Fact-checked by Grok 2 weeks ago

Microbial genetics

Microbial genetics is the subdiscipline of genetics that examines the heredity, variation, and evolution of microorganisms, including bacteria, archaea, viruses, and fungi, with a focus on their genetic material—primarily DNA, though some viruses use RNA—and the mechanisms governing its replication, expression, and transfer. This field reveals how microbial genomes, often circular and compact in prokaryotes like Escherichia coli (with approximately 4.6 million base pairs encoding around 4,300 genes), enable rapid adaptation through processes such as semiconservative DNA replication, which occurs bidirectionally from a single origin in bacteria at speeds of up to 1,000 nucleotides per second. Central to microbial genetics are the molecular processes of , where DNA is transcribed into RNA by —using a single in prokaryotes for polycistronic mRNAs—and translated into proteins via ribosomes, following the nearly universal triplet that starts with AUG for . Regulation of these processes occurs through operons, such as the in E. coli, which coordinates inducible expression of genes for in response to environmental cues, and the for repressible synthesis. , arising spontaneously at rates of about 10^{-10} per per generation or induced by mutagens like UV light or chemicals, introduce heritable changes that drive diversity, often repaired by mechanisms including proofreading during replication and mismatch excision repair. A hallmark of microbial genetics is , which contrasts with vertical inheritance and accelerates evolution in microbial populations through three primary mechanisms: , the uptake of free DNA by competent cells as demonstrated in Streptococcus pneumoniae; , where bacteriophages shuttle DNA between hosts in generalized or specialized forms; and conjugation, involving direct cell-to-cell transfer via plasmids like the F factor in E. coli. These processes, combined with point mutations, gene duplications, and transposon activity, contribute to genomic dynamism, enabling microorganisms to acquire traits like antibiotic resistance via R plasmids or explore new ecological niches. The study of microbial genetics has profound implications for understanding , as microbes' short times and vast sizes allow real-time of genetic changes, and for , where techniques like the using Salmonella typhimurium detect mutagens, and leverages these mechanisms for applications in and .

Fundamentals and History

Definition and Scope

Microbial genetics is the branch of genetics that examines heredity, variation, and gene function in microorganisms, encompassing the structure, organization, replication, and expression of their genetic material. This field primarily investigates prokaryotic organisms such as bacteria and archaea, as well as eukaryotic microbes including fungi and protozoa, and viruses, which serve as model systems due to their simple genetic architectures. For instance, Escherichia coli stands out as a key model organism in bacterial genetics, valued for its well-characterized genome and ease of genetic manipulation. The scope of microbial genetics is delimited to unicellular or simply organized microorganisms, excluding complex multicellular organisms, which allows for focused study of genetic processes in isolation from higher-order developmental complexities. Prokaryotic microbes, lacking membrane-bound nuclei, feature compact genomes often arranged on single circular chromosomes, while eukaryotic microbes possess nuclei and more compartmentalized genetic systems. This unicellular nature facilitates rapid reproduction cycles—such as the 20-minute generation time in E. coli under optimal conditions—and enhances experimental tractability through techniques like and genetic mapping. Microbial genetics provides foundational insights into microbial diversity and evolutionary dynamics, revealing how genetic mechanisms underpin and across vast microbial populations. It also elucidates human-relevant impacts, including microbial roles in infectious diseases through virulence gene and in via harnessing genetic tools for applications like antibiotic production.

Historical Development

The foundations of microbial genetics were laid in the 19th century through the pioneering work of Louis Pasteur and Robert Koch, who established the germ theory of disease and demonstrated the hereditary stability of microbial traits in processes like fermentation and pathogenesis. Pasteur's experiments in the 1860s refuted spontaneous generation, showing that microbes reproduce true to type, while Koch's isolation of pure cultures in the 1880s enabled observations of consistent inheritance in bacterial strains. A major breakthrough occurred in 1928 when Frederick Griffith reported the "transforming principle" in Streptococcus pneumoniae, observing that heat-killed virulent bacteria could transfer virulence to live non-virulent strains in mice, suggesting a heritable factor. This discovery implied genetic material could be transferred between microbes. In 1944, Oswald Avery, Colin MacLeod, and Maclyn McCarty confirmed that deoxyribonucleic acid (DNA) was the transforming agent, providing the first direct evidence that DNA serves as the genetic material in bacteria. The molecular era advanced rapidly in the mid-20th century. and Crick's 1953 elucidation of DNA's double-helix structure provided a model applicable to microbial genomes, explaining how genetic information could be stored and replicated. In 1958, and Franklin Stahl's experiments with demonstrated semi-conservative , confirming the mechanism by which bacterial genetic fidelity is maintained during . The 1970s marked the advent of genetic engineering with the development of recombinant DNA technology. Paul Berg's 1972 construction of the first recombinant DNA molecule using virus and laid the groundwork, while Stanley Cohen and Herbert Boyer's 1973 experiments successfully cloned and expressed foreign genes in E. coli, enabling manipulation of microbial genomes. In 1983, invented the (), a technique to amplify specific DNA segments exponentially, revolutionizing microbial genetic analysis and cloning. Modern microbial genetics took shape in the 1990s with the first complete sequencing of a free-living organism's . In 1995, and colleagues published the 1.83 million (1,830,138 bp) sequence of , ushering in the genomic era and revealing insights into gene organization and function. Subsequent milestones included the 1996 sequencing of the first archaeal , Methanococcus jannaschii, highlighting distinctions between and , and the identification of CRISPR-Cas systems in bacterial genomes in 2007, which enabled precise technologies by the 2010s.

Core Genetic Mechanisms

DNA Replication and Repair

DNA replication in microbial genetics ensures the accurate duplication of genetic material during cell division, primarily through a semi-conservative mechanism where each parental DNA strand serves as a template for synthesizing a complementary daughter strand. This process is highly conserved across microbes, with bacteria like Escherichia coli serving as a model organism due to their well-characterized machinery. Key enzymes include DNA polymerase III, which catalyzes the addition of nucleotides in the 5' to 3' direction; DnaB helicase, which unwinds the double helix at the replication fork; DnaG primase, which synthesizes short RNA primers to initiate DNA synthesis; and DNA ligase, which seals nicks between Okazaki fragments on the lagging strand. These components form the replisome, a dynamic complex that coordinates unwinding, priming, and polymerization to maintain genomic integrity. In E. coli, replication initiates at the (oriC), where proteins bind and facilitate the unwinding of DNA, recruiting the to form bidirectional replication forks. During elongation, the leading strand is synthesized continuously by DNA polymerase III, while the lagging strand is produced discontinuously in short , each primed by and later joined by after primer removal and gap filling. Termination occurs when the converging forks meet in the terminus region, opposed by Tus proteins binding to Ter sites to prevent over-replication. This entire process duplicates the ~4.6 million genome in approximately 40 minutes under optimal conditions, with replication forks progressing at a speed of ≈1000 per second. In contrast, eukaryotic microbes like exhibit slower replication rates, around 100 per second, and longer cycle times due to larger genomes and multiple origins. Microbial DNA repair mechanisms correct errors and damage to preserve replication fidelity, with employing excision-based pathways to address specific lesions. (BER) removes damaged or modified bases via , creating an abasic site that is processed by apurinic/apyrimidinic endonuclease, , and to restore the correct nucleotide. (NER) handles bulky distortions, such as UV-induced dimers, through the UvrABC system: UvrA and UvrB recognize the damage, UvrC excises a short segment, and the gap is filled by and sealed by . Mismatch repair (MMR) targets replication errors, with MutS detecting base mismatches, MutL recruiting MutH to nick the unmethylated daughter strand, and UvrD facilitating strand removal for resynthesis. These pathways achieve error rates as low as 10^{-10} per , far surpassing uncorrected replication fidelity. In prokaryotes, the rapid replication cycle (20-40 minutes) enables quick adaptation but increases vulnerability to errors, mitigated by high-fidelity repair; however, under stress like exposure, activate the response, inducing error-prone polymerases (e.g., Pol II, , ) that prioritize survival over accuracy, facilitating for resistance evolution. Failures in these repair systems can lead to persistent mutations, contributing to in microbial populations.

Mutation and Genetic Variation

Mutations in microbial genetics refer to heritable changes in the DNA sequence that can alter gene function and contribute to genetic variation within populations. These changes arise spontaneously during DNA replication or are induced by environmental agents, serving as a primary source of genetic diversity in bacteria, archaea, viruses, and eukaryotic microbes. Unlike higher organisms, microbes' rapid reproduction cycles amplify the impact of mutations, allowing for quick adaptation to selective pressures such as antibiotics or host defenses. Point mutations, the most common type, involve the substitution of a single base for another, categorized as transitions ( to or to , e.g., A to G) or transversions ( to or vice versa, e.g., A to C). Insertions add one or more , while deletions remove them; both can cause frameshift mutations if occurring in non-multiples of three bases, shifting the during and often leading to nonfunctional proteins. Spontaneous mutations occur naturally due to replication errors, tautomeric shifts in bases, or spontaneous (e.g., to uracil), with error rates partially offset by mechanisms. Induced mutations result from external mutagens, such as (UV) radiation causing thymine dimers or chemical agents like alkylating compounds that modify bases, leading to mispairing during replication. Microbial mutation rates typically range from 10^{-10} to 10^{-9} per per generation in like Escherichia coli, reflecting a balance between replication fidelity and evolutionary flexibility; for a of approximately 4.6 million s, this equates to about 0.002–0.005 mutations per per generation. Factors influencing rates include (smaller genomes exhibit higher per-base rates for equivalent genomic load) and environmental stressors, with viruses showing higher rates (up to 10^{-5} per base per generation) due to error-prone RNA-dependent RNA polymerases. In , hypermutation—often 100- to 1,000-fold elevated rates—arises from defects in methyl-directed mismatch repair (MMR) systems, such as mutations in mutS or mutL genes, facilitating rapid evolution in chronic infections like . Genetic variation in microbes stems from these mutations, including adaptive mutations where stressed, non-growing cells exhibit elevated mutagenesis targeted to specific loci, as observed in E. coli lac reversion systems under lactose selection. In viruses, the quasispecies model describes populations as dynamic clouds of closely related variants arising from high mutation rates, enabling rapid adaptation and persistence in heterogeneous environments. A key example is the role of mutations in antibiotic resistance, such as point mutations in the rpoB gene encoding RNA polymerase β-subunit, which confer rifampicin resistance by altering the drug-binding pocket; these mutations, often at codons 516, 526, or 531, arise at frequencies around 10^{-7} to 10^{-8} per cell and are prevalent in clinical isolates of Mycobacterium tuberculosis and other pathogens. Detection of mutations and mutagens relies on classic assays like the Luria-Delbrück fluctuation test (1943), which demonstrated the random, pre-selective origin of mutations by showing jackpot events in parallel bacterial cultures exposed to , distinguishing physiological adaptation from genetic change. The (1975), using histidine-requiring Salmonella typhimurium strains, detects mutagens by measuring reversion to prototrophy on minimal , often with added rat liver enzymes to mimic ; it has identified thousands of carcinogens with over 90% correlation to . Mutation rates are estimated using fluctuation analysis from the Luria-Delbrück experiment, such as the p0 method: \mu \approx -\frac{\ln(p_0)}{N_t} where p_0 is the proportion of cultures with no mutants and N_t is the final population size per culture. More precise methods, like maximum likelihood estimators, account for growth dynamics and clonal expansion.

Horizontal Gene Transfer

(HGT) refers to the movement of genetic material between microbial organisms other than by vertical inheritance from parent to offspring, playing a pivotal role in microbial by enabling rapid to environmental pressures. In and , HGT is a major driver of genetic diversity, allowing the acquisition of novel traits such as metabolic capabilities or survival advantages. This process was first evidenced in 1946 when and Edward L. Tatum demonstrated in through mixing auxotrophic mutants, revealing that could exchange genetic information in a manner analogous to in eukaryotes. Unlike , HGT facilitates the spread of beneficial genes across species boundaries, contributing significantly to microbial plasticity. The three primary mechanisms of HGT in microbes are , , and conjugation, each involving distinct molecular processes. involves the uptake of naked DNA from the environment by competent bacterial cells, a state induced by environmental signals such as nutrient limitation or . is mediated by proteins like ComEC, which forms a DNA import channel across the , and ComFA, an that translocates single-stranded DNA into the for recombination via homologous sequences of 25–200 base pairs. A classic example is , where observed in 1928 that heat-killed virulent cells could transform non-virulent strains into pathogenic ones, later confirmed as DNA-mediated by , Colin , and in 1944. This mechanism is prevalent in about 1–10% of bacterial species, particularly those in nutrient-rich environments like or host-associated niches. Transduction occurs through bacteriophage-mediated transfer of bacterial DNA, where phages accidentally package host genetic material during lytic cycles and deliver it to new host cells upon infection. There are two main types: generalized transduction, in which any bacterial DNA segment can be transferred randomly, and specialized transduction, involving specific genes adjacent to the prophage integration site, such as in lambda phage of E. coli. This process was discovered in 1952 by Norton Zinder and Joshua Lederberg while studying Salmonella typhimurium, where filtered lysates induced genetic changes without direct cell contact. Transduction is widespread in phage-abundant ecosystems, with up to 50% of bacterial genomes containing prophages, and it facilitates the horizontal spread of virulence factors, exemplified by Shiga toxin genes (stx) encoded in lambdoid prophages of enterohemorrhagic E. coli (EHEC), converting non-toxigenic strains into pathogens. Conjugation entails direct cell-to-cell transfer of DNA via a conjugation pilus, typically involving self-transmissible plasmids or integrative conjugative elements (ICEs). In Gram-negative bacteria like E. coli, the F (fertility) plasmid encodes the Tra (transfer) operon, which assembles a type IV secretion system to form a sex pilus that bridges donor and recipient cells, rolling-circle replication then exporting single-stranded DNA. This mechanism, building on Lederberg and Tatum's discovery, was mechanistically elucidated by William Hayes in 1953. Conjugation is highly efficient in dense populations, such as biofilms or the gut microbiome, and is a primary vector for antibiotic resistance genes carried on plasmids like RP4, which exhibit broad host ranges across bacterial genera. Mobile elements such as integrons and transposons further enhance HGT by capturing and mobilizing gene cassettes; integrons, with their integrase (IntI) and attI site, enable cassette excision and recombination, while transposons like IS elements promote plasmid-chromosome shuffling, amplifying resistance dissemination. HGT is particularly prevalent in prokaryotes, with genomic analyses indicating that up to 20% of genes in bacterial and archaeal genomes have been acquired horizontally, far exceeding rates in eukaryotic microbes where physical barriers like nuclei limit such exchanges. This prevalence underscores HGT's role in shaping pangenomes and fostering adaptability, as seen in the rapid evolution of pathogens. The impacts are profound: beyond antibiotic resistance, HGT disseminates virulence determinants, such as the in EHEC, which enhances pathogenicity and complicates responses. In clinical settings, conjugative plasmids have accelerated the global rise of multidrug-resistant strains, highlighting HGT as a key evolutionary force in microbial genetics.

Gene Expression and Regulation

Transcription and Translation

In microbial genetics, transcription and translation represent the core processes of the central dogma, converting genetic information from DNA to RNA to proteins, with notable variations between prokaryotic and eukaryotic microorganisms. Transcription involves the synthesis of RNA from a DNA template by RNA polymerase, while translation decodes messenger RNA (mRNA) into polypeptide chains using ribosomes and transfer RNAs (tRNAs). These processes are highly efficient in microbes, enabling rapid adaptation to environmental changes, and differ fundamentally in their machinery and coupling between bacteria and archaea (prokaryotes) versus eukaryotic microbes (such as fungi and protozoa).

Transcription

In bacteria, transcription is mediated by a single multisubunit RNA polymerase (RNAP) core enzyme, consisting of subunits α₂ββ'ω, which associates with a sigma (σ) factor to form the holoenzyme responsible for promoter recognition and initiation. The σ factor, such as the housekeeping σ⁷⁰ in Escherichia coli, directs the holoenzyme to promoter regions upstream of genes, where it binds specific DNA sequences: the -35 box (consensus TTGACA) and the -10 box (consensus TATAAT), spaced approximately 17 base pairs apart. These elements facilitate the formation of a closed promoter complex, followed by DNA unwinding to create an open complex, allowing RNA synthesis to begin at the transcription start site (+1 position). Elongation proceeds at rates of about 20–80 nucleotides per second, with the σ factor typically dissociating after promoter clearance.85006-a) In , transcription uses a single RNAP structurally similar to eukaryotic , composed of 11–13 subunits (including homologs of eukaryotic A', A'', B, etc.), which requires transcription factors TBP () and TFB (transcription factor B) for promoter recognition at TATA-box elements, without sigma factors. Transcription termination in occurs via two main mechanisms. Intrinsic (rho-independent) termination involves the formation of a GC-rich stem-loop structure in the nascent , followed by a uracil-rich sequence that destabilizes the RNAP--DNA complex, causing dissociation without additional proteins. Rho-dependent termination requires the Rho hexameric protein, which binds to C-rich rut sites on the nascent RNA, translocates along it in a 5' to 3' direction using ATP hydrolysis, and catches up to the RNAP at pause sites to induce termination. Bacterial transcripts are often polycistronic, encoding multiple proteins from a single mRNA, which supports coordinated in operons. In contrast, eukaryotic microbes employ three distinct RNA polymerases, each with specialized functions: (Pol I) transcribes most ribosomal RNAs (rRNAs), Pol II synthesizes mRNA precursors and some small nuclear RNAs, and Pol III produces RNAs (tRNAs) and 5S rRNA. Unlike the single bacterial RNAP, these eukaryotic enzymes require multiple general transcription factors for , and promoters vary by polymerase (e.g., TATA boxes for Pol II). Eukaryotic mRNAs are typically monocistronic, encoding a single protein, reflecting compartmentalized transcription separated from cytoplasmic translation.

Translation

Bacterial translation utilizes 70S ribosomes, composed of a small subunit and a 50S large subunit, to decode mRNA into proteins. Initiation begins with the subunit binding the mRNA via base-pairing between the Shine-Dalgarno (SD) sequence (consensus AGGAGG, located 8–10 nucleotides upstream of the AUG ) and the anti-SD sequence in 16S rRNA, aided by initiation factors IF1, IF2 (bound to fMet-tRNA^fMet^), and IF3. This forms the preinitiation , to which the 50S subunit joins, releasing factors and creating the 70S initiation with the initiator tRNA in the . Archaea also use 70S ribosomes but primarily initiate on leaderless mRNAs, where the small subunit binds near the 5' end and the large subunit joins directly at the , facilitated by archaeal factors (aIF1, aIF1A, aIF2, aIF5B) that are homologous to eukaryotic counterparts; Shine-Dalgarno sequences are used less frequently. During in , (EF-Tu), complexed with GTP and , delivers the cognate tRNA to the A site, where GTP enables and peptidyl transfer from the P-site tRNA to the new via the ribosome's center. Translocation of tRNAs and mRNA to the E, P, and A sites follows, driven by and GTP , advancing the by one . Termination occurs when a (UAA, UAG, or UGA) enters the A site, recruiting release factors RF1 or RF2 (recognizing specific codons) to hydrolyze the completed peptidyl-tRNA bond, with RF3 facilitating factor dissociation. In , is tightly coupled to transcription, as ribosomes can bind and initiate on nascent mRNA emerging from RNAP, enhancing efficiency and allowing real-time regulation. Protein synthesis proceeds at approximately 20 per second in E. coli. Eukaryotic microbes use ribosomes ( small and 60S large subunits) for translation, which is spatially separated from transcription in the . Initiation employs a cap-dependent scanning : the 43S preinitiation complex ( subunit with eIFs and Met-tRNA^i^Met^) binds the 7-methylguanosine cap via eIF4F, then scans the for the start codon, lacking the SD sequence used in . This process requires over 10 eukaryotic initiation factors (eIFs), contrasting with the simpler bacterial system, and results in subunit joining to form the elongation-competent ribosome. and termination s are analogous but involve eukaryotic elongation factors (eEF1A for aa-tRNA delivery, for translocation) and release factors eRF1/eRF3.

Regulatory Mechanisms

Microbial regulatory mechanisms primarily operate at the transcriptional level to control in response to environmental cues, ensuring efficient resource allocation and adaptation. In , these mechanisms often involve coordinated regulation of gene clusters through , where a single promoter directs the transcription of multiple into a polycistronic mRNA. This allows rapid, synchronized responses to nutrient availability or stress. The operon model, first proposed by Jacob and Monod, exemplifies negative and positive control, with repressors binding sites to block transcription initiation and activators enhancing recruitment. The lac operon in Escherichia coli illustrates inducible regulation, where the lac repressor protein, encoded by the lacI gene, binds the operator sequence in the absence of lactose, preventing transcription of genes encoding β-galactosidase, lactose permease, and transacetylase. Upon lactose addition, allolactose binds the repressor, releasing it from the operator and allowing transcription. Positive regulation occurs via the catabolite activator protein (CAP), which, when bound to cyclic AMP (cAMP) during glucose scarcity, interacts with the promoter to facilitate RNA polymerase binding and increase transcription up to 50-fold. In contrast, the trp operon demonstrates repressible control coupled with attenuation, a mechanism that fine-tunes transcription termination based on tryptophan levels. The trp repressor, activated by tryptophan binding, inhibits the promoter, while attenuation involves a leader sequence in the mRNA forming alternative hairpin structures: high tryptophan promotes a terminator hairpin, halting transcription before structural genes, whereas low tryptophan stalls the ribosome, favoring an antiterminator structure for full operon expression. This dual control represses the operon over 600-fold when tryptophan is abundant. Bacterial transcription initiation is further modulated by sigma factors, which associate with the core to recognize specific promoter sequences. The housekeeping sigma factor σ⁷⁰ in E. coli directs transcription of most constitutive genes under normal conditions, binding -10 (TATAAT) and -35 (TTGACA) consensus sequences to initiate basal expression. Alternative sigma factors, such as σ³⁸ for stationary phase or σ³² for heat shock, compete with σ⁷⁰ to redirect to stress-responsive promoters, enabling rapid shifts in without altering levels. Two-component systems provide environmental sensing through sensor kinases and response regulators; the sensor autophosphorylates upon stimulus detection (e.g., osmolarity or ), transferring the phosphate to the regulator's aspartate residue, which then binds DNA to activate or repress target genes. This histidine kinase-response regulator paradigm, ubiquitous in bacteria, coordinates responses like production in pathogens. In , regulation often involves TFB variants and , with organization similar to but initiation controlled by eukaryotic-like factors. In eukaryotic microbes, incorporates . Fungi, such as , utilize enhancers—distal DNA elements that boost transcription when bound by activators like Gal4, which recruit coactivators to loop and contact promoters, often synergizing with upstream activating sequences (UAS) for up to 1,000-fold activation of genes like those in metabolism. Protozoan parasites, including , employ complexes like ISWI or to alter positioning, exposing or occluding promoters during stages; for instance, facilitates variant surface glycoprotein expression switches essential for immune evasion. Global regulatory networks integrate multiple inputs for population-level control. Quorum sensing in uses autoinducers like N-acyl homoserine lactones (AHLs) to monitor cell density; at low densities, AHLs diffuse away, but accumulation at high densities activates LuxR-type receptors, inducing in Vibrio fischeri or formation in others, coordinating behaviors like . Small regulatory RNAs (sRNAs), typically 50-500 , provide fine-tuning by base-pairing with target mRNAs to modulate stability or translation, often Hfq-assisted; for example, RyhB sRNA represses iron genes during scarcity, conserving resources.00643-5) Virulence regulation exemplifies these mechanisms' integration, as in , where the ToxR transmembrane regulator senses environmental signals like temperature and pH to activate the ctxAB encoding and tcp genes for toxin-coregulated , forming a cascade with TcpP that ensures toxin production only in the host intestine.

Post-Transcriptional Control

Post-transcriptional control in microbial genetics encompasses mechanisms that modify, stabilize, or degrade RNA transcripts after synthesis, thereby fine-tuning at the RNA and protein levels. In prokaryotes such as and , these processes are streamlined to support rapid responses to environmental cues, often involving small regulatory RNAs (sRNAs) and ribonucleases that influence mRNA turnover and . In contrast, eukaryotic microorganisms like fungi and employ more complex RNA processing pathways, including capping, , and splicing, which ensure mRNA maturation and export from the . These controls integrate with upstream to enable adaptive responses, such as nutrient scavenging or stress tolerance, without altering the genome. RNA processing begins immediately after transcription and varies significantly between microbial domains. In bacteria, mRNA typically requires minimal processing and is polycistronic, allowing coordinated translation of multiple genes; however, ribosomal RNA (rRNA) and transfer RNA (tRNA) precursors undergo cleavage by specific endoribonucleases to generate mature forms. For instance, in Escherichia coli, RNase III processes the primary rRNA transcript into precursors that are further trimmed by exonucleases. In archaea, similar processing occurs, but with additional eukaryotic-like features in some rRNA modifications. In eukaryotic microbes, such as the yeast Saccharomyces cerevisiae, pre-mRNA acquires a 5' cap (7-methylguanosine) shortly after transcription initiation to protect against degradation and facilitate translation, while a poly-A tail is added at the 3' end by poly-A polymerase to enhance stability and nuclear export. Splicing removes introns via the spliceosome, a process essential for generating functional mRNAs in fungi and protozoa. tRNA and rRNA maturation in these organisms involves similar endonucleolytic cleavages but within nucleolar compartments. mRNA stability and degradation represent a primary layer of post-transcriptional control, determining the lifespan of transcripts and thus protein output. In bacteria, RNase E serves as a central endoribonuclease that initiates most mRNA decay by cleaving single-stranded regions, often in a 5'-monophosphate-dependent manner, and assembles the RNA degradosome complex with exonucleases like PNPase for complete degradation. This process is modulated by sRNAs, which base-pair with target mRNAs to expose RNase E cleavage sites, accelerating turnover under stress conditions. In archaea, degradation pathways involve homologs of eukaryotic exonucleases, with less characterized sRNA roles. In eukaryotic microbes, mRNA decay pathways involve decapping enzymes (e.g., Dcp2 in yeast) followed by 5'-3' exonucleolytic degradation by Xrn1, or deadenylation-dependent 3'-5' decay by the exosome complex; quality control mechanisms, such as nonsense-mediated decay, further eliminate aberrant transcripts. In protozoan parasites like Trypanosoma brucei, siRNAs derived from double-stranded RNA precursors mediate gene silencing by guiding the RNA-induced silencing complex (RISC) to cleave target mRNAs, contributing to transcriptome homeostasis.30064-4) Translational regulation occurs through RNA elements that modulate ribosome binding or initiation without altering mRNA levels. Riboswitches, structured RNA domains in the 5' untranslated region (UTR) of bacterial mRNAs, bind metabolites to undergo conformational changes that either sequester the ribosome binding site or promote mRNA degradation. The thiamine pyrophosphate (TPP) riboswitch, prevalent in bacteria like Bacillus subtilis and even in eukaryotic microbes such as plants and fungi, represses thiamine biosynthesis genes upon TPP binding, preventing unnecessary cofactor production. Antisense RNAs and sRNAs further regulate translation by pairing with mRNA targets to block ribosome access or recruit ribonucleases; for example, in Gram-positive bacteria, sRNAs form extensive posttranscriptional regulons affecting dozens of genes. In archaea, similar sRNA-mediated controls influence translation efficiency, though less characterized. CRISPR-associated RNAs (crRNAs) exemplify post-transcriptional processing in microbial defense systems. In , pre-crRNA transcripts from CRISPR arrays are cleaved by endoribonucleases (e.g., Cas6 in Type I systems) into mature crRNAs, which then guide proteins to degrade invading nucleic acids, ensuring immunity without ongoing transcription. This maturation step allows precise spatiotemporal control of antiviral responses. These mechanisms enable microbes to adapt rapidly to fluctuating environments. In , sRNAs like RyhB paralogs regulate iron homeostasis by repressing non-essential iron-utilizing proteins during scarcity, conserving the metal for core processes like and ; this posttranscriptional layer complements iron-responsive transcriptional regulators, enhancing survival in host niches. Similarly, sRNA networks in biofilms modulate and , promoting community resilience. Overall, post-transcriptional controls provide a dynamic, energy-efficient means for microbes to optimize in response to ecological pressures.

Microbial Genomes and Diversity

Genome Structure in Prokaryotes

Prokaryotic genomes, particularly those of bacteria and archaea, exhibit a compact organization that supports efficient replication and gene expression in diverse environments. Bacterial genomes are typically composed of a single, circular chromosome, as exemplified by Escherichia coli, which has a genome size of approximately 4.6 megabases (Mb) containing about 4,400 genes. This circular structure facilitates bidirectional replication from a single origin, oriC, and lacks the linear telomeres and introns common in eukaryotic genomes. The guanine-cytosine (GC) content in bacterial genomes varies widely, ranging from as low as 13% in some obligate symbionts to up to 75% in certain actinobacteria, influencing DNA stability, codon usage, and adaptation to environmental stresses such as temperature or salinity. A key organizational feature is the presence of operons, clusters of functionally related genes transcribed together under a single promoter, which enable coordinated regulation of metabolic pathways, as seen in the lac operon of E. coli for lactose metabolism. Archaeal genomes share structural similarities with bacterial ones, including a predominantly circular and compact arrangement, but they incorporate -like proteins that aid in DNA compaction and organization, akin to eukaryotic . These small, basic proteins, such as HMfA and HMfB in hyperthermophilic , form nucleosome-like structures that wrap DNA and regulate access during transcription, particularly in extremophiles adapted to harsh conditions like high salt or heat. For instance, the halophilic Halobacterium salinarum possesses a of about 2.6 Mb, divided into a main and megaplasmids, with adaptations including high (around 68%) and genes for via compatible solutes. Archaeal genomes generally range from 0.5 to 5.8 Mb, reflecting their prokaryotic simplicity while accommodating specialized variants that enhance stability in extreme niches. Common features across prokaryotic genomes include a core set of essential genes required for basic cellular functions, estimated at around 300 in minimal genomes such as that of Mycoplasma genitalium, which spans only 0.58 Mb and encodes 482 protein-coding genes, many of which are vital for replication, transcription, and translation. Genomes also contain pseudogenes—non-functional relics of gene duplication or decay—and insertion sequences (IS), short mobile elements (typically 700–2500 base pairs) that promote genomic plasticity through transposition and recombination, as observed in genomes like Escherichia coli where IS elements number over 10. Prokaryotic genome sizes overall vary from 0.5 Mb in endosymbionts to over 10 Mb in free-living species with complex lifestyles, allowing for metabolic versatility. The pan-genome concept captures this diversity, comprising a core genome of universally shared genes (e.g., for housekeeping functions) and an accessory genome of strain-specific genes acquired via processes like horizontal transfer, enabling adaptation; for example, in Staphylococcus aureus, the core represents about 70% of genes, with the accessory fraction varying by 20–30% across strains. Insights from genome sequencing have illuminated prokaryotic structure, beginning with the first complete bacterial genome: Rd, sequenced in 1995 at 1.83 Mb using whole-genome shotgun assembly, which revealed 1,743 protein-coding and organization without prior . This milestone demonstrated the feasibility of sequencing compact prokaryotic genomes, paving the way for understanding features like gene density (averaging one per kilobase) and the absence of large intergenic regions.

Genome Structure in Eukaryotic Microbes

Eukaryotic microbes, including fungi and , organize their genetic material within a membrane-bound , a defining feature that enables compartmentalized and larger sizes compared to prokaryotes. These nuclear typically span 10 to 100 Mb, comprising multiple linear equipped with specialized structures such as telomeres at the ends and centromeres at central regions to facilitate and during . Fungal exemplify this organization; for instance, the Saccharomyces cerevisiae possesses a compact 12 Mb distributed across 16 linear , ranging from 0.23 Mb to over 2 Mb in length. This structure includes a modest number of introns, with approximately 250 ribosomal protein genes and other essential loci containing them, contributing to . Additionally, fungal feature mating-type loci, such as the MAT locus on III in S. cerevisiae, which encodes regulators determining a or α and influences and cell identity. Telomeres in these fungi consist of repetitive TG1-3 sequences bound by shelterin-like complexes, protecting ends from degradation, while centromeres are typically short, point-like sequences (100-120 bp) recognized by proteins for attachment. Protozoan genomes display greater variability in ploidy and composition, often adapted to parasitic lifestyles. The human malaria parasite Plasmodium falciparum has a 23 Mb haploid nuclear genome organized into 14 linear chromosomes (0.7-3.4 Mb each), characterized by extreme AT richness (~80%) that challenges sequencing and influences gene expression. This genome encodes about 5,300 genes, including families involved in host-pathogen interactions. In contrast, trypanosomes like Trypanosoma brucei exhibit a ~35 Mb genome with variable ploidy across life stages, featuring extensive arrays of variant surface glycoprotein (vsg) genes—over 1,000 copies in subtelomeric expression sites and silent archives—that enable antigenic variation to evade mammalian immune responses through periodic switching of surface coats. Protozoan chromosomes similarly bear telomeres with TTAGGG repeats and centromeres that vary from regional (several kb) to point-like, often embedded in heterochromatin to silence nearby genes. Repeat elements, such as transposons and subtelomeric repeats, constitute 10-30% of these genomes and drive expansions of pathogenesis-related gene families, like those for immune evasion or host invasion. Beyond the nucleus, eukaryotic microbial genomes encompass organellar components. Mitochondria in both fungi and maintain small, circular genomes (15-100 kb) encoding a subset of respiratory chain proteins, with S. cerevisiae mtDNA at ~85 kb containing 8 protein-coding genes. Algae-like , such as chromalveolates (e.g., apicomplexans or dinoflagellates), harbor genomes like (cpDNA), which are compact (30-200 kb) and encode photosynthetic or metabolic genes, though reduced in non-photosynthetic parasites like Plasmodium (, ~35 kb). These organellar genomes share reoccurring themes of gene loss, linear-to-circular transitions, and scarcity, reflecting endosymbiotic origins. The sequencing of the S. cerevisiae genome in 1996, completed by an international consortium, represented the first fully assembled eukaryotic genome at 12,068 kb, identifying ~6,225 open reading frames and setting a benchmark for microbial genomics. This achievement highlighted the prevalence of repeat elements (3.1%) and gene duplications in fungal evolution, informing subsequent protozoan projects.

Viral Genomes

Viral genomes exhibit remarkable diversity in structure and composition, distinguishing them from the cellular genomes of bacteria, archaea, and eukaryotic microbes. Unlike prokaryotic or eukaryotic genomes, which are typically double-stranded DNA, viral genomes can consist of single-stranded or double-stranded DNA or RNA, and they are often linear or circular, packaged within protein capsids. This diversity is systematically classified by the Baltimore classification system, proposed in 1971, which divides viruses into seven groups based on the nature of their nucleic acid and the mechanism of mRNA production for protein synthesis. Group I includes double-stranded DNA (dsDNA) viruses, such as herpesviruses, which replicate using host DNA polymerases. Group II comprises single-stranded DNA (ssDNA) viruses, exemplified by parvoviruses, whose small genomes (around 4-6 kb) require conversion to dsDNA for replication. Group III features double-stranded RNA (dsRNA) viruses, like rotaviruses, which use viral RNA-dependent RNA polymerases to transcribe mRNA from their segmented genomes. Groups IV and V encompass positive-sense single-stranded RNA (+ssRNA) and negative-sense single-stranded RNA (-ssRNA) viruses, respectively, with +ssRNA genomes (e.g., many picornaviruses) directly serving as mRNA, while -ssRNA genomes (e.g., influenza viruses) require transcription to positive sense. Groups VI and VII involve reverse transcription: ssRNA reverse-transcribing viruses like HIV (Group VI) and dsDNA reverse-transcribing viruses like hepatitis B (Group VII). Viral genome sizes generally range from 3 to over 300 , allowing for compact encoding of essential genes while maintaining high evolutionary flexibility; however, their rates are exceptionally elevated compared to cellular organisms, typically 10^{-3} to 10^{-5} substitutions per site per replication cycle for viruses, driven by error-prone polymerases lacking mechanisms. This high mutability contributes to rapid adaptation but constrains complexity. Distinct structural features further characterize genomes, including overlapping genes, where the same sequence encodes multiple proteins in different reading frames, a prevalent in compact genomes to maximize capacity without increasing size. For instance, overlapping open reading frames are common in ssRNA viruses like . Segmented genomes, consisting of multiple independent molecules, occur in viruses such as , which has eight -ssRNA segments totaling about 13.5 , facilitating genetic reassortment during co-infection. Additionally, some viruses integrate their genomes into host chromosomes as proviruses or prophages; the bacteriophage , a dsDNA , exemplifies this by site-specifically recombining its 48.5 genome into the chromosome during lysogeny. In microbial contexts, —viruses infecting and —highlight the functional significance of these genomic features. The T4 , a well-studied dsDNA phage, possesses a 169 kb linear genome encoding approximately 300 genes, including those for a complex tail structure and lysis functions. like T4 and contribute to microbial genetics by mediating , a form of where viral particles package and transfer host DNA between . Evolutionarily, viral populations evolve as dynamic quasispecies—clouds of closely related rather than uniform clones—due to high rates and large population sizes, enabling rapid diversification and to hosts or environments. This quasispecies underlies viral persistence and , as described in foundational studies on populations.

Organisms in Microbial Genetics

Bacteria

Bacteria represent a of microbial genetics research due to their genetic simplicity, rapid reproduction, and diverse physiological adaptations. Escherichia coli strain K-12 serves as the preeminent , prized for its well-characterized and ease of genetic manipulation, enabling foundational studies on gene regulation, replication, and metabolism. Similarly, Bacillus subtilis is a key model for investigating sporulation genetics, where the process involves a cascade of sigma factors that coordinate developmental . Unique genetic mechanisms in bacteria highlight their adaptive versatility. Endospore formation in B. subtilis is governed by the sigF regulon, which activates early sporulation genes in the forespore compartment, ensuring survival under harsh conditions. Biofilm development, critical for community behavior and persistence, is regulated by systems; in Pseudomonas aeruginosa, the Las and Rhl systems control expression of adhesins and exopolysaccharides via autoinducer signaling, promoting matrix production and antibiotic resistance. Pathogenic bacteria exemplify specialized genetic adaptations for host interaction. In Salmonella enterica serovar Typhi, the Vi antigen—a capsular polysaccharide—is encoded by the viaB locus, including genes like tviB and vexB, which enhance immune evasion and virulence during typhoid fever. Mycobacterium tuberculosis employs latency-associated genes, such as the DosR regulon, to enter dormancy within host granulomas, downregulating metabolism and upregulating survival factors like nitrate reductase to persist asymptomatically. Genetic tools have advanced bacterial studies profoundly. Lambda phage vectors, derived from bacteriophage λ, facilitate cloning of DNA fragments up to 20 kb by integrating into the E. coli chromosome via site-specific recombination, allowing stable propagation and mutagenesis. Bacterial artificial chromosomes (BACs), based on the F-plasmid, enable cloning of large inserts (100-300 kb) with low chimerism, supporting comprehensive genomic libraries and functional analyses in bacteria. Bacterial diversity manifests in cell wall genetics, particularly peptidoglycan biosynthesis. Gram-positive bacteria, like B. subtilis, possess genes encoding thick layers with extensive cross-linking via rodA and pbp families, incorporating teichoic acids for structural reinforcement. In contrast, Gram-negative bacteria, such as E. coli, feature thinner synthesized by similar core enzymes but with additional regulators like mrcA for periplasmic coordination, flanked by an outer membrane that influences for lipopolysaccharide integration. These differences underscore evolutionary divergences in envelope maintenance, with bacterial genomes generally comprising a single circular harboring these loci.

Archaea

Archaea possess a distinct genetic framework that bridges prokaryotic simplicity with eukaryotic-like molecular processes, particularly in information processing pathways. Their genomes, typically ranging from 0.5 to 5 , encode for unique adaptations enabling survival in extreme environments, such as high temperatures, , and acidity. Unlike , archaeal shares more similarities with eukaryotes, including the use of multiple origins of replication and homologs, though it retains prokaryotic efficiency. Key model organisms in archaeal genetics include Methanococcus maripaludis, a hydrogenotrophic whose 1.7 Mb genome contains about 1,700 protein-coding genes, many dedicated to pathways involving coenzyme M and methanofuran biosynthesis. Genes like mcr (methyl-coenzyme M reductase) and (heterodisulfide reductase) are central to this process, highlighting archaea's role in global carbon cycling. Similarly, NRC-1 serves as a model for halophilic adaptations, with its 2.6 Mb genome featuring genes for compatible solute accumulation, such as and glycine betaine transporters (e.g., betP), and for light-driven proton pumping to cope with hypersaline conditions. These organisms facilitate genetic manipulation studies due to their tractability and relevance to biotechnological applications like biofuel production. Archaea exhibit eukaryotic-like transcription machinery, utilizing (TBP) and transcription factor B (TFB) homologs to recruit II-like enzymes to promoters, differing from bacterial sigma factors. This system supports precise initiation at TATA-box elements, with multiple TBP and TFB variants in some species enabling combinatorial regulation. Membrane lipids in archaea are uniquely ether-linked isoprenoids, synthesized via the and enzymes like geranylfarnesyl diphosphate synthase (Ggdps) and digeranylglyceryl phosphate synthase (Dggps), conferring stability in extreme conditions compared to bacterial ester lipids. Extremophile archaea showcase specialized genetic responses to stress. In Thermococcus species, such as T. kodakarensis, heat-shock proteins like small Hsps (e.g., Hsp20 family) and chaperonins prevent at temperatures above 80°C, with genes upregulated via sigma-like factors during thermal stress. species, thermoacidophiles from the Crenarchaeota, possess early CRISPR-Cas systems, including type I-A and III-B variants with cas genes (e.g., cas1, cas2) that acquire spacers from viral invaders, providing adaptive immunity predating bacterial versions. Archaeal genomes generally harbor fewer plasmids than bacterial counterparts, with extrachromosomal elements often integrated as megaplasmids in but rare in other lineages, limiting autonomous replication events. (HGT) from is prevalent, particularly for metabolic genes like those for carbon fixation (e.g., Wood-Ljungdahl pathway variants) in methanogens, evidenced by phylogenetic incongruences in up to 20% of archaeal genes. Genetic diversity between major phyla underscores archaeal versatility: , encompassing methanogens and halophiles, feature genes for unique metabolisms like acetogenesis (e.g., synthase clusters), while Crenarchaeota, including thermoacidophiles, emphasize sulfur oxidation pathways (e.g., genes) and lack methanogenic capabilities, reflecting adaptations to distinct niches. These differences, analyzed through , reveal a core set of shared informational genes but divergent operational genes shaping ecological roles.

Fungi and Protozoa

Fungi and represent diverse eukaryotic microbes whose genetics underpin unique adaptations such as dimorphism and . In fungi, genetic mechanisms enable transitions between unicellular forms and multicellular filamentous growth, facilitating environmental colonization and host interactions. , including parasitic species, exhibit specialized genetic strategies for immune evasion and tissue invasion, often involving dynamic at telomeres or surface antigens. These organisms' genomes, typically larger and more compartmentalized than prokaryotic counterparts, feature introns, organelles like plastids in apicomplexans, and mating-type loci that regulate . Fungal genetic diversity manifests in unicellular yeasts, such as , which propagate via budding and maintain compact suited for rapid division, versus filamentous species like that form hyphae for nutrient foraging and produce secondary metabolites through clustered biosynthetic genes. These gene clusters, often regulated by global transcription factors like LaeA, encode enzymes for compounds such as aflatoxins, enabling ecological niches but also posing risks in . In contrast, protozoan diversity includes free-living forms and obligate parasites; apicomplexans like harbor apicoplasts—non-photosynthetic plastids derived from —containing a ~35 kb with ~50 genes essential for isoprenoid and , targeted by antibiotics like fosmidomycin.01921-5) Model fungi illustrate key genetic phenomena, including dimorphism in Candida albicans, where hyphal switching is governed by genes like UME6, HGC1, and EFG1 that activate filamentation under environmental cues such as serum or neutral pH, enhancing biofilm formation and tissue penetration. In Aspergillus species, secondary metabolite production involves velvet complex regulators (e.g., VeA, VelB) that coordinate ~30-50 gene clusters, with epigenetic modifications silencing or activating pathways in response to nutrient stress.00180-2) Protozoan genetics highlight parasitism strategies, as in Trypanosoma brucei, where variant surface glycoprotein (VSG) switching evades host immunity through ~1,000 VSG genes in subtelomeric arrays; expression-site switching via transcriptional recombination at telomeres allows rapid coat replacement every 7-10 days. Similarly, Entamoeba histolytica relies on invasion genes encoding cysteine proteinases (e.g., ACP1-5) and galectin-like lectins that degrade extracellular matrix and trigger host inflammation, with virulence correlated to higher expression in pathogenic strains versus non-invasive E. dispar.01043-0) Unique genetic features include mating types in yeasts, controlled by the MAT locus on chromosome III in S. cerevisiae, where MATa and MATα idiomorphs encode transcription factors that dictate a/α cell identity and suppress homothallic switching via HO endonuclease-mediated recombination with silent HML/HMR cassettes. In protozoa, telomere variation supports antigenic diversity; for instance, T. brucei telomeres (~100-300 bp of irregular TTAGGG repeats) facilitate VSG mosaics through RAD51-dependent recombination, while ciliate protozoa like Tetrahymena thermophila exhibit developmental telomere elongation from 300-400 bp to 20 kb post-division.00730-9) Genetic tools developed in these microbes have broad impact, notably the yeast two-hybrid system introduced in S. cerevisiae, which detects protein-protein interactions by fusing bait () and prey (activation domain) proteins to reconstitute GAL4 transcription, enabling high-throughput mapping of interactomes since its seminal description.

Evolutionary and Comparative Aspects

Role in Evolutionary Studies

Microbial genetics provides critical insights into by leveraging the short s and high mutation rates of microorganisms, allowing researchers to observe and manipulate evolutionary processes in laboratory settings that would be infeasible in multicellular organisms. For example, exhibits a minimal generation time of 20 minutes in rich media, facilitating the study of adaptation over thousands of generations within months. This rapid turnover enables , where selective pressures can be applied to track genetic changes in real time. A landmark study is the Long-Term Evolution Experiment (LTEE), started by Richard Lenski in 1988, which propagates 12 initially identical E. coli populations daily, amassing over 80,000 generations as of 2025 and documenting innovations like aerobic citrate metabolism in one lineage after 31,500 generations. Key evolutionary mechanisms in microbes align with broader theories, such as the , which asserts that most genetic variation arises from neutral mutations fixed by drift rather than selection, a pattern observable in microbial populations due to their vast effective sizes that amplify subtle selective effects. and subsequent divergence further drive innovation, as duplicated copies can evolve new functions without disrupting the original; in antibiotic resistance, this process amplifies resistance genes, enabling to adapt quickly to environmental stressors like drugs. Phylogenetic analyses rooted in microbial genetics have reshaped understandings of life's history, with 16S rRNA serving as a to infer divergence times based on sequence conservation, a tool pioneered by to delineate evolutionary relationships among prokaryotes. Woese's work culminated in the 1990 proposal of the —Bacteria, , and Eukarya—using 16S rRNA differences to reveal Archaea as a distinct lineage, overturning prior two-kingdom classifications and establishing rRNA as a universal chronometer for microbial . Insights into the (LUCA) emerge from conserved microbial genes, particularly ribosomal proteins among a core set of about 30 translation-related genes shared across domains, indicating LUCA had a DNA-based and complex metabolic capabilities around 4.2 billion years ago. Viral genetics within the microbial realm illustrates explosive evolutionary dynamics, as seen in influenza A viruses, where antigenic drift—gradual point mutations in and neuraminidase genes—allows seasonal evasion of host immunity, while —reassortment of segments between strains—can spawn pandemics by creating novel subtypes. These processes highlight how microbial-scale informs predictions of emergence and the co-evolution of hosts and microbes.

Comparative Microbial Genomics

Comparative microbial genomics involves the systematic comparison of complete or near-complete microbial sequences to elucidate evolutionary relationships, functional adaptations, and genetic exchanges among microorganisms. This field leverages high-throughput sequencing and bioinformatics to align and analyze genomes from diverse taxa, revealing patterns of conservation, variation, and innovation that underpin microbial phylogeny and . By integrating genomic data across species or strains, researchers can reconstruct phylogenetic trees, identify shared core functions, and detect events like (HGT) that challenge traditional vertical inheritance models. Key advancements have enabled the study of uncultured microbes through , expanding the scope beyond isolate-based comparisons to community-level insights. Central methods in comparative microbial genomics include whole-genome alignment and ortholog identification. Whole-genome alignment tools, such as , facilitate the detection of conserved syntenic regions while accommodating rearrangements, inversions, and insertions common in microbial evolution. employs a progressive alignment strategy based on locally collinear blocks (LCBs) to identify homologous segments across multiple genomes, providing a framework for visualizing structural variations. For ortholog identification, sequence similarity searches using (Basic Local Alignment Search Tool) are foundational, allowing reciprocal best-hit approaches to infer orthologous genes between genomes. Complementary resources like Clusters of Orthologous Groups (COGs) catalog orthologs across prokaryotic genomes, enabling functional annotations and comparative analyses by grouping genes into evolutionary lineages based on bidirectional best hits and phylogenetic consistency. Significant findings from comparative analyses highlight HGT signatures and phylogenomic revisions. Parametric methods detect HGT by identifying genomic regions with atypical compositional features, such as deviations in GC content, dinucleotide frequencies, or codon usage bias from the host genome's norms; for instance, genes acquired via HGT often exhibit GC profiles closer to the donor organism. These approaches, benchmarked against simulated datasets, reveal that up to 10-20% of genes in some bacterial genomes may result from HGT, influencing pathogenicity and metabolic versatility. In phylogenomics, comparative genome analyses have revised the tree of life; the discovery of Asgard archaea, through metagenome-assembled genomes, positioned them as the closest prokaryotic relatives to eukaryotes, with shared eukaryotic signature proteins like actin and ESCRT machinery supporting an archaeal host in eukaryotic origins. Core genome analyses, which focus on genes present in all strains of a species, have identified minimal gene sets essential for bacterial life, comprising approximately 250 genes involved in replication, transcription, translation, and basic metabolism across diverse phyla. Tools like modeling and further advance comparative studies. The concept, introduced by Tettelin et al. in their analysis of isolates, describes the full gene repertoire of a as the union of a conserved genome and a variable accessory genome, with mathematical models predicting open pan-genomes that expand indefinitely with additional strains sequenced. enables comparative genomics of uncultured microbes by reconstructing genomes directly from , bypassing cultivation biases and revealing novel lineages; for example, binning algorithms assemble metagenome-assembled genomes (MAGs) that can be aligned with cultured representatives to infer functional diversity in microbial communities. These approaches collectively underscore the dynamic nature of microbial genomes, informing broader evolutionary and functional inferences.

Applications and Impacts

Biotechnology and Genetic Engineering

Microbial genetics has revolutionized biotechnology by providing the foundational tools and model systems for genetic engineering, enabling the precise manipulation of DNA in microorganisms to produce valuable compounds industrially. Recombinant DNA technology, pioneered through discoveries in bacterial systems, allows the insertion of foreign genes into microbial hosts for scalable protein production and metabolic engineering. This approach leverages the genetic malleability of microbes, such as bacteria and yeast, to serve as cellular factories for biopharmaceuticals, biofuels, and other products. The development of restriction enzymes in the 1970s marked a pivotal advancement in recombinant techniques. , along with Hamilton Smith and , discovered these bacterial enzymes that cleave DNA at specific sequences, enabling the precise cutting and joining of genetic material; their work earned the 1978 in Physiology or Medicine. Shortly thereafter, the plasmid emerged as one of the first versatile cloning vectors for , featuring unique restriction sites for inserting foreign DNA and selectable markers for and resistance, facilitating efficient gene propagation and expression. E. coli serves as a premier model for protein expression due to its rapid growth and well-characterized genetics, with the T7 RNA polymerase system providing tight inducible control for high-yield production. Developed by F. William Studier, this bacteriophage-derived system uses T7 promoters to drive gene transcription selectively, minimizing leaky expression and enabling up to gram-per-liter yields of recombinant proteins. In parallel, yeast species like Pichia pastoris and are favored for expressing glycoproteins requiring eukaryotic post-translational modifications, such as proper N-linked , which is essential for protein stability and function in industrial applications. A landmark application was the production of recombinant human insulin by in 1978, where the insulin A and B chain genes were synthesized, cloned into E. coli using pBR322-derived vectors, and expressed separately before chemical assembly, marking the first commercial recombinant protein and demonstrating microbial genetics' potential for therapeutic manufacturing. In biofuel production, engineered microbes have been optimized for , such as E. coli strains modified to convert biomass-derived sugars into advanced fuels like at titers exceeding 20 g/L through pathway engineering of pyruvate decarboxylase and genes. Similarly, has been engineered for biodiesel precursors, with S. cerevisiae expressing plant-derived fatty acid synthases to produce convertible to fatty acid ethyl esters. The discovery of CRISPR-Cas9, derived from bacterial adaptive immunity systems, has transformed microbial genetic engineering by enabling precise, programmable genome editing. In 2012, Martin Jinek and colleagues demonstrated that the Cas9 endonuclease, guided by a chimeric single-guide RNA, cleaves target DNA sequences in vitro, adapting this bacterial mechanism for in vivo editing in microbes to introduce mutations, delete genes, or insert pathways for synthetic biology applications. This tool has accelerated the construction of microbial chassis for complex metabolic networks, building on natural bacterial processes like horizontal gene transfer that inspire modular DNA assembly. Synthetic genomics represents the pinnacle of microbial engineering, exemplified by the 2010 creation of the first synthetic bacterial by J. Craig Venter's team. They chemically synthesized a 1.08 million genome of Mycoplasma mycoides JCVI-syn1.0, transplanted it into a recipient M. capricolum , and achieved a self-replicating controlled by the artificial , validating bottom-up design principles for custom microbes in .

Medical and Pharmaceutical Uses

Microbial genetics has profoundly impacted medicine by revealing the molecular mechanisms underlying pathogen-host interactions, facilitating the development of diagnostics, and driving innovative therapeutics. The study of microbial genomes identifies virulence factors—genes or gene products essential for disease causation—that enable pathogens to colonize hosts, evade immunity, and cause tissue damage. For example, in Yersinia species, the type III secretion system (T3SS), encoded by genes on the pYV virulence plasmid, forms a needle-like apparatus to inject Yop effector proteins into host macrophages, disrupting phagocytosis and promoting systemic infection such as plague. This genetic framework has allowed targeted disruption of T3SS in preclinical models, highlighting its role in pathogenesis. Outbreak investigations rely on microbial genetic tools like (MLST), which sequences multiple housekeeping genes to assign allelic profiles and reconstruct phylogenetic relationships among isolates, enabling precise tracking of pathogen dissemination. MLST has been pivotal in resolving clonal relationships during epidemics of and , informing containment strategies and source attribution. Complementing this, whole-genome sequencing (WGS) provides higher resolution for real-time surveillance, as demonstrated in outbreaks where wgMLST identified transmission chains with greater sensitivity than traditional methods. Diagnostics have been revolutionized by genetic approaches, with PCR amplification of the 16S rRNA serving as a for identifying in clinical specimens, achieving up to 90% in culture-negative cases like endocarditis or prosthetic joint . WGS extends this by directly detecting resistance genes, such as those encoding beta-lactamases or efflux pumps, to predict phenotypic resistance and streamline empirical therapy. In Escherichia coli bacteremia, WGS accurately forecasted resistance to extended-spectrum beta-lactams in over 95% of cases, reducing inappropriate use. These tools are particularly vital for polymicrobial , where 16S sequencing distinguishes pathogens from commensals. Therapeutic applications draw directly from microbial genetics, notably in , where genome sequencing of bacteriophages ensures specificity to target pathogens without disrupting —a approach revived post-2000s amid multidrug-resistant infections. Clinical trials have shown phages lysing in patients, with safety profiles comparable to antibiotics and promising results in compassionate-use cases. Similarly, vaccine design exploits viral genetics; the human papillomavirus (HPV) vaccines, such as , utilize the L1 sequence to produce virus-like particles that mimic native virions, inducing neutralizing antibodies that prevent 90-100% of infections from high-risk types like HPV-16 and -18. Genetic analysis of HPV variants has further refined second-generation vaccines for broader coverage. Antibiotic resistance mechanisms are decoded through microbial , exemplified by the mecA gene in methicillin-resistant Staphylococcus aureus (MRSA), which encodes penicillin-binding protein 2a (PBP2a) with low affinity for beta-lactams, allowing synthesis to proceed under drug pressure. This staphylococcal cassette chromosome mec (SCCmec) element, acquired horizontally, underpins MRSA's global prevalence, with WGS revealing its spread in over 80% of hospital-associated strains. integrates these insights to tailor dosing, such as higher levels for mecA-positive isolates, minimizing treatment failures. The underscored microbial genetics' role in real-time response, with the Wuhan-Hu-1 (GenBank MN908947) enabling variant tracking via next-generation sequencing, which identified over 1,000 mutations by mid-2020 and informed vaccine updates against strains like Alpha and . This genomic surveillance, coordinated globally, traced transmission chains and predicted immune escape, averting widespread diagnostic delays.

Environmental and Ecological Applications

Microbial genetics plays a pivotal role in understanding and harnessing dynamics through , which enables the analysis of genetic material directly from environmental samples without the need for culturing individual organisms. In and environments, metagenomic approaches have revealed vast microbial diversity, including the functional genes driving nutrient cycling and plant-microbe interactions. For instance, early work demonstrated the potential of metagenomes to access the collective genomes of uncultured soil microbes, providing insights into their biosynthetic capabilities. This technique has been instrumental in analysis, where microbial communities interact closely with plant roots to influence growth and resilience. Additionally, has highlighted that over 99% of microbial in remain unculturable using traditional methods, underscoring the genetic richness of these ecosystems and the limitations of culture-based studies. In , microbial genetics facilitates the degradation of environmental pollutants by identifying and engineering key genes in . Dehalogenase genes in like Pseudomonas encode enzymes that cleave carbon-halogen bonds in halogenated organic compounds, enabling the breakdown of persistent pollutants such as chlorinated solvents. These genes have been studied for their role in detoxifying contaminated sites, with enhancing their efficiency in Pseudomonas putida strains for degrading compounds like . A notable application occurred during the 2010 in the , where metagenomic analyses identified hydrocarbon-degrading , including Alcanivorax and Cycloclasticus, whose alkane monooxygenase and other catabolic genes rapidly responded to the influx of crude oil, contributing to natural attenuation. Ecological roles of microbial genetics are evident in essential biogeochemical processes, such as nitrogen fixation mediated by nif genes in symbiotic bacteria like Rhizobium. These genes encode the nitrogenase enzyme complex, which converts atmospheric N₂ into ammonia, supporting plant nutrition in legume-rhizobia symbioses and enhancing soil fertility without synthetic fertilizers. In carbon cycling, methanogenic archaea utilize genes in the mcr operon to reduce CO₂ or acetate to methane, closing anaerobic degradation loops in wetlands and sediments and influencing global carbon flux. Links between microbial genetics and are pronounced in the of gases and the spread of resistance traits. The genetics of in , governed by pathways involving methyl-coenzyme M reductase encoded by mcr genes, contributes significantly to atmospheric CH₄ levels, with biological by methanogens accounting for about 74% of global and exacerbating warming. Furthermore, resistance genes (ARGs) in environmental microbes, such as those for efflux pumps (tet genes) and enzymatic inactivation, are naturally abundant in soils and water bodies, facilitating and posing risks to and human-impacted environments. Tools like stable isotope probing (SIP) advance the study of microbial genetics by linking specific functional genes to active community members . DNA- or RNA-SIP incorporates stable isotopes (e.g., ¹³C) into substrates, allowing the separation and sequencing of labeled nucleic acids from microbes assimilating those substrates, thus identifying genes involved in processes like pollutant degradation or nutrient turnover. This method has been refined for metagenomic integration, enabling targeted recovery of functional gene clusters from complex environmental consortia.

References

  1. [1]
    Genetics - Medical Microbiology - NCBI Bookshelf - NIH
    The genetic material of bacteria and plasmids is DNA. Bacterial viruses (bacteriophages or phages) have DNA or RNA as genetic material.Genetic Information in Microbes · Exchange of Genetic Information
  2. [2]
    Microbial Genetics and Evolution - PMC - NIH
    Jun 23, 2022 · The diversity observed in the microbial world allows the comprehension of the processes behind genetic diversity; bacteria, archaea, viruses, ...
  3. [3]
    [PDF] Mechanisms of Microbial Genetics
    We now know that within the shared overall theme of the genetic mechanism, there are significant differences among the three domains of life: Eukarya, Archaea, ...
  4. [4]
    Microbial Genetics - an overview | ScienceDirect Topics
    Microbial genetics is defined as the study of the hereditary material (genomes) of microorganisms, focusing on their evolution, genetic expression, and the ...
  5. [5]
  6. [6]
  7. [7]
    The Genetic Theory of Infectious Diseases: A Brief History and ...
    Compelling experimental evidence established the role of microbes (from Louis Pasteur to Robert Koch), leading to the germ theory of infectious diseases (~1870) ...
  8. [8]
    A structural view of bacterial DNA replication - PMC - PubMed Central
    Bacterial cells can replicate DNA with remarkable speed and fidelity: in Escherichia coli, the in vitro rate is estimated at ~1000 bp/s (based on chromosome ...
  9. [9]
    Escherichia coli DNA replication: the old model organism still holds ...
    This review provides an updated view on the Escherichia coli replication fork, in particular its structure, dynamics, and factors influencing DNA replication ...
  10. [10]
    Mechanisms of DNA replication termination - PMC - PubMed Central
    Feb 22, 2019 · DNA replication finishes when converging replication forks meet. During this process, called replication termination, DNA synthesis is completed.
  11. [11]
    Single-Molecule Studies of Fork Dynamics of Escherichia coli DNA ...
    The enzymes of the E. coli replisome duplicate DNA with remarkable efficiency: the replication fork moves at a rate approaching 1000 nucleotides per second ...Missing: speed | Show results with:speed
  12. [12]
    Bacterial DNA excision repair pathways - PMC - NIH
    Bacteria have evolved several DNA repair pathways to correct and repair the different types of DNA damages and non-canonical bases that occur.
  13. [13]
    The DNA Damage Inducible SOS Response Is a Key ... - Frontiers
    Aug 3, 2020 · The SOS error-prone polymerases promote an elevated mutation rate, generating genetic diversity and adaptation, including antibiotic resistance.
  14. [14]
    RecA and Specialized Error-Prone DNA Polymerases Are Not ...
    To cope with stressful conditions, including antibiotic exposure, bacteria activate the SOS response, a pathway that induces error-prone DNA repair and ...
  15. [15]
    Mutation, Repair and Recombination - Genomes - NCBI Bookshelf
    Many mutations are point mutations that replace one nucleotide with another; others involve insertion or deletion of one or a few nucleotides. Mutations result ...
  16. [16]
    Mutations | Microbiology - Lumen Learning
    Mutations also result from the addition of one or more bases, known as an insertion, or the removal of one or more bases, known as a deletion.
  17. [17]
    Rates of spontaneous mutation - PubMed - NIH
    Mutation rates in higher eukaryotes are roughly 0.1-100 per genome per sexual generation but are currently indistinguishable from 1/300 per cell division per ...Missing: bacterial | Show results with:bacterial
  18. [18]
  19. [19]
    Quasispecies as a matter of fact: Viruses and beyond - PMC
    We review the origins of the quasispecies concept and its relevance for RNA virus evolution, viral pathogenesis and antiviral treatment strategies.
  20. [20]
    Resistance to rifampicin: a review | The Journal of Antibiotics - Nature
    Aug 13, 2014 · Areas of study include: resistance mechanisms (primarily acquired resistance because of mutation in the rpoB gene, which encodes the β subunit ...
  21. [21]
    [PDF] Mutations of bacteria from virus sensitivity to virus resistance ...
    In 1943, Salvador E. Luria and. Max Delbrück showed that apparent examples of Lamarckian inheritance were actually due to true genetic mutation, and in 1946.
  22. [22]
    Methods for detecting carcinogens and mutagens with the ... - PubMed
    Methods for detecting carcinogens and mutagens with the Salmonella/mammalian-microsome mutagenicity test. Mutat Res. 1975 Dec;31(6):347-64. doi: 10.1016/0165 ...Missing: paper | Show results with:paper
  23. [23]
  24. [24]
  25. [25]
  26. [26]
    Redefining fundamental concepts of transcription initiation in bacteria
    For example, the λ phage PRE promoter has −10 and −35 elements that differ from the consensus sequence, and it requires cII protein for Eσ binding. Therefore, ...
  27. [27]
    Translation in Prokaryotes - PMC - PubMed Central
    This review summarizes our current understanding of translation in prokaryotes, focusing on the mechanistic and structural aspects of each phase of translation: ...
  28. [28]
  29. [29]
  30. [30]
    50+ years of eukaryotic transcription: an expanding universe of ...
    Aug 22, 2019 · The landmark 1969 discovery of nuclear RNA polymerases I, II and III in diverse eukaryotes represented a major turning point in the field.Missing: polycistronic | Show results with:polycistronic
  31. [31]
  32. [32]
  33. [33]
  34. [34]
  35. [35]
  36. [36]
    Regulation of Translation Initiation in Eukaryotes: Mechanisms and ...
    Coupled release of eukaryotic translation initiation factors 5B and 1A from 80S ribosomes following subunit joining. Mol Cell Biol. 2007;27:2384–2397. doi ...
  37. [37]
  38. [38]
  39. [39]
    Genetic regulatory mechanisms in the synthesis of proteins - PubMed
    Genetic regulatory mechanisms in the synthesis of proteins. J Mol Biol. 1961 Jun:3:318-56. doi: 10.1016/s0022-2836(61)80072-7.
  40. [40]
    Bacterial Sigma Factors and Anti-Sigma Factors: Structure, Function ...
    Jun 26, 2015 · In this review, the structure and function of the σ70 family is considered, along with selected mechanisms for their control by anti-σ factors.
  41. [41]
    The Evolution of Combinatorial Gene Regulation in Fungi
    Our study centers on Mcm1, a transcriptional regulator that, in combination with five cofactors, binds roughly 4% of the genes in Saccharomyces cerevisiae.<|separator|>
  42. [42]
    Chromatin modifications, epigenetics, and how protozoan parasites ...
    Chromatin structure plays a vital role in epigenetic regulation of protozoan parasite gene expression. Epigenetic gene regulation impacts parasite virulence.Missing: microbial | Show results with:microbial
  43. [43]
    Mechanisms of post-transcriptional gene regulation in bacterial ...
    In this review, we discuss post-transcriptional mechanisms that influence the biofilm developmental cycle in a variety of pathogenic bacteria.Abstract · Biofilm Structure: What Does it... · Mechanisms of Post... · Conclusions
  44. [44]
    Small RNAs, Large Networks: Posttranscriptional Regulons in Gram ...
    Sep 15, 2023 · Small regulatory RNA (sRNAs) are key mediators of posttranscriptional gene control in bacteria. Assisted by RNA-binding proteins, a single sRNA ...
  45. [45]
    Processing Endoribonucleases and mRNA Degradation in Bacteria
    Many reviews and papers assume that RNase E is the primary RNase in mRNA degradation. Contrary evidence and alternate mechanisms are ignored.
  46. [46]
    Review RNA Quality Control in Eukaryotes - ScienceDirect.com
    Nov 16, 2007 · Eukaryotic cells contain numerous RNA quality-control systems that are important for shaping the transcriptome of eukaryotic cells.
  47. [47]
    Eukaryotic pre-mRNA processing | RNA splicing (article)
    In this article, we'll take a closer look at the cap, tail, and splicing modifications that eukaryotic RNA transcripts receive, seeing how they're carried out ...
  48. [48]
    RNase E: at the interface of bacterial RNA processing and decay
    Dec 14, 2012 · RNase E is an essential endoribonuclease involved in most aspects of RNA processing and degradation in many bacteria.
  49. [49]
    The emerging world of small silencing RNAs in protozoan parasites
    siRNAs are processed from long dsRNAs by the endonuclease Dicer (Box 1), an RNase III-family enzyme, in a complex with a dsRNA-binding protein (dsRBP). Long ...
  50. [50]
    Small-Molecule-Binding Riboswitches | Microbiology Spectrum
    THE TPP RIBOSWITCH—AN EXAMPLE ACROSS ALL DOMAINS OF LIFE. TPP-binding riboswitches belong to the most widely distributed riboswitch class. They are found in ...
  51. [51]
    Riboswitches: A Common RNA Regulatory Element - Nature
    This schematic is an example of a riboswitch that controls transcription. ... (TPP) riboswitch regulates splicing at the 3' end (Wachter et al. 2007) ...
  52. [52]
    CRISPR-Cas systems: new players in gene regulation and bacterial ...
    Apr 4, 2014 · CRISPR-Cas systems are bacterial defenses against foreign nucleic acids derived from bacteriophages, plasmids or other sources.
  53. [53]
    Bacterial Iron Homeostasis Regulation by sRNAs - ASM Journals
    In this review, we will summarize knowledge on how sRNAs control iron homeostasis mainly through studies on RyhB in Escherichia coli.
  54. [54]
    Roles of two RyhB paralogs in the physiology of Salmonella enterica
    In the previous decade RyhB sRNA, which is involved in iron homeostasis, was intensively characterized in an E. coli strain (Masse and Gottesman, 2002 · An E.
  55. [55]
    Circular permutation of a synthetic eukaryotic chromosome ... - PNAS
    Nov 5, 2014 · The S. cerevisiae genome is composed of 12 Mb of DNA organized as 16 linear chromosomes ranging in size from 230 kb to more than 2 Mb (1).
  56. [56]
    The Reference Genome Sequence of Saccharomyces cerevisiae
    The 16 chromosomes of Saccharomyces cerevisiae comprise the first completely finished eukaryotic genome and were sequenced in the early 1990s by an ...Missing: linear | Show results with:linear
  57. [57]
    Mating-type Gene Switching in Saccharomyces cerevisiae
    The MAT locus lies in the middle of the right arm of chromosome 3, ∼100 kb from both the centromere and the telomere. The two mating-type alleles, MATα and MATa ...
  58. [58]
    Telomere Roles in Fungal Genome Evolution and Adaptation - PMC
    Aug 9, 2021 · Telomeres form the ends of linear chromosomes and usually comprise protein complexes that bind to simple repeated sequence motifs that are ...
  59. [59]
    Centromere-driven genomic innovations in fungal pathogens - NIH
    Mar 28, 2024 · Here we highlight centromere-associated genomic innovations that often remain unnoticed but facilitate landmark events such as fungal species diversification.
  60. [60]
    Genome sequence of the human malaria parasite Plasmodium ...
    Oct 3, 2002 · Here we report an analysis of the genome sequence of P. falciparum clone 3D7. The 23-megabase nuclear genome consists of 14 chromosomes, encodes ...
  61. [61]
    Antigenic diversity is generated by distinct evolutionary mechanisms ...
    Antigenic variation enables pathogens to avoid the host immune response by continual switching of surface proteins. The protozoan blood parasite Trypanosoma ...
  62. [62]
    Repetitive Elements in Genomes of Parasitic Protozoa - PMC
    In this review, we focus on repetitive elements in the genomes of five pathogenic protozoa with genome sequencing projects that are either complete or nearing ...Missing: fungal | Show results with:fungal
  63. [63]
    Mitochondrial and plastid genome architecture: Reoccurring themes ...
    Mitochondrial and plastid genomes show a wide array of architectures, varying immensely in size, structure, and content.
  64. [64]
    Structure and Classification of Viruses - Medical Microbiology - NCBI
    Single-stranded linear DNA, 4–6 kb in size, is found with the members of the Parvovirus family that comprises the parvo-, the erythro- and the dependoviruses.Missing: rotavirus | Show results with:rotavirus
  65. [65]
    Mutation Rate - an overview | ScienceDirect Topics
    Studies done for animal viruses in vitro or in cell culture have estimated rates of 10−3 to 10−5 substitutions per nucleotide per round of replication. ... That ...
  66. [66]
    Properties and abundance of overlapping genes in viruses - PMC
    Feb 13, 2020 · Segmented and non-segmented viruses had different proportions of gene overlap presence within most genome types.Missing: influenza lambda
  67. [67]
    THE BIOLOGY OF INFLUENZA VIRUSES - PMC - NIH
    The influenza viruses are characterized by segmented, negative-strand RNA genomes requiring an RNA-dependent RNA polymerase of viral origin for replication.Missing: lambda phage<|separator|>
  68. [68]
    Bacteriophage Lambda Site-Specific Recombination - PMC
    Bacteriophage λ uses site-specific recombination to integrate and excise the vial chromosome into and out of the bacterial chromosome.Missing: provirus | Show results with:provirus
  69. [69]
    Bacteriophage T4 Genome | Microbiology and Molecular Biology ...
    All of these functions serve to enhance phage promoter recognition and transcription; no DNA-binding transcriptional repressor protein has been identified in ...
  70. [70]
    Viral Quasispecies Evolution - PMC - PubMed Central - NIH
    Viral quasispecies evolution refers to RNA viral populations consisting of mutant spectra, not all with the same nucleotide sequence.<|control11|><|separator|>
  71. [71]
    Transcriptomic profiling of Escherichia coli K-12 in response to a ...
    May 24, 2022 · It is a model organism for biological research due to its non-pathogenic properties, easy handling, and a wide nutritional palate. Non ...
  72. [72]
    Diverse Mechanisms Regulate Sporulation Sigma Factor Activity in ...
    In Bacillus subtilis, the sporulation-specific sigma factors, σ F , σ E , σ G , and σ K , activate compartment-specific transcriptional programs that drive ...Missing: paper | Show results with:paper
  73. [73]
    Genome-Wide Analysis of the Stationary-Phase Sigma Factor ...
    The sporulation program of gene expression in B. subtilis is carried out under the direction of five alternative sigma factors whose activities are subject to ...
  74. [74]
    Quorum-Sensing Genes in Pseudomonas aeruginosa Biofilms - NIH
    P. aeruginosa utilizes QS to regulate expression of a number of virulence genes and gene products in a density-dependent fashion, presumably to ensure that ...
  75. [75]
    The Salmonella enterica Serotype Typhi Vi Capsular Antigen Is ...
    We report that tviB, a gene necessary for Vi production in S. Typhi, was significantly upregulated during invasion of intestinal epithelial cells in vitro.
  76. [76]
    An Overview of Genetic Information of Latent Mycobacterium ... - NIH
    This study aimed to briefly examine the genes involved in the latent state as well as the changes that are caused by Mycobacterium tuberculosis.
  77. [77]
    Bacteriophage Lambda Charon Vectors for DNA Cloningt
    Twenty hybrid lambda phages especially designed for molecular cloning have been constructed and named Charon phages. These phages differ in the ranges of.
  78. [78]
    The development and applications of the bacterial artificial ...
    The development of the Bacterial Artificial Chromosome (BAC) system was driven in part by the Human Genome Project as a means to construct genomic DNA ...<|separator|>
  79. [79]
    Peptidoglycan: Structure, Synthesis, and Regulation | EcoSal Plus
    Gram-positive bacteria possess a thick multilayered peptidoglycan that is exposed to the cell exterior with covalently bound glycopolymers—teichoic acids— ...<|control11|><|separator|>
  80. [80]
    Recent Advances in Peptidoglycan Synthesis and Regulation in ...
    In this review, we highlight recent progress in our understanding of peptidoglycan synthesis, remodeling, repair, and regulation in two model bacteria: the Gram ...
  81. [81]
    Genome sequence of a model prokaryote: Current Biology - Cell Press
    The complete genome sequence of Escherichia coli K-12​​ , which used a deliberately high cut-off, but they still indicate that functional annotation of distantly ...
  82. [82]
    The evolution of TBP in archaea and their eukaryotic offspring - NIH
    A fundamental difference between bacterial and eukaryotic/archaeal transcription initiation lies in the way the RNA polymerase (RNAP) is recruited to the ...Missing: similarity review paper
  83. [83]
    Complete Genome Sequence of the Genetically Tractable ...
    The genome sequence of the genetically tractable, mesophilic, hydrogenotrophic methanogen Methanococcus maripaludis contains 1722 protein-coding genes in a ...Missing: review | Show results with:review
  84. [84]
    Genome sequence of Halobacterium species NRC-1 - PNAS
    The sequence of Halobacterium NRC-1 has revealed 3 large replicons, a large chromosome and 2 novel minichromosomes, and 2,682 putative genes, including 972 ...
  85. [85]
    Transcription Regulation in Archaea | Journal of Bacteriology
    Jun 27, 2016 · All three of the aforementioned transcription factors have close eukaryotic homologs: archaeal TBPs are nearly identical to eukaryotic TBPs (45); ...Basal Transcription Factors · Histone-Based Regulation Of... · Nucleosome Occupancy At The...
  86. [86]
    Biosynthesis of archaeal membrane ether lipids - PMC
    This review describes the current knowledge of the biosynthetic pathway of archaeal ether lipids; insights on the stability and robustness of archaeal lipid ...
  87. [87]
    Minimal Yet Powerful: The Role of Archaeal Small Heat Shock ...
    sHsps play a crucial role in preventing protein aggregation and holding unfolded protein substrates in a folding-competent form.Introduction · Different Functions of Small... · Regulation of Archaeal Small...
  88. [88]
    Dynamic properties of the Sulfolobus CRISPR/Cas and ... - NIH
    The adaptive immune CRISPR/Cas and CRISPR/Cmr systems of the crenarchaeal thermoacidophile Sulfolobus were challenged by a variety of viral and plasmid genes, ...
  89. [89]
    Archaeal Extrachromosomal Genetic Elements - PMC
    Plasmids occur widely in archaea and are most common in haloarchaea. Among 15 haloarchaea for which complete genomes have been determined, only one lacks a ...
  90. [90]
    Nanoarchaea: representatives of a novel archaeal phylum or a fast ...
    Apr 14, 2005 · Cultivable archaeal species are assigned to two phyla - the Crenarchaeota and the Euryarchaeota - by a number of important genetic differences,
  91. [91]
    Comparative genomics reveals the origin of fungal hyphae and ...
    Sep 9, 2019 · Hypha morphogenesis gene families, in general, show more change in domain composition between unicellular and filamentous fungi than do randomly ...
  92. [92]
    Genetic complementation in apicomplexan parasites - PNAS
    A robust forward genetic model for Apicomplexa could greatly enhance functional analysis of genes in these important protozoan pathogens.
  93. [93]
    Accurate prediction of secondary metabolite gene clusters in ... - PNAS
    Dec 17, 2012 · In this study, we design and apply a DNA expression array for Aspergillus nidulans in combination with legacy data to form a comprehensive gene ...
  94. [94]
    Aspergillus Secondary Metabolite Database, a resource to ... - Nature
    Aug 4, 2017 · A2MDB provides an easy access to unbiased, comprehensive information about Aspergillosis, Aspergillus species, their secondary metabolites and cellular targets.
  95. [95]
    Linking secondary metabolites to gene clusters through genome ...
    Jan 9, 2018 · Our study delivers six fungal genomes, showing the large diversity found in the Aspergillus genus; highlights the potential for discovery of beneficial or ...
  96. [96]
    Variant surface glycoprotein density defines an immune evasion ...
    Oct 10, 2017 · Following a genetic VSG switch, trypanosomes must replace their entire VSG coat. During this period, trypanosomes simultaneously display both ...
  97. [97]
    African trypanosomes expressing multiple VSGs are rapidly ... - PNAS
    Sep 25, 2019 · Trypanosoma brucei parasites successfully evade the host immune system by periodically switching the dense coat of variant surface glycoprotein ...
  98. [98]
    Evolution of the MAT locus and its Ho endonuclease in yeast species
    The mating type of a haploid cell is determined by its genotype at the mating-type (MAT) locus on chromosome III. The two variants of the MAT locus, MATα and ...
  99. [99]
    RAD51-mediated R-loop formation acts to repair transcription ...
    RAD51-directed recombination of silent variant surface glycoprotein (VSG) genes allows Trypanosoma brucei to evade host immunity. How such VSG switching is ...Sign Up For Pnas Alerts · Results · T. Brucei Rad51 Binds...<|separator|>
  100. [100]
    A novel genetic system to detect protein–protein interactions - Nature
    Jul 20, 1989 · We have generated a novel genetic system to study these interactions by taking advantage of the properties of the GAL4 protein of the yeast ...
  101. [101]
  102. [102]
    Introduction to the Long-Term Evolution Experiment (LTEE)
    The LTEE has 12 populations of E. coli, all started in 1988 from the same non-pathogenic ancestral strain. (While some strains of E. coli can cause disease, ...
  103. [103]
    Experimental evolution and the dynamics of adaptation and genome ...
    May 16, 2017 · The long-term evolution experiment, or LTEE, is simple both conceptually and practically. Twelve populations were started the same ancestral ...
  104. [104]
    Neutral Theory, Microbial Practice: Challenges in Bacterial ...
    Apr 19, 2018 · “The neutral mutation-random drift hypothesis (or the neutral theory for short) holds that at the molecular level most evolutionary change and ...
  105. [105]
    Duplicated antibiotic resistance genes reveal ongoing selection and ...
    Feb 16, 2024 · Our findings indicate that duplicated genes often encode functions undergoing positive selection and horizontal gene transfer in microbial communities.
  106. [106]
    Evolutionary Pathways and Trajectories in Antibiotic Resistance
    Jun 30, 2021 · Gene amplification (gene duplication in its simplest version) is likely relevant in the adaptation to antibiotic exposure because it generates ...
  107. [107]
    The last universal common ancestor between ancient Earth ...
    Aug 16, 2018 · What one finds is a collection of about 30 genes, mostly for ribosomal proteins, telling us that LUCA had a ribosome and had the genetic code, ...
  108. [108]
    The nature of the last universal common ancestor and its impact on ...
    Jul 12, 2024 · The last universal common ancestor (LUCA) is the node on the tree of life from which the fundamental prokaryotic domains (Archaea and Bacteria) diverge.
  109. [109]
    Influenza Virus: Dealing with a Drifting and Shifting Pathogen
    Mar 1, 2018 · Antigenic drift is the process by which minor changes are introduced into key viral epitopes through point mutations in the viral genome (5).<|control11|><|separator|>
  110. [110]
    The Nobel Prize in Physiology or Medicine 1978 - Press release
    Werner Arber, Dan Nathans and Hamilton Smith​​ for the discovery of “restriction enzymes and their application to problems of molecular genetics”.
  111. [111]
    Construction and characterization of new cloning vehicles. II. A ...
    1977;2(2):95-113. Authors. F Bolivar, R L Rodriguez, P J Greene, M C ... The vector pBR322 was constructed in order to have a plasmid with a single ...Missing: citation | Show results with:citation
  112. [112]
    Use of bacteriophage T7 RNA polymerase to direct selective high ...
    A gene expression system based on bacteriophage T7 RNA polymerase has been developed. T7 RNA polymerase is highly selective for its own promoters.
  113. [113]
    Glycosylation engineering in yeast: the advent of fully humanized ...
    Major advances in the glycoengineering of the yeast Pichia pastoris, have culminated in the production of fully humanized sialylated glycoproteins.
  114. [114]
    A Programmable Dual-RNA–Guided DNA Endonuclease ... - Science
    Jun 28, 2012 · Our study reveals a family of endonucleases that use dual-RNAs for site-specific DNA cleavage and highlights the potential to exploit the system for RNA- ...
  115. [115]
    How restriction enzymes became the workhorses of molecular biology
    Arber had provided the theoretical framework that described the biology of restriction and modification and had successfully isolated the very first type I ...
  116. [116]
    Creation of a Bacterial Cell Controlled by a Chemically Synthesized ...
    May 20, 2010 · We report the design, synthesis, and assembly of the 1.08–mega–base pair Mycoplasma mycoides JCVI-syn1.0 genome starting from digitized genome sequence ...
  117. [117]
    Yersinia Type III Secretion System Master Regulator LcrF - PMC - NIH
    Dec 7, 2015 · Human-pathogenic Yersinia species share a virulence plasmid, called pCD1 in Y. pestis and pYV in enteropathogenic yersiniae, encoding the Ysc ...
  118. [118]
    Targeting Type III Secretion in Yersinia pestis - PMC
    T3SS is absolutely required for the virulence of Y. pestis, making it a potential target for new therapeutics. Using a novel and simple high-throughput ...
  119. [119]
    Multi-locus sequence typing: a tool for global epidemiology - PubMed
    Multi-locus sequence typing (MLST) was proposed as a nucleotide sequence-based approach that could be applied to many bacterial pathogens.
  120. [120]
    Whole genome multilocus sequence typing as an epidemiologic tool ...
    Our findings indicate wgMLST is a simplified, sensitive, and scalable tool for epidemiologic analysis of Y. pestis strains.
  121. [121]
    16S rRNA Gene Sequencing for Bacterial Identification in the ... - NIH
    16S rRNA sequencing helps identify bacteria, especially those with ambiguous profiles, but it is not foolproof and has limitations in species identification.
  122. [122]
    Whole-Genome Sequencing Accurately Identifies Resistance to ...
    A whole-genome sequencing approach predicted phenotypic resistance to extended spectrum β-lactams for 4 leading causes of gram-negative bacteremia in ...Whole-Genome Sequencing And... · Comparison Of Bmd With Data... · Discussion
  123. [123]
    Diagnostic Yield and Impact on Antimicrobial Management of 16S ...
    Nov 14, 2022 · 16S rRNA gene sequencing is increasingly used in clinical practice for bacterial identification of clinical specimens.
  124. [124]
    Phage Therapy: From Biologic Mechanisms to Future Directions - PMC
    This review provides a comprehensive view on the state of the art in phage therapy, covering biologic mechanisms, clinical applications, remaining challenges, ...
  125. [125]
    The HPV Vaccine Story - PMC - NIH
    We put together a plan to make an infectious papillomavirus in the laboratory, using the mammalian cell gene expression techniques that had recently become ...Missing: microbial | Show results with:microbial
  126. [126]
    Second-Generation Prophylactic HPV Vaccines: Successes and ...
    This review will discuss efforts to develop second generation HPV vaccines that will provide broader protection against the HPV types associated with cancer.Missing: microbial | Show results with:microbial
  127. [127]
    mecA Gene Is Widely Disseminated in Staphylococcus aureus ... - NIH
    High-level resistance to methicillin is caused by the mecA gene, which encodes an alternative penicillin-binding protein, PBP 2a.
  128. [128]
    Methicillin-Resistant Staphylococcus aureus (MRSA)
    Sep 1, 2021 · mecA in the S. aureus is a marker of MRSA. The main objective of this study was to detect mecA and vanA genes conferring resistance in S. aureus ...
  129. [129]
    Mechanisms of Methicillin Resistance in Staphylococcus aureus
    Resistance is usually conferred by the acquisition of a nonnative gene encoding a penicillin-binding protein (PBP2a), with significantly lower affinity for β- ...
  130. [130]
    Novel coronavirus complete genome from the Wuhan outbreak now ...
    Jan 13, 2020 · The GenBank record of Wuhan-Hu-1 includes sequence data, annotation and metadata from this virus isolated approximately two weeks ago from a ...
  131. [131]
    SARS-CoV-2 variants evolved during the early stage of the ...
    In this review, we summarize the mutations in SARS-CoV-2 during the early phase of virus evolution and discuss the significance of the gene alterations in ...
  132. [132]
    Molecular biological access to the chemistry of unknown soil microbes
    The concept of cloning the metagenome to access the collective genomes and the biosynthetic machinery of soil microflora is explored here.Missing: rhizosphere analysis
  133. [133]
    Cultivation of unculturable soil bacteria: Trends in Biotechnology
    Despite the abundance of bacterial species in soil, more than 99% of these species cannot be cultured by traditional techniques. In addition, the less than ...
  134. [134]
    Complete genome sequence of Pseudomonas sp. PP3, a ... - NIH
    Apr 21, 2025 · This well-characterized organism continues to provide key insights into adaptive dehalogenase-mediated bioremediation of halogenated organic ...
  135. [135]
    A Pseudomonas putida Strain Genetically Engineered for 1,2,3 ...
    The results demonstrate the successful use of a laboratory-evolved dehalogenase and genetic engineering to produce an effective, plasmid-free, and stable whole- ...
  136. [136]
    Hydrocarbon-degrading bacteria enriched by the Deepwater ...
    Jun 20, 2013 · We identified several aliphatic (Alcanivorax, Marinobacter)- and polycyclic aromatic hydrocarbon (Alteromonas, Cycloclasticus, Colwellia)-degrading bacteria.
  137. [137]
    Nitrogen fixation (nif) genes and large plasmids of Rhizobium ... - NIH
    The location of structural nitrogen-fixation genes was determined for the slow- and fast-growing types of Rhizobium japonicum.
  138. [138]
    Methanogenesis - ScienceDirect.com
    Jul 9, 2018 · First, all methanogens are archaea, and there is little evidence for horizontal gene transfer of this metabolism. Presumably, the large number ...
  139. [139]
    Expanding the phylogenetic distribution of cytochrome b-containing ...
    Jul 9, 2022 · Biological methanogenesis by methanogenic archaea (methanogens) accounts for ~74% of global methane emissions [2]. For many years it was assumed ...
  140. [140]
    Antibiotic resistance in the environment | Nature Reviews Microbiology
    Nov 4, 2021 · Antibiotic resistance is a global health challenge, involving the transfer of bacteria and genes between humans, animals and the environment.
  141. [141]
    RNA Stable Isotope Probing, a Novel Means of Linking Microbial ...
    SIP provides access to the relationship between environmental functions and the specific microbial community members pivotal to function performance (12). SIP ...
  142. [142]
    Advances and perspectives of using stable isotope probing (SIP)
    Sep 1, 2023 · Stable isotope probing (SIP) is a powerful tool to study microbial community structure and function in both nature and engineered ...