Neurogenomics

Last updated
Areas of focus in neurogenomics. This figure highlights the different sources of data and areas of research that guide the field of neurogenomics. Neuro overview.png
Areas of focus in neurogenomics. This figure highlights the different sources of data and areas of research that guide the field of neurogenomics.

Neurogenomics is the study of how the genome of an organism influences the development and function of its nervous system. [1] This field intends to unite functional genomics and neurobiology in order to understand the nervous system as a whole from a genomic perspective.

Contents

The nervous system in vertebrates is made up of two major types of cells – neuroglial cells and neurons. Hundreds of different types of neurons exist in humans, with varying functions – some of them process external stimuli; others generate a response to stimuli; others organize in centralized structures (brain, spinal ganglia) that are responsible for cognition, perception, and regulation of motor functions. Neurons in these centralized locations tend to organize in giant networks and communicate extensively with each other. Prior to the availability of expression arrays and DNA sequencing methodologies, researchers sought to understand the cellular behaviour of neurons (including synapse formation and neuronal development and regionalization in the human nervous system) in terms of the underlying molecular biology and biochemistry, without any understanding of the influence of a neuron's genome on its development and behaviour. As our understanding of the genome has expanded, the role of networks of gene interactions in the maintenance of neuronal function and behaviour has garnered interest in the neuroscience research community. Neurogenomics allows scientists to study the nervous system of organisms in the context of these underlying regulatory and transcriptional networks. This approach is distinct from neurogenetics, which emphasizes the role of single genes without a network-interaction context when studying the nervous system. [2]

Approaches

Advent of high-throughput biology

In 1999, Cirelli & Tononi [3] first reported the association of genome-wide brain gene expression profiling (using microarrays) with a behavioural phenotype in mice. Since then, global brain gene expression data, derived from microarrays, has been aligned to various behavioural quantitative trait loci (QTLs) and reported in several publications. [4] [5] [6] However, microarray based approaches have their own problems that confound analysis – probe saturation can result in very small measurable variance of gene expression between genetically unique individuals, [7] and the presence of single nucleotide polymorphisms (SNPs) can result in hybridization artifacts. [8] [9] Furthermore, due to their probe-based nature, microarrays can miss out on many types of transcripts (ncRNAs, miRNAs, and mRNA isoforms). Probes can also have species-specific binding affinities that can confound comparative analysis.

Notably, the association between behavioural patterns and high penetrance single gene loci falls under the purview of neurogenetics research, wherein the focus is to identify a simple causative relationship between a single, high penetrance gene and an observed function/behaviour. However, it has been shown that several neurological diseases tend to be polygenic, being influenced by multiple different genes and regulatory regions instead of one gene alone. There has hence been a shift from single gene approaches to network approaches for studying neurological development and diseases, a shift that has been greatly propelled by the advent of next generation sequencing methodologies.

Next-generation sequencing approaches

Twin studies have revealed that schizophrenia, [10] bipolar disorder, [11] autism spectrum disorder (ASD), [12] [13] and attention deficit hyperactivity disorder [14] (ADHD) are highly heritable, genetically complex psychiatric disorders. However, linkage studies have largely failed at identifying causative variants for psychiatric disorders such as these, primarily because of their complex genetic architecture. Multiple low penetrance risk variants can be aggregated in affected individuals and families, and sets of causative variants could vary across families. Studies along these lines have determined a polygenic basis for several psychiatric disorders. [15] Several independently occurring de novo mutations in patients Alzheimer's disease have been found to disrupt a shared set of functional pathways involved with neuronal signalling, for example. [16] The quest to understand the causative biology of psychiatric disorders is hence greatly assisted by the ability to analyse entire genomes of affected and unaffected individuals in an unbiased manner. [17]

With the availability of massively parallel next generation sequencing methodologies, scientists have been able to look beyond the probe based captures of expressed genes. RNA-seq, for example, identifies 25-60% more expressed genes than microarrays do. In the upcoming field of neurogenomics, it is hoped that by understanding the genomic profiles of different parts of the brain, we might be able to improve our understanding of how the interactions between genes and pathways influence cellular function and development. This approach is expected to be able to identify the secondary gene networks that are disrupted in neurological disorders, subsequently assisting drug development stratagems for brain diseases. [18] The BRAIN initiative launched in 2013, for example, seeks to "inform the development of future treatments for brain disorders, including Alzheimer’s disease, epilepsy, and traumatic brain injury" .

Rare variant association studies (RVAS) have highlighted the role of de novo mutations in several congenital and early-childhood-onset disorders like autism. [19] [20] Several of these protein disrupting mutations have been able to be identified only with the aid of whole genome sequencing efforts, and validated with RNA-Seq. Additionally, these mutations are not statistically enriched in individual genes, but rather, exhibit patterns of statistical enrichment in groups of genes associated with networks regulating neurological development and maintenance. Such a discovery would have been impossible with prior gene-centric approaches (neurogenetics, behavioural neuroscience). Neurogenomics allows for a high-throughput system-based approach for understanding the polygenic basis of neuropsychiatric disorders. [17]

Imaging studies and optical mapping

When autism was identified as a distinct biological disorder in the 1980s, researchers found that autistic individuals showed a brain growth abnormality in the cerebellum in their early developmental years. [21] Subsequent research has indicated that 90% of autistic children have a larger brain volume than their peers by 2 to 4 years of age, and show an expansion in the white and gray matter content in the cerebrum. [22] The white and gray matter in the cerebrum is associated with learning and cognition respectively, and the formation of amyloid plaques in the white matter has been associated with Alzheimer's disease. These findings highlighted the influence of structural variance in the brain on psychiatric disorders, and have motivated the use of imaging technologies to map regions of divergence between healthy and diseased brains. Furthermore, while it may not always be possible to retrieve biological specimens from different areas live human brains, neuroimaging techniques offer a noninvasive means to understanding the biological basis of neurological disorders. It is hoped that an understanding of localization patterns of different psychiatric diseases could in turn inform network analysis studies in neurogenomics.

MRI

Structural Magnetic Resonance Imaging (MRI) can be used to identify the structural composition of the brain. Particularly in the context of neurogenomics, MRI has played an extensive role in the study of Alzheimer's disease over the past four decades. It was initially used to rule out other causes of dementia, [16] but recent studies indicated the presence of characteristic changes in patients with Alzheimer's disease. As a result, MRI scans are currently being used as a neuroimaging tool to help identify the temporal and spatial pathophysiology of Alzheimer's disease, such as specific cerebral alterations and amyloid imaging. [16]

The ease and non-invasive nature of MRI scans has motivated research projects that trace the development and onset of psychiatric diseases in the brain. Alzheimer disease has become a key candidate in this topographical approach to psychiatric diseases. For example, MRI scans are currently being used to track the resting and task-dependent functional profiles of brains in children with autosomal dominant Alzheimer disease. [23] These studies have found indications of early onset brain alterations in at-risk individuals for Alzheimer's disease. [16] The Autism Center of Excellence at University of California, San Diego, is also conducting MRI studies with children between 12 and 42 months, in the hopes of characterizing brain development abnormalities in children who present behavioural symptoms of autism. [24]

Additional research has indicated that there are specifics patterns of atrophy in the cerebrum (as a repercussion of neurodegeneration) in different neurological disorders and diseases. These disease-specific patterns of progression of atrophy can be identified with MRI scans, and provide a clinical phenotype context to neurogenomic research. The temporal information about disease progression provided by this approach can also potentially inform the interpretation of gene network-level perturbations in psychiatric diseases. [16]

Optical mapping

One prohibitive feature of 2nd generation sequencing methodologies is the upper limit on the genomic range accessible by mate-pairing. Optical mapping is an emerging methodology used to span large-scale variants that cannot usually be detected using paired end reads. This approach has been successfully applied to detect structural variants in oligodendroglioma, a type of brain cancer. [25] Recent work has also highlighted the versatility of optical maps in improving existing genome assemblies. Chromosomal rearrangements, microdeletions, and large-scale translocations have been associated with impaired neurological and cognitive function, for example in hereditary neuropathy and neurofibromatosis. Optical mapping can significantly improve variant detection and inform gene interaction network models for the diseased state in neurological disorders.

Studying other brain diseases

Apart from neurological disorders, there are additional diseases that manifest in the brain and have formed exemplar use-case scenarios for the application of brain imaging in network analysis. In a classic example of imaging-genomic analyses, a research study in 2012 compared MRI scans and gene expression profiles of 104 glioma patients in order to distinguish treatment outcomes and identify novel targetable genomic pathways in Glioblastoma Multiforme (GBM). Researchers found two distinct groups of patients with significantly different organization of white matter (invasive vs non-invasive). Subsequent pathway analysis of the gene expression data indicated mitochondrial dysfunction as the top canonical pathway in an aggressive, low-mortality GBM phenotype. [26]

Expansion of brain imaging approaches to other diseases can be used to rule out other medical illnesses while diagnosing psychiatric disorders, but cannot be used to inform the presence or absence of a psychiatric disorder.

Research developmental models

In humans

The current approaches in collecting gene expression data in human brains are to use either microarrays or RNA-seq. Currently, it is rare to gather "live" brain tissue – only when treatments involve brain surgery is there a chance that brain tissue is collected during the procedure. This is the case with epilepsy.

Currently, gene expression data is usually collected on post mortem brains and this is often a barrier to neurogenomics research in humans. [27] [28] After death, the amount of time between death and when the data from the post mortem brain is collected is known as the post mortem interval (PMI). Since RNA degrades after death, a fresh brain is optimal – but not always available. This in turn can influence a variety of downstream analyses. Consideration should be taken of the following factors when working with 'omics data collected from post-mortem brains:

Differential diagnosis also remains a critical pre-analytical confounder of cohort-wide studies of spectrum neurological disorders. Specifically, this has been noted to be a problem for Alzheimer's disease and autism spectrum disorder studies. Furthermore, as our understanding of the diverse symptoms and genomic underpinnings of various neurogenomic disorders improves, the diagnostic criteria itself undergoes rearrangements and review. [32]

Animal models

Ongoing genomics research in neurological disorders tends to use animal models (and corresponding gene homologs) to understand the network interactions underlying a particular disorder due to ethical issues surrounding the retrieval of biological specimens from live human brains. This, too, is not without its roadblocks.

Neurogenomic research with a model organism is contingent on the availability of a fully sequenced and annotated reference genome. Additionally, the RNA profiles (miRNA, ncRNA, mRNA) of the model organism need to be well catalogued, and any inferences applied from them to humans must have a basis in functional/sequence homology. [33]

Zebrafish

Zebrafish development relies on gene networks that are highly conserved among all vertebrates. [34] Additionally, with an extremely well annotated set of 12,000 genes and 1,000 early development mutants that are actually visible in the optically clear zebrafish embryos and larvae, zebrafish offer a sophisticated system for mutagenesis and real-time imaging of developing pathologies. This early development model has been employed to study the nervous system at cellular resolution. [35] [36] The zebrafish model system has already been used to study neuroregeneration [37] and severe polygenic human diseases like cancer and heart disease. [38] Several zebrafish mutants with behavioural variations in response to cocaine and alcohol dosage have been isolated and can also form a basis for studying the pathogenesis of behavioural disorders. [39] [40]

Rodent

Rodent models have been preeminent in studying human disorders. These models have been extensively annotated with gene homologs of several monogenic disorders in humans. Knockout studies of these homologs have led to expansion of our understanding of network interactions of genes in human tissues. For example, the FMR1 gene has been implicated with autism from a number of network studies. [41] [42] Using a knockout of FMR 1 in mice creates the model for Fragile X Syndrome, one of the disorders in the Autism spectrum. [43]

Mice xenografts are particularly useful for drug discovery, [44] and were extremely important in the discovery of early anti-psychotic drugs. The development of animal models for complex psychiatric diseases has also improved over the last few years. Rodent models have demonstrated behavioural phenotype changes resembling a positive schizophrenia state, either after genetic manipulation or after treatment with drugs that target the areas of the brain suspected to influence hyperactivity or neurodevelopment. [45] Interest has been generated in identifying the network disruptions mediated by these laboratory manipulations, and collection of genomic data from rodent studies has contributed significantly to a better understanding of the genomics of psychiatric diseases.

The first mouse brain transcriptome was generated in 2008. [46] Since then, extensive work has been done with building social-stress mice models to study the pathway level expression signatures of various psychiatric diseases. A recent paper simulated features of Post Traumatic Stress Disorder (PTSD) in mice, and profiled the entire transcriptome of these mice. [47] The authors found differential regulation in many biological pathways, some of which were implicated in anxiety disorders (hyperactivity, fear response), mood disorders, and impaired cognition. These findings are backed by extensive transcriptomic analyses of anxiety disorders, and expression level changes in biological pathways involved with fear learning and memory are thought to contribute to the behavioural manifestations of these disorders. [47] It is thought that functional enrichment of genes involved in long term synaptic potentiation, depression, and plasticity has an important role to play in the acquisition, consolidation, and maintenance of traumatic memories underlying anxiety disorders. [47] [48]

Experimental mice models for psychiatric disorders

A common approach to using a mouse model is to apply an experimental treatment to a pregnant mouse in order to affect a whole litter. However, a key issue in the field is the treatment of litters in a statistical analysis. Most studies consider the total number of offspring produced as that may lead to an increase in statistical power. However, the correct way is to count by the number of litters and to normalize based on litter size. It was found that several autism studies incorrectly performed their statistical analyses based on total number of offspring instead of number of litters. [49]

Several anxiety disorders such as post-traumatic stress disorder (PTSD) involve heterogeneous changes in several different brain regions, such as the hippocampus, amygdala, and nucleus accumbens. The cellular encoding of traumatic events and the behavioral responses triggered by such events has been shown to lie primarily in changes in signaling molecules associated with synaptic transmission.

Global gene expression profiling of the various gene regions implicated in fear and anxiety processing, using mice models, has led to the identification of temporally and spatially distinct sets of differentially expressed genes. Pathway analysis of these genes has indicated possible roles in neurogenesis and anxiety-related behavioural responses, alongside other functional and phenotypic observations. [47]

Mice models for brain research have contributed significantly to drug development and increased our understanding of the genomic underpinnings of several neurological diseases in the last generation. Chlorpromazine, the first antipsychotic drug (discovered in 1951), was identified as a viable treatment option after it was shown to suppress response to aversive stimuli in rats in a behavioural screen.

Challenges

The modelling and assessment of latent symptoms (thoughts, verbal learning, social interactions, cognitive behaviour) remains a challenge when using model organisms to study psychiatric disorders with a complex genetic pathology. For example, a given genotype+phenotype in a mouse model must imitate the genomic underpinnings of a phenotype observed in a human.

This is a particularly crucial item of consideration in spectrum disorders such as autism. Autism is a disorder whose symptoms can be divided into two categories: (i) deficits of social interactions and (ii) repetitive behaviours and restricted interests. Since mice tend to be more social creatures amongst all members of the order Rodentia currently being used as model organisms, mice are generally used to model human psychiatric disorders as closely as possible. Particularly for autism, the following work-arounds are currently in place to emulate human behavioural symptoms:

  • For the first diagnostic category of impaired social behaviour, mice are subject to a social assay intended to represent typical autistic social deficits. Normal social behaviour for mice includes sniffing, following, physical contact and allogrooming. Vocal communication could be used as well.
  • There are a number of ways the second diagnostic category can be observed in mice. Examples of repetitive behaviours can include excessive circling, self-grooming and excessive digging. Usually these behaviours would be performed consistently within a long measurement of time (i.e. self-grooming for 10 minutes). [50]
    • While repetitive behaviours are easily observable, it is difficult to characterize actual restricted interests of mice. One aspect of restricted interests of autistic individuals is the "insistence of sameness"—the concept that autistic individuals require their environment to remain consistent. If that environment should change, the individual would experience stress and anxiety. There has been reported success in confirming a mouse model of autism by changing the mouse's environment. [51]

In any of these experiments, the ‘autistic’ mice have a ‘normal’ socializing partner and the scientists observing the mice are unaware ("blind") to the genotypes of the mice.

Gene expression in the brain

The gene expression profile of the central nervous system (CNS) is unique. Eighty percent of all human genes are expressed in the brain; 5,000 of these genes are solely expressed in the CNS. The human brain has the highest amount of gene expression of all studied mammalian brains. In comparison, tissues outside of the brain will have more similar expression levels in comparison to their mammalian counterparts. One source of the increased expression levels in the human brain is from the non-protein coding region of the genome. Numerous studies have indicated that the human brain have a higher level of expression in regulatory regions in comparison to other mammalian brains. There is also notable enrichment for more alternative splicing events in the human brain. [2]

Spatial differences

Gene expression profiles also vary within specific regions of the brain. A microarray study showed that the transcriptome profile of the CNS clusters together based on region. A different study characterized the regulation of gene expression across 10 different regions based on their eQTL signals. [52] The cause of the varying expression profiles relates to function, neuron migration and cellular heterogeneity of the region. Even the three layers of the cerebral cortex have distinct expression profiles. [53]

A study completed at Harvard Medical School in 2014 was able to identify developmental lineages stemming from single base neuronal mutations. The researchers sequenced 36 neurons from the cerebral cortex of three normal individuals, and found that highly expressed genes, and neural associated genes, were significantly enriched for single-neuron SNVs. These SNVs, in turn, were found to be correlated with chromatin markers of transcription from fetal brain. [54]

Development patterns in humans

Gene expression of the brain changes throughout the different phases of life. The most significant levels of expression are found during early development, with the rate of gene expression being highest during fetal development. This results from the rapid growth of neurons in the embryo. Neurons at this stage are undergoing neuronal differentiation, cell proliferation, migration events and dendritic and synaptic development. [55] Gene expression patterns shift closer towards specialized functional profiles during embryonic development, however, certain developmental steps are still ongoing at parturition. Consequently, gene expression profiles of the two brain hemispheres appear asymmetrical at birth. At birth, gene expression profiles appear asymmetrical between brain hemispheres. As development continues, the gene expression profiles become similar between the hemispheres. Given a healthy adult, expression profiles stay relatively consistent from the late twenties into the late forties. From the fifties onwards, there is significant decrease in the expression of genes important for regular function. Despite this, there is an increase in the diversity of genes being expressed across the brain. This age related change in expression may be correlated with GC content. At later stages of life, there is an increase in the induction of low GC-content pivotal genes as well as an increase in the repression of high GC-content pivotal genes. [53] Another cause of the shift in gene diversity is the accumulation of mutations and DNA damage. Gene expression studies show that genes that accrue these age-related mutations are consistent between individuals in the aging population. Genes that are highly expressed at development decrease significantly at late stages in life, whereas genes that are highly repressed at development increase significantly at the late stages. [54]

Evolution of the mammalian brain

The evolution of Homo sapiens since the divergence from the primate common ancestor has shown a marked expansion in the size and complexity of the brain, especially in the cerebral cortex. [56] [57] [58] [59] In comparison to primates, the human cerebral cortex has a larger surface area but differs only slightly in thickness. Many large scale studies in understanding the differences of the human brain from other species have indicated expansion of gene families and changes in alternative splicing to be responsible for the corollary increase in cognitive capabilities and cooperative behaviour in humans. [60] [61] However, we are yet to determine the exact phenotypic consequences of all these changes. One difficulty is that only primates have developed subdivisions in their cerebral cortex, making the modeling of human specific neurological problems difficult to mimic in rodents. [58] [62] [63]

Sequence data is used to understand the evolutionary genetic changes which led to the development of the human CNS. We can then understand how the neurological phenotypes differ between species. Comparative genomics entails comparison of sequence data across a phylogeny to pinpoint the genotypic changes that occur within specific lineages, and understand how these changes might have arisen. The increase in high quality mammalian reference sequences generally makes comparative analysis better as it increases statistical power. However, the increase in number of species in a phylogeny does risk adding unnecessary noise as the alignments of the orthologous sequences usually decrease in quality. Furthermore, different classes of species will have significant differences in their phenotypes. [64]

Despite this, comparative genomics has allowed us to connect the genetic changes found in a phylogeny to specific pathways. In order to determine this, lineages are tested for the functional changes that accrue over time. This is often measured as a ratio of nonsynonymous substitutions to synonymous substitutions or the dN/dS ratio (sometimes, further abbreviated to ω). When the dN/dS ratio is greater than 1, this indicates positive selection. A dN/dS ratio equal to 1 is evidence of no selective pressures. A dN/dS ratio less than 1 indicates negative selection. For example, the conserved regions of the genome will generally have a dN/dS ratio of less than 1 since any changes to those positions will likely be detrimental. [65] Of the genes expressed in the human brain, it is estimated that 342 of them have a dN/dS ratio greater than 1 in the human lineage in comparison to other primate lineages. [64] This indicates positive selection on the human lineage for brain phenotypes. Understanding the significance of the positive selection is generally the next step. For example, ASPM , CDK5RAP2 and NIN are genes that are positively selected for on the human lineage and have been directly correlated with brain size. This finding may help elucidate why human brains are larger than other mammalian brains. [65]

Network level expression differences between species

It is thought that gene expression changes, being the ultimate response for any genetic changes, are a good proxy for understanding phenotypic differences within biological samples. Comparative studies have revealed a range of differences in the transcriptional controls between primates and rodents. For example, the gene CNTNAP2 is specifically enriched for in the prefrontal cortex. The mouse homolog of CNTNAP2 is not expressed in the mouse brain. CNTNAP2 has been implicated in cognitive functions of language as well as neurodevelopmental disorders such as Autism Spectrum Disorder. This suggests that the control of expression plays a significant role in the development in unique human cognitive function. As a consequence, a number of studies have investigated the brain specific enhancers. Transcription factors such as SOX5 have been found to be positively selected for on the human lineage. Gene expression studies in humans, chimpanzees and rhesus macaques, have identified human specific co-expression networks, and an elevation in gene expression in the human cortex in comparison to primates. [66]

Disorders

Neurogenomic disorders manifest themselves as neurological disorders with a complex genetic architecture and a non-Mendelian-like pattern of inheritance. [18] Some examples of these disorders include Bipolar disorder and Schizophrenia. [15] Several genes may be involved in the manifestation of the disorder, and mutations in such disorders are generally rare and de novo. Hence it becomes extremely unlikely to observe the same (potentially causative) variant in two unrelated individuals affected with the same neurogenomic disorder. [15] Ongoing research has implicated several de novo exonic variations and structural variations in Autism Spectrum Disorder (ASD), for example. [15] The allelic spectrum of the rare and common variants in neurogenomic disorders therefore necessitates a need for large cohort studies in order to effectively exclude low effect variants and identify the overarching pathways frequently mutated in the different disorders, rather than specific genes and specific high penetrance mutations.

Whole genome sequencing (WGS) and whole exome sequencing (WES) has been used in Genome Wide Association Studies (GWAS) to characterize genetic variants associated with neurogenomic disorders. However, the impact of these variants cannot always be verified because of the non-Mendelian inheritance patterns observed in several of these disorders. [15] Another prohibitive feature in network analysis is the lack of large-scale datasets for many psychiatric (neurogenomic) diseases. Since several diseases with neurogenomic underpinnings tend to have a polygenic basis, several nonspecific, rare, and partially penetrant de novo mutations in different patients can contribute to the same observed range of phenotypes, as is the case with Autism Spectrum Disorder and schizophrenia. [67] Extensive research in alcohol dependence has also highlighted the need for high-quality genomic profiling of large sample sets [68] [69] when studying polygenic, spectrum disorders.

The 1000 Genomes Project was a successful demonstration of how a concerted effort to acquire representative genomic data from the broad spectrum of humans can result in identification of actionable biological insights for different diseases. [70] However, a large-scale initiative like this is still lacking in the field of neurogenomic disorders specifically.

Modelling psychiatric disorders in neurogenomics research – issues

One major GWAS study identified 13 new risk loci for schizophrenia. [71] Studying the impact of these candidates would ideally demonstrate a schizophrenia phenotype in animal models, which is usually difficult to observe due to its manifestation as a latent personality. This approach is able to determine the molecular impact the candidate gene. Ideally the candidate genes would have a neurological impact, which in turn would suggest that it plays a role in the neurological disorder. For example, in the aforementioned schizophrenia GWAS study, Ripke and colleagues [71] determined that these candidate genes were all involved in calcium signalling. Alternatively, one can study these variants in model organisms in the context of affected neurological function. It is important to note that the high penetrance variants of these disorders tend to be de novo mutations.

A further complication to studying neurogenomic disorders is the heterogeneous nature of the disorder. In many of these disorders, the mutations observed from case to case do not stay consistent. In autism, an affected individual may experience a large amount of deleterious mutations in gene X. A different affected individual may not have any significant mutations on gene X but have a large amount of mutations in gene Y. The alternative is to determine if gene X and gene Y impact the same biochemical pathway—one that influences a neurological function. A bioinformatics network analysis is one approach to this problem. Network analyses methodologies provide a generalized, systems overview of a molecular pathway.

One final complication to consider is the comorbidity of neurogenomic genes. Several disorders, especially at the more severe ends of the spectrum tend to be comorbid with each other. For example, more severe cases of ASD tend to be associated with intellectual disability (ID). This raises the question of whether or not there are true, unique ASD genes and unique ID genes or if there are just genes just associated with neurological function that can be mutated into an abnormal phenotype. One confounding factor may be the actual diagnostic category and methods of the spectrum disorders as symptoms between severe disorders may be similar. One study investigated the comorbid symptoms between groups of ID and ASD, and found no significant difference between the symptoms of ID children, ASD children with ID and ASD children without ID. Future research may help establish a more stringent genetic basis for the diagnoses of these disorders.

Network analysis

A typical network analysis pipeline in neurogenomics Network analysis pipeline.tif
A typical network analysis pipeline in neurogenomics

The main goal of network analysis in neurogenomics is to identify statistically significant nonrandom associations between genes that contain risk variants. [15] While several algorithm implementations of this approach already exist, [72] [73] the general steps for network analysis remain the same.

The underlying principle of this approach is that the genes that cluster together, will also jointly affect the same molecular pathway. Again, they would ideally be part of a neurological function. The candidate genes can then be used to prioritize variants for wet lab validation.

Neuropharmacology

Historically, due to the behavioural stimulation manifested as a symptom in several the neurogenomic disorders, the therapies would rely mostly on anti-psychotics or antidepressants. These classes of medications would suppress common symptoms of the disorders, but with questionable efficacy. The biggest barrier to neruopharmacogenomic research was the cohort sizes. Given newly available large-cohort sequencing data, there has been a recent push to expand therapeutic options. The heterogenous nature of neurological diseases is the key motivation for personalized medicine approaches to their therapies. It is rare to find single high penetrance causative genes in neurological diseases. The genomic profiles understandably vary between cases, and logically, the therapies would need to vary between cases. Further complicating the issue is that many of these disorders are spectrum disorders. Their genetic etiology will vary within this spectrum. For example, severe ASD is associated with high penetrance de novo mutations. Milder forms of ASD is usually associated with a mixture of common variants.

The key issue then is the translation of these newly identified genetic variants (from Copy Number Variant studies, candidate gene sequencing and high throughput sequencing technologies) into an intervention for patients with neurogenomic disorders. One aspect will be if the neurological disorder are medically actionable (i.e. is there a simple metabolic pathway that a therapy can target). For example, specific cases of ASD have been associated with microdeletions on TMLHE gene. This gene codes for the enzyme of carnitine biosynthesis. Supplements to elevate carnitine levels appeared to alleviate certain ASD symptoms but the study was confounded by many influencing factors. As mentioned earlier, using a gene network approach will help identify relevant pathways of interest. Many neuropharmacogenomic approaches have focused on targeting the downstream products of these pathways. [75] [76]

Blood brain barrier

Studies in animal models for several brain diseases has shown that the blood brain barrier (BBB) undergoes modification at many levels; for example, the surface glycoprotein composition can influence the types of HIV-1 strains transported by the BBB. The BBB has been found to be key in the onset of Alzheimer's disease. [77] It is extremely difficult, however, to be able to study this in humans due to obvious restrictions with accessing the brain and retrieving biological specimens for sequencing or morphological analysis. Mice models of the BBB and models of disease states have served well in conceptualizing the BBB as a regulatory interface between disease and good health in the brain.

Personalized neurobiology

The heterogenous nature of neurological diseases is the key motivation for personalized medicine approaches to their therapies. [75] Genomic samples of individual patients could be used to identify predictive factors, or to better understand the specific prognosis of a neurogenomic disease, and use this information to guide treatment options. [78] While there is a clear clinical utility to this approach, the adaptation of this approach is still nonexistent.

There are various issues prohibiting the application of personalized genomics to the assessment, diagnosis, and treatment of psychiatric disorders.

See also

Related Research Articles

<span class="mw-page-title-main">Fragile X syndrome</span> X-linked dominant genetic disorder

Fragile X syndrome (FXS) is a genetic disorder characterized by mild-to-moderate intellectual disability. The average IQ in males with FXS is under 55, while about two thirds of affected females are intellectually disabled. Physical features may include a long and narrow face, large ears, flexible fingers, and large testicles. About a third of those affected have features of autism such as problems with social interactions and delayed speech. Hyperactivity is common, and seizures occur in about 10%. Males are usually more affected than females.

<span class="mw-page-title-main">Rett syndrome</span> Genetic brain disorder

Rett syndrome (RTT) is a genetic disorder that typically becomes apparent after 6–18 months of age and almost exclusively in females. Symptoms include impairments in language and coordination, and repetitive movements. Those affected often have slower growth, difficulty walking, and a smaller head size. Complications of Rett syndrome can include seizures, scoliosis, and sleeping problems. The severity of the condition is variable.

<span class="mw-page-title-main">FOXP2</span> Transcription factor gene of the forkhead box family

Forkhead box protein P2 (FOXP2) is a protein that, in humans, is encoded by the FOXP2 gene. FOXP2 is a member of the forkhead box family of transcription factors, proteins that regulate gene expression by binding to DNA. It is expressed in the brain, heart, lungs and digestive system.

The Allen Mouse and Human Brain Atlases are projects within the Allen Institute for Brain Science which seek to combine genomics with neuroanatomy by creating gene expression maps for the mouse and human brain. They were initiated in September 2003 with a $100 million donation from Paul G. Allen and the first atlas went public in September 2006. As of May 2012, seven brain atlases have been published: Mouse Brain Atlas, Human Brain Atlas, Developing Mouse Brain Atlas, Developing Human Brain Atlas, Mouse Connectivity Atlas, Non-Human Primate Atlas, and Mouse Spinal Cord Atlas. There are also three related projects with data banks: Glioblastoma, Mouse Diversity, and Sleep. It is the hope of the Allen Institute that their findings will help advance various fields of science, especially those surrounding the understanding of neurobiological diseases. The atlases are free and available for public use online.

<span class="mw-page-title-main">Pleiotropy</span> Influence of a single gene on multiple phenotypic traits

Pleiotropy occurs when one gene influences two or more seemingly unrelated phenotypic traits. Such a gene that exhibits multiple phenotypic expression is called a pleiotropic gene. Mutation in a pleiotropic gene may have an effect on several traits simultaneously, due to the gene coding for a product used by a myriad of cells or different targets that have the same signaling function.

<span class="mw-page-title-main">Heritability of autism</span>

The heritability of autism is the proportion of differences in expression of autism that can be explained by genetic variation; if the heritability of a condition is high, then the condition is considered to be primarily genetic. Autism has a strong genetic basis. Although the genetics of autism are complex, autism spectrum disorder (ASD) is explained more by multigene effects than by rare mutations with large effects.

<span class="mw-page-title-main">Causes of autism</span> Proposed causes of autism

The causes of autism are environmental or genetic factors that predispose an individual to develop autism, also known as autism spectrum disorder (ASD). Many causes of autism have been proposed, but understanding of the theory of causation of autism is incomplete. Attempts have been made to incorporate the known genetic and environmental causes into a comprehensive causative framework. ASD is a neurodevelopmental disorder marked by impairments in communicative ability and social interaction and restricted/repetitive behaviors, interests, or activities not suitable for the individual's developmental stage. The severity of symptoms and functional impairment vary between individuals.

<span class="mw-page-title-main">22q13 deletion syndrome</span> Rare genetic syndrome

22q13 deletion syndrome, also known as Phelan–McDermid syndrome (PMS), is a genetic disorder caused by deletions or rearrangements on the q terminal end of chromosome 22. Any abnormal genetic variation in the q13 region that presents with significant manifestations (phenotype) typical of a terminal deletion may be diagnosed as 22q13 deletion syndrome. There is disagreement among researchers as to the exact definition of 22q13 deletion syndrome. The Developmental Synaptopathies Consortium defines PMS as being caused by SHANK3 mutations, a definition that appears to exclude terminal deletions. The requirement to include SHANK3 in the definition is supported by many but not by those who first described 22q13 deletion syndrome.

<span class="mw-page-title-main">DISC1</span> Protein-coding gene in the species Homo sapiens

Disrupted in schizophrenia 1 is a protein that in humans is encoded by the DISC1 gene. In coordination with a wide array of interacting partners, DISC1 has been shown to participate in the regulation of cell proliferation, differentiation, migration, neuronal axon and dendrite outgrowth, mitochondrial transport, fission and/or fusion, and cell-to-cell adhesion. Several studies have shown that unregulated expression or altered protein structure of DISC1 may predispose individuals to the development of schizophrenia, clinical depression, bipolar disorder, and other psychiatric conditions. The cellular functions that are disrupted by permutations in DISC1, which lead to the development of these disorders, have yet to be clearly defined and are the subject of current ongoing research. Although, recent genetic studies of large schizophrenia cohorts have failed to implicate DISC1 as a risk gene at the gene level, the DISC1 interactome gene set was associated with schizophrenia, showing evidence from genome-wide association studies of the role of DISC1 and interacting partners in schizophrenia susceptibility.

<span class="mw-page-title-main">GABRB3</span> Protein-coding gene in the species Homo sapiens

Gamma-aminobutyric acid receptor subunit beta-3 is a protein that in humans is encoded by the GABRB3 gene. It is located within the 15q12 region in the human genome and spans 250kb. This gene includes 10 exons within its coding region. Due to alternative splicing, the gene codes for many protein isoforms, all being subunits in the GABAA receptor, a ligand-gated ion channel. The beta-3 subunit is expressed at different levels within the cerebral cortex, hippocampus, cerebellum, thalamus, olivary body and piriform cortex of the brain at different points of development and maturity. GABRB3 deficiencies are implicated in many human neurodevelopmental disorders and syndromes such as Angelman syndrome, Prader-Willi syndrome, nonsyndromic orofacial clefts, epilepsy and autism. The effects of methaqualone and etomidate are mediated through GABBR3 positive allosteric modulation.

<span class="mw-page-title-main">Kalirin</span> Protein-coding gene in the species Homo sapiens

Kalirin, also known as Huntingtin-associated protein-interacting protein (HAPIP), protein duo (DUO), or serine/threonine-protein kinase with Dbl- and pleckstrin homology domain, is a protein that in humans is encoded by the KALRN gene. Kalirin was first identified in 1997 as a protein interacting with huntingtin-associated protein 1. Is also known to play an important role in nerve growth and axonal development.

<span class="mw-page-title-main">Neurogenetics</span>

Neurogenetics studies the role of genetics in the development and function of the nervous system. It considers neural characteristics as phenotypes, and is mainly based on the observation that the nervous systems of individuals, even of those belonging to the same species, may not be identical. As the name implies, it draws aspects from both the studies of neuroscience and genetics, focusing in particular how the genetic code an organism carries affects its expressed traits. Mutations in this genetic sequence can have a wide range of effects on the quality of life of the individual. Neurological diseases, behavior and personality are all studied in the context of neurogenetics. The field of neurogenetics emerged in the mid to late 20th century with advances closely following advancements made in available technology. Currently, neurogenetics is the center of much research utilizing cutting edge techniques.

Potocki–Lupski syndrome (PTLS), also known as dup(17)p11.2p11.2 syndrome, trisomy 17p11.2 or duplication 17p11.2 syndrome, is a contiguous gene syndrome involving the microduplication of band 11.2 on the short arm of human chromosome 17 (17p11.2). The duplication was first described as a case study in 1996. In 2000, the first study of the disease was released, and in 2007, enough patients had been gathered to complete a comprehensive study and give it a detailed clinical description. PTLS is named for two researchers involved in the latter phases, Drs. Lorraine Potocki and James R. Lupski of Baylor College of Medicine.

Cognitive genomics is the sub-field of genomics pertaining to cognitive function in which the genes and non-coding sequences of an organism's genome related to the health and activity of the brain are studied. By applying comparative genomics, the genomes of multiple species are compared in order to identify genetic and phenotypical differences between species. Observed phenotypical characteristics related to the neurological function include behavior, personality, neuroanatomy, and neuropathology. The theory behind cognitive genomics is based on elements of genetics, evolutionary biology, molecular biology, cognitive psychology, behavioral psychology, and neurophysiology.

Autism spectrum disorder (ASD) refers to a variety of conditions typically identified by challenges with social skills, communication, speech, and repetitive sensory-motor behaviors. The 11th International Classification of Diseases (ICD-11), released in January 2021, characterizes ASD by the associated deficits in the ability to initiate and sustain two-way social communication and restricted or repetitive behavior unusual for the individual's age or situation. Although linked with early childhood, the symptoms can appear later as well. Symptoms can be detected before the age of two and experienced practitioners can give a reliable diagnosis by that age. However, official diagnosis may not occur until much older, even well into adulthood. There is a large degree of variation in how much support a person with ASD needs in day-to-day life. This can be classified by a further diagnosis of ASD level 1, level 2, or level 3. Of these, ASD level 3 describes people requiring very substantial support and who experience more severe symptoms. ASD-related deficits in nonverbal and verbal social skills can result in impediments in personal, family, social, educational, and occupational situations. This disorder tends to have a strong correlation with genetics along with other factors. More research is identifying ways in which epigenetics is linked to autism. Epigenetics generally refers to the ways in which chromatin structure is altered to affect gene expression. Mechanisms such as cytosine regulation and post-translational modifications of histones. Of the 215 genes contributing, to some extent in ASD, 42 have been found to be involved in epigenetic modification of gene expression. Some examples of ASD signs are specific or repeated behaviors, enhanced sensitivity to materials, being upset by changes in routine, appearing to show reduced interest in others, avoiding eye contact and limitations in social situations, as well as verbal communication. When social interaction becomes more important, some whose condition might have been overlooked suffer social and other exclusion and are more likely to have coexisting mental and physical conditions. Long-term problems include difficulties in daily living such as managing schedules, hypersensitivities, initiating and sustaining relationships, and maintaining jobs.

Joseph D. Buxbaum is an American molecular and cellular neuroscientist, autism researcher, and the Director of the Seaver Autism Center at the Icahn School of Medicine at Mount Sinai. Buxbaum is also, along with Simon Baron-Cohen, the co-editor of the BioMed Central journal Molecular Autism, and is a member of the scientific advisory board of the Autism Science Foundation. Buxbaum is a Professor of Psychiatry, Neuroscience, and Genetics and Genomic Sciences. He is also the Vice Chair for Research and for Mentoring in the Department of Psychiatry at the Icahn School of Medicine at Mount Sinai.

The development of an animal model of autism is one approach researchers use to study potential causes of autism. Given the complexity of autism and its etiology, researchers often focus only on single features of autism when using animal models.

Benjamin Michael Neale is a statistical geneticist with a specialty in psychiatric genetics. He is an institute member at the Broad Institute as well as an associate professor at both Harvard Medical School and the Analytic and Translational Genetics Unit at Massachusetts General Hospital. Neale specializes in genome-wide association studies (GWAS). He was responsible for the data analysis of the first GWAS on attention-deficit/hyperactivity-disorder, and he developed new analysis software such as PLINK, which allows for whole-genome data to be analyzed for specific gene markers. Related to his work on GWAS, Neale is the lead of the ADHD psychiatric genetics and also a member of the Psychiatric GWAS Consortium analysis committee.

Personality traits are patterns of thoughts, feelings and behaviors that reflect the tendency to respond in certain ways under certain circumstances.

Sagiv Shifman is an Israeli scientist, professor in the field of neurogenetics at the Alexander Silberman Institute of Life Sciences, The Hebrew University of Jerusalem. He holds the Arnold and Bess Zeldich Ungerman chair in Neurobiology.

References

  1. Boguski, Mark S.; Jones, Allan R. (2004-05-01). "Neurogenomics: at the intersection of neurobiology and genome sciences". Nature Neuroscience. 7 (5): 429–433. doi:10.1038/nn1232. ISSN   1097-6256. PMID   15114353. S2CID   14310435.
  2. 1 2 Jain, Kewal K. (2013-01-01). "Neurogenetics and Neurogenomics". Applications of Biotechnology in Neurology. Humana Press. pp. 7–16. doi:10.1007/978-1-62703-272-8_2. ISBN   9781627032711.
  3. Cirelli, Chiara; Tononi, Giulio (1999). "Differences in gene expression during sleep and wakefulness". Annals of Medicine. 31 (2): 117–124. doi: 10.3109/07853899908998787 . PMID   10344584.
  4. Matthews, Douglas B.; Bhave, Sanjiv V.; Belknap, John K.; Brittingham, Cynthia; Chesler, Elissa J.; Hitzemann, Robert J.; Hoffmann, Paula L.; Lu, Lu; McWeeney, Shannon (2005-09-01). "Complex genetics of interactions of alcohol and CNS function and behavior". Alcoholism: Clinical and Experimental Research. 29 (9): 1706–1719. doi:10.1097/01.alc.0000179209.44407.df. ISSN   0145-6008. PMID   16205371.
  5. Hoffman, Paula L.; Miles, Michael; Edenberg, Howard J.; Sommer, Wolfgang; Tabakoff, Boris; Wehner, Jeanne M.; Lewohl, Joanne (2003-02-01). "Gene expression in brain: a window on ethanol dependence, neuroadaptation, and preference". Alcoholism: Clinical and Experimental Research. 27 (2): 155–168. doi: 10.1097/01.ALC.0000060101.89334.11 . ISSN   0145-6008. PMID   12605065.
  6. Farris, Sean P.; Miles, Michael F. (2012-01-01). "Ethanol modulation of gene networks: implications for alcoholism". Neurobiology of Disease. 45 (1): 115–121. doi:10.1016/j.nbd.2011.04.013. ISSN   1095-953X. PMC   3158275 . PMID   21536129.
  7. Pozhitkov, Alex E.; Boube, Idrissa; Brouwer, Marius H.; Noble, Peter A. (2010-03-01). "Beyond Affymetrix arrays: expanding the set of known hybridization isotherms and observing pre-wash signal intensities". Nucleic Acids Research. 38 (5): e28. doi:10.1093/nar/gkp1122. ISSN   0305-1048. PMC   2836560 . PMID   19969547.
  8. Walter, Nicole A. R.; McWeeney, Shannon K.; Peters, Sandra T.; Belknap, John K.; Hitzemann, Robert; Buck, Kari J. (2007-09-01). "SNPs matter: impact on detection of differential expression". Nature Methods. 4 (9): 679–680. doi:10.1038/nmeth0907-679. ISSN   1548-7091. PMC   3410665 . PMID   17762873.
  9. Walter, Nicole A. R.; Bottomly, Daniel; Laderas, Ted; Mooney, Michael A.; Darakjian, Priscila; Searles, Robert P.; Harrington, Christina A.; McWeeney, Shannon K.; Hitzemann, Robert (2009-01-01). "High throughput sequencing in mice: a platform comparison identifies a preponderance of cryptic SNPs". BMC Genomics. 10: 379. doi:10.1186/1471-2164-10-379. ISSN   1471-2164. PMC   2743714 . PMID   19686600.
  10. 1 2 Sullivan PF; Kendler KS; Neale MC (2003-12-01). "Schizophrenia as a complex trait: Evidence from a meta-analysis of twin studies". Archives of General Psychiatry. 60 (12): 1187–1192. doi: 10.1001/archpsyc.60.12.1187 . ISSN   0003-990X. PMID   14662550.
  11. Smoller, Jordan W.; Finn, Christine T. (2003-11-15). "Family, twin, and adoption studies of bipolar disorder". American Journal of Medical Genetics Part C. 123C (1): 48–58. CiteSeerX   10.1.1.456.6790 . doi:10.1002/ajmg.c.20013. ISSN   1552-4868. PMID   14601036. S2CID   29453951.
  12. 1 2 Rosenberg, Rebecca E.; Law, J. Kiely; Yenokyan, Gayane; McGready, John; Kaufmann, Walter E.; Law, Paul A. (2009-10-01). "Characteristics and concordance of autism spectrum disorders among 277 twin pairs". Archives of Pediatrics & Adolescent Medicine. 163 (10): 907–914. doi: 10.1001/archpediatrics.2009.98 . ISSN   1538-3628. PMID   19805709.
  13. Frazier, Thomas W.; Thompson, Lee; Youngstrom, Eric A.; Law, Paul; Hardan, Antonio Y.; Eng, Charis; Morris, Nathan (2014-08-01). "A twin study of heritable and shared environmental contributions to autism". Journal of Autism and Developmental Disorders. 44 (8): 2013–2025. doi:10.1007/s10803-014-2081-2. ISSN   1573-3432. PMC   4104233 . PMID   24604525.
  14. Boomsma, Dorret; Busjahn, Andreas; Peltonen, Leena (2002-11-01). "Classical twin studies and beyond" (PDF). Nature Reviews Genetics. 3 (11): 872–882. doi:10.1038/nrg932. ISSN   1471-0056. PMID   12415317. S2CID   9318812.
  15. 1 2 3 4 5 6 7 8 Sullivan, Patrick F.; Daly, Mark J.; O'Donovan, Michael (2012-08-01). "Genetic architectures of psychiatric disorders: the emerging picture and its implications". Nature Reviews Genetics. 13 (8): 537–551. doi:10.1038/nrg3240. ISSN   1471-0056. PMC   4110909 . PMID   22777127.
  16. 1 2 3 4 5 Johnson, Keith A.; Fox, Nick C.; Sperling, Reisa A.; Klunk, William E. (2012-04-01). "Brain Imaging in Alzheimer Disease". Cold Spring Harbor Perspectives in Medicine. 2 (4): a006213. doi:10.1101/cshperspect.a006213. ISSN   2157-1422. PMC   3312396 . PMID   22474610.
  17. 1 2 McCarroll, Steven A.; Feng, Guoping; Hyman, Steven E. (2014-06-01). "Genome-scale neurogenetics: methodology and meaning". Nature Neuroscience. 17 (6): 756–763. doi:10.1038/nn.3716. ISSN   1546-1726. PMC   4912829 . PMID   24866041.
  18. 1 2 "Opinion: The Present and Future of Neurogenomics | The Scientist Magazine®". The Scientist. Retrieved 2016-02-23.
  19. Malhotra, Dheeraj; Sebat, Jonathan (2012-03-16). "CNVs: harbingers of a rare variant revolution in psychiatric genetics". Cell. 148 (6): 1223–1241. doi:10.1016/j.cell.2012.02.039. ISSN   1097-4172. PMC   3351385 . PMID   22424231.
  20. McClellan, Jon; King, Mary-Claire (2010-06-23). "Genomic analysis of mental illness: a changing landscape". JAMA. 303 (24): 2523–2524. doi:10.1001/jama.2010.869. ISSN   1538-3598. PMID   20571020.
  21. Courchesne, E.; Yeung-Courchesne, R.; Press, G. A.; Hesselink, J. R.; Jernigan, T. L. (1988-05-26). "Hypoplasia of cerebellar vermal lobules VI and VII in autism". The New England Journal of Medicine. 318 (21): 1349–1354. doi:10.1056/NEJM198805263182102. ISSN   0028-4793. PMID   3367935.
  22. Courchesne, E.; Karns, C. M.; Davis, H. R.; Ziccardi, R.; Carper, R. A.; Tigue, Z. D.; Chisum, H. J.; Moses, P.; Pierce, K. (2001-07-24). "Unusual brain growth patterns in early life in patients with autistic disorder: an MRI study". Neurology. 57 (2): 245–254. doi:10.1212/wnl.57.2.245. ISSN   0028-3878. PMID   11468308. S2CID   11234938.
  23. Quiroz, Yakeel T.; Schultz, Aaron P.; Chen, Kewei; Protas, Hillary D.; Brickhouse, Michael; Fleisher, Adam S.; Langbaum, Jessica B.; Thiyyagura, Pradeep; Fagan, Anne M. (2015-08-01). "Brain Imaging and Blood Biomarker Abnormalities in Children With Autosomal Dominant Alzheimer Disease: A Cross-Sectional Study". JAMA Neurology. 72 (8): 912–919. doi:10.1001/jamaneurol.2015.1099. ISSN   2168-6157. PMC   4625544 . PMID   26121081.
  24. "UC San Diego Autism Center of Excellence". autism-center.ucsd.edu. Retrieved 2016-02-24.
  25. Ray, Mohana; Goldstein, Steve; Zhou, Shiguo; Potamousis, Konstantinos; Sarkar, Deepayan; Newton, Michael A; Esterberg, Elizabeth; Kendziorski, Christina; Bogler, Oliver (2013-07-26). "Discovery of structural alterations in solid tumor oligodendroglioma by single molecule analysis". BMC Genomics. 14 (1): 505. doi:10.1186/1471-2164-14-505. PMC   3727977 . PMID   23885787.
  26. Colen, Rivka R.; Vangel, Mark; Wang, Jixin; Gutman, David A.; Hwang, Scott N.; Wintermark, Max; Jain, Rajan; Jilwan-Nicolas, Manal; Chen, James Y. (2014-01-01). "Imaging genomic mapping of an invasive MRI phenotype predicts patient outcome and metabolic dysfunction: a TCGA glioma phenotype research group project". BMC Medical Genomics. 7: 30. doi:10.1186/1755-8794-7-30. ISSN   1755-8794. PMC   4057583 . PMID   24889866.
  27. Lipska, Barbara K.; Deep-Soboslay, Amy; Weickert, Cynthia Shannon; Hyde, Thomas M.; Martin, Catherine E.; Herman, Mary M.; Kleinman, Joel E. (2006-09-15). "Critical Factors in Gene Expression in Postmortem Human Brain: Focus on Studies in Schizophrenia". Biological Psychiatry. 60 (6): 650–658. doi:10.1016/j.biopsych.2006.06.019. PMID   16997002. S2CID   39379510.
  28. Stan, Ana D.; Ghose, Subroto; Gao, Xue-Min; Roberts, Rosalinda C.; Lewis-Amezcua, Kelly; Hatanpaa, Kimmo J.; Tamminga, Carol A. (2006-12-06). "Human postmortem tissue: What quality markers matter?". Brain Research. 1123 (1): 1–11. doi:10.1016/j.brainres.2006.09.025. PMC   1995236 . PMID   17045977.
  29. Duric, Vanja; Banasr, Mounira; Stockmeier, Craig A.; Simen, Arthur A.; Newton, Samuel S.; Overholser, James C.; Jurjus, George J.; Dieter, Lesa; Duman, Ronald S. (2013-02-01). "Altered expression of synapse and glutamate related genes in post-mortem hippocampus of depressed subjects". International Journal of Neuropsychopharmacology. 16 (1): 69–82. doi:10.1017/S1461145712000016. ISSN   1461-1457. PMC   3414647 . PMID   22339950.
  30. Nagy, Corina; Maheu, Marissa; Lopez, Juan Pablo; Vaillancourt, Kathryn; Cruceanu, Cristiana; Gross, Jeffrey A.; Arnovitz, Mitchell; Mechawar, Naguib; Turecki, Gustavo (2015-05-01). "Effects of Postmortem Interval on Biomolecule Integrity in the Brain". Journal of Neuropathology & Experimental Neurology. 74 (5): 459–469. doi: 10.1097/NEN.0000000000000190 . ISSN   0022-3069. PMID   25868148.
  31. Darmanis, Spyros; Sloan, Steven A.; Zhang, Ye; Enge, Martin; Caneda, Christine; Shuer, Lawrence M.; Gephart, Melanie G. Hayden; Barres, Ben A.; Quake, Stephen R. (2015-06-09). "A survey of human brain transcriptome diversity at the single cell level". Proceedings of the National Academy of Sciences. 112 (23): 7285–7290. Bibcode:2015PNAS..112.7285D. doi: 10.1073/pnas.1507125112 . ISSN   0027-8424. PMC   4466750 . PMID   26060301.
  32. Cicognola, Claudia; Chiasserini, Davide; Parnetti, Lucilla (2015-06-29). "Preanalytical Confounding Factors in the Analysis of Cerebrospinal Fluid Biomarkers for Alzheimer's Disease: The Issue of Diurnal Variation". Frontiers in Neurology. 6: 143. doi: 10.3389/fneur.2015.00143 . ISSN   1664-2295. PMC   4483516 . PMID   26175714.
  33. Rinkwitz, Silke; Mourrain, Philippe; Becker, Thomas S. (2011-02-01). "Zebrafish: an integrative system for neurogenomics and neurosciences". Progress in Neurobiology. 93 (2): 231–243. doi:10.1016/j.pneurobio.2010.11.003. ISSN   1873-5118. PMID   21130139. S2CID   34169168.
  34. Cañestro, Cristian; Postlethwait, John H. (2007-05-15). "Development of a chordate anterior-posterior axis without classical retinoic acid signaling". Developmental Biology. 305 (2): 522–538. doi: 10.1016/j.ydbio.2007.02.032 . ISSN   0012-1606. PMID   17397819.
  35. Tallafuss, Alexandra; Trepman, Alissa; Eisen, Judith S. (2009-12-01). "DeltaA mRNA and protein distribution in the zebrafish nervous system". Developmental Dynamics. 238 (12): 3226–3236. doi:10.1002/dvdy.22136. ISSN   1097-0177. PMC   2882441 . PMID   19924821.
  36. Russek-Blum, Niva; Gutnick, Amos; Nabel-Rosen, Helit; Blechman, Janna; Staudt, Nicole; Dorsky, Richard I.; Houart, Corinne; Levkowitz, Gil (2008-10-01). "Dopaminergic neuronal cluster size is determined during early forebrain patterning". Development. 135 (20): 3401–3413. doi:10.1242/dev.024232. ISSN   0950-1991. PMC   2692842 . PMID   18799544.
  37. Reimer, Michell M.; Sörensen, Inga; Kuscha, Veronika; Frank, Rebecca E.; Liu, Chong; Becker, Catherina G.; Becker, Thomas (2008-08-20). "Motor neuron regeneration in adult zebrafish". The Journal of Neuroscience. 28 (34): 8510–8516. doi: 10.1523/JNEUROSCI.1189-08.2008 . ISSN   1529-2401. PMC   6671064 . PMID   18716209.
  38. White, Richard; Rose, Kristin; Zon, Leonard (2013-09-01). "Zebrafish cancer: the state of the art and the path forward". Nature Reviews Cancer. 13 (9): 624–636. doi:10.1038/nrc3589. ISSN   1474-175X. PMC   6040891 . PMID   23969693.
  39. Darland, T.; Dowling, J. E. (2001). "Behavioral screening for cocaine sensitivity in mutagenized zebrafish". Proc. Natl. Acad. Sci. USA. 98 (20): 11691–11696. Bibcode:2001PNAS...9811691D. doi: 10.1073/pnas.191380698 . PMC   58791 . PMID   11553778.
  40. Lockwood, B., Bjerke, S., Kobayashi, K. & Guo, S. "Acute effects of alcohol on larval zebrafish
    a genetic system for large-scale screening" Pharmacol. Biochem. Behav 2004; 77, 647–654
  41. Bourgeron, Thomas (2015-09-01). "From the genetic architecture to synaptic plasticity in autism spectrum disorder". Nature Reviews Neuroscience. 16 (9): 551–563. doi:10.1038/nrn3992. ISSN   1471-003X. PMID   26289574. S2CID   12742356.
  42. 1 2 Just, Marcel Adam; Cherkassky, Vladimir L.; Keller, Timothy A.; Kana, Rajesh K.; Minshew, Nancy J. (2007-04-01). "Functional and Anatomical Cortical Underconnectivity in Autism: Evidence from an fMRI Study of an Executive Function Task and Corpus Callosum Morphometry". Cerebral Cortex. 17 (4): 951–961. doi:10.1093/cercor/bhl006. ISSN   1047-3211. PMC   4500121 . PMID   16772313.
  43. Oddi, D.; Crusio, W. E.; D’Amato, F. R.; Pietropaolo, S. (2013-08-15). "Monogenic mouse models of social dysfunction: Implications for autism". Behavioural Brain Research. SI:Neurobiology of Autism. 251: 75–84. doi:10.1016/j.bbr.2013.01.002. PMID   23327738. S2CID   10384899.
  44. Gould, Stephen E.; Junttila, Melissa R.; de Sauvage, Frederic J. (2015-05-01). "Translational value of mouse models in oncology drug development". Nature Medicine. 21 (5): 431–439. doi:10.1038/nm.3853. ISSN   1078-8956. PMID   25951530. S2CID   5789096.
  45. Jones, CA; Watson, DJG; Fone, KCF (2011-10-01). "Animal models of schizophrenia". British Journal of Pharmacology. 164 (4): 1162–1194. doi:10.1111/j.1476-5381.2011.01386.x. ISSN   0007-1188. PMC   3229756 . PMID   21449915.
  46. Mortazavi, Ali; Williams, Brian A.; McCue, Kenneth; Schaeffer, Lorian; Wold, Barbara (2008-07-01). "Mapping and quantifying mammalian transcriptomes by RNA-Seq". Nature Methods. 5 (7): 621–628. doi:10.1038/nmeth.1226. ISSN   1548-7105. PMID   18516045. S2CID   205418589.
  47. 1 2 3 4 Muhie, Seid; Gautam, Aarti; Meyerhoff, James; Chakraborty, Nabarun; Hammamieh, Rasha; Jett, Marti (2015-02-28). "Brain transcriptome profiles in mouse model simulating features of post-traumatic stress disorder". Molecular Brain. 8 (1): 14. doi:10.1186/s13041-015-0104-3. PMC   4359441 . PMID   25888136.
  48. Nutt, David J.; Malizia, Andrea L. (2004-01-01). "Structural and functional brain changes in posttraumatic stress disorder". The Journal of Clinical Psychiatry. 65 Suppl 1: 11–17. ISSN   0160-6689. PMID   14728092.
  49. Lazic, Stanley E; Essioux, Laurent (2013-03-22). "Improving basic and translational science by accounting for litter-to-litter variation in animal models". BMC Neuroscience. 14: 37. doi:10.1186/1471-2202-14-37. ISSN   1471-2202. PMC   3661356 . PMID   23522086.
  50. Crawley, Jacqueline N. (2012-09-01). "Translational animal models of autism and neurodevelopmental disorders". Dialogues in Clinical Neuroscience. 14 (3): 293–305. doi:10.31887/DCNS.2012.14.3/jcrawley. ISSN   1294-8322. PMC   3513683 . PMID   23226954.
  51. Gotham, Katherine; Bishop, Somer L.; Hus, Vanessa; Huerta, Marisela; Lund, Sabata; Buja, Andreas; Krieger, Abba; Lord, Catherine (2013-02-01). "Exploring the Relationship Between Anxiety and Insistence on Sameness in Autism Spectrum Disorders". Autism Research. 6 (1): 33–41. doi:10.1002/aur.1263. ISSN   1939-3806. PMC   4373663 . PMID   23258569.
  52. Ramasamy, Adaikalavan; Trabzuni, Daniah; Guelfi, Sebastian; Varghese, Vibin; Smith, Colin; Walker, Robert; De, Tisham; UK Brain Expression Consortium; North American Brain Expression Consortium (2014-10-01). "Genetic variability in the regulation of gene expression in ten regions of the human brain". Nature Neuroscience. 17 (10): 1418–1428. doi:10.1038/nn.3801. ISSN   1097-6256. PMC   4208299 . PMID   25174004.
  53. 1 2 Naumova, Oksana Yu.; Lee, Maria; Rychkov, Sergei Yu.; Vlasova, Natalia V.; Grigorenko, Elena L. (2013-01-01). "Gene Expression in the Human Brain: The Current State of the Study of Specificity and Spatiotemporal Dynamics". Child Development. 84 (1): 76–88. doi:10.1111/cdev.12014. ISSN   1467-8624. PMC   3557706 . PMID   23145569.
  54. 1 2 Lodato, Michael A.; Woodworth, Mollie B.; Lee, Semin; Evrony, Gilad D.; Mehta, Bhaven K.; Karger, Amir; Lee, Soohyun; Chittenden, Thomas W.; D'Gama, Alissa M. (2015-10-02). "Somatic mutation in single human neurons tracks developmental and transcriptional history". Science. 350 (6256): 94–98. Bibcode:2015Sci...350...94L. doi:10.1126/science.aab1785. ISSN   1095-9203. PMC   4664477 . PMID   26430121.
  55. Miller, Jeremy A.; Ding, Song-Lin; Sunkin, Susan M.; Smith, Kimberly A.; Ng, Lydia; Szafer, Aaron; Ebbert, Amanda; Riley, Zackery L.; Royall, Joshua J. (2014-04-10). "Transcriptional landscape of the prenatal human brain". Nature. 508 (7495): 199–206. Bibcode:2014Natur.508..199M. doi:10.1038/nature13185. ISSN   0028-0836. PMC   4105188 . PMID   24695229.
  56. Carroll, Sean B. (April 2003). "Genetics and the making of Homo sapiens". Nature. 422 (6934): 849–857. doi:10.1038/nature01495. PMID   12712196. S2CID   4333307.
  57. Hill, Robert Sean; Walsh, Christopher A. (2005-09-01). "Molecular insights into human brain evolution". Nature. 437 (7055): 64–67. Bibcode:2005Natur.437...64H. doi:10.1038/nature04103. ISSN   1476-4687. PMID   16136130. S2CID   4406401.
  58. 1 2 Rakic, Pasko (2009-10-01). "Evolution of the neocortex: a perspective from developmental biology". Nature Reviews Neuroscience. 10 (10): 724–735. doi:10.1038/nrn2719. ISSN   1471-003X. PMC   2913577 . PMID   19763105.
  59. Geschwind, Daniel H.; Rakic, Pasko (2013-10-30). "Cortical evolution: judge the brain by its cover". Neuron. 80 (3): 633–647. doi:10.1016/j.neuron.2013.10.045. ISSN   1097-4199. PMC   3922239 . PMID   24183016.
  60. Calarco, John A.; Xing, Yi; Cáceres, Mario; Calarco, Joseph P.; Xiao, Xinshu; Pan, Qun; Lee, Christopher; Preuss, Todd M.; Blencowe, Benjamin J. (2007-11-15). "Global analysis of alternative splicing differences between humans and chimpanzees". Genes & Development. 21 (22): 2963–2975. doi:10.1101/gad.1606907. ISSN   0890-9369. PMC   2049197 . PMID   17978102.
  61. Zhang, Xiao-Ou; Yin, Qing-Fei; Wang, Hai-Bin; Zhang, Yang; Chen, Tian; Zheng, Ping; Lu, Xuhua; Chen, Ling-Ling; Yang, Li (2014-01-01). "Species-specific alternative splicing leads to unique expression of sno-lncRNAs". BMC Genomics. 15: 287. doi:10.1186/1471-2164-15-287. ISSN   1471-2164. PMC   4234469 . PMID   24734784.
  62. Somel, Mehmet; Liu, Xiling; Khaitovich, Philipp (2013-02-01). "Human brain evolution: transcripts, metabolites and their regulators". Nature Reviews Neuroscience. 14 (2): 112–127. doi:10.1038/nrn3372. ISSN   1471-003X. PMID   23324662. S2CID   12572416.
  63. Qureshi, Irfan A.; Mehler, Mark F. (2012-08-01). "Emerging roles of non-coding RNAs in brain evolution, development, plasticity and disease". Nature Reviews Neuroscience. 13 (8): 528–541. doi:10.1038/nrn3234. ISSN   1471-003X. PMC   3478095 . PMID   22814587.
  64. 1 2 Geschwind, Daniel H.; Rakic, Pasko (2013-10-30). "Cortical Evolution: Judge the Brain by Its Cover". Neuron. 80 (3): 633–647. doi:10.1016/j.neuron.2013.10.045. ISSN   0896-6273. PMC   3922239 . PMID   24183016.
  65. 1 2 Enard, Wolfgang (2014-01-01). "Comparative genomics of brain size evolution". Frontiers in Human Neuroscience. 8: 345. doi: 10.3389/fnhum.2014.00345 . PMC   4033227 . PMID   24904382.
  66. Wang, Guang-Zhong; Konopka, Genevieve (2013-06-01). "Decoding human gene expression signatures in the brain". Transcription. 4 (3): 102–108. doi:10.4161/trns.24885. ISSN   2154-1272. PMC   4042582 . PMID   23665540.
  67. Kirov, G.; Pocklington, A. J.; Holmans, P.; Ivanov, D.; Ikeda, M.; Ruderfer, D.; Moran, J.; Chambert, K.; Toncheva, D. (2012-02-01). "De novo CNV analysis implicates specific abnormalities of postsynaptic signalling complexes in the pathogenesis of schizophrenia". Molecular Psychiatry. 17 (2): 142–153. doi:10.1038/mp.2011.154. ISSN   1476-5578. PMC   3603134 . PMID   22083728.
  68. Bierut, Laura J.; Agrawal, Arpana; Bucholz, Kathleen K.; Doheny, Kimberly F.; Laurie, Cathy; Pugh, Elizabeth; Fisher, Sherri; Fox, Louis; Howells, William (2010-03-16). "A genome-wide association study of alcohol dependence". Proceedings of the National Academy of Sciences. 107 (11): 5082–5087. Bibcode:2010PNAS..107.5082B. doi: 10.1073/pnas.0911109107 . ISSN   0027-8424. PMC   2841942 . PMID   20202923.
  69. Juraeva, Dilafruz; Treutlein, Jens; Scholz, Henrike; Frank, Josef; Degenhardt, Franziska; Cichon, Sven; Ridinger, Monika; Mattheisen, Manuel; Witt, Stephanie H. (2015-01-01). "XRCC5 as a risk gene for alcohol dependence: evidence from a genome-wide gene-set-based analysis and follow-up studies in Drosophila and humans". Neuropsychopharmacology. 40 (2): 361–371. doi:10.1038/npp.2014.178. ISSN   1740-634X. PMC   4443948 . PMID   25035082.
  70. The 1000 Genomes Project Consortium (2015-10-01). "A global reference for human genetic variation". Nature. 526 (7571): 68–74. Bibcode:2015Natur.526...68T. doi:10.1038/nature15393. ISSN   0028-0836. PMC   4750478 . PMID   26432245.
  71. 1 2 Ripke, Stephan; O'Dushlaine, Colm; Chambert, Kimberly; Moran, Jennifer L.; Kähler, Anna K.; Akterin, Susanne; Bergen, Sarah E.; Collins, Ann L.; Crowley, James J. (2013-10-01). "Genome-wide association analysis identifies 13 new risk loci for schizophrenia". Nature Genetics. 45 (10): 1150–1159. doi:10.1038/ng.2742. ISSN   1546-1718. PMC   3827979 . PMID   23974872.
  72. Lee, Phil H.; O'Dushlaine, Colm; Thomas, Brett; Purcell, Shaun M. (2012-07-01). "INRICH: interval-based enrichment analysis for genome-wide association studies". Bioinformatics. 28 (13): 1797–1799. doi:10.1093/bioinformatics/bts191. ISSN   1367-4811. PMC   3381960 . PMID   22513993.
  73. Morris, Andrew P; Voight, Benjamin F; Teslovich, Tanya M; Ferreira, Teresa; Segrè, Ayellet V; Steinthorsdottir, Valgerdur; Strawbridge, Rona J; Khan, Hassan; Grallert, Harald (2012-09-01). "Large-scale association analysis provides insights into the genetic architecture and pathophysiology of type 2 diabetes". Nature Genetics. 44 (9): 981–990. doi:10.1038/ng.2383. ISSN   1061-4036. PMC   3442244 . PMID   22885922.
  74. Gillis, Jesse; Pavlidis, Paul (2012). ""Guilt by Association" Is the Exception Rather Than the Rule in Gene Networks". PLOS Computational Biology. 8 (3): e1002444. Bibcode:2012PLSCB...8E2444G. doi:10.1371/journal.pcbi.1002444. PMC   3315453 . PMID   22479173.
  75. 1 2 McMahon, Francis J.; Insel, Thomas R. (2012-06-07). "Pharmacogenomics and Personalized Medicine in Neuropsychiatry". Neuron. 74 (5): 773–776. doi:10.1016/j.neuron.2012.05.004. PMC   3407812 . PMID   22681682.
  76. Brandler, William M.; Sebat, Jonathan (2015-01-01). "From De Novo Mutations to Personalized Therapeutic Interventions in Autism". Annual Review of Medicine. 66 (1): 487–507. doi:10.1146/annurev-med-091113-024550. PMID   25587659.
  77. Banks, William A. (2010-10-01). "Mouse models of neurological disorders: a view from the blood-brain barrier". Biochimica et Biophysica Acta (BBA) - Molecular Basis of Disease. 1802 (10): 881–888. doi:10.1016/j.bbadis.2009.10.011. ISSN   0006-3002. PMC   2891624 . PMID   19879356.
  78. 1 2 3 4 Biesecker, Barbara Bowles; Peay, Holly Landrum (2013-08-01). "Genomic sequencing for psychiatric disorders: Promise and challenge". The International Journal of Neuropsychopharmacology. 16 (7): 1667–1672. doi:10.1017/S146114571300014X. ISSN   1461-1457. PMC   3703499 . PMID   23575420.
  79. Bunnik, Eline M; Schermer, Maartje HN; Janssens, A Cecile JW (2012-01-19). "The role of disease characteristics in the ethical debate on personal genome testing". BMC Medical Genomics. 5 (1): 4. doi:10.1186/1755-8794-5-4. PMC   3293088 . PMID   22260407.