcia, I., & Miska, E. A. (2005). MicroRNA functions in animal development and human disease. Development, 132, 4653–4662.

50

Brent Wilkinson and Daniel B. Campbell

Amaral, P. P., Neyt, C., Wilkins, S. J., Askarian-Amiri, M. E., Sunkin, S. M., Perkins, A. C., et al. (2009). Complex architecture and regulated expression of the Sox2ot locus during vertebrate development. RNA, 15, 2013–2027. Amiri, A., Cho, W., Zhou, J., Birnbaum, S. G., Sinton, C. M., McKay, R. M., et al. (2012). Pten deletion in adult hippocampal neural stem/progenitor cells causes cellular abnormalities and alters neurogenesis. Journal of Neuroscience, 32, 5880–5890. Anderson, S. A., Eisenstat, D. D., Shi, L., & Rubenstein, J. L. (1997). Interneuron migration from basal forebrain to neocortex: Dependence on Dlx genes. Science, 278, 474–476. Ashley-Koch, A. E., Jaworski, J., Ma, de, Q., Mei, H., Ritchie, M. D., Skaar, D. A., et al. (2007). Investigation of potential gene-gene interactions between APOE and RELN contributing to autism risk. Psychiatric Genetics, 17, 221–226. Atwal, J. K., Chen, Y., Chiu, C., Mortensen, D. L., Meilandt, W. J., Liu, Y., et al. (2011). A therapeutic antibody targeting BACE1 inhibits amyloid-b production in vivo. Science Translational Medicine, 3, 84ra43. Babiarz, J. E., Ruby, J. G., Wang, Y., Bartel, D. P., & Blelloch, R. (2008). Mouse ES cells express endogenous shRNAs, siRNAs, and other Microprocessor-independent, Dicerdependent small RNAs. Genes and Development, 22, 2773–2785. Barry, G., Briggs, J. A., Vanichkina, D. P., Poth, E. M., Beveridge, N. J., Ratnu, V. S., et al. (2013). The long non-coding RNA Gomafu is acutely regulated in response to neuronal activation and involved in schizophrenia-associated alternative splicing. Molecular Psychiatry, http://dx.doi.org/10.1038/mp.2013.45. Berezikov, E., Guryev, V., van de Belt, J., Wienholds, E., Plasterk, R. H., & Cuppen, E. (2005). Phylogenetic shadowing and computational identification of human microRNA genes. Cell, 120, 21–24. Bernard, D., Prasanth, K. V., Tripathi, V., Colasse, S., Nakamura, T., Xuan, Z., et al. (2010). A long nuclear-retained non-coding RNA regulates synaptogenesis by modulating gene expression. The EMBO Journal, 29, 3082–3093. Bertone, P., Stolc, V., Royce, T. E., Rozowsky, J. S., Urban, A. E., Zhu, X., et al. (2004). Global identification of human transcribed sequences with genome tiling arrays. Science, 306, 2242–2246. Birney, E., Stamatoyannopoulos, J. A., Dutta, A., Guigo´, R., Gingeras, T. R., Margulies, E. H., et al. (2007). Identification and analysis of functional elements in 1% of the human genome by the ENCODE pilot project. Nature, 447, 799–816. Bond, A. M., Vangompel, M. J., Sametsky, E. A., Clark, M. F., Savage, J. C., Disterhoft, J. F., et al. (2009). Balanced gene regulation by an embryonic brain ncRNA is critical for adult hippocampal GABA circuitry. Nature Neuroscience, 12, 1020–1027. Bourgeron, T. (2009). A synaptic trek to autism. Current Opinion in Neurobiology, 19, 231–234. Brown, S. D. (1991). XIST and the mapping of the X chromosome inactivation centre. Bioessays, 13, 607–612. Brown, C. J., Hendrich, B. D., Rupert, J. L., Lafrenie`re, R. G., Xing, Y., Lawrence, J., et al. (1992). The human XIST gene: Analysis of a 17 kb inactive X-specific RNA that contains conserved repeats and is highly localized within the nucleus. Cell, 71, 527–542. Burnett, J. C., & Rossi, J. J. (2012). RNA-based therapeutics: Current progress and future prospects. Chemistry and Biology, 19, 60–71. Butler, M. G., Dasouki, M. J., Zhou, X. P., Talebizadeh, Z., Brown, M., Takahashi, T. N., et al. (2005). Subset of individuals with autism spectrum disorders and extreme macrocephaly associated with germline PTEN tumour suppressor gene mutations. Journal of Medical Genetics, 42, 318–321. Cai, H., Wang, Y., McCarthy, D., Wen, H., Borchelt, D. R., Price, D. L., et al. (2001). BACE1 is the major beta-secretase for generation of Abeta peptides by neurons. Nature Neuroscience, 4, 233–234.

LncRNAs in Autism

51

Cairns, P., Okami, K., Halachmi, S., Halachmi, N., Esteller, M., Herman, J. G., et al. (1997). Frequent inactivation of PTEN/MMAC1 in primary prostate cancer. Cancer Research, 57, 4997–5000. Carninci, P., Kasukawa, T., Katayama, S., Gough, J., Frith, M. C., Maeda, N., et al. (2005). The transcriptional landscape of the mammalian genome. Science, 309, 1559–1563. Carninci, P., Sandelin, A., Lenhard, B., Katayama, S., Shimokawa, K., Ponjavic, J., et al. (2006). Genome-wide analysis of mammalian promoter architecture and evolution. Nature Genetics, 38, 626–635. Carthew, R. W., & Sontheimer, E. J. (2009). Origins and mechanisms of miRNAs and siRNAs. Cell, 136, 642–655. Chan, A. S., Thorner, P. S., Squire, J. A., & Zielenska, M. (2002). Identification of a novel gene NCRMS on chromosome 12q21 with differential expression between rhabdomyosarcoma subtypes. Oncogene, 21, 3029–3037. Chan, W. L., Yuo, C. Y., Yang, W. K., Hung, S. Y., Chang, Y. S., Chiu, C. C., et al. (2013). Transcribed pseudogene cPPM1K generates endogenous siRNA to suppress oncogenic cell growth in hepatocellular carcinoma. Nucleic Acids Research, 41, 3734–3747. Cheng, J., Kapranov, P., Drenkow, J., Dike, S., Brubaker, S., Patel, S., et al. (2005). Transcriptional maps of 10 human chromosomes at 5-nucleotide resolution. Science, 308, 1149–1154. Chodroff, R. A., Goodstadt, L., Sirey, T. M., Oliver, P. L., Davies, K. E., Green, E. D., et al. (2010). Long noncoding RNA genes: Conservation of sequence and brain expression among diverse amniotes. Genome Biology, 11, R72. Chung, W. J., Okamura, K., Martin, R., & Lai, E. C. (2008). Endogenous RNA interference provides a somatic defense against Drosophila transposons. Current Biology, 18, 795–802. Core, L. J., Waterfall, J. J., & Lis, J. T. (2008). Nascent RNA sequencing reveals widespread pausing and divergent initiation at human promoters. Science, 322, 1845–1848. de la Grange, P., Gratadou, L., Delord, M., Dutertre, M., & Auboeuf, D. (2010). Splicing factor and exon profiling across human tissues. Nucleic Acids Research, 38, 2825–2838. Djebali, S., Davis, C. A., Merkel, A., Dobin, A., Lassmann, T., Mortazavi, A., et al. (2012). Landscape of transcription in human cells. Nature, 489, 101–108. Eastwood, S. L., & Harrison, P. J. (2006). Cellular basis of reduced cortical reelin expression in schizophrenia. The American Journal of Psychiatry, 163, 540–542. Egan, M. F., Kojima, M., Callicott, J. H., Goldberg, T. E., Kolachana, B. S., Bertolino, A., et al. (2003). The BDNF val66met polymorphism affects activity-dependent secretion of BDNF and human memory and hippocampal function. Cell, 112, 257–269. Elbashir, S. M., Martinez, J., Patkaniowska, A., Lendeckel, W., & Tuschl, T. (2001). Functional anatomy of siRNAs for mediating efficient RNAi in Drosophila melanogaster embryo lysate. The EMBO Journal, 20, 6877–6888. Ellis, P., Fagan, B. M., Magness, S. T., Hutton, S., Taranova, O., Hayashi, S., et al. (2004). SOX2, a persistent marker for multipotential neural stem cells derived from embryonic stem cells, the embryo or the adult. Developmental Neuroscience, 26, 148–165. Faghihi, M. A., Modarresi, F., Khalil, A. M., Wood, D. E., Sahagan, B. G., Morgan, T. E., et al. (2008). Expression of a noncoding RNA is elevated in Alzheimer’s disease and drives rapid feed-forward regulation of beta-secretase. Nature Medicine, 14, 723–730. Fantes, J., Ragge, N. K., Lynch, S. A., McGill, N. I., Collin, J. R., Howard-Peebles, P. N., et al. (2003). Mutations in SOX2 cause anophthalmia. Nature Genetics, 33, 461–463. Fatemi, S. H., Folsom, T. D., Kneeland, R. E., & Liesch, S. B. (2011). Metabotropic glutamate receptor 5 upregulation in children with autism is associated with underexpression of both Fragile X mental retardation protein and GABAA receptor beta 3 in adults with autism. Anatomical Record, 294, 1635–1645. Favaro, R., Valotta, M., Ferri, A. L., Latorre, E., Mariani, J., Giachino, C., et al. (2009). Hippocampal development and neural stem cell maintenance require Sox2-dependent regulation of Shh. Nature Neuroscience, 12, 1248–1256.

52

Brent Wilkinson and Daniel B. Campbell

Feng, J., Bi, C., Clark, B. S., Mady, R., Shah, P., & Kohtz, J. D. (2006). The Evf-2 noncoding RNA is transcribed from the Dlx-5/6 ultraconserved region and functions as a Dlx-2 transcriptional coactivator. Genes and Development, 20, 1470–1484. Ferri, A. L., Cavallaro, M., Braida, D., Di Cristofano, A., Canta, A., Vezzani, A., et al. (2004). Sox2 deficiency causes neurodegeneration and impaired neurogenesis in the adult mouse brain. Development, 131, 3805–3819. Friedman, R. C., Farh, K. K., Burge, C. B., & Bartel, D. P. (2009). Most mammalian mRNAs are conserved targets of microRNAs. Genome Research, 19, 92–105. Gadow, K. D., Roohi, J., DeVincent, C. J., Kirsch, S., & Hatchwell, E. (2009). Association of COMT (Val158Met) and BDNF (Val66Met) gene polymorphisms with anxiety, ADHD and tics in children with autism spectrum disorder. Journal of Autism and Developmental Disorders, 39, 1542–1551. Garber, K. B., Visootsak, J., & Warren, S. T. (2008). Fragile X syndrome. European Journal of Human Genetics, 16, 666–672. Grant, S. G. (2012). Synaptopathies: Diseases of the synaptome. Current Opinion in Neurobiology, 22, 522–529. Grosso, A. R., Gomes, A. Q., Barbosa-Morais, N. L., Caldeira, S., Thorne, N. P., Grech, G., et al. (2008). Tissue-specific splicing factor gene expression signatures. Nucleic Acids Research, 36, 4823–4832. Guidi, S., Bonasoni, P., Ceccarelli, C., Santini, D., Gualtieri, F., Ciani, E., et al. (2008). Neurogenesis impairment and increased cell death reduce total neuron number in the hippocampal region of fetuses with Down syndrome. Brain Pathology, 18, 180–197. Guo, H., Ingolia, N. T., Weissman, J. S., & Bartel, D. P. (2010). Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature, 466, 835–840. Gupta, R. A., Shah, N., Wang, K. C., Kim, J., Horlings, H. M., Wong, D. J., et al. (2010). Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature, 464, 1071–1076. Hock, C., Heese, K., Hulette, C., Rosenberg, C., & Otten, U. (2000). Region-specific neurotrophin imbalances in Alzheimer disease: Decreased levels of brain-derived neurotrophic factor and increased levels of nerve growth factor in hippocampus and cortical areas. Archives of Neurology, 57, 846–851. Hogart, A., Nagarajan, R. P., Patzel, K. A., Yasui, D. H., & Lasalle, J. M. (2007). 15q11-13 GABAA receptor genes are normally biallelically expressed in brain yet are subject to epigenetic dysregulation in autism-spectrum disorders. Human Molecular Genetics, 16, 691–703. Horch, H. W., & Katz, L. C. (2002). BDNF release from single cells elicits local dendritic growth in nearby neurons. Nature Neuroscience, 5, 1177–1184. Horder, J., Lavender, T., Mendez, M. A., O’Gorman, R., Daly, E., Craig, M. C., et al. (2013). Reduced subcortical glutamate/glutamine in adults with autism spectrum disorders: A [(1)H]MRS study. Translational Psychiatry, 3, e279. Horsthemke, B., & Wagstaff, J. (2008). Mechanisms of imprinting of the Prader-Willi/ Angelman region. American Journal of Medical Genetics Part A, 146A, 2041–2052. Howells, D. W., Porritt, M. J., Wong, J. Y., Batchelor, P. E., Kalnins, R., Hughes, A. J., et al. (2000). Reduced BDNF mRNA expression in the Parkinson’s disease substantia nigra. Experimental Neurology, 166, 127–135. Hsieh, J., & Eisch, A. J. (2010). Epigenetics, hippocampal neurogenesis, and neuropsychiatric disorders: Unraveling the genome to understand the mind. Neurobiology of Disease, 39, 73–84. Hu, V. W. (2013). The expanding genomic landscape of autism: Discovering the ‘forest’ beyond the ‘trees’. Future Neurology, 8, 29–42. Hu, V. W., Sarachana, T., Kim, K. S., Nguyen, A., Kulkarni, S., Steinberg, M. E., et al. (2009). Gene expression profiling differentiates autism case-controls and phenotypic

LncRNAs in Autism

53

variants of autism spectrum disorders: Evidence for circadian rhythm dysfunction in severe autism. Autism Research, 2, 78–97. Huttenlocher, P. R., & Dabholkar, A. S. (1997). Regional differences in synaptogenesis in human cerebral cortex. Journal of Comparative Neurology, 387, 167–178. Hutva´gner, G., & Zamore, P. D. (2002). A microRNA in a multiple-turnover RNAi enzyme complex. Science, 297, 2056–2060. Iba´n˜ez-Ventoso, C., Vora, M., & Driscoll, M. (2008). Sequence relationships among C. elegans, D. melanogaster and human microRNAs highlight the extensive conservation of microRNAs in biology. PLoS One, 3, e2818. Impagnatiello, F., Guidotti, A. R., Pesold, C., Dwivedi, Y., Caruncho, H., Pisu, M. G., et al. (1998). A decrease of reelin expression as a putative vulnerability factor in schizophrenia. Proceedings of the National Academy of Sciences of the United States of America, 95, 15718–15723. Investigators, Autism and Developmental Disabilities Monitoring Network Surveillance Year 2008 Principal, & Prevention, Centers for Disease Control (2012). Prevalence of autism spectrum disorders—Autism and Developmental Disabilities Monitoring Network, 14 sites, United States, 2008. MMWR Surveillance Summaries, 61, 1–19. Ishibashi, M., Kogo, R., Shibata, K., Sawada, G., Takahashi, Y., Kurashige, J., et al. (2013). Clinical significance of the expression of long non-coding RNA HOTAIR in primary hepatocellular carcinoma. Oncology Reports, 29, 946–950. Jalali, S., Jayaraj, G. G., & Scaria, V. (2012). Integrative transcriptome analysis suggest processing of a subset of long non-coding RNAs to small RNAs. Biology Direct, 7, 25. Ji, P., Diederichs, S., Wang, W., Bo¨ing, S., Metzger, R., Schneider, P. M., et al. (2003). MALAT-1, a novel noncoding RNA, and thymosin beta4 predict metastasis and survival in early-stage non-small cell lung cancer. Oncogene, 22, 8031–8041. Jiang, J., Jing, Y., Cost, G., Chiang, J., Kolpa, H., Cotton, A., et al. (2013). Translating dosage compensation to trisomy 21. Nature, 500, 296–300. Johnsson, P., Ackley, A., Vidarsdottir, L., Lui, W. O., Corcoran, M., Grande´r, D., et al. (2013). A pseudogene long-noncoding-RNA network regulates PTEN transcription and translation in human cells. Nature Structural and Molecular Biology, 20, 440–446. Kapranov, P., Cheng, J., Dike, S., Nix, D. A., Duttagupta, R., Willingham, A. T., et al. (2007). RNA maps reveal new RNA classes and a possible function for pervasive transcription. Science, 316, 1484–1488. Katoh-Semba, R., Wakako, R., Komori, T., Shigemi, H., Miyazaki, N., Ito, H., et al. (2007). Age-related changes in BDNF protein levels in human serum: Differences between autism cases and normal controls. International Journal of Developmental Neuroscience, 25, 367–372. Keniry, A., Oxley, D., Monnier, P., Kyba, M., Dandolo, L., Smits, G., et al. (2012). The H19 lincRNA is a developmental reservoir of miR-675 that suppresses growth and Igf1r. Nature Cell Biology, 14, 659–665. Kerin, T., Ramanathan, A., Rivas, K., Grepo, N., Coetzee, G. A., & Campbell, D. B. (2012). A noncoding RNA antisense to moesin at 5p14.1 in autism. Science Translational Medicine, 4, 128ra140. Khalil, A. M., Faghihi, M. A., Modarresi, F., Brothers, S. P., & Wahlestedt, C. (2008). A novel RNA transcript with antiapoptotic function is silenced in fragile X syndrome. PLoS One, 3, e1486. Kim, V. N., Han, J., & Siomi, M. C. (2009). Biogenesis of small RNAs in animals. Nature Reviews Molecular Cell Biology, 10, 126–139. Kishino, T., Lalande, M., & Wagstaff, J. (1997). UBE3A/E6-AP mutations cause Angelman syndrome. Nature Genetics, 15, 70–73. Kogo, R., Shimamura, T., Mimori, K., Kawahara, K., Imoto, S., Sudo, T., et al. (2011). Long noncoding RNA HOTAIR regulates polycomb-dependent chromatin

54

Brent Wilkinson and Daniel B. Campbell

modification and is associated with poor prognosis in colorectal cancers. Cancer Research, 71, 6320–6326. Krebs, M. O., Guillin, O., Bourdell, M. C., Schwartz, J. C., Olie, J. P., Poirier, M. F., et al. (2000). Brain derived neurotrophic factor (BDNF) gene variants association with age at onset and therapeutic response in schizophrenia. Molecular Psychiatry, 5, 558–562. Krek, A., Gru¨n, D., Poy, M. N., Wolf, R., Rosenberg, L., Epstein, E. J., et al. (2005). Combinatorial microRNA target predictions. Nature Genetics, 37, 495–500. Kuwajima, T., Nishimura, I., & Yoshikawa, K. (2006). Necdin promotes GABAergic neuron differentiation in cooperation with Dlx homeodomain proteins. Journal of Neuroscience, 26, 5383–5392. Ladd, P. D., Smith, L. E., Rabaia, N. A., Moore, J. M., Georges, S. A., Hansen, R. S., et al. (2007). An antisense transcript spanning the CGG repeat region of FMR1 is upregulated in premutation carriers but silenced in full mutation individuals. Human Molecular Genetics, 16, 3174–3187. Lai, E. C. (2002). Micro RNAs are complementary to 3’ UTR sequence motifs that mediate negative post-transcriptional regulation. Nature Genetics, 30, 363–364. Lander, E. S., Linton, L. M., Birren, B., Nusbaum, C., Zody, M. C., Baldwin, J., et al. (2001). Initial sequencing and analysis of the human genome. Nature, 409, 860–921. Law, A. J., Kleinman, J. E., Weinberger, D. R., & Weickert, C. S. (2007). Disease-associated intronic variants in the ErbB4 gene are related to altered ErbB4 splice-variant expression in the brain in schizophrenia. Human Molecular Genetics, 16, 129–141. Lee, R. C., Feinbaum, R. L., & Ambros, V. (1993). The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell, 75, 843–854. Lewis, D. A., Hashimoto, T., & Volk, D. W. (2005). Cortical inhibitory neurons and schizophrenia. Nature Reviews Neuroscience, 6, 312–324. Li, H., Li, W. X., & Ding, S. W. (2002). Induction and suppression of RNA silencing by an animal virus. Science, 296, 1319–1321. Li, J., Yen, C., Liaw, D., Podsypanina, K., Bose, S., Wang, S. I., et al. (1997). PTEN, a putative protein tyrosine phosphatase gene mutated in human brain, breast, and prostate cancer. Science, 275, 1943–1947. Lin, M., Pedrosa, E., Shah, A., Hrabovsky, A., Maqbool, S., Zheng, D., et al. (2011). RNA-Seq of human neurons derived from iPS cells reveals candidate long noncoding RNAs involved in neurogenesis and neuropsychiatric disorders. PLoS One, 6, e23356. Lipovich, L., Dachet, F., Cai, J., Bagla, S., Balan, K., Jia, H., et al. (2012). Activitydependent human brain coding/noncoding gene regulatory networks. Genetics, 192, 1133–1148. Lipovich, L., Tarca, A. L., Cai, J., Jia, H., Chugani, H. T., Sterner, K. N., et al. (2013). Developmental changes in the transcriptome of human cerebral cortex tissue: Long noncoding RNA transcripts. Cerebral Cortex, http://dx.doi.org/10.1093/cercor/bhs414. Liu, Q. R., Lu, L., Zhu, X. G., Gong, J. P., Shaham, Y., & Uhl, G. R. (2006). Rodent BDNF genes, novel promoters, novel splice variants, and regulation by cocaine. Brain Research, 1067, 1–12. Loesch, D. Z., Godler, D. E., Evans, A., Bui, Q. M., Gehling, F., Kotschet, K. E., et al. (2011). Evidence for the toxicity of bidirectional transcripts and mitochondrial dysfunction in blood associated with small CGG expansions in the FMR1 gene in patients with parkinsonism. Genetics in Medicine, 13, 392–399. Luteijn, M. J., & Ketting, R. F. (2013). PIWI-interacting RNAs: From generation to transgenerational epigenetics. Nature Reviews Genetics, 14, 523–534. Lytle, J. R., Yario, T. A., & Steitz, J. A. (2007). Target mRNAs are repressed as efficiently by microRNA-binding sites in the 5’ UTR as in the 3’ UTR. Proceedings of the National Academy of Sciences of the United States of America, 104, 9667–9672.

LncRNAs in Autism

55

Martignetti, J. A., & Brosius, J. (1993). BC200 RNA: A neural RNA polymerase III product encoded by a monomeric Alu element. Proceedings of the National Academy of Sciences of the United States of America, 90, 11563–11567. Martinez, J., Patkaniowska, A., Urlaub, H., Lu¨hrmann, R., & Tuschl, T. (2002). Single-stranded antisense siRNAs guide target RNA cleavage in RNAi. Cell, 110, 563–574. Matera, A. G., Terns, R. M., & Terns, M. P. (2007). Non-coding RNAs: Lessons from the small nuclear and small nucleolar RNAs. Nature Reviews Molecular Cell Biology, 8, 209–220. Mattick, J. S., & Makunin, I. V. (2005). Small regulatory RNAs in mammals. Human Molecular Genetics, 14(Spec. No. 1), R121–R132. Meng, L., Person, R. E., & Beaudet, A. L. (2012). Ube3a-ATS is an atypical RNA polymerase II transcript that represses the paternal expression of Ube3a. Human Molecular Genetics, 21, 3001–3012. Mercer, T. R., Dinger, M. E., & Mattick, J. S. (2009). Long non-coding RNAs: Insights into functions. Nature Reviews Genetics, 10, 155–159. Mercer, T. R., Dinger, M. E., Sunkin, S. M., Mehler, M. F., & Mattick, J. S. (2008). Specific expression of long noncoding RNAs in the mouse brain. Proceedings of the National Academy of Sciences of the United States of America, 105, 716–721. Millar, J. K., Wilson-Annan, J. C., Anderson, S., Christie, S., Taylor, M. S., Semple, C. A., et al. (2000). Disruption of two novel genes by a translocation co-segregating with schizophrenia. Human Molecular Genetics, 9, 1415–1423. Ming, G. L., & Song, H. (2005). Adult neurogenesis in the mammalian central nervous system. Annual Review of Neuroscience, 28, 223–250. Morikawa, T., & Manabe, T. (2010). Aberrant regulation of alternative pre-mRNA splicing in schizophrenia. Neurochemistry International, 57, 691–704. Moseley, M. L., Zu, T., Ikeda, Y., Gao, W., Mosemiller, A. K., Daughters, R. S., et al. (2006). Bidirectional expression of CUG and CAG expansion transcripts and intranuclear polyglutamine inclusions in spinocerebellar ataxia type 8. Nature Genetics, 38, 758–769. Muddashetty, R., Khanam, T., Kondrashov, A., Bundman, M., Iacoangeli, A., Kremerskothen, J., et al. (2002). Poly(A)-binding protein is associated with neuronal BC1 and BC200 ribonucleoprotein particles. Journal of Molecular Biology, 321, 433–445. Mus, E., Hof, P. R., & Tiedge, H. (2007). Dendritic BC200 RNA in aging and in Alzheimer’s disease. Proceedings of the National Academy of Sciences of the United States of America, 104, 10679–10684. Nagano, T., & Fraser, P. (2011). No-nonsense functions for long noncoding RNAs. Cell, 145, 178–181. Nagano, T., Mitchell, J. A., Sanz, L. A., Pauler, F. M., Ferguson-Smith, A. C., Feil, R., et al. (2008). The Air noncoding RNA epigenetically silences transcription by targeting G9a to chromatin. Science, 322, 1717–1720. Nakata, K., Lipska, B. K., Hyde, T. M., Ye, T., Newburn, E. N., Morita, Y., et al. (2009). DISC1 splice variants are upregulated in schizophrenia and associated with risk polymorphisms. Proceedings of the National Academy of Sciences of the United States of America, 106, 15873–15878. Neves-Pereira, M., Cheung, J. K., Pasdar, A., Zhang, F., Breen, G., Yates, P., et al. (2005). BDNF gene is a risk factor for schizophrenia in a Scottish population. Molecular Psychiatry, 10, 208–212. Ng, S. Y., Johnson, R., & Stanton, L. W. (2012). Human long non-coding RNAs promote pluripotency and neuronal differentiation by association with chromatin modifiers and transcription factors. The EMBO Journal, 31, 522–533. Nicholls, R. D., & Knepper, J. L. (2001). Genome organization, function, and imprinting in Prader-Willi and Angelman syndromes. Annual Review of Genomics and Human Genetics, 2, 153–175.

56

Brent Wilkinson and Daniel B. Campbell

Nieto, R., Kukuljan, M., & Silva, H. (2013). BDNF and schizophrenia: From neurodevelopment to neuronal plasticity, learning, and memory. Frontiers in Psychiatry, 4, 45. O’Roak, B. J., Vives, L., Fu, W., Egertson, J. D., Stanaway, I. B., Phelps, I. G., et al. (2012). Multiplex targeted sequencing identifies recurrently mutated genes in autism spectrum disorders. Science, 338, 1619–1622. Pang, K. C., Frith, M. C., & Mattick, J. S. (2006). Rapid evolution of noncoding RNAs: Lack of conservation does not mean lack of function. Trends in Genetics, 22, 1–5. Pearson, J. C., Lemons, D., & McGinnis, W. (2005). Modulating Hox gene functions during animal body patterning. Nature Reviews Genetics, 6, 893–904. Penny, G. D., Kay, G. F., Sheardown, S. A., Rastan, S., & Brockdorff, N. (1996). Requirement for Xist in X chromosome inactivation. Nature, 379, 131–137. Persico, A. M., D’Agruma, L., Maiorano, N., Totaro, A., Militerni, R., Bravaccio, C., et al. (2001). Reelin gene alleles and haplotypes as a factor predisposing to autistic disorder. Molecular Psychiatry, 6, 150–159. Petanjek, Z., Judasˇ, M., Sˇimic, G., Rasin, M. R., Uylings, H. B., Rakic, P., et al. (2011). Extraordinary neoteny of synaptic spines in the human prefrontal cortex. Proceedings of the National Academy of Sciences of the United States of America, 108, 13281–13286. Plath, K., Fang, J., Mlynarczyk-Evans, S. K., Cao, R., Worringer, K. A., Wang, H., et al. (2003). Role of histone H3 lysine 27 methylation in X inactivation. Science, 300, 131–135. Poliseno, L., Salmena, L., Zhang, J., Carver, B., Haveman, W. J., & Pandolfi, P. P. (2010). A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature, 465, 1033–1038. Pollard, K. S., Salama, S. R., Lambert, N., Lambot, M. A., Coppens, S., Pedersen, J. S., et al. (2006). An RNA gene expressed during cortical development evolved rapidly in humans. Nature, 443, 167–172. Ponjavic, J., Oliver, P. L., Lunter, G., & Ponting, C. P. (2009). Genomic and transcriptional co-localization of protein-coding and long non-coding RNA pairs in the developing brain. PLoS Genetics, 5, e1000617. Ponting, C. P., Oliver, P. L., & Reik, W. (2009). Evolution and functions of long noncoding RNAs. Cell, 136, 629–641. Pruunsild, P., Kazantseva, A., Aid, T., Palm, K., & Timmusk, T. (2007). Dissecting the human BDNF locus: Bidirectional transcription, complex splicing, and multiple promoters. Genomics, 90, 397–406. Qin, R., Chen, Z., Ding, Y., Hao, J., Hu, J., & Guo, F. (2013). Long non-coding RNA MEG3 inhibits the proliferation of cervical carcinoma cells through the induction of cell cycle arrest and apoptosis. Neoplasma, 60, 486–492. Reif, A., Fritzen, S., Finger, M., Strobel, A., Lauer, M., Schmitt, A., et al. (2006). Neural stem cell proliferation is decreased in schizophrenia, but not in depression. Molecular Psychiatry, 11, 514–522. Ricci, S., Businaro, R., Ippoliti, F., Lo Vasco, V. R., Massoni, F., Onofri, E., et al. (2013). Altered cytokine and BDNF levels in autism spectrum disorder. Neurotoxicity Research, 24, 491–501. Rinn, J. L., & Chang, H. Y. (2012). Genome regulation by long noncoding RNAs. Annual Review of Biochemistry, 81, 145–166. Rinn, J. L., Kertesz, M., Wang, J. K., Squazzo, S. L., Xu, X., Brugmann, S. A., et al. (2007). Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell, 129, 1311–1323. Ross-Inta, C., Omanska-Klusek, A., Wong, S., Barrow, C., Garcia-Arocena, D., Iwahashi, C., et al. (2010). Evidence of mitochondrial dysfunction in fragile X-associated tremor/ataxia syndrome. The Biochemical Journal, 429, 545–552.

LncRNAs in Autism

57

Rougeulle, C., Cardoso, C., Fonte´s, M., Colleaux, L., & Lalande, M. (1998). An imprinted antisense RNA overlaps UBE3A and a second maternally expressed transcript. Nature Genetics, 19, 15–16. Samaco, R. C., Hogart, A., & LaSalle, J. M. (2005). Epigenetic overlap in autism-spectrum neurodevelopmental disorders: MECP2 deficiency causes reduced expression of UBE3A and GABRB3. Human Molecular Genetics, 14, 483–492. Sayed, D., & Abdellatif, M. (2011). MicroRNAs in development and disease. Physiological Reviews, 91, 827–887. Seila, A. C., Calabrese, J. M., Levine, S. S., Yeo, G. W., Rahl, P. B., Flynn, R. A., et al. (2008). Divergent transcription from active promoters. Science, 322, 1849–1851. Singer, O., Marr, R. A., Rockenstein, E., Crews, L., Coufal, N. G., Gage, F. H., et al. (2005). Targeting BACE1 with siRNAs ameliorates Alzheimer disease neuropathology in a transgenic model. Nature Neuroscience, 8, 1343–1349. Sleutels, F., Zwart, R., & Barlow, D. P. (2002). The non-coding Air RNA is required for silencing autosomal imprinted genes. Nature, 415, 810–813. Song, R., Hennig, G. W., Wu, Q., Jose, C., Zheng, H., & Yan, W. (2011). Male germ cells express abundant endogenous siRNAs. Proceedings of the National Academy of Sciences of the United States of America, 108, 13159–13164. St Pourcain, B., Wang, K., Glessner, J. T., Golding, J., Steer, C., Ring, S. M., et al. (2010). Association between a high-risk autism locus on 5p14 and social communication spectrum phenotypes in the general population. The American Journal of Psychiatry, 167, 1364–1372. Sullivan, P. F., Kendler, K. S., & Neale, M. C. (2003). Schizophrenia as a complex trait: Evidence from a meta-analysis of twin studies. Archives of General Psychiatry, 60, 1187–1192. Taft, R. J., Pheasant, M., & Mattick, J. S. (2007). The relationship between non-proteincoding DNA and eukaryotic complexity. Bioessays, 29, 288–299. Tam, O. H., Aravin, A. A., Stein, P., Girard, A., Murchison, E. P., Cheloufi, S., et al. (2008). Pseudogene-derived small interfering RNAs regulate gene expression in mouse oocytes. Nature, 453, 534–538. Tanaka, J., Horiike, Y., Matsuzaki, M., Miyazaki, T., Ellis-Davies, G. C., & Kasai, H. (2008). Protein synthesis and neurotrophin-dependent structural plasticity of single dendritic spines. Science, 319, 1683–1687. Taylor, M. S., Devon, R. S., Millar, J. K., & Porteous, D. J. (2003). Evolutionary constraints on the disrupted in schizophrenia locus. Genomics, 81, 67–77. Tiedge, H., Chen, W., & Brosius, J. (1993). Primary structure, neural-specific expression, and dendritic location of human BC200 RNA. Journal of Neuroscience, 13, 2382–2390. Tripathi, V., Ellis, J. D., Shen, Z., Song, D. Y., Pan, Q., Watt, A. T., et al. (2010). The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Molecular Cell, 39, 925–938. Uhde, C. W., Vives, J., Jaeger, I., & Li, M. (2010). Rmst is a novel marker for the mouse ventral mesencephalic floor plate and the anterior dorsal midline cells. PLoS One, 5, e8641. Vallot, C., Huret, C., Lesecque, Y., Resch, A., Oudrhiri, N., Bennaceur-Griscelli, A., et al. (2013). XACT, a long noncoding transcript coating the active X chromosome in human pluripotent cells. Nature Genetics, 45, 239–241. Vlietstra, R. J., van Alewijk, D. C., Hermans, K. G., van Steenbrugge, G. J., & Trapman, J. (1998). Frequent inactivation of PTEN in prostate cancer cell lines and xenografts. Cancer Research, 58, 2720–2723. Voineagu, I., Wang, X., Johnston, P., Lowe, J. K., Tian, Y., Horvath, S., et al. (2011). Transcriptomic analysis of autistic brain reveals convergent molecular pathology. Nature, 474, 380–384.

58

Brent Wilkinson and Daniel B. Campbell

Wang, X. H., Aliyari, R., Li, W. X., Li, H. W., Kim, K., Carthew, R., et al. (2006). RNA interference directs innate immunity against viruses in adult Drosophila. Science, 312, 452–454. Wang, K. C., & Chang, H. Y. (2011). Molecular mechanisms of long noncoding RNAs. Molecular Cell, 43, 904–914. Wang, P., Ren, Z., & Sun, P. (2012). Overexpression of the long non-coding RNA MEG3 impairs in vitro glioma cell proliferation. Journal of Cellular Biochemistry, 113, 1868–1874. Wang, K., Zhang, H., Ma, D., Bucan, M., Glessner, J. T., Abrahams, B. S., et al. (2009). Common genetic variants on 5p14.1 associate with autism spectrum disorders. Nature, 459, 528–533. Watanabe, T., Totoki, Y., Toyoda, A., Kaneda, M., Kuramochi-Miyagawa, S., Obata, Y., et al. (2008). Endogenous siRNAs from naturally formed dsRNAs regulate transcripts in mouse oocytes. Nature, 453, 539–543. Watson, P., Black, G., Ramsden, S., Barrow, M., Super, M., Kerr, B., et al. (2001). Angelman syndrome phenotype associated with mutations in MECP2, a gene encoding a methyl CpG binding protein. Journal of Medical Genetics, 38, 224–228. Wegiel, J., Kuchna, I., Nowicki, K., Imaki, H., Marchi, E., Ma, S. Y., et al. (2010). The neuropathology of autism: Defects of neurogenesis and neuronal migration, and dysplastic changes. Acta Neuropathologica, 119, 755–770. Wheeler, E., Huang, N., Bochukova, E. G., Keogh, J. M., Lindsay, S., Garg, S., et al. (2013). Genome-wide SNP and CNV analysis identifies common and low-frequency variants associated with severe early-onset obesity. Nature Genetics, 45, 513–517. Wienholds, E., & Plasterk, R. H. (2005). MicroRNA function in animal development. FEBS Letters, 579, 5911–5922. Winter, J., Jung, S., Keller, S., Gregory, R. I., & Diederichs, S. (2009). Many roads to maturity: MicroRNA biogenesis pathways and their regulation. Nature Cell Biology, 11, 228–234. Wonders, C. P., & Anderson, S. A. (2006). The origin and specification of cortical interneurons. Nature Reviews Neuroscience, 7, 687–696. Xia, J., Joyce, C. E., Bowcock, A. M., & Zhang, W. (2013). Noncanonical microRNAs and endogenous siRNAs in normal and psoriatic human skin. Human Molecular Genetics, 22, 737–748. Yamasaki, K., Joh, K., Ohta, T., Masuzaki, H., Ishimaru, T., Mukai, T., et al. (2003). Neurons but not glial cells show reciprocal imprinting of sense and antisense transcripts of Ube3a. Human Molecular Genetics, 12, 837–847. Yang, N., & Kazazian, H. H. (2006). L1 retrotransposition is suppressed by endogenously encoded small interfering RNAs in human cultured cells. Nature Structural & Molecular Biology, 13, 763–771. Ying, L., Huang, Y., Chen, H., Wang, Y., Xia, L., Chen, Y., et al. (2013). Downregulated MEG3 activates autophagy and increases cell proliferation in bladder cancer. Molecular BioSystems, 9, 407–411. Yotova, I. Y., Vlatkovic, I. M., Pauler, F. M., Warczok, K. E., Ambros, P. F., Oshimura, M., et al. (2008). Identification of the human homolog of the imprinted mouse Air noncoding RNA. Genomics, 92, 464–473. Zhang, X., Gejman, R., Mahta, A., Zhong, Y., Rice, K. A., Zhou, Y., et al. (2010). Maternally expressed gene 3, an imprinted noncoding RNA gene, is associated with meningioma pathogenesis and progression. Cancer Research, 70, 2350–2358. Zhang, X., Zhou, Y., Mehta, K. R., Danila, D. C., Scolavino, S., Johnson, S. R., et al. (2003). A pituitary-derived MEG3 isoform functions as a growth suppressor in tumor cells. Journal of Clinical Endocrinology and Metabolism, 88, 5119–5126.

LncRNAs in Autism

59

Zhao, J., Dahle, D., Zhou, Y., Zhang, X., & Klibanski, A. (2005). Hypermethylation of the promoter region is associated with the loss of MEG3 gene expression in human pituitary tumors. Journal of Clinical Endocrinology and Metabolism, 90, 2179–2186. Ziats, M. N., & Rennert, O. M. (2013). Aberrant expression of long noncoding RNAs in autistic brain. Journal of Molecular Neuroscience, 49, 589–593. Zoghbi, H. Y. (2003). Postnatal neurodevelopmental disorders: Meeting at the synapse? Science, 302, 826–830.

Contribution of long noncoding RNAs to autism spectrum disorder risk.

Accumulating evidence indicates that long noncoding RNAs (lncRNAs) contribute to autism spectrum disorder (ASD) risk. Although a few lncRNAs have long...
570KB Sizes 0 Downloads 0 Views