Messenger RNA

Last updated
The "life cycle" of an mRNA in a eukaryotic cell. RNA is transcribed in the nucleus; after processing, it is transported to the cytoplasm and translated by the ribosome. Finally, the mRNA is degraded. MRNA-interaction.svg
The "life cycle" of an mRNA in a eukaryotic cell. RNA is transcribed in the nucleus; after processing, it is transported to the cytoplasm and translated by the ribosome. Finally, the mRNA is degraded.

In molecular biology, messenger ribonucleic acid (mRNA) is a single-stranded molecule of RNA that corresponds to the genetic sequence of a gene, and is read by a ribosome in the process of synthesizing a protein.

Contents

mRNA is created during the process of transcription, where an enzyme (RNA polymerase) converts the gene into primary transcript mRNA (also known as pre-mRNA). This pre-mRNA usually still contains introns, regions that will not go on to code for the final amino acid sequence. These are removed in the process of RNA splicing, leaving only exons, regions that will encode the protein. This exon sequence constitutes mature mRNA. Mature mRNA is then read by the ribosome, and the ribosome creates the protein utilizing amino acids carried by transfer RNA (tRNA). This process is known as translation. All of these processes form part of the central dogma of molecular biology, which describes the flow of genetic information in a biological system.

As in DNA, genetic information in mRNA is contained in the sequence of nucleotides, which are arranged into codons consisting of three ribonucleotides each. Each codon codes for a specific amino acid, except the stop codons, which terminate protein synthesis. The translation of codons into amino acids requires two other types of RNA: transfer RNA, which recognizes the codon and provides the corresponding amino acid, and ribosomal RNA (rRNA), the central component of the ribosome's protein-manufacturing machinery.

The concept of mRNA was developed by Sydney Brenner and Francis Crick in 1960 during a conversation with François Jacob. In 1961, mRNA was identified and described independently by one team consisting of Brenner, Jacob, and Matthew Meselson, and another team led by James Watson. While analyzing the data in preparation for publication, Jacob and Jacques Monod coined the name "messenger RNA".

Synthesis

RNA polymerase transcribes a DNA strand to form mRNA DNA transcription.svg
RNA polymerase transcribes a DNA strand to form mRNA

The brief existence of an mRNA molecule begins with transcription, and ultimately ends in degradation. During its life, an mRNA molecule may also be processed, edited, and transported prior to translation. Eukaryotic mRNA molecules often require extensive processing and transport, while prokaryotic mRNA molecules do not. A molecule of eukaryotic mRNA and the proteins surrounding it are together called a messenger RNP.[ citation needed ]

Transcription

Transcription is when RNA is copied from DNA. During transcription, RNA polymerase makes a copy of a gene from the DNA to mRNA as needed. This process differs slightly in eukaryotes and prokaryotes. One notable difference is that prokaryotic RNA polymerase associates with DNA-processing enzymes during transcription so that processing can proceed during transcription. Therefore, this causes the new mRNA strand to become double stranded by producing a complementary strand known as the tRNA strand, which when combined are unable to form structures from base-pairing. Moreover, the template for mRNA is the complementary strand of tRNA, which is identical in sequence to the anticodon sequence that the DNA binds to. The short-lived, unprocessed or partially processed product is termed precursor mRNA, or pre-mRNA ; once completely processed, it is termed mature mRNA .[ citation needed ]

Uracil substitution for thymine

mRNA uses uracil (U) instead of thymine (T) in DNA. uracil (U) is the complimentary base to adenine (A) during transcription instead of thymine (T). So when using a template strand of DNA to build RNA, thymine is replaced with uracil. This substitution allows the mRNA to carry the appropriate genetic information from DNA to the ribosome for translation. Regarding the natural history, uracil came first than thymine; evidence suggests that RNA came before DNA in evolution. [1] The RNA World hypothesis proposes that life began with RNA molecules, before the emergence of DNA genomes and coded proteins. In DNA, the evolutionary substitution of thymine for uracil may have increased DNA stability and improved the efficiency of DNA replication. [2] [3]

Eukaryotic pre-mRNA processing

DNA gene is transcribed to pre-mRNA, which is then processed to form a mature mRNA, and then lastly translated by a ribosome to a protein Gene structure eukaryote 2 annotated.svg
DNA gene is transcribed to pre-mRNA, which is then processed to form a mature mRNA, and then lastly translated by a ribosome to a protein

Processing of mRNA differs greatly among eukaryotes, bacteria, and archaea. Non-eukaryotic mRNA is, in essence, mature upon transcription and requires no processing, except in rare cases. [4] Eukaryotic pre-mRNA, however, requires several processing steps before its transport to the cytoplasm and its translation by the ribosome.

Splicing

The extensive processing of eukaryotic pre-mRNA that leads to the mature mRNA is the RNA splicing, a mechanism by which introns or outrons (non-coding regions) are removed and exons (coding regions) are joined.[ citation needed ]

5' cap addition

5' cap structure 5' cap labeled.svg
5' cap structure

A 5' cap (also termed an RNA cap, an RNA 7-methylguanosine cap, or an RNA m7G cap) is a modified guanine nucleotide that has been added to the "front" or 5' end of a eukaryotic messenger RNA shortly after the start of transcription. The 5' cap consists of a terminal 7-methylguanosine residue that is linked through a 5'-5'-triphosphate bond to the first transcribed nucleotide. Its presence is critical for recognition by the ribosome and protection from RNases.[ citation needed ]

Cap addition is coupled to transcription, and occurs co-transcriptionally, such that each influences the other. Shortly after the start of transcription, the 5' end of the mRNA being synthesized is bound by a cap-synthesizing complex associated with RNA polymerase. This enzymatic complex catalyzes the chemical reactions that are required for mRNA capping. Synthesis proceeds as a multi-step biochemical reaction.[ citation needed ]

Editing

In some instances, an mRNA will be edited, changing the nucleotide composition of that mRNA. An example in humans is the apolipoprotein B mRNA, which is edited in some tissues, but not others. The editing creates an early stop codon, which, upon translation, produces a shorter protein.

Polyadenylation

Polyadenylation Polyadenylation.png
Polyadenylation

Polyadenylation is the covalent linkage of a polyadenylyl moiety to a messenger RNA molecule. In eukaryotic organisms most messenger RNA (mRNA) molecules are polyadenylated at the 3' end, but recent studies have shown that short stretches of uridine (oligouridylation) are also common. [5] The poly(A) tail and the protein bound to it aid in protecting mRNA from degradation by exonucleases. Polyadenylation is also important for transcription termination, export of the mRNA from the nucleus, and translation. mRNA can also be polyadenylated in prokaryotic organisms, where poly(A) tails act to facilitate, rather than impede, exonucleolytic degradation.[ citation needed ]

Polyadenylation occurs during and/or immediately after transcription of DNA into RNA. After transcription has been terminated, the mRNA chain is cleaved through the action of an endonuclease complex associated with RNA polymerase. After the mRNA has been cleaved, around 250 adenosine residues are added to the free 3' end at the cleavage site. This reaction is catalyzed by polyadenylate polymerase. Just as in alternative splicing, there can be more than one polyadenylation variant of an mRNA.

Polyadenylation site mutations also occur. The primary RNA transcript of a gene is cleaved at the poly-A addition site, and 100–200 A's are added to the 3' end of the RNA. If this site is altered, an abnormally long and unstable mRNA construct will be formed.

Transport

Another difference between eukaryotes and prokaryotes is mRNA transport. Because eukaryotic transcription and translation is compartmentally separated, eukaryotic mRNAs must be exported from the nucleus to the cytoplasm—a process that may be regulated by different signaling pathways. [6] Mature mRNAs are recognized by their processed modifications and then exported through the nuclear pore by binding to the cap-binding proteins CBP20 and CBP80, [7] as well as the transcription/export complex (TREX). [8] [9] Multiple mRNA export pathways have been identified in eukaryotes. [10]

In spatially complex cells, some mRNAs are transported to particular subcellular destinations. In mature neurons, certain mRNA are transported from the soma to dendrites. One site of mRNA translation is at polyribosomes selectively localized beneath synapses. [11] The mRNA for Arc/Arg3.1 is induced by synaptic activity and localizes selectively near active synapses based on signals generated by NMDA receptors. [12] Other mRNAs also move into dendrites in response to external stimuli, such as β-actin mRNA. [13] For export from the nucleus, actin mRNA associates with ZBP1 [14] and later with 40S subunit. The complex is bound by a motor protein and is transported to the target location (neurite extension) along the cytoskeleton. Eventually ZBP1 is phosphorylated by Src in order for translation to be initiated. [15] In developing neurons, mRNAs are also transported into growing axons and especially growth cones. Many mRNAs are marked with so-called "zip codes", which target their transport to a specific location. [16] [17] mRNAs can also transfer between mammalian cells through structures called tunneling nanotubes. [18] [19]

Translation

Translation of mRNA to protein Peptide syn.svg
Translation of mRNA to protein

Because prokaryotic mRNA does not need to be processed or transported, translation by the ribosome can begin immediately after the end of transcription. Therefore, it can be said that prokaryotic translation is coupled to transcription and occurs co-transcriptionally.[ citation needed ]

Eukaryotic mRNA that has been processed and transported to the cytoplasm (i.e., mature mRNA) can then be translated by the ribosome. Translation may occur at ribosomes free-floating in the cytoplasm, or directed to the endoplasmic reticulum by the signal recognition particle. Therefore, unlike in prokaryotes, eukaryotic translation is not directly coupled to transcription. It is even possible in some contexts that reduced mRNA levels are accompanied by increased protein levels, as has been observed for mRNA/protein levels of EEF1A1 in breast cancer. [20] [ non-primary source needed ]

Structure

The structure of a mature eukaryotic mRNA. A fully processed mRNA includes a 5' cap, 5' UTR, coding region, 3' UTR, and poly(A) tail. MRNA structure.svg
The structure of a mature eukaryotic mRNA. A fully processed mRNA includes a 5' cap, 5' UTR, coding region, 3' UTR, and poly(A) tail.

Coding regions

Coding regions are composed of codons, which are decoded and translated into proteins by the ribosome; in eukaryotes usually into one and in prokaryotes usually into several. Coding regions begin with the start codon and end with a stop codon. In general, the start codon is an AUG triplet and the stop codon is UAG ("amber"), UAA ("ochre"), or UGA ("opal"). The coding regions tend to be stabilised by internal base pairs; this impedes degradation. [21] [22] In addition to being protein-coding, portions of coding regions may serve as regulatory sequences in the pre-mRNA as exonic splicing enhancers or exonic splicing silencers.

Untranslated regions

Universal structure of eukaryotic mRNA, showing the structure of the 5' and 3' UTRs. Fbioe-09-718753-g002.jpg
Universal structure of eukaryotic mRNA, showing the structure of the 5' and 3' UTRs.

Untranslated regions (UTRs) are sections of the mRNA before the start codon and after the stop codon that are not translated, termed the five prime untranslated region (5' UTR) and three prime untranslated region (3' UTR), respectively. These regions are transcribed with the coding region and thus are exonic as they are present in the mature mRNA. Several roles in gene expression have been attributed to the untranslated regions, including mRNA stability, mRNA localization, and translational efficiency. The ability of a UTR to perform these functions depends on the sequence of the UTR and can differ between mRNAs. Genetic variants in 3' UTR have also been implicated in disease susceptibility because of the change in RNA structure and protein translation. [23]

The stability of mRNAs may be controlled by the 5' UTR and/or 3' UTR due to varying affinity for RNA degrading enzymes called ribonucleases and for ancillary proteins that can promote or inhibit RNA degradation. (See also, C-rich stability element.)

Translational efficiency, including sometimes the complete inhibition of translation, can be controlled by UTRs. Proteins that bind to either the 3' or 5' UTR may affect translation by influencing the ribosome's ability to bind to the mRNA. MicroRNAs bound to the 3' UTR also may affect translational efficiency or mRNA stability.

Cytoplasmic localization of mRNA is thought to be a function of the 3' UTR. Proteins that are needed in a particular region of the cell can also be translated there; in such a case, the 3' UTR may contain sequences that allow the transcript to be localized to this region for translation.

Some of the elements contained in untranslated regions form a characteristic secondary structure when transcribed into RNA. These structural mRNA elements are involved in regulating the mRNA. Some, such as the SECIS element, are targets for proteins to bind. One class of mRNA element, the riboswitches, directly bind small molecules, changing their fold to modify levels of transcription or translation. In these cases, the mRNA regulates itself.

Poly(A) tail

The 3' poly(A) tail is a long sequence of adenine nucleotides (often several hundred) added to the 3' end of the pre-mRNA. This tail promotes export from the nucleus and translation, and protects the mRNA from degradation.

Monocistronic versus polycistronic mRNA

An mRNA molecule is said to be monocistronic when it contains the genetic information to translate only a single protein chain (polypeptide). This is the case for most of the eukaryotic mRNAs. [24] [25] On the other hand, polycistronic mRNA carries several open reading frames (ORFs), each of which is translated into a polypeptide. These polypeptides usually have a related function (they often are the subunits composing a final complex protein) and their coding sequence is grouped and regulated together in a regulatory region, containing a promoter and an operator. Most of the mRNA found in bacteria and archaea is polycistronic, [24] as is the human mitochondrial genome. [26] Dicistronic or bicistronic mRNA encodes only two proteins.

mRNA circularization

mRNA circularisation and regulation Fgene-10-00006-g001.jpg
mRNA circularisation and regulation

In eukaryotes mRNA molecules form circular structures due to an interaction between the eIF4E and poly(A)-binding protein, which both bind to eIF4G, forming an mRNA-protein-mRNA bridge. [27] Circularization is thought to promote cycling of ribosomes on the mRNA leading to time-efficient translation, and may also function to ensure only intact mRNA are translated (partially degraded mRNA characteristically have no m7G cap, or no poly-A tail). [28]

Other mechanisms for circularization exist, particularly in virus mRNA. Poliovirus mRNA uses a cloverleaf section towards its 5' end to bind PCBP2, which binds poly(A)-binding protein, forming the familiar mRNA-protein-mRNA circle. Barley yellow dwarf virus has binding between mRNA segments on its 5' end and 3' end (called kissing stem loops), circularizing the mRNA without any proteins involved.

RNA virus genomes (the + strands of which are translated as mRNA) are also commonly circularized. [29] During genome replication the circularization acts to enhance genome replication speeds, cycling viral RNA-dependent RNA polymerase much the same as the ribosome is hypothesized to cycle.

Degradation

Different mRNAs within the same cell have distinct lifetimes (stabilities). In bacterial cells, individual mRNAs can survive from seconds to more than an hour. However, the lifetime averages between 1 and 3 minutes, making bacterial mRNA much less stable than eukaryotic mRNA. [30] In mammalian cells, mRNA lifetimes range from several minutes to days. [31] The greater the stability of an mRNA the more protein may be produced from that mRNA. The limited lifetime of mRNA enables a cell to alter protein synthesis rapidly in response to its changing needs. There are many mechanisms that lead to the destruction of an mRNA, some of which are described below.

Prokaryotic mRNA degradation

Overview of mRNA decay pathways in the different life domains. 1-s2.0-S1874939913000436-gr1 lrg.jpg
Overview of mRNA decay pathways in the different life domains.

In general, in prokaryotes the lifetime of mRNA is much shorter than in eukaryotes. Prokaryotes degrade messages by using a combination of ribonucleases, including endonucleases, 3' exonucleases, and 5' exonucleases. In some instances, small RNA molecules (sRNA) tens to hundreds of nucleotides long can stimulate the degradation of specific mRNAs by base-pairing with complementary sequences and facilitating ribonuclease cleavage by RNase III. It was recently shown that bacteria also have a sort of 5' cap consisting of a triphosphate on the 5' end. [32] Removal of two of the phosphates leaves a 5' monophosphate, causing the message to be destroyed by the exonuclease RNase J, which degrades 5' to 3'.

Eukaryotic mRNA turnover

Inside eukaryotic cells, there is a balance between the processes of translation and mRNA decay. Messages that are being actively translated are bound by ribosomes, the eukaryotic initiation factors eIF-4E and eIF-4G, and poly(A)-binding protein. eIF-4E and eIF-4G block the decapping enzyme (DCP2), and poly(A)-binding protein blocks the exosome complex, protecting the ends of the message. The balance between translation and decay is reflected in the size and abundance of cytoplasmic structures known as P-bodies. [33] The poly(A) tail of the mRNA is shortened by specialized exonucleases that are targeted to specific messenger RNAs by a combination of cis-regulatory sequences on the RNA and trans-acting RNA-binding proteins. Poly(A) tail removal is thought to disrupt the circular structure of the message and destabilize the cap binding complex. The message is then subject to degradation by either the exosome complex or the decapping complex. In this way, translationally inactive messages can be destroyed quickly, while active messages remain intact. The mechanism by which translation stops and the message is handed-off to decay complexes is not understood in detail.

AU-rich element decay

The presence of AU-rich elements in some mammalian mRNAs tends to destabilize those transcripts through the action of cellular proteins that bind these sequences and stimulate poly(A) tail removal. Loss of the poly(A) tail is thought to promote mRNA degradation by facilitating attack by both the exosome complex [34] and the decapping complex. [35] Rapid mRNA degradation via AU-rich elements is a critical mechanism for preventing the overproduction of potent cytokines such as tumor necrosis factor (TNF) and granulocyte-macrophage colony stimulating factor (GM-CSF). [36] AU-rich elements also regulate the biosynthesis of proto-oncogenic transcription factors like c-Jun and c-Fos. [37]

Nonsense-mediated decay

Eukaryotic messages are subject to surveillance by nonsense-mediated decay (NMD), which checks for the presence of premature stop codons (nonsense codons) in the message. These can arise via incomplete splicing, V(D)J recombination in the adaptive immune system, mutations in DNA, transcription errors, leaky scanning by the ribosome causing a frame shift, and other causes. Detection of a premature stop codon triggers mRNA degradation by 5' decapping, 3' poly(A) tail removal, or endonucleolytic cleavage. [38]

Small interfering RNA (siRNA)

In metazoans, small interfering RNAs (siRNAs) processed by Dicer are incorporated into a complex known as the RNA-induced silencing complex or RISC. This complex contains an endonuclease that cleaves perfectly complementary messages to which the siRNA binds. The resulting mRNA fragments are then destroyed by exonucleases. siRNA is commonly used in laboratories to block the function of genes in cell culture. It is thought to be part of the innate immune system as a defense against double-stranded RNA viruses. [39]

MicroRNA (miRNA)

MicroRNAs (miRNAs) are small RNAs that typically are partially complementary to sequences in metazoan messenger RNAs. [40] [41] Binding of a miRNA to a message can repress translation of that message and accelerate poly(A) tail removal, thereby hastening mRNA degradation. The mechanism of action of miRNAs is the subject of active research. [42] [43]

Other decay mechanisms

There are other ways by which messages can be degraded, including non-stop decay and silencing by Piwi-interacting RNA (piRNA), among others.

Applications

The administration of a nucleoside-modified messenger RNA sequence can cause a cell to make a protein, which in turn could directly treat a disease or could function as a vaccine; more indirectly the protein could drive an endogenous stem cell to differentiate in a desired way. [44] [45]

The primary challenges of RNA therapy center on delivering the RNA to the appropriate cells. [46] Challenges include the fact that naked RNA sequences naturally degrade after preparation; they may trigger the body's immune system to attack them as an invader; and they are impermeable to the cell membrane. [45] Once within the cell, they must then leave the cell's transport mechanism to take action within the cytoplasm, which houses the necessary ribosomes. [44]

Overcoming these challenges, mRNA as a therapeutic was first put forward in 1989 "after the development of a broadly applicable in vitro transfection technique." [47] In the 1990s, mRNA vaccines for personalized cancer have been developed, relying on non-nucleoside modified mRNA. mRNA based therapies continue to be investigated as a method of treatment or therapy for both cancer as well as auto-immune, metabolic, and respiratory inflammatory diseases. Gene editing therapies such as CRISPR may also benefit from using mRNA to induce cells to make the desired Cas protein. [48]

Since the 2010s, RNA vaccines and other RNA therapeutics have been considered to be "a new class of drugs". [49] The first mRNA-based vaccines received restricted authorization and were rolled out across the world during the COVID-19 pandemic by Pfizer–BioNTech COVID-19 vaccine and Moderna, for example. [50] The 2023 Nobel Prize in Physiology or Medicine was awarded to Katalin Karikó and Drew Weissman for the development of effective mRNA vaccines against COVID-19. [51] [52] [53]

History

Several molecular biology studies during the 1950s indicated that RNA played some kind of role in protein synthesis, but that role was not clearly understood. For instance, in one of the earliest reports, Jacques Monod and his team showed that RNA synthesis was necessary for protein synthesis, specifically during the production of the enzyme β-galactosidase in the bacterium E. coli . [54] Arthur Pardee also found similar RNA accumulation in 1954. [55] In 1953, Alfred Hershey, June Dixon, and Martha Chase described a certain cytosine-containing DNA (indicating it was RNA) that disappeared quickly after its synthesis in E. coli. [56] In hindsight, this may have been one of the first observations of the existence of mRNA but it was not recognized at the time as such. [57]

The idea of mRNA was first conceived by Sydney Brenner and Francis Crick on 15 April 1960 at King's College, Cambridge, while François Jacob was telling them about a recent experiment conducted by Arthur Pardee, himself, and Monod (the so-called PaJaMo experiment, which did not prove mRNA existed but suggested the possibility of its existence). With Crick's encouragement, Brenner and Jacob immediately set out to test this new hypothesis, and they contacted Matthew Meselson at the California Institute of Technology for assistance. During the summer of 1960, Brenner, Jacob, and Meselson conducted an experiment in Meselson's laboratory at Caltech which was the first to prove the existence of mRNA. That fall, Jacob and Monod coined the name "messenger RNA" and developed the first theoretical framework to explain its function. [57]

In February 1961, James Watson revealed that his Harvard-based research group had been right behind them with a series of experiments whose results pointed in roughly the same direction. Brenner and the others agreed to Watson's request to delay publication of their research findings. As a result, the Brenner and Watson articles were published simultaneously in the same issue of Nature in May 1961, while that same month, Jacob and Monod published their theoretical framework for mRNA in the Journal of Molecular Biology . [57]

See also

Related Research Articles

<span class="mw-page-title-main">Protein biosynthesis</span> Assembly of proteins inside biological cells

Protein biosynthesis is a core biological process, occurring inside cells, balancing the loss of cellular proteins through the production of new proteins. Proteins perform a number of critical functions as enzymes, structural proteins or hormones. Protein synthesis is a very similar process for both prokaryotes and eukaryotes but there are some distinct differences.

<span class="mw-page-title-main">Ribosome</span> Synthesizes proteins in cells

Ribosomes are macromolecular machines, found within all cells, that perform biological protein synthesis. Ribosomes link amino acids together in the order specified by the codons of messenger RNA molecules to form polypeptide chains. Ribosomes consist of two major components: the small and large ribosomal subunits. Each subunit consists of one or more ribosomal RNA molecules and many ribosomal proteins. The ribosomes and associated molecules are also known as the translational apparatus.

The central dogma of molecular biology deals with the flow of genetic information within a biological system. It is often stated as "DNA makes RNA, and RNA makes protein", although this is not its original meaning. It was first stated by Francis Crick in 1957, then published in 1958:

The Central Dogma. This states that once "information" has passed into protein it cannot get out again. In more detail, the transfer of information from nucleic acid to nucleic acid, or from nucleic acid to protein may be possible, but transfer from protein to protein, or from protein to nucleic acid is impossible. Information here means the precise determination of sequence, either of bases in the nucleic acid or of amino acid residues in the protein.

<span class="mw-page-title-main">Gene expression</span> Conversion of a genes sequence into a mature gene product or products

Gene expression is the process by which information from a gene is used in the synthesis of a functional gene product that enables it to produce end products, proteins or non-coding RNA, and ultimately affect a phenotype. These products are often proteins, but in non-protein-coding genes such as transfer RNA (tRNA) and small nuclear RNA (snRNA), the product is a functional non-coding RNA. The process of gene expression is used by all known life—eukaryotes, prokaryotes, and utilized by viruses—to generate the macromolecular machinery for life.

<span class="mw-page-title-main">Translation (biology)</span> Cellular process of protein synthesis

In biology, translation is the process in living cells in which proteins are produced using RNA molecules as templates. The generated protein is a sequence of amino acids. This sequence is determined by the sequence of nucleotides in the RNA. The nucleotides are considered three at a time. Each such triple results in addition of one specific amino acid to the protein being generated. The matching from nucleotide triple to amino acid is called the genetic code. The translation is performed by a large complex of functional RNA and proteins called ribosomes. The entire process is called gene expression.

<span class="mw-page-title-main">Three prime untranslated region</span> Sequence at the 3 end of messenger RNA that does not code for product

In molecular genetics, the three prime untranslated region (3′-UTR) is the section of messenger RNA (mRNA) that immediately follows the translation termination codon. The 3′-UTR often contains regulatory regions that post-transcriptionally influence gene expression.

Polyadenylation is the addition of a poly(A) tail to an RNA transcript, typically a messenger RNA (mRNA). The poly(A) tail consists of multiple adenosine monophosphates; in other words, it is a stretch of RNA that has only adenine bases. In eukaryotes, polyadenylation is part of the process that produces mature mRNA for translation. In many bacteria, the poly(A) tail promotes degradation of the mRNA. It, therefore, forms part of the larger process of gene expression.

In molecular biology, the five-prime cap is a specially altered nucleotide on the 5′ end of some primary transcripts such as precursor messenger RNA. This process, known as mRNA capping, is highly regulated and vital in the creation of stable and mature messenger RNA able to undergo translation during protein synthesis. Mitochondrial mRNA and chloroplastic mRNA are not capped.

The 5′ untranslated region is the region of a messenger RNA (mRNA) that is directly upstream from the initiation codon. This region is important for the regulation of translation of a transcript by differing mechanisms in viruses, prokaryotes and eukaryotes. While called untranslated, the 5′ UTR or a portion of it is sometimes translated into a protein product. This product can then regulate the translation of the main coding sequence of the mRNA. In many organisms, however, the 5′ UTR is completely untranslated, instead forming a complex secondary structure to regulate translation.

<span class="mw-page-title-main">Ribosomal RNA</span> RNA component of the ribosome, essential for protein synthesis in all living organisms

Ribosomal ribonucleic acid (rRNA) is a type of non-coding RNA which is the primary component of ribosomes, essential to all cells. rRNA is a ribozyme which carries out protein synthesis in ribosomes. Ribosomal RNA is transcribed from ribosomal DNA (rDNA) and then bound to ribosomal proteins to form small and large ribosome subunits. rRNA is the physical and mechanical factor of the ribosome that forces transfer RNA (tRNA) and messenger RNA (mRNA) to process and translate the latter into proteins. Ribosomal RNA is the predominant form of RNA found in most cells; it makes up about 80% of cellular RNA despite never being translated into proteins itself. Ribosomes are composed of approximately 60% rRNA and 40% ribosomal proteins, though this ratio differs between prokaryotes and eukaryotes.

Bacterial translation is the process by which messenger RNA is translated into proteins in bacteria.

Eukaryotic translation is the biological process by which messenger RNA is translated into proteins in eukaryotes. It consists of four phases: initiation, elongation, termination, and recapping.

The Kozak consensus sequence is a nucleic acid motif that functions as the protein translation initiation site in most eukaryotic mRNA transcripts. Regarded as the optimum sequence for initiating translation in eukaryotes, the sequence is an integral aspect of protein regulation and overall cellular health as well as having implications in human disease. It ensures that a protein is correctly translated from the genetic message, mediating ribosome assembly and translation initiation. A wrong start site can result in non-functional proteins. As it has become more studied, expansions of the nucleotide sequence, bases of importance, and notable exceptions have arisen. The sequence was named after the scientist who discovered it, Marilyn Kozak. Kozak discovered the sequence through a detailed analysis of DNA genomic sequences.

Gene structure is the organisation of specialised sequence elements within a gene. Genes contain most of the information necessary for living cells to survive and reproduce. In most organisms, genes are made of DNA, where the particular DNA sequence determines the function of the gene. A gene is transcribed (copied) from DNA into RNA, which can either be non-coding (ncRNA) with a direct function, or an intermediate messenger (mRNA) that is then translated into protein. Each of these steps is controlled by specific sequence elements, or regions, within the gene. Every gene, therefore, requires multiple sequence elements to be functional. This includes the sequence that actually encodes the functional protein or ncRNA, as well as multiple regulatory sequence regions. These regions may be as short as a few base pairs, up to many thousands of base pairs long.

<span class="mw-page-title-main">Directionality (molecular biology)</span> End-to-end chemical orientation of a single strand of nucleic acid

Directionality, in molecular biology and biochemistry, is the end-to-end chemical orientation of a single strand of nucleic acid. In a single strand of DNA or RNA, the chemical convention of naming carbon atoms in the nucleotide pentose-sugar-ring means that there will be a 5′ end, which frequently contains a phosphate group attached to the 5′ carbon of the ribose ring, and a 3′ end, which typically is unmodified from the ribose -OH substituent. In a DNA double helix, the strands run in opposite directions to permit base pairing between them, which is essential for replication or transcription of the encoded information.

A ribosome binding site, or ribosomal binding site (RBS), is a sequence of nucleotides upstream of the start codon of an mRNA transcript that is responsible for the recruitment of a ribosome during the initiation of translation. Mostly, RBS refers to bacterial sequences, although internal ribosome entry sites (IRES) have been described in mRNAs of eukaryotic cells or viruses that infect eukaryotes. Ribosome recruitment in eukaryotes is generally mediated by the 5' cap present on eukaryotic mRNAs.

<span class="mw-page-title-main">Non-stop decay</span>

Non-stop decay (NSD) is a cellular mechanism of mRNA surveillance to detect mRNA molecules lacking a stop codon and prevent these mRNAs from translation. The non-stop decay pathway releases ribosomes that have reached the far 3' end of an mRNA and guides the mRNA to the exosome complex, or to RNase R in bacteria for selective degradation. In contrast to nonsense-mediated decay (NMD), polypeptides do not release from the ribosome, and thus, NSD seems to involve mRNA decay factors distinct from NMD.

<span class="mw-page-title-main">Untranslated region</span> Non-coding regions on either end of mRNA

In molecular genetics, an untranslated region refers to either of two sections, one on each side of a coding sequence on a strand of mRNA. If it is found on the 5' side, it is called the 5' UTR, or if it is found on the 3' side, it is called the 3' UTR. mRNA is RNA that carries information from DNA to the ribosome, the site of protein synthesis (translation) within a cell. The mRNA is initially transcribed from the corresponding DNA sequence and then translated into protein. However, several regions of the mRNA are usually not translated into protein, including the 5' and 3' UTRs.

mRNA surveillance mechanisms are pathways utilized by organisms to ensure fidelity and quality of messenger RNA (mRNA) molecules. There are a number of surveillance mechanisms present within cells. These mechanisms function at various steps of the mRNA biogenesis pathway to detect and degrade transcripts that have not properly been processed.

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

In molecular biology, a riboregulator is a ribonucleic acid (RNA) that responds to a signal nucleic acid molecule by Watson-Crick base pairing. A riboregulator may respond to a signal molecule in any number of manners including, translation of the RNA into a protein, activation of a ribozyme, release of silencing RNA (siRNA), conformational change, and/or binding other nucleic acids. Riboregulators contain two canonical domains, a sensor domain and an effector domain. These domains are also found on riboswitches, but unlike riboswitches, the sensor domain only binds complementary RNA or DNA strands as opposed to small molecules. Because binding is based on base-pairing, a riboregulator can be tailored to differentiate and respond to individual genetic sequences and combinations thereof.

References

  1. "The Information in DNA Is Decoded by Transcription | Learn Science at Scitable". www.nature.com. Retrieved 2024-05-03.
  2. "RNA world (article) | Natural selection". Khan Academy. Retrieved 2024-05-03.
  3. "The presence of thymine the place of uracil also confers additional stability to DNA. How?". Toppr Ask. Retrieved 2024-05-04.
  4. Watson JD (February 22, 2013). Molecular Biology of the Gene, 7th edition. Pearson Higher Ed USA. ISBN   9780321851499.
  5. Choi YS, Patena W, Leavitt AD, McManus MT (March 2012). "Widespread RNA 3'-end oligouridylation in mammals". RNA. 18 (3): 394–401. doi:10.1261/rna.029306.111. PMC   3285928 . PMID   22291204.
  6. Quaresma AJ, Sievert R, Nickerson JA (April 2013). "Regulation of mRNA export by the PI3 kinase/AKT signal transduction pathway". Molecular Biology of the Cell. 24 (8): 1208–1221. doi:10.1091/mbc.E12-06-0450. PMC   3623641 . PMID   23427269.
  7. Kierzkowski D, Kmieciak M, Piontek P, Wojtaszek P, Szweykowska-Kulinska Z, Jarmolowski A (September 2009). "The Arabidopsis CBP20 targets the cap-binding complex to the nucleus, and is stabilized by CBP80". The Plant Journal. 59 (5): 814–825. doi: 10.1111/j.1365-313X.2009.03915.x . PMID   19453442.
  8. Strässer K, Masuda S, Mason P, Pfannstiel J, Oppizzi M, Rodriguez-Navarro S, Rondón AG, Aguilera A, Struhl K, Reed R, Hurt E (May 2002). "TREX is a conserved complex coupling transcription with messenger RNA export". Nature. 417 (6886): 304–308. Bibcode:2002Natur.417..304S. doi:10.1038/nature746. PMID   11979277. S2CID   1112194.
  9. Katahira J, Yoneda Y (27 October 2014). "Roles of the TREX complex in nuclear export of mRNA". RNA Biology. 6 (2): 149–152. doi: 10.4161/rna.6.2.8046 . PMID   19229134.
  10. Cenik C, Chua HN, Zhang H, Tarnawsky SP, Akef A, Derti A, Tasan M, Moore MJ, Palazzo AF, Roth FP (April 2011). "Genome analysis reveals interplay between 5'UTR introns and nuclear mRNA export for secretory and mitochondrial genes". PLOS Genetics. 7 (4): e1001366. doi: 10.1371/journal.pgen.1001366 . PMC   3077370 . PMID   21533221.
  11. Steward O, Levy WB (March 1982). "Preferential localization of polyribosomes under the base of dendritic spines in granule cells of the dentate gyrus". The Journal of Neuroscience. 2 (3): 284–291. doi:10.1523/JNEUROSCI.02-03-00284.1982. PMC   6564334 . PMID   7062109.
  12. Steward O, Worley PF (April 2001). "Selective targeting of newly synthesized Arc mRNA to active synapses requires NMDA receptor activation". Neuron. 30 (1): 227–240. doi: 10.1016/s0896-6273(01)00275-6 . PMID   11343657. S2CID   13395819.
  13. Job C, Eberwine J (December 2001). "Localization and translation of mRNA in dendrites and axons". Nature Reviews. Neuroscience. 2 (12): 889–898. doi:10.1038/35104069. PMID   11733796. S2CID   5275219.
  14. Oleynikov Y, Singer RH (February 2003). "Real-time visualization of ZBP1 association with beta-actin mRNA during transcription and localization". Current Biology. 13 (3): 199–207. Bibcode:2003CBio...13..199O. doi:10.1016/s0960-9822(03)00044-7. PMC   4765734 . PMID   12573215.
  15. Hüttelmaier S, Zenklusen D, Lederer M, Dictenberg J, Lorenz M, Meng X, et al. (November 2005). "Spatial regulation of beta-actin translation by Src-dependent phosphorylation of ZBP1". Nature. 438 (7067): 512–515. Bibcode:2005Natur.438..512H. doi:10.1038/nature04115. PMID   16306994. S2CID   2453397.
  16. Oleynikov Y, Singer RH (October 1998). "RNA localization: different zipcodes, same postman?". Trends in Cell Biology. 8 (10): 381–383. doi:10.1016/s0962-8924(98)01348-8. PMC   2136761 . PMID   9789325.
  17. Ainger K, Avossa D, Diana AS, Barry C, Barbarese E, Carson JH (September 1997). "Transport and localization elements in myelin basic protein mRNA". The Journal of Cell Biology. 138 (5): 1077–1087. doi:10.1083/jcb.138.5.1077. PMC   2136761 . PMID   9281585.
  18. Haimovich G, Ecker CM, Dunagin MC, Eggan E, Raj A, Gerst JE, Singer RH (November 2017). "Intercellular mRNA trafficking via membrane nanotube-like extensions in mammalian cells". Proceedings of the National Academy of Sciences of the United States of America. 114 (46): E9873–E9882. Bibcode:2017PNAS..114E9873H. doi: 10.1073/pnas.1706365114 . PMC   5699038 . PMID   29078295.
  19. Haimovich G, Dasgupta S, Gerst JE (February 2021). "RNA transfer through tunneling nanotubes". Biochemical Society Transactions. 49 (1): 145–160. doi:10.1042/BST20200113. PMID   33367488. S2CID   229689880.
  20. Lin CY, Beattie A, Baradaran B, Dray E, Duijf PH (September 2018). "Contradictory mRNA and protein misexpression of EEF1A1 in ductal breast carcinoma due to cell cycle regulation and cellular stress". Scientific Reports. 8 (1): 13904. Bibcode:2018NatSR...813904L. doi:10.1038/s41598-018-32272-x. PMC   6141510 . PMID   30224719.
  21. Shabalina SA, Ogurtsov AY, Spiridonov NA (2006). "A periodic pattern of mRNA secondary structure created by the genetic code". Nucleic Acids Research. 34 (8): 2428–2437. doi:10.1093/nar/gkl287. PMC   1458515 . PMID   16682450.
  22. Katz L, Burge CB (September 2003). "Widespread selection for local RNA secondary structure in coding regions of bacterial genes". Genome Research. 13 (9): 2042–2051. doi:10.1101/gr.1257503. PMC   403678 . PMID   12952875.
  23. Lu YF, Mauger DM, Goldstein DB, Urban TJ, Weeks KM, Bradrick SS (November 2015). "IFNL3 mRNA structure is remodeled by a functional non-coding polymorphism associated with hepatitis C virus clearance". Scientific Reports. 5: 16037. Bibcode:2015NatSR...516037L. doi:10.1038/srep16037. PMC   4631997 . PMID   26531896.
  24. 1 2 Kozak M (March 1983). "Comparison of initiation of protein synthesis in procaryotes, eucaryotes, and organelles". Microbiological Reviews. 47 (1): 1–45. doi:10.1128/MMBR.47.1.1-45.1983. PMC   281560 . PMID   6343825.
  25. Niehrs C, Pollet N (December 1999). "Synexpression groups in eukaryotes". Nature. 402 (6761): 483–487. Bibcode:1999Natur.402..483N. doi:10.1038/990025. PMID   10591207. S2CID   4349134.
  26. Mercer TR, Neph S, Dinger ME, Crawford J, Smith MA, Shearwood AM, Haugen E, Bracken CP, Rackham O, Stamatoyannopoulos JA, Filipovska A, Mattick JS (August 2011). "The human mitochondrial transcriptome". Cell. 146 (4): 645–658. doi:10.1016/j.cell.2011.06.051. PMC   3160626 . PMID   21854988.
  27. Wells SE, Hillner PE, Vale RD, Sachs AB (July 1998). "Circularization of mRNA by eukaryotic translation initiation factors". Molecular Cell. 2 (1): 135–140. CiteSeerX   10.1.1.320.5704 . doi:10.1016/S1097-2765(00)80122-7. PMID   9702200.
  28. López-Lastra M, Rivas A, Barría MI (2005). "Protein synthesis in eukaryotes: the growing biological relevance of cap-independent translation initiation". Biological Research. 38 (2–3): 121–146. doi: 10.4067/S0716-97602005000200003 . PMID   16238092.
  29. Zhang X, Liang Z, Wang C, Shen Z, Sun S, Gong C, Hu X (2022). "Viral Circular RNAs and Their Possible Roles in Virus-Host Interaction". Frontiers in Immunology. 13: 939768. doi: 10.3389/fimmu.2022.939768 . PMC   9247149 . PMID   35784275.
  30. Lewin B, Krebs JE, Kilpatrick ST, Goldstein ES, eds. (2011). Lewin's genes X (10th ed.). Sudbury, Mass.: Jones and Bartlett. ISBN   9780763766320. OCLC   456641931.
  31. Yu J, Russell JE (September 2001). "Structural and functional analysis of an mRNP complex that mediates the high stability of human beta-globin mRNA". Molecular and Cellular Biology. 21 (17): 5879–5888. doi:10.1128/mcb.21.17.5879-5888.2001. PMC   87307 . PMID   11486027.
  32. Deana A, Celesnik H, Belasco JG (January 2008). "The bacterial enzyme RppH triggers messenger RNA degradation by 5' pyrophosphate removal". Nature. 451 (7176): 355–358. Bibcode:2008Natur.451..355D. doi:10.1038/nature06475. PMID   18202662. S2CID   4321451.
  33. Parker R, Sheth U (March 2007). "P bodies and the control of mRNA translation and degradation". Molecular Cell. 25 (5): 635–646. doi: 10.1016/j.molcel.2007.02.011 . PMID   17349952.
  34. Chen CY, Gherzi R, Ong SE, Chan EL, Raijmakers R, Pruijn GJ, Stoecklin G, Moroni C, Mann M, Karin M (November 2001). "AU binding proteins recruit the exosome to degrade ARE-containing mRNAs". Cell. 107 (4): 451–464. doi: 10.1016/S0092-8674(01)00578-5 . PMID   11719186. S2CID   14817671.
  35. Fenger-Grøn M, Fillman C, Norrild B, Lykke-Andersen J (December 2005). "Multiple processing body factors and the ARE binding protein TTP activate mRNA decapping". Molecular Cell. 20 (6): 905–915. doi: 10.1016/j.molcel.2005.10.031 . PMID   16364915.
  36. Shaw G, Kamen R (August 1986). "A conserved AU sequence from the 3' untranslated region of GM-CSF mRNA mediates selective mRNA degradation". Cell. 46 (5): 659–667. doi:10.1016/0092-8674(86)90341-7. PMID   3488815. S2CID   40332253.
  37. Chen CY, Shyu AB (November 1995). "AU-rich elements: characterization and importance in mRNA degradation". Trends in Biochemical Sciences. 20 (11): 465–470. doi:10.1016/S0968-0004(00)89102-1. PMID   8578590.
  38. Isken O, Maquat LE (August 2007). "Quality control of eukaryotic mRNA: safeguarding cells from abnormal mRNA function". Genes & Development. 21 (15): 1833–1856. doi: 10.1101/gad.1566807 . PMID   17671086.
  39. Obbard DJ, Gordon KH, Buck AH, Jiggins FM (January 2009). "The evolution of RNAi as a defence against viruses and transposable elements". Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 364 (1513): 99–115. doi:10.1098/rstb.2008.0168. PMC   2592633 . PMID   18926973.
  40. Robert E. Farrell, Jr. RNA Methodologies, 5th Edition. Academic Press, 2017
  41. Brennecke J, Stark A, Russell RB, Cohen SM (March 2005). "Principles of microRNA-target recognition". PLOS Biology. 3 (3): e85. doi: 10.1371/journal.pbio.0030085 . PMC   1043860 . PMID   15723116.
  42. Tasuku Honjo, Michael Reth, Andreas Radbruch, Frederick Alt. Molecular Biology of B Cells, 2nd Edition. Academic Press, 2014 (including "updated research on microRNAs")
  43. Eulalio A, Huntzinger E, Nishihara T, Rehwinkel J, Fauser M, Izaurralde E (January 2009). "Deadenylation is a widespread effect of miRNA regulation". RNA. 15 (1): 21–32. doi:10.1261/rna.1399509. PMC   2612776 . PMID   19029310.
  44. 1 2 Hajj KA, Whitehead KA (12 September 2017). "Tools for translation: non-viral materials for therapeutic mRNA delivery". Nature Reviews Materials. 2 (10): 17056. Bibcode:2017NatRM...217056H. doi: 10.1038/natrevmats.2017.56 .
  45. 1 2 Gousseinov E, Kozlov M, Scanlan C (September 15, 2015). "RNA-Based Therapeutics and Vaccines". Genetic Engineering News.
  46. Kaczmarek JC, Kowalski PS, Anderson DG (June 2017). "Advances in the delivery of RNA therapeutics: from concept to clinical reality". Genome Medicine. 9 (1): 60. doi: 10.1186/s13073-017-0450-0 . PMC   5485616 . PMID   28655327.
  47. Schlake T, Thess A, Fotin-Mleczek M, Kallen KJ (November 2012). "Developing mRNA-vaccine technologies". RNA Biology. 9 (11): 1319–30. doi:10.4161/rna.22269. PMC   3597572 . PMID   23064118.
  48. Haridi R (2021-04-23). "The mRNA revolution: How COVID-19 hit fast-forward on an experimental technology". New Atlas. Retrieved 2021-04-26.
  49. Kowalska J, Wypijewska del Nogal A, Darzynkiewicz ZM, Buck J, Nicola C, Kuhn AN, Lukaszewicz M, Zuberek J, Strenkowska M, Ziemniak M, Maciejczyk M, Bojarska E, Rhoads RE, Darzynkiewicz E, Sahin U, Jemielity J (2014), "mRNA-based therapeutics–developing a new class of drugs.", Nature Reviews Drug Discovery , vol. 13, no. 10, pp. 759–780, doi: 10.1093/nar/gku757 , PMC   4176373 , PMID   25150148
  50. Barbier AJ, Jiang AY, Zhang P, Wooster R, Anderson DG (June 2022). "The clinical progress of mRNA vaccines and immunotherapies". Nature Biotechnology. 40 (6): 840–854. doi: 10.1038/s41587-022-01294-2 . PMID   35534554. S2CID   248667843.
  51. "The Nobel Prize in Physiology or Medicine 2023". NobelPrize.org. Retrieved 2023-10-03.
  52. "Hungarian and US scientists win Nobel for COVID-19 vaccine discoveries". Reuters. 2023-10-02. Retrieved 2023-10-03.
  53. "The Nobel Prize in Physiology or Medicine 2023". NobelPrize.org. Retrieved 2023-10-03.
  54. Monod J, Pappenheimer AM, Cohen-Bazire G (1952). "La cinétique de la biosynthèse de la β-galactosidase chez E. coli considérée comme fonction de la croissance". Biochimica et Biophysica Acta (in French). 9 (6): 648–660. doi:10.1016/0006-3002(52)90227-8. PMID   13032175.
  55. Pardee AB (May 1954). "Nucleic Acid Precursors and Protein Synthesis". Proceedings of the National Academy of Sciences of the United States of America. 40 (5): 263–270. Bibcode:1954PNAS...40..263P. doi: 10.1073/pnas.40.5.263 . PMC   534118 . PMID   16589470.
  56. Hershey AD, Dixon J, Chase M (July 1953). "Nucleic acid economy in bacteria infected with bacteriophage T2. I. Purine and pyrimidine composition". The Journal of General Physiology. 36 (6): 777–789. doi:10.1085/jgp.36.6.777. PMC   2147416 . PMID   13069681.
  57. 1 2 3 Cobb M (29 June 2015). "Who discovered messenger RNA?". Current Biology. 25 (13): R526–R532. Bibcode:2015CBio...25.R526C. doi: 10.1016/j.cub.2015.05.032 . PMID   26126273.

Further reading