Fact-checked by Grok 2 weeks ago

Microsatellite

A microsatellite is a short tandem repeat of DNA motifs, typically consisting of 1–6 base pairs repeated 5–50 times, that occurs ubiquitously in prokaryotic and eukaryotic genomes, particularly in noncoding regions such as intergenic spaces and introns. These repetitive sequences, also known as simple sequence repeats (SSRs) or short tandem repeats (STRs), exhibit high genetic variability due to their inherent instability during DNA replication, where strand slippage can lead to expansions or contractions in repeat number. Microsatellites are distinguished by their elevated mutation rates, often ranging from 10⁻³ to 10⁻⁶ per locus per generation, which far exceed those of other genomic regions and make them polymorphic markers ideal for genetic analysis. Structurally, microsatellites can be mononucleotide (e.g., poly-A tracts like (A)₁₁), dinucleotide (e.g., (GT)₆), trinucleotide (e.g., (CTG)₄), or tetranucleotide (e.g., (ACTC)₄) repeats, with longer motifs up to six base pairs also common. They are scattered throughout the , with abundance varying by organism; for instance, the contains over 200,000 such loci in analyzed regions, predominantly in non-exonic areas. This distribution contributes to their role in genomic , as mutations in these repeats can influence , structure, and even susceptibility when expansions disrupt coding sequences. In research and applications, microsatellites serve as powerful tools for linkage analysis, , kinship determination, and due to their codominant inheritance and multiallelic nature, allowing discrimination between individuals with high resolution. They are detected primarily through (PCR) amplification followed by gel or , enabling cost-effective . Notably, —a hallmark of defective —plays a critical role in cancer diagnostics, where it indicates potential responsiveness to in tumors like colorectal . Beyond medicine, these markers facilitate (QTL) mapping, evolutionary studies, and assessment across species, underscoring their versatility in modern .

Definition and Characteristics

Basic Definition

Microsatellites, also known as short tandem repeats (STRs) or simple sequence repeats (SSRs), are tandemly repeated DNA sequences consisting of short motifs of 1–6 base pairs that are typically repeated 5–50 times (with minimum thresholds varying by motif, e.g., ≥10 for mononucleotides and ≥5 for longer motifs), resulting in alleles ranging from approximately 10 to 300 base pairs in length. These repeats are ubiquitous in eukaryotic genomes and are classified based on the length of their core motif, distinguishing them from other repetitive elements. Microsatellites differ from minisatellites, which feature longer repeat units of 10–100 base pairs, and from single nucleotide polymorphisms (SNPs), which involve single base substitutions without repetitive structure. The term "microsatellite" reflects their relatively short size and overall tract length compared to these longer repeats. Common types include mononucleotides, such as polyadenine (A)_n exemplified by AAAAA; dinucleotides, such as (CA)_n shown as CACACA; and trinucleotides, such as ()_n represented by CAGCAGCAG. Due to their high mutation rates—often ranging from 10^{-3} to 10^{-6} per locus per generation, orders of magnitude higher than typical point s—microsatellites display hypervariability, making them polymorphic markers useful in genetic studies. This variability arises primarily from changes in repeat number but is not detailed in mechanisms here.

Structural Features

Microsatellites, also known as short tandem repeats (STRs), are composed of tandemly arrayed DNA motifs consisting of 1 to 6 base pairs (bp) in length, flanked on both sides by unique, non-repetitive sequences. These core repeat units are repeated consecutively multiple times, forming the polymorphic region of the locus, while the overall allele length, encompassing the repeat tract and flanking regions, typically spans 100 to 400 bp in standard genotyping applications. This structure allows for precise amplification and analysis of the variable repeat array using polymerase chain reaction (PCR) techniques. The variability of microsatellites primarily arises from differences in the number of repeat units, which define distinct alleles within a . For instance, alleles may differ by having 10 versus 15 repeats of a dinucleotide motif such as CA, leading to length polymorphisms that are detectable by or capillary sequencing. Microsatellites are classified by the length of their repeat motif into mononucleotide (1 ), dinucleotide (2 ), trinucleotide (3 ), tetranucleotide (4 ), pentanucleotide (5 ), and hexanucleotide (6 ) types, with dinucleotide repeats being the most prevalent in eukaryotic genomes due to their high abundance and mutability. The flanking regions surrounding the repeat tract consist of conserved, non-repetitive DNA sequences that are essential for the design of locus-specific PCR primers, ensuring targeted amplification without interference from similar repeats elsewhere in the genome. Additionally, microsatellite tracts may occasionally contain rare interruptions—single or few non-repeat bases inserted within the array—which disrupt the perfect tandem structure and contribute to sequence stability by impeding replication slippage mechanisms.

Genomic Locations and Prevalence

Microsatellites are predominantly distributed across non-coding regions of the , with approximately 60-70% located in intergenic spaces, 20-30% within introns, and only 5-10% in exons. This uneven distribution reflects their tendency to accumulate in areas less constrained by protein-coding requirements, while their density is notably higher in than in , facilitating accessibility for replication and transcription processes. In the human genome, microsatellites comprise about 3% of the total DNA sequence and number approximately 1–2 million loci (as of 2023), depending on the minimum repeat threshold used for identification. Their prevalence varies across organisms, with plants generally exhibiting higher overall abundance due to larger genome sizes, though density per megabase is often lower compared to animals. For instance, land plants and mammals show similar proportions of genome coverage by microsatellites (around 11%), but plant genomes tend to harbor more total instances owing to polyploidy and expansion events. Trinucleotide repeats are particularly enriched in coding regions across eukaryotes, as their length (divisible by three) minimizes the risk of frameshift mutations that could disrupt protein translation. Organism-specific patterns further highlight this variability: dinucleotide repeats, such as (GT/CA)_n, are more prevalent in mammals, where they constitute a significant portion of polymorphic loci used in genetic studies. In contrast, show a bias toward mononucleotide poly-A/T tracts, which are overrepresented and contribute to phase variation and adaptive . Microsatellites also cluster in heterochromatic regions like centromeres and telomeres, where they form structural components such as telomeric TTAGGG repeats, yet they are functionally relevant in euchromatic promoters, influencing through length polymorphisms. Genome-wide identification of microsatellites relies on bioinformatic tools, such as Tandem Repeats Finder (TRF), which scans DNA sequences for tandemly repeated motifs of 1-2000 bases, outputting details on location, copy number, and consensus patterns without requiring user-specified parameters. This tool has been instrumental in mapping microsatellites across diverse genomes, enabling precise annotation of their positional prevalence.

History and Discovery

Early Identification

The origins of microsatellite identification trace back to the mid-1980s, when studies on human DNA variability first highlighted sequences. In 1985, and colleagues described hypervariable regions composed of with motif lengths of 10-60 base pairs, terming them "minisatellites" or variable number (VNTRs), which were dispersed throughout the and exhibited high polymorphism useful for individual identification. These findings laid the groundwork for recognizing shorter repetitive elements, though the specific class of microsatellites—defined by motifs of 1-6 base pairs—emerged later in the decade. The term "microsatellite" was introduced in 1989 by Mark Litt and Joseph A. Luty, who identified a highly polymorphic dinucleotide repeat (TG)_n within the actin using (PCR) amplification, revealing 12 alleles among 37 unrelated individuals. Concurrently, J.L. Weber and P.E. May reported an abundant class of (CA)_n/(GT)_n dinucleotide repeats that could be efficiently genotyped via , emphasizing their potential as polymorphic markers across the . These works shifted terminology from earlier descriptors like "simple repetitive DNA" to "microsatellites," distinguishing them from longer minisatellites. Early identification often occurred in the context of disease-linked hypervariable regions, such as those studied in , where variable simple sequence motifs near the DM1 locus on were probed as genetic markers in 1989. Initial challenges in microsatellite recognition stemmed from overlap with minisatellites, leading to confusion in classifying repeat lengths and variability; this was resolved through direct sequencing, which confirmed the short structure and high of microsatellites. Early observations also noted elevated rates in these repeats, attributed to replication slippage, setting the stage for their use in genetic mapping.

Key Milestones in Research

In the late and early , the development of (PCR) amplification techniques revolutionized microsatellite analysis, enabling the reliable detection and genotyping of these repetitive sequences from small DNA samples. This breakthrough was pioneered by Litt and Luty in 1989, who first described a hypervariable dinucleotide microsatellite within the cardiac gene and demonstrated its amplification via PCR. By the early , these methods facilitated the widespread use of microsatellites as polymorphic markers for genetic mapping, particularly in the (HGP) from 1990 to 2003. Microsatellites served as key second-generation markers, with comprehensive genetic maps constructed using over 8,000 such loci to achieve high-resolution linkage analysis across the . The 1993 identification of expanded trinucleotide repeats as the genetic basis for marked a pivotal advancement in understanding microsatellite instability's role in hereditary disorders. Researchers from the Collaborative Research Group isolated the huntingtin gene (HTT) and revealed that pathological expansions of repeats (beyond 36 units) cause the disease through a toxic gain-of-function mechanism. In the 2000s, microsatellites began integrating with emerging (SNP) technologies, appearing in combined genotyping panels for enhanced and linkage studies, though SNPs increasingly supplemented them due to higher throughput. The launch of the in 2008 represented a global effort to catalog , including microsatellites, by sequencing over 1,000 individuals from diverse populations. This initiative identified nearly 700,000 short tandem repeat () loci, providing a comprehensive reference for germline microsatellite polymorphisms and revealing patterns of repeat length variation across ancestries. During the , next-generation sequencing (NGS) technologies expanded microsatellite research into microbial ecosystems, uncovering abundant repeats in gut microbiomes that influence bacterial evolution and host interactions. Concurrently, microsatellites gained prominence in genetics, enabling fine-scale population structure analysis in , such as monitoring in fish stocks through multiplex panels. In the 2020s, emerged as a transformative tool for modeling microsatellite-related diseases, allowing precise contraction or interruption of expanded repeats in cellular and animal models of disorders like Huntington's. For instance, studies have used dual-guide designs to excise repeats in HTT, reducing toxicity in neuronal cultures and mouse brains. Additionally, () models have advanced the prediction of () in cancer genomes, with 2025 approaches analyzing whole-slide images or genomic data to forecast MSI-high status in colorectal and lung tumors, aiding selection.

Functions and Biological Roles

Role in Gene Regulation

Microsatellites within promoter and enhancer regions play a key role in modulating by altering the binding affinity or number of sites for transcription factors through variations in repeat length. These repeats can serve as flexible spacers or direct binding motifs, where expansions or contractions influence the spacing between regulatory elements or the strength of protein-DNA interactions. For instance, repeat length variations in promoters have been shown to affect transcriptional activity in s involved in stress response pathways. In the oxygenase-1 (HO-1) , polymorphic (GT)n repeats in the promoter inversely correlate with basal and induced expression levels, where longer repeats reduce promoter activity compared to shorter alleles. Epigenetic regulation is another mechanism by which microsatellites influence , particularly through repeat expansions that recruit to alter structure. Expanded repeats can form abnormal DNA or structures that attract complexes including histone deacetylases (HDACs) and methyltransferases, resulting in formation and transcriptional silencing, or in some cases, activation via enhancer-like effects. In , the expanded CGG microsatellite in the 5' UTR of the gene recruits HDACs and DNA methyltransferases, leading to hypermethylation of the promoter and near-complete silencing of FMR1 expression. Similarly, in imprinting disorders such as fragile X-associated / (FXTAS), these expansions contribute to RNA-mediated recruitment of modifiers, disrupting normal epigenetic marks and affecting expression of nearby imprinted genes. Specific examples highlight the regulatory impact of microsatellites on gene expression. The length of CAG repeats in exon 1 of the androgen receptor (AR) gene modulates AR transcriptional activity, with shorter repeats (e.g., <20) associated with higher AR protein levels and enhanced transactivation of target genes, while longer repeats reduce this activity due to altered protein stability and recruitment efficiency. Microsatellites in untranslated regions (UTRs) further contribute by influencing post-transcriptional regulation, particularly through interactions with microRNAs (miRNAs). Polymorphic short tandem repeats in 3' UTRs can disrupt or enhance miRNA binding sites, altering mRNA stability and translation. Overall, the repeat number in these microsatellites often correlates with expression variability, underscoring their role as tunable regulatory elements.

Evolutionary Significance

Microsatellites play a pivotal role in neutral evolution due to their exceptionally high mutation rates, ranging from 10^{-2} to 10^{-6} per locus per generation, which far exceed those of point mutations in coding regions and generate substantial allelic diversity subject primarily to genetic drift rather than selection. This hypervariability positions microsatellites as ideal neutral markers for tracing evolutionary processes, as their patterns reflect the balance between mutational input and stochastic loss through drift in populations. In isolated or small populations, this dynamic fosters rapid divergence without adaptive pressures, contributing to overall genomic variability that can influence long-term evolutionary trajectories. Present in both prokaryotic and eukaryotic genomes, microsatellites trace their ancient origins to early cellular life, with evidence of conservation spanning over 450 million years across diverse taxa, suggesting an enduring role in genome architecture. Their prevalence expanded notably in eukaryotes, where they enhance genome plasticity by facilitating structural rearrangements and insertions that promote evolutionary flexibility. This expansion likely supported the complexity of eukaryotic genomes, allowing microsatellites to act as mutable elements that buffer against or enable responses to environmental shifts over evolutionary timescales. While largely neutral, certain microsatellite variations exhibit adaptive potential by influencing phenotypic traits, such as differences in flowering time in plants through expansions or contractions in promoter regions that modulate gene expression timing. For instance, repeat length polymorphisms in regulatory sequences have been associated with adaptive shifts in reproductive phenology, enabling populations to align flowering with local climates and enhancing fitness in heterogeneous environments. In hybrid zones, microsatellite divergence driven by can accelerate speciation by creating barriers to gene flow. Conservation under selection is evident in specific contexts, particularly trinucleotide repeats within exons, where their length and motif are constrained to maintain open reading frames and avoid frameshift mutations that could disrupt protein coding. This selective pressure favors in-frame repeats, such as CAG tracts aligned to preserve translational fidelity, thereby stabilizing essential gene functions across evolutionary lineages despite the inherent mutability of microsatellites. Such mechanisms underscore how selection can counteract instability to retain functional repeats in critical genomic regions.

Mutation Processes

Mechanisms of Instability

Microsatellites exhibit instability primarily through slipped-strand mispairing during DNA replication, where the DNA polymerase temporarily dissociates from the template strand within the repetitive sequence, allowing realignment that results in insertions or deletions of repeat units (indels). This process, known as , occurs because the repetitive nature of microsatellites facilitates non-B DNA conformations that stall the replication fork, leading to polymerase stuttering and the incorporation of extra or fewer nucleotides in the nascent strand. Defects in DNA mismatch repair (MMR) exacerbate microsatellite instability by failing to correct these replication errors, particularly in conditions like , where germline mutations in MMR genes such as MLH1 or MSH2 impair the recognition and excision of mismatched loops formed during slippage. In proficient cells, MMR proteins detect and resolve these quasi-stable mispairs, but in deficient systems, uncorrected indels accumulate, promoting expansions especially in coding microsatellites. Microsatellite mutations typically occur as single-step changes involving the gain or loss of 1-2 repeat units, though multi-step alterations involving larger shifts can arise in highly unstable contexts, such as MMR-deficient tumors. Contractions predominate in longer repeat tracts, while expansions are more frequent in shorter ones, reflecting allele length-dependent biases in slippage resolution. Key factors influencing instability include repeat purity, where uninterrupted tracts are far more prone to slippage than those containing base interruptions that disrupt misalignment; for instance, even a single nucleotide variant can reduce mutation rates by stabilizing the duplex. Additionally, trinucleotide repeats often form stable hairpin secondary structures during replication, which impede polymerase progression and favor expansions, as seen in disease-associated loci like CAG repeats. In the slippage model, the probability of a mutation event is proportional to the repeat tract length n, as longer tracts increase opportunities for misalignment: P(\text{error}) \propto n This relationship underscores the exponential rise in instability with increasing repeat number, without requiring detailed derivation here.

Rates and Factors Influencing Mutations

Microsatellite mutation rates in humans typically range from $10^{-3} to $10^{-4} per locus per generation, though estimates vary across loci and studies due to differences in repeat structure and assay methods. Mononucleotide repeats exhibit higher mutation rates than trinucleotide repeats, with mononucleotide instability often exceeding dinucleotide rates by factors of 2-10 in both germline and somatic contexts. Several factors influence these mutation rates, including the length of the repeat tract, where longer microsatellites mutate more frequently than shorter ones, often showing a positive correlation with allele size. Replication timing also plays a role, with hotspots during S-phase associated with elevated instability due to increased polymerase slippage opportunities. Defects in mismatch repair (MMR) genes, such as MSH2 mutations, dramatically increase rates by 100-fold or more, as MMR normally corrects slippage errors during replication. Mutation rates are commonly measured through pedigree studies, which track intergenerational changes; data from 1990s analyses reported frequencies of 0.001 to 0.02 mutations per meiosis across various loci. In model organisms like yeast, rates are generally faster than in humans, often reaching $10^{-2} to $10^{-4} per locus per generation, reflecting differences in replication fidelity and repair efficiency. Microsatellite mutations largely follow a stepwise model, in which approximately 80% of events involve single-repeat unit gains or losses, though larger changes occur occasionally. Recent whole-genome sequencing studies from the 2020s have refined these estimates, revealing average germline rates around $5 \times 10^{-5} per microsatellite per generation while highlighting environmental influences like oxidative stress, which can accelerate instability by promoting replication errors. These findings underscore how external factors interact with intrinsic sequence properties to modulate mutation dynamics.

Biological Consequences

Impacts on Protein Function

Microsatellites located within protein-coding regions, particularly trinucleotide repeats, can significantly alter protein sequences by encoding expanded poly-amino acid tracts. For instance, CAG trinucleotide repeats in the huntingtin gene (HTT) translate into polyglutamine tracts; expansions exceeding 35 repeats are pathogenic and promote protein aggregation, leading to loss of normal function and gain of toxic properties in Huntington's disease. Non-triplet microsatellites, such as dinucleotide repeats, rarely occur in coding regions due to strong purifying selection against frameshift mutations that disrupt the reading frame. When such contractions or expansions do arise, they can introduce premature stop codons or produce aberrant proteins with toxic effects, though these are infrequent compared to triplet repeat disorders. In spinocerebellar ataxias (SCAs), CAG expansions in genes like ATXN1 (SCA1), ATXN2 (SCA2), and ATXN3 (SCA3) generate elongated polyglutamine tracts that confer length-dependent instability, with longer repeats (>35-40) increasing protein insolubility, misfolding, and aggregation, thereby disrupting neuronal and causing cerebellar degeneration. These impacts exhibit threshold effects, where repeat lengths of 10-30 are typically normal and polymorphic without phenotypic consequences, but expansions beyond 40 often trigger pathogenicity. Inheritance of these expansions can show , with intergenerational increases in repeat length leading to earlier disease onset and greater severity, particularly in paternal transmissions for polyglutamine disorders.

Effects in Non-Coding Regions

Microsatellite expansions in non-coding regions can profoundly disrupt gene expression and genome stability without altering protein sequences directly. These regions, including introns, untranslated regions (UTRs), and intergenic areas, harbor variable numbers of tandem repeats that, when expanded, often lead to RNA toxicity, altered regulatory processes, or structural instability. Such changes contribute to various pathologies by interfering with splicing, translation, mRNA stability, and chromosomal integrity. In intronic regions, microsatellite expansions frequently impair pre-mRNA splicing by sequestering key splicing factors. For instance, in type 2 (DM2), an expanded CCTG repeat in the first of the CNBP produces a toxic that binds and depletes muscleblind-like splicing regulator 1 (MBNL1), resulting in widespread missplicing of exons across multiple genes, which manifests as and other systemic symptoms. This gain-of-function mechanism exemplifies how intronic repeats can deregulate pathways essential for tissue-specific . Interactions between microsatellites and transposable elements, particularly Alu sequences, can enhance genomic instability through increased recombination or mobility. Alu elements, which are short interspersed nuclear elements comprising about 11% of the , often contain or are adjacent to microsatellite repeats; these associations promote unequal recombination events during or , leading to insertions, deletions, or copy number variations that disrupt nearby non-coding regulatory elements. Studies have shown that the presence of Alu insertions correlates with elevated local recombination rates within 2 kb, facilitating the genesis and propagation of microsatellite alleles in primate genomes. Microsatellite variations in UTRs exert post-transcriptional control over . In the 5' UTR, expansions such as CGG repeats in the gene inhibit translation initiation by forming stable secondary structures that impede ribosomal scanning and cap-dependent initiation, reducing FMRP protein levels and contributing to cognitive impairments. Similarly, in the 3' UTR, expanded CTG repeats in the DMPK gene, as seen in type 1 (), promote nuclear retention of the mRNA and enhance degradation via mechanisms involving -binding proteins, thereby destabilizing transcripts and amplifying splicing defects through RNA foci formation. These UTR effects highlight the role of repeats in fine-tuning mRNA translation efficiency and half-life without coding sequence changes. Non-coding microsatellite expansions also drive genome-wide instability, particularly at fragile sites prone to breakage. The FRAXA locus on the , associated with , features CGG repeat expansions in the 5' UTR of that induce chromosomal fragility under stress, leading to gaps or breaks visible in spreads and increased recombination or deletion events nearby. This instability arises from the formation of non-B DNA structures like hairpins during replication, which stall forks and recruit repair machinery, potentially propagating mutations across the genome. Somatic expansions of non-coding microsatellites exhibit tissue-specific patterns that accumulate with aging and contribute to oncogenesis. In normal tissues, microsatellite instability rises progressively with age, with higher rates observed in brain and colon cells, where expanded repeats in intergenic or intronic regions foster localized genomic rearrangements. In cancer, somatic expansions of tandem repeats, including those in non-coding areas, occur recurrently and drive clonal evolution; for example, in colorectal tumors, such expansions correlate with mismatch repair deficiencies, promoting tumor heterogeneity and progression in a tissue-dependent manner. These dynamic changes underscore the role of environmental and replicative stresses in exacerbating non-coding repeat instability over time.

Applications

Forensic Identification and Kinship Testing

Microsatellites, also known as short tandem repeats (STRs), serve as the cornerstone of forensic DNA profiling through systems like the Combined DNA Index System (CODIS), which utilizes 20 core autosomal STR loci, including D3S1358, to generate unique genetic fingerprints for individual identification. These loci are selected for their high polymorphism and low mutation rates, enabling the creation of DNA profiles that exhibit an extraordinarily low random match probability, approximately 1 in 10^18 for unrelated individuals in the general population. This discriminatory power allows forensic laboratories to link biological evidence from crime scenes to suspects or databases with high confidence, facilitating the resolution of criminal investigations. In paternity testing, enables exclusion of a putative father if there is an mismatch at one or more loci, as the must inherit one from each parent. For inclusion, likelihood ratios quantify the probability of the observed genotypes assuming paternity versus non-paternity, with the paternity index (PI) calculated per locus; for instance, when the and alleged father share a single , the PI is often 0.5 divided by the frequency of that allele in the population. Combined across multiple loci, these indices yield a combined paternity index that supports probabilistic statements of relationship, typically exceeding thresholds for legal or personal confirmation. Beyond direct parentage, STR-based testing extends to grandparentage and relationships by analyzing patterns across multiple loci to compute likelihood ratios for complex pedigrees. In grandparentage tests, the absence of a direct parent requires evaluating the transmission of alleles through intermediate generations, often achieving reliable results with 15-20 loci when combined with maternal data. tests similarly rely on shared alleles at multiple loci to distinguish full from half-siblings, with higher numbers of loci improving resolution for ambiguous cases. These methods are routinely applied in forensic contexts, such as evidence linking perpetrators to victims or disaster victim identification, where reference samples from relatives aid in matching fragmented remains. To address degraded DNA from environmental exposure or time, mini-STRs—variants with shorter amplicon sizes targeting the same core loci—enhance recovery by reducing inhibition and dropout. This approach has proven effective in analyzing from crime scenes or skeletal remains in mass disasters, yielding partial profiles sufficient for kinship matching when full profiles fail. Despite their utility, STR profiling has limitations, including the inability to distinguish identical monozygotic twins, who share identical genotypes at all loci, necessitating alternative markers like SNPs for differentiation. Additionally, population substructure can introduce biases in match probability estimates if allele frequencies are not adjusted for ethnic subgroups, potentially inflating or deflating likelihood ratios in kinship assessments.

Population Genetics and Biodiversity

Microsatellites serve as powerful genetic markers in owing to their high levels of polymorphism, which enable the detection of subtle differences in frequencies across populations. This polymorphism arises from variations in repeat number, allowing researchers to quantify and events. For instance, calculations of F_ST, a measure of genetic , rely on microsatellite frequencies to identify recent in structured populations, such as in studies of human continental groups where only 5–10% of variation occurs between major regions. In conservation biology, microsatellites are instrumental for detecting population bottlenecks, characterized by reduced heterozygosity due to historical demographic contractions. A classic example is the cheetah (Acinonyx jubatus), where microsatellite analyses have revealed persistently low genetic diversity stemming from bottlenecks approximately 10,000–12,000 years ago, leading to elevated inbreeding and reduced adaptability. Such markers help prioritize conservation efforts by highlighting populations at risk of further erosion in genetic variation. Microsatellites also facilitate phylogeographic studies by tracing migration patterns through repeat length variations that accumulate over generations. In human populations, Y-chromosome microsatellite data support the out-of-Africa model, showing higher diversity in African groups and a serial in non-African lineages, consistent with migrations beginning around 50,000–70,000 years ago. Similarly, in biodiversity assessments, simple sequence repeats (SSRs, synonymous with microsatellites) are used to monitor spread; for example, they reconstruct invasion routes and source populations in plants and animals, aiding management strategies to mitigate ecological impacts. Key statistical tools like analysis of molecular variance (AMOVA) leverage microsatellite data to partition genetic variance into components attributable to within-population, between-population, and among-group differences, providing a hierarchical view of structure. Typically, 10–20 microsatellite loci are sufficient for robust population-level analyses, as fewer highly polymorphic markers can resolve major structures while minimizing costs.

Medical Diagnostics and Breeding

Microsatellites play a crucial role in medical diagnostics, particularly through the assessment of (MSI), a hallmark of certain hereditary and sporadic cancers. In , MSI testing is routinely used to screen for Lynch , an inherited condition caused by mutations in mismatch repair genes. The revised Bethesda guidelines recommend evaluating tumors from patients under 50 years or with specific histopathological features using a panel of five microsatellite loci, including mononucleotide repeats BAT-25 and BAT-26, and dinucleotide repeats D5S346, D2S123, and D17S250; instability in two or more loci indicates MSI-high (MSI-H) status, prompting further and testing for Lynch syndrome mutations. High MSI status also serves as a predictive for response to , as MSI-H tumors exhibit a high mutational burden that enhances tumor immunogenicity and susceptibility to inhibitors like . In the diagnosis of repeat expansion disorders, (PCR) amplification and sizing of microsatellite repeats enable precise for conditions like , where expansions of the CAG trinucleotide repeat in the HTT gene beyond 36 repeats confer full of the neurodegenerative . This PCR-based method, often employing fluorescent primers and , confirms diagnosis in symptomatic individuals and supports presymptomatic testing in at-risk adults, with alleles of 36-39 repeats showing reduced . Prenatal screening via PCR on chorionic villus samples or amniocytes identifies expanded alleles early, allowing informed reproductive decisions; noninvasive approaches using from maternal plasma have also demonstrated feasibility for detecting paternal CAG expansions. Microsatellites, particularly simple sequence repeat (SSR) markers, are integral to in and through (QTL) mapping and (MAS). In crop improvement, SSR markers have facilitated the identification of QTLs for resistance in ; for instance, a major QTL (qDTY1.1) on , flanked by SSR markers RM431 and RM11943, explains up to 17% of phenotypic variance in grain yield under reproductive-stage stress, enabling the of tolerance alleles into elite varieties. In livestock, microsatellite markers support MAS by linking genetic variants to traits like milk yield or disease resistance; panels of 30-50 bovine microsatellites have been used to construct linkage maps for QTL detection, accelerating selection for economically important traits while preserving . In , microsatellite variants influence and efficacy, with the variable number (VNTR) in the (TYMS) gene promoter serving as a key example. The TYMS VNTR, consisting of 2- or 3-repeat alleles, modulates TYMS expression levels, where the 3-repeat variant is associated with higher enzyme activity and poorer response to 5-fluorouracil-based in colorectal and cancers, guiding personalized dosing to optimize therapeutic outcomes and minimize . Advances in the 2020s have expanded microsatellite applications through liquid biopsies, which analyze for somatic in advanced cancers without invasive tissue sampling. Techniques like targeted next-generation sequencing of monomorphic microsatellite panels in detect MSI-H with over 90% concordance to tissue-based assays, enabling real-time monitoring of tumor evolution and response in colorectal and pancreatic cancers; this noninvasive approach has improved accessibility for patients with metastatic disease.

Analytical Methods

PCR-Based Detection

Polymerase chain reaction () is the primary method for amplifying microsatellite loci, enabling the detection of length variations in tandem repeats. In standard PCR protocols for forensic applications, such as those targeting the 20 (CODIS) core loci, fluorescently labeled primers are used to tag amplicons for subsequent analysis. These primers incorporate dyes like FAM, , , and , allowing multiplex detection of alleles differing by as little as one repeat unit. The thermal cycling conditions typically involve an initial denaturation at 94–95°C for 2–5 minutes to separate DNA strands, followed by 25–35 cycles of denaturation at 94–95°C for 30–60 seconds, annealing at 55–60°C for 30–60 seconds to allow primer binding, and extension at 72°C for 30–60 seconds to synthesize new strands using a thermostable DNA polymerase like Taq. A final extension at 72°C for 5–10 minutes ensures complete product formation. These parameters balance specificity and yield, minimizing non-specific amplification while accommodating the short amplicon sizes (100–400 base pairs) common in microsatellites. Multiplexing enhances efficiency by co-amplifying 10 or more loci in a single reaction, reducing sample consumption and processing time in applications like forensic and . Commercial kits, such as GlobalFiler Express, enable simultaneous amplification of the 20 CODIS core loci plus additional markers using carefully balanced primer concentrations and buffer components to avoid competition and ensure uniform amplification across loci. This approach has become standard, supporting high-throughput of thousands of samples annually in forensic databases. Following amplification, amplicons are sized using , where fluorescently labeled products are separated by size in a polymer-filled capillary under an . Detection occurs via , producing electropherograms with peaks corresponding to allele lengths; allele calling is performed by comparing peak positions to a size standard like GeneScan 600 LIZ, with software identifying stutter peaks (typically 1–4 bases shorter than the true allele due to polymerase slippage) for accurate . This method offers high resolution (better than 0.5 base pairs) and automation, essential for distinguishing homozygotes from heterozygotes. Optimization of conditions is crucial to reduce artifacts like stutter, which can complicate interpretation. Adjusting Mg²⁺ concentration to 1.5–2.5 mM stabilizes the polymerase-DNA interaction while minimizing slippage; higher levels (>3 mM) increase stutter by enhancing non-specific priming, whereas lower levels reduce . Other tweaks, such as touchdown annealing (starting 5–10°C above the primer Tm and decreasing gradually), further improve specificity without altering cycle times significantly. A variant, real-time PCR, quantifies () by monitoring in using fluorescent probes or intercalating dyes, often coupled with to detect shifts in product length indicative of insertions or deletions. This approach is particularly useful in cancer diagnostics, where MSI-high tumors show altered at mononucleotide loci like BAT-26, enabling rapid screening without post-PCR separation.

Primer Design and Optimization

Effective primer design is crucial for the specific amplification of microsatellite loci, as primers must anneal to unique flanking sequences to avoid non-specific products and ensure reliable genotyping. Primers are typically 18-25 base pairs (bp) in length, selected from non-repetitive regions immediately adjacent to the microsatellite repeat to flank the variable region precisely. This positioning allows for amplicon sizes of 100-500 bp, which balances specificity with efficient PCR amplification. The GC content of primers should be maintained between 40% and 60% to promote stable hybridization without excessive secondary structure, as deviations can lead to poor annealing or dimer formation. Computational tools such as Primer3 are widely used for designing microsatellite primers, incorporating parameters like melting temperature (Tm) calculations to ensure optimal annealing. Primer3 defaults recommend a Tm of 57-63°C (optimum 60°C), but for microsatellite , annealing temperatures are often set to 50-60°C to accommodate variable flanking sequences and reduce non-specific binding. The software also facilitates avoidance of repetitive motifs within primers by limiting mononucleotide runs (e.g., no more than 5 identical bases) and screening against repeat libraries to prevent mispriming. These features help generate locus-specific primer pairs that amplify the target microsatellite without cross-reactivity. Optimization of primer performance involves empirical adjustments to PCR conditions, particularly annealing temperature, which can be determined using gradient to test a range (e.g., 50-65°C) in a single run for the highest specificity and yield. For primers flanking GC-rich regions, additives like 5-10% (DMSO) are incorporated to lower the Tm and disrupt secondary structures, improving amplification efficiency without altering primer sequences. These strategies ensure robust product formation across diverse templates, though they must be validated per locus to account for sequence variability. Key challenges in microsatellite primer design include heteroduplex formation during of heterozygous samples, especially in longer amplicons (>300 ), where partially annealed products create artifacts that obscure peaks in . This issue arises from re-annealing of strands with differing repeat lengths post-denaturation, complicating interpretation and requiring shorter extension times or touchdown to mitigate. Additionally, for degraded DNA samples common in forensics or , primers are redesigned as mini-STRs by shifting them closer to the repeat , reducing amplicon sizes to <150 to enhance recovery of partial profiles from fragmented templates. Best practices emphasize post-design validation using reference samples with known alleles to confirm primer specificity, sizing accuracy, and absence of stutter peaks that could mimic variants. Controls for alleles—non-amplifying variants due to in flanking regions—are essential; these include testing multiple individuals per and redesigning primers if amplification failure exceeds 5-10% in heterozygotes. Such validation ensures reliable locus utility across applications, minimizing errors.

Limitations and Emerging Techniques

One major limitation in microsatellite analysis is the occurrence of stutter artifacts, which arise from slippage during amplification and can mimic true mutations, leading to errors, particularly in heterozygous samples. Null alleles, resulting from primer mismatches with variant flanking sequences, further complicate analysis by causing apparent homozygotes or dropout, particularly in diverse populations where such variants are common. Homozygote dropout, often linked to preferential amplification of shorter alleles, exacerbates these issues, biasing heterozygosity estimates and inflating coefficients in population studies. Ascertainment bias also poses a significant challenge, as loci are typically selected based on high polymorphism in the source , leading to overestimation of when applied cross-species and underrepresentation of rarer alleles. Additionally, the high cost of developing and typing microsatellites across whole genomes—often exceeding that of arrays due to labor-intensive primer design and validation—limits their scalability for large-scale or routine applications. Emerging techniques address these limitations through next-generation sequencing (NGS), enabling high-throughput of over 100 loci simultaneously via Illumina-based panels that reduce stutter through improved read depths and error correction algorithms. As of 2025, EMQN guidelines recommend combined fragment and sequencing methods for (MSI) assessment, with NGS facilitating comprehensive genomic profiling in cancer diagnostics. offers particular advantages for long repeats, providing long-read data that accurately resolves expansions beyond 100 repeats, which traditional methods often fail to characterize due to slippage. As alternatives, inter-simple sequence repeat (ISSR-) generates anonymous multilocus markers from microsatellite-flanking regions without prior sequencing, offering a cost-effective option for screening in non-model organisms. CRISPR-based editing has emerged to study directly, with targeted knockouts in organoids revealing mutation rates and repair mechanisms in controlled models. Looking ahead, integration of microsatellites with in panels—developed in the —combines the high of repeats for kinship analysis with SNP stability, as seen in monitoring arrays that enhance hybrid detection accuracy.

References

  1. [1]
    What Is a Microsatellite - NIH
    Microsatellites are tandem repeats of short (1–6 bp) DNA motifs and are ubiquitous in eukaryotic genomes. Germline microsatellite mutation rates are high in ...
  2. [2]
    Definition of microsatellite - NCI Dictionary of Genetics Terms
    Repetitive segments of DNA scattered throughout the genome in noncoding regions between genes or within genes (introns).
  3. [3]
    Microsatellite - an overview | ScienceDirect Topics
    A microsatellite is a stretch of DNA with mono-, di-, tri-, or tetra-nucleotide units repeated. Microsatellites are short sequences of nucleotides (typically 1 ...
  4. [4]
    Functional Mechanisms of Microsatellite DNA in Eukaryotic Genomes
    Microsatellites, or short tandem repeats (STRs), also often called short sequence repeats (SSRs), consist of tandem duplications of 1–6 bp motifs. They are ...
  5. [5]
    Microsatellite DNA - an overview | ScienceDirect Topics
    Simple sequence repeats or short tandem repeats, also known as microsatellites, are repeating sequences of 2–6 base pairs of DNA. STRs are typically codominant.
  6. [6]
    Mini- and microsatellite expansions: the recombination connection
    Microsatellites are tandem arrays of short (usually <10 bp) units, while minisatellites are tandem arrays of longer units (>10 and <100 bp).
  7. [7]
    Minisatellite - an overview | ScienceDirect Topics
    Genomes and evolution​​ Minisatellites are similar to microsatellites, but with a longer repeat unit, typically 10–150 bp in length.
  8. [8]
    Microsatellite DNA - an overview | ScienceDirect Topics
    Microsatellites are short stretches of genomic DNA that contain multiple tandem copies of a mononucleotide, dinucleotide, or trinucleotide repeat motif. Due ...
  9. [9]
    Microsatellite evolution: Mutations, sequence variation, and ...
    In general, microsatellites have a high mutation rate (10-2–10-6) as compared to point mutations in coding genes [4]. It is accepted that the most common ...
  10. [10]
    Microsatellite mutations in the germline - ScienceDirect.com
    Microsatellite DNA sequences mutate at rates several orders of magnitude higher than that of the bulk of DNA. Such high rates mean that spontaneous ...
  11. [11]
    A Comprehensive Survey of Human Y-Chromosomal Microsatellites
    ... lengths (∼100–400 bp). DNA Samples. DNA samples of three male and two female human individuals were used for locus evaluation. DNA samples from eight males ...
  12. [12]
    Microsatellite markers: what they mean and why they are so useful
    Aug 4, 2016 · Microsatellites or Single Sequence Repeats (SSRs) are extensively employed in plant genetics studies, using both low and high throughput genotyping approaches.
  13. [13]
    Microsatellites in Different Eukaryotic Genomes: Survey and Analysis
    In all vertebrates and arthropods, AC is the most frequent dinucleotide repeat motif (Tables 2–4). C. elegans prefers AG in intergenic regions, AT in ...
  14. [14]
    Genome-wide analysis of microsatellite polymorphism in chicken ...
    Interruptions within perfect repeat arrays reduce the likelihood for a microsatellite locus being polymorphic.
  15. [15]
    Comparison of the Microsatellite Distribution Patterns in ... - Frontiers
    Feb 25, 2021 · Here, we conducted a genome-wide characterization of microsatellite distribution patterns at different taxonomic levels in 153 Euarchontoglires genomes.
  16. [16]
    Patterns of microsatellite distribution across eukaryotic genomes
    Feb 22, 2019 · Though a majority of SSRs in genomes are present at intergenic and non-coding regions, a small proportion of SSRs occur within exons [3, 9].
  17. [17]
    Functional Mechanisms of Microsatellite DNA in Eukaryotic Genomes
    Aug 24, 2017 · They are highly abundant in the noncoding DNA of all eukaryotic genomes studied, covering 1–3% of the human genome, depending on how they are ...
  18. [18]
    Genome-wide analysis of microsatellite repeats in humans
    Simple sequence repeats are found in most organisms, and occupy about 3% of the human genome. The densities of simple sequence repeats across the human ...Abundance Of Ssrs In The... · Dinucleotide Repeats · Figure 7
  19. [19]
    Conservation of Human Microsatellites across 450 Million Years of ...
    Out of 696,016 microsatellites found in human sequences, 85.39% were conserved in at least one other species, whereas 28.65% and 5.98% were found in at least ...
  20. [20]
    Patterns of microsatellite distribution reflect the evolution of ... - bioRxiv
    Jan 25, 2018 · The distribution of SSRs in coding and non-coding regions reveals taxon-specific variations in their exonic, intronic and intergenic densities.
  21. [21]
    The abundance of various polymorphic microsatellite motifs differs ...
    The GT/CA motif being the most abundant dinucleotide repeat in mammals was found to be considerably less frequent in plants.Missing: higher | Show results with:higher
  22. [22]
    Homopolymeric tracts represent a general regulatory mechanism in ...
    Feb 9, 2010 · Analyses of 81 bacterial and 18 archaeal genomes showed that poly(A) and poly(T) HTs are overrepresented in these genomes and preferentially ...Discussion · Methods · Homopolymeric Tract Search
  23. [23]
    Sequence, Chromatin and Evolution of Satellite DNA - MDPI
    Most telomeric repeats fall into the category of microsatellites, whereas subtelomeric repeats are categorized as satellites. Telomeres consist of short tandem ...
  24. [24]
    TRF website - Tandem Repeats Finder
    Tandem Repeats Finder locates and displays tandem repeats in DNA sequences, which are two or more adjacent, approximate copies of a pattern of nucleotides.Submit a Sequence · Download TRF Executable · TRF Definitions
  25. [25]
  26. [26]
    Hypervariable 'minisatellite' regions in human DNA - Nature
    Mar 7, 1985 · The human genome contains many dispersed tandem-repetitive 'minisatellite' regions detected via a shared 10–15-base pair 'core' sequence.
  27. [27]
    A hypervariable microsatellite revealed by in vitro amplification of a ...
    Using the polymerase chain reaction to amplify a (TG)n microsatellite in the human cardiac actin gene, we detected 12 different allelic fragments in 37 ...
  28. [28]
    Abundant class of human DNA polymorphisms which can be typed ...
    Use of the polymerase chain reaction to detect DNA polymorphisms offers improved sensitivity and speed compared with standard blotting and hybridization.
  29. [29]
    Use of variable simple sequence motifs as genetic markers
    May 8, 1989 · Use of variable simple sequence motifs as genetic markers: application to study of myotonic dystrophy ... myotonic dystrophy gene region at 19q.
  30. [30]
    The Human Genome Project: from mapping to sequencing - PubMed
    The genetic map comprises about 8000 highly informative second generation markers of the microsatellite type. ... genome radiation hybrids that enable ...
  31. [31]
    A novel gene containing a trinucleotide repeat that is expanded and ...
    A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. ... 1993 Mar 26;72(6):971-83.Missing: discovery | Show results with:discovery
  32. [32]
    The Application of Single Nucleotide Polymorphism Microarrays in ...
    Integration of global SNP-based mapping and expression arrays reveals key regions, mechanisms, and genes important in the pathogenesis of multiple myeloma.
  33. [33]
    Microsatellites in Pursuit of Microbial Genome Evolution - Frontiers
    Microsatellites or short sequence repeats are widespread genetic markers which are hypermutable 1–6 bp long short nucleotide motifs.
  34. [34]
    Application of Microsatellite Markers in Conservation Genetics and ...
    Apr 7, 2014 · Microsatellites are very powerful genetic markers for identifying fish stock structure and pedigree analysis and to study the genetic variation of closely ...Missing: key milestones
  35. [35]
    Precise CAG repeat contraction in a Huntington's Disease mouse ...
    Jun 23, 2021 · The most straightforward application of CRISPR/Cas9 is to excise the expanded repeat tracts. This was accomplished by designing two gRNAs ...
  36. [36]
    Deepath-MSI: a clinic-ready deep learning model for microsatellite ...
    Aug 28, 2025 · Deep learning models developed from WSIs of tumor H&E slides have demonstrated significant potential in predicting MSI status in CRC, with area ...
  37. [37]
    Beyond Junk-Variable Tandem Repeats as Facilitators of Rapid ...
    Variable TRs in promoters can affect gene expression by altering the number of transcription factor binding sites. Frequent TR variability of repeat ...
  38. [38]
    RNA biology of disease-associated microsatellite repeat expansions
    Aug 29, 2017 · R-loops associated with triplet repeat expansions promote gene silencing in Friedreich ataxia and fragile X syndrome. PLoS Genet. 2014;10 ...
  39. [39]
    Polymorphic CAG Repeat and Protein Expression of Androgen ...
    The CAG repeat length of AR inversely affects its transactivation potential, either as a directly altered receptor function (7, 8) or indirectly reduced AR ...
  40. [40]
    Interplay Between Polymorphic Short Tandem Repeats and Gene ...
    Mar 31, 2023 · Furthermore, we found that distant eSTRs might affect gene expression by disrupting miRNA binding in the 3′ UTRs of genes encoding TFs, such as ...
  41. [41]
    Constraints on Allele Size at Microsatellite Loci - NIH
    The evolution of selectively neutral markers is governed by the interaction of mutation and random genetic drift. Mutation pressure has the inherent ...
  42. [42]
    Microsatellites as Molecular Markers with Applications in ...
    Microsatellites have been known as hypervariable neutral molecular markers with the highest resolution power in comparison with any other markers.
  43. [43]
    Microsatellites as Targets of Natural Selection - PMC - NIH
    Microsatellites have long been used as markers in population genetics and forensic analysis because they are often highly variable (Oliveira et al. 2006). An ...
  44. [44]
    Simple sequence repeats and their expansions: role in plant ...
    May 5, 2025 · Soon after, plant geneticists picked up the usage of microsatellites ... flowering time protein correlates with Island age in a Hawaiian plant ...
  45. [45]
    A hybrid zone of the genus Ctenomys: A case study in southern Brazil
    Microsatellite markers have been used in studies of hybrid zones, where they have provided insights into introgression, population structure and gene flow ( ...Material And Methods · Dna Extraction And... · Results
  46. [46]
    A Role for Selection in Regulating the Evolutionary Emergence of ...
    There is no a priori reason to expect tandem repeats of CAG to lie in any particular reading frame of an exon unless selection has constrained the reading ...
  47. [47]
    Full article: Transcription-induced DNA toxicity at trinucleotide repeats
    Microsatellites in general tend to be excluded from the coding regions of genes because their instability plays havoc with the reading frame, but trinucleotide ...
  48. [48]
    Slipped-strand Mispairing: A Major Mechanism for DNA Sequence ...
    We propose that slipped-strand mispairing events, in concert with unequal crossing-over, can readily account for all of these features.
  49. [49]
    Replication stalling and DNA microsatellite instability - PMC - NIH
    Noncanonical microsatellite DNA structures are sites of polymerase stalling. Stalling at microsatellites is associated with repeat length instability and DNA ...Missing: papers | Show results with:papers
  50. [50]
    Microsatellite Instability in Cancer of the Proximal Colon - Science
    Microsatellite instability was significantly correlated with the tumor's location in the proximal colon (P = 0.003), with increased patient survival.
  51. [51]
    Mismatch Repair Pathway, Genome Stability and Cancer - Frontiers
    The postulated causative mechanisms of high microsatellite instability are DNA polymerase slippage during replication, deficient repair processes, and unequal ...
  52. [52]
    Microsatellite instability in tumors as a model to study the process of ...
    These results demonstrate that microsatellite mutations in unstable tumors show similarities to germline mutations and suggest that their study may be useful in ...
  53. [53]
    Relationship Between Microsatellite Slippage Mutation Rate and the ...
    When slippage mutations happen, expansion occurs more frequently for short microsatellites and contraction occurs more frequently for long microsatellites.
  54. [54]
    Sequence interruptions confer differential stability at microsatellite ...
    The inherently unstable nature of microsatellites results in frequent alterations in the length of the repeat tracts making many of them highly polymorphic (1,3) ...
  55. [55]
    Microsatellite evolution inferred from human– chimpanzee genomic ...
    The human genome is composed of 40–50% repetitive DNA, an important class being simple tandem repeats or microsatellite DNA sequences (1). Microsatellites are ...
  56. [56]
    Features of Evolution and Expansion of Modern Humans, Inferred ...
    ... mutation rate as 1.52×10−3 per dinucleotide locus per generation. Comparison of variation at tri- and tetranucleotide repeat loci with that at dinucleotide loci ...
  57. [57]
    Two Distinct Modes of Microsatellite Mutation Processes
    Tri-nucleotide repeats are overrepresented in coding se- quences, but less frequent than mono- and di-nucleotide repeats in noncoding regions (Tóth et al ...
  58. [58]
    Every Microsatellite is Different: Intrinsic DNA Features Dictate ... - NIH
    Microsatellite sequences are tandem repeats of short (1-6 base pair) DNA motifs that are ubiquitous in eukaryotic genomes. Approximately 3% of the human genome ...
  59. [59]
    The sequence of the repetitive motif influences the frequency of ...
    Jun 24, 2023 · So far, diverse studies have shown the influence of several factors on STRs mutation rates, such as the allele length, repeat motif size and ...
  60. [60]
    Comprehensive analysis of indels in whole-genome microsatellite ...
    Mar 16, 2020 · Third, we found that replication timing and DNA shape were significantly associated with mutation rates of microsatellites. Last, analysis of ...
  61. [61]
    Defective Mismatch Repair, Microsatellite Mutation Bias, and ... - NIH
    Jan 15, 2010 · Mutations arising within microsatellites associated with critical target genes are believed to play a causative role in the evolution of MMR- ...
  62. [62]
    The nucleotide composition of microsatellites impacts both ... - NIH
    In the eukaryotic genome, microsatellites are prone to frameshift mutations as a result of polymerase slippage during DNA replication (22). Such events involve ...Results · Egfp Fluorescence Pattern... · Mutation Spectra Of Mono...Missing: seminal | Show results with:seminal
  63. [63]
    Mutation Rate in Human Microsatellites: Influence of the Structure ...
    Our data demonstrate that mutation rates of different loci can differ by several orders of magnitude and that dif- ferent alleles at one locus exhibit different ...
  64. [64]
    Mutation Rates, Spectra, and Genome-Wide Distribution of ...
    We find that mutations occurred randomly across the genome, with no chromosomal, gene, or replication timing biases; however, mismatch repair defective cells ...
  65. [65]
    Heterozygosity increases microsatellite mutation rate, linking it to ...
    Biochemical experiments in yeast suggest a possible mechanism that would cause heterozygous sites to mutate faster than equivalent homozygous sites.
  66. [66]
    Microsatellite Mutation Models: Insights From a Comparison of ... - NIH
    Microsatellites interrupted by even a single point mutation exhibit a twofold decrease in their mutation rate when compared to pure AC repeats.
  67. [67]
    Microsatellites Within Genes: Structure, Function, and Evolution
    Like the SSRs in untranscribed regions, the SSRs in genes also show a higher mutation rate (instability) than non-repetitive regions. The fact that in human ...
  68. [68]
    Sequence variants affecting the genome-wide rate of germline ...
    Jun 29, 2023 · Around 3% of the human genome are short tandem repeats (STRs), some of which are polymorphic, i.e. microsatellites, and mutate several orders of ...
  69. [69]
    Oxidative stress accelerates repeat sequence instability and base ...
    Oct 23, 2025 · To better understand how oxidative stress interacts with the MMR pathway to influence mutagenesis and tumorigenesis, we employed a multifaceted ...
  70. [70]
    Trinucleotide Repeat Disorders - StatPearls - NCBI Bookshelf
    Dec 11, 2024 · Trinucleotide repeat disorders are caused by an abnormal number of triplet repeat sequences, either in the coding or noncoding regions, and ...
  71. [71]
    Understanding the molecular basis of fragile X syndrome
    Apr 1, 2000 · In patients with fragile X syndrome, the expanded CGG triplet repeats are hypermethylated and the expression of the FMR1 gene is repressed, ...
  72. [72]
    Selection Against Frameshift Mutations Limits Microsatellite ... - NIH
    Microsatellite enrichment is an excess of repetitive sequences characteristic to all studied eukaryotes. It is thought to result from the accumulated ...
  73. [73]
    Polyglutamine Ataxias: Our Current Molecular Understanding and ...
    Oct 20, 2021 · Polyglutamine (polyQ) ataxias are a heterogenous group of neurological disorders all caused by an expanded CAG trinucleotide repeat located in the coding ...
  74. [74]
    Repeat expansion diseases - PMC - NIH
    Not all repeat expansions diseases show significant anticipation or a clear parent of origin effect. OPMD, for example, does not show anticipation, which may ...
  75. [75]
    Short Tandem Repeat Expansions and RNA-Mediated ...
    Jul 9, 2019 · Although DM1 is caused by a DMPK 3' untranslated region (3'UTR) CTGexp and DM2 by an intronic CCTGexp in CNBP, they share a number of ...
  76. [76]
    Myotonic Dystrophy Type 2 - GeneReviews® - NCBI Bookshelf
    Sep 21, 2006 · CNBP intron 1 contains the complex repeat motif of (TG)n ... The presence of an expanded (TG)n(TCTG)n(CCTG)n repeat within CNBP causes DM2.
  77. [77]
    An Overview of Alternative Splicing Defects Implicated in Myotonic ...
    This review focuses on the cause and effects of MBNL and CELF1 deregulation in DM1, describing the molecular mechanisms underlying alternative splicing ...
  78. [78]
    Alu elements: know the SINEs | Genome Biology | Full Text
    Dec 28, 2011 · Alu elements are primate-specific repeats and comprise 11% of the human genome. They have wide-ranging influences on gene expression.
  79. [79]
    Alu repeats increase local recombination rates - PubMed
    Nov 16, 2009 · We show that the presence of a fixed AluY insertion is significantly predictive of an elevated local recombination rate within 2 kb of the insertion.Missing: microsatellite promote
  80. [80]
    Alu Repeats: A Source for the Genesis of Primate Microsatellites
    The association of an Alu element with a microsatellite repeat could result from the integration of an Alu element within a preexisting microsatellite repeat.
  81. [81]
    Initiation of translation of the FMR1 mRNA occurs predominantly ...
    Highly-stable secondary/tertiary structure within the 5′UTR region is expected to block translation initiation that occurs via scanning. In the current instance ...
  82. [82]
    Molecular Effects of the CTG Repeats in Mutant Dystrophia ... - NIH
    A body of work demonstrates that DMPK mRNAs containing abnormally expanded CUG repeats are toxic to several cell types.
  83. [83]
    Non-canonical DNA/RNA structures associated with the ...
    Aug 30, 2022 · Abnormal expansion of CGG repeat tracts above a certain threshold confers instability and chromosome fragility, resulting in various clinical ...
  84. [84]
    Chromatin changes in the development and pathology of the Fragile ...
    This review will discuss recent work done in our lab and elsewhere to better understand FXS gene silencing as well as the role of chromatin in repeat expansion, ...
  85. [85]
    Microsatellite instability (MSI) increases with age in normal somatic ...
    Small pool PCR (SP-PCR) is a sensitive method for the detection and quantification of microsatellite instability (MSI) in somatic cells.Missing: expansions | Show results with:expansions
  86. [86]
    Recurrent repeat expansions in human cancer genomes - Nature
    Dec 14, 2022 · In some cancers, mutations accumulate in short tracts of TRs, a phenomenon termed microsatellite instability; however, larger repeat expansions ...Missing: common | Show results with:common
  87. [87]
    Somatic mutations, genome mosaicism, cancer and aging - PMC - NIH
    In conclusion, somatic mutations accumulate with age in a tissue-specific manner, turning tissues into genome mosaics. Direct evidence for such mosaics has ...
  88. [88]
    DNA Amplification | CODIS Core Loci - National Institute of Justice
    Jul 31, 2023 · The purpose of this project was to evaluate various STR loci and to establish core loci for the CODIS. Crime laboratories from Alabama, Arizona, ...Missing: profiling | Show results with:profiling
  89. [89]
    Law Enforcement Databases: Limited Genetic Information and ...
    CODIS allows for a forensic DNA sample to be compared against a database of DNA profiles for a match (Fig. 17A). Since 2017, the DNA profiles uploaded into ...
  90. [90]
    Forensic DNA Profiling: Autosomal Short Tandem Repeat as a ... - NIH
    Aug 19, 2020 · Short tandem repeat (STR) typing continues to be the primary workhorse in forensic DNA profiling. Therefore, the present review discusses the prominent role of ...
  91. [91]
    [PDF] Recommendations on biostatistics in paternity testing - ISFG
    If the weight of the evidence is calculated, it shall be based on likelihood ratio principles. The paternity index (PI) is a likelihood ratio: PI ¼.
  92. [92]
    ARTICLE Paternity testing and forensic DNA typing by multiplex STR ...
    Paternity index (PI) was calculated for each STR locus, then the combined paternity index (CPI) was estimated by multiplying the individual paternity index ...
  93. [93]
    [PDF] Kinship and Parentage Analysis - NYC.gov
    Jul 1, 2024 · 1. Due to the possibility of mutation at STR loci between generations, two or more loci of the alleged father must not contain the obligate ...
  94. [94]
    Forensic Kinship and Paternity Testing: A Comprehensive Guide
    Apr 19, 2025 · Through advanced DNA profiling, scientists can now statistically assess not only parentage but also sibling, grandparent, avuncular, and distant ...
  95. [95]
    Development of a new screening method for faster kinship analyses ...
    Nov 27, 2022 · Wenk et al. used three independent polymorphic VNTRs loci to establish a sibling comparison test. STR multiplex markers are the predominantly ...
  96. [96]
    Forensic trace DNA: a review | Investigative Genetics - BioMed Central
    Dec 1, 2010 · DNA analysis is frequently used to acquire information from biological material to aid enquiries associated with criminal offences, disaster ...
  97. [97]
    Mini-STRs: A powerful tool to identify genetic profiles in samples ...
    Mini-STRs (using reduced-size STR amplicons) can help recovering information from degraded DNA samples by generating small PCR products.
  98. [98]
    An overview of DNA degradation and its implications in forensic ...
    Mar 15, 2024 · However, mini-STRs, which are more likely to amplify from degraded DNA samples, can be used to obtain usable DNA profiles from highly ...
  99. [99]
    Identical twins in forensic genetics — Epidemiology and risk based ...
    However, increased numbers of STR systems will not solve the problems caused by identical, monozygotic (MZ) twins, DNA contamination, errors, etc. Recent ...Missing: limitations | Show results with:limitations
  100. [100]
    Overview - The Evaluation of Forensic DNA Evidence - NCBI - NIH
    The main reason for departures from random-mating proportions in forensic DNA markers is population structure due to incomplete mixing of ancestral stocks.
  101. [101]
    Genetics in geographically structured populations: defining ...
    For example, recent analyses based on more than 370 short tandem repeat loci (microsatellites) and 600,000 SNPs suggest that only 5–10% of human genetic ...
  102. [102]
    The estimation of population differentiation with microsatellite markers
    Aug 6, 2025 · Microsatellite markers are routinely used to investigate the genetic structuring of natural populations. The knowledge of how genetic ...
  103. [103]
    Conservation Genetics of the Cheetah - PubMed Central - NIH
    While microsatellite markers demonstrated levels of heterozygosity in the cheetah that were not always significantly lower than for other species (Table 6.1), ...
  104. [104]
    A View of Modern Human Origins from Y Chromosome Microsatellite ...
    We analyze variation at 10 Y chromosome microsatellite loci that were typed in 506 males representing 49 populations and every inhabited continent and find ...Missing: migration | Show results with:migration
  105. [105]
    Microsatellites as Molecular Markers with Applications in ... - MDPI
    The use of molecular markers makes it possible to identify invasive species, reconstruct their migration routes and source populations, reveal their ...
  106. [106]
    Analysis of molecular variance inferred from metric distances among ...
    AMOVA uses DNA haplotype divergence to analyze variance, producing estimates of variance components and phi-statistics, reflecting diversity at different ...Missing: seminal microsatellites
  107. [107]
    Identifying the minimum number of microsatellite loci needed ... - NIH
    More than 95% of individuals can be correctly assigned using eight loci, and major population structure is visible with two highly polymorphic loci.
  108. [108]
    Revised Bethesda Guidelines for Hereditary Nonpolyposis ...
    Revised Bethesda Guidelines for Hereditary Nonpolyposis Colorectal Cancer (Lynch Syndrome) and Microsatellite Instability
  109. [109]
    Comparison of the Microsatellite Instability Analysis System ... - NIH
    14 The reference panel, referred to as the Bethesda panel, consists of two mononucleotide loci (Big Adenine Tract [BAT]-25and BAT-26) and three dinucleotide ...Bethesda Panel Assay · Figure 1 · Figure 2
  110. [110]
    Detection of microsatellite instability-high (MSI-H) by liquid biopsy ...
    Detection of microsatellite instability-high (MSI-H) by liquid biopsy predicts robust and durable response to immunotherapy in patients with pancreatic cancer.Missing: 2020s | Show results with:2020s
  111. [111]
    Huntington Disease (HD) CAG Repeat Expansion | Test Fact Sheet
    Oct 20, 2025 · Use to confirm the diagnosis of HD in symptomatic individuals. Use for presymptomatic testing in adults with a family history of HD.
  112. [112]
    [PDF] Genetic Testing Protocol for Huntington's Disease
    The presence of a CAG repeat expansion in a person with HD symptoms confirms the clinical impression and supports a diagnosis of HD. The absence of a CAG repeat.
  113. [113]
    detection of the paternally inherited expanded CAG repeat in ...
    Mar 12, 2015 · All fetal HD (n = 7) and intermediate (n = 3) CAG repeats could be detected in maternal plasma. Detection of repeats in the normal range (n = 10) ...
  114. [114]
    Identification and mapping of QTLs associated with drought ... - NIH
    These novel QTLs can be used for marker assisted breeding to develop new drought-tolerant rice varieties and fine mapping can be used to explore the functional ...
  115. [115]
    Mapping QTLs for Drought Tolerance at Seedling Stage in Rice ...
    A genetic linkage map with 226 SSR marker loci was constructed. Single-locus analysis following composite interval mapping (CIM) detected a total of five QTLs ...
  116. [116]
    [PDF] marker-assisted selection in livestock – case studies
    This chapter reviews the principles, opportunities and limitations for detection of quantitative trait loci (QTL) in livestock and for their use in genetic ...
  117. [117]
    Identification and Functional Analysis of Single Nucleotide ...
    In this study we identified a functional SNP in the VNTR of TS. TS VNTR has been considered a novel predictor of clinical outcome of 5-FU-based chemotherapy.
  118. [118]
    Pharmacogenomics DNA Biomarkers in Colorectal Cancer - Frontiers
    Oct 11, 2017 · Thymidylate synthase gene polymorphism determines response and toxicity of 5-FU chemotherapy. Pharmacogenomics J. 1, 65–70. doi: 10.1038/sj ...
  119. [119]
    Microsatellite Instable Colorectal Adenocarcinoma Diagnostics
    Jun 27, 2022 · The aim of this review is to summarize the main technical aspects and clinical applications, the benefits, and limitations of the use of liquid biopsy in MSI/ ...Missing: 2020s | Show results with:2020s
  120. [120]
    [PDF] AmpFlSTR™ Identifiler™ Plus PCR Amplification Kit
    All 13 of the required loci for the Combined DNA Index System (CODIS) loci are included in this kit for known-offender databasing in the United States. (Budowle ...
  121. [121]
    Introduction to Microsatellite and Microsatellite Genotyping
    Thermal cycling is performed in a PCR instrument under the following conditions: initial denaturation at 94°C for 5 min, 36 cycles of denaturation at 94°C for ...
  122. [122]
    PCR Amplification | An Introduction to PCR Methods
    In the next step of a cycle, the temperature is reduced to approximately 40–60°C. At this temperature, the oligonucleotide primers can form stable associations ...<|control11|><|separator|>
  123. [123]
    [PDF] Microsatellite DNA: Population Genetics and Forensic Applications
    A PCR procedure involving amplification of multiple loci simultaneously in one reaction is known as multiplexing. In a multiplex reaction, multiple pairs of ...
  124. [124]
    Performance comparison of gel and capillary electrophoresis-based ...
    Dec 7, 2021 · The gold standard for microsatellite genotyping is capillary electrophoresis (CE), a technology that accurately scores the alleles owing to its ...
  125. [125]
    [PDF] Technical Focus - Thermo Fisher Scientific
    In general, higher concentrations of magnesium result in higher stutter percentages thought to result from a lowered binding stringency allowing more efficient ...
  126. [126]
    Detection of microsatellite instability by real time PCR and ... - PubMed
    We have established a technique to detect MSI by LightCycler PCR and melting point analysis using sequence-specific hybridization probes (HyProbes) labeled ...
  127. [127]
    PCR Primer Design Tips - Behind the Bench - Thermo Fisher Scientific
    Sep 25, 2019 · Aim for the GC content to be between 40 and 60% with the 3' of a primer ending in G or C to promote binding. · A good length for PCR primers is ...
  128. [128]
    A review for researchers using microsatellites in the 21st century
    Jun 16, 2016 · Microsatellites have been used for a wide variety of applications, including genome mapping, forensics, parentage analysis, conservation ...
  129. [129]
    [PDF] Type-it® Microsatellite PCR Handbook - QIAGEN
    This handbook contains 2 protocols. Multiplex PCR for Amplification of Microsatellite Loci (Subsequent Analysis on. Sequencing Instruments). Choose this ...
  130. [130]
    Primer3 - Manual
    Summary of each segment:
  131. [131]
    New softwares for automated microsatellite marker development
    The program then analyses Primer3 output to cluster repeats being flanked by the same primer pairs. When all sequences have been analysed, the program outputs a ...<|separator|>
  132. [132]
    PCR Assay Design and Optimization | Bio-Rad
    The gradient feature allows you to test a range of temperatures simultaneously, optimizing the annealing temperature in a single experiment. To find the optimal ...
  133. [133]
    Optimization of PCR Conditions for Amplification of GC‐Rich EGFR ...
    Results showed that addition of 5% dimethyl sulfoxide (DMSO), as well as DNA concentration in PCR reaction of at least 2 μg/ml, were necessary for successful ...Missing: microsatellite | Show results with:microsatellite
  134. [134]
    Reduction of heteroduplex formation in PCR amplification
    Aug 6, 2025 · A common way to eliminate heteroduplex formation is to use reconditioning PCR. Because we detected that reconditioning PCR was not always ...
  135. [135]
    A study on the effects of degradation and template ... - PubMed
    Miniplex primer sets produced more complete profiles from degraded DNA compared to commercial kits, even at low template concentrations, increasing usable ...
  136. [136]
    [PDF] Validation of 15 microsatellites for parentage testing in North ...
    High PIC values, high heterozygosity and a large number of alleles. 2. Lack of known null alleles. 3. Loci non-syntenic or separated by more than. 40 cM.
  137. [137]
    [PDF] A practical approach to microsatellite genotyping with special ...
    D) Basic PCR protocol ... Following PCR amplification of a locus, polymorphisms are detected by.<|control11|><|separator|>
  138. [138]
    [PDF] Microsatellite genotyping errors: detection approaches, common ...
    Abstract. Microsatellite genotyping errors will be present in all but the smallest data sets and have the potential to undermine the conclusions of most ...Missing: limitations artifacts
  139. [139]
    Challenges in analysis and interpretation of microsatellite data for ...
    We present here a literature review that synthesizes the limitations of microsatellites in population genetic studies.
  140. [140]
    [PDF] Significant deviations from Hardy–Weinberg equilibrium caused by ...
    Four of the most common types of genotyping errors create a bias towards increased homozygosity. These include allelic dropout; null alleles; misinterpretation ...Missing: artifacts | Show results with:artifacts
  141. [141]
    Quantifying Ascertainment Bias and Species-Specific Length ...
    Because ascertainment bias should not exist if a microsatellite selected in one species is as likely to be longer as it is to be shorter than its homologue, we ...
  142. [142]
    High-Throughput Sequencing Strategy for Microsatellite Genotyping ...
    Mar 8, 2018 · We present a novel method for microsatellite genotyping using Illumina combinatorial barcoding that dispenses exhaustive PCR calibrations.
  143. [143]
    NanoSatellite: accurate characterization of expanded tandem repeat ...
    Nov 14, 2019 · We show that long-read sequencing with a single Oxford Nanopore Technologies PromethION flow cell per individual achieves 30× human genome ...
  144. [144]
    Inter-Simple Sequence Repeats (ISSR), Microsatellite-Primed ...
    Inter-simple sequence repeat (ISSR) markers are highly polymorphic, relatively easy to develop, and inexpensive compared to other methods and have numerous ...
  145. [145]
    Use of CRISPR-modified human stem cell organoids to study the ...
    Sep 14, 2017 · We used CRISPR-Cas9 technology to delete key DNA repair genes in human colon organoids, followed by delayed subcloning and whole-genome sequencing.
  146. [146]
    Reliable wolf-dog hybrid detection in Europe using a reduced SNP ...
    Jun 25, 2021 · We showed that the proposed SNP panel is an efficient tool for detecting hybrids up to the third-generation backcrosses to wolves across Europe.