Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Dec 1.
Published in final edited form as: Mol Psychiatry. 2015 Mar 31;20(6):677–684. doi: 10.1038/mp.2015.30

Noncoding RNAs and neurobehavioral mechanisms in psychiatric disease

Jannet Kocerha 1,#, Yogesh Dwivedi 2, Kristen J Brennand 3,#
PMCID: PMC4440836  NIHMSID: NIHMS658911  PMID: 25824307

Abstract

The human genome project has revolutionized our understanding of the underlying mechanisms in psychiatric disease. It is now abundantly clear that neurobehavioral phenotypes are epigenetically controlled by noncoding RNAs (ncRNAs). The microRNA (miRNA) class of ncRNAs are ubiquitously expressed throughout the brain and govern all major neuronal pathways. The attractive therapeutic potential of miRNAs is underscored by their pleiotropic capacities, putatively targeting multiple pathways within a single neuron. Many psychiatric diseases stem from a multi-factorial origin, thus conventional drug targeting of single proteins may not prove most effective. In this exciting post-genome sequencing era, many new epigenetic targets are emerging for therapeutic investigation. Here we review the reported roles of miRNAs, as well as other ncRNA classes, in the pathology of psychiatric disorders; there are both common and unique ncRNA mechanisms that influence the various diagnoses. Collectively, these potent epigenetic regulators may clarify the disrupted signaling networks in psychiatric phenotypes.

Introduction

A groundbreaking paradigm shift, the discovery of noncoding RNAs (ncRNAs), established the conventional “one gene, one protein” model to be an oversimplified view of gene regulation. Numerous ncRNA classes are now implicated in central nervous system (CNS) functions; microRNAs (miRNAs), natural antisense transcripts (NATs) and long intergenic RNAs (LincRNAs) all have reported regulatory activities in the brain 1-4. While pleiotropic ncRNAs, including miRNAs, can target large numbers of genes and signaling pathways simultaneously 5, 6, there are also ncRNAs, such as NATs 2, which hybridize to a limited and precise subset of candidates. Growing evidence indicates that distinct neuronal ncRNA mechanisms, particularly miRNAs, likely influence the development of psychiatric disease. Here we review the role of miRNAs in the neurobehavioral deficits of CNS disorders and also discuss the reported contributions of other ncRNA classes.

Neuronal and synaptic miRNA biology

miRNAs have recently emerged as a global regulator of gene expression and, ultimately, effector of synaptic physiology. These miRNAs function by several mechanisms, including ribosomal RNA modifications, repression of mRNA expression by RNA interference, alternative splicing, and regulatory mechanisms mediated by RNA-RNA interactions. The processing from a primary (pri) transcript to precursor (pre) and mature miRNA in the brain requires the standard miRNA biogenesis machinery, including Drosha, Dicer, DGCR8, and argonaute (Ago) proteins 7, 8. Functional knockdown of Dicer, which processes the pre-miRNA to its mature transcript, leads to reduced neuronal size and branching as well as aberrant axonal pathfinding 9-11. Correspondingly, mice with genetic knockout of the pri-miRNA processing protein DGCR8 display a loss of synaptic connectivity and reduced number and size of dendritic spines 12, 13. At the behavioral level, these mice display impaired spatial working memory-dependent tasks 12. Interestingly, it has been shown that miRNAs are formed in part by processing of pre-miRNAs locally within dendritic spines 14, 15. Furthermore, synaptic stimulation leads to local processing of pre-miRNAs in proximity to the synapse 16-18; a cohort of miRNAs are localized within the synapse, including miR-219-5p, miR-124, miR-134, miR-138, and miR-125b 19, 20. These miRNAs directly impact learning and memory behaviors, neurotransmission, neurogenesis among other functions 1, 21-24 and their disruption contribute to psychiatric impairments 25.

A subset of miRNAs mediate neuronal specification, maturation and function. The transition from neural stem cells (NSCs) to neural progenitors and ultimately to fully differentiated neurons is highly regulated by a complex interaction of miRNAs and other factors 24. In general, let-7 26, 27, miR-124 28-30 and miR-9 31 are thought to reduce NSC proliferation and promote neuronal differentiation. It is typically held that miR-134 32 and miR-25 33 induce the proliferation and/or inhibit the differentiation of NSCs and neural progenitors. In parallel, miR-137 both decreases 34, 35 and increases 36, 37 NSC proliferation, either enhancing or opposing neuronal maturation. There is an intricate overlap and feedback occurring between these key miRNAs, mediated in part by their target genes. Notably, miRNAs may mediate neurogenesis throughout development from embryo to adult. Adult neurogenesis is reportedly decreased in neurodegenerative disease 38, 39 and depression 40, and is modulated by antidepressant therapeutics 41. Thus, the following discussions regarding miRNA pathways in psychiatric disorders (outlined in Table 1) may involve neurons derived at all stages of brain maturity.

Table 1. miRNAs implicated in psychiatric phenotypes.

miRNA Biological finding References Diagnosis
miR-137 schizophrenia risk allele 44-47 schizophrenia
alters neuronal connectivity, and size of brain tissues 48-49
direct and indirect targets of miR-137 are linked to schizophrenia 50-53
miRs- 137, 548 associated with schizophrenia (along with miR-137) in largest study to date on
common genetic variants
46
miRs- 181b, 26b the premature and mature transcripts were upregulated in human prefrontal
cortex
54
miR-181b, miR-30e associated with schizophrenia through plasma analysis of human patients 55, 56
miR-219-5p alters schizophrenia behaviors in mice through NMDA-R signaling; targets
CaMKIIgamma
57
most downregulated miR in cortical synaptosomes from schizophrenia
patients; upregulated in tissue (total) from cortex
20
upregulated in cortical tissue (total) from schizophrenia patients 54
miR-132 repressed in mice deficient for the NMDA-R coagonist D-serine 64
miR-134 altered expression in plasma of bipolar patients 79 bipolar disorder
miRs- let-7b, let-7c, 128a, 24a, 30c,
34a, 221, 144
dysregulated in hippocampus of rats by the mood stabilizers valproate and
lithium
78
miRs- 219-5p, 132, 279, 142-3p, 185,
138, let-7b, 125a, 206, 182
linked to circadian rhythm behaviors 65-66; 68-76 circadian function, potential
links to multiple psychiatric
diagnoses
miR-124a regulates depression-associated behaviors in rats through BDNF 83 depression
miR-16 mediates depression behaviors through regulation of SERT; mechanistic
connection to depression through neurogenesis pathways
88-89
miRs- 96, 141, 182, 183, 183*, 198,
200a, 200a*, 200b, 200b*, 200c, 429
different expression profiles of these miRs between rats with learned helpless
(LH) versus non-learned helpless behavior (NLH); possible link to depression
phenotypes
114
miRs- 142-5p, 142-3p, 494, 376a*, 496,
369-3p, 23b, 27b, 24-1*, 34b*, 34c, 17*,
20a, 424, 20b, 142, 301a, 324-5p, 497
downregulated in MDD patients, possibly due to alterered transcription from
the pri-miRNA source
85
miRs- let-7b, mir-132, 181b, 338-3p,
486, 650
networked together in MDD patients 85
rs76481776 polymorphism in the pre-
miR-182
associated with Clock genes 74
miR-135 genetic manipulation of miR-135 leads to precipitation of anxiety and
depression and antidepressant response in mice
91 depression/anxiety
miRs- 19b, 223 disrupted in human and animal models of PTSD 97, 101 PTSD
miR-34c overexpression in mice prevents anxiety-associated responses in stressed
mice; may control these behaviors through CRFR1 gene
102 stress/anxiety
miR-34a controls fear-responses in mice through Notch signaling 103
miRs- 608, 330-3p, genetic variations identified for these miR binding sites in target mRNAs with
association to anxiety
104, 107
miR-124 link to anxiety through glucocorticoid signaling 105
link to stress through corticosteroid signaling 110
miR-16 linked to anxiety in medical students through peripheral blood profiling 109
binds to SERT mRNA 108 implicated via SERT-mediated
pathways
miRs- 22, 138-2, 148a, 488 genetic variations identified in miR binidng sites of gene targets associated
with panic disorders
112 panic disorder

miRNA modulation of behavioral phenotypes in psychiatric diseases

Schizophrenia and bipolar disorder

Substantial evidence indicates that miRNAs mediate schizophrenia phenotypes through a variety of mechanistic pathways. The copy number variant (CNV) at 22q11.2, resulting in DiGeorge syndrome, is the most common rare variant identified in schizophrenia, occurring in approximately one percent of patients 42. Many children with 22q11.2 deletion syndrome have developmental delays and learning disabilities, with a substantially elevated risk of developing schizophrenia (30%) 43. Though the 22q11.2 region contains 30 to 40 genes, many have not been well characterized; one notable gene is DiGeorge syndrome critical region gene 8 (DGCR8), an essential contributor to microRNA biogenesis 8. Microdeletion of 22q11.2 in mice results in the downregulation of a cluster of miRNAs with corresponding changes in cognitive and behavioral functions12.

Of the common genetic variants enriched in patients with schizophrenia, there is a well-established signal at miR-137 and several of its target genes 44-46. Schizophrenia patients with a miR-137 risk allele exhibit greater symptom severity 47, significantly altered functional connectivity 48, reduced white matter integrity, smaller hippocampi and larger lateral ventricles 49. Functionally validated targets of miR-137 include such schizophrenia risk genes as CACNA1C 50, TCF4 50 and ZNF804A 51, while additional putative targets include ERBB4, GABRA1, GRIN2A, GRM5, GSK3B, NRG2, and HTR2C 52. Overexpression of miR-137 in a human NSC line identified direct and indirect miR-137 targets in neural cells. While direct miR-137 targets were enriched for transcription factors and cell cycle genes, indirect targets included pathways enriched in schizophrenia genome wide association studies, particularly major histocompatibility complex, synapses, FMRP interacting RNAs and calcium channels 53. More recently, the largest study of common variance identified 108 regions significantly associated with schizophrenia, including significant hits at miR-137 and miR-548 46.

Post-mortem profiling of miRNA expression in the prefrontal cortex of schizophrenia patients identified a global increase in miRNA expression compared to control populations 54. More specifically, although the mature and pre-miRNA species were increased (particularly of miR-181b and miR-26b), there was no significant difference in transcription of the source pri-miRNA. These results suggest that the changes were due to increased miRNA biogenesis rather than altered miRNA transcription. The authors further observed upregulated expression of Drosha and DGCR8, both of which are involved in pri-miRNA processing, and speculate this is the ultimate cause of aberrant miRNA levels. Additionally, human analyses have started to reveal a profile signature of miRNA dysregulation in peripheral tissues such as plasma and serum of schizophrenia patients. Intriguingly, plasma levels of miR-181b (discussed above) were found to predict response to antipsychotic treatment 55; correspondingly, miR-30e has also been postulated as a plasma biomarker of schizophrenia 56.

NMDA-R signaling, one of the most consistently implicated pathways in schizophrenia, is strongly controlled by miR-219 57. Notably, miR-219 is the most enriched miRNA in the human synapse, the most downregulated in cortical synaptosomes of schizophrenia patients 20, and was dysregulated in cortical tissue from two patient cohorts 20, 54. Moreover, neuronal inhibition of miR-219-5p in mice significantly altered the precipitation of schizophrenia behavior by NMDA-R antagonists 57. Taken together, miR-219-5p actively mediates synaptic functions and the development of psychiatric phenotypes. NMDA-R also regulates miR-132 21, 58-63, which is implicated in learning and memory functions 21, 59, long-term potentiation 58-61 and neurotransmission 61. miR-132 expression is repressed in mice with genetic disruption of the enzyme which produces the NMDA-R co-agonist D-serine 64, and these D-serine deficient mice recapitulate some of the neurobehavioral and cognitive impairments presented in patients with schizophrenia. Expression of miR-132 was dysregulated in human cortical tissue from two schizophrenia datasets but those expression changes were in opposite directions 20, 21; it remains possible that miR-132 is disrupted in specific neuronal cell types or distinct subcellular regions of neurons in schizophrenia. Interestingly, the NMDA-R regulated transcripts miR-132 and miR-219-5p (discussed above) 65, 66 regulate circadian rhythm, which is frequently disrupted in schizophrenia and other psychiatric disorders 67. In addition, a subset of other miRNAs also reportedly regulate circadian functions, including miRs – 279, 142-3p, 185, 138, let-7b, 125a, 206 and 182 68-76. Indeed, miRNA regulators of NMDA-R signaling and circadian rhythm could yield new therapeutic targets for treatment of neurobehavioral deficits.

miRNA expression is responsive to current therapeutics administered for bipolar disorder, which is known to share overlapping genetic links with schizophrenia 77. In rats treated with either lithium or valproate, there are a cohort of miRNAs altered in the hippocampus, including let-7b, let-7c, miR-128a, miR-24a, miR-30c, miR-34a, miR-221 and miR-144 78. Additionally, miR-134 is altered in the plasma of treated bipolar disorder patients 79. Valproate and lithium significantly modulate Brain-Derived Neurotrophic Factor (BDNF) levels, a critical regulator of neuronal homeostasis 80, 81, which is itself regulated by both short and long ncRNAs 82-84, such as miR-124a. Notably, miR-124a is linked to depression-related behaviors 83, as discussed in the section below.

As more post-mortem datasets become publicly available, we believe that groups of consistently altered transcripts in schizophrenia and bipolar disorder will emerge and potentially converge. Combining multimodal SNP and exome sequencing genotype information with RNA and miRNA expression datasets will facilitate miRNA-mRNA correlations. Furthermore, in vivo studies will allow for comprehensive investigation of identified candidates. Arguably, some of the more persistent miRNA associations in schizophrenia to date implicate miR-137, miR-181b, and miR-219-5p (Table 1).

Depression

There is accumulating evidence for significant contribution of miRNA mechanisms in mood disorders such as depression. In the prefrontal cortex of depressed subjects who had died by suicide 85, 21 miRNAs were significantly down-regulated in the major depressive disorder (MDD) group. More miRNAs were down-regulated than upregulated, implying a global down-regulation of miRNA levels in MDD. Furthermore, almost half of the down-regulated miRNAs were transcribed by the same pri-miRNA gene transcripts (mir-142-5p and 142-3p; mir-494, 376a*, 496, and 369-3p; mir-23b, 27b and 24-1*; mir-34b* and 34c; mir-17* and 20a) or found within the same chromosomal region (mir-424 and 20b at Xq26.2-3, 377 kb apart; mir-142 and 301a at 17q22, 820 kb apart; mir-324-5p and 497 at 17p13.1, 205 kb apart), suggesting that the down-regulated miRNA expression may be due to decreased transcription. In addition, a set of 29 miRNAs formed an inter-connected network in the MDD group: let-7b, mir-132, 181b, 338-3p, 486-5p, and 650 were “hubs”. A recent study of genotyping polymorphisms from three miRNA processing genes (DGCR8, AGO1, and GEMIN4) found that DGCR8 rs3757 was associated with increased risk of suicidal tendency and improvement response to antidepressant treatment, whereas AGO1 rs636832 showed decreased risk of suicidal tendency, suicidal behavior, and recurrence 86. Thus, polymorphisms in miRNA processing genes may influence depression risk and treatment.

Molecular targets of anti-depressants frequently engage the transporters of serotonin, a monoamine neurotransmitter 87. miRNAs regulate the serotonin transporter (SERT) and its response to serotonin reuptake inhibitor (SSRI) therapeutics. Specifically, SSRI antidepressant fluoxetine (Prozac) treatment in mice induces expression of miR-16 while repressing the miR-16 target SERT 88. Notably, miR-16 mediates the depression-related behavior through precise control of neurogenesis 89. Another SERT associated miRNA, miR-135, was found to control the onset of co-existing depression and anxiety symptoms in mice 90, 91 as well as the response to anti-depressant treatment 90, 91.

Because of limited studies in depressed patients, it is difficult to pinpoint specific miRNAs consistently implicated in the pathogenesis of depression; nevertheless, miRNAs that influence BDNF, including miR-132 and miR-34, and neuroinflammation, such as the let-7 family92, may be highly relevant (Table 1). Intriguingly, miR-124a repression of BDNF provokes depression-related behaviors 83 and significantly mediates neurogenesis. Future studies could investigate if current anti-depressant therapeutics signal through miR-124a or other disease-associated miRNAs to enhance adult hippocampal neurogenesis 40, 41, 93, 94.

Stress, anxiety, and fear disorders

The link between ncRNAs and stress, anxiety, or fear related responses opens new avenues for therapeutic intervention95. While a range of genetic associations have been loosely implicated in Post-Traumatic Stress Disorder (PTSD) 96, epigenetic factors likely play a defining role in disease progression. A panel of disrupted miRNAs in the blood of military veterans with active PTSD symptoms have been identified 97; these PTSD-associated miRNAs were significantly associated with immunological pathways, suggestive of their pathogenic mechanisms in the disorder. Indeed, abnormal systemic immune responses are routinely reported in stress-related conditions 98-100, consistent with its regulation by the PTSD-linked miRNAs. Intriguingly, a few of the miRNA transcripts altered in the veterans suffering from PTSD, including miR-19b and miR-223, were also perturbed in the serum and amygdala of a PTSD animal model 101.

Stress and anxiety, however, are not unique behaviors to PTSD and are provoked through many distinct triggers. Overexpression of miR-34c in the central amygdala elicited an anxiolytic-like effect in mice stressed through acute restraint 102; furthermore, miR-34c directly targets a key mediator of stress responses, the corticotropin releasing factor receptor type 1 (CRFR1) gene. Notably, miR-34a was recently reported to regulate fear related responses through Notch signaling in mice 103. Although miR-34a didn’t alter anxiety-related parameters in the same manner as miR-34c, it is possible that the miR-34 family is central in behavioral manifestations.

In addition to the above mentioned miRNAs, miRs - 608, 124a, 132, 330-3p, and 16 are also implicated in anxiety phenotypes through genetic associations or in vivo analyses 104-109. Specifically, miR-124 is linked to both anxiety and stress related behaviors through glucocorticoid and corticosteroid signaling, respectively 105, 110. Moreover, ablation of the miRNA biogenesis enzyme Dicer in the central amygdala of mice provoked the onset of anxiety-like symptoms 102. Mice with Dicer knockout specifically in dopaminoceptive neurons also exhibit behavioral changes, including ataxia as well as front and hind limb clasping 111. These Dicer-related phenotypes may also be independent of neurodegenerative mechanisms 111.

Human genetic linkage studies and animal models also revealed miRNA pathways in panic or fear responses. At least four miRNAs have single- nucleotide polymorphisms (SNPs) in miRNA sequences located within panic disorder associated genes, including miR-22, miR-138-2, miR-148a, and miR-488 112. Furthermore, miR-128 expression is increased with the formation of fear-extinction memory in mice, which is the re-conditioning of the memory to overcome established fear behaviors 113.

An individual’s ability to cope with stress is critical in the development of MDD. There is contrasting miRNA expression between rats who developed learned helpless (LH), a behavior that resembles stress-induced depression, compared to rats who did not develop depression (non-learned helpless [NLH]) in spite of receiving similar inescapable shocks 114. One set of miRNAs showed large, significant, and consistent down-regulation in the frontal cortex of NLH rats compared to a blunted response in LH rats (miR-96, miR-141, miR-182, miR-183, miR-183*, miR-198, miR-200a, miR-200a*, miR-200b, miR-200b*, miR-200c, and miR-429). These synaptically enriched miRNAs16 are encoded at a few shared polycistronic loci, suggesting coordinated control of their transcription, and they share 5′-seed motifs which indicate similar or overlapping sets of target mRNAs. Interestingly, half of this set are predicted to hit Creb1 as a target, and binding sites for CREB lie upstream of miR-96, miR-182, miR-183, miR-200a, miR-200b, miR-200c, miR-220a*, and miR-200b*. This suggests that a feedback loop arrangement may also exist between Creb and Creb-stimulated miRNAs and target genes 115. Because these miRNAs are down regulated in NLH rats, but not LH rats, this can be interpreted as a homeostatic response intended to minimize repressive effects on Creb1.

Although more studies are needed to identify a list of miRNAs associated with stress, anxiety and fear, some putative candidates are emerging (Table 1). As discussed previously, miR19b and miR-223 were dysregulated in human and animal models of PTSD 97, 101 and should be evaluated further for therapeutic targeting in trauma-induced stress. Notably, these two transcripts share over 80 bioinformatic mRNA targets through Targetscan prediction. Additionally, the miR-34 family may exhibit an overlapping role in anxiety and fear responses 102, 103. Overall, replication studies in human and animals will further define persistent miRNA mechanisms in these neurobehavioral deficits.

Long ncRNAs and the interaction between ncRNA transcripts in psychiatric disease

Although miRNAs are arguably the most extensively characterized class of ncRNA in neurons, long ncRNAs (lncRNA) are increasingly implicated in CNS functions 4, 116-118. Long ncRNAs acts as miRNA sponges and can bind proteins and RNAs that regulate transcriptional changes and epigenetic modifications of chromatin. Indeed, these transcripts regulate basic neuronal biology as well as genes with strong disease-association 4, 116-122. BC1, one of the first lncRNA transcripts characterized in the brain, is now known to modulate metabotropic receptor signaling 123, 124. Select members of the NAT class of ncRNAs in the brain are functional and modulate expression of their sense partner 2, 82; NATs are endogenously transcribed and exhibit at least partial sequence complimentarity to protein-coding genes, ranging from short to long in nucleotide length 2. An emerging area of investigation indicates that long and short neuronal ncRNAs are co-regulatory 125, which is discussed more throughout the following sections. For example, the BACE1 gene linked to psychiatric and cognitive deficits in Alzheimer’s Disease (AD)] is reportedly modulated by a network of competitive ncRNA interactions 126. More specifically, it was shown that miR-485-5p binding sites in BACE1 are masked through a long antisense transcript, BACE1-AS 126.

Evidence demonstrates that lncRNAs participate in neural plasticity, supported by the expression of a large number of these transcripts in dendrites 127. In addition, a genome-wide analysis show that lncRNAs are modulated by neuronal activity in human brain 128. LncRNAs also mediate neurogenesis, neurodevelopment and related pathways 129, 130. Reports suggest that fine-tuned regulatory control of the embryonic brain by these transcripts is required for development of mature CNS functions 130. Mouse knockout studies involving long-intergenic ncRNAs (lincRNAs) found that loss of function for linc–Brn1b results in a reduction of cerebral cortex progenitor cells and abnormal cortical lamination amongst other pathologies 3. Furthermore, a subset of lincRNAs bind miRNAs and inhibit their functions (i.e. miRNA sponge), which can reportedly disrupt brain development 125. Circular ncRNAs (circRNA) were also recently reported to function as miRNA sponges in neurons. For example, a circRNA was found to have multiple miRNA binding sites, including for miR-7, and both were co-expressed in neocortical and hippocampal neurons 131. These studies further indicate the importance of interactions between ncRNAs in the brain and the need to uncover their epigenetic networks.

Neurotrophins, which play a key role in maintaining homeostatic activity in the CNS, reportedly respond to lncRNA signaling. For example, the neurotrophin BDNF has sequence complementarity with a conserved NAT, termed BDNF-AS. Functionally, BDNF-AS was reported to modulate neuronal growth in vitro and in vivo 82. The sequence complementarity of this noncoding transcript appears to extend into the 3′ untranslated region (3′UTR) of BDNF, which contains regulatory miRNA binding sites such as miR-124a. miR-124, as discussed in previous sections, modulates depression and related neurobehavioral deficits; it is possible that the BDNF-AS partly functions by preventing miR-124 from binding to BDNF.

Susceptibility genes that precipitate psychiatric symptoms are also subject to lncRNA modulation. For example, an endogenous noncoding antisense transcript to the HTT gene (HTTAS) functionally regulates HTT expression119, which is the primary genetic aberration in HD and associated cognitive and neurobehavioral pathology. Susceptibility genes for schizophrenia are also epigenetically controlled by lncRNAs. Genetic variations in the Disrupted in Schizophrenia 1 (DISC1) gene are consistently linked with schizophrenia-associated behaviors 132, 133. A recent report suggests the lncRNA Gomafu mediates DISC1 splicing events, resulting in splice variants linked to schizophrenia 134. Gomafu, is implicated in neural development 129 and is downregulated in cortical tissue of schizophrenia patients and in activated human neuronal cells 134. Additionally, DISC1 is regulated by its lncRNA antisense transcript DISC2 120-122, 132. Several groups have reported genetic association of DISC2 with schizophrenia as well as other psychiatric disorders 120-122, 132; however, DISC2 is not conserved among species 132, hampering investigation of its regulatory mechanisms. One intriguing possibility is the existence of regulatory loop between DISC1 and the two ncRNA transcripts DISC2 and Gomafu.

Recent peripheral blood profiling studies also identified significant disruption of lncRNAs in depression, specifically patients with MDD, 17 of which were documented as depression-related gene in previous studies 135. In parallel, this study uncovered potential miRNA-mRNA networks that are consistent with genes previously associated with depression. Future studies could investigate co-regulatory mechanisms between the dysregulated miRNAs and lncRNAs in the disease. Additionally, the lncRNA antisense transcript coded by LOC285758 has been implicated in violent suicide completers 136, likely triggered through depression or other psychiatric disorders.

Similar to miRNAs, it is possible some of the lncRNAs listed above may mediate the development of the distinct psychiatric symptoms through control of circadian genes. Global transcriptome profiling studies indicate that lncRNAs, including NATs, epigenetically regulate circadian biology 137, 138. Indeed, an antisense RNA to a circadian gene in Neurospora, termed qrf, is regulated by light and represses the expression of its sense partner frq, through chromatin modifications 139.

Summary

Regulating neural plasticity, neurogenesis, and numerous behavioral phenotypes, it is clear that ncRNAs contribute to the pathogenesis of many psychiatric disorders. Moreover, direct evidence comes from human postmortem brain and animal studies indicating perturbed ncRNA levels in disparate disease such as Huntington’s, schizophrenia, depression, PTSD among others.

Despite these findings, one needs to find an integrated view of these ncRNA network(s). It is known that differential co-expression of distinct miRNA groups can directly mediate human disease pathogenesis as well as serve as a biomarker profile for disease diagnosis140-143. The disrupted miRNAs, and their corresponding mRNA target genes in each psychiatric disease, are likely to interact with and regulate each other, both directly as targets and indirectly as part of larger regulatory networks. Furthermore, correlated miRNAs and mRNAs may be coordinately regulated by a (possibly overlapping) set of transcription factors or other epigenetic influences. It remains to be determined whether the changes in the miRNA/mRNA network are i) similar or different across distinct brain regions ii) cell type-specific and iii) reversible mechanisms.

The underlying reasons for altered ncRNA expression remain unresolved and could result from a number of factors, including genetic changes in the promoter region or other locations within the gene. Additionally, defects in RNA editing or epigenetic suppression of the chromosomal region encoding the ncRNAs can also occur. Finally, miRNAs, lncRNAs, and their processing genes are susceptible to regulation by well-established signaling modulators such as BDNF, CREB, calcium, or calcium responsive neurotransmitters. For example, recently it was reported that Dicer is activated by proteolytic cleavage under conditions of elevated calcium levels 14, 16. The networking of miRNAs and other ncRNAs into critical pathways such as neurotransmission, neurogenesis, and neurodevelopment may open up an entirely new understanding of psychiatric disease. Ultimately, we expect that for many of the complex psychiatric disorders we have considered, novel links between ncRNA mechanisms and the disease pathology will continue to emerge.

Acknowledgements

We would like to acknowledge that some of the studies mentioned in the review were funded by National Institute of Mental Health (R01MH100616) to Dr. Dwivedi. Kristen Brennand is a New York Stem Cell Foundation - Robertson Investigator. The Brennand Laboratory is supported by a Brain and Behavior Young Investigator Grant, National Institute of Health (NIH) grant R01 MH101454 and the New York Stem Cell Foundation. The authors would also like to thank Matthew Daniel for helping to prepare Table 1 on miRNAs in psychiatric disease.

Footnotes

Conflict of Interest: The authors declare no conflict of interest.

References

  • 1.Kocerha J, Kauppinen S, Wahlestedt C. microRNAs in CNS disorders. Neuromolecular Med. 2009;11(3):162–172. doi: 10.1007/s12017-009-8066-1. [DOI] [PubMed] [Google Scholar]
  • 2.Faghihi MA, Wahlestedt C. Regulatory roles of natural antisense transcripts. Nat Rev Mol Cell Biol. 2009;10(9):637–643. doi: 10.1038/nrm2738. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Sauvageau M, Goff LA, Lodato S, Bonev B, Groff AF, Gerhardinger C, et al. Multiple knockout mouse models reveal lincRNAs are required for life and brain development. Elife. 2013;2:e01749. doi: 10.7554/eLife.01749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Barry G, Mattick JS. The role of regulatory RNA in cognitive evolution. Trends Cogn Sci. 2012;16(10):497–503. doi: 10.1016/j.tics.2012.08.007. [DOI] [PubMed] [Google Scholar]
  • 5.Bartel DP. MicroRNAs: target recognition and regulatory functions. Cell. 2009;136(2):215–233. doi: 10.1016/j.cell.2009.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Nam JW, Rissland OS, Koppstein D, Abreu-Goodger C, Jan CH, Agarwal V, et al. Global analyses of the effect of different cellular contexts on microRNA targeting. Molecular cell. 2014;53(6):1031–1043. doi: 10.1016/j.molcel.2014.02.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Gregory RI, Chendrimada TP, Shiekhattar R. MicroRNA biogenesis: isolation and characterization of the microprocessor complex. Methods Mol Biol. 2006;342:33–47. doi: 10.1385/1-59745-123-1:33. [DOI] [PubMed] [Google Scholar]
  • 8.Gregory RI, Yan KP, Amuthan G, Chendrimada T, Doratotaj B, Cooch N, et al. The Microprocessor complex mediates the genesis of microRNAs. Nature. 2004;432(7014):235–240. doi: 10.1038/nature03120. [DOI] [PubMed] [Google Scholar]
  • 9.Davis TH, Cuellar TL, Koch SM, Barker AJ, Harfe BD, McManus MT, et al. Conditional loss of Dicer disrupts cellular and tissue morphogenesis in the cortex and hippocampus. J Neurosci. 2008;28(17):4322–4330. doi: 10.1523/JNEUROSCI.4815-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.De Pietri Tonelli D, Pulvers JN, Haffner C, Murchison EP, Hannon GJ, Huttner WB. miRNAs are essential for survival and differentiation of newborn neurons but not for expansion of neural progenitors during early neurogenesis in the mouse embryonic neocortex. Development. 2008;135(23):3911–3921. doi: 10.1242/dev.025080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Schaefer A, O’Carroll D, Tan CL, Hillman D, Sugimori M, Llinas R, et al. Cerebellar neurodegeneration in the absence of microRNAs. J Exp Med. 2007;204(7):1553–1558. doi: 10.1084/jem.20070823. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Stark KL, Xu B, Bagchi A, Lai WS, Liu H, Hsu R, et al. Altered brain microRNA biogenesis contributes to phenotypic deficits in a 22q11-deletion mouse model. Nat Genet. 2008;40(6):751–760. doi: 10.1038/ng.138. [DOI] [PubMed] [Google Scholar]
  • 13.Olde Loohuis NF, Kos A, Martens GJ, Van Bokhoven H, Nadif Kasri N, Aschrafi A. MicroRNA networks direct neuronal development and plasticity. Cellular and molecular life sciences: CMLS. 2012;69(1):89–102. doi: 10.1007/s00018-011-0788-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Lugli G, Larson J, Martone ME, Jones Y, Smalheiser NR. Dicer and eIF2c are enriched at postsynaptic densities in adult mouse brain and are modified by neuronal activity in a calpain-dependent manner. J Neurochem. 2005;94(4):896–905. doi: 10.1111/j.1471-4159.2005.03224.x. [DOI] [PubMed] [Google Scholar]
  • 15.Glanzer J, Miyashiro KY, Sul JY, Barrett L, Belt B, Haydon P, et al. RNA splicing capability of live neuronal dendrites. Proc Natl Acad Sci U S A. 2005;102(46):16859–16864. doi: 10.1073/pnas.0503783102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Lugli G, Torvik VI, Larson J, Smalheiser NR. Expression of microRNAs and their precursors in synaptic fractions of adult mouse forebrain. J Neurochem. 2008;106(2):650–661. doi: 10.1111/j.1471-4159.2008.05413.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Smalheiser NR. Regulation of mammalian microRNA processing and function by cellular signaling and subcellular localization. Biochim Biophys Acta. 2008;1779(11):678–681. doi: 10.1016/j.bbagrm.2008.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Smalheiser NR. Synaptic enrichment of microRNAs in adult mouse forebrain is related to structural features of their precursors. Biol Direct. 2008;3:44. doi: 10.1186/1745-6150-3-44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Siegel G, Saba R, Schratt G. microRNAs in neurons: manifold regulatory roles at the synapse. Curr Opin Genet Dev. 2011;21(4):491–497. doi: 10.1016/j.gde.2011.04.008. [DOI] [PubMed] [Google Scholar]
  • 20.Smalheiser NR, Lugli G, Zhang H, Rizavi H, Cook EH, Dwivedi Y. Expression of microRNAs and other small RNAs in prefrontal cortex in schizophrenia, bipolar disorder and depressed subjects. PLoS One. 2014;9(1):e86469. doi: 10.1371/journal.pone.0086469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Bicker S, Lackinger M, Weiss K, Schratt G. MicroRNA-132, -134, and -138: a microRNA troika rules in neuronal dendrites. Cellular and molecular life sciences: CMLS. 2014 doi: 10.1007/s00018-014-1671-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Smalheiser NR, Lugli G. microRNA regulation of synaptic plasticity. Neuromolecular Med. 2009;11(3):133–140. doi: 10.1007/s12017-009-8065-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Sun AX, Crabtree GR, Yoo AS. MicroRNAs: regulators of neuronal fate. Current opinion in cell biology. 2013;25(2):215–221. doi: 10.1016/j.ceb.2012.12.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Meza-Sosa KF, Pedraza-Alva G, Perez-Martinez L. microRNAs: key triggers of neuronal cell fate. Frontiers in cellular neuroscience. 2014;8:175. doi: 10.3389/fncel.2014.00175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Pittenger C. Disorders of memory and plasticity in psychiatric disease. Dialogues Clin Neurosci. 2013;15(4):455–463. doi: 10.31887/DCNS.2013.15.4/cpittenger. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Zhao C, Sun G, Li S, Lang MF, Yang S, Li W, et al. MicroRNA let-7b regulates neural stem cell proliferation and differentiation by targeting nuclear receptor TLX signaling. Proc Natl Acad Sci U S A. 2010;107(5):1876–1881. doi: 10.1073/pnas.0908750107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Rybak A, Fuchs H, Smirnova L, Brandt C, Pohl EE, Nitsch R, et al. A feedback loop comprising lin-28 and let-7 controls pre-let-7 maturation during neural stem-cell commitment. Nature cell biology. 2008;10(8):987–993. doi: 10.1038/ncb1759. [DOI] [PubMed] [Google Scholar]
  • 28.Cheng LC, Pastrana E, Tavazoie M, Doetsch F. miR-124 regulates adult neurogenesis in the subventricular zone stem cell niche. Nat Neurosci. 2009;12(4):399–408. doi: 10.1038/nn.2294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Liu K, Liu Y, Mo W, Qiu R, Wang X, Wu JY, et al. MiR-124 regulates early neurogenesis in the optic vesicle and forebrain, targeting NeuroD1. Nucleic Acids Res. 2011;39(7):2869–2879. doi: 10.1093/nar/gkq904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Papagiannakopoulos T, Kosik KS. MicroRNA-124: micromanager of neurogenesis. Cell stem cell. 2009;4(5):375–376. doi: 10.1016/j.stem.2009.04.007. [DOI] [PubMed] [Google Scholar]
  • 31.Zhao C, Sun G, Li S, Shi Y. A feedback regulatory loop involving microRNA-9 and nuclear receptor TLX in neural stem cell fate determination. Nat Struct Mol Biol. 2009;16(4):365–371. doi: 10.1038/nsmb.1576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Gaughwin P, Ciesla M, Yang H, Lim B, Brundin P. Stage-specific modulation of cortical neuronal development by Mmu-miR-134. Cerebral cortex. 2011;21(8):1857–1869. doi: 10.1093/cercor/bhq262. [DOI] [PubMed] [Google Scholar]
  • 33.Brett JO, Renault VM, Rafalski VA, Webb AE, Brunet A. The microRNA cluster miR-106b~25 regulates adult neural stem/progenitor cell proliferation and neuronal differentiation. Aging. 2011;3(2):108–124. doi: 10.18632/aging.100285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Sun G, Ye P, Murai K, Lang MF, Li S, Zhang H, et al. miR-137 forms a regulatory loop with nuclear receptor TLX and LSD1 in neural stem cells. Nat Commun. 2011;2:529. doi: 10.1038/ncomms1532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Silber J, Lim DA, Petritsch C, Persson AI, Maunakea AK, Yu M, et al. miR-124 and miR-137 inhibit proliferation of glioblastoma multiforme cells and induce differentiation of brain tumor stem cells. BMC medicine. 2008;6:14. doi: 10.1186/1741-7015-6-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Smrt RD, Szulwach KE, Pfeiffer RL, Li X, Guo W, Pathania M, et al. MicroRNA miR-137 regulates neuronal maturation by targeting ubiquitin ligase mind bomb-1. Stem Cells. 2010;28(6):1060–1070. doi: 10.1002/stem.431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Szulwach KE, Li X, Smrt RD, Li Y, Luo Y, Lin L, et al. Cross talk between microRNA and epigenetic regulation in adult neurogenesis. J Cell Biol. 2010;189(1):127–141. doi: 10.1083/jcb.200908151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Mu Y, Gage FH. Adult hippocampal neurogenesis and its role in Alzheimer’s disease. Mol Neurodegener. 2011;6:85. doi: 10.1186/1750-1326-6-85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Winner B, Rockenstein E, Lie DC, Aigner R, Mante M, Bogdahn U, et al. Mutant alpha-synuclein exacerbates age-related decrease of neurogenesis. Neurobiology of aging. 2008;29(6):913–925. doi: 10.1016/j.neurobiolaging.2006.12.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Zhao C, Deng W, Gage FH. Mechanisms and functional implications of adult neurogenesis. Cell. 2008;132(4):645–660. doi: 10.1016/j.cell.2008.01.033. [DOI] [PubMed] [Google Scholar]
  • 41.Jacobs BL, van Praag H, Gage FH. Adult brain neurogenesis and psychiatry: a novel theory of depression. Mol Psychiatry. 2000;5(3):262–269. doi: 10.1038/sj.mp.4000712. [DOI] [PubMed] [Google Scholar]
  • 42.Rees E, Kirov G, Sanders A, Walters JT, Chambert KD, Shi J, et al. Evidence that duplications of 22q11.2 protect against schizophrenia. Mol Psychiatry. 2014;19(1):37–40. doi: 10.1038/mp.2013.156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Karayiorgou M, Simon TJ, Gogos JA. 22 q11.2 microdeletions: linking DNA structural variation to brain dysfunction and schizophrenia. Nat Rev Neurosci. 2010;11(6):402–416. doi: 10.1038/nrn2841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Ripke S, Sanders AR, Kendler KS, Levinson DF, Sklar P, Holmans PA, et al. Genome-wide association study identifies five new schizophrenia loci. Nat Genet. 2011;43(10):969–976. doi: 10.1038/ng.940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Ripke S, O’Dushlaine C, Chambert K, Moran JL, Kahler AK, Akterin S, et al. Genome-wide association analysis identifies 13 new risk loci for schizophrenia. Nat Genet. 2013;45(10):1150–1159. doi: 10.1038/ng.2742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Schizophrenia Working Group of the Psychiatric Genomics C Biological insights from 108 schizophrenia-associated genetic loci. Nature. 2014;511(7510):421–427. doi: 10.1038/nature13595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Green MJ, Cairns MJ, Wu J, Dragovic M, Jablensky A, Tooney PA, et al. Genome-wide supported variant MIR137 and severe negative symptoms predict membership of an impaired cognitive subtype of schizophrenia. Mol Psychiatry. 2013;18(7):774–780. doi: 10.1038/mp.2012.84. [DOI] [PubMed] [Google Scholar]
  • 48.Liu B, Zhang X, Hou B, Li J, Qiu C, Qin W, et al. The Impact of MIR137 on Dorsolateral Prefrontal-Hippocampal Functional Connectivity in Healthy Subjects. Neuropsychopharmacology. 2014;39(9):2153–2160. doi: 10.1038/npp.2014.63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lett TA, Chakravarty MM, Felsky D, Brandl EJ, Tiwari AK, Goncalves VF, et al. The genome-wide supported microRNA-137 variant predicts phenotypic heterogeneity within schizophrenia. Mol Psychiatry. 2013;18(4):443–450. doi: 10.1038/mp.2013.17. [DOI] [PubMed] [Google Scholar]
  • 50.Kwon E, Wang W, Tsai LH. Validation of schizophrenia-associated genes CSMD1, C10orf26, CACNA1C and TCF4 as miR-137 targets. Mol Psychiatry. 2013;18(1):11–12. doi: 10.1038/mp.2011.170. [DOI] [PubMed] [Google Scholar]
  • 51.Kim AH, Parker EK, Williamson V, McMichael GO, Fanous AH, Vladimirov VI. Experimental validation of candidate schizophrenia gene ZNF804A as target for hsa-miR-137. Schizophr Res. 2012;141(1):60–64. doi: 10.1016/j.schres.2012.06.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Wright C, Turner JA, Calhoun VD, Perrone-Bizzozero N. Potential Impact of miR-137 and Its Targets in Schizophrenia. Front Genet. 2013;4:58. doi: 10.3389/fgene.2013.00058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Collins AL, Kim Y, Bloom RJ, Kelada SN, Sethupathy P, Sullivan PF. Transcriptional targets of the schizophrenia risk gene MIR137. Transl Psychiatry. 2014;4:e404. doi: 10.1038/tp.2014.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Beveridge NJ, Gardiner E, Carroll AP, Tooney PA, Cairns MJ. Schizophrenia is associated with an increase in cortical microRNA biogenesis. Mol Psychiatry. 2010;15(12):1176–1189. doi: 10.1038/mp.2009.84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Song HT, Sun XY, Zhang L, Zhao L, Guo ZM, Fan HM, et al. A preliminary analysis of association between the down-regulation of microRNA-181b expression and symptomatology improvement in schizophrenia patients before and after antipsychotic treatment. J Psychiatr Res. 2014;54:134–140. doi: 10.1016/j.jpsychires.2014.03.008. [DOI] [PubMed] [Google Scholar]
  • 56.Sun XY, Lu J, Zhang L, Song HT, Zhao L, Fan HM, et al. Aberrant microRNA expression in peripheral plasma and mononuclear cells as specific blood-based biomarkers in schizophrenia patients. Journal of clinical neuroscience: official journal of the Neurosurgical Society of Australasia. 2014 doi: 10.1016/j.jocn.2014.08.018. [DOI] [PubMed] [Google Scholar]
  • 57.Kocerha J, Faghihi MA, Lopez-Toledano MA, Huang J, Ramsey AJ, Caron MG, et al. MicroRNA-219 modulates NMDA receptor-mediated neurobehavioral dysfunction. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(9):3507–3512. doi: 10.1073/pnas.0805854106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Remenyi J, van den Bosch MW, Palygin O, Mistry RB, McKenzie C, Macdonald A, et al. miR-132/212 knockout mice reveal roles for these miRNAs in regulating cortical synaptic transmission and plasticity. PLoS One. 2013;8(4):e62509. doi: 10.1371/journal.pone.0062509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Scott HL, Tamagnini F, Narduzzo KE, Howarth JL, Lee YB, Wong LF, et al. MicroRNA-132 regulates recognition memory and synaptic plasticity in the perirhinal cortex. Eur J Neurosci. 2012;36(7):2941–2948. doi: 10.1111/j.1460-9568.2012.08220.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Wibrand K, Pai B, Siripornmongcolchai T, Bittins M, Berentsen B, Ofte ML, et al. MicroRNA regulation of the synaptic plasticity-related gene Arc. PLoS One. 2012;7(7):e41688. doi: 10.1371/journal.pone.0041688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Wibrand K, Panja D, Tiron A, Ofte ML, Skaftnesmo KO, Lee CS, et al. Differential regulation of mature and precursor microRNA expression by NMDA and metabotropic glutamate receptor activation during LTP in the adult dentate gyrus in vivo. Eur J Neurosci. 2010;31(4):636–645. doi: 10.1111/j.1460-9568.2010.07112.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Pathania M, Torres-Reveron J, Yan L, Kimura T, Lin TV, Gordon V, et al. miR-132 enhances dendritic morphogenesis, spine density, synaptic integration, and survival of newborn olfactory bulb neurons. PLoS One. 2012;7(5):e38174. doi: 10.1371/journal.pone.0038174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Edbauer D, Neilson JR, Foster KA, Wang CF, Seeburg DP, Batterton MN, et al. Regulation of synaptic structure and function by FMRP-associated microRNAs miR-125b and miR-132. Neuron. 2010;65(3):373–384. doi: 10.1016/j.neuron.2010.01.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Balu DT, Li Y, Puhl MD, Benneyworth MA, Basu AC, Takagi S, et al. Multiple risk pathways for schizophrenia converge in serine racemase knockout mice, a mouse model of NMDA receptor hypofunction. Proc Natl Acad Sci U S A. 2013;110(26):E2400–2409. doi: 10.1073/pnas.1304308110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Cheng HY, Papp JW, Varlamova O, Dziema H, Russell B, Curfman JP, et al. microRNA modulation of circadian-clock period and entrainment. Neuron. 2007;54(5):813–829. doi: 10.1016/j.neuron.2007.05.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Mohawk JA, Green CB, Takahashi JS. Central and peripheral circadian clocks in mammals. Annu Rev Neurosci. 2012;35:445–462. doi: 10.1146/annurev-neuro-060909-153128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Asarnow LD, Soehner AM, Harvey AG. Circadian rhythms and psychiatric illness. Curr Opin Psychiatry. 2013;26(6):566–571. doi: 10.1097/YCO.0b013e328365a2fa. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Shende VR, Neuendorff N, Earnest DJ. Role of miR-142-3p in the post-transcriptional regulation of the clock gene Bmal1 in the mouse SCN. PLoS One. 2013;8(6):e65300. doi: 10.1371/journal.pone.0065300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Lee KH, Kim SH, Lee HR, Kim W, Kim DY, Shin JC, et al. MicroRNA-185 oscillation controls circadian amplitude of mouse Cryptochrome 1 via translational regulation. Mol Biol Cell. 2013;24(14):2248–2255. doi: 10.1091/mbc.E12-12-0849. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Davis CJ, Clinton JM, Krueger JM. MicroRNA 138, let-7b, and 125a inhibitors differentially alter sleep and EEG delta-wave activity in rats. J Appl Physiol (1985) 2012;113(11):1756–1762. doi: 10.1152/japplphysiol.00940.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Zhou W, Li Y, Wang X, Wu L, Wang Y. MiR-206-mediated dynamic mechanism of the mammalian circadian clock. BMC Syst Biol. 2011;5:141. doi: 10.1186/1752-0509-5-141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Davis CJ, Clinton JM, Taishi P, Bohnet SG, Honn KA, Krueger JM. MicroRNA 132 alters sleep and varies with time in brain. J Appl Physiol (1985) 2011;111(3):665–672. doi: 10.1152/japplphysiol.00517.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Alvarez-Saavedra M, Antoun G, Yanagiya A, Oliva-Hernandez R, Cornejo-Palma D, Perez-Iratxeta C, et al. miRNA-132 orchestrates chromatin remodeling and translational control of the circadian clock. Hum Mol Genet. 2011;20(4):731–751. doi: 10.1093/hmg/ddq519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Saus E, Soria V, Escaramis G, Vivarelli F, Crespo JM, Kagerbauer B, et al. Genetic variants and abnormal processing of pre-miR-182, a circadian clock modulator, in major depression patients with late insomnia. Hum Mol Genet. 2010;19(20):4017–4025. doi: 10.1093/hmg/ddq316. [DOI] [PubMed] [Google Scholar]
  • 75.Luo W, Sehgal A. Regulation of circadian behavioral output via a MicroRNA-JAK/STAT circuit. Cell. 2012;148(4):765–779. doi: 10.1016/j.cell.2011.12.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Kadener S, Menet JS, Sugino K, Horwich MD, Weissbein U, Nawathean P, et al. A role for microRNAs in the Drosophila circadian clock. Genes Dev. 2009;23(18):2179–2191. doi: 10.1101/gad.1819509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Cardno AG, Owen MJ. Genetic relationships between schizophrenia, bipolar disorder, and schizoaffective disorder. Schizophr Bull. 2014;40(3):504–515. doi: 10.1093/schbul/sbu016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Zhou R, Yuan P, Wang Y, Hunsberger JG, Elkahloun A, Wei Y, et al. Evidence for selective microRNAs and their effectors as common long-term targets for the actions of mood stabilizers. Neuropsychopharmacology. 2009;34(6):1395–1405. doi: 10.1038/npp.2008.131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Rong H, Liu TB, Yang KJ, Yang HC, Wu DH, Liao CP, et al. MicroRNA-134 plasma levels before and after treatment for bipolar mania. J Psychiatr Res. 2011;45(1):92–95. doi: 10.1016/j.jpsychires.2010.04.028. [DOI] [PubMed] [Google Scholar]
  • 80.Croce N, Mathe AA, Gelfo F, Caltagirone C, Bernardini S, Angelucci F. Effects of lithium and valproic acid on BDNF protein and gene expression in an in vitro human neuron-like model of degeneration. J Psychopharmacol. 2014 doi: 10.1177/0269881114529379. [DOI] [PubMed] [Google Scholar]
  • 81.Yasuda S, Liang MH, Marinova Z, Yahyavi A, Chuang DM. The mood stabilizers lithium and valproate selectively activate the promoter IV of brain-derived neurotrophic factor in neurons. Mol Psychiatry. 2009;14(1):51–59. doi: 10.1038/sj.mp.4002099. [DOI] [PubMed] [Google Scholar]
  • 82.Modarresi F, Faghihi MA, Lopez-Toledano MA, Fatemi RP, Magistri M, Brothers SP, et al. Inhibition of natural antisense transcripts in vivo results in gene-specific transcriptional upregulation. Nat Biotechnol. 2012;30(5):453–459. doi: 10.1038/nbt.2158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Bahi A, Chandrasekar V, Dreyer JL. Selective lentiviral-mediated suppression of microRNA124a in the hippocampus evokes antidepressants-like effects in rats. Psychoneuroendocrinology. 2014;46:78–87. doi: 10.1016/j.psyneuen.2014.04.009. [DOI] [PubMed] [Google Scholar]
  • 84.Li YJ, Xu M, Gao ZH, Wang YQ, Yue Z, Zhang YX, et al. Alterations of serum levels of BDNF-related miRNAs in patients with depression. PLoS One. 2013;8(5):e63648. doi: 10.1371/journal.pone.0063648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Smalheiser NR, Lugli G, Rizavi HS, Torvik VI, Turecki G, Dwivedi Y. MicroRNA expression is down-regulated and reorganized in prefrontal cortex of depressed suicide subjects. PloS one. 2012;7(3):e33201. doi: 10.1371/journal.pone.0033201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.He Y, Zhou Y, Xi Q, Cui H, Luo T, Song H, et al. Genetic variations in microRNA processing genes are associated with susceptibility in depression. DNA and cell biology. 2012;31(9):1499–1506. doi: 10.1089/dna.2012.1660. [DOI] [PubMed] [Google Scholar]
  • 87.Araragi N, Lesch KP. Serotonin (5-HT) in the regulation of depression-related emotionality: insight from 5-HT transporter and tryptophan hydroxylase-2 knockout mouse models. Curr Drug Targets. 2013;14(5):549–570. doi: 10.2174/1389450111314050005. [DOI] [PubMed] [Google Scholar]
  • 88.Baudry A, Mouillet-Richard S, Schneider B, Launay JM, Kellermann O. miR-16 targets the serotonin transporter: a new facet for adaptive responses to antidepressants. Science. 2010;329(5998):1537–1541. doi: 10.1126/science.1193692. [DOI] [PubMed] [Google Scholar]
  • 89.Launay JM, Mouillet-Richard S, Baudry A, Pietri M, Kellermann O. Raphe-mediated signals control the hippocampal response to SRI antidepressants via miR-16. Transl Psychiatry. 2011;1:e56. doi: 10.1038/tp.2011.54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Asan E, Steinke M, Lesch KP. Serotonergic innervation of the amygdala: targets, receptors, and implications for stress and anxiety. Histochem Cell Biol. 2013;139(6):785–813. doi: 10.1007/s00418-013-1081-1. [DOI] [PubMed] [Google Scholar]
  • 91.Issler O, Haramati S, Paul ED, Maeno H, Navon I, Zwang R, et al. MicroRNA 135 Is Essential for Chronic Stress Resiliency, Antidepressant Efficacy, and Intact Serotonergic Activity. Neuron. 2014 doi: 10.1016/j.neuron.2014.05.042. [DOI] [PubMed] [Google Scholar]
  • 92.Thounaojam MC, Kaushik DK, Basu A. MicroRNAs in the brain: it’s regulatory role in neuroinflammation. Mol Neurobiol. 2013;47(3):1034–1044. doi: 10.1007/s12035-013-8400-3. [DOI] [PubMed] [Google Scholar]
  • 93.Rotheneichner P, Lange S, O’Sullivan A, Marschallinger J, Zaunmair P, Geretsegger C, et al. Hippocampal Neurogenesis and Antidepressive Therapy: Shocking Relations. Neural Plast. 2014;2014:723915. doi: 10.1155/2014/723915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Anacker C. Adult Hippocampal Neurogenesis in Depression: Behavioral Implications and Regulation by the Stress System. Curr Top Behav Neurosci. 2014 doi: 10.1007/7854_2014_275. [DOI] [PubMed] [Google Scholar]
  • 95.Scott KA, Hoban AE, Clarke G, Moloney GM, Dinan TG, Cryan JF. Thinking small: towards microRNA-based therapeutics for anxiety disorders. Expert opinion on investigational drugs. 2015:1–14. doi: 10.1517/13543784.2014.997873. [DOI] [PubMed] [Google Scholar]
  • 96.Domschke K. Patho-genetics of posttraumatic stress disorder. Psychiatr Danub. 2012;24(3):267–273. [PubMed] [Google Scholar]
  • 97.Zhou J, Nagarkatti P, Zhong Y, Ginsberg JP, Singh NP, Zhang J, et al. Dysregulation in microRNA Expression Is Associated with Alterations in Immune Functions in Combat Veterans with Post-Traumatic Stress Disorder. PLoS One. 2014;9(4):e94075. doi: 10.1371/journal.pone.0094075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Dhabhar FS. Effects of stress on immune function: the good, the bad, and the beautiful. Immunol Res. 2014;58(2-3):193–210. doi: 10.1007/s12026-014-8517-0. [DOI] [PubMed] [Google Scholar]
  • 99.Slavich GM, Irwin MR. From stress to inflammation and major depressive disorder: a social signal transduction theory of depression. Psychol Bull. 2014;140(3):774–815. doi: 10.1037/a0035302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Jones KA, Thomsen C. The role of the innate immune system in psychiatric disorders. Molecular and cellular neurosciences. 2013;53:52–62. doi: 10.1016/j.mcn.2012.10.002. [DOI] [PubMed] [Google Scholar]
  • 101.Balakathiresan NS, Chandran R, Bhomia M, Jia M, Li H, Maheshwari RK. Serum and amygdala microRNA signatures of posttraumatic stress: Fear correlation and biomarker potential. Journal of psychiatric research. 2014 doi: 10.1016/j.jpsychires.2014.05.020. [DOI] [PubMed] [Google Scholar]
  • 102.Haramati S, Navon I, Issler O, Ezra-Nevo G, Gil S, Zwang R, et al. MicroRNA as repressors of stress-induced anxiety: the case of amygdalar miR-34. J Neurosci. 2011;31(40):14191–14203. doi: 10.1523/JNEUROSCI.1673-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Dias BG, Goodman JV, Ahluwalia R, Easton AE, Andero R, Ressler KJ. Amygdala-Dependent Fear Memory Consolidation via miR-34a and Notch Signaling. Neuron. 2014;83(4):906–918. doi: 10.1016/j.neuron.2014.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Hanin G, Shenhar-Tsarfaty S, Yayon N, Hoe YY, Bennett ER, Sklan EH, et al. Competing targets of microRNA-608 affect anxiety and hypertension. Hum Mol Genet. 2014;23(17):4569–4580. doi: 10.1093/hmg/ddu170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Durairaj RV, Koilmani ER. Environmental enrichment modulates glucocorticoid receptor expression and reduces anxiety in Indian field male mouse Mus booduga through up-regulation of microRNA-124a. General and comparative endocrinology. 2014;199:26–32. doi: 10.1016/j.ygcen.2014.01.005. [DOI] [PubMed] [Google Scholar]
  • 106.Shaltiel G, Hanan M, Wolf Y, Barbash S, Kovalev E, Shoham S, et al. Hippocampal microRNA-132 mediates stress-inducible cognitive deficits through its acetylcholinesterase target. Brain structure & function. 2013;218(1):59–72. doi: 10.1007/s00429-011-0376-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Jensen KP, Kranzler HR, Stein MB, Gelernter J. The effects of a MAP2K5 microRNA target site SNP on risk for anxiety and depressive disorders. American journal of medical genetics Part B, Neuropsychiatric genetics: the official publication of the International Society of Psychiatric Genetics. 2014;165B(2):175–183. doi: 10.1002/ajmg.b.32219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Yoon Y, McKenna MC, Rollins DA, Song M, Nuriel T, Gross SS, et al. Anxiety-associated alternative polyadenylation of the serotonin transporter mRNA confers translational regulation by hnRNPK. Proc Natl Acad Sci U S A. 2013;110(28):11624–11629. doi: 10.1073/pnas.1301485110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Honda M, Kuwano Y, Katsuura-Kamano S, Kamezaki Y, Fujita K, Akaike Y, et al. Chronic academic stress increases a group of microRNAs in peripheral blood. PLoS One. 2013;8(10):e75960. doi: 10.1371/journal.pone.0075960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Mannironi C, Camon J, De Vito F, Biundo A, De Stefano ME, Persiconi I, et al. Acute stress alters amygdala microRNA miR-135a and miR-124 expression: inferences for corticosteroid dependent stress response. PLoS One. 2013;8(9):e73385. doi: 10.1371/journal.pone.0073385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Cuellar TL, Davis TH, Nelson PT, Loeb GB, Harfe BD, Ullian E, et al. Dicer loss in striatal neurons produces behavioral and neuroanatomical phenotypes in the absence of neurodegeneration. Proc Natl Acad Sci U S A. 2008;105(14):5614–5619. doi: 10.1073/pnas.0801689105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Muinos-Gimeno M, Espinosa-Parrilla Y, Guidi M, Kagerbauer B, Sipila T, Maron E, et al. Human microRNAs miR-22, miR-138-2, miR-148a, and miR-488 are associated with panic disorder and regulate several anxiety candidate genes and related pathways. Biol Psychiatry. 2011;69(6):526–533. doi: 10.1016/j.biopsych.2010.10.010. [DOI] [PubMed] [Google Scholar]
  • 113.Lin Q, Wei W, Coelho CM, Li X, Baker-Andresen D, Dudley K, et al. The brain-specific microRNA miR-128b regulates the formation of fear-extinction memory. Nat Neurosci. 2011;14(9):1115–1117. doi: 10.1038/nn.2891. [DOI] [PubMed] [Google Scholar]
  • 114.Smalheiser NR, Lugli G, Rizavi HS, Zhang H, Torvik VI, Pandey GN, et al. MicroRNA expression in rat brain exposed to repeated inescapable shock: differential alterations in learned helplessness vs. non-learned helplessness. Int J Neuropsychopharmacol. 2011;14(10):1315–1325. doi: 10.1017/S1461145710001628. [DOI] [PubMed] [Google Scholar]
  • 115.Wu J, Xie X. Comparative sequence analysis reveals an intricate network among REST, CREB and miRNA in mediating neuronal gene expression. Genome Biol. 2006;7(9):R85. doi: 10.1186/gb-2006-7-9-r85. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Clark BS, Blackshaw S. Long non-coding RNA-dependent transcriptional regulation in neuronal development and disease. Frontiers in genetics. 2014;5:164. doi: 10.3389/fgene.2014.00164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ng SY, Lin L, Soh BS, Stanton LW. Long noncoding RNAs in development and disease of the central nervous system. Trends Genet. 2013;29(8):461–468. doi: 10.1016/j.tig.2013.03.002. [DOI] [PubMed] [Google Scholar]
  • 118.Bian S, Sun T. Functions of noncoding RNAs in neural development and neurological diseases. Mol Neurobiol. 2011;44(3):359–373. doi: 10.1007/s12035-011-8211-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Chung DW, Rudnicki DD, Yu L, Margolis RL. A natural antisense transcript at the Huntington’s disease repeat locus regulates HTT expression. Hum Mol Genet. 2011;20(17):3467–3477. doi: 10.1093/hmg/ddr263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Chubb JE, Bradshaw NJ, Soares DC, Porteous DJ, Millar JK. The DISC locus in psychiatric illness. Mol Psychiatry. 2008;13(1):36–64. doi: 10.1038/sj.mp.4002106. [DOI] [PubMed] [Google Scholar]
  • 121.Devon RS, Anderson S, Teague PW, Burgess P, Kipari TM, Semple CA, et al. Identification of polymorphisms within Disrupted in Schizophrenia 1 and Disrupted in Schizophrenia 2, and an investigation of their association with schizophrenia and bipolar affective disorder. Psychiatr Genet. 2001;11(2):71–78. doi: 10.1097/00041444-200106000-00003. [DOI] [PubMed] [Google Scholar]
  • 122.Millar JK, James R, Brandon NJ, Thomson PA. DISC1 and DISC2: discovering and dissecting molecular mechanisms underlying psychiatric illness. Ann Med. 2004;36(5):367–378. doi: 10.1080/07853890410033603. [DOI] [PubMed] [Google Scholar]
  • 123.Zhong J, Chuang SC, Bianchi R, Zhao W, Lee H, Fenton AA, et al. BC1 regulation of metabotropic glutamate receptor-mediated neuronal excitability. The Journal of neuroscience: the official journal of the Society for Neuroscience. 2009;29(32):9977–9986. doi: 10.1523/JNEUROSCI.3893-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Maccarrone M, Rossi S, Bari M, De Chiara V, Rapino C, Musella A, et al. Abnormal mGlu 5 receptor/endocannabinoid coupling in mice lacking FMRP and BC1 RNA. Neuropsychopharmacology: official publication of the American College of Neuropsychopharmacology. 2010;35(7):1500–1509. doi: 10.1038/npp.2010.19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Tay Y, Rinn J, Pandolfi PP. The multilayered complexity of ceRNA crosstalk and competition. Nature. 2014;505(7483):344–352. doi: 10.1038/nature12986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Faghihi MA, Zhang M, Huang J, Modarresi F, Van der Brug MP, Nalls MA, et al. Evidence for natural antisense transcript-mediated inhibition of microRNA function. Genome Biol. 2010;11(5):R56. doi: 10.1186/gb-2010-11-5-r56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Mercer TR, Dinger ME, Sunkin SM, Mehler MF, Mattick JS. Specific expression of long noncoding RNAs in the mouse brain. Proc Natl Acad Sci U S A. 2008;105(2):716–721. doi: 10.1073/pnas.0706729105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Lipovich L, Dachet F, Cai J, Bagla S, Balan K, Jia H, et al. Activity-dependent human brain coding/noncoding gene regulatory networks. Genetics. 2012;192(3):1133–1148. doi: 10.1534/genetics.112.145128. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Mercer TR, Qureshi IA, Gokhan S, Dinger ME, Li G, Mattick JS, et al. Long noncoding RNAs in neuronal-glial fate specification and oligodendrocyte lineage maturation. BMC Neurosci. 2010;11:14. doi: 10.1186/1471-2202-11-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Bond AM, Vangompel MJ, Sametsky EA, Clark MF, Savage JC, Disterhoft JF, et al. Balanced gene regulation by an embryonic brain ncRNA is critical for adult hippocampal GABA circuitry. Nature neuroscience. 2009;12(8):1020–1027. doi: 10.1038/nn.2371. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Hansen TB, Jensen TI, Clausen BH, Bramsen JB, Finsen B, Damgaard CK, et al. Natural RNA circles function as efficient microRNA sponges. Nature. 2013;495(7441):384–388. doi: 10.1038/nature11993. [DOI] [PubMed] [Google Scholar]
  • 132.Taylor MS, Devon RS, Millar JK, Porteous DJ. Evolutionary constraints on the Disrupted in Schizophrenia locus. Genomics. 2003;81(1):67–77. doi: 10.1016/s0888-7543(02)00026-5. [DOI] [PubMed] [Google Scholar]
  • 133.Johnstone M, Thomson PA, Hall J, McIntosh AM, Lawrie SM, Porteous DJ. DISC1 in schizophrenia: genetic mouse models and human genomic imaging. Schizophr Bull. 2011;37(1):14–20. doi: 10.1093/schbul/sbq135. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Barry G, Briggs JA, Vanichkina DP, Poth EM, Beveridge NJ, Ratnu VS, et al. The long non-coding RNA Gomafu is acutely regulated in response to neuronal activation and involved in schizophrenia-associated alternative splicing. Mol Psychiatry. 2014;19(4):486–494. doi: 10.1038/mp.2013.45. [DOI] [PubMed] [Google Scholar]
  • 135.Liu Z, Li X, Sun N, Xu Y, Meng Y, Yang C, et al. Microarray profiling and co-expression network analysis of circulating lncRNAs and mRNAs associated with major depressive disorder. PLoS One. 2014;9(3):e93388. doi: 10.1371/journal.pone.0093388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Punzi G, Ursini G, Shin JH, Kleinman JE, Hyde TM, Weinberger DR. Increased expression of MARCKS in post-mortem brain of violent suicide completers is related to transcription of a long, noncoding, antisense RNA. Mol Psychiatry. 2014;19(10):1057–1059. doi: 10.1038/mp.2014.41. [DOI] [PubMed] [Google Scholar]
  • 137.Coon SL, Munson PJ, Cherukuri PF, Sugden D, Rath MF, Moller M, et al. Circadian changes in long noncoding RNAs in the pineal gland. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(33):13319–13324. doi: 10.1073/pnas.1207748109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Vollmers C, Schmitz RJ, Nathanson J, Yeo G, Ecker JR, Panda S. Circadian oscillations of protein-coding and regulatory RNAs in a highly dynamic mammalian liver epigenome. Cell Metab. 2012;16(6):833–845. doi: 10.1016/j.cmet.2012.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Xue Z, Ye Q, Anson SR, Yang J, Xiao G, Kowbel D, et al. Transcriptional interference by antisense RNA is required for circadian clock function. Nature. 2014 doi: 10.1038/nature13671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Staehler CF, Keller A, Leidinger P, Backes C, Chandran A, Wischhusen J, et al. Whole miRNome-wide differential co-expression of microRNAs. Genomics, proteomics & bioinformatics. 2012;10(5):285–294. doi: 10.1016/j.gpb.2012.08.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Xu J, Li CX, Li YS, Lv JY, Ma Y, Shao TT, et al. MiRNA-miRNA synergistic network: construction via co-regulating functional modules and disease miRNA topological features. Nucleic Acids Res. 2011;39(3):825–836. doi: 10.1093/nar/gkq832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Choi JK, Yu U, Yoo OJ, Kim S. Differential coexpression analysis using microarray data and its application to human cancer. Bioinformatics. 2005;21(24):4348–4355. doi: 10.1093/bioinformatics/bti722. [DOI] [PubMed] [Google Scholar]
  • 143.Mo WJ, Fu XP, Han XT, Yang GY, Zhang JG, Guo FH, et al. A stochastic model for identifying differential gene pair co-expression patterns in prostate cancer progression. BMC Genomics. 2009;10:340. doi: 10.1186/1471-2164-10-340. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES