Jump to content

Gene: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Senpai dont care
Tags: blanking Possible vandalism
m Reverting possible vandalism by GalaxyGamer2015 to version by Ettrig. False positive? Report it. Thanks, ClueBot NG. (2186720) (Bot)
Line 5: Line 5:
[[File:DNA chemical structure.svg|thumb|upright=1.3|The chemical structure of a four-base fragment of a [[DNA]] [[double helix]].]]
[[File:DNA chemical structure.svg|thumb|upright=1.3|The chemical structure of a four-base fragment of a [[DNA]] [[double helix]].]]


A '''gene''' is the molecular unit of [[heredity]] of a living [[organism]]. The word is used extensively by the scientific community for stretches of [[DNA|deoxyribonucleic acids]] (DNA) and [[RNA|ribonucleic acids]] (RNA) that code for a [[polypeptide]] or for an RNA chain that has a function in the organism. Living beings depend on genes, as they specify all proteins and functional RNA chains. Genes hold the information to build and maintain an organism's [[cell (biology)|cells]] and pass genetic [[trait (biology)|traits]] to offspring. All organisms have genes corresponding to various biological traits, some of which are instantly visible, such as [[eye color]] or number of limbs, and some of which are not, such as [[blood type]], increased risk for specific diseases, or the thousands of basic [[Biochemistry|biochemical]] processes that comprise [[life]]. The word ''gene'' was coined by [[Wilhelm Johannsen]] in 1909 and is indirectly derived (via ''[[pangenesis
A '''gene''' is the molecular unit of [[heredity]] of a living [[organism]]. The word is used extensively by the scientific community for stretches of [[DNA|deoxyribonucleic acids]] (DNA) and [[RNA|ribonucleic acids]] (RNA) that code for a [[polypeptide]] or for an RNA chain that has a function in the organism. Living beings depend on genes, as they specify all proteins and functional RNA chains. Genes hold the information to build and maintain an organism's [[cell (biology)|cells]] and pass genetic [[trait (biology)|traits]] to offspring. All organisms have genes corresponding to various biological traits, some of which are instantly visible, such as [[eye color]] or number of limbs, and some of which are not, such as [[blood type]], increased risk for specific diseases, or the thousands of basic [[Biochemistry|biochemical]] processes that comprise [[life]]. The word ''gene'' was coined by [[Wilhelm Johannsen]] in 1909 and is indirectly derived (via ''[[pangenesis|pangene]]'') from the [[Ancient Greek]] word γένος (''génos'') meaning "race, offspring".<ref>{{OED|gene}}</ref>

...........................................................................
A modern working definition of a gene is "''a [[Locus (genetics)|locatable region]] of [[genome|genomic]] sequence, corresponding to a unit of inheritance, which is associated with regulatory regions, transcribed regions, and/or other functional sequence regions'' ".<ref name=Pearson_2006>{{vcite2 journal | vauthors = Pearson H | title = Genetics: what is a gene? | journal = Nature | volume = 441 | issue = 7092 | pages = 398–401 | date = May 2006 | pmid = 16724031 | doi = 10.1038/441398a | bibcode = 2006Natur.441..398P }}</ref><ref name = "Rethink">{{vcite2 journal | vauthors = Pennisi E | title = Genomics. DNA study forces rethink of what it means to be a gene | journal = Science | volume = 316 | issue = 5831 | pages = 1556–1557 | date = June 2007 | pmid = 17569836 | doi = 10.1126/science.316.5831.1556 }}</ref> Colloquial usage of the term ''gene'' (e.g., "good genes", "hair color gene") may actually refer to an [[allele]]: a ''gene'' is the basic instruction— a sequence of nucleic acids (DNA or, in the case of certain [[virus]]es RNA), while an ''allele'' is one variant of that gene. Thus, when the mainstream press refers to "having a gene" for a specific trait, this is customarily inaccurate. In most cases, all people would have a gene for the trait in question, although certain people will have a specific allele of that gene, which results in the trait variant. Further, genes code for proteins, which might result in identifiable traits, but it is the gene ([[genotype]]), not the trait ([[phenotype]]), which is inherited.

Big genes are a class of genes whose [[cell nucleus|nuclear]] [[Transcription (genetics)|transcript]] spans 500 [[Kilo base pair|kb]] (1 kb = 1,000 [[base pairs]]) or more of [[chromosomal]] [[DNA]]. The largest of the big genes is the gene for [[dystrophin]], which spans 2.3 Mb. Many big genes have modestly sized [[mRNA]]s; the [[exon]]s [[encoding]] these [[RNAs]] typically encompass about 1% of the total chromosomal gene region in which they occur.

==History==
[[File:Gregor Mendel.png|thumb|200px|Gregor Mendel]]
{{main|History of genetics}}

The<!--<ref>{{vcite2 journal | vauthors = Noble D | title = Genes and causation | journal = Philosophical Transactions. Series A, Mathematical, Physical, and Engineering Sciences | volume = 366 | issue = 1878 | pages = 3001–3015 | date = September 2008 | pmid = 18559318 | doi = 10.1098/rsta.2008.0086 | url = http://rsta.royalsocietypublishing.org/cgi/pmidlookup?view=long&pmid=18559318 | format = Free full text | bibcode = 2008RSPTA.366.3001N }}</ref>--> existence of discrete inheritable units was first suggested by [[Gregor Mendel]] (1822–1884). From 1857 to 1864, he studied inheritance patterns in 8000 common edible [[pea plant|pea plants]] (''Pisum''), tracking distinct traits from parent to offspring. He described these mathematically as 2<sup>n</sup> combinations where n is the number of differing characteristics in the original peas. Although he did not use the term ''gene'', he explained his results in terms of discrete inherited units that give rise to distinct [[phenotype|phenotypes]]. Mendel was also the first to show [[independent assortment]], the distinction between [[dominant gene|dominant]] and [[recessive]] traits, the distinction between a [[heterozygote]] and [[homozygote]], the phenomenon of discontinuous inheritance and what would later be described as [[genotype]] (the genetic material of an organism) and [[phenotype]] (the visible traits of that organism) and the conversion of one form into another within few generations.

Prior to Mendel's work, the dominant theory of heredity was one of [[blending inheritance]], which suggested that each parent contributed fluids to the fertilisation process and that the traits of the parents blended and mixed to produce the offspring. [[Charles Darwin]] developed a theory of inheritance he termed [[pangenesis]], which used the term ''[[gemmule (pangenesis)|gemmule]]'' to describe hypothetical particles that would mix during reproduction.

Although Mendel's work was largely unrecognized after its first publication in 1866, it was 'rediscovered' in 1900 by three European scientists, [[Hugo de Vries]], [[Carl Correns]], and [[Erich von Tschermak]], who claimed to have reached similar conclusions in their own research. [[Denmark|Danish]] [[botanist]] [[Wilhelm Johannsen]] coined the word "gene" ("gen" in Danish and German) in 1909 to describe the fundamental physical and functional units of heredity,<ref name="genome">{{cite web | url=http://www.genome.gov/25019879 | title=The Human Genome Project Timeline | accessdate=13 September 2006 }}</ref> while the related word [[genetics]] was first used by [[William Bateson]] in 1905.<ref name="Gerstein"/> In 1910, [[Thomas Hunt Morgan]] showed that genes reside on specific [[chromosome]]s. He later showed that genes occupy specific locations on the chromosome. With this knowledge, Morgan and his students began the first chromosomal map of the fruit fly ''[[Drosophila melanogaster|Drosophila]]''. In 1928, [[Frederick Griffith]] showed that genes could be transferred. In what is now known as [[Griffith's experiment]], injections into a mouse of a deadly strain of [[bacteria]] that had been heat-killed transferred genetic information to a safe strain of the same bacteria, killing the mouse.

A series of subsequent discoveries led to the realization decades later that the genetic material is made of [[DNA]] (deoxyribonucleic acid). In 1941, [[George Wells Beadle]] and [[Edward Lawrie Tatum]] showed that mutations in genes caused errors in specific steps in [[metabolic pathway]]s. This showed that specific genes code for specific proteins, leading to the "[[one gene, one enzyme]]" hypothesis.<ref name="Gerstein"/> [[Oswald Avery]], [[Colin Munro MacLeod]], and [[Maclyn McCarty]] [[Avery-MacLeod-McCarty experiment|showed in 1944]] that DNA holds the gene's information.<ref>{{vcite2 journal | vauthors = Steinman RM, Moberg CL | title = A triple tribute to the experiment that transformed biology | journal = The Journal of Experimental Medicine | volume = 179 | issue = 2 | pages = 379–84 | date = February 1994 | pmid = 8294854 | pmc = 2191359 | doi = 10.1084/jem.179.2.379 }}</ref> In 1952, [[Rosalind Franklin]] and Raymond Gosling produced a strikingly clear x-ray diffraction pattern indicating a helical form, and in 1953, [[James D. Watson]] and [[Francis Crick]] demonstrated the molecular structure of [[DNA]]. Together, these discoveries established the [[central dogma of molecular biology]], which states that proteins are translated from [[RNA]] which is transcribed from DNA. This dogma has since been shown to have exceptions, such as [[reverse transcription]] in [[retrovirus]]es.

In 1972, [[Walter Fiers]] and his team at the [[University of Ghent]] were the first to determine the sequence of a gene: the gene for [[Bacteriophage MS2]] coat protein.<ref name=Min_1972>{{vcite2 journal | vauthors = Min Jou W, Haegeman G, Ysebaert M, Fiers W | title = Nucleotide sequence of the gene coding for the bacteriophage MS2 coat protein | journal = Nature | volume = 237 | issue = 5350 | pages = 82–8 | date = May 1972 | pmid = 4555447 | doi = 10.1038/237082a0 | bibcode = 1972Natur.237...82J }}</ref> [[Richard J. Roberts]] and [[Phillip Sharp]] discovered in 1977 that genes can be split into segments. This led to the idea that one gene can make several proteins. The successful sequencing of many organisms' [[genome|genomes]] has complicated the molecular definition of genes. In particular, genes do not seem to sit side by side on [[DNA]] like discrete beads. Instead, [[region]]s of the DNA producing distinct proteins may overlap, so that the idea emerges that "genes are one long [[Continuum (theory)|continuum]]".<ref name=Pearson_2006 /> It was first hypothesized in 1986 by [[Walter Gilbert]] that neither DNA nor protein would be required in such a primitive system as that of a very early stage of the earth if RNA could perform as simply a catalyst and genetic information storage processor.

The modern study of [[genetics]] at the level of DNA is known as [[molecular genetics]] and the synthesis of molecular genetics with traditional [[Charles Darwin|Darwinian]] [[evolution]] is known as the [[modern evolutionary synthesis]].

==Mendelian inheritance and classical genetics==
[[File:Punnett square mendel flowers.svg|thumb|Crossing between two pea plants [[heterozygous]] for purple (B, dominant) and white (b, recessive) blossoms]]
{{main|Mendelian inheritance|Classical genetics}}

According to the theory of [[Mendelian inheritance]], variations in [[phenotype]]—the observable physical and behavioral characteristics of an organism—are due in part to variations in [[genotype]], or the organism's particular set of genes, each of which specifies a particular trait. Different forms of a gene, which may give rise to different phenotypes, are known as [[alleles]]. Organisms such as the pea plants Mendel worked on, along with many plants and animals, have two alleles for each trait, one inherited from each parent. Alleles may be [[dominant gene|dominant]] or [[recessive gene|recessive]]; dominant alleles give rise to their corresponding phenotypes when paired with any other allele for the same trait, whereas recessive alleles give rise to their corresponding phenotype only when paired with another copy of the same allele. For example, if the allele specifying tall stems in pea plants is dominant over the allele specifying short stems, then pea plants that inherit one tall allele from one parent and one short allele from the other parent will also have tall stems. Mendel's work demonstrated that alleles assort independently in the production of [[gamete]]s, or [[germ cell]]s, ensuring variation in the next generation.

==Physical definitions==

===Functional structure of a gene===
[[File:Gene structure 2 annotated.svg|thumb|600px|Diagram of the "typical" [[Eukaryote|eukaryotic]] protein-coding '''gene'''. [[promoter (biology)|Promoter]]s and [[Enhancer (genetics)|enhancers]] determine what portions of the [[DNA]] will be [[transcription (genetics)|transcribed]] into the [[precursor mRNA]] (pre-mRNA). The pre-mRNA is then spliced into [[messenger RNA]] (mRNA) which is later [[Translation (biology)|translated]] into [[protein]].]]

The vast majority of living organisms encode their genes in long strands of DNA (deoxyribonucleic acid). DNA consists of a chain made from four types of [[nucleotide]] subunits, each composed of: a five-carbon sugar ([[deoxyribose|2'-deoxyribose]]), a [[phosphate]] group, and one of the four [[nucleobase|bases]] [[adenine]], [[cytosine]], [[guanine]], and [[thymine]]. The most common form of DNA in a cell is in a [[double helix]] structure, in which two individual DNA strands twist around each other in a right-handed spiral. In this structure, the [[Watson-Crick base pair|base pairing]] rules specify that guanine pairs with cytosine and adenine pairs with thymine. The base pairing between guanine and cytosine forms three [[hydrogen bond]]s, whereas the base pairing between adenine and thymine forms two hydrogen bonds. The two strands in a double helix must therefore be ''complementary'', that is, their bases must align such that the adenines of one strand are paired with the thymines of the other strand, and so on.

Due to the chemical composition of the [[pentose]] residues of the bases, DNA strands have directionality. One end of a DNA polymer contains an exposed [[hydroxyl]] group on the [[deoxyribose]]; this is known as the [[3' end]] of the molecule. The other end contains an exposed [[phosphate]] group; this is the [[5' end]]. The directionality of DNA is vitally important to many cellular processes, since double helices are necessarily directional (a strand running 5'-3' pairs with a complementary strand running 3'-5'), and processes such as [[DNA replication]] occur in only one direction. All nucleic acid synthesis in a cell occurs in the 5'-3' direction, because new monomers are added via a [[dehydration]] reaction that uses the exposed 3' hydroxyl as a [[nucleophile]].

The [[gene expression|expression]] of genes encoded in DNA begins by [[transcription (genetics)|transcribing]] the gene into [[RNA]], a second type of [[nucleic acid]] that is very similar to DNA, but whose monomers contain the sugar [[ribose]] rather than [[deoxyribose]]. RNA also contains the base [[uracil]] in place of [[thymine]]. RNA molecules are less stable than DNA and are typically single-stranded. Genes that encode proteins are composed of a series of three-[[nucleotide]] sequences called [[codon]]s, which serve as the ''words'' in the genetic ''language''. The [[genetic code]] specifies the correspondence during [[translation (genetics)|protein translation]] between codons and [[amino acid]]s. The genetic code is nearly the same for all known organisms.

All genes have regulatory regions in addition to regions that explicitly code for a protein or RNA product. A [[Regulatory sequence|regulatory region]] shared by almost all genes is known as the [[promoter (biology)|promoter]], which provides a position that is recognized by the transcription machinery when a gene is about to be transcribed and expressed. A gene can have more than one promoter, resulting in RNAs that differ in how far they extend in the 5' end.<ref>{{vcite2 journal | vauthors = Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B | title = Mapping and quantifying mammalian transcriptomes by RNA-Seq | journal = Nature Methods | volume = 5 | issue = 7 | pages = 621–8 | date = July 2008 | pmid = 18516045 | doi = 10.1038/nmeth.1226 }}</ref> Although promoter regions have a [[consensus sequence]] that is the most common sequence at this position, some genes have "strong" promoters that bind the transcription machinery well, and others have "weak" promoters that bind poorly. These weak promoters usually permit a lower rate of transcription than the strong promoters, because the transcription machinery binds to them and initiates transcription less frequently. Other possible regulatory regions include [[enhancer (genetics)|enhancers]], which can compensate for a weak promoter. Most regulatory regions are "upstream"—that is, before or toward the 5' end of the transcription initiation site. [[Eukaryote|Eukaryotic]] [[promoter (biology)|promoter]] regions are much more complex and difficult to identify than [[prokaryote|prokaryotic]] promoters.

Many prokaryotic genes are organized into [[operon]]s, or groups of genes whose products have related functions and which are transcribed as a unit. By contrast, [[eukaryote|eukaryotic]] genes are transcribed only one at a time, but may include long stretches of DNA called [[intron]]s which are transcribed but never translated into protein (they are spliced out before translation). Splicing can also occur in prokaryotic genes, but is less common than in eukaryotes.<ref>{{vcite2 journal | vauthors = Woodson SA | title = Ironing out the kinks: splicing and translation in bacteria | journal = Genes & Development | volume = 12 | issue = 9 | pages = 1243–7 | date = May 1998 | pmid = 9573040 | doi = 10.1101/gad.12.9.1243 | url = http://genesdev.cshlp.org/content/12/9/1243.full }}</ref>

===Chromosomes===
The total complement of genes in an organism or cell is known as its [[genome]], which may be stored on one or more [[chromosome]]s; the region of the chromosome at which a particular gene is located is called its [[locus (genetics)|locus]]. A chromosome consists of a single, very long DNA helix on which thousands of genes are encoded. [[Prokaryote]]s—[[bacteria]] and [[archaea]]—typically store their genomes on a single large,circular chromosome, sometimes supplemented by additional small circles of DNA called [[plasmid]]s, which usually encode only a few genes and are easily transferable between individuals. For example, the genes for [[antibiotic resistance]] are usually encoded on bacterial plasmids and can be passed between individual cells, even those of different species, via [[horizontal gene transfer]].
<!-- this would be a good place for a table of % 'junk' or similar -->
Although some simple eukaryotes also possess plasmids with small numbers of genes, the majority of eukaryotic genes are stored on multiple linear chromosomes, which are packed within the [[cell nucleus|nucleus]] in complex with storage proteins called [[histone]]s. The manner in which DNA is stored on the histone, as well as chemical modifications of the histone itself, are regulatory mechanisms governing whether a particular region of DNA is accessible for [[gene expression]]. The ends of eukaryotic chromosomes are capped by long stretches of repetitive sequences called [[telomere]]s, which do not code for any gene product but are present to prevent degradation of coding and regulatory regions during [[DNA replication]]. The length of the telomeres tends to decrease each time the genome is replicated in preparation for cell division; the loss of telomeres has been proposed as an explanation for cellular [[senescence]], or the loss of the ability to divide, and by extension for the [[aging]] process in organisms.<ref name="Braig">{{vcite2 journal | vauthors = Braig M, Schmitt CA | title = Oncogene-induced senescence: putting the brakes on tumor development | journal = Cancer Research | volume = 66 | issue = 6 | pages = 2881–4 | date = March 2006 | pmid = 16540631 | doi = 10.1158/0008-5472.CAN-05-4006 }}</ref>

Whereas the chromosomes of prokaryotes are relatively gene-dense, those of eukaryotes often contain so-called "[[junk DNA]]", or regions of DNA that serve no obvious function. Simple single-celled eukaryotes have relatively small amounts of such DNA, whereas the genomes of complex [[multicellular organism]]s, including humans, contain an absolute majority of DNA without an identified function.<ref name="IHSGC2004">{{vcite2 journal | vauthors = International Human Genome Sequencing Consortium | title = Finishing the euchromatic sequence of the human genome | journal = Nature | volume = 431 | issue = 7011 | pages = 931–45 | date = October 2004 | pmid = 15496913 | doi = 10.1038/nature03001 | url = http://www.nature.com/nature/journal/v431/n7011/full/nature03001.html | bibcode = 2004Natur.431..931H }}</ref> However, it now appears that, although protein-coding DNA makes up barely 2% of the [[human genome]], about 80% of the bases in the genome may be expressed, so the term "junk DNA" may be a misnomer.<ref name="Rethink" />

===RNA genes===
[[File:DNA to protein or ncRNA.svg|thumb|300px|Protein coding genes are transcribed to an [[mRNA]] intermediate, then translated to a functional [[protein]]. RNA-coding genes are transcribed to a functional [[non-coding RNA]]. ({{PDB|3BSE}}, {{PDB2|1OBB}}, {{PDB2|3TRA}})]]

When proteins are manufactured, the gene is first copied into [[RNA]] as an intermediate product.<ref name = "Alberts_2002">{{cite book | first1 = Bruce | last1 = Alberts | first2 = Alexander | last2 = Johnson | first3 = Julian | last3 = Lewis | first4 = Martin | last4 = Raff | first5 = Keith | last5 = Roberts | first6 = Peter | last6 = Walter | name-list-format = vanc | author1-link = Bruce Alberts | author3-link = Julian Lewis (biologist) | author4-link = Martin Raff | author6-link = Peter Walter | title = Molecular Biology of the Cell | edition = Fourth | publisher = Garland Science | location = New York | year = 2002 | isbn = 978-0-8153-3218-3 | url = http://www.ncbi.nlm.nih.gov/books/NBK21054/ }}</ref> In other cases, the RNA molecules are the actual functional products, as in the synthesis of [[ribosomal RNA]] and [[transfer RNA]]. Some RNAs known as [[ribozyme]]s are capable of [[enzyme|enzymatic function]], and [[microRNA]] has a regulatory role. The [[DNA]] sequences from which such RNAs are transcribed are known as [[RNA gene]]s.

Some [[virus]]es store their entire genomes in the form of [[RNA]], and contain no DNA at all. Because they use RNA to store genes, their [[cell (biology)|cellular]] [[host (biology)|hosts]] may synthesize their proteins as soon as they are [[infection|infected]] and without the delay in waiting for transcription. On the other hand, RNA [[retrovirus]]es, such as [[HIV]], require the [[reverse transcription]] of their [[genome]] from RNA into DNA before their proteins can be synthesized. RNA-mediated [[epigenetic]] inheritance has also been observed in plants and very rarely in animals.<ref name=morris>{{cite journal|last1=Morris|first1=KV|last2=Mattick|first2=JS|title=The rise of regulatory RNA.|journal=Nature reviews. Genetics|date=June 2014|volume=15|issue=6|pages=423-37|pmid=24776770}}</ref>

==Gene expression==
{{main|Gene expression}}

In all organisms, there are two major steps separating a protein-coding gene from its protein: First, the DNA on which the gene resides must be ''[[transcription (genetics)|transcribed]]'' from DNA to [[messenger RNA]] (mRNA); and, second, it must be ''[[translation (genetics)|translated]]'' from mRNA to protein. RNA-coding genes must still go through the first step, but are not translated into protein. The process of producing a biologically functional molecule of either RNA or protein is called [[gene expression]], and the resulting molecule itself is called a [[gene product]].

===Genetic code===
[[File:RNA-codons-aminoacids.svg|thumb|300px|Schematic diagram of a single-stranded RNA molecule illustrating the position of three-base codons.]]
{{main|Genetic code}}

The genetic code is the set of rules by which information encoded within a gene is translated into a functional protein. Each gene consists of a specific sequence of nucleotides encoded in DNA or RNA. The nucleotide being made up of a sugar, a phosphate molecule and a specific base (adenine, thymine, cytosine, guanine or sometimes uracil; thymine is replaced with uracil in some viruses<ref>{{MeSH name| RNA Viruses}}</ref>); a correspondence between nucleotides, the basic building blocks of genetic material, and amino acids, the basic building blocks of proteins, must be established for genes to be successfully translated into functional proteins. Sets of three nucleotides, known as [[codon]]s, each correspond to a specific amino acid or to a signal; three codons are known as "[[stop codon]]s" and, instead of specifying a new amino acid, alert the translation machinery that the end of the gene has been reached, just as a specific set of 3 bases, "AUG", known as the "[[start codon]]", signifies the gene to start transcribing. There are 64 possible codons (four possible nucleotides at each of three positions, hence 4<sup>3</sup> possible codons) and only 20 standard amino acids; hence the code is redundant and multiple codons can specify the same amino acid. The correspondence between codons and amino acids is nearly universal among all known living organisms.

===Transcription===
The process of genetic [[transcription (genetics)|transcription]] produces a single-stranded [[RNA]] molecule known as [[messenger RNA]], whose nucleotide sequence is complementary to the DNA from which it was transcribed. The DNA strand whose sequence matches that of the RNA is known as the [[coding strand]] and the strand from which the RNA was synthesized is the [[template strand]]. Transcription is performed by an [[enzyme]] called an [[RNA polymerase]], which reads the template strand in the [[3' end|3']] to [[5' end|5']] direction and synthesizes the RNA from [[5' end|5']] to [[3' end|3']]. To initiate transcription, the polymerase first recognizes and binds a [[promoter (biology)|promoter]] region of the gene. Thus, a major mechanism of [[gene regulation]] is the blocking or sequestering of the promoter region, either by tight binding by [[repressor]] molecules that physically block the polymerase, or by organizing the DNA so that the promoter region is not accessible.

In [[prokaryote]]s, [[Transcription (genetics)|transcription]] occurs in the [[cytoplasm]]; for very long transcripts, translation may begin at the 5' end of the RNA while the 3' end is still being transcribed. In [[eukaryote]]s, transcription necessarily occurs in the nucleus, where the cell's DNA is sequestered; the RNA molecule produced by the polymerase is known as the [[primary transcript]] and must undergo [[post-transcriptional modification]]s before being exported to the cytoplasm for translation. The [[splicing (genetics)|splicing]] of [[intron]]s present within the transcribed region is a modification unique to eukaryotes; [[alternative splicing]] mechanisms can result in mature transcripts from the same gene having different sequences and thus coding for different proteins. This is a major form of regulation in eukaryotic cells.

===Translation===
[[Translation (genetics)|Translation]] is the process by which a [[Mature messenger RNA|mature mRNA]] molecule is used as a template for synthesizing a new [[protein]]. Translation is carried out by [[ribosome]]s, large complexes of RNA and protein responsible for carrying out the chemical reactions to add new [[amino acid]]s to a growing [[polypeptide chain]] by the formation of [[peptide bond]]s. The genetic code is read three nucleotides at a time, in units called [[codon]]s, via interactions with specialized RNA molecules called [[transfer RNA]] (tRNA). Each tRNA has three unpaired bases known as the [[anticodon]] that are complementary to the codon it reads; the tRNA is also [[covalent]]ly attached to the [[amino acid]] specified by the complementary codon. When the tRNA binds to its complementary codon in an mRNA strand, the ribosome ligates its amino acid cargo to the new polypeptide chain, which is synthesized from [[N-terminus|amino terminus]] to [[C-terminus|carboxyl terminus]]. During and after its synthesis, the new protein must [[protein folding|fold]] to its active [[tertiary structure|three-dimensional structure]] before it can carry out its cellular function.

==DNA replication and inheritance==
The growth, development, and reproduction of organisms relies on [[cell division]], or the process by which a single [[cell (biology)|cell]] divides into two usually identical [[daughter cell]]s. This requires first making a duplicate copy of every gene in the [[genome]] in a process called [[DNA replication]]. The copies are made by specialized [[enzyme]]s known as [[DNA polymerase]]s, which "read" one strand of the double-helical DNA, known as the template strand, and synthesize a new complementary strand. Because the DNA double helix is held together by [[base pair]]ing, the sequence of one strand completely specifies the sequence of its complement; hence only one strand needs to be read by the enzyme to produce a faithful copy. The process of DNA replication is [[semiconservative replication|semiconservative]]; that is, the copy of the genome inherited by each daughter cell contains one original and one newly synthesized strand of DNA.<ref name=Watson_2004>{{cite book | author = Watson JD, Baker TA, Bell SP, Gann A, Levine M, Losick R | title = Molecular Biology of the Gene | edition = 5th | publisher = Peason Benjamin Cummings (Cold Spring Harbor Laboratory Press) | year = 2004 | isbn = 0-8053-4635-X }}</ref>

After DNA replication is complete, the cell must physically separate the two copies of the genome and divide into two distinct membrane-bound cells. In [[prokaryote]]s&nbsp;– [[bacteria]] and [[archaea]]&nbsp;– this usually occurs via a relatively simple process called [[binary fission]], in which each circular genome attaches to the [[cell membrane]] and is separated into the daughter cells as the membrane [[invagination|invaginates]] to split the [[cytoplasm]] into two membrane-bound portions. Binary fission is extremely fast compared to the rates of cell division in [[eukaryote]]s. Eukaryotic cell division is a more complex process known as the [[cell cycle]]; DNA replication occurs during a phase of this cycle known as [[S phase]], whereas the process of segregating [[chromosome]]s and splitting the [[cytoplasm]] occurs during [[M phase]]. In many single-celled eukaryotes such as [[yeast]], reproduction by [[budding]] is common, which results in asymmetrical portions of cytoplasm in the two daughter cells.

===Molecular inheritance===
The duplication and transmission of genetic material from one generation of cells to the next is the basis for molecular inheritance, and the link between the classical and molecular pictures of genes. Organisms inherit the characteristics of their parents because the cells of the offspring contain copies of the genes in their parents' cells. In [[asexual reproduction|asexually reproducing]] organisms, the offspring will be a genetic copy or [[Clone (genetics)|clone]] of the parent organism. In [[sexual reproduction|sexually reproducing]] organisms, a specialized form of cell division called [[meiosis]] produces cells called [[gamete]]s or [[germ cell]]s that are [[haploid]], or contain only one copy of each gene. The gametes produced by females are called [[egg (biology)|eggs]] or ova, and those produced by males are called [[sperm]]. Two gametes fuse to form a [[fertilized egg]], a single cell that once again has a [[diploid]] number of genes—each with one copy from the mother and one copy from the father.

During the process of meiotic cell division, an event called [[genetic recombination]] or ''crossing-over'' can sometimes occur, in which a length of DNA on one [[chromatid]] is swapped with a length of DNA on the corresponding sister chromatid. This has no effect if the [[allele]]s on the chromatids are the same, but results in reassortment of otherwise linked alleles if they are different. The Mendelian principle of independent assortment asserts that each of a parent's two genes for each trait will sort independently into gametes; which allele an organism inherits for one trait is unrelated to which allele it inherits for another trait. This is in fact only true for genes that do not reside on the same chromosome, or are located very far from one another on the same chromosome. The closer two genes lie on the same chromosome, the more closely they will be associated in gametes and the more often they will appear together; genes that are very close are essentially never separated because it is extremely unlikely that a crossover point will occur between them. This is known as [[genetic linkage]].

==Mutation==
{{main|Mutation}}

DNA replication is for the most part extremely accurate, with an error rate per site of around 10<sup>−6</sup> to 10<sup>−10</sup> in [[eukaryote]]s.<ref name=Watson_2004 /> (Although in prokaryotes and viruses, the rate is much higher.) Rare, spontaneous alterations in the base sequence of a particular gene arise from a number of sources, such as errors in [[DNA replication]] and the aftermath of [[DNA damage]]. These errors are called [[mutation]]s. The cell contains many [[DNA repair]] mechanisms for preventing mutations and maintaining the integrity of the genome; however, in some cases—such as breaks in both DNA strands of a chromosome—repairing the physical damage to the molecule is a higher priority than producing an exact copy. Due to the degeneracy of the genetic code, some mutations in protein-coding genes are ''silent'', or produce no change in the [[Peptide sequence|amino acid sequence]] of the protein for which they code; for example, the codons [[codon|UCU]] and [[codon|UCC]] both code for [[serine]], so the U→C mutation has no effect on the protein. Mutations that do have phenotypic effects are most often neutral or deleterious to the organism. Variants may confer benefits to the organism's [[fitness (biology)|fitness]]; it is commonly thought that mutations may produce beneficial variants. The most common mutations include point mutations in which a single codon is replaced, frame shift mutation where a single nucleotide base is inserted or deleted from the DNA strand so that all bases are shifted over, silent mutations where a single nucleotide base is replaced but without causing a change for the amino acid being coded for, and nonsense mutations, where a change in a single nucleotide base causes a codon to be turned into a stop codon hence terminating transcription at this point.

Mutations propagated to the next [[generation]] lead to variations within a species' population. Variants of a single gene are known as [[allele]]s, and differences in [[allele]]s may give rise to differences in traits. Although it is rare for the variants in a single gene to have clearly distinguishable phenotypic effects, certain well-defined traits are in fact controlled by single genetic loci. A gene's most common allele is called the [[wild type]] allele, and rare alleles are called [[mutant]]s. However, this does not imply that the wild-type allele is the [[ancestor]] from which the [[mutant]]s are descended. For the most part, these mutations are recessive and are phased out quickly. However, on occasion these mutations appear as dominant to other alleles, becoming predominant and increasing in the rate they are seen in a population.

==Genome==

===Chromosomal organization===
The total complement of genes in an organism or cell is known as its [[genome]]. In [[prokaryote]]s, the vast majority of genes are located on a single chromosome of [[circular DNA]], while [[eukaryote]]s usually possess multiple individual linear DNA helices packed into dense DNA-protein complexes called [[chromosome]]s. Genes that appear together on one chromosome of one species may appear on separate chromosomes in another species. Many species carry more than one copy of their genome within each of their [[somatic cell]]s. Cells or organisms with only one copy of each chromosome are called [[haploid]]; those with two copies are called [[diploid]]; and those with more than two copies are called [[polyploidy|polyploid]]. The copies of genes on the chromosomes are not necessarily identical. In sexually reproducing organisms, one copy is normally inherited from each parent.

===Number of genes===
[[File:Human genome by functions.svg|thumb|480px|The protein-encoding component of the [[human genome]], categorized by function of each gene product, given both as number of genes and as percentage of all genes.<ref>[http://www.pantherdb.org/chart/summary/pantherChart.jsp?filterLevel=1&chartType=1&listType=1&type=5&species=Homo%20sapiens PANTHER Pie Chart] at the PANTHER Classification System homepage. Retrieved 25 May 2011</ref>]]

Early estimates of the number of human genes that used [[expressed sequence tag]] data put it at 50 000–100 000.<ref>{{vcite2 journal | vauthors = Schuler GD, Boguski MS, Stewart EA, Stein LD, Gyapay G, Rice K, White RE, Rodriguez-Tomé P, Aggarwal A, Bajorek E, Bentolila S, Birren BB, Butler A, Castle AB, Chiannilkulchai N, Chu A, Clee C, Cowles S, Day PJ, Dibling T, Drouot N, Dunham I, Duprat S, East C, Edwards C, Fan JB, Fang N, Fizames C, Garrett C, Green L, Hadley D, Harris M, Harrison P, Brady S, Hicks A, Holloway E, Hui L, Hussain S, Louis-Dit-Sully C, Ma J, MacGilvery A, Mader C, Maratukulam A, Matise TC, McKusick KB, Morissette J, Mungall A, Muselet D, Nusbaum HC, Page DC, Peck A, Perkins S, Piercy M, Qin F, Quackenbush J, Ranby S, Reif T, Rozen S, Sanders C, She X, Silva J, Slonim DK, Soderlund C, Sun WL, Tabar P, Thangarajah T, Vega-Czarny N, Vollrath D, Voyticky S, Wilmer T, Wu X, Adams MD, Auffray C, Walter NA, Brandon R, Dehejia A, Goodfellow PN, Houlgatte R, Hudson JR, Ide SE, Iorio KR, Lee WY, Seki N, Nagase T, Ishikawa K, Nomura N, Phillips C, Polymeropoulos MH, Sandusky M, Schmitt K, Berry R, Swanson K, Torres R, Venter JC, Sikela JM, Beckmann JS, Weissenbach J, Myers RM, Cox DR, James MR, Bentley D, Deloukas P, Lander ES, Hudson TJ | title = A gene map of the human genome | journal = Science | volume = 274 | issue = 5287 | pages = 540–6 | date = October 1996 | pmid = 8849440 | doi = 10.1126/science.274.5287.540 | url = http://www.sciencemag.org/cgi/content/full/274/5287/540 | bibcode = 1996Sci...274..540S }}</ref> Following the [[Human Genome Project|sequencing of the human genome]] and other genomes, it has been found that rather few genes (~20 000 in human, mouse and fly, ~13 000 in roundworm, >46,000 in rice<ref name="pmid11935017">{{vcite2 journal | vauthors = Yu J, Hu S, Wang J, Wong GK, Li S, Liu B, Deng Y, Dai L, Zhou Y, Zhang X, Cao M, Liu J, Sun J, Tang J, Chen Y, Huang X, Lin W, Ye C, Tong W, Cong L, Geng J, Han Y, Li L, Li W, Hu G, Huang X, Li W, Li J, Liu Z, Li L, Liu J, Qi Q, Liu J, Li L, Li T, Wang X, Lu H, Wu T, Zhu M, Ni P, Han H, Dong W, Ren X, Feng X, Cui P, Li X, Wang H, Xu X, Zhai W, Xu Z, Zhang J, He S, Zhang J, Xu J, Zhang K, Zheng X, Dong J, Zeng W, Tao L, Ye J, Tan J, Ren X, Chen X, He J, Liu D, Tian W, Tian C, Xia H, Bao Q, Li G, Gao H, Cao T, Wang J, Zhao W, Li P, Chen W, Wang X, Zhang Y, Hu J, Wang J, Liu S, Yang J, Zhang G, Xiong Y, Li Z, Mao L, Zhou C, Zhu Z, Chen R, Hao B, Zheng W, Chen S, Guo W, Li G, Liu S, Tao M, Wang J, Zhu L, Yuan L, Yang H | title = A draft sequence of the rice genome (Oryza sativa L. ssp. indica) | journal = Science | volume = 296 | issue = 5565 | pages = 79–92 | date = April 2002 | pmid = 11935017 | doi = 10.1126/science.1068037 | bibcode = 2002Sci...296...79Y }}</ref>) encode all the [[protein]]s in an organism.<ref name=Carninci2007>{{vcite2 journal | vauthors = Carninci P, Hayashizaki Y | title = Noncoding RNA transcription beyond annotated genes | journal = Current Opinion in Genetics & Development | volume = 17 | issue = 2 | pages = 139–44 | date = April 2007 | pmid = 17317145 | doi = 10.1016/j.gde.2007.02.008 }}</ref> These protein-coding sequences make up 1–2% of the human genome.<ref name=Claverie2005>{{vcite2 journal | vauthors = Claverie JM | title = Fewer genes, more noncoding RNA | journal = Science | volume = 309 | issue = 5740 | pages = 1529–30 | date = September 2005 | pmid = 16141064 | doi = 10.1126/science.1116800 | bibcode = 2005Sci...309.1529C }}</ref> A large part of the genome is transcribed however, to [[intron]]s, [[retrotransposon]]s and seemingly a large array of [[noncoding RNA]]s.<ref name=Carninci2007/><ref name=Claverie2005/> Total number of proteins (the Earth's [[proteome]]) is estimated to be 5 million sequences.<ref>{{vcite2 journal | vauthors = Perez-Iratxeta C, Palidwor G, Andrade-Navarro MA | title = Towards completion of the Earth's proteome | journal = EMBO Reports | volume = 8 | issue = 12 | pages = 1135–1141 | date = December 2007 | pmid = 18059312 | pmc = 2267224 | doi = 10.1038/sj.embor.7401117 | url = http://www.nature.com/embor/journal/v8/n12/full/7401117.html | first2 = Gareth | first3 = Miguel A }}</ref>

===Genetic and genomic nomenclature===
[[Gene nomenclature]] has been established by the [[Human Genome Organisation|HUGO]] Gene Nomenclature Committee ([[HGNC]]) for each known human gene in the form of an approved gene name and [[symbol]] (short-form [[abbreviation]]). All approved symbols are stored in the [http://www.genenames.org/cgi-bin/search?search_type=symbols&search=&submit=Submit HGNC Database]. Each symbol is unique and each gene is only given one approved gene symbol. This also facilitates [[electronics|electronic]] [[data]] retrieval from publications. In preference each symbol maintains parallel construction in different members of a [[gene family]] and can be used in other [[species]], especially the [[mouse]].

===Essential genes===
{{main| Essential genes}}
Essential genes are those genes of an organism that are thought to be critical for its survival. Surprisingly few genes have been shown to be absolutely essential for the survival of bacteria, e.g. only about 10% of the ~4,200 genes of [[Escherichia coli]].

==Evolutionary concept of a gene==
[[George C. Williams]] first explicitly advocated the [[gene-centered view of evolution|gene-centric view of evolution]] in his 1966 book ''[[Adaptation and Natural Selection]]''. He proposed an evolutionary concept of gene to be used when we are talking about [[natural selection]] favoring some genes. The definition is: "that which segregates and recombines with appreciable frequency." According to this definition, even an [[asexual reproduction|asexual]] genome could be considered a gene, insofar that it have an appreciable permanency through many generations.

The difference is: the molecular gene ''transcribes'' as a unit, and the evolutionary gene ''inherits'' as a unit.

[[Richard Dawkins]]' books ''[[The Selfish Gene]]'' (1976) and ''[[The Extended Phenotype]]'' (1982) defended the idea that the gene is the only [[DNA replication|replicator]] in living systems. This means that only genes transmit their structure largely intact and are potentially immortal in the form of copies. So, genes should be the [[unit of selection]]. In ''[[River Out of Eden]]'', Dawkins further refined the idea of gene-centric selection by describing life as a river of compatible genes flowing through [[geological time]]. Scoop up a bucket of genes from the river of genes, and we have an [[organism]] serving as temporary bodies or [[survival machine]]s. A river of genes may fork into two branches representing two non-[[Hybrid (biology)|interbreeding]] [[species]] as a result of geographical separation.

==Gene targeting and implications==
{{main|Gene targeting}}

Gene targeting is commonly referred to techniques for altering or disrupting mouse genes and provides the mouse models for studying the roles of individual genes in [[embryonic development]], human disorders, aging and diseases. The mouse models, where one or more of its genes are deactivated or made inoperable, are called [[knockout mice]]. Since the first reports in which [[homologous recombination]] among [[homologous chromosome]]s in [[embryonic stem cell]]s was used to generate gene-targeted mice,<ref>{{vcite2 journal | vauthors = Thomas KR, Capecchi MR | title = Site-directed mutagenesis by gene targeting in mouse embryo-derived stem cells | journal = Cell | volume = 51 | issue = 3 | pages = 503–12 | date = November 1987 | pmid = 2822260 | doi = 10.1016/0092-8674(87)90646-5 }}</ref> gene targeting has proven to be a powerful means of precisely manipulating the mammalian genome, producing at least ten thousand mutant mouse strains and it is now possible to introduce mutations that can be activated at specific time points, or in specific cells or organs, both during development and in the adult animal.<ref>[http://nobelprize.org/nobel_prizes/medicine/laureates/2007/press.html The 2007 Nobel Prize in Physiology or Medicine&nbsp;– Press Release<!-- Bot generated title -->]</ref><ref name=Deng_2007>{{vcite2 journal | vauthors = Deng C | title = In celebration of Dr. Mario R. Capecchi's Nobel Prize | journal = International Journal of Biological Sciences | volume = 3 | issue = 7 | pages = 417–419 | year = 2007 | pmid = 17998949 | doi = 10.7150/ijbs.3.417 | url = http://www.biolsci.org/v03p0417.htm }}</ref>

Gene targeting strategies have been expanded to all kinds of modifications, including [[point mutation]]s, isoform deletions, mutant allele correction, large pieces of chromosomal DNA [[Genetic insertion|insertion]] and [[Deletion (genetics)|deletion]], tissue specific disruption combined with spatial and temporal regulation and so on. It is predicted that the ability to generate mouse models with predictable phenotypes will have a major impact on studies of all phases of development, [[immunology]], [[neurobiology]], [[oncology]], [[physiology]], [[metabolism]], and human diseases. Gene targeting is also in theory applicable to species from which [[Totipotency|totipotent]] embryonic stem cells can be established, and therefore may offer a potential to the improvement of domestic animals and plants.<ref name="Deng_2007"/><ref>[http://www.hhmi.org/research/investigators/capecchi.html Mario R. Capecchi<!-- Bot generated title -->]</ref>

==Changing concept==
The concept of the gene has changed considerably (see [[gene#History|history section]]). From the original definition of a "unit of inheritance", the term evolved to mean a [[DNA]]-based unit that can exert its effects on the organism through [[RNA]] or [[protein]] products. It was also previously believed that one gene makes one protein; this concept was overthrown by the discovery of [[alternative splicing]] and [[trans-splicing]].<ref name="Gerstein"/>

The definition of a gene is still changing. The first cases of RNA-based [[biological inheritance|inheritance]] have been discovered in mammals.<ref name="rass">{{vcite2 journal | vauthors = Rassoulzadegan M, Grandjean V, Gounon P, Vincent S, Gillot I, Cuzin F | title = RNA-mediated non-mendelian inheritance of an epigenetic change in the mouse | journal = Nature | volume = 441 | issue = 7092 | pages = 469–74 | date = May 2006 | pmid = 16724059 | doi = 10.1038/nature04674 | bibcode = 2006Natur.441..469R }}</ref> Evidence is also accumulating that the [[Enhancer (genetics)|control regions]] of a gene do not necessarily have to be close to the [[coding sequence]] on the linear molecule or even on the same chromosome. Spilianakis and colleagues discovered that the [[promoter region]] of the [[interferon-gamma]] gene on chromosome 10 and the regulatory regions of the T(H)2 [[cytokine]] locus on chromosome 11 come into close proximity in the [[cell nucleus|nucleus]] possibly to be jointly regulated.<ref>{{vcite2 journal | vauthors = Spilianakis CG, Lalioti MD, Town T, Lee GR, Flavell RA | title = Interchromosomal associations between alternatively expressed loci | journal = Nature | volume = 435 | issue = 7042 | pages = 637–45 | date = June 2005 | pmid = 15880101 | doi = 10.1038/nature03574 | bibcode = 2005Natur.435..637S }}</ref> Even the coding sequence of a gene itself doesn't have to be all on the same chromosome: Marande and Burger showed that, in the [[mitochondria]] of the [[protist]] ''Diplonema papillatum'', "genes are systematically fragmented into small pieces that are encoded on separate chromosomes, transcribed individually, and then concatenated into contiguous messenger RNA molecules".<ref>{{vcite2 journal | vauthors = Marande W, Burger G | title = Mitochondrial DNA as a genomic jigsaw puzzle | journal = Science | volume = 318 | issue = 5849 | pages = 415 | date = October 2007 | pmid = 17947575 | doi = 10.1126/science.1148033 | publisher = AAAS | bibcode = 2007Sci...318..415M }}</ref>

The concept that genes are clearly delimited is also being eroded. There is evidence for fused proteins stemming from two adjacent genes that can produce two separate protein products. While it is not clear whether these fusion proteins are functional, the phenomenon is more frequent than previously thought.<ref>{{vcite2 journal | vauthors = Parra G, Reymond A, Dabbouseh N, Dermitzakis ET, Castelo R, Thomson TM, Antonarakis SE, Guigó R | title = Tandem chimerism as a means to increase protein complexity in the human genome | journal = Genome Research | volume = 16 | issue = 1 | pages = 37–44 | date = January 2006 | pmid = 16344564 | pmc = 1356127 | doi = 10.1101/gr.4145906 }}</ref> Even more ground-breaking than the discovery of fused genes is the observation that some proteins can be composed of [[exon]]s from far away regions and even different chromosomes.<ref name="Rethink"/><ref>{{vcite2 journal | vauthors = Kapranov P, Drenkow J, Cheng J, Long J, Helt G, Dike S, Gingeras TR | title = Examples of the complex architecture of the human transcriptome revealed by RACE and high-density tiling arrays | journal = Genome Research | volume = 15 | issue = 7 | pages = 987–97 | date = July 2005 | pmid = 15998911 | pmc = 1172043 | doi = 10.1101/gr.3455305 }}</ref> This new data has led to an updated, and probably tentative, definition of a gene as "a union of genomic sequences encoding a coherent set of potentially overlapping functional products".<ref name="Gerstein">{{vcite2 journal | vauthors = Gerstein MB, Bruce C, Rozowsky JS, Zheng D, Du J, Korbel JO, Emanuelsson O, Zhang ZD, Weissman S, Snyder M | title = What is a gene, post-ENCODE? History and updated definition | journal = Genome Research | volume = 17 | issue = 6 | pages = 669–681 | date = June 2007 | pmid = 17567988 | doi = 10.1101/gr.6339607 | first2 = C. | first3 = J. S. }}</ref> This new definition categorizes genes by functional products, whether they be proteins or RNA, rather than specific DNA loci; all regulatory elements of DNA are therefore classified as ''gene-associated'' regions.<ref name="Gerstein"/>

== See also ==
{{Columns-list|2|
* [[Copy number variation]]
* [[DNA]]
* [[Epigenetics]]
* [[Full genome sequencing]]
* [[Gene-centered view of evolution|Gene-centric view of evolution]] {{nb5}}
* [[Gene dosage]]
* [[Gene expression]]
* [[Gene family]]
* [[Gene nomenclature]]
* [[Gene patent]]
* [[Gene pool]]
* [[Gene redundancy]]
* [[Gene therapy]]
* [[Genetic algorithm]]
* [[Genetic engineering]] {{nb10}}{{nb5}}
* [[Genetics]]
* [[Genomics]]
* [[List of gene prediction software]]
* [[List of notable genes]]
* [[Population genetics]]
* [[Predictive medicine]]
* [[Pseudogene]]
}}

==Notes and references==
{{Reflist|30em}}

==Bibliography==
* {{cite book | first1 = Bruce | last1 = Alberts | first2 = Alexander | last2 = Johnson | first3 = Julian | last3 = Lewis | first4 = Martin | last4 = Raff | first5 = Keith | last5 = Roberts | first6 = Peter | last6 = Walter | name-list-format = vanc | author1-link = Bruce Alberts | author3-link = Julian Lewis (biologist) | author4-link = Martin Raff | author6-link = Peter Walter | title = Molecular Biology of the Cell | edition = Fourth | publisher = Garland Science | location = New York | year = 2002 | isbn = 978-0-8153-3218-3 | url = http://www.ncbi.nlm.nih.gov/books/NBK21054/ }}
* {{cite book | first1 = James D. | last1 = Watson | first2 = Tania A. | last2 = Baker | first3 = Stephen P. | last3 = Bell | first4 = Alexander | last4 = Gann | first5 = Michael | last5 = Levine | first6 = Richard | last6 = Losick | name-list-format = vanc | author1-link = James Watson | author2-link = Tania A. Baker | author-link5 = Michael Levine (biologist) | title = Molecular Biology of the Gene | edition = 7th | date = 2013 | publisher = Benjamin Cummings | location = | isbn = 978-0-321-90537-6 }}
* {{cite book | first = Richard | last = Dawkins | authorlink = Richard Dawkins | title = [[The Selfish Gene]] | publisher = Oxford University Press | year = 1990 | isbn = 0-19-286092-5 | name-list-format = vanc }} [http://books.google.com/print?id=WkHO9HI7koEC Google Book Search]; first published 1976.
* {{cite book | first = Richard | last = Dawkins | authorlink = Richard Dawkins | title = [[River Out of Eden]] | publisher = Basic Books | year = 1995 | isbn = 0-465-06990-8 | name-list-format = vanc }}
* {{cite book | first = Matt | last = Ridley | authorlink = Matt Ridley | title = [[Genome: The Autobiography of a Species in 23 Chapters]] | publisher = Fourth Estate | year = 1999 | isbn = 0-00-763573-7 | name-list-format = vanc }}
* {{cite book | author = Hartwell L, Hood L, Goldberg ML, Reynolds AE, Silver LM, Veres R | title = Genetics: from genes to genomes | edition = Second | publisher = McGraw-Hill Higher Education | location = Boston | year = 2004 | origyear = | pages = | quote = | isbn = 0-07-291930-2 }}
* {{vcite2 journal | vauthors = Guerzoni D, McLysaght A | title = De novo origins of human genes | journal = PLoS Genetics | volume = 7 | issue = 11 | date = November 2011 | pmid = 22102832 | doi = 10.1371/journal.pgen.1002381 }}

== External links ==
{{Spoken Wikipedia|Gene.ogg|2005-04-21}}
* [http://ctdbase.org/ Comparative Toxicogenomics Database]
* [http://www.dnaftb.org/ DNA From The Beginning&nbsp;– a primer on genes and DNA]
* [http://www.bioinformaticstutorials.com/?p=6 Genes And DNA&nbsp;– Introduction to genes and DNA aimed at non-biologist]
* [http://www.ncbi.nlm.nih.gov/sites/entrez?db=gene Entrez Gene&nbsp;– a searchable database of genes]
* [http://idconverter.bioinfo.cnio.es/ IDconverter&nbsp;– converts gene IDs between public databases]
* [http://www.ihop-net.org/UniPub/iHOP/ iHOP&nbsp;– Information Hyperlinked over Proteins]
* [http://tagc.univ-mrs.fr/tbrowser TranscriptomeBrowser&nbsp;– Gene expression profile analysis]
* [http://www.jcvi.org/pn-utility The Protein Naming Utility, a database to identify and correct deficient gene names]
* [http://www.mdpi.com/journal/genes/ ''Genes'']&nbsp;– an Open Access journal
* [http://www.mousephenotype.org/ IMPC (International Mouse Phenotyping Consortium)]&nbsp;– Encyclopedia of mammalian gene function
* [http://www.globalgenes.org/ Global Genes Project]&nbsp;– Leading non-profit organization supporting people living with genetic diseases
* [http://www.nature.com/encode/#/threads/characterization-of-intergenic-regions-and-gene-definition ENCODE threads Explorer] Characterization of intergenic regions and gene definition. [[Nature (journal)]]

[[Category:Cloning]]
[[Category:Genes| ]]
[[Category:Molecular biology]]

Revision as of 20:06, 7 April 2015

This stylistic diagram shows a gene in relation to the double helix structure of DNA and to a chromosome (right). The chromosome is X-shaped because it is dividing. Introns are regions often found in eukaryote genes that are removed in the splicing process (after the DNA is transcribed into RNA): Only the exons encode the protein. The diagram labels a region of only 55 or so bases as a gene. In reality, most genes are hundreds of times longer.
The chemical structure of a four-base fragment of a DNA double helix.

A gene is the molecular unit of heredity of a living organism. The word is used extensively by the scientific community for stretches of deoxyribonucleic acids (DNA) and ribonucleic acids (RNA) that code for a polypeptide or for an RNA chain that has a function in the organism. Living beings depend on genes, as they specify all proteins and functional RNA chains. Genes hold the information to build and maintain an organism's cells and pass genetic traits to offspring. All organisms have genes corresponding to various biological traits, some of which are instantly visible, such as eye color or number of limbs, and some of which are not, such as blood type, increased risk for specific diseases, or the thousands of basic biochemical processes that comprise life. The word gene was coined by Wilhelm Johannsen in 1909 and is indirectly derived (via pangene) from the Ancient Greek word γένος (génos) meaning "race, offspring".[1]

A modern working definition of a gene is "a locatable region of genomic sequence, corresponding to a unit of inheritance, which is associated with regulatory regions, transcribed regions, and/or other functional sequence regions ".[2][3] Colloquial usage of the term gene (e.g., "good genes", "hair color gene") may actually refer to an allele: a gene is the basic instruction— a sequence of nucleic acids (DNA or, in the case of certain viruses RNA), while an allele is one variant of that gene. Thus, when the mainstream press refers to "having a gene" for a specific trait, this is customarily inaccurate. In most cases, all people would have a gene for the trait in question, although certain people will have a specific allele of that gene, which results in the trait variant. Further, genes code for proteins, which might result in identifiable traits, but it is the gene (genotype), not the trait (phenotype), which is inherited.

Big genes are a class of genes whose nuclear transcript spans 500 kb (1 kb = 1,000 base pairs) or more of chromosomal DNA. The largest of the big genes is the gene for dystrophin, which spans 2.3 Mb. Many big genes have modestly sized mRNAs; the exons encoding these RNAs typically encompass about 1% of the total chromosomal gene region in which they occur.

History

Gregor Mendel

The existence of discrete inheritable units was first suggested by Gregor Mendel (1822–1884). From 1857 to 1864, he studied inheritance patterns in 8000 common edible pea plants (Pisum), tracking distinct traits from parent to offspring. He described these mathematically as 2n combinations where n is the number of differing characteristics in the original peas. Although he did not use the term gene, he explained his results in terms of discrete inherited units that give rise to distinct phenotypes. Mendel was also the first to show independent assortment, the distinction between dominant and recessive traits, the distinction between a heterozygote and homozygote, the phenomenon of discontinuous inheritance and what would later be described as genotype (the genetic material of an organism) and phenotype (the visible traits of that organism) and the conversion of one form into another within few generations.

Prior to Mendel's work, the dominant theory of heredity was one of blending inheritance, which suggested that each parent contributed fluids to the fertilisation process and that the traits of the parents blended and mixed to produce the offspring. Charles Darwin developed a theory of inheritance he termed pangenesis, which used the term gemmule to describe hypothetical particles that would mix during reproduction.

Although Mendel's work was largely unrecognized after its first publication in 1866, it was 'rediscovered' in 1900 by three European scientists, Hugo de Vries, Carl Correns, and Erich von Tschermak, who claimed to have reached similar conclusions in their own research. Danish botanist Wilhelm Johannsen coined the word "gene" ("gen" in Danish and German) in 1909 to describe the fundamental physical and functional units of heredity,[4] while the related word genetics was first used by William Bateson in 1905.[5] In 1910, Thomas Hunt Morgan showed that genes reside on specific chromosomes. He later showed that genes occupy specific locations on the chromosome. With this knowledge, Morgan and his students began the first chromosomal map of the fruit fly Drosophila. In 1928, Frederick Griffith showed that genes could be transferred. In what is now known as Griffith's experiment, injections into a mouse of a deadly strain of bacteria that had been heat-killed transferred genetic information to a safe strain of the same bacteria, killing the mouse.

A series of subsequent discoveries led to the realization decades later that the genetic material is made of DNA (deoxyribonucleic acid). In 1941, George Wells Beadle and Edward Lawrie Tatum showed that mutations in genes caused errors in specific steps in metabolic pathways. This showed that specific genes code for specific proteins, leading to the "one gene, one enzyme" hypothesis.[5] Oswald Avery, Colin Munro MacLeod, and Maclyn McCarty showed in 1944 that DNA holds the gene's information.[6] In 1952, Rosalind Franklin and Raymond Gosling produced a strikingly clear x-ray diffraction pattern indicating a helical form, and in 1953, James D. Watson and Francis Crick demonstrated the molecular structure of DNA. Together, these discoveries established the central dogma of molecular biology, which states that proteins are translated from RNA which is transcribed from DNA. This dogma has since been shown to have exceptions, such as reverse transcription in retroviruses.

In 1972, Walter Fiers and his team at the University of Ghent were the first to determine the sequence of a gene: the gene for Bacteriophage MS2 coat protein.[7] Richard J. Roberts and Phillip Sharp discovered in 1977 that genes can be split into segments. This led to the idea that one gene can make several proteins. The successful sequencing of many organisms' genomes has complicated the molecular definition of genes. In particular, genes do not seem to sit side by side on DNA like discrete beads. Instead, regions of the DNA producing distinct proteins may overlap, so that the idea emerges that "genes are one long continuum".[2] It was first hypothesized in 1986 by Walter Gilbert that neither DNA nor protein would be required in such a primitive system as that of a very early stage of the earth if RNA could perform as simply a catalyst and genetic information storage processor.

The modern study of genetics at the level of DNA is known as molecular genetics and the synthesis of molecular genetics with traditional Darwinian evolution is known as the modern evolutionary synthesis.

Mendelian inheritance and classical genetics

Crossing between two pea plants heterozygous for purple (B, dominant) and white (b, recessive) blossoms

According to the theory of Mendelian inheritance, variations in phenotype—the observable physical and behavioral characteristics of an organism—are due in part to variations in genotype, or the organism's particular set of genes, each of which specifies a particular trait. Different forms of a gene, which may give rise to different phenotypes, are known as alleles. Organisms such as the pea plants Mendel worked on, along with many plants and animals, have two alleles for each trait, one inherited from each parent. Alleles may be dominant or recessive; dominant alleles give rise to their corresponding phenotypes when paired with any other allele for the same trait, whereas recessive alleles give rise to their corresponding phenotype only when paired with another copy of the same allele. For example, if the allele specifying tall stems in pea plants is dominant over the allele specifying short stems, then pea plants that inherit one tall allele from one parent and one short allele from the other parent will also have tall stems. Mendel's work demonstrated that alleles assort independently in the production of gametes, or germ cells, ensuring variation in the next generation.

Physical definitions

Functional structure of a gene

Diagram of the "typical" eukaryotic protein-coding gene. Promoters and enhancers determine what portions of the DNA will be transcribed into the precursor mRNA (pre-mRNA). The pre-mRNA is then spliced into messenger RNA (mRNA) which is later translated into protein.

The vast majority of living organisms encode their genes in long strands of DNA (deoxyribonucleic acid). DNA consists of a chain made from four types of nucleotide subunits, each composed of: a five-carbon sugar (2'-deoxyribose), a phosphate group, and one of the four bases adenine, cytosine, guanine, and thymine. The most common form of DNA in a cell is in a double helix structure, in which two individual DNA strands twist around each other in a right-handed spiral. In this structure, the base pairing rules specify that guanine pairs with cytosine and adenine pairs with thymine. The base pairing between guanine and cytosine forms three hydrogen bonds, whereas the base pairing between adenine and thymine forms two hydrogen bonds. The two strands in a double helix must therefore be complementary, that is, their bases must align such that the adenines of one strand are paired with the thymines of the other strand, and so on.

Due to the chemical composition of the pentose residues of the bases, DNA strands have directionality. One end of a DNA polymer contains an exposed hydroxyl group on the deoxyribose; this is known as the 3' end of the molecule. The other end contains an exposed phosphate group; this is the 5' end. The directionality of DNA is vitally important to many cellular processes, since double helices are necessarily directional (a strand running 5'-3' pairs with a complementary strand running 3'-5'), and processes such as DNA replication occur in only one direction. All nucleic acid synthesis in a cell occurs in the 5'-3' direction, because new monomers are added via a dehydration reaction that uses the exposed 3' hydroxyl as a nucleophile.

The expression of genes encoded in DNA begins by transcribing the gene into RNA, a second type of nucleic acid that is very similar to DNA, but whose monomers contain the sugar ribose rather than deoxyribose. RNA also contains the base uracil in place of thymine. RNA molecules are less stable than DNA and are typically single-stranded. Genes that encode proteins are composed of a series of three-nucleotide sequences called codons, which serve as the words in the genetic language. The genetic code specifies the correspondence during protein translation between codons and amino acids. The genetic code is nearly the same for all known organisms.

All genes have regulatory regions in addition to regions that explicitly code for a protein or RNA product. A regulatory region shared by almost all genes is known as the promoter, which provides a position that is recognized by the transcription machinery when a gene is about to be transcribed and expressed. A gene can have more than one promoter, resulting in RNAs that differ in how far they extend in the 5' end.[8] Although promoter regions have a consensus sequence that is the most common sequence at this position, some genes have "strong" promoters that bind the transcription machinery well, and others have "weak" promoters that bind poorly. These weak promoters usually permit a lower rate of transcription than the strong promoters, because the transcription machinery binds to them and initiates transcription less frequently. Other possible regulatory regions include enhancers, which can compensate for a weak promoter. Most regulatory regions are "upstream"—that is, before or toward the 5' end of the transcription initiation site. Eukaryotic promoter regions are much more complex and difficult to identify than prokaryotic promoters.

Many prokaryotic genes are organized into operons, or groups of genes whose products have related functions and which are transcribed as a unit. By contrast, eukaryotic genes are transcribed only one at a time, but may include long stretches of DNA called introns which are transcribed but never translated into protein (they are spliced out before translation). Splicing can also occur in prokaryotic genes, but is less common than in eukaryotes.[9]

Chromosomes

The total complement of genes in an organism or cell is known as its genome, which may be stored on one or more chromosomes; the region of the chromosome at which a particular gene is located is called its locus. A chromosome consists of a single, very long DNA helix on which thousands of genes are encoded. Prokaryotesbacteria and archaea—typically store their genomes on a single large,circular chromosome, sometimes supplemented by additional small circles of DNA called plasmids, which usually encode only a few genes and are easily transferable between individuals. For example, the genes for antibiotic resistance are usually encoded on bacterial plasmids and can be passed between individual cells, even those of different species, via horizontal gene transfer. Although some simple eukaryotes also possess plasmids with small numbers of genes, the majority of eukaryotic genes are stored on multiple linear chromosomes, which are packed within the nucleus in complex with storage proteins called histones. The manner in which DNA is stored on the histone, as well as chemical modifications of the histone itself, are regulatory mechanisms governing whether a particular region of DNA is accessible for gene expression. The ends of eukaryotic chromosomes are capped by long stretches of repetitive sequences called telomeres, which do not code for any gene product but are present to prevent degradation of coding and regulatory regions during DNA replication. The length of the telomeres tends to decrease each time the genome is replicated in preparation for cell division; the loss of telomeres has been proposed as an explanation for cellular senescence, or the loss of the ability to divide, and by extension for the aging process in organisms.[10]

Whereas the chromosomes of prokaryotes are relatively gene-dense, those of eukaryotes often contain so-called "junk DNA", or regions of DNA that serve no obvious function. Simple single-celled eukaryotes have relatively small amounts of such DNA, whereas the genomes of complex multicellular organisms, including humans, contain an absolute majority of DNA without an identified function.[11] However, it now appears that, although protein-coding DNA makes up barely 2% of the human genome, about 80% of the bases in the genome may be expressed, so the term "junk DNA" may be a misnomer.[3]

RNA genes

Protein coding genes are transcribed to an mRNA intermediate, then translated to a functional protein. RNA-coding genes are transcribed to a functional non-coding RNA. (PDB: 3BSE​, 1OBB​, 3TRA​)

When proteins are manufactured, the gene is first copied into RNA as an intermediate product.[12] In other cases, the RNA molecules are the actual functional products, as in the synthesis of ribosomal RNA and transfer RNA. Some RNAs known as ribozymes are capable of enzymatic function, and microRNA has a regulatory role. The DNA sequences from which such RNAs are transcribed are known as RNA genes.

Some viruses store their entire genomes in the form of RNA, and contain no DNA at all. Because they use RNA to store genes, their cellular hosts may synthesize their proteins as soon as they are infected and without the delay in waiting for transcription. On the other hand, RNA retroviruses, such as HIV, require the reverse transcription of their genome from RNA into DNA before their proteins can be synthesized. RNA-mediated epigenetic inheritance has also been observed in plants and very rarely in animals.[13]

Gene expression

In all organisms, there are two major steps separating a protein-coding gene from its protein: First, the DNA on which the gene resides must be transcribed from DNA to messenger RNA (mRNA); and, second, it must be translated from mRNA to protein. RNA-coding genes must still go through the first step, but are not translated into protein. The process of producing a biologically functional molecule of either RNA or protein is called gene expression, and the resulting molecule itself is called a gene product.

Genetic code

Schematic diagram of a single-stranded RNA molecule illustrating the position of three-base codons.

The genetic code is the set of rules by which information encoded within a gene is translated into a functional protein. Each gene consists of a specific sequence of nucleotides encoded in DNA or RNA. The nucleotide being made up of a sugar, a phosphate molecule and a specific base (adenine, thymine, cytosine, guanine or sometimes uracil; thymine is replaced with uracil in some viruses[14]); a correspondence between nucleotides, the basic building blocks of genetic material, and amino acids, the basic building blocks of proteins, must be established for genes to be successfully translated into functional proteins. Sets of three nucleotides, known as codons, each correspond to a specific amino acid or to a signal; three codons are known as "stop codons" and, instead of specifying a new amino acid, alert the translation machinery that the end of the gene has been reached, just as a specific set of 3 bases, "AUG", known as the "start codon", signifies the gene to start transcribing. There are 64 possible codons (four possible nucleotides at each of three positions, hence 43 possible codons) and only 20 standard amino acids; hence the code is redundant and multiple codons can specify the same amino acid. The correspondence between codons and amino acids is nearly universal among all known living organisms.

Transcription

The process of genetic transcription produces a single-stranded RNA molecule known as messenger RNA, whose nucleotide sequence is complementary to the DNA from which it was transcribed. The DNA strand whose sequence matches that of the RNA is known as the coding strand and the strand from which the RNA was synthesized is the template strand. Transcription is performed by an enzyme called an RNA polymerase, which reads the template strand in the 3' to 5' direction and synthesizes the RNA from 5' to 3'. To initiate transcription, the polymerase first recognizes and binds a promoter region of the gene. Thus, a major mechanism of gene regulation is the blocking or sequestering of the promoter region, either by tight binding by repressor molecules that physically block the polymerase, or by organizing the DNA so that the promoter region is not accessible.

In prokaryotes, transcription occurs in the cytoplasm; for very long transcripts, translation may begin at the 5' end of the RNA while the 3' end is still being transcribed. In eukaryotes, transcription necessarily occurs in the nucleus, where the cell's DNA is sequestered; the RNA molecule produced by the polymerase is known as the primary transcript and must undergo post-transcriptional modifications before being exported to the cytoplasm for translation. The splicing of introns present within the transcribed region is a modification unique to eukaryotes; alternative splicing mechanisms can result in mature transcripts from the same gene having different sequences and thus coding for different proteins. This is a major form of regulation in eukaryotic cells.

Translation

Translation is the process by which a mature mRNA molecule is used as a template for synthesizing a new protein. Translation is carried out by ribosomes, large complexes of RNA and protein responsible for carrying out the chemical reactions to add new amino acids to a growing polypeptide chain by the formation of peptide bonds. The genetic code is read three nucleotides at a time, in units called codons, via interactions with specialized RNA molecules called transfer RNA (tRNA). Each tRNA has three unpaired bases known as the anticodon that are complementary to the codon it reads; the tRNA is also covalently attached to the amino acid specified by the complementary codon. When the tRNA binds to its complementary codon in an mRNA strand, the ribosome ligates its amino acid cargo to the new polypeptide chain, which is synthesized from amino terminus to carboxyl terminus. During and after its synthesis, the new protein must fold to its active three-dimensional structure before it can carry out its cellular function.

DNA replication and inheritance

The growth, development, and reproduction of organisms relies on cell division, or the process by which a single cell divides into two usually identical daughter cells. This requires first making a duplicate copy of every gene in the genome in a process called DNA replication. The copies are made by specialized enzymes known as DNA polymerases, which "read" one strand of the double-helical DNA, known as the template strand, and synthesize a new complementary strand. Because the DNA double helix is held together by base pairing, the sequence of one strand completely specifies the sequence of its complement; hence only one strand needs to be read by the enzyme to produce a faithful copy. The process of DNA replication is semiconservative; that is, the copy of the genome inherited by each daughter cell contains one original and one newly synthesized strand of DNA.[15]

After DNA replication is complete, the cell must physically separate the two copies of the genome and divide into two distinct membrane-bound cells. In prokaryotes – bacteria and archaea – this usually occurs via a relatively simple process called binary fission, in which each circular genome attaches to the cell membrane and is separated into the daughter cells as the membrane invaginates to split the cytoplasm into two membrane-bound portions. Binary fission is extremely fast compared to the rates of cell division in eukaryotes. Eukaryotic cell division is a more complex process known as the cell cycle; DNA replication occurs during a phase of this cycle known as S phase, whereas the process of segregating chromosomes and splitting the cytoplasm occurs during M phase. In many single-celled eukaryotes such as yeast, reproduction by budding is common, which results in asymmetrical portions of cytoplasm in the two daughter cells.

Molecular inheritance

The duplication and transmission of genetic material from one generation of cells to the next is the basis for molecular inheritance, and the link between the classical and molecular pictures of genes. Organisms inherit the characteristics of their parents because the cells of the offspring contain copies of the genes in their parents' cells. In asexually reproducing organisms, the offspring will be a genetic copy or clone of the parent organism. In sexually reproducing organisms, a specialized form of cell division called meiosis produces cells called gametes or germ cells that are haploid, or contain only one copy of each gene. The gametes produced by females are called eggs or ova, and those produced by males are called sperm. Two gametes fuse to form a fertilized egg, a single cell that once again has a diploid number of genes—each with one copy from the mother and one copy from the father.

During the process of meiotic cell division, an event called genetic recombination or crossing-over can sometimes occur, in which a length of DNA on one chromatid is swapped with a length of DNA on the corresponding sister chromatid. This has no effect if the alleles on the chromatids are the same, but results in reassortment of otherwise linked alleles if they are different. The Mendelian principle of independent assortment asserts that each of a parent's two genes for each trait will sort independently into gametes; which allele an organism inherits for one trait is unrelated to which allele it inherits for another trait. This is in fact only true for genes that do not reside on the same chromosome, or are located very far from one another on the same chromosome. The closer two genes lie on the same chromosome, the more closely they will be associated in gametes and the more often they will appear together; genes that are very close are essentially never separated because it is extremely unlikely that a crossover point will occur between them. This is known as genetic linkage.

Mutation

DNA replication is for the most part extremely accurate, with an error rate per site of around 10−6 to 10−10 in eukaryotes.[15] (Although in prokaryotes and viruses, the rate is much higher.) Rare, spontaneous alterations in the base sequence of a particular gene arise from a number of sources, such as errors in DNA replication and the aftermath of DNA damage. These errors are called mutations. The cell contains many DNA repair mechanisms for preventing mutations and maintaining the integrity of the genome; however, in some cases—such as breaks in both DNA strands of a chromosome—repairing the physical damage to the molecule is a higher priority than producing an exact copy. Due to the degeneracy of the genetic code, some mutations in protein-coding genes are silent, or produce no change in the amino acid sequence of the protein for which they code; for example, the codons UCU and UCC both code for serine, so the U→C mutation has no effect on the protein. Mutations that do have phenotypic effects are most often neutral or deleterious to the organism. Variants may confer benefits to the organism's fitness; it is commonly thought that mutations may produce beneficial variants. The most common mutations include point mutations in which a single codon is replaced, frame shift mutation where a single nucleotide base is inserted or deleted from the DNA strand so that all bases are shifted over, silent mutations where a single nucleotide base is replaced but without causing a change for the amino acid being coded for, and nonsense mutations, where a change in a single nucleotide base causes a codon to be turned into a stop codon hence terminating transcription at this point.

Mutations propagated to the next generation lead to variations within a species' population. Variants of a single gene are known as alleles, and differences in alleles may give rise to differences in traits. Although it is rare for the variants in a single gene to have clearly distinguishable phenotypic effects, certain well-defined traits are in fact controlled by single genetic loci. A gene's most common allele is called the wild type allele, and rare alleles are called mutants. However, this does not imply that the wild-type allele is the ancestor from which the mutants are descended. For the most part, these mutations are recessive and are phased out quickly. However, on occasion these mutations appear as dominant to other alleles, becoming predominant and increasing in the rate they are seen in a population.

Genome

Chromosomal organization

The total complement of genes in an organism or cell is known as its genome. In prokaryotes, the vast majority of genes are located on a single chromosome of circular DNA, while eukaryotes usually possess multiple individual linear DNA helices packed into dense DNA-protein complexes called chromosomes. Genes that appear together on one chromosome of one species may appear on separate chromosomes in another species. Many species carry more than one copy of their genome within each of their somatic cells. Cells or organisms with only one copy of each chromosome are called haploid; those with two copies are called diploid; and those with more than two copies are called polyploid. The copies of genes on the chromosomes are not necessarily identical. In sexually reproducing organisms, one copy is normally inherited from each parent.

Number of genes

The protein-encoding component of the human genome, categorized by function of each gene product, given both as number of genes and as percentage of all genes.[16]

Early estimates of the number of human genes that used expressed sequence tag data put it at 50 000–100 000.[17] Following the sequencing of the human genome and other genomes, it has been found that rather few genes (~20 000 in human, mouse and fly, ~13 000 in roundworm, >46,000 in rice[18]) encode all the proteins in an organism.[19] These protein-coding sequences make up 1–2% of the human genome.[20] A large part of the genome is transcribed however, to introns, retrotransposons and seemingly a large array of noncoding RNAs.[19][20] Total number of proteins (the Earth's proteome) is estimated to be 5 million sequences.[21]

Genetic and genomic nomenclature

Gene nomenclature has been established by the HUGO Gene Nomenclature Committee (HGNC) for each known human gene in the form of an approved gene name and symbol (short-form abbreviation). All approved symbols are stored in the HGNC Database. Each symbol is unique and each gene is only given one approved gene symbol. This also facilitates electronic data retrieval from publications. In preference each symbol maintains parallel construction in different members of a gene family and can be used in other species, especially the mouse.

Essential genes

Essential genes are those genes of an organism that are thought to be critical for its survival. Surprisingly few genes have been shown to be absolutely essential for the survival of bacteria, e.g. only about 10% of the ~4,200 genes of Escherichia coli.

Evolutionary concept of a gene

George C. Williams first explicitly advocated the gene-centric view of evolution in his 1966 book Adaptation and Natural Selection. He proposed an evolutionary concept of gene to be used when we are talking about natural selection favoring some genes. The definition is: "that which segregates and recombines with appreciable frequency." According to this definition, even an asexual genome could be considered a gene, insofar that it have an appreciable permanency through many generations.

The difference is: the molecular gene transcribes as a unit, and the evolutionary gene inherits as a unit.

Richard Dawkins' books The Selfish Gene (1976) and The Extended Phenotype (1982) defended the idea that the gene is the only replicator in living systems. This means that only genes transmit their structure largely intact and are potentially immortal in the form of copies. So, genes should be the unit of selection. In River Out of Eden, Dawkins further refined the idea of gene-centric selection by describing life as a river of compatible genes flowing through geological time. Scoop up a bucket of genes from the river of genes, and we have an organism serving as temporary bodies or survival machines. A river of genes may fork into two branches representing two non-interbreeding species as a result of geographical separation.

Gene targeting and implications

Gene targeting is commonly referred to techniques for altering or disrupting mouse genes and provides the mouse models for studying the roles of individual genes in embryonic development, human disorders, aging and diseases. The mouse models, where one or more of its genes are deactivated or made inoperable, are called knockout mice. Since the first reports in which homologous recombination among homologous chromosomes in embryonic stem cells was used to generate gene-targeted mice,[22] gene targeting has proven to be a powerful means of precisely manipulating the mammalian genome, producing at least ten thousand mutant mouse strains and it is now possible to introduce mutations that can be activated at specific time points, or in specific cells or organs, both during development and in the adult animal.[23][24]

Gene targeting strategies have been expanded to all kinds of modifications, including point mutations, isoform deletions, mutant allele correction, large pieces of chromosomal DNA insertion and deletion, tissue specific disruption combined with spatial and temporal regulation and so on. It is predicted that the ability to generate mouse models with predictable phenotypes will have a major impact on studies of all phases of development, immunology, neurobiology, oncology, physiology, metabolism, and human diseases. Gene targeting is also in theory applicable to species from which totipotent embryonic stem cells can be established, and therefore may offer a potential to the improvement of domestic animals and plants.[24][25]

Changing concept

The concept of the gene has changed considerably (see history section). From the original definition of a "unit of inheritance", the term evolved to mean a DNA-based unit that can exert its effects on the organism through RNA or protein products. It was also previously believed that one gene makes one protein; this concept was overthrown by the discovery of alternative splicing and trans-splicing.[5]

The definition of a gene is still changing. The first cases of RNA-based inheritance have been discovered in mammals.[26] Evidence is also accumulating that the control regions of a gene do not necessarily have to be close to the coding sequence on the linear molecule or even on the same chromosome. Spilianakis and colleagues discovered that the promoter region of the interferon-gamma gene on chromosome 10 and the regulatory regions of the T(H)2 cytokine locus on chromosome 11 come into close proximity in the nucleus possibly to be jointly regulated.[27] Even the coding sequence of a gene itself doesn't have to be all on the same chromosome: Marande and Burger showed that, in the mitochondria of the protist Diplonema papillatum, "genes are systematically fragmented into small pieces that are encoded on separate chromosomes, transcribed individually, and then concatenated into contiguous messenger RNA molecules".[28]

The concept that genes are clearly delimited is also being eroded. There is evidence for fused proteins stemming from two adjacent genes that can produce two separate protein products. While it is not clear whether these fusion proteins are functional, the phenomenon is more frequent than previously thought.[29] Even more ground-breaking than the discovery of fused genes is the observation that some proteins can be composed of exons from far away regions and even different chromosomes.[3][30] This new data has led to an updated, and probably tentative, definition of a gene as "a union of genomic sequences encoding a coherent set of potentially overlapping functional products".[5] This new definition categorizes genes by functional products, whether they be proteins or RNA, rather than specific DNA loci; all regulatory elements of DNA are therefore classified as gene-associated regions.[5]

See also

2

Notes and references

  1. ^ "gene". Oxford English Dictionary (Online ed.). Oxford University Press. (Subscription or participating institution membership required.)
  2. ^ a b Pearson H (May 2006). "Genetics: what is a gene?". Nature. 441 (7092): 398–401. Bibcode:2006Natur.441..398P. doi:10.1038/441398a. PMID 16724031.
  3. ^ a b c Pennisi E (June 2007). "Genomics. DNA study forces rethink of what it means to be a gene". Science. 316 (5831): 1556–1557. doi:10.1126/science.316.5831.1556. PMID 17569836.
  4. ^ "The Human Genome Project Timeline". Retrieved 13 September 2006.
  5. ^ a b c d e "What is a gene, post-ENCODE? History and updated definition". Genome Research. 17 (6): 669–681. June 2007. doi:10.1101/gr.6339607. PMID 17567988. {{cite journal}}: |first2= missing |last2= (help); |first3= missing |last3= (help); More than one of author-name-list parameters specified (help)
  6. ^ Steinman RM, Moberg CL (February 1994). "A triple tribute to the experiment that transformed biology". The Journal of Experimental Medicine. 179 (2): 379–84. doi:10.1084/jem.179.2.379. PMC 2191359. PMID 8294854.
  7. ^ Min Jou W, Haegeman G, Ysebaert M, Fiers W (May 1972). "Nucleotide sequence of the gene coding for the bacteriophage MS2 coat protein". Nature. 237 (5350): 82–8. Bibcode:1972Natur.237...82J. doi:10.1038/237082a0. PMID 4555447.
  8. ^ Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B (July 2008). "Mapping and quantifying mammalian transcriptomes by RNA-Seq". Nature Methods. 5 (7): 621–8. doi:10.1038/nmeth.1226. PMID 18516045.
  9. ^ Woodson SA (May 1998). "Ironing out the kinks: splicing and translation in bacteria". Genes & Development. 12 (9): 1243–7. doi:10.1101/gad.12.9.1243. PMID 9573040.
  10. ^ Braig M, Schmitt CA (March 2006). "Oncogene-induced senescence: putting the brakes on tumor development". Cancer Research. 66 (6): 2881–4. doi:10.1158/0008-5472.CAN-05-4006. PMID 16540631.
  11. ^ International Human Genome SC (October 2004). "Finishing the euchromatic sequence of the human genome". Nature. 431 (7011): 931–45. Bibcode:2004Natur.431..931H. doi:10.1038/nature03001. PMID 15496913. {{cite journal}}: Vancouver style error: initials in name 1 (help)
  12. ^ Alberts, Bruce; Johnson, Alexander; Lewis, Julian; Raff, Martin; Roberts, Keith; Walter, Peter (2002). Molecular Biology of the Cell (Fourth ed.). New York: Garland Science. ISBN 978-0-8153-3218-3. {{cite book}}: Unknown parameter |name-list-format= ignored (|name-list-style= suggested) (help)
  13. ^ Morris, KV; Mattick, JS (June 2014). "The rise of regulatory RNA". Nature reviews. Genetics. 15 (6): 423–37. PMID 24776770.
  14. ^ RNA Viruses at the U.S. National Library of Medicine Medical Subject Headings (MeSH)
  15. ^ a b Watson JD, Baker TA, Bell SP, Gann A, Levine M, Losick R (2004). Molecular Biology of the Gene (5th ed.). Peason Benjamin Cummings (Cold Spring Harbor Laboratory Press). ISBN 0-8053-4635-X.{{cite book}}: CS1 maint: multiple names: authors list (link)
  16. ^ PANTHER Pie Chart at the PANTHER Classification System homepage. Retrieved 25 May 2011
  17. ^ Schuler GD, Boguski MS, Stewart EA, Stein LD, Gyapay G, Rice K, White RE, Rodriguez-Tomé P, Aggarwal A, Bajorek E, Bentolila S, Birren BB, Butler A, Castle AB, Chiannilkulchai N, Chu A, Clee C, Cowles S, Day PJ, Dibling T, Drouot N, Dunham I, Duprat S, East C, Edwards C, Fan JB, Fang N, Fizames C, Garrett C, Green L, Hadley D, Harris M, Harrison P, Brady S, Hicks A, Holloway E, Hui L, Hussain S, Louis-Dit-Sully C, Ma J, MacGilvery A, Mader C, Maratukulam A, Matise TC, McKusick KB, Morissette J, Mungall A, Muselet D, Nusbaum HC, Page DC, Peck A, Perkins S, Piercy M, Qin F, Quackenbush J, Ranby S, Reif T, Rozen S, Sanders C, She X, Silva J, Slonim DK, Soderlund C, Sun WL, Tabar P, Thangarajah T, Vega-Czarny N, Vollrath D, Voyticky S, Wilmer T, Wu X, Adams MD, Auffray C, Walter NA, Brandon R, Dehejia A, Goodfellow PN, Houlgatte R, Hudson JR, Ide SE, Iorio KR, Lee WY, Seki N, Nagase T, Ishikawa K, Nomura N, Phillips C, Polymeropoulos MH, Sandusky M, Schmitt K, Berry R, Swanson K, Torres R, Venter JC, Sikela JM, Beckmann JS, Weissenbach J, Myers RM, Cox DR, James MR, Bentley D, Deloukas P, Lander ES, Hudson TJ (October 1996). "A gene map of the human genome". Science. 274 (5287): 540–6. Bibcode:1996Sci...274..540S. doi:10.1126/science.274.5287.540. PMID 8849440.
  18. ^ Yu J, Hu S, Wang J, Wong GK, Li S, Liu B, Deng Y, Dai L, Zhou Y, Zhang X, Cao M, Liu J, Sun J, Tang J, Chen Y, Huang X, Lin W, Ye C, Tong W, Cong L, Geng J, Han Y, Li L, Li W, Hu G, Huang X, Li W, Li J, Liu Z, Li L, Liu J, Qi Q, Liu J, Li L, Li T, Wang X, Lu H, Wu T, Zhu M, Ni P, Han H, Dong W, Ren X, Feng X, Cui P, Li X, Wang H, Xu X, Zhai W, Xu Z, Zhang J, He S, Zhang J, Xu J, Zhang K, Zheng X, Dong J, Zeng W, Tao L, Ye J, Tan J, Ren X, Chen X, He J, Liu D, Tian W, Tian C, Xia H, Bao Q, Li G, Gao H, Cao T, Wang J, Zhao W, Li P, Chen W, Wang X, Zhang Y, Hu J, Wang J, Liu S, Yang J, Zhang G, Xiong Y, Li Z, Mao L, Zhou C, Zhu Z, Chen R, Hao B, Zheng W, Chen S, Guo W, Li G, Liu S, Tao M, Wang J, Zhu L, Yuan L, Yang H (April 2002). "A draft sequence of the rice genome (Oryza sativa L. ssp. indica)". Science. 296 (5565): 79–92. Bibcode:2002Sci...296...79Y. doi:10.1126/science.1068037. PMID 11935017.
  19. ^ a b Carninci P, Hayashizaki Y (April 2007). "Noncoding RNA transcription beyond annotated genes". Current Opinion in Genetics & Development. 17 (2): 139–44. doi:10.1016/j.gde.2007.02.008. PMID 17317145.
  20. ^ a b Claverie JM (September 2005). "Fewer genes, more noncoding RNA". Science. 309 (5740): 1529–30. Bibcode:2005Sci...309.1529C. doi:10.1126/science.1116800. PMID 16141064.
  21. ^ "Towards completion of the Earth's proteome". EMBO Reports. 8 (12): 1135–1141. December 2007. doi:10.1038/sj.embor.7401117. PMC 2267224. PMID 18059312. {{cite journal}}: |first2= missing |last2= (help); |first3= missing |last3= (help); More than one of author-name-list parameters specified (help)
  22. ^ Thomas KR, Capecchi MR (November 1987). "Site-directed mutagenesis by gene targeting in mouse embryo-derived stem cells". Cell. 51 (3): 503–12. doi:10.1016/0092-8674(87)90646-5. PMID 2822260.
  23. ^ The 2007 Nobel Prize in Physiology or Medicine – Press Release
  24. ^ a b Deng C (2007). "In celebration of Dr. Mario R. Capecchi's Nobel Prize". International Journal of Biological Sciences. 3 (7): 417–419. doi:10.7150/ijbs.3.417. PMID 17998949.
  25. ^ Mario R. Capecchi
  26. ^ Rassoulzadegan M, Grandjean V, Gounon P, Vincent S, Gillot I, Cuzin F (May 2006). "RNA-mediated non-mendelian inheritance of an epigenetic change in the mouse". Nature. 441 (7092): 469–74. Bibcode:2006Natur.441..469R. doi:10.1038/nature04674. PMID 16724059.
  27. ^ Spilianakis CG, Lalioti MD, Town T, Lee GR, Flavell RA (June 2005). "Interchromosomal associations between alternatively expressed loci". Nature. 435 (7042): 637–45. Bibcode:2005Natur.435..637S. doi:10.1038/nature03574. PMID 15880101.
  28. ^ Marande W, Burger G (October 2007). "Mitochondrial DNA as a genomic jigsaw puzzle". Science. 318 (5849). AAAS: 415. Bibcode:2007Sci...318..415M. doi:10.1126/science.1148033. PMID 17947575.
  29. ^ Parra G, Reymond A, Dabbouseh N, Dermitzakis ET, Castelo R, Thomson TM, Antonarakis SE, Guigó R (January 2006). "Tandem chimerism as a means to increase protein complexity in the human genome". Genome Research. 16 (1): 37–44. doi:10.1101/gr.4145906. PMC 1356127. PMID 16344564.
  30. ^ Kapranov P, Drenkow J, Cheng J, Long J, Helt G, Dike S, Gingeras TR (July 2005). "Examples of the complex architecture of the human transcriptome revealed by RACE and high-density tiling arrays". Genome Research. 15 (7): 987–97. doi:10.1101/gr.3455305. PMC 1172043. PMID 15998911.

Bibliography

Listen to this article
(2 parts, 17 minutes)
Spoken Wikipedia icon
These audio files were created from a revision of this article dated
Error: no date provided
, and do not reflect subsequent edits.