Skip to content

Abnormality Of Movement

Abnormality of movement encompasses a wide range of conditions characterized by involuntary or impaired voluntary movements, significantly impacting an individual’s physical function and quality of life. These abnormalities can manifest in various forms, from subtle eye movement dysfunctions to more pronounced motor disturbances. Understanding the underlying mechanisms of these conditions is crucial for diagnosis, treatment, and improving patient outcomes.

Movement is a complex biological process orchestrated by intricate neural networks involving the brain, spinal cord, and peripheral nervous system. Genetic factors play a significant role in predisposing individuals to various movement abnormalities. Genome-wide association studies (GWAS) have identified numerous susceptibility loci associated with abnormal movements. For instance, research into eye movement dysfunction, often considered an endophenotype of schizophrenia, has pinpointed specific single nucleotide polymorphisms (SNPs) in genomic regions like 1q21.3, 7p12.1, and 20q13.12.[1] These SNPs, frequently located in non-coding regions, can act as expression quantitative trait loci (eQTLs), influencing the expression levels of nearby genes such as THEM4, S100A10, and CDH22 in specific brain regions like the frontal cortex and cerebellum. [1] These genes are implicated in neurotransmission and brain morphogenesis, highlighting the molecular underpinnings of motor control. [1]

Another example is periodic leg movement in sleep, which has been linked to polymorphisms in genes like BTBD9, TOX3/BC034767, MEIS1, MAP2K5/SKOR1, and PTPRD. [2] Furthermore, movement-related adverse drug reactions (ADRs) to antipsychotics like aripiprazole have been associated with genetic variants in genes such as SLC22A8, ADCYAP1R1, KCNIP4, SMAD9, NAP1L4, and ERBB4, suggesting important roles for ion transporters and channels in these effects. [3] The high heritability of certain eye movement characteristics, such as smooth pursuit eye movement, further underscores the strong genetic component in these conditions. [1]

Abnormalities of movement are clinically relevant across numerous medical fields. In psychiatry, specific eye movement dysfunctions, such as impaired exploratory eye movements (EEM) or smooth pursuit eye movements (SPEM), are recognized as robust biological markers for schizophrenia, often persisting even when clinical symptoms are alleviated[4]. [1]Identifying genetic loci associated with these dysfunctions can aid in early diagnosis and understanding disease pathophysiology.[4]

In pharmacology, genetic predispositions to movement-related adverse drug reactions, particularly with antipsychotic medications, are critical for personalized medicine. Genetic screening for SNPs in genes like SLC22A8 and KCNIP4 can help predict the risk of developing conditions such as akathisia, enabling clinicians to adjust treatment strategies and improve patient safety. [3] Movement disorders also significantly impact sleep, as seen in periodic leg movement in sleep, affecting overall health and well-being. [2]

The social importance of addressing abnormalities of movement is profound. These conditions can severely impair an individual’s ability to perform daily activities, maintain employment, and engage in social interactions, leading to reduced independence and a diminished quality of life. The economic burden associated with long-term care, specialized therapies, and lost productivity is substantial. Genetic research into these conditions offers hope for developing more effective diagnostic tools, targeted pharmacological interventions, and personalized treatment plans, moving towards precision medicine. By elucidating the genetic architecture of movement abnormalities, researchers aim to uncover novel therapeutic targets and ultimately improve the lives of affected individuals and their families. Continued research, including large-scale international studies, is essential to validate findings and translate genetic insights into clinical practice. [1]

Methodological and Statistical Constraints

Section titled “Methodological and Statistical Constraints”

Genetic studies of abnormality of movement are subject to several methodological and statistical limitations that can influence the robustness and interpretation of findings. While large cohorts, such as those within the UK Biobank, provide substantial statistical power, specific analyses or sub-cohorts may still operate with smaller sample sizes, potentially leading to inflated effect sizes or insufficient power to detect subtle genetic associations.[5] Furthermore, the replication of initial findings across independent cohorts is critical for validation, yet challenges can arise from differences in genotyping platforms, imputation quality (e.g., exclusion of variants with R2 < 0.6), or the absence of specific lead variants in replication datasets, which can impact the transferability and reliability of reported associations. [5]

Rigorous statistical filtering is applied in genome-wide association studies (GWAS) to ensure data quality, including the exclusion of single nucleotide polymorphisms (SNPs) with low call rates, deviations from Hardy–Weinberg equilibrium (p < 10−6), or low minor allele frequency (MAF < 0.01 or MAF < 0.1% for imputed variants).[6] While essential, these stringent filters might inadvertently exclude true rare variants or those with modest effects, potentially contributing to the phenomenon of missing heritability. Moreover, the presence of population structure or unmodeled relatedness within study populations can confound association signals, necessitating careful adjustment through methods like principal component analysis (PCA) or mixed models to prevent spurious findings. [6]

Phenotypic Heterogeneity and Measurement Precision

Section titled “Phenotypic Heterogeneity and Measurement Precision”

The characterization and measurement of abnormality of movement present inherent complexities that contribute to phenotypic heterogeneity across studies. Traits related to movement can be derived from diverse sources, such as structural magnetic resonance imaging (MRI), diffusion MRI, or task functional MRI, each employing distinct acquisition protocols and being susceptible to varying degrees of measurement error.[5] For instance, some traits may exhibit a limited range of unique values or, conversely, an excessively large number, occasionally suggesting data errors, thereby necessitating rigorous quality control processes and normalization techniques like rank-based inverse-normal transformation or quantile normalization. [5] Consistency in applying specific scoring criteria, such as those used for periodic leg movements in sleep (PLMS), is paramount across different research settings to ensure the comparability and validity of identified genetic associations. [2]

Beyond measurement, the expression of movement abnormalities is influenced by numerous non-genetic factors that can confound genetic analyses. Researchers routinely adjust for known confounders such as age, sex, head motion, head volume, head position, temporal imaging effects, imaging center, and genetic principal components. [5] However, despite these adjustments, residual confounding from environmental factors not fully accounted for, or unmodeled gene-environment interactions, could still obscure genuine genetic signals or introduce bias into effect estimates. Additionally, interpreting the functional implications of identified genetic variants, particularly those located in non-coding regions, remains a significant challenge, requiring further investigation to elucidate their precise roles in protein function or gene regulatory pathways. [1]

Generalizability and Unexplained Heritability

Section titled “Generalizability and Unexplained Heritability”

A notable limitation in the genetic study of abnormality of movement is the predominant focus on populations of European descent, including “self-reported white British individuals,” “European ancestry,” and “Northern European populations”.[7] While efforts are made to identify and exclude genetic outliers or conduct cross-ancestry analyses, findings derived primarily from these cohorts may not be directly generalizable to other ancestral groups due to variations in allele frequencies, linkage disequilibrium patterns, and environmental exposures. [3] This lack of ancestral diversity in study populations can restrict the discovery of novel genetic variants specific to underrepresented populations and impede the development of equitable genetic insights and clinical applications globally.

Despite significant advancements in GWAS, a substantial portion of the heritability for complex traits, including various movement abnormalities, often remains unexplained, a phenomenon known as “missing heritability”. [8] This gap may stem from several factors, including the aggregate effect of numerous common variants, each contributing a very small effect, the involvement of rare variants, structural variants, complex gene-gene interactions (epistasis), gene-environment interactions, and non-additive genetic effects such as dominance. [3]While some studies have begun to explore non-additive effects, a comprehensive dissection of these intricate genetic architectures and their dynamic interplay with environmental factors represents a significant knowledge gap, hindering a complete understanding of the genetic etiology of abnormality of movement.

Genetic variations can significantly influence an individual’s susceptibility to various health conditions, including those that manifest as abnormalities of movement, by altering gene function and physiological pathways. Variants within genes like BANK1 and SLC39A8 exemplify how diverse genetic roles can converge to impact complex traits. BANK1 encodes a B-cell scaffold protein, playing a crucial role in immune cell signaling and activation; dysregulation here can lead to autoimmune conditions that sometimes present with neurological or neuromuscular symptoms affecting movement . Similarly, SLC39A8 is involved in transporting essential trace elements, particularly zinc, across cell membranes, which is vital for numerous enzymatic processes and neurological health; impaired zinc homeostasis due to variants like rs35518360 and rs35225200 can contribute to neurodevelopmental issues and motor dysfunction. These single nucleotide polymorphisms (SNPs) can alter protein structure or expression, subtly shifting the balance of cellular processes and potentially contributing to a spectrum of movement-related challenges .

The FTOgene, primarily known for its strong association with obesity and metabolic traits, also holds relevance for neurological function and movement control through its role as an RNA demethylase. Variants withinFTO, such as rs1421085 , are implicated in altering metabolic pathways that can indirectly influence brain health and motor coordination . Beyond its metabolic links, FTO expression in the brain suggests a direct role in neuronal function, including dopaminergic pathways that are critical for motor control, potentially affecting an individual’s predisposition to conditions characterized by atypical movements or gait disturbances. The widespread impact of FTO on energy balance and neural circuits means that its variants can have broad, pleiotropic effects on overall physiological resilience, including the ability to perform coordinated movements.

Another set of variants with potential implications for movement abnormalities involves the non-coding RNA RN7SL643P and the serotonin receptor gene HTR1E. While RN7SL643P belongs to a family of small cytoplasmic RNAs with diverse functions, including stress response and protein synthesis regulation, its direct contribution to movement disorders is an area of ongoing investigation. In contrast, HTR1E encodes a receptor for serotonin, a neurotransmitter fundamental to mood, cognition, and motor function . Variations like rs9450487 in HTR1Ecan modify the efficiency of serotonin signaling, which in turn can affect the excitability of motor neurons and the coordination of muscle movements, potentially contributing to conditions like dystonia, tremors, or other motor control impairments. Altered serotonin pathways are frequently implicated in neurological and psychiatric disorders that feature prominent motor symptoms.

Finally, the CNTNAP5 gene, or Contactin Associated Protein Family Member 5, is crucial for proper neuronal development and function, particularly in establishing and maintaining synaptic connections and cell adhesion within the brain. Variants in CNTNAP5, such as rs181670972 , are frequently associated with neurodevelopmental disorders, including autism spectrum disorder and intellectual disability, which often manifest with significant motor challenges like poor coordination, repetitive movements, or atypical gait . Disruptions inCNTNAP5 can impair the intricate wiring of neural circuits essential for precise motor execution and learning. Therefore, understanding the impact of these genetic variations is key to elucidating the underlying mechanisms of movement abnormalities and developing targeted interventions.

RS IDGeneRelated Traits
rs35518360
rs35225200
BANK1 - SLC39A8schizophrenia
ST2 protein measurement
body mass index
health trait
cholesteryl esters:totallipids ratio, intermediate density lipoprotein measurement
rs1421085 FTObody mass index
obesity
energy intake
pulse pressure measurement
lean body mass
rs9450487 RN7SL643P - HTR1Eabnormality of movement
rs181670972 CNTNAP5abnormality of movement

Abnormality of movement arises from a complex interplay of genetic predispositions, physiological changes, and external influences. Understanding these causal factors is crucial for diagnosis and intervention.

Genetic Architecture and Neural Mechanisms

Section titled “Genetic Architecture and Neural Mechanisms”

Genetic factors play a significant role in the etiology of movement abnormalities, with both inherited variants and polygenic risk contributing to susceptibility. For instance, periodic leg movements in sleep (PLMS) are associated with polymorphisms in genes such as BTBD9, TOX3/BC034767, MEIS1, MAP2K5/SKOR1, and PTPRD. [2] Similarly, eye movement dysfunction, including smooth pursuit eye movements (SPEMs) and predictive pursuit gain, exhibits high heritability. [1]Genome-wide association studies (GWAS) have identified susceptibility loci for eye movement dysfunction in schizophrenia, with specific single nucleotide polymorphisms (SNPs) at 1q21.3, 7p12.1, and 20q13.12 being relevant to horizontal position gain (HPG).[1] These SNPs, often intronic or intergenic, are found near genes like THEM4, S100A10, and CDH22, which are critical for neurotransmission and brain morphogenesis. [1]

The functional implications of these genetic associations highlight underlying neural mechanisms. For example, THEM4 (also known as CTMP) acts as a negative regulator of the AKT1 gene, influencing signal transduction in neurons, while S100A10 modulates the transport of neurotransmitters such as calcium ions and serotonin. [1] CDH22 is a cell-adhesion factor predominantly expressed in the brain, contributing to its structural development. [1] These genes underscore how specific molecular pathways and brain regions are intrinsically linked to the regulation of eye movements. [1]Beyond these, exploratory eye movement (EEM) dysfunction in schizophrenia is associated with a susceptibility locus at 5q21.3, and abnormal SPEMs have been linked to genes includingCOMT, ZDHHC8, ERBB4, RANBP1, and NRG1. [4] The observation of EEM impairments in healthy siblings of schizophrenic patients further supports the notion of a strong genetic predisposition acting as a biological marker for this movement abnormality. [4]

Section titled “Physiological Context and Age-Related Factors”

The physiological context and an individual’s age significantly contribute to the manifestation and severity of movement abnormalities. The average periodic leg movement index (PLMI), a measure of leg movement abnormality during sleep, demonstrates variation across different 5-year age intervals. [2] This suggests that age-related physiological changes can influence the expression of such movements. Furthermore, the occurrence of sleep-disordered breathing, which can impact overall physiological regulation and potentially contribute to movement disturbances, is notably prevalent among middle-aged adults. [9]Studies in older populations have also investigated associations between sleep architecture, sleep-disordered breathing, and cognitive function, indicating a broader physiological context in which movement abnormalities may arise or be exacerbated.[10]

External factors such as medication effects and the presence of comorbid conditions are important contributors to movement abnormalities. Certain pharmacological agents can induce or worsen movement-related adverse effects. For example, aripiprazole monotherapy has been associated with such effects, particularly in individuals carrying specific genetic variants. [3] Carriers of the G allele of rs4149181 in SLC22A8 or the decreased function allele of CYP2D6*10 exhibit higher adverse reactions to aripiprazole. [3] These drug-induced abnormalities are often linked to genes involved in ion transport and ion channel function, including SLC22A8, KCNIP4, KCNA1, and CACNG1, highlighting the impact of medication on neuronal excitability and movement control. [3]

Moreover, various medical conditions can coexist with or directly contribute to movement abnormalities. Periodic leg movements during sleep are a notable feature in patients with narcolepsy/cataplexy and restless legs syndrome, although the specific patterns and periodicity of these movements may differ between the conditions.[11]Eye movement dysfunction, particularly exploratory eye movement (EEM) dysfunction, is a well-established physiological abnormality and a potential biological marker in individuals with schizophrenia, demonstrating a clear link between neurological and psychiatric disorders and specific movement impairments.[4]

Abnormalities of movement encompass a range of dysfunctions, from subtle eye tracking impairments to more pronounced periodic leg movements during sleep. These conditions often reflect underlying disruptions in complex biological systems, involving intricate genetic factors, neural pathways, molecular signaling, and cellular functions. Understanding these biological underpinnings is crucial for elucidating the etiology and developing targeted interventions for related neurological and psychiatric disorders.

Genetic Foundations of Movement Abnormalities

Section titled “Genetic Foundations of Movement Abnormalities”

Genetic factors play a significant role in the predisposition and manifestation of movement abnormalities. For instance, eye movement dysfunctions, such as smooth pursuit eye movements (SPEMs), are recognized as heritable characteristics. In schizophrenia, abnormal exploratory eye movements (EEM) are considered a biological marker and have been linked to genetic factors influencing the disorder’s pathology.[4] Specific genes have been associated with eye movement abnormalities, including COMT, ZDHHC8, ERBB4, RANBP1, NRG1, and MAN2A1. [4] These genes are implicated in various neural processes that contribute to the precise control of eye movements, highlighting the genetic complexity underlying these motor functions.

Beyond eye movements, genetic associations have also been identified for periodic leg movements during sleep (PLMS). Polymorphisms in genes such as BTBD9, TOX3/BC034767, MEIS1, MAP2K5/SKOR1, and PTPRD have been associated with PLMS. [2]Furthermore, genome-wide association studies (GWAS) have identified single nucleotide polymorphisms (SNPs) near genes likeTHEM4, S100A10, and CDH22 that are relevant to eye movement dysfunction. [1] These genetic markers, often found in non-coding regions, can influence gene expression patterns in brain regions critical for movement control, underscoring the role of regulatory elements and epigenetic modifications in shaping motor phenotypes. [1]

Movement abnormalities are often rooted in structural impairments and functional disabilities within specific brain regions. Exploratory eye movement (EEM) abnormalities, for example, have been attributed to such brain deficits in schizophrenia.[4] The precise coordination of movements relies on the integrity of neural substrates, with distinct brain areas responsible for different aspects of motor control. For instance, the V5 region of the visual cortex is critical for guiding smooth pursuit eye movements, and lesions in this area can lead to significant defects. [1]

The activity of various brain regions is intimately linked to the execution of eye movements. Studies have investigated the expression levels of genes in different brain regions, including the cerebellar cortex, frontal cortex, hippocampus, medulla, putamen, substantia nigra, thalamus, occipital cortex, intralobular white matter, and temporal cortex, revealing their interconnected roles in motor function. [1] Genes like CDH22 (cadherin 22), predominantly expressed in the brain, are crucial cell-adhesion factors involved in brain morphogenesis, indicating that proper brain development is fundamental for normal movement control. [1]

At the molecular and cellular level, a complex network of signaling pathways, key biomolecules, and cellular functions orchestrates movement. Signal transduction in neurons is vital, with proteins such as THEM4 (thioesterase superfamily member 4), also known as CTMP, acting as a negative regulator of the AKT1 gene, thereby influencing neuronal signaling. [1] The precise transport of neurotransmitters, including calcium ions and serotonin, is modulated by proteins like S100A10 (S100 calcium binding protein A10), which are essential for proper neural communication and motor command execution. [1]

Ion channels and transporters play a critical role in neuronal excitability and synaptic transmission, which are fundamental to movement. Genes such as SLC22A8, KCNIP4, KCNA1, and CACNG1 encode ion transporters or channels that are involved in molecular functions and cellular pathways related to movement. [3] For example, KCNIP4, a potassium channel-interacting protein, acts as a calcium regulator and has been implicated in a range of mental disorders including ADHD, autism, schizophrenia, bipolar disorder, and major depressive disorder, suggesting a common genetic basis involving K+ channel-related calcium regulation in various neuropsychiatric conditions.[3] Furthermore, RBFOX1 (RNA binding fox-1 homolog 1), a regulator of tissue-specific splicing, controls neuronal excitation in the mammalian brain, highlighting the importance of post-transcriptional regulation in maintaining neuronal function and preventing movement-related abnormalities. [12]

Pathophysiological Processes and Systemic Effects

Section titled “Pathophysiological Processes and Systemic Effects”

Movement abnormalities often serve as indicators of underlying pathophysiological processes and can have systemic consequences. Exploratory eye movement (EEM) dysfunction is considered a specific biological marker for schizophrenia, and importantly, these impairments can also be observed in healthy siblings of affected individuals, suggesting a genetic predisposition to the disorder.[4]The presence of periodic leg movements during sleep (PLMS) is associated with conditions such as narcolepsy/cataplexy and restless legs syndrome, reflecting disruptions in sleep-wake regulation and motor control during sleep.[2] These movements can involve alterations in cerebral and autonomic activity [13] and may be influenced by serotonergic antidepressants and even provoked by certain medications like mirtazapine. [14]

Beyond diagnostic markers, movement-related adverse drug reactions represent a significant clinical challenge. For instance, polymorphisms in genes such as CYP2D6*10 and SLC22A8 are associated with adverse reactions to antipsychotic medications like aripiprazole. [3] These genes, particularly ion transporter genes and their associated regulatory proteins, are critical in the molecular signaling pathways that contribute to movement-related side effects, such as akathisia. [3] The disruption of key biomolecules, such as the RBFOX1gene, has also been linked to abnormal phenotypes including mental retardation and epilepsy, further illustrating the broad impact of these molecular disruptions on neurological health and complex motor behaviors.[12]

Abnormalities of movement arise from a complex interplay of genetic, molecular, and cellular dysregulations affecting neuronal function and circuit integrity. These mechanisms span from specific signaling cascades and metabolic processes to broader regulatory networks and systems-level integration within the central nervous system. Understanding these pathways provides insights into the etiology of various movement dysfunctions, including drug-induced adverse effects and intrinsic neurological conditions.

Neurotransmitter and Ion Channel Regulation

Section titled “Neurotransmitter and Ion Channel Regulation”

Movement abnormalities often stem from dysregulation in neuronal excitability and communication, fundamentally governed by neurotransmitter systems and ion channels. Key genes identified in movement-related adverse drug reactions, such as SLC22A8, KCNIP4, KCNA1, and CACNG1, highlight the critical role of ion transporters and voltage-gated ion channels. [3] For instance, SLC22A8 functions as an organic anion transporter, influencing drug disposition and potentially neurochemical balance, while KCNIP4acts as a K+ channel-interacting protein, modulating the function of voltage-gated potassium channels like Kv4. Its interaction with Kv4 channels and presenilin 2 suggests involvement in regulating neuronal repolarization and calcium homeostasis, essential for precise motor control.[3] Dysregulation of these channels can alter neuronal firing patterns, contributing to conditions like akathisia.

Further illustrating the importance of signaling, ADCYAP1R, a receptor, is a candidate gene implicated in molecular signaling pathways of mental and nervous system diseases, underscoring the role of G-protein coupled receptor signaling in mediating neuronal responses. [3] Similarly, S100A10 modulates the transport of neurotransmitters, including calcium ions and serotonin, thereby affecting synaptic transmission and neuronal excitability. [1] Genes like COMT, ZDHHC8, ERBB4, RANBP1, and NRG1 have also been associated with abnormal smooth pursuit eye movements, suggesting their involvement in neurotransmitter metabolism, receptor signaling, and neuronal development that collectively influence oculomotor control. [4] Alterations in these signaling cascades, including receptor activation and subsequent intracellular events, can lead to imbalanced neurotransmission, manifesting as various forms of movement dysfunction.

Cellular Signaling and Structural Plasticity

Section titled “Cellular Signaling and Structural Plasticity”

Beyond immediate electrical signaling, the structural and functional integrity of neuronal circuits relies on intricate cellular signaling pathways that govern development, plasticity, and maintenance. Genes like THEM4 (CTMP) play a role as a negative regulator of the AKT1 gene, influencing signal transduction pathways crucial for neuronal survival, growth, and synaptic plasticity. [1] Dysregulation in this pathway can impair the proper functioning and connectivity of neurons, affecting coordinated movements. Moreover, cell adhesion molecules like CDH22, predominantly expressed in the brain, are vital for brain morphogenesis and establishing appropriate neuronal connections, the foundation for complex motor behaviors. [1]

The Bone Morphogenetic Protein (BMP) signaling pathway, involving transcriptional regulators likeSMAD9, also contributes to cellular regulation and development. [15] Such pathways involve receptor activation leading to intracellular cascades that regulate gene expression through transcription factors, influencing neuronal differentiation and circuit formation. Similarly, EphA7 modulates apical constriction during hindbrain neuroepithelium development, highlighting its role in the precise architectural formation of brain regions critical for motor control. [16] Disruptions in these fundamental developmental and signaling pathways can lead to emergent properties of dysfunctional neural networks, underlying various movement disorders.

Protein Homeostasis and Glycosylation Pathways

Section titled “Protein Homeostasis and Glycosylation Pathways”

The proper folding, modification, and trafficking of proteins are fundamental to cellular function, and their disruption can significantly impact movement. The N-glycan maturation pathway, for instance, is implicated in conditions like exploratory eye movement dysfunction in schizophrenia.[4]This pathway involves alpha-glucosidases that participate in glycoprotein folding, mediated by chaperones like calnexin and calreticulin.[4] MAN2A1 (alpha-mannosidase II), a key enzyme in this pathway, is highly expressed in specific brain regions and is essential for advanced brain functions, suggesting its role in maintaining the integrity of neuronal glycoproteins. [4]

Furthermore, PACS2 (phosphofurin acidic cluster sorting protein 2) regulates the trafficking of ion channels and mediates the subcellular distribution of calnexin, thereby influencing protein quality control and the availability of functional ion channels at the cell surface. [4]These mechanisms ensure that critical proteins, including those involved in neurotransmission and neuronal excitability, are correctly processed and localized. Impairments in protein modification and post-translational regulation, such as those within the N-glycan pathway, can lead to misfolded or mislocalized proteins, ultimately compromising synaptic function and contributing to abnormal movement. Metabolic regulation, exemplified by drug-metabolizing enzymes likeCYP2D6, also plays a role in the catabolism and detoxification of neuroactive substances, directly influencing drug-induced movement abnormalities. [3]

Genetic Regulatory Networks and Systemic Integration

Section titled “Genetic Regulatory Networks and Systemic Integration”

Movement is an emergent property of complex genetic regulatory networks and their systemic integration across brain regions. Genetic variants in genes such as BTBD9, TOX3/BC034767, MEIS1, MAP2K5/SKOR1, and PTPRD are associated with periodic leg movements in sleep (PLMS), illustrating how specific genetic architectures predispose individuals to distinct movement disorders. [2] These genes likely participate in various regulatory processes, from transcriptional control to synaptic development, and their collective influence shapes neuronal circuit function. The identification of RBFOX1 (A2BP1) as a regulator of tissue-specific splicing, controlling neuronal excitation in the mammalian brain, further underscores the importance of gene regulation at the post-transcriptional level. [12]

RBFOX1’s role in neuronal excitation suggests that precise control over gene expression and alternative splicing is critical for maintaining balanced neuronal activity, and its dysregulation is associated with neurological conditions including autism, mental retardation, and epilepsy.[12] This highlights hierarchical regulation where splicing factors dictate the repertoire of proteins expressed in specific neuronal types, profoundly impacting network interactions and emergent behaviors. The interplay between these genetic factors and environmental triggers, such as drug exposure, demonstrates pathway crosstalk and network interactions that lead to complex phenotypes like adverse drug reactions. [3] Understanding these integrated genetic and molecular networks is crucial for identifying the underlying causes of movement abnormalities and potential therapeutic targets.

Genetic Influences on Drug Metabolism and Transport

Section titled “Genetic Influences on Drug Metabolism and Transport”

Variants in drug-metabolizing enzymes and transporters significantly impact the pharmacokinetics of medications used to treat conditions associated with movement abnormalities, such as antipsychotics. For aripiprazole, a commonly prescribed antipsychotic, a decreased function allele of CYP2D6, specifically CYP2D6*10, has been linked to a higher incidence of adverse reactions. [3] This enzyme is crucial for the hepatic metabolism of many drugs, and reduced activity can lead to elevated plasma concentrations of aripiprazole and its active metabolites, thereby increasing the risk of movement-related side effects like akathisia.

Beyond metabolic enzymes, drug transporters also play a critical role in drug disposition and clearance. The organic anion transporter SLC22A8 (also known as OAT3), predominantly expressed in the kidney and brain, is involved in the absorption, distribution, and excretion of various drugs . A genome-wide association study identified rs4149181 in SLC22A8 as a significant locus associated with movement-related adverse drug reactions to aripiprazole. [3] Genetic variations within the SLC22 family can alter transporter activity, such as rs45566039 , which reduces transport capacity, potentially contributing to altered drug exposure. [17] It is hypothesized that a synergistic effect between risk variants in CYP2D6 and SLC22A8 could further exacerbate adverse reactions by impairing both hepatic metabolism and renal clearance of aripiprazole. [3]

Receptor and Ion Channel Polymorphisms Affecting Drug Response

Section titled “Receptor and Ion Channel Polymorphisms Affecting Drug Response”

Pharmacogenetic variants can also influence drug pharmacodynamics by affecting drug targets or related signaling pathways, impacting therapeutic response and susceptibility to adverse movement abnormalities. Several genes encoding ion channels and their interacting proteins have been associated with movement-related adverse drug reactions. For instance, rs73258503 in KCNIP4, a gene encoding a Kv channel-interacting protein, shows a significant association with these adverse effects. [3] KCNIP4is primarily expressed in the brain, where it plays roles in neurodevelopment and modulates voltage-gated potassium channels, which are critical for neuronal excitability and neurotransmission.[18] Variants in KCNIP4could alter the function of potassium channels, potentially influencing how aripiprazole, known to block Kv1.4 potassium channels, impacts neuronal activity and contributes to movement disorders.[19]

Other genetic factors influencing neuronal function and signaling pathways have also been implicated. The gene KCNA1, which encodes a voltage-gated potassium channel subunit, is associated with dyskinesia when its activity is completely lost, and variants have been identified in individuals with episodic ataxia and neurodevelopmental disorders.[20] Furthermore, ADCYAP1R1, encoding the PAC1 receptor, has variants like rs2284223 linked to movement-related adverse reactions and is known to influence startle response and fear conditioning. [21] Polymorphisms in genes such as SMAD9, involved in bone morphogenetic protein signaling,NAP1L4, which interacts with DGKζ proteins in the brain, and ERBB4, a receptor tyrosine kinase, have also been significantly associated with movement-related adverse reactions, suggesting a broader genetic landscape influencing individual responses to aripiprazole. [22]

Clinical Utility for Personalized Treatment

Section titled “Clinical Utility for Personalized Treatment”

The identification of genetic variants influencing both the pharmacokinetics and pharmacodynamics of drugs like aripiprazole offers significant potential for personalized medicine in managing movement abnormalities. A combined prediction model incorporating six significant loci (SLC22A8, ADCYAP1R1, KCNIP4, SMAD9, NAP1L4, and ERBB4) demonstrated acceptable performance in predicting serious movement-related adverse reactions. [3] This suggests that a multi-gene approach could be valuable in identifying patients at higher risk for conditions such as akathisia, abnormal involuntary movements, and extrapyramidal side effects before treatment initiation. [3]

Integrating pharmacogenetic information into clinical practice can guide drug selection and dosing strategies. For instance, for CYP2D6 variants, consensus recommendations exist from organizations like the Clinical Pharmacogenetics Implementation Consortium (CPIC) and the Dutch Pharmacogenetics Working Group (DPWG) to translate genotype into phenotype, facilitating dose adjustments or alternative drug choices. [23] By leveraging these genetic insights, clinicians may be able to proactively mitigate the risk of adverse movement events, optimize therapeutic outcomes, and improve patient safety and adherence. Furthermore, the genes identified, particularly SLC22A8, ADCYAP1R, and KCNIP4, represent important candidates for future research into molecular signaling pathways and potential specific drug targets to address challenging clinical problems like akathisia. [3]

Frequently Asked Questions About Abnormality Of Movement

Section titled “Frequently Asked Questions About Abnormality Of Movement”

These questions address the most important and specific aspects of abnormality of movement based on current genetic research.


1. My family has shaky hands; will I get them too?

Section titled “1. My family has shaky hands; will I get them too?”

Yes, many movement abnormalities, including tremors, have a strong genetic component. If your family has a history, you might be predisposed due to inherited genetic variations. However, it’s not a certainty, and other factors can also play a role in whether you develop the condition.

2. Why do some medicines make my movements weird but not my friend’s?

Section titled “2. Why do some medicines make my movements weird but not my friend’s?”

Your genetic makeup influences how your body processes medications. Specific genetic variants in genes like SLC22A8 and KCNIP4 can affect your risk of developing movement-related adverse drug reactions, such as akathisia, when taking certain medications like antipsychotics. This means you might react differently than someone else with a different genetic profile.

3. Why do my legs twitch so much when I sleep?

Section titled “3. Why do my legs twitch so much when I sleep?”

This sounds like periodic leg movement in sleep, which has significant genetic links. Polymorphisms in genes such as BTBD9 and MEIS1 are associated with an increased likelihood of experiencing these involuntary movements during sleep. Understanding these genetic factors can help in diagnosis and management.

4. Could a genetic test explain why I struggle with coordination?

Section titled “4. Could a genetic test explain why I struggle with coordination?”

Yes, genetic testing can help identify specific variations linked to various movement abnormalities and coordination issues. These tests can provide insights into the underlying biological basis of your symptoms, potentially aiding in early diagnosis and guiding more personalized treatment approaches.

5. Is there a reason my clumsy movements affect my social life?

Section titled “5. Is there a reason my clumsy movements affect my social life?”

Absolutely. Abnormalities of movement can severely impair your ability to perform daily activities, engage in social interactions, and maintain independence. These challenges can lead to a diminished quality of life, highlighting the profound social importance of understanding and treating these conditions.

6. Why do I struggle with simple visual tracking tasks more than others?

Section titled “6. Why do I struggle with simple visual tracking tasks more than others?”

Your ability to track objects with your eyes, like smooth pursuit eye movement, has high heritability and is influenced by your genes. Genetic variations in regions impacting genes like THEM4 and CDH22 can affect brain areas responsible for motor control and visual processing, making these tasks harder for you.

7. Can I prevent my kids from inheriting my movement issues?

Section titled “7. Can I prevent my kids from inheriting my movement issues?”

While a genetic predisposition means an increased risk, it doesn’t always guarantee your children will develop the same movement issues. Genetic research helps identify these risks, and ongoing studies aim to uncover novel therapeutic targets and prevention strategies, though specific preventions are still an active area of research.

8. Are my eye movements linked to my mental health struggles?

Section titled “8. Are my eye movements linked to my mental health struggles?”

Yes, specific eye movement dysfunctions, such as impaired exploratory or smooth pursuit eye movements, are recognized as robust biological markers for conditions like schizophrenia. Genetic factors influencing these movements are implicated in neurotransmission and brain morphogenesis, connecting motor control to mental health.

9. Will my movement problems definitely get worse as I get older?

Section titled “9. Will my movement problems definitely get worse as I get older?”

The progression of movement abnormalities can vary significantly among individuals, and not everyone experiences worsening symptoms with age. Your genetic makeup can influence the course and severity of these conditions, and ongoing research aims to better understand how these factors interact over time.

10. Can my daily routine or work be impacted by these subtle movements?

Section titled “10. Can my daily routine or work be impacted by these subtle movements?”

Yes, even subtle abnormalities of movement can significantly impact your ability to perform daily activities, maintain employment, and engage in social interactions. These conditions can reduce independence and diminish overall quality of life, making genetic research crucial for improving outcomes.


This FAQ was automatically generated based on current genetic research and may be updated as new information becomes available.

Disclaimer: This information is for educational purposes only and should not be used as a substitute for professional medical advice. Always consult with a healthcare provider for personalized medical guidance.

[1] Kikuchi M et al. “Genome-wide Association Analysis of Eye Movement Dysfunction in Schizophrenia.”Sci Rep, vol. 8, 2018, p. 12397.

[2] Edelson JL et al. “The Genetic Etiology of Periodic Leg Movement in Sleep.” Sleep, vol. 45, no. 6, 2022, zsac090.

[3] Wang X et al. “Cross-ancestry analyses identify new genetic loci associated with 25-hydroxyvitamin D.” PLoS Genet, vol. 19, no. 11, 2023, e1011082.

[4] Ma, Y., et al. “Association of chromosome 5q21.3 polymorphisms with the exploratory eye movement dysfunction in schizophrenia.”Sci Rep, 2015.

[5] Backman JD et al. “Exome sequencing and analysis of 454,787 UK Biobank participants.” Nature, vol. 599, 2021, pp. 82–87.

[6] Chernus JM et al. “A candidate gene analysis and GWAS for genes associated with maternal nondisjunction of chromosome 21.” PLoS Genet, vol. 15, no. 12, 2019, e1008501.

[7] Liu F. “Genetics of skin color variation in Europeans: genome-wide association studies with functional follow-up.” Hum Genet, vol. 134, 2015, pp. 823–835.

[8] Nolte IM et al. “Genetic variance estimation with imputed variants finds negligible missing heritability for human height and body mass index.”Nat Genet, vol. 47, 2015, pp. 1114–20.

[9] Young, T., et al. “The occurrence of sleep-disordered breathing among middle-aged adults.” N Engl J Med, vol. 328, no. 17, 1993, pp. 1230–1235.

[10] Blackwell, T., et al. “Associations between sleep architecture and sleep-disordered breathing and cognition in older community-dwelling men: The Osteoporotic Fractures in Men Sleep Study.” J Am Geriatr Soc, vol. 59, no. 12, 2011, pp. 2217–2225.

[11] Ferri, R., et al. “Different periodicity and time structure of leg movements during sleep in narcolepsy/cataplexy and restless legs syndrome.”Sleep, vol. 29, no. 12, 2006, pp. 1587–1594.

[12] Stambolian D et al. “Meta-analysis of genome-wide association studies in five cohorts reveals common variants in RBFOX1, a regulator of tissue-specific splicing, associated with refractive error.” Hum Mol Genet, vol. 22, no. 13, 2013, pp. 2727–2738.

[13] Sforza, E, et al. “Time-dependent variation in cerebral and autonomic activity during periodic leg movements in sleep: implications for arousal mechanisms.” Neurophysiol, 2002.

[14] Zhang, B, et al. “Sertraline and periodic limb movements during sleep: an 8-week open-label study in depressed patients with insomnia.” Sleep Med, 2013.

[15] Gregson, C. L., et al. “A rare mutation in SMAD9 associated with high bone mass identifies the SMAD-dependent BMP signaling pathway as a potential anabolic target for osteoporosis.”J. Bone Miner. Res., vol. 35, 2020, pp. 92–105.

[16] Galvan-Femenia, I., et al. “Multitrait genome association analysis identifies new susceptibility genes for human anthropometric variation in the GCAT cohort.” J Med Genet, vol. 55, no. 12, 2018, pp. 803-812.

[17] Lozano, E., et al. “Genetic heterogeneity of SLC22 family of transporters in drug disposition.” J. Pers. Med., vol. 8, 2018, p. 14.

[18] Morohashi, Y., et al. “Molecular cloning and characterization of CALP/KChIP4, a novel EF-hand protein interacting with presenilin 2 and voltage-gated potassium channel subunit Kv4.”J Biol Chem, vol. 277, no. 16, 2002, pp. 14965-14975.

[19] Park, J., Cho, K. H., Lee, H. J., Choi, J. S. & Rhie, D. J. “Open channel block of Kv1.4 potassium channels by aripiprazole.”Korean J. Physiol. Pharmacol., vol. 24, 2020, pp. 545–553.

[20] Verdura, E., et al. “Complete loss of KCNA1 activity causes neonatal epileptic encephalopathy and dyskinesia.” J. Med. Genet., vol. 57, 2020, pp. 132–137.

[21] Jovanovic, T., et al. “Impact of ADCYAP1R1 genotype on longitudinal fear conditioning in children: interaction with trauma and sex.” Neuropsychopharmacology, vol. 45, 2020, pp. 1603–1608.

[22] Takahashi, N., et al. “Cellular expression and localization of DGKζ-interacting NAP1-like proteins in the brain and functional implications under hypoxic stress.” Histochem. Cell Biol., vol. 142, 2014, pp. 461–471.

[23] Caudle, K. E., et al. “Standardizing CYP2D6 genotype to phenotype translation: consensus recommendations from the Clinical Pharmacogenetics Implementation Consortium and Dutch Pharmacogenetics Working Group.” Clin. Transl. Sci., vol. 13, 2020, pp. 116–124.