Fact-checked by Grok 2 weeks ago

RNA virus

An RNA virus is a virus that has ribonucleic acid (RNA) as its genetic material and is capable of infecting a wide range of hosts, including humans, animals, , and , by replicating inside host cells using virally encoded enzymes. These viruses are distinguished from DNA viruses by their RNA genomes, which can be single-stranded (ssRNA) or double-stranded (dsRNA), linear or segmented, and range in size from approximately 3 to more than 40 kilobases. Unlike cellular organisms, RNA viruses lack the machinery for independent replication and must hijack host cellular processes, relying on an (RdRp) to synthesize their genomes and messenger RNAs. RNA viruses are classified primarily using the Baltimore classification system, which groups them based on genome type and mRNA synthesis strategy into four main categories: Group III (dsRNA viruses, such as reoviruses), Group IV (positive-sense ssRNA viruses, including picornaviruses and coronaviruses), Group V (negative-sense ssRNA viruses, like orthomyxoviruses and rhabdoviruses), and Group VI (positive-sense ssRNA viruses with reverse transcriptase, such as retroviruses). This system, established in 1971, highlights the diversity in replication mechanisms, with positive-sense RNA directly serving as mRNA for translation, while negative-sense RNA requires initial transcription into positive-sense intermediates. The International Committee on Taxonomy of Viruses (ICTV) further organizes RNA viruses into numerous families, such as Coronaviridae, Picornaviridae, Retroviridae, and Orthomyxoviridae, reflecting their morphological, genomic, and ecological differences. A defining feature of RNA viruses is their exceptionally high mutation rates, typically 10^{-3} to 10^{-5} errors per per replication cycle, driven by the error-prone nature of RdRp lacking activity. This genetic variability enables rapid evolution, quasispecies formation, and adaptation to new hosts or immune pressures, contributing to challenges in development and antiviral therapies. viruses account for a significant portion of emerging infectious diseases, with examples including viruses (), human immunodeficiency virus (Retroviridae), (), Ebola virus (), and severe acute respiratory syndrome coronavirus 2 (). Their ability to cause pandemics, such as the 1918 outbreak and the , underscores their public health impact, while also making them valuable models for studying and host-pathogen interactions.

Overview

Definition and Historical Discovery

RNA viruses are a diverse group of viruses that utilize ribonucleic acid (RNA) as their genetic material, in contrast to DNA viruses which employ deoxyribonucleic acid (DNA). Their genomes can be single-stranded (ssRNA) or double-stranded (dsRNA), and may exist as linear molecules or in segmented forms consisting of multiple RNA segments. Unlike most cellular organisms, RNA viruses lack the machinery for DNA synthesis and instead rely on RNA-dependent RNA polymerases encoded by their own genomes to replicate. This fundamental distinction places RNA viruses within the realm Riboviria in modern taxonomy, encompassing a wide array of pathogens that infect humans, animals, plants, and other organisms. The historical discovery of RNA viruses began in the late with investigations into plant diseases. In 1892, Russian microbiologist Dmitri Ivanovsky demonstrated that the causative agent of could pass through filters that retained bacteria, suggesting a submicroscopic infectious entity; this agent was later identified as the (TMV), the first recognized and a single-stranded virus. Building on this, Dutch microbiologist in 1898 coined the term "" to describe the filterable, non-bacterial pathogen of TMV, characterizing it as a "contagium vivum fluidum" (living infectious fluid) that multiplied only in living cells. These foundational observations marked the birth of , though the nature of the TMV was not yet known. Advancements in the confirmed the particulate nature and genetic composition of viruses. In the 1930s, the invention of the enabled the first visualizations of virus particles, including TMV in 1939 by and colleagues, providing direct evidence of their submicroscopic structure and distinguishing them from fluid-like agents. By the 1950s, biochemical analyses revealed the basis of animal virus genomes; notably, in 1955, researchers Frederick Schaffer and Carlton Schwerdt identified the genome of —a major —as single-stranded , solidifying RNA viruses as a distinct class. These discoveries shifted toward molecular understanding, highlighting RNA's role in . Early efforts to classify viruses laid the groundwork for systematic . In , André Lwoff, Robert Horne, and Paul Tournier proposed a foundational emphasizing viral properties like type, symmetry, and replication site, which first delineated viruses as a category. This was refined in 1971 by , who introduced a genome-based classification scheme grouping viruses by their type and mRNA synthesis method; viruses were divided into classes such as positive-sense ssRNA (e.g., TMV), negative-sense ssRNA, dsRNA, and reverse-transcribing viruses, providing a enduring framework for understanding viral diversity. Baltimore's remains central to virus today.

Biological and Medical Significance

RNA viruses play a pivotal role in due to their exceptionally high rates and genetic variability, which arise from error-prone RNA-dependent RNA polymerases, making them ideal models for studying rapid , , and in real time. This variability enables RNA viruses to serve as experimental systems for investigating fundamental processes like , recombination, and host-virus co-, providing insights into broader microbial and eukaryotic . Additionally, their diverse strategies for replication and , such as and ribosomal frameshifting, position RNA viruses as key models for understanding RNA processing, , and unconventional mechanisms of protein synthesis in cellular contexts. Medically, RNA viruses are major pathogens responsible for both acute respiratory illnesses and chronic infections, exemplified by viruses causing seasonal epidemics, human immunodeficiency virus () leading to acquired immunodeficiency syndrome (AIDS), and severe acute respiratory syndrome coronavirus 2 () triggering the . These viruses account for significant global morbidity and mortality, with alone resulting in hundreds of thousands of deaths annually and infecting 40.8 million [37.0 million–45.6 million] people worldwide as of 2024. Advances in countermeasures include vaccines like inactivated shots and live-attenuated options, as well as highly active antiretroviral therapy (HAART) for ; the crisis accelerated development, with platforms like those for demonstrating rapid deployment and high efficacy in eliciting immune responses. Antivirals such as for neuraminidase inhibition and for RNA polymerase blockade further highlight targeted therapies against RNA virus replication. The rapid evolutionary potential of RNA viruses poses substantial challenges, facilitating antigenic drift and shift that drive recurrent outbreaks and pandemics, as seen with influenza's ability to evade immunity through surface protein mutations. The 1918 H1N1 , caused by an RNA virus, resulted in an estimated 50 million deaths worldwide, underscoring their potential for catastrophic impact. Similarly, the , ongoing into 2025, has caused over 7 million confirmed deaths and economic losses exceeding $13 trillion globally, disrupting health systems and temporarily reversing a decade of progress in global while affecting billions through direct infections, lockdowns, and socioeconomic fallout. Beyond threats, RNA viruses have transformative research applications; the discovery of RNA interference (RNAi) stemmed from studies of plant virus infections, where double-stranded RNA from viruses triggers silencing of homologous sequences, revealing a conserved antiviral defense mechanism later harnessed for gene knockdown tools. In oncology, oncolytic RNA viruses like reoviruses and vesicular stomatitis virus are engineered for virotherapy, selectively replicating in and lysing tumor cells while stimulating antitumor immunity, with clinical trials showing promise for cancers such as melanoma and ovarian tumors.

Genomic Features

Types of RNA Genomes

RNA virus genomes exhibit structural diversity primarily in terms of strandedness and segmentation, which influence their packaging, transmission, and evolutionary dynamics. The majority of RNA viruses contain single-stranded (ssRNA) genomes, which can be either linear or, less commonly, circular. ssRNA genomes predominate across numerous viral families, such as Picornaviridae and for positive-sense linear forms, and represent the most common configuration among RNA viruses. In contrast, a notable exception is the hepatitis delta virus (HDV), which possesses a circular, negative-sense ssRNA genome of approximately 1.7 , enabling unique replication strategies akin to viroids. Double-stranded RNA (dsRNA) genomes are rarer among RNA viruses, occurring mainly in families like Reoviridae and Birnaviridae. These dsRNA genomes are typically linear and consist of multiple discrete segments, with Reoviridae viruses featuring 10 to 12 segments that encode structural and non-structural proteins. The segmented nature of dsRNA genomes facilitates selective packaging, where each segment is independently transcribed and packaged into virions. Segmentation is a key feature in both ssRNA and dsRNA viruses, with genomes divided into 1 to 12 independent RNA segments, though unsegmented (monopartite) forms are more prevalent in many ssRNA families. For instance, influenza viruses () have eight negative-sense ssRNA segments, allowing for efficient virion assembly and genetic exchange. This segmentation provides advantages in genome packaging by permitting the incorporation of specific segments via dedicated signals and enables reassortment during co-infection, which generates novel viral variants and contributes to rapid evolution. Overall, RNA virus genomes range in size from about 1.7 to over 40 kb, considerably smaller than those of many DNA viruses, which can exceed hundreds of kilobases. This size constraint arises from the error-prone nature of RNA-dependent RNA polymerases, which lack 3'–5' exonuclease proofreading activity, resulting in high mutation rates that limit genome expansion to prevent mutational meltdown.

Sense and Antisense Mechanisms

RNA viruses exhibit diverse genome polarities that determine their initial interactions with cells, primarily categorized as positive-sense single-stranded RNA (+ssRNA), negative-sense single-stranded RNA (-ssRNA), and ambisense RNA. In positive-sense ssRNA viruses, the genomic has the same polarity and nucleotide sequence as (mRNA), allowing it to be directly recognized and translated by the host cell's ribosomes upon entry. This immediate translation produces viral proteins, including the (RdRp) necessary for subsequent genome replication, where the +ssRNA serves as a template to synthesize complementary negative-sense RNA intermediates. Representative examples include viruses from the Picornaviridae family, such as , and , like . Negative-sense ssRNA viruses possess genomic RNA that is complementary to mRNA, rendering it non-translatable by host machinery and necessitating the prior synthesis of positive-sense mRNAs for . To initiate , these viruses must package the RdRp within the virion, which uses the -ssRNA as a template to transcribe mRNAs directly in the shortly after entry. Examples include orthomyxoviruses, such as , and rhabdoviruses, like . Ambisense RNA viruses employ a hybrid strategy, where individual genome segments contain coding regions of both positive and negative polarities, requiring transcription from both the genomic RNA and its complementary strand to express all genes. This mixed approach is exemplified by arenaviruses, such as Lassa virus, in the family Arenaviridae, and certain viruses in the order like virus (genus , family Phenuiviridae), where the small (S) segment typically codes for in the negative sense and precursor in the positive sense. Like -ssRNA viruses, ambisense viruses package RdRp in the virion to enable initial transcription. These polarity mechanisms profoundly influence host-virus interactions: +ssRNA viruses benefit from a simpler entry process, leveraging the host's apparatus without needing pre-packaged enzymes, which facilitates rapid initial protein synthesis. In contrast, -ssRNA and ambisense viruses require virion-encapsidated RdRp for the first round of transcription, adding complexity to their lifecycle but enabling tight control over through nucleoprotein associations.

Virion Structure

Capsid and Genome Packaging

The of RNA viruses is a protein shell composed of self-assembling structural proteins known as capsomeres, which protect the enclosed and facilitate transmission between host cells. These capsids typically exhibit one of two primary symmetries: icosahedral, formed by 60 or multiples of 60 identical protein subunits arranged in a roughly spherical structure, or helical, where protein subunits and the wind around each other in a rod-like configuration. Icosahedral capsids are prevalent in many non-enveloped RNA viruses, providing efficient enclosure with minimal protein usage, while helical capsids accommodate elongated genomes by coiling the RNA along the protein lattice. Genome packaging within the capsid involves the selective incorporation of the viral , often guided by specific motifs on the RNA molecule that interact with capsid proteins. In positive-sense single-stranded (+ssRNA) viruses, the RNA is typically coiled or folded inside the capsid, with packaging signals such as 5' cap structures or internal entry sites (IRES) in the untranslated regions facilitating specific binding and stabilization. These motifs ensure that only the full-length genomic RNA is encapsidated, excluding host or subgenomic RNAs, through electrostatic interactions between positively charged protein regions and the negatively charged RNA backbone. The process is highly specific, with capsid proteins often multimerizing around the RNA to drive spontaneous assembly. Icosahedral capsids typically have diameters of 20 to 100 , while helical capsids feature diameters of about 10 to 20 and lengths ranging from 100 to over 1,000 , depending on the length. Stability is conferred by inter-subunit bonds and the packaged itself, which can reinforce the structure; for instance, some non-enveloped virus capsids, such as those of enteroviruses, remain intact in acidic environments (pH 3-5), aiding survival in the . Non-enveloped viruses feature bare s exposed to the , enhancing to detergents and , whereas enveloped viruses surround the with a derived from the host, providing additional protection but reducing overall stability.

Envelope and Surface Proteins

Many RNA viruses possess a lipid envelope surrounding their capsid and genome, which is derived from modified portions of the host cell's plasma membrane during the viral budding process. This envelope consists of a phospholipid bilayer incorporating host lipids and is studded with virus-encoded glycoproteins that project outward as spikes or peplomers. In contrast, non-enveloped RNA viruses, such as caliciviruses, lack this lipid layer and rely solely on their protein capsid for external protection. The surface embedded in the are critical for mediating interactions with cells, primarily through receptor and activities. These proteins are typically heavily glycosylated, with carbohydrate moieties aiding in proper folding, stability, and shielding from immune responses. For instance, in viruses, the (HA) forms trimeric spikes that bind to residues on cell surfaces, facilitating viral attachment. Similarly, the (GP) of virus serves dual roles in receptor engagement and low-pH-induced of the viral and endosomal membranes. In flaviviruses, such as , E performs both attachment and functions, highlighting the multifunctional nature of these surface proteins across enveloped RNA virus families. Glycosylation on these surface glycoproteins plays a key role in immune evasion by forming a glycan shield that masks underlying epitopes, reducing recognition by antibodies and reducing . This dense layer, often comprising up to 50% of the glycoprotein's mass in viruses like , sterically hinders access to conserved neutralization sites. Such modifications enable enveloped viruses to persist in host environments despite immune pressures, as observed in the surface proteins of coronaviruses and paramyxoviruses. While the provides an external barrier, it works in concert with the underlying to enclose and protect the viral .

Replication Cycle

Entry and Uncoating

The entry and uncoating of RNA viruses represent the initial critical phases of , enabling the viral genome to access the host cell's for subsequent replication. These processes involve specific interactions between viral surface proteins and host cell receptors, followed by internalization via or direct membrane fusion, and culminate in the disassembly of the virion to liberate the RNA genome. Most RNA viruses target the for replication, necessitating efficient uncoating mechanisms that respond to cellular cues such as pH changes or proteolytic processing. Attachment begins with the binding of viral envelope glycoproteins or capsid proteins to host cell surface receptors, a highly specific interaction that determines tissue tropism. For instance, influenza A viruses use hemagglutinin (HA) to bind sialic acid residues on glycoproteins or glycolipids, initiating receptor-mediated uptake. Similarly, SARS-CoV-2 employs its spike (S) protein to engage the angiotensin-converting enzyme 2 (ACE2) receptor on respiratory epithelial cells, often with assistance from co-receptors like neuropilin-1. In non-enveloped RNA viruses, such as picornaviruses (e.g., poliovirus or rhinovirus), capsid proteins like VP1 interact with receptors including CD155 or intercellular adhesion molecule 1 (ICAM-1), respectively. These binding events not only anchor the virus but also trigger signaling cascades that promote cytoskeletal rearrangements for efficient internalization. Entry into the host cell typically occurs through endocytosis, with clathrin-mediated endocytosis being a common route for many RNA viruses, though macropinocytosis or caveolar pathways are also utilized depending on the virus and cell type. For pH-dependent enveloped RNA viruses like influenza, the virion is trafficked to late endosomes where the acidic environment (pH ~5.0–6.0) induces conformational changes in HA, driving fusion of the viral envelope with the endosomal membrane and release of the viral ribonucleoprotein (vRNP) complex into the cytosol. Coronaviruses such as SARS-CoV-2 predominantly enter via endocytosis as well, but fusion can occur at the plasma membrane if the S protein is cleaved by surface proteases like TMPRSS2; otherwise, endosomal cathepsins (e.g., cathepsin L) process the protein in a pH-sensitive manner to trigger fusion. Non-enveloped RNA viruses, lacking an envelope for fusion, rely on endosomal acidification to initiate entry without membrane merger. Uncoating follows entry and involves the disassembly of the or complex to free the . In enveloped RNA viruses, during entry effectively uncoats the by delivering it directly to the , often with additional cues like ubiquitination or dynein-mediated transport aiding vRNP dissociation; for , low endosomal and high concentrations further destabilize the protein shell surrounding the vRNP. For non-enveloped picornaviruses, uncoating is triggered by low in late endosomes, causing irreversible expansion of the into an "A-particle" conformation; this exposes the of , which inserts into the endosomal to form a ~10 pore at the icosahedral two-fold symmetry axis, facilitating externalization and translocation of the positive-sense through the pore into the . These mechanisms ensure rapid release while minimizing exposure to host nucleases, highlighting the evolutionary adaptations of viruses to exploit host vesicular trafficking.

Genome Replication and Transcription

The (RdRp), also known as the RNA replicase, serves as the core for genome replication and transcription in RNA viruses of Baltimore Groups III, IV, and V, catalyzing the synthesis of RNA from an RNA template without requiring DNA intermediates. Unlike DNA polymerases, RdRp lacks 3' to 5' proofreading activity, rendering the replication process inherently error-prone and contributing to the high rates observed in RNA viruses. This is typically encoded within the viral and forms part of a replication-transcription complex (RTC) that associates with host cell membranes or cytoplasmic structures to shield viral RNA synthesis from innate immune detection. In positive-sense single-stranded RNA (+ssRNA) viruses, such as picornaviruses and flaviviruses, the genomic functions directly as mRNA upon entry into the host cell, undergoing to produce viral proteins including RdRp. Replication then proceeds in two main steps: first, the RdRp synthesizes a complementary negative-sense RNA (-RNA) strand using the +RNA genome as a template, forming a double-stranded RNA intermediate; subsequently, the -RNA serves as the template for producing new +RNA genomes and additional mRNAs. This process often occurs within virus-induced membrane-bound replication organelles that concentrate viral and host factors for efficient synthesis. For negative-sense single-stranded (-ssRNA) viruses, including orthomyxoviruses and paramyxoviruses, the genomic cannot serve as mRNA due to its antisense ; instead, the virion-associated RdRp initiates primary transcription immediately after uncoating, producing positive-sense mRNAs (+mRNA) from the - within the ribonucleoprotein . Once sufficient viral proteins, including nucleoproteins, accumulate, replication shifts to full-length antigenome synthesis (+ complementary to the ), which then the production of new - progeny . This sequential transcription-replication switch is regulated by the availability of nucleoproteins and access to promoter sequences. Double-stranded RNA (dsRNA) viruses, such as reoviruses, package their RdRp within the virion core, enabling transcription of one strand of the dsRNA into +mRNA shortly after cell entry, without initial host . replication occurs conservatively in subviral particles—immature capsid-like structures—where parental dsRNA segments are transcribed into full-length +RNA intermediates that are then packaged with newly synthesized -RNA strands to form progeny dsRNA without mixing with host ribosomes. This enclosed mechanism ensures fidelity and segregation of viral segments during replication. Many RNA viruses with polycistronic , particularly +ssRNA viruses like coronaviruses, employ subgenomic RNAs (sgRNAs) to express downstream genes efficiently, avoiding the need for ribosomal frameshifting or polyprotein cleavage alone. In coronaviruses, sgRNAs are generated through discontinuous transcription, where the RdRp pauses at transcription-regulatory sequences (TRS) on the negative-sense template and switches to the 5' leader sequence of the , producing a nested set of 3'-coterminal mRNAs with common 5' and 3' ends. This strategy allows coordinated expression of structural and accessory proteins from a single promoter, with sgRNA abundance regulated by TRS strength and polymerase processivity. Reverse-transcribing RNA viruses of Group VI, such as retroviruses (e.g., human immunodeficiency virus), differ by using a DNA intermediate in their replication. Upon entry, the positive-sense ss genome is reverse-transcribed into double-stranded DNA by virion-packaged , an RNA-dependent . This proviral DNA is then integrated into the host cell genome by viral integrase, becoming a permanent . Host transcribes the provirus to produce full-length viral RNAs, which function as both mRNAs for translating viral proteins (via splicing for some) and as genomes packaged into new virions during assembly. This reliance on host transcription machinery and reverse transcription sets Group VI apart from other RNA viruses.

Genetic Variability

Mutation Rates and Error-Prone Polymerases

RNA viruses are characterized by exceptionally high mutation rates, typically ranging from 10^{-3} to 10^{-6} substitutions per per replication cycle, which is orders of magnitude higher than those observed in DNA-based organisms. This elevated variability stems primarily from their reliance on RNA-dependent RNA polymerases (RdRps), enzymes that synthesize RNA from RNA templates without the 3'–5' proofreading activity common in many DNA polymerases. As a result, errors introduced during replication accumulate rapidly, generating diverse viral populations known as quasispecies that enhance adaptability to defenses and environmental pressures. The error-prone nature of RdRps is an inherent property of these viral enzymes, which prioritize replication speed over fidelity to support the rapid life cycles of RNA viruses. For instance, in , the 3D introduces approximately 10^{-4} mutations per copied, leading to about one per per replication cycle in its ~7.5 . Similarly, HIV-1's , an RNA-dependent DNA , exhibits a of around 10^{-5} substitutions per , compounded by frequent template switching. These rates reflect a evolutionary : while might reduce deleterious mutations, it could slow replication and impair in dynamic host environments, as demonstrated by experiments where engineered higher-fidelity variants showed reduced despite lower loads. Not all RNA viruses conform to this pattern of extreme error proneness. Coronaviruses, such as , possess an accessory proofreading exonuclease (nsp14) that enhances RdRp fidelity, reducing the to approximately 10^{-6} per —about 10- to 15-fold lower than in proofreading-deficient relatives like murine hepatitis virus mutants. This mechanism allows for larger genome sizes (up to ~30 kb) without exceeding Eigen's error threshold, beyond which viral populations collapse due to excessive . Overall, the dynamics driven by these polymerases underpin the evolutionary success of RNA viruses, enabling rapid diversification while posing challenges for antiviral strategies that exploit their mutational vulnerability, such as lethal .

Recombination and Segment Reassortment

Recombination in RNA viruses involves the exchange of genetic material between viral genomes, primarily occurring through template switching by the viral (RdRp) during replication. This process, often termed copy-choice recombination, allows the polymerase to dissociate from one RNA template and associate with another, generating chimeric genomes. In positive-sense single-stranded RNA (+ssRNA) viruses, such as coronaviruses, this mechanism is frequent and facilitates adaptation by reshuffling genes, as seen in the evolution of variants through inter-genomic recombination in the spike gene. In contrast, recombination is rarer in negative-sense single-stranded RNA (-ssRNA) viruses due to the separation of replication and transcription compartments, though it can occur via similar template switching under co-infection conditions. Recombination mechanisms are classified as homologous or non-homologous, both driven by co-infection of the same host cell by two or more viral strains, which provides the necessary templates for exchange. Homologous recombination, the more common type, involves template switching at sites of high sequence similarity, preserving genome structure and often producing viable progeny, as exemplified in +ssRNA viruses like enteroviruses and coronaviruses. Non-homologous recombination, occurring without sequence homology, is less frequent and typically generates defective or rearranged genomes, potentially contributing to viral diversity but at lower efficiency. Recombination rates vary by virus but are estimated around 10^{-4} per site in susceptible +ssRNA viruses, exceeding baseline mutation rates and enabling rapid genetic diversification during mixed infections. Segment reassortment, a distinct form of genetic exchange unique to segmented RNA viruses, involves the packaging of genome segments from different parental viruses into a single virion during co-infection. This process shuffles entire segments without altering their sequences, leading to novel viral genotypes that can enhance transmissibility or host range. In influenza A viruses, which have eight negative-sense segments, reassortment drives antigenic shifts, as observed in the 2009 H1N1 pandemic where segments from human, avian, and swine strains combined. Rotaviruses, with 11 double-stranded segments, similarly undergo reassortment in the gut , contributing to strain diversity and vaccine escape, with frequencies reaching up to 10% of progeny in experimental mixed infections. Reassortment efficiency depends on segment and viral but is generally higher than recombination in non-segmented viruses due to independent segment packaging.

Classification

Baltimore Classification System

The Baltimore classification system, proposed by David Baltimore in 1971, categorizes viruses into seven groups based on their genomic nucleic acid type—whether DNA or RNA, single- or double-stranded, and sense (positive or negative)—and the specific pathway used to synthesize messenger RNA (mRNA) from the genome during replication. This framework emphasizes the central role of mRNA as the intermediary for protein synthesis, distinguishing viruses by how they bridge the gap between their genetic material and host translational machinery, such as through direct use of the genome as mRNA or reliance on viral polymerases for transcription or reverse transcription. Originally outlined for animal viruses, the system has since been broadly applied to all viruses, providing a functional rather than phylogenetic basis for organization. RNA viruses are assigned to groups III through VI within this scheme, reflecting their reliance on RNA-directed RNA polymerases (RdRps) or s for genome expression. Group III includes double-stranded RNA (dsRNA) viruses, which package segmented genomes and use virion-associated RdRp to transcribe one strand into mRNA. Group IV encompasses positive-sense single-stranded RNA (+ssRNA) viruses, whose non-segmented or segmented genomes function directly as mRNA upon entry into the host cell. Group V comprises negative-sense single-stranded RNA (-ssRNA) viruses, typically segmented, that carry RdRp in the virion to transcribe the genomic RNA into complementary mRNA. Group VI covers positive-sense ssRNA viruses that employ to generate a DNA from the RNA genome, which then serves as a template for mRNA production. These groupings underscore the diversity in RNA virus replication strategies while excluding DNA viruses (groups I, II, and VII). A primary advantage of the Baltimore system lies in its predictive power for viral replication mechanisms, enabling inferences about required enzymes, genome stability, and potential therapeutic targets without needing full genomic sequencing. It also integrates with the International Committee on Taxonomy of Viruses (ICTV) hierarchical taxonomy, where RNA virus groups III, IV, and V align with the realm Riboviria, a monophyletic assemblage defined by shared RdRp usage, thus bridging molecular function with evolutionary classification. Nonetheless, the system has notable limitations: it prioritizes replication pathways over phylogenetic relationships, host specificity, or ecological roles, and does not incorporate concepts like viral quasispecies dynamics that influence RNA virus evolution and adaptation.

Group III: Double-Stranded RNA Viruses

Group III viruses, also known as double-stranded RNA (dsRNA) viruses, are characterized by genomes consisting of segmented dsRNA molecules, typically packaged within non-enveloped, icosahedral capsids with multiple protein layers. These viruses replicate entirely in the host cell cytoplasm, utilizing viral RNA-dependent RNA polymerase (RdRp) enzymes packaged within the virion to initiate transcription without relying on host nuclear machinery. The genome segmentation, ranging from 2 to 12 linear dsRNA segments depending on the family, encodes structural proteins, polymerases, and non-structural factors essential for assembly and replication. The primary family within Group III is Reoviridae, which includes genera such as Orthoreovirus, , and Orbivirus, featuring 10–12 genome segments enclosed in a double-layered structure. Representative examples include rotaviruses, which infect mammals including humans, and bluetongue virus, which primarily affects ruminants. Another key family, Birnaviridae, possesses a bisegmented dsRNA (segments A and B) packaged in a single-shelled, non-enveloped icosahedral virion lacking the inner core typical of Reoviridae. Examples from this family, such as infectious bursal disease virus, target birds and aquatic animals. A smaller family, Picobirnaviridae, also features bisegmented genomes but remains less characterized. Replication of Group III viruses follows a conservative model, where the parental dsRNA duplex remains intact as a template for both transcription and replication within cytoplasmic subviral particles. Upon entry into the host cell via , the outer capsid layer is proteolytically removed to generate transcriptionally active cores, which contain the RdRp complex associated with each . The RdRp transcribes the negative-sense strand of each dsRNA into positive-sense mRNA, which exits the core through channels at the icosahedral vertices for into proteins in the . Subsequent replication within these cores synthesizes full-length complementary strands to form new dsRNA s, packaged into progeny virions. The segmented nature of the facilitates genetic reassortment during co-infection, contributing to diversity. These viruses predominantly infect and , with Reoviridae spanning a broad host range including mammals, birds, , and , while Birnaviridae mainly affects vertebrates such as , amphibians, and . infections are rare and typically limited to specific Reoviridae members like certain rotaviruses.

Group IV: Positive-Sense Single-Stranded RNA Viruses

Group IV viruses possess a single-stranded RNA genome of positive polarity, which functions directly as (mRNA) and can be translated by host ribosomes to produce viral proteins immediately upon infection. This genomic feature distinguishes them from other RNA virus groups, enabling rapid initiation of the replication cycle without requiring prior transcription. sizes typically range from 3.5 to 40 kilobases, with most being non-segmented, though some exceptions exist. The viral encodes one or more large polyproteins that are proteolytically processed into structural and non-structural components, including the essential (RdRp). Unlike negative-sense viruses, Group IV virions do not package RdRp or other enzymes; these are synthesized de novo from the genomic . Replication occurs in the host , often on rearranged intracellular membranes, where the RdRp first generates a complementary negative-sense intermediate that serves as a template for producing new positive-sense genomic and, in certain families, subgenomic mRNAs for downstream open reading frames. Some members, like coronaviruses, employ mechanisms via a 3'-5' to maintain larger stability. Prominent families within Group IV include Picornaviridae, such as , which causes acute in humans through non-enveloped, icosahedral virions. Flaviviridae encompasses enveloped viruses like and , transmitted by vectors and leading to febrile illnesses in humans. Coronaviridae features enveloped pathogens including , which emerged as a major human respiratory threat. In plants, Potyviridae represents the largest family, with examples like infecting solanaceous crops and causing mosaic symptoms and yield losses. These viruses trace their origins to ancient RNA replicators, potentially from a primordial , as evidenced by the deep phylogenetic roots of their RdRp enzymes within the realm. Diversification is driven by high rates of approximately 10^{-3} to 10^{-5} substitutions per per replication cycle, owing to the error-prone nature of viral RdRp. Recombination serves as a key evolutionary mechanism, particularly in families like , where template-switching during replication creates hotspots in structural genes such as the , facilitating host jumps and variant emergence. Group IV viruses exhibit broad host , infecting organisms across multiple domains, including (e.g., Leviviridae in Enterobacteriaceae), fungi (e.g., Narnaviridae in yeasts), (e.g., Potyviridae in angiosperms), (e.g., Flaviviridae in mosquitoes), and vertebrates (e.g., Picornaviridae and in mammals). This versatility underscores their ecological significance and potential.

Group V: Negative-Sense Single-Stranded RNA Viruses

Group V viruses, also known as negative-sense single-stranded (-ssRNA) viruses, possess genomes that are complementary to (mRNA), rendering them incapable of direct translation by host ribosomes upon entry into a . Unlike positive-sense viruses, these genomes require transcription into positive-sense mRNA before protein synthesis can occur, a process mediated by a virion-packaged (RdRp). This polymerase, along with nucleoproteins, forms a ribonucleoprotein complex that protects the fragile genome and initiates primary transcription immediately after uncoating. Consequently, the purified genomic of -ssRNA viruses is non-infectious when introduced into cells, as it lacks the enzymatic machinery for transcription without the accompanying viral proteins. Key families within Group V include the , which encompass influenza viruses; the , including the ; the , such as the measles virus; the , featuring Ebola virus; and the Bunyaviridae (now reclassified into several families like Peribunyaviridae), which include viruses like La Crosse virus. These viruses are predominantly enveloped and exhibit diverse morphologies, from bullet-shaped (rhabdoviruses) to filamentous (filoviruses). Genome segmentation varies across families: most, such as those in and , have non-segmented genomes ranging from 10 to 19 kilobases, while genomes consist of eight distinct segments encoding up to 11 proteins in influenza A virus. Segmented -ssRNA viruses, including orthomyxoviruses and bunyaviruses, employ a unique cap-snatching mechanism, where the viral endonuclease cleaves 5' cap structures from host mRNAs to prime viral transcription, enhancing mRNA stability and translation efficiency.30321-X) -ssRNA viruses primarily infect vertebrates, causing significant diseases in humans, , and , with examples including respiratory from paramyxoviruses and hemorrhagic fevers from filoviruses. Certain families, notably Bunyaviridae, are arthropod-borne, utilizing mosquitoes, ticks, or sandflies as vectors that serve both as replicative hosts and transmission intermediaries to vertebrate hosts. In segmented viruses like , genome reassortment during co-infection can generate novel strains, contributing to and potential.

Group VI: Reverse-Transcribing RNA Viruses

Group VI viruses, also known as reverse-transcribing RNA viruses, are characterized by a positive-sense single-stranded (+ssRNA) genome that is replicated through a DNA intermediate, distinguishing them from other RNA viruses that directly use RNA-dependent RNA polymerases for replication. These viruses package two copies of their linear +ssRNA , typically 7–13 in length, as a dimer within enveloped virions. Upon entry into the host , the viral RNA serves as a template for reverse transcription, producing a double-stranded DNA (dsDNA) copy that integrates into the host . Central to their replication is the enzyme (RT), which possesses both and RNase H activities, enabling the synthesis of (cDNA) from the template while degrading the RNA strand in RNA-DNA hybrids. Reverse transcription begins in the using a host tRNA as a primer, generating a linear dsDNA molecule through a series of strand transfers and elongation steps. The resulting viral dsDNA is then processed by the viral integrase enzyme, which catalyzes its insertion into the host cell's chromosomal DNA, forming a that serves as a permanent template for viral . Integrase performs two key reactions: 3'-processing to expose reactive ends on the viral DNA and strand transfer to join these ends to the host DNA. The primary family in Group VI is Retroviridae, which includes subfamilies Orthoretrovirinae and , encompassing 11 genera without segmented genomes. Representative examples include human virus () from the genus and human T-lymphotropic virus (HTLV) from the Deltaretrovirus genus, both causing persistent infections in humans. The proviral integration enables a latent phase where the virus evades immune detection, with viral genes transcribed by host only upon cellular activation, leading to lifelong infection and potential oncogenesis or . Due to the error-prone nature of reverse transcriptase, lacking 3'–5' proofreading activity, these viruses exhibit high mutation rates, approximately 10^{-4} to 10^{-5} errors per , contributing to and immune evasion.

Evolution and Origins

Ancient Origins and Fossil Evidence

The evolutionary origins of RNA viruses are hypothesized to predate those of DNA viruses, consistent with the RNA world hypothesis, which proposes that self-replicating RNA molecules functioned as both genetic material and catalysts in primordial life, potentially including early RNA replicases that gave rise to viral polymerases. This scenario suggests RNA viruses emerged during an ancient phase of life dominated by -based biochemistry, before the transition to DNA genomes in cellular organisms. Although direct evidence for such early viruses is scarce due to RNA's instability, indirect genetic traces support their deep antiquity. Endogenous viral elements (EVEs)—genomic integrations of ancient viral sequences—serve as molecular fossils revealing past RNA virus infections. For instance, endogenous retroviral elements derived from reverse-transcribing viruses are widespread in mammalian genomes, with some integrations predating the divergence of placental mammals by over 100 million years. Similarly, bornavirus-like EVEs from negative-sense single-stranded viruses indicate infections exceeding 93 million years in s. In aquatic lineages, filovirus-like EVEs have been identified in fish genomes, suggesting RNA viruses infected early ancestors potentially hundreds of millions of years ago, based on divergence timelines. Metagenomic analyses of environmental samples, particularly from , have uncovered a vast and diverse RNA virosphere, expanding known viral lineages and implying ancient origins tied to aquatic ecosystems. These studies reveal RNA viruses in and other microbes with phylogenetic depths reaching at least 600 million years, as seen in the order Articulavirales associated with early aquatic animals. Such findings highlight the as a cradle for RNA virus diversification, with enabling reconstruction of evolutionary histories far beyond traditional sampling. A key indicator of common ancestry among RNA viruses is the highly conserved (RdRp), an essential enzyme present across diverse groups from positive-sense to , tracing back to a shared replicase. Evidence of , detected through metagenomic phylogenies, further underscores how ancient RNA viruses exchanged genetic material, contributing to their early radiation and adaptation.

Mechanisms of Diversification and Host Adaptation

RNA viruses exhibit exceptionally high rates, primarily due to the error-prone nature of their RNA-dependent RNA polymerases, which lack mechanisms, leading to frequent substitutions during replication. These rates, often exceeding 10^{-4} mutations per per replication cycle, generate a diverse array of genetic variants within a single infected host, forming what is known as a quasispecies—a dynamic population of closely related but non-identical genomes subjected to continuous variation, competition, and selection. This quasispecies structure enhances the virus's adaptability by providing a of mutants that can rapidly respond to selective pressures, such as immune responses or antiviral drugs. In addition to point mutations, recombination plays a crucial role in RNA virus diversification by allowing the exchange of genetic material between co-infecting viral genomes, producing chimeric variants with novel combinations of traits. This process is particularly prevalent in positive-sense single-stranded RNA viruses and segmented viruses, where it can accelerate the of new serotypes or enhance in diverse environments. For instance, recombination contributes to the of coronaviruses, enabling the virus to explore broader and evade host defenses more effectively. Host adaptation in RNA viruses often occurs through spillover events, where viruses jump from reservoir species to new hosts, followed by rapid evolutionary changes that optimize transmission and replication in the novel environment. A prominent example is the 2019 emergence of , which likely spilled over from bats to humans via an intermediate host at a wildlife market in , , allowing the virus to adapt key mutations for efficient human ACE2 receptor binding. Similarly, highly pathogenic H5N1 has demonstrated repeated spillover from birds to mammals, including recent transmissions to in 2024, driven by mutations that alter receptor specificity and enhance mammalian cell tropism. Zoonotic reservoirs, primarily wildlife such as bats and birds, serve as the source for approximately 75% of emerging infectious diseases, providing a persistent pool from which RNA viruses can spill over to humans or other animals under conditions of increased contact, such as habitat encroachment. Climate change exacerbates these risks by expanding the geographic range and activity seasons of arthropod vectors like mosquitoes, which transmit RNA viruses such as dengue and West Nile, thereby facilitating more frequent spillover opportunities through altered vector-pathogen dynamics and extended transmission windows. Phylodynamic analyses, employing Bayesian statistical frameworks to integrate genetic sequences with epidemiological data, have been instrumental in quantifying RNA virus evolution rates and reconstructing adaptation timelines. For example, these models estimate the evolutionary rate of A viruses at approximately 10^{-3} substitutions per site per year, revealing how antigenic drift accumulates over seasons to drive host and vaccine escape. Such approaches highlight the interplay between , selection, and in shaping viral diversification and .

Role in Disease and Ecology

Major Human Pathogens

RNA viruses are responsible for a wide array of significant human diseases, ranging from acute infections to chronic conditions and pandemics. Key families include Picornaviridae, , , and Retroviridae, among others, which collectively cause millions of cases annually and substantial mortality. These pathogens exploit diverse routes and pose ongoing challenges due to their , which complicates development and therapeutic interventions. Picornaviruses, such as and , exemplify acute infections with fecal-oral transmission. , once a major cause of , has been nearly eradicated through global vaccination efforts, with only isolated wild poliovirus type 1 cases reported in and in 2025, totaling fewer than 100 confirmed instances. The inactivated and oral have reduced global incidence by over 99% since 1988, though vaccine-derived strains remain a concern in under-vaccinated areas. causes self-limiting liver , affecting an estimated 1.4 million people yearly worldwide, primarily in regions with poor , and is preventable by a highly effective administered in two doses. Flaviviruses like hepatitis C virus (HCV) lead to chronic liver disease through bloodborne transmission, often via shared needles or unsafe medical practices. Globally, an estimated 50 million people have chronic hepatitis C virus infection, with approximately 242,000 deaths in 2022 from related liver cirrhosis and hepatocellular carcinoma, though direct-acting antivirals achieve cure rates exceeding 95% in treated cases. Norovirus, from the Caliciviridae family, is the leading cause of viral gastroenteritis in humans, infecting about 685 million people annually and resulting in roughly 200,000 deaths, mainly in vulnerable populations; it spreads via fecal-oral route, contaminated food, or surfaces, with no specific vaccine available but hygiene measures reducing outbreaks by up to 50%. Coronaviruses, particularly , have caused unprecedented global impact through respiratory droplet transmission. The , declared in 2020, resulted in over 7 million confirmed deaths worldwide by mid-2025, though excess mortality estimates suggest higher figures due to underreporting. mRNA-based vaccines, such as those from Pfizer-BioNTech and , demonstrated 90-95% efficacy against severe disease in initial trials and have been administered to billions, though evolving variants like necessitate boosters to maintain protection above 70%. Influenza viruses (), also respiratory pathogens, cause seasonal epidemics and occasional pandemics; the 1918 H1N1 pandemic killed an estimated 50 million people, while the 2009 H1N1 outbreak affected over 60 million in the U.S. alone, with annual vaccines updated yearly to match circulating strains, reducing hospitalization by 40-60%. Retroviruses, notably from the Retroviridae family, establish lifelong infections via blood, sexual, or perinatal , with 40.8 million people living with globally in 2024. Antiretroviral therapy suppresses to undetectable levels in over 29 million treated individuals, preventing progression to AIDS and reducing by 96%, though no curative exists despite ongoing trials. These pathogens highlight the need for integrated strategies, including , , and antivirals, to mitigate their burden amid mutation-driven variants that evade immunity.

Impacts on Animals, Plants, and Ecosystems

RNA viruses exert profound effects on health, particularly through economically devastating diseases in and . , caused by a , is a highly contagious affecting cloven-hoofed animals such as , sheep, and pigs, leading to blisters, fever, and reduced productivity. Globally, it results in direct production losses and vaccination costs exceeding $21 billion annually in endemic regions. Similarly, , a , causes fatal neurological disease in a wide range of mammals, including dogs, , and , with outbreaks leading to significant culling and population declines in affected communities. While human fatalities from number approximately 59,000 per year, the virus inflicts even greater losses on populations through direct mortality and control measures. In plants, RNA viruses represent major agricultural threats, often lacking envelopes which facilitates their mechanical transmission via tools, sap, or . The (TMV), the first virus identified in plants in 1857, infects over 350 species including , tomatoes, and peppers, causing mottled leaves, stunted growth, and yield reductions. Potyviruses, the largest group of plant-infecting RNA viruses, affect crops like potatoes, beans, and cereals, with infections leading to mosaic symptoms, , and crop losses ranging from 10% to 53% depending on the host and strain. These viruses contribute to substantial global agricultural impacts, undermining in vulnerable regions. Beyond direct , RNA viruses play key ecological roles in regulating host populations and maintaining . By infecting and controlling microbial and animal communities, they influence primary productivity and biogeochemical cycles in terrestrial and aquatic ecosystems. Arboviruses, transmitted by vectors such as mosquitoes, exemplify this dynamic, as they modulate vector populations and host interactions, potentially preventing overpopulation of species and promoting through selective pressure. In marine environments, diverse viruses shape community structures, with their abundance correlating to and . Many viruses bridge animal and human health via zoonotic transmission, underscoring their ecosystem-wide implications. Approximately 75% of emerging infectious diseases in humans originate from animal reservoirs, facilitating spillover events that disrupt dynamics. For instance, , carried by fruit bats, spills over to pigs and humans, causing severe outbreaks that affect bat populations indirectly through habitat changes and control efforts. Such zoonoses highlight how RNA viral circulation in animal hosts can destabilize ecosystems while posing risks to and .

References

  1. [1]
    Structure and Classification of Viruses - Medical Microbiology - NCBI
    Viruses are small obligate intracellular parasites, which by definition contain either a RNA or DNA genome surrounded by a protective, virus-coded protein coat.<|control11|><|separator|>
  2. [2]
    Introduction to RNA Viruses - PMC - PubMed Central
    Definition and Basic Properties of RNA Viruses. RNA viruses replicate their genomes using virally encoded RNA-dependent RNA polymerase (RdRp). The RNA genome ...
  3. [3]
    The Baltimore Classification of Viruses 50 Years Later
    The five classes of RNA viruses and reverse-transcribing viruses share an origin, whereas both the single-stranded DNA viruses and double-stranded DNA (dsDNA) ...
  4. [4]
    Classification of Human Viruses - PMC - PubMed Central - NIH
    Virus families are designated with the suffix -viridae. Families are distinguished largely on the basis of physiochemical properties, genome structure, size, ...
  5. [5]
    RNA virus mutations and fitness for survival - PubMed - NIH
    RNA viruses exploit all known mechanisms of genetic variation to ensure their survival. Distinctive features of RNA virus replication include high mutation ...
  6. [6]
    RNA Viruses: RNA Roles in Pathogenesis, Coreplication and Viral ...
    1.1. RNA Viruses. Human diseases causing RNA viruses include Orthomyxoviruses, Hepatitis C Virus (HCV), Ebola disease, SARS, influenza, polio measles and ...
  7. [7]
    RNA viruses: a case study of the biology of emerging infectious ...
    Sep 26, 2018 · The only two Level 4 RNA viruses which are vector borne – dengue and yellow fever – are also relatively virulent, in line with ideas that vector ...
  8. [8]
    [PDF] The Discovery of the Causal Agent of the Tobacco Mosaic Disease
    This review documents the contributions of the three men who were the pioneers in this work, namely, Adolph Mayer, Dmitrii Iwanowski and Martinus Beijerinck. To ...
  9. [9]
    Historical lessons from the first discovery of a virus - Sage Journals
    Feb 15, 2021 · It was Russian plant physiologist Dmitri Ivanovsky (1864–1920) who eventually discovered that the causative agent of TMD could pass through a ...
  10. [10]
    Tobacco Mosaic Virus - an overview | ScienceDirect Topics
    The discovery of viruses is attributed to Dmitry Ivanovsky, a Russian microbiologist who, between 1887 and 1890, investigated the mosaic disease of tobacco ...
  11. [11]
    Modern Uses of Electron Microscopy for Detection of Viruses - PMC
    Eight years later, Ruska and colleagues Kausche and Pfankuch were the first to visualize viruses (tobacco mosaic virus) with the EM (47), and in 1986, Ruska ...
  12. [12]
    7.10: Discovery and Origin of Viruses - Biology LibreTexts
    Mar 5, 2021 · Tobacco Mosaic Virus. This tobacco mosaic virus was the first virus to be discovered. It was first seen with an electron microscope in 1935.
  13. [13]
    Genetics of poliovirus. - Document - Gale Academic OneFile
    Schwerdt & Schaffer reported in 1955 (342) that the poliovirus genome is RNA. The virus was crystallized in 1955 (339), and its three-dimensional structure was ...
  14. [14]
    Virus Classification - an overview | ScienceDirect Topics
    The first taxonomic system that received broad attention is that of Lwoff, Horne, and Tournier (the LHT system), which was first published in 1962 and finalized ...
  15. [15]
    Virus Taxonomy - PMC - PubMed Central - NIH
    virus”). Baltimore Classification (1971). In 1971, David Baltimore published a working classification of viruses that is still used today in parallel with ...
  16. [16]
    The evolution of RNA viruses: A population genetics view - PMC
    RNA viruses are excellent experimental models for studying evolution under the theoretical framework of population genetics.
  17. [17]
    The population genetics and evolutionary epidemiology of RNA ...
    mutation, recombination, natural selection, genetic drift and migration, ...
  18. [18]
    Unconventional viral gene expression mechanisms as therapeutic ...
    May 19, 2021 · In this Review, we describe the types and mechanisms of unconventional gene and protein expression in viruses, and provide a perspective on how future basic ...
  19. [19]
    RNA levers and switches controlling viral gene expression - PMC
    Oct 24, 2025 · RNA viruses that enter the host nucleus can exploit alternative splicing to express multiple transcripts from one genomic sequence and regulate ...
  20. [20]
    A Tale of Three Recent Pandemics: Influenza, HIV and SARS-CoV-2
    Jun 2, 2022 · We will discuss the major determinants of HIV, pandemic influenza, and SARS-CoV-2, focusing on the host, the infectious agent and the ...
  21. [21]
    Molecular strategies to inhibit the replication of RNA viruses - PMC
    Drugs approved for the treatment of RNA virus infections (other than HIV) are the influenza M2 channel inhibitors, amantadine and rimantadine; the influenza ...
  22. [22]
    Decades in the Making: mRNA COVID-19 Vaccines | NIAID
    Apr 4, 2024 · These vaccines were developed with NIH support and research on a protein found on SARS-CoV-2, the virus that causes COVID-19.<|separator|>
  23. [23]
    The evolution of seasonal influenza viruses - Nature
    Oct 30, 2017 · The key process underlying these recurrent epidemics is the evolution of the viruses to escape the immunity that is induced by prior infection or vaccination.Missing: COVID- | Show results with:COVID-
  24. [24]
    History of 1918 Flu Pandemic - CDC Archive
    The number of deaths was estimated to be at least 50 million worldwide with about 675,000 occurring in the United States. Mortality was high in people younger ...
  25. [25]
    COVID-19 eliminated a decade of progress in global level of life ...
    May 24, 2024 · The pandemic wiped out nearly a decade of progress in improving life expectancy within just two years. Between 2019 and 2021, global life expectancy dropped by ...
  26. [26]
    The COVID-19 pandemic and continuing challenges to global health
    The International Monetary Fund has estimated cumulative economic loss to 2024 as a consequence of the pandemic at US$ 13.8 trillion.3 The pandemic has set back ...
  27. [27]
    RNA silencing in plants - Nature
    Sep 15, 2004 · Here, I describe three natural pathways of RNA silencing in plants that have been revealed by genetic and molecular analysis.Abstract · Main · Author Information
  28. [28]
    RNAi Nobel ignores vital groundwork on plants - Nature
    Oct 25, 2006 · The discovery of RNA interference (RNAi) changed the face of gene regulation, a feat deservedly recognized with this year's Nobel Prize in Physiology or ...
  29. [29]
    Therapeutic Applications for Oncolytic Self-Replicating RNA Viruses
    Dec 9, 2022 · Both naturally and engineered oncolytic self-replicating RNA viruses providing specific replication in tumor cells have been evaluated for cancer therapy.
  30. [30]
    A short biased history of RNA viruses - PMC - NIH
    Among its many roles, RNA can also act as a viral genome. In the majority of cases these genomes are single-stranded RNA, and both the form that can be ...
  31. [31]
    The hepatitis delta (δ) virus possesses a circular RNA - Nature
    Oct 9, 1986 · Here we present evidence based on electron microscopy and electrophoretic behaviour that HDV contains a single stranded circular RNA molecule.
  32. [32]
    Critical role of segment-specific packaging signals in genetic ... - PNAS
    Sep 16, 2013 · We show that packaging signals are crucial for genetic reassortment and that suboptimal compatibility between the segment-specific packaging signals of the two ...Abstract · Sign Up For Pnas Alerts · ResultsMissing: advantages | Show results with:advantages
  33. [33]
    Reassortment in segmented RNA viruses: mechanisms and outcomes
    Reassortment can create viral progeny that contain genes that are derived from more than one parent, potentially conferring important fitness advantages or ...
  34. [34]
    Genome Size - an overview | ScienceDirect Topics
    For RNA viruses, typically 3–33 Kb (see Chapter 1 for comparative genome size of viruses and cellular organisms). The limited genome size render RNA viruses ...
  35. [35]
    Pushing the limits on RNA virus genome size in the absence of ... - NIH
    Aug 26, 2024 · Mathematically speaking, smaller genomes will accumulate fewer total mutations and therefore are less likely to encounter an error catastrophe.
  36. [36]
    Structure and Organization of Virus Genomes - PMC - PubMed Central
    Nov 18, 2019 · Additionally, single stranded virus genomes may be either positive sense (+) where the RNA present in the genome will of the same polarity as ...
  37. [37]
    Ambisense polarity of genome RNA of orthomyxoviruses and ...
    Sep 25, 2021 · It means that genome RNA of negative sense polarity is transcribed by the virus polymerase to produce positive sense mRNAs, which recognized by ...
  38. [38]
    Structure and Composition of Viruses - PMC - PubMed Central
    The nucleocapsids of several RNA viruses have a different type of symmetry: the capsomers and nucleic acid molecule(s) self-assemble as a helix (Fig. 1-1C,D; ...
  39. [39]
    Virus Nucleocapsid - an overview | ScienceDirect Topics
    Capsids are protein coats that package viral genomes. Simple capsids are either rods (helical) or spheres. Spherical viruses have icosahedral symmetry. An ...
  40. [40]
    Genome packaging in viruses - PMC - PubMed Central - NIH
    Mechanism for dsRNA packaging ... In dsRNA viruses such as φ6 and φ12, the positive-sense strand is packaged and then replicated to form dsRNA inside the capsid.
  41. [41]
    Mechanisms of assembly and genome packaging in an RNA virus ...
    Dec 10, 2015 · Cryo-EM of human rhinovirus reveals capsid-RNA duplex interactions that provide insights into virus assembly and genome uncoating. Article ...
  42. [42]
    The curious case of genome packaging and assembly in RNA ...
    Jun 7, 2023 · Genome packaging is the crucial step for maturation of plant viruses containing an RNA genome. Viruses exhibit a remarkable degree of packaging ...CPs with a novel ATPase... · A model for genome... · Strategies to combat plant...
  43. [43]
    Viral capsids: Mechanical characteristics, genome packaging and ...
    In such a mechanism cleavage of the concatemeric DNA and the termination of packaging takes place after a certain genome density is reached inside the capsid.Viral Capsids: Mechanical... · Capsid Mechanics · Genome Ejection
  44. [44]
    Principles for enhancing virus capsid capacity and stability from a ...
    Oct 2, 2019 · The capsids of double-stranded DNA viruses protect the viral genome from the harsh extracellular environment, while maintaining stability ...
  45. [45]
    The Importance of Glycans of Viral and Host Proteins in Enveloped ...
    Here, we review the effects of host glycans and viral proteins on biological behaviors of viruses, and the opportunities for prevention and treatment of viral ...Abstract · Introduction · Glycans of Viral Proteins and... · Host Glycans Affect Viral...
  46. [46]
    Ebolavirus glycoprotein structure and mechanism of entry - PMC
    The EBOV glycoprotein (GP) is the only virally expressed protein on the virion surface and is critical for attachment to host cells and catalysis of membrane ...
  47. [47]
    Mapping glycoprotein structure reveals Flaviviridae evolutionary ...
    Sep 4, 2024 · Viral glycoproteins drive membrane fusion in enveloped viruses and determine host range, tissue tropism and pathogenesis.
  48. [48]
    Role of surface glycans in enveloped RNA virus infections - PubMed
    Nov 22, 2023 · In this review, we summarize findings regarding structure-function correlation of glycans on enveloped RNA virus proteins.
  49. [49]
    Virus Entry: Looking Back and Moving Forward - PMC
    Successive steps in entry include binding to receptors, endocytosis, passage through one or more membranes, targeting to specific sites within the cell, and ...
  50. [50]
    Principles of Virus Uncoating: Cues and the Snooker Ball - PMC
    This review covers a wide range of cellular processes that enhance viral uncoating. The underlying mechanisms provide deep insights into cell biological and ...
  51. [51]
    The Enterovirus 71 A-particle Forms a Gateway to Allow Genome ...
    Mar 21, 2013 · One current model for picornavirus uncoating states that the Ig-like receptor binds into the canyon, dislodges the pocket factor, and allows the ...
  52. [52]
    RNA virus mutation rates: viral & cellular determinants
    Apr 27, 2017 · Hence, RNA virus replication is error prone due to the lack of proofreading activity, not because of an intrinsically lower fidelity polymerase.
  53. [53]
    Review Positive-strand RNA virus genome replication organelles
    The infectious virions of (+)RNA viruses transmit the viral genome as a single-stranded, messenger-sense RNA that is replicated intracellularly via a double- ...Viral Rna Genetics Is Driven... · Nodaviruses As Model (+)rna... · Concluding Remarks
  54. [54]
    Positive-Strand RNA Virus - an overview | ScienceDirect Topics
    Positive-strand RNA virus replication occurs via synthesis of a complementary negative-strand RNA, which serves as a template to produce new viral genomes, ...<|control11|><|separator|>
  55. [55]
    Viruses with Single-Stranded, Positive-Sense RNA Genomes - PMC
    Eight virus families whose members infect vertebrates are currently known to possess single-stranded, positive-sense RNA genomes.
  56. [56]
    Negative-Strand RNA Virus - an overview | ScienceDirect Topics
    After infiltrating the cell, the first mission of a negative-strand RNA virus is to make its RNA double-stranded by synthesizing the corresponding positive RNA ...
  57. [57]
    Structures and Mechanisms of Nonsegmented, Negative-Strand ...
    Sep 29, 2023 · In this review, we consider each of the steps involved in nsNSV transcription and replication and suggest how these relate to solved polymerase structures.
  58. [58]
    Negative and ambisense RNA virus ribonucleocapsids - ASM Journals
    Sep 26, 2023 · Reduced nucleoprotein availability impairs negative-sense RNA virus replication and promotes host recognition. ... Common mechanism for RNA ...
  59. [59]
    Replication and Expression Strategies of Viruses - PubMed Central
    For some RNA viruses, the infecting genome acts as mRNA. For other RNA and DNA viruses, viral mRNA is synthesized upon entry into the host cell. Figure 3.1.
  60. [60]
    RNA Virus Replication Complexes | PLOS Pathogens
    Jul 22, 2010 · Genome replication by dsRNA viruses occurs in subviral particles. These subviral particles, also called the cores, have an intact viral capsid ...
  61. [61]
    Structure of RNA polymerase complex and genome within a dsRNA ...
    Jun 25, 2018 · Most double-stranded RNA (dsRNA) viruses transcribe RNA plus strands within a common innermost capsid shell. This process requires coordinated ...
  62. [62]
    Continuous and Discontinuous RNA Synthesis in Coronaviruses - NIH
    Replication of the coronavirus genome requires continuous RNA synthesis, whereas transcription is a discontinuous process unique among RNA viruses.
  63. [63]
    Discontinuous template switching generates coronavirus ...
    Sep 25, 2025 · Discontinuous template switching generates coronavirus subgenomic RNAs from the 3ʹ viral genome end by 5ʹ to 3ʹ transcription | Journal of ...
  64. [64]
    Discontinuous and non‐discontinuous subgenomic RNA ...
    The best studied nidoviruses, the corona‐ and arteriviruses, employ a unique transcription mechanism, which involves discontinuous RNA synthesis.
  65. [65]
    Mutation Rates, Mutation Frequencies, and Proofreading-Repair ...
    Measurements during infections and with purified viral polymerases indicate that mutation rates for RNA viruses are in the range of 10−3 to 10−6 copying errors ...
  66. [66]
    Complexities of Viral Mutation Rates | Journal of Virology
    Jun 29, 2018 · Finally, for unclear reasons, single-stranded viruses tend to mutate more rapidly than double-stranded viruses, causing some single-stranded DNA ...Missing: percentage | Show results with:percentage
  67. [67]
    Why are RNA virus mutation rates so damn high? | PLOS Biology
    Aug 13, 2018 · RNA viruses have high mutation rates—up to a million times higher than their hosts—and these high rates are correlated with enhanced virulence ...
  68. [68]
    The cost of replication fidelity in an RNA virus - PNAS
    RNA viruses are characterized by high mutation rates compared with most DNA systems (1), due mainly to the lack of exonuclease proofreading activity displayed ...
  69. [69]
  70. [70]
  71. [71]
    Recombination in Positive-Strand RNA Viruses - Frontiers
    May 17, 2022 · The recombination rates vary markedly for different viral species. For example, positive single-stranded RNA viruses, such as picornaviruses ...
  72. [72]
    Why do RNA viruses recombine? | Nature Reviews Microbiology
    Jul 4, 2011 · The process of recombination that takes place in RNA viruses corresponds to the formation of chimeric molecules from parental genomes of mixed origin.
  73. [73]
    Influenza Virus Reassortment Occurs with High Frequency in the ...
    Our results indicate that reassortment between two like influenza viruses is efficient but also strongly dependent on dose and timing of the infections.Missing: rotavirus | Show results with:rotavirus<|separator|>
  74. [74]
  75. [75]
    Double-Stranded RNA Virus - an overview | ScienceDirect Topics
    This is particularly documented for the Reoviridae family of viruses which contain 10 (for reoviruses) or 11 (for rotaviruses) segments of genomic dsRNA.
  76. [76]
    Birnaviridae - an overview | ScienceDirect Topics
    Genome replication (and possibly mRNA synthesis) is initiated with a protein primer. •. In contrast to the members of the family Reoviridae the birnaviruses ...
  77. [77]
    The Formation and Function of Birnaviridae Virus Factories - NIH
    May 9, 2023 · The Birnaviridae is a family of viruses that have a single protein capsid and a double stranded RNA (dsRNA) genome divided into two segments [1] ...Missing: characteristics | Show results with:characteristics
  78. [78]
    Conservative transcription in three steps visualized in a double ... - NIH
    Nov 6, 2019 · Our results support an ouroboros model for endogenous conservative transcription in dsRNA viruses. RNA transcription is an ancient process ...
  79. [79]
    RNA Virus Replication Complexes - PMC - PubMed Central - NIH
    Jul 22, 2010 · Reoviruses have ∼12 copies of the polymerase in each virion, each associated with one of the ten to 12 dsRNA gene segments that make up the ...
  80. [80]
    Function, Architecture, and Biogenesis of Reovirus Replication ...
    Mar 21, 2019 · Most viruses that replicate in the cytoplasm of host cells form neoorganelles that serve as sites of viral genome replication and particle assembly.
  81. [81]
    A Glimpse on the Evolution of RNA Viruses - NIH
    RNA viruses are obligate intracellular parasites characterised by extremely high genetic variability and phenotypic diversity, facilitating infection of an ...
  82. [82]
    The Positive Sense Single Stranded RNA Viruses - PMC
    Members of the Leviviridae family occur worldwide and are abundantly present in sewage, waste water, animal and human faeces. In Asia a particular geographic ...
  83. [83]
    The mystery remains: How do potyviruses move within and between ...
    Aug 12, 2023 · The genus Potyvirus is considered as the largest among plant single‐stranded (positive‐sense) RNA viruses, causing considerable economic damage
  84. [84]
    Reduced Nucleoprotein Availability Impairs Negative-Sense RNA ...
    Negative-sense RNA viruses (NSVs) rely on prepackaged viral RNA-dependent RNA polymerases (RdRp) to replicate and transcribe their viral genomes.
  85. [85]
    Segmented, negative sense RNA viruses of humans - NIH
    Negative stranded RNA viruses are a large group of viruses that encode their genomes in RNA across multiple segments in an orientation antisense to messenger ...
  86. [86]
    Minus-Strand RNA Viruses - PMC - PubMed Central
    In contrast, some (−)RNA viruses, such as rabies and Ebola viruses, cause illnesses with high fatality rates but (fortunately) infect only a small fraction of ...
  87. [87]
    The structure of the influenza A virus genome | Nature Microbiology
    Jul 22, 2019 · The IAV genome consists of eight single-stranded viral RNA segments contained in separate viral ribonucleoprotein (vRNP) complexes that are packaged together ...
  88. [88]
  89. [89]
    Reverse Transcriptase and the Generation of Retroviral DNA - NCBI
    Reverse transcription—the reverse or “retro” flow of genetic information from RNA to DNA—is a hallmark of the retroviral replication cycle.Missing: Retroviridae | Show results with:Retroviridae
  90. [90]
    Structure and function of retroviral integrase - PMC - PubMed Central
    Integrase is the viral enzyme responsible for the catalytic steps involved in this process, and integrase strand transfer inhibitors are widely used to treat ...
  91. [91]
    Human Retroviruses - Medical Microbiology - NCBI Bookshelf - NIH
    Since the provirus is duplicated along with the cellular DNA during the replication cycle of the cell, infection of the cell will persist throughout the ...
  92. [92]
    Mutation Rates and Intrinsic Fidelity of Retroviral Reverse ...
    On the other hand, an error rate as low as 1.9 × 10−7 to 2.3 × 10−7 mutations per copied nucleotide has been determined for the yellow fever virus RNA-dependent ...
  93. [93]
    The RNA World and the Origins of Life - Molecular Biology of the Cell
    The hypothesis that RNA preceded DNA and proteins in evolution. In the earliest cells, pre-RNA molecules would have had combined genetic, structural, and ...Polynucleotides Can Both... · Pre-RNA World Probably... · Single-stranded RNA...
  94. [94]
    Investigating the Concept and Origin of Viruses - PMC
    Nov 3, 2020 · It is therefore logical to think that RNA viruses evolved first from RNA cells, and that later, DNA viruses evolved directly from RNA viruses, ...Origins Of Viruses: Which... · Figure 1 · Pathways To Dna Cells And...<|separator|>
  95. [95]
    Human Endogenous Retroviruses Are Ancient Acquired Elements ...
    HERVs are relics of ancient infections that affected the primates' germ line along the last 100 million of years, and became stable elements at the interface ...
  96. [96]
    Endogenous viruses: Connecting recent and ancient viral evolution
    This method has been used to estimate the integration dates of primate and rodent endogenous bornaviruses to be ~19–40 Myr ago (Ma) (Belyi et al., 2010a) and ...Missing: birnavirus- | Show results with:birnavirus-<|separator|>
  97. [97]
    PLOS Pathogens - Research journals
    Sep 3, 2024 · The finding of filovirus-like paleoviruses in fish genomes in each of the major lineages of viruses proposed to be associated with fishes (XILV ...Missing: age | Show results with:age
  98. [98]
    Evidence for an ancient aquatic origin of the RNA viral order ...
    Oct 31, 2023 · These data suggest that the Articulavirales has evolved over at least 600 million years, first emerging in aquatic animals.
  99. [99]
    Cryptic and abundant marine viruses at the evolutionary origins of ...
    Apr 7, 2022 · ... 70% of sequences for some families were ocean derived (fig. ... single-stranded RNA (+ssRNA) viruses (7). Previously, all viruses in ...Expanding The Rna Catalog · Marine Rna Viruses Double... · Marine Rna Viruses Revise...<|control11|><|separator|>
  100. [100]
    Origin and Evolution of RNA-Dependent RNA Polymerase - Frontiers
    Sep 19, 2017 · RNA-dependent RNA polymerases (RdRp) are very ancient enzymes and are essential for all viruses with RNA genomes. We reconstruct the origin and evolution of ...
  101. [101]
    Origins and Evolution of the Global RNA Virome - PubMed Central
    This reconstruction reveals the relationships between different Baltimore classes of viruses and indicates extensive transfer of viruses between distantly ...
  102. [102]
    Pacing a small cage: mutation and RNA viruses - Cell Press
    RNA viruses have an extremely high mutation rate, and we argue that the most plausible explanation for this is a trade-off with replication speed.
  103. [103]
    Quasispecies Theory and the Behavior of RNA Viruses - PMC
    Jul 22, 2010 · RNA viruses exist as a quasispecies. A virus replicating with a high mutation rate will generate a diverse mutant repertoire over the course of ...
  104. [104]
    Quasispecies Structure and Persistence of RNA Viruses - CDC
    Viral quasispecies are closely related (but nonidentical) mutant and recombinant viral genomes subjected to continuous genetic variation, competition, and ...
  105. [105]
    Recombination in Positive-Strand RNA Viruses - PMC
    May 18, 2022 · RNA copy-choice recombination of CHIKV is responsible for genome diversification. Moreover, RNA recombination can generate new viral ...
  106. [106]
    Why do RNA viruses recombine? - PMC - PubMed Central
    Recombination can be an important evolutionary force for RNA viruses, but the rate of recombination varies greatly between different RNA viruses. In this ...
  107. [107]
    Extensive Recombination-driven Coronavirus Diversification ...
    Most focus on RNA virus evolution concerns the role of point mutation, but recombination plays an underappreciated role in generating coronavirus genetic ...
  108. [108]
    Cross-Species Virus Transmission and the Emergence of New ...
    The SARS CoV appeared to gain some host-adaptive changes during its spread ... Evolution and adaptation of H5N1 influenza virus in avian and human hosts in ...
  109. [109]
    Ecology, evolution and spillover of coronaviruses from bats - Nature
    Nov 19, 2021 · Since the emergence of SARS-CoV in 2002, and the evidence that it originated from a bat reservoir, coronaviruses have been detected in 16% of ...Missing: credible | Show results with:credible
  110. [110]
    Spillover of highly pathogenic avian influenza H5N1 virus to dairy ...
    Jul 25, 2024 · Infectious virus and viral RNA were consistently detected in milk from affected cows. Viral distribution in tissues via immunohistochemistry and ...Missing: SARS- | Show results with:SARS-
  111. [111]
    About Zoonotic Diseases | One Health - CDC
    Apr 7, 2025 · Scientists estimate that more than 6 out of every 10 known infectious diseases in people can be spread from animals, and 3 out of every 4 new or ...
  112. [112]
    Vector-Borne Diseases | Climate and Health - CDC
    Mar 2, 2024 · Climate changes can lead to vector/pathogen adaptations, causing shifts or expansions in their geographic ranges. Such shifts can alter disease ...
  113. [113]
    Bayesian phylodynamics reveals the transmission dynamics of ... - NIH
    Apr 17, 2023 · Our study used a phylodynamic model incorporating geographical and host structure to uncover critical information about the transmission dynamics of H7N9 virus ...
  114. [114]
    Divergent evolutionary trajectories of influenza B viruses underlie ...
    Dec 16, 2019 · The mean nucleotide substitution rates (evolutionary rate) was 1.93 × 10−3 (95% HPD 1.76–2.1 × 10−3) and 2.41 × 10−3 (2.18–2.66 × 10−3) ...
  115. [115]
    Phylodynamics of H1N1/2009 influenza reveals the transition from ...
    Aug 6, 2015 · Here we show that natural selection acting on H1N1/2009 directly after introduction into humans was driven by adaptation to the new host.<|control11|><|separator|>
  116. [116]
    The diversity of human RNA viruses - PMC - PubMed Central - NIH
    We find that 158 RNA virus species as recognised by the ICTV [10] have been reported to infect humans (Table 1). These species are distributed among 47 genera ...
  117. [117]
    Statement of the Forty-second meeting of the Polio IHR Emergency ...
    Jul 28, 2025 · A total of 275 WPV1 positive environmental samples have been reported in 2025 so far (as of 04 June), 30 from Afghanistan and 245 from Pakistan.
  118. [118]
    Poliomyelitis (polio) - World Health Organization (WHO)
    26 August 2025. Polio eradication strategy 2022-2026: delivering on a promise, extension to 2029. Based on today's epidemiology and after critical analysis ...
  119. [119]
    Hepatitis A - World Health Organization (WHO)
    Feb 12, 2025 · The disease is closely associated with unsafe water or food, inadequate sanitation, poor personal hygiene and oral-anal sex. Unlike hepatitis B ...
  120. [120]
    Hepatitis C - World Health Organization (WHO)
    Jul 25, 2025 · Hepatitis C is a viral infection that affects the liver. It can cause both acute (short term) and chronic (long term) illness. It can be life-threatening.
  121. [121]
    About Norovirus - CDC
    Apr 24, 2024 · Norovirus spreads very easily and quickly in different ways. You can get norovirus by: Having direct contact with someone with norovirus, like ...Site Index | Norovirus | CDC · Laboratory Testing for Norovirus
  122. [122]
    Norovirus Facts and Stats - CDC
    May 8, 2024 · Norovirus is the leading cause of vomiting and diarrhea from acute gastroenteritis among people of all ages in the United States.Missing: transmission | Show results with:transmission
  123. [123]
    COVID-19 deaths | WHO COVID-19 dashboard - WHO Data
    Total cumulative. Count; Rate per 100 000. Number of COVID-19 deaths reported to WHO. World, 28 days to 19 October 2025. Show table. Open full-screen mode ...
  124. [124]
    COVID-19 epidemiological update – 17 January 2025
    Jan 17, 2025 · Thirty-one (13%) countries reported COVID-19 deaths, and 81 (35%) countries reported COVID-19 cases globally during the 28-day period from 11 November to 8 ...
  125. [125]
    The Deadliest Flu: 1918 Pandemic Virus Discovery & Reconstruction
    The 1918 H1N1 flu pandemic, sometimes referred to as the “Spanish flu,” killed an estimated 50 million people worldwide, including an estimated 675,000 people ...Missing: toll | Show results with:toll
  126. [126]
    2009 H1N1 Pandemic (H1N1pdm09 virus) - CDC Archive
    Jun 11, 2019 · In the spring of 2009, a novel influenza A (H1N1) virus emerged. It was detected first in the United States and spread quickly across the United States and the ...
  127. [127]
    Global HIV & AIDS statistics — Fact sheet - UNAIDS
    Global HIV statistics: People living with HIV: In 2024, there were 40.8 million [37.0 million–45.6 million] people living with HIV.
  128. [128]
    HIV and AIDS - World Health Organization (WHO)
    Jul 15, 2025 · By 2025, 95% of all people living with HIV should have a diagnosis, 95% of whom should be taking lifesaving antiretroviral treatment, and 95% of ...
  129. [129]
    Foot and mouth disease - World Organisation for Animal Health
    Foot and mouth disease (FMD) is a severe, highly contagious viral disease of livestock that has a significant economic impact. The disease affects cattle, swine ...
  130. [130]
    FAO warns: Enhanced awareness and action needed amid foot-and ...
    May 5, 2025 · The economic losses are substantial, with global direct production losses and vaccination costs in endemic regions estimated to be USD 21 ...Missing: per | Show results with:per
  131. [131]
    [PDF] Rabies and Rabies-Related Lyssaviruses
    Importance. Rabies is a severe viral disease that affects the central nervous system (CNS) of mammals. Survival is very rare once the clinical signs develop ...
  132. [132]
    Rabies - World Health Organization (WHO)
    Jun 5, 2024 · Globally there are an estimated 59 000 deaths from rabies annually; however, due to underreporting, documented case numbers often differ from ...WHO's work on rabies · Rabies vaccines · Rabies Vaccine Inactivated...
  133. [133]
    The Virus was First Discovered in Plants | Gardening in the Panhandle
    Apr 15, 2020 · The first virus discovered was in plants, not in humans. As early as 1857, tobacco farmers in the Netherlands recognized a new disease of tobacco.<|separator|>
  134. [134]
    Estimation of losses caused by Potato virus Y in ... - SciELO México
    Yield losses caused by PVY in Fianna potato crops in the experimental plot ranged from 9.4% to 53%, while the estimated losses in the Coahuila region during the ...
  135. [135]
    Aphid Transmission of Potyvirus: The Largest Plant-Infecting RNA ...
    Jul 17, 2020 · Potyviruses are the largest group of plant infecting RNA viruses that cause significant losses in a wide range of crops across the globe.
  136. [136]
    Understanding the roles of viruses as key players in environmental ...
    Apr 17, 2025 · Viruses control microbial community composition thereby playing crucial roles in regulating primary productivity and biogeochemical processes in an ecosystem.
  137. [137]
    Arbovirus-Mosquito Vector-Host Interactions and the Impact on ...
    Jan 23, 2019 · This review focuses on mosquito-borne arboviruses and discusses current knowledge of the factors and underlying mechanisms that influence infection and ...
  138. [138]
    Diversity and ecological footprint of Global Ocean RNA viruses
    Jun 9, 2022 · In this study, we uncover patterns and predictors of marine RNA virus community- and “species”-level diversity and contextualize their ecological impacts from ...
  139. [139]
    Nipah virus, an emerging zoonotic disease causing fatal encephalitis
    Nipah virus is an acute febrile illness that can cause fatal encephalitis. It is an emerging zoonotic paramyxovirus endemic to south-east Asia and the western ...
  140. [140]
    Wildmeat consumption and zoonotic spillover - ScienceDirect.com
    Zoonotic diseases are estimated to constitute 75% of all emerging infectious diseases, of which more than 70% come from wild species.