What is the structure and function of nucleic acid?

What is the structure and function of nucleic acid?

In biochemistry and biology, a nucleic acid is a complex chemical substance contained in living cells, especially DNA or RNA, whose molecules comprise numerous nucleotides connected in a long chain.

Discovery of Nucleic acid

  • In the pus of infected wounds, nitrogen-containing molecules have previously been found.
  • Then, a few years later, a Swiss biochemist found these nitrogen-containing substances in the cell nucleus, and he referred to them as nucleic acids.
  • Nucleic acid is a component of every living cell.
  • The brains of a cell are nucleic acids. They are nucleotide polymers.

Structure and Function of Nucleic acid

Nucleotides: Building blocks of nucleic acids

Basic Structure

Polynucleotides, which include nucleic acids, are long chain-like molecules made up of a number of chemical building blocks that are substantially identical to one another. Each nucleotide consists of a nitrogen-containing aromatic base connected to a pentose (five-carbon) sugar, which is in turn attached to a phosphate group. Four of the five nitrogen-containing bases (adenine (A), guanine (G), cytosine (C), thymine (T), and uracil (U)) are present in every nucleic acid (U). A and G are classed as purines, and C, T, and U are jointly called pyrimidines. The nucleic acid bases A, C, and G are shared by all RNA and DNA, but the nucleic acid bases T and U are specific to their respective molecule types. DNA contains a pentose sugar called 2′-deoxyribose, which is chemically distinct from RNA’s ribose because it lacks a hydroxyl group (OH) on the 2′ carbon of the sugar ring.

Without an associated phosphate group, the sugar connected to one of the bases is known as a nucleoside. Phosphate groups link together chains of sugar residues by linking the 5′-hydroxyl group of one sugar to the 3′-hydroxyl group of the next sugar in the chain. RNA and DNA share the same nucleoside connections, called phosphodiester bonds.

Biosynthesis and degradation of Nucleotides

The cell synthesises nucleotides from easily accessible sources. Both purine and pyrimidine nucleotides have a ribose phosphate component that is generated from glucose via the pentose phosphate pathway. The six-atom pyrimidine ring is produced initially and afterwards linked to the ribose phosphate. The two rings of purines are generated while linked to the ribose phosphate during the construction of adenine or guanine nucleosides. Both processes result in a nucleotide with a phosphate bonded to the sugar’s 5′ carbon. Finally, a specialised enzyme called a kinase adds two phosphate groups utilising adenosine triphosphate (ATP) as the phosphate donor to generate ribonucleoside triphosphate, the immediate precursor of RNA. Deoxyribonucleoside diphosphate, the building block of DNA, is derived from ribonucleoside diphosphate by hydrolyzing the 2′-position of the ribose sugar. An extra phosphate group from ATP is subsequently added by another kinase to generate a deoxyribonucleoside triphosphate, the immediate precursor of DNA.

RNA is always being synthesised and degraded as part of a cell’s regular metabolic processes. A number of salvage processes recycle the purine and pyrimidine residues into new DNA. Nucleosides for purine and pyrimidine are recovered, respectively.

Structure of Nucleotide:

There are three parts to a nucleotide’s chemical structure, and they are as follows:

  • Nitrogenous bases
  • Pentose sugar (either ribose or deoxyribose)
  • Phosphate group
structure of nucleotide

“Classification of Nucleic Acids”

Nucleic acids are classified as;

  • Deoxyribonucleic acid (DNA)
  • Ribonucleic acid (RNA)

Deoxyribonucleic acid (DNA)

structure of dna and rna
  • DNA is the genetic substance that makes up a cell’s nucleus.
  • It stores genetic information and transmits it to the organism as it grows.
  • For instance, in humans, the fertilized egg transmits the instructions for creating the head, liver, heart, kidney, hands, and legs.
  • DNA has a double helix structure, which is a spiral made of two strands that spiral over one another with the aid of hydrogen bonds.
  • J. Watson and Francis Crick discovered the structure of DNA in 1953.
  • They won the Nobel Prize for their efforts.
  • Each DNA helix is constructed of nucleotides, which comprise phosphate groups, nitrogenous bases, and deoxyribose sugar (pentose sugar).
  • Nitrogenous bases are organized in pairs in a strand.
  • Each base pair in a strand represents a code.
  • Each code contains genetic data, which is utilized to synthesize proteins.
  • DNA is the carrier of genetic information from one generation to the next.
structure of dna nucleotide

Ribonucleic acid (RNA):

  • RNA is a single-stranded structure.
  • It is also a polymer of nucleotides that contain phosphate group, ribose sugar, and nitrogenous bases (in RNA ribose sugar and DNA deoxyribose sugar).
  • DNA transmits genetic data to RNA. RNA reads the data and immediately produces protein.
structure of rna nucleotide

Types of RNA

Messenger RNA (mRNA)

Information from DNA is sent to the ribosome, a specialised structure or organelle, where it is decoded to create a protein. A terminal 5′-triphosphate group and a 3′-hydroxyl residue are added to the transcribed DNA sequence in mRNAs from prokaryotes. The molecules of messenger RNA are more complex in eukaryotes. A cap is formed by further esterification of the 5′-triphosphate residue. Eukaryotic messenger RNAs (mRNAs) generally have long lines of adenosine residues (polyA) at their 3′ ends. These polyA tracts are not present in the DNA but are added enzymatically after transcription. mRNA in eukaryotes is often made up of smaller pieces of the original gene and is created by cleaving and rejoining a precursor RNA (pre-mRNA) molecule that is an identical duplicate of the gene (as described in the section Splicing). The cap structure and the polyA tail of eukaryotic mRNAs significantly increase their stability, while bacterial mRNAs are swiftly destroyed.

Ribosomal RNA (rRNA)

The ribosome’s structural components are molecules of ribosomal RNA (rRNA). Secondary structures formed by rRNAs are crucial for the recognition of conserved regions of messenger RNAs and transfer RNAs. They also play a role in protein synthesis catalysis. Seven copies of the rRNA genes in the prokaryotic cell E. coli are responsible for producing roughly 15,000 ribosomes. Eukaryotes have a substantially higher population density. The number of ribosomes in a single cell can range from 50 to 5,000 sets of genes, and there can be as many as 10 million ribosomes. These rRNA genes are separated from the main chromosomal fibres in eukaryotes, and they join together with the help of proteins to form an organelle called the nucleolus. The nucleolus is responsible for the transcription of rRNA genes and the initial stages of ribosome assembly.

Transfer RNA (tRNA)

Amino acids are transported to the ribosome by transfer RNA (tRNA) and assembled into the polypeptide chain. The tRNA molecules fold into a distinct cloverleaf shape and contain between 70 and 80 nucleotides. Each of the 20 amino acids required for protein synthesis has its own specialised tRNA, and in many cases, multiple tRNAs for the same amino acid are present. Each three-base sequence (a codon) in the nucleotide sequence corresponds to a specific amino acid in the translated protein. Much less than 61 unique tRNAs can decode the 61 codons necessary to code for amino acids (as described in the section Translation). There are 61 codons in the genome of E. coli, and 40 distinct tRNAs are utilised to translate them. A group of enzymes known as aminoacyl tRNA synthetases are responsible for attaching amino acids to tRNAs. Certain organisms need fewer than 20 synthetases since some amino acids, such as glutamine and asparagine, can be produced on their corresponding tRNAs. Because they must all interact with the same locations on the ribosome, all tRNAs take on structurally similar forms.

Ribozymes

In the cell, proteins are not always the catalysts of choice. In 1989, Thomas Cech and Sidney Altman—who were jointly given the Nobel Prize in Chemistry for their discovery of RNAs with enzymatic activity—became widely recognised as pioneers in the field. Cech demonstrated that a noncoding sequence (intron) in protozoan small subunit rRNA may extract itself from a much longer precursor RNA molecule and reunite the two ends in an autocatalytic event, allowing the rRNA to function. According to Altman’s research, a complex of RNA and proteins known as ribonuclease P can split a precursor tRNA into a mature tRNA. Artificial RNAs with a wide range of catalytic reactions, including self-splicing RNAs like the one identified by Cech, have been created. Most scientists now agree that there was a time in evolutionary history when RNA alone catalysed reactions and stored genetic data. This historical period, commonly referred to as “the RNA world,” is thought to have occurred before DNA served as genetic material.

Antisense RNAs

Most therapeutically promising antisense RNAs are chemically modified forms of RNA or DNA. Antisense RNAs are found naturally and contain sequences that are the antisense of the coding sequences found in messenger RNAs (also called sense RNAs). Antisense RNAs are also single-stranded like messenger RNAs, however, they cannot be translated into proteins. Through the formation of a stable double-stranded structure, they are able to render their complementary mRNA non-functional and halt the process of base-sequence translation. Selective gene silencing is achieved through the disruption of normal RNA metabolism by the artificial introduction of antisense RNAs into cells.

Note:

  • DNA codes for the protein insulin.
  • It is necessary for cells to utilize glucose properly.
  • Previously, insulin from cattle had to be used by diabetics.
  • The production of human insulin now uses recombinant DNA technology.
  • The human gene for producing insulin is taken by scientists and pasted into the DNA of E. coli, a bacterium frequently present in the human digestive tract.
  • Each new E. coli cell has a human insulin gene embedded in its DNA, and the bacterial cell multiplies quickly, creating billions of copies.

Note:

  • By using a sample of a person’s hair cells, skin cells, or other body fluids, we may quickly and readily identify each and every individual thanks to the variances in DNA.
  • Everyone has a unique fingerprint that differs from everyone else’s. Fingerprinting, which involves distinct DNA sequencing, is the cause of the variations in fingerprints.

Nucleic acid metabolism

DNA metabolism

DNA metabolism consists of three basic activities that are each carried out by specialised equipment within the cell: replication, repair, and recombination. Proper DNA replication is essential for protecting the stability of the genome. Errors introduced during replication, or those caused by damage after replication, require correction. Last but not least, recombination between genomes is a crucial method for providing variety within a species and aiding in the repair of damaged DNA. In prokaryotes, where the machinery is more streamlined and simpler, the intricacies of each step have been figured out. It appears that many fundamental principles are shared among eukaryotes.

Replication

Mechanical Principles
In the semiconservative process of DNA replication, the two strands are physically separated and new complementary strands are created independently, resulting in two identical copies of the original DNA molecule. This means that both the parent strand and a synthetic strand are present in each clone. Each chromosome has a unique starting place, or origin, from which replication spreads out in both directions down the strand and eventually halts. When the two elongating chains of a circular chromosome meet, some proteins attach themselves to the strands, marking the end of the chromosome. DNA polymerases can only stretch oligonucleotide segments called primers, and cannot start replication from the end of a DNA strand. In order to prevent data loss, linear chromosomes have unique processes to start and stop DNA synthesis. A short RNA primer is typically synthesised by a specialised RNA polymerase called primase before DNA synthesis is initiated. The RNAs that served as the replication origins of DNA are eventually degraded after DNA synthesis is complete.

The orientation of the phosphodiester bond governs the mode of replication for each DNA strand. Every new nucleotide is added to the 3′ end of the chain, resulting in a continual replication of the leading strand. Bases are always added in the 5′ to 3′ direction as small RNA primers are laid down and then filled in by DNA polymerase to synthesis the lagging strand in an interrupted fashion. When their tiny RNA fragments created during lagging strand copying are no longer required, they are destroyed. DNA ligase is an enzyme that joins together the two pieces of synthetic DNA. This allows replication to occur in both directions, with two leading strands and two lagging strands leaving the origin in opposite directions.

Replicating enzymes

DNA polymerase is an enzyme that can add single nucleotides to the 3′ end of a strand of DNA or RNA. There are three DNA polymerases in the prokaryotic cell E. coli, one of which replicates chromosomes while the other two are engaged in DNA resynthesis during damage repair. The DNA polymerases of eukaryotes are even more intricate. For example, more than five distinct DNA polymerases have been identified in human cells. In human cells, leading and lagging strand synthesis are catalysed by distinct polymerases, and mitochondrial DNA replication is carried out by yet another polymerase. Some of the other polymerases are involved in fixing broken DNA.

Many other proteins play important roles in the replication process as well. DNA helicase proteins aid in unwinding the double helix, whereas single-stranded DNA binding proteins hold the strands together while they unwind in preparation for duplication. Stress in the form of supercoiling is introduced during DNA helix unwinding and is relieved by enzymes known as topoisomerases (see above Supercoiling). Primase, a kind of RNA polymerase, synthesises the primers used to initiate transcription at the origin, while DNA ligase repairs the nicks created between the individual DNA molecules.

The DNA polymerase telomerase is responsible for the synthesis of telomeres, which are unique sequences found at the ends of linear eukaryotic chromosomes. An RNA subunit of this enzyme provides a blueprint for the precise sequence of bases at chromosomal ends. In order to lengthen chromosomes, telomerase RNA replicates a small sequence several times and appends these copies to the telomeres. This prevents the DNA chain from becoming shorter than necessary during the replication process.

Particularly intricate mechanisms are used in the replication of single-stranded viral genomes, mitochondrial genomes, and certain viral genomes. A nucleotide attached to a protein serves as a primer for certain viruses, including adenoviruses; the protein continues to be attached to the DNA even after replication is complete. The rolling circle mechanism of replication is used by many single-stranded viruses to produce a double-stranded copy of the virus. The replication machinery then continuously replicates the nonviral strand, producing long stretches of single-stranded DNA from which specific nucleases cut off the viral DNA in its entirety.

Recombination

To introduce new traits into populations, recombination is crucial.

For example: during meiosis, the process that creates sperm and eggs, homologous chromosomes (one from the mother and one from the father) join up and undergo recombination (crossing over). By shuffling around identical chromosome segments from both DNA molecules, we can create two new chromosomes that are a mosaic of the two originals. Each sperm or egg will take with it only one of the newly rearranged chromosomes when the pair splits up. By combining an egg and sperm, nature reestablishes a complete set of chromosomes.

The term “recombination” can refer to either a broad or local phenomenon. Cleavage and rejoining at identical or extremely similar sequences characterise the vast majority of general recombination. Cleavage happens at a particular spot where DNA is often introduced in site-specific recombination. General recombination occurs during viral infection, bacterial conjugation, transformation (the process by which DNA is transferred directly into cells), and certain types of repair. Parasitic DNA segments often spread throughout genomes via site-specific recombination. Site-specific recombination is crucial to the reproduction and spread of many viruses and transposons, which are unique DNA sequences. More information on the two procedures is provided below.

General recombination

Two DNA molecules with very similar base sequences can undergo general recombination, also known as homologous recombination. Nips in the DNA molecules generate single strands, which then invade the complementary DNA in the other duplex, where base pairing creates a four-stranded DNA structure. A Holliday junction is the cruciform junction found within this structure; it was named after Robin Holliday, the scientist who first presented the hypothesis for homologous recombination in 1964. In order to move along the DNA duplex, the Holliday junction “unzips” one strand and reforms the hydrogen bonds on the other strand. After this fork migration, nicking the two duplexes again will allow them to unwind. DNA ligase is then used to seal the nicks in the DNA. The end result is a pair of DNA duplexes in which the region between the two nicks has been swapped out. It is in the prokaryotic organism E. coli that the enzymes involved in recombination have been studied the most thoroughly. RecA is an essential enzyme because it catalyses strand invasion. RecA covers single-stranded DNA and makes it easier for it to couple with a complementary double-stranded DNA molecule, resulting in a loop structure.

RecBC is another protein that plays an integral role in recombination. RecBC catalyses an unwinding-rewinding process at DNA’s free ends and functions its way along the molecule. Due to the fact that unwinding is more rapid than rewinding, an enzyme creates a loop in the DNA’s backbone that can then be used to mate with another strand of DNA. Single-stranded DNA binding proteins, which serve to maintain single-stranded DNA, DNA polymerase, which repairs any gaps that may be created, and DNA ligase, which reseals the nicks left behind after recombination, are also crucial. Eukaryotic recombination is thought to be very similar to that of E. coli, despite the fact that it is more difficult due to the very compact chromatin structure of eukaryotes.

Recombination between two comparable but not identical regions of DNA results in a “heteroduplex,” or a molecule with mismatched bases at some locations in the helix. Therefore, the mosaic chromosomes generated by recombination during meiosis require one cycle of replication before they are correctly matched. Sometimes, the early results of recombination can be fixed before they are duplicated, thanks to enzymes present in cells that precisely recognise and repair mismatches. In such circumstances, one of the original parental molecules will appear to have been retained to the exclusion of the other, a process known as gene conversion, and the final results of replication will not be true reciprocal events.

Lesions in DNA can sometimes be repaired through recombination. With recombination, the information from one chromosome can be duplicated and placed into the other to provide a correct replacement for the damaged area if one of the chromosomes in a pair is destroyed irreparably. The essential principle here is that replication can copy the proper sequence and repair the lesion if sequences bordering the damage on a sister chromosome base pair with the equivalent sequences on the damaged chromosome.

Site-specific recombination

During site-specific recombination, proteins look for extremely tiny stretches of DNA. Insertion, deletion, or inversion of long DNA sequences like viral genomes, drug-resistance genes, or the mating type locus in yeast can have dramatic regulatory implications. Site-specific recombination is the primary force behind the modification of genomes. Evidence of transposable element insertions across the genome and even into one another may be found in the genomes of many higher organisms, including plants and humans.

The incorporation of bacteriophage DNA into the chromosome of Escherichia coli is a good illustration of site-specific recombination. In this procedure, the circularised phage attachment site is cleaved by the enzyme -integrase. Normally, phage DNA is linear and cannot undergo this transformation. Integrase makes a break in the bacterial chromosome at a comparable location, producing ends with the same in-frame extension. These two pieces complement one another, allowing the original circular chromosome to be put into the E. coli chromosome. Once integrated, the phage can be kept dormant until signals are released to uncouple the phage genome from the host cell, allowing the phage to multiply and infect new bacteria as usual. Only -integrase and a single DNA-binding protein from the host, the integration host factor, are needed for site-specific recombination. Excisionase, a third protein, works with integrase to catalyse the excision of the chromosome from the bacterial chromosome by recognising the hybrid sites generated during integration.

A similar but more ubiquitous type of DNA integration and excision is displayed by the transposons, the so-called hopping genes. These components can have anywhere from under a thousand to over forty thousand base pairs. Transposons are able to migrate from one region in a genome to another, as initially identified in corn (maize) during the 1940s and ’50s by Barbara McClintock, whose discovery garnered her a Nobel Prize in 1983. Like -integrase, transposase cleaves the transposon ends and the target site and is encoded by the vast majority of transposons. In contrast to bacteriophages, which can exist outside of the chromosome, transposons are always kept in an integrated position within the genome. There are two distinct types of transposition: one in which the element merely travels from one place in the chromosome to another, and the other in which the transposon replicates first. This second form of transposition leaves behind the original copy of the transposon and makes a second copy that is put elsewhere in the genome. Known as replicative transposition, this process is the mechanism responsible for the massive spread of transposable elements in many higher species.

Many transposons lack any extra genes to save the transposase itself, making them the simplest type. They exhibit parasitic behaviour and are generally unhelpful to the host organism. Oftentimes, transposable elements have auxiliary genes like those for antibiotic resistance. The development of antibiotic resistance often results from the acquisition by an infecting bacterium of a plasmid harbouring a gene expressing resistance to one or more antibiotics. These resistance genes are typically located on transposable elements that have migrated into plasmids and can be easily transmitted from one organism to another. The ability to continue growing while exposed to an antibiotic provides a significant selective advantage to any bacterium that acquires such a gene. The development of plasmids and strains resistant to many antibiotics is facilitated by their widespread use.

Repair

The correct functioning of a cell throughout its existence and the appropriate transmission of genetic information from one generation to the next rely heavily on the integrity of DNA, which must be preserved at all costs. Repair processes achieve this upkeep by continually monitoring the DNA for damages and activating relevant repair enzymes. Recombination mechanisms, as explained in General recombination, can fix significant lesions in DNA such as pyrimidine dimers or gaps, but they aren’t the only ones.

Mismatch repair is an essential mechanism that has received a lot of attention thanks to research on it in E. coli. The presence of a methyl group inside the template strand sequence GATC controls the mechanism. Eukaryotes have functional mismatch repair systems similar to those of prokaryotes, despite the absence of methyl groups on the template strand. Many malignancies are known to be caused by alterations in the genes encoding human mismatch repair mechanisms. Due to the absence of a mismatch repair system, mutations accumulate rapidly, threatening the function of the genes responsible for cell division. That’s why cancer develops when cells divide without control.

The most prevalent form of damage to nucleic acids occurs after replication is complete and consists of chemical modifications to the regular A, C, G, and T bases, resulting in bases that are usually very different from their natural counterparts. The only exceptions are the deamination of cytosine to uracil and the deamination of 5-methylcytosine to thymine. There will be a G:U or G:T mismatch as a result. DNA glycosylases are enzymes that specifically remove bases by cleaving the link between the base and the deoxyribose sugar, and they can identify uracil in DNA or the thymine in a G:T mismatch. Several of these enzymes target only one of the many possible chemically altered DNA bases.

Excision repair is a common alternative for fixing DNA damage. The enzymes identify the site of DNA damage, most likely by sensing a change in DNA’s structure and then nick the strands on either side of the lesion to create a short segment of single-stranded DNA. The single-stranded gap is then fixed by DNA polymerase and DNA ligase. To repair the correct strand, each of these methods uses the complementary strand as a template and relies on the presence of an aberrant base to determine which strand needs fixing.

Metabolic processes involving RNA

The DNA code is translated by RNA into a form that may be used by the cell. Some RNA molecules, like rRNAs and snRNAs (explained in the section Types of RNA), become components of intricate ribonucleoprotein complexes that serve distinct functions in the cell. The ribosome synthesises proteins in response to instructions from messenger RNAs (mRNAs). Three separate steps of RNA metabolism occur. The precursor RNAs are made by first copying specific regions of the genome via transcription. Second, these molecules go through a maturation process to become fully-formed RNAs. When these RNAs are mRNAs, they are subsequently employed for translation. The bases are recycled, and the RNAs are degraded after they have served their purpose. To put it another way, transcription is the process by which a gene is copied into an RNA that codes for a single protein or performs some other structural or catalytic function. The ribosome is a specialised structure responsible for the process of translation, which involves the decoding of information contained in mRNA molecules. The transcription and translation processes in prokaryotic and eukaryotic species are quite different from one another.

Transcription

The enzyme RNA polymerase is responsible for the precise duplication of short stretches of DNA into RNA. Locating the promoter sequence on DNA is the first step in getting the gene started. Once the DNA strands have been split, RNA polymerase uses a ribonucleoside 5′-triphosphate to start a copying reaction at a specified location on one strand. Substrate-like ribonucleoside triphosphates are used to generate ribonucleoside monophosphates, which are then integrated into the expanding RNA chain via high-energy phosphate bond cleavage. Each subsequent ribonucleotide is directed by the complementary base pairing principles of DNA. Therefore, a G is replicated into C, T into A, and U into RNA when a C is present in DNA to direct its incorporation. When an endpoint is reached in the synthesis process, RNA polymerase detaches from the DNA and the resulting RNA molecule is released. In certain circumstances, this RNA molecule is the final mRNA. In other circumstances, the RNA is a pre-mRNA that must undergo additional processing before the ribosome can begin translation. Ahead of many genes in prokaryotes, there are signals termed “operators” where specific proteins called repressors bind to the DNA immediately upstream of the start point of transcription and inhibit access to the DNA by RNA polymerase. These repressor proteins impede RNA polymerase’s ability to transcribe a gene by physically preventing the enzyme from doing its job. To allow the gene to be expressed, the repressor often waits to receive signals from other molecules in the cell. Certain bacterial genes have signals upstream that activator proteins can attach to and activate transcription.

The process of transcription in multicellular organisms is more involved. To begin, eukaryotic RNA polymerase is a more complex enzyme than prokaryotic RNA polymerase, which only consists of five subunits. Further, the activity of specific promoters is modulated by a wide variety of accessory variables. Transcription factors are proteins that help determine if the transcription is needed in response to signals from within the cell. The transcription of many human genes may require the cooperation of many components. In eukaryotes, a transcription factor can either repress or activate gene expression.

Typically, during transcription, only one DNA strand is duplicated. The RNA molecules synthesised use this strand as their “template” and thus are single-stranded. It is common for the coding or sense strand of DNA to switch between genes, as this is the strand that would encode the mRNA. In eukaryotes, the first product of transcription is a pre-mRNA, which undergoes significant splicing to become the final, mature mRNA.

Translation

The process of translation employs the information inherent in the nucleotide sequence of mRNA to drive the synthesis of a specific protein for usage by the cell. Translation takes occurs on the ribosomes—complex particles in the cell that holds RNA and protein. Ribosomes are loaded onto mRNA during transcription in prokaryotes. A brief sequence of nucleotides towards the 5′ end of the mRNA indicates the initiation of translation. It contains a few nucleotides called a ribosome binding site, or Shine-Dalgarno sequence. The tetranucleotide sequence GAGG is sufficient as a binding site in E. coli. This is usually found within the first 5–8 bases after a start codon. The mRNA sequence is read three bases at a time from its 5′ end toward its 3′ end, and one amino acid is added to the expanding chain from its respective aminoacyl tRNA until the complete protein chain is produced. When the ribosome reaches a termination codon, usually UAG, UAA, or UGA, translation is halted. In reaction to these codons, specific release factors bind to the ribosome, and the newly created protein, tRNAs, and mRNA dissociate. After that, the ribosome is free to bind to yet another mRNA molecule.

Even while eukaryotic ribosomes are more complex, the process of protein synthesis is fundamentally the same. A similar interaction occurs between the signal sequence and the 3′ end of the small subunit rRNA during bacterial initiation.

When compared to the importance of fidelity during replication, the issue of fidelity during protein synthesis is less pressing. One mRNA molecule can be translated repeatedly to give numerous copies of the protein. It is common for proteins to be destroyed by the cell’s machinery if they fail to fold properly after being mistranslated. However, the ribosome incorporates proofreading mechanisms to guarantee proper codon-anticodon matching.

Foods that contain high Nucleic acid

  • Seafood
  • nuts
  • vegetables
  • mushrooms
  • yeast
  • meat
  • stocks and soups

Seafood

Nucleic acids can be found in a variety of foods, but fish is especially rich in this nutrient. Gordon Research Institute found that sardines (1.5% of their overall composition) contain the greatest quantities of nucleic acids. Nucleic acids can be found in both seafoods derived from marine animals and plant-based choices. AGM Foods claims that its product lineup includes chlorella, a form of single-celled algae. Algae are well-known for the high concentrations of fatty acids and carbohydrates they contain.

Nuts

Nuts are packed with heart-healthy proteins and unsaturated fats. The Gordon Research Institute found that most species also have especially high quantities of nucleic acids.

Vegetables

Vegetables are essential for good health because they help control blood sugar and blood pressure and protect against stroke, heart disease, digestive issues, vision difficulties, and cancer, as reported by Harvard University. The United States National Library of Medicine reports that numerous vegetables are rich in nucleic acids. Chinese cabbage, leeks, spinach, cauliflower, beans, and soybeans are just a few examples.

Mushrooms

Mushrooms are edible fungi that are high in nutrients like vitamin E and selenium and low in cholesterol, fat, calories, and sodium. The U.S. National Library of Medicine reports that numerous types of mushrooms, including the flat, whitecap (button), cep, and oyster mushrooms, contain significant amounts of nucleic acids.

Yeast

Hydrolyzed and autolyzed yeast, which are frequently added to microwavable vegetarian meals, are two excellent sources of nucleic acids, according to the U.S. National Library of Medicine. They can also promote an increase in purine synthesis, which is necessary for the body’s creation of uric acid.

Beef

Similarly to pork, beef is a rich supplier of nucleic acids. The Gordon Research Institute found that red meat has the highest nucleic acid concentration, at 0.05 percent on average.

Soups and Broths

The Gordon Research Institute claims that the nucleic acids in vegetable, mushroom, and/or beef-based soups and broths are of high quality.


->Frequently asked questions:

What are nucleic acids?

Molecules of nucleic acid, such as DNA or RNA, are made up of a lengthy chain of nucleotides joined together.
Nucleic acids are chemical compounds found naturally in the body and are the primary molecules in cells responsible for transporting genetic information. They are crucial in regulating the production of proteins. Deoxyribonucleic acid (DNA) and ribonucleic acid (RNA) are the most common types of nucleic acids (RNA).

What is the basic structure of nucleic acid?

Nucleic acids are molecules that are long and chainlike, made up of smaller molecules called nucleotides. A nitrogen-containing aromatic base is linked to a pentose (five-carbon) sugar, which is linked to a phosphate group, and so on for each and every nucleotide.

What is the basic structure of nucleic acids DNA and RNA?

Nucleotides are the building blocks of eukaryotic cellular nucleic acids (DNA and RNA). Polynucleotides, such as deoxyribonucleic acid (DNA) or ribonucleic acid (RNA), are created when nucleotides bind together. A nitrogenous base, a pentose (five-carbon) sugar, and a phosphate group make up each nucleotide.

What is nucleic acid made of?

Biopolymers, the building blocks from which nucleic acids are constructed, are made up of recurring sets of monomers (the polymers’ building blocks). Understanding the structure of nucleotides is essential for comprehending the structure of nucleic acid.

What nitrogen-containing bases occur in nucleic acids?

Four of the five nitrogen-containing bases (adenine (A), guanine (G), cytosine (C), thymine (T), and uracil (U)) are present in every nucleic acid (U). Purines include the nucleobases A and G, while pyrimidines include the nucleobases C, T, and U. The nucleic acid bases A, C, and G are shared by all RNA and DNA, but the nucleic acid bases T and U are specific to their respective molecule types.

When were nucleic acids discovered?

Friedrich Miescher, a Swiss biochemist, discovered nucleic acids in 1869.

What are examples of nucleic acids in food?

The foods you eat, like all other living things, contain nucleic acids. The highest concentrations of these chemicals have been found in meat, fish, seafood, legumes, and mushrooms.

Does milk have nucleic acids?

Nucleic acids (primarily RNA) and nucleotides are also present in milk.


Related Biochemistry Posts

Biochemistry

Carbohydrates

Proteins

What are Lipids

1 thought on “What is the structure and function of nucleic acid?”

  1. Pingback: Why are Vitamins important | Vitamins in Biochemistry - Best Online Free Chemistry Learning

Leave a Reply

%d bloggers like this: