Gene structure

Last updated

Gene structure is the organisation of specialised sequence elements within a gene. Genes contain most of the information necessary for living cells to survive and reproduce. [1] [2] In most organisms, genes are made of DNA, where the particular DNA sequence determines the function of the gene. A gene is transcribed (copied) from DNA into RNA, which can either be non-coding (ncRNA) with a direct function, or an intermediate messenger (mRNA) that is then translated into protein. Each of these steps is controlled by specific sequence elements, or regions, within the gene. Every gene, therefore, requires multiple sequence elements to be functional. [2] This includes the sequence that actually encodes the functional protein or ncRNA, as well as multiple regulatory sequence regions. These regions may be as short as a few base pairs, up to many thousands of base pairs long.

Contents

Much of gene structure is broadly similar between eukaryotes and prokaryotes. These common elements largely result from the shared ancestry of cellular life in organisms over 2 billion years ago. [3] Key differences in gene structure between eukaryotes and prokaryotes reflect their divergent transcription and translation machinery. [4] [5] Understanding gene structure is the foundation of understanding gene annotation, expression, and function. [6]

Common features

The structures of both eukaryotic and prokaryotic genes involve several nested sequence elements. Each element has a specific function in the multi-step process of gene expression. The sequences and lengths of these elements vary, but the same general functions are present in most genes. [2] Although DNA is a double-stranded molecule, typically only one of the strands encodes information that the RNA polymerase reads to produce protein-coding mRNA or non-coding RNA. This 'sense' or 'coding' strand, runs in the 5' to 3' direction where the numbers refer to the carbon atoms of the backbone's ribose sugar. The open reading frame (ORF) of a gene is therefore usually represented as an arrow indicating the direction in which the sense strand is read. [7]

Regulatory sequences are located at the extremities of genes. These sequence regions can either be next to the transcribed region (the promoter) or separated by many kilobases (enhancers and silencers). [8] The promoter is located at the 5' end of the gene and is composed of a core promoter sequence and a proximal promoter sequence. The core promoter marks the start site for transcription by binding RNA polymerase and other proteins necessary for copying DNA to RNA. The proximal promoter region binds transcription factors that modify the affinity of the core promoter for RNA polymerase. [9] [10] Genes may be regulated by multiple enhancer and silencer sequences that further modify the activity of promoters by binding activator or repressor proteins. [11] [12] Enhancers and silencers may be distantly located from the gene, many thousands of base pairs away. The binding of different transcription factors, therefore, regulates the rate of transcription initiation at different times and in different cells. [13]

Regulatory elements can overlap one another, with a section of DNA able to interact with many competing activators and repressors as well as RNA polymerase. For example, some repressor proteins can bind to the core promoter to prevent polymerase binding. [14] For genes with multiple regulatory sequences, the rate of transcription is the product of all of the elements combined. [15] Binding of activators and repressors to multiple regulatory sequences has a cooperative effect on transcription initiation. [16]

Although all organisms use both transcriptional activators and repressors, eukaryotic genes are said to be 'default off', whereas prokaryotic genes are 'default on'. [5] The core promoter of eukaryotic genes typically requires additional activation by promoter elements for expression to occur. The core promoter of prokaryotic genes, conversely, is sufficient for strong expression and is regulated by repressors. [5]

An additional layer of regulation occurs for protein coding genes after the mRNA has been processed to prepare it for translation to protein. Only the region between the start and stop codons encodes the final protein product. The flanking untranslated regions (UTRs) contain further regulatory sequences. [18] The 3' UTR contains a terminator sequence, which marks the endpoint for transcription and releases the RNA polymerase. [19] The 5’ UTR binds the ribosome, which translates the protein-coding region into a string of amino acids that fold to form the final protein product. In the case of genes for non-coding RNAs, the RNA is not translated but instead folds to be directly functional. [20] [21]

Eukaryotes

The structure of eukaryotic genes includes features not found in prokaryotes. Most of these relate to post-transcriptional modification of pre-mRNAs to produce mature mRNA ready for translation into protein. Eukaryotic genes typically have more regulatory elements to control gene expression compared to prokaryotes. [5] This is particularly true in multicellular eukaryotes, humans for example, where gene expression varies widely among different tissues. [11]

A key feature of the structure of eukaryotic genes is that their transcripts are typically subdivided into exon and intron regions. Exon regions are retained in the final mature mRNA molecule, while intron regions are spliced out (excised) during post-transcriptional processing. [22] Indeed, the intron regions of a gene can be considerably longer than the exon regions. Once spliced together, the exons form a single continuous protein-coding regions, and the splice boundaries are not detectable. Eukaryotic post-transcriptional processing also adds a 5' cap to the start of the mRNA and a poly-adenosine tail to the end of the mRNA. These additions stabilise the mRNA and direct its transport from the nucleus to the cytoplasm, although neither of these features are directly encoded in the structure of a gene. [18]

Prokaryotes

The overall organisation of prokaryotic genes is markedly different from that of the eukaryotes. The most obvious difference is that prokaryotic ORFs are often grouped into a polycistronic operon under the control of a shared set of regulatory sequences. These ORFs are all transcribed onto the same mRNA and so are co-regulated and often serve related functions. [23] [24] Each ORF typically has its own ribosome binding site (RBS) so that ribosomes simultaneously translate ORFs on the same mRNA. Some operons also display translational coupling, where the translation rates of multiple ORFs within an operon are linked. [25] This can occur when the ribosome remains attached at the end of an ORF and simply translocates along to the next without the need for a new RBS. [26] Translational coupling is also observed when translation of an ORF affects the accessibility of the next RBS through changes in RNA secondary structure. [27] Having multiple ORFs on a single mRNA is only possible in prokaryotes because their transcription and translation take place at the same time and in the same subcellular location. [23] [28]

The operator sequence next to the promoter is the main regulatory element in prokaryotes. Repressor proteins bound to the operator sequence physically obstructs the RNA polymerase enzyme, preventing transcription. [29] [30] Riboswitches are another important regulatory sequence commonly present in prokaryotic UTRs. These sequences switch between alternative secondary structures in the RNA depending on the concentration of key metabolites. The secondary structures then either block or reveal important sequence regions such as RBSs. Introns are extremely rare in prokaryotes and therefore do not play a significant role in prokaryotic gene regulation. [31]

Related Research Articles

<span class="mw-page-title-main">Messenger RNA</span> RNA that is read by the ribosome to produce a protein

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

<span class="mw-page-title-main">Promoter (genetics)</span> Region of DNA encouraging transcription

In genetics, a promoter is a sequence of DNA to which proteins bind to initiate transcription of a single RNA transcript from the DNA downstream of the promoter. The RNA transcript may encode a protein (mRNA), or can have a function in and of itself, such as tRNA or rRNA. Promoters are located near the transcription start sites of genes, upstream on the DNA . Promoters can be about 100–1000 base pairs long, the sequence of which is highly dependent on the gene and product of transcription, type or class of RNA polymerase recruited to the site, and species of organism.

Non-coding DNA (ncDNA) sequences are components of an organism's DNA that do not encode protein sequences. Some non-coding DNA is transcribed into functional non-coding RNA molecules. Other functional regions of the non-coding DNA fraction include regulatory sequences that control gene expression; scaffold attachment regions; origins of DNA replication; centromeres; and telomeres. Some non-coding regions appear to be mostly nonfunctional such as introns, pseudogenes, intergenic DNA, and fragments of transposons and viruses.

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

Gene expression is the process by which information from a gene is used in the synthesis of a functional gene product that enables it to produce end products, protein or non-coding RNA, and ultimately affect a phenotype, as the final effect. These products are often proteins, but in non-protein-coding genes such as transfer RNA (tRNA) and small nuclear RNA (snRNA), the product is a functional non-coding RNA. Gene expression is summarized in the central dogma of molecular biology first formulated by Francis Crick in 1958, further developed in his 1970 article, and expanded by the subsequent discoveries of reverse transcription and RNA replication.

In genetics, an operon is a functioning unit of DNA containing a cluster of genes under the control of a single promoter. The genes are transcribed together into an mRNA strand and either translated together in the cytoplasm, or undergo splicing to create monocistronic mRNAs that are translated separately, i.e. several strands of mRNA that each encode a single gene product. The result of this is that the genes contained in the operon are either expressed together or not at all. Several genes must be co-transcribed to define an operon.

In molecular biology and genetics, transcriptional regulation is the means by which a cell regulates the conversion of DNA to RNA (transcription), thereby orchestrating gene activity. A single gene can be regulated in a range of ways, from altering the number of copies of RNA that are transcribed, to the temporal control of when the gene is transcribed. This control allows the cell or organism to respond to a variety of intra- and extracellular signals and thus mount a response. Some examples of this include producing the mRNA that encode enzymes to adapt to a change in a food source, producing the gene products involved in cell cycle specific activities, and producing the gene products responsible for cellular differentiation in multicellular eukaryotes, as studied in evolutionary developmental biology.

<span class="mw-page-title-main">Regulation of gene expression</span> Modifying mechanisms used by cells to increase or decrease the production of specific gene products

Regulation of gene expression, or gene regulation, includes a wide range of mechanisms that are used by cells to increase or decrease the production of specific gene products. Sophisticated programs of gene expression are widely observed in biology, for example to trigger developmental pathways, respond to environmental stimuli, or adapt to new food sources. Virtually any step of gene expression can be modulated, from transcriptional initiation, to RNA processing, and to the post-translational modification of a protein. Often, one gene regulator controls another, and so on, in a gene regulatory network.

A transcriptional activator is a protein that increases transcription of a gene or set of genes. Activators are considered to have positive control over gene expression, as they function to promote gene transcription and, in some cases, are required for the transcription of genes to occur. Most activators are DNA-binding proteins that bind to enhancers or promoter-proximal elements. The DNA site bound by the activator is referred to as an "activator-binding site". The part of the activator that makes protein–protein interactions with the general transcription machinery is referred to as an "activating region" or "activation domain".

<span class="mw-page-title-main">Primary transcript</span> RNA produced by transcription

A primary transcript is the single-stranded ribonucleic acid (RNA) product synthesized by transcription of DNA, and processed to yield various mature RNA products such as mRNAs, tRNAs, and rRNAs. The primary transcripts designated to be mRNAs are modified in preparation for translation. For example, a precursor mRNA (pre-mRNA) is a type of primary transcript that becomes a messenger RNA (mRNA) after processing.

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

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

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

<span class="mw-page-title-main">Repressor</span> Sort of RNA-binding protein in molecular genetics

In molecular genetics, a repressor is a DNA- or RNA-binding protein that inhibits the expression of one or more genes by binding to the operator or associated silencers. A DNA-binding repressor blocks the attachment of RNA polymerase to the promoter, thus preventing transcription of the genes into messenger RNA. An RNA-binding repressor binds to the mRNA and prevents translation of the mRNA into protein. This blocking or reducing of expression is called repression.

<span class="mw-page-title-main">Silencer (genetics)</span> Type of DNA sequence

In genetics, a silencer is a DNA sequence capable of binding transcription regulation factors, called repressors. DNA contains genes and provides the template to produce messenger RNA (mRNA). That mRNA is then translated into proteins. When a repressor protein binds to the silencer region of DNA, RNA polymerase is prevented from transcribing the DNA sequence into RNA. With transcription blocked, the translation of RNA into proteins is impossible. Thus, silencers prevent genes from being expressed as proteins.

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

A regulator gene, regulator, or regulatory gene is a gene involved in controlling the expression of one or more other genes. Regulatory sequences, which encode regulatory genes, are often at the five prime end (5') to the start site of transcription of the gene they regulate. In addition, these sequences can also be found at the three prime end (3') to the transcription start site. In both cases, whether the regulatory sequence occurs before (5') or after (3') the gene it regulates, the sequence is often many kilobases away from the transcription start site. A regulator gene may encode a protein, or it may work at the level of RNA, as in the case of genes encoding microRNAs. An example of a regulator gene is a gene that codes for a repressor protein that inhibits the activity of an operator.

In genetics, attenuation is a regulatory mechanism for some bacterial operons that results in premature termination of transcription. The canonical example of attenuation used in many introductory genetics textbooks, is ribosome-mediated attenuation of the trp operon. Ribosome-mediated attenuation of the trp operon relies on the fact that, in bacteria, transcription and translation proceed simultaneously. Attenuation involves a provisional stop signal (attenuator), located in the DNA segment that corresponds to the leader sequence of mRNA. During attenuation, the ribosome becomes stalled (delayed) in the attenuator region in the mRNA leader. Depending on the metabolic conditions, the attenuator either stops transcription at that point or allows read-through to the structural gene part of the mRNA and synthesis of the appropriate protein.

<span class="mw-page-title-main">Gene</span> Sequence of DNA or RNA that codes for an RNA or protein product

In biology, the word gene can have several different meanings. The Mendelian gene is a basic unit of heredity and the molecular gene is a sequence of nucleotides in DNA that is transcribed to produce a functional RNA. There are two types of molecular genes: protein-coding genes and noncoding genes.

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

Bacterial transcription is the process in which a segment of bacterial DNA is copied into a newly synthesized strand of messenger RNA (mRNA) with use of the enzyme RNA polymerase.

<span class="mw-page-title-main">Eukaryotic transcription</span> Transcription is heterocatalytic function of DNA

Eukaryotic transcription is the elaborate process that eukaryotic cells use to copy genetic information stored in DNA into units of transportable complementary RNA replica. Gene transcription occurs in both eukaryotic and prokaryotic cells. Unlike prokaryotic RNA polymerase that initiates the transcription of all different types of RNA, RNA polymerase in eukaryotes comes in three variations, each translating a different type of gene. A eukaryotic cell has a nucleus that separates the processes of transcription and translation. Eukaryotic transcription occurs within the nucleus where DNA is packaged into nucleosomes and higher order chromatin structures. The complexity of the eukaryotic genome necessitates a great variety and complexity of gene expression control.

The gal operon is a prokaryotic operon, which encodes enzymes necessary for galactose metabolism. Repression of gene expression for this operon works via binding of repressor molecules to two operators. These repressors dimerize, creating a loop in the DNA. The loop as well as hindrance from the external operator prevent RNA polymerase from binding to the promoter, and thus prevent transcription. Additionally, since the metabolism of galactose in the cell is involved in both anabolic and catabolic pathways, a novel regulatory system using two promoters for differential repression has been identified and characterized within the context of the gal operon.

<span class="mw-page-title-main">5′ flanking region</span>

The 5′ flanking region is a region of DNA that is adjacent to the 5′ end of the gene. The 5′ flanking region contains the promoter, and may contain enhancers or other protein binding sites. It is the region of DNA that is not transcribed into RNA. Not to be confused with the 5′ untranslated region, this region is not transcribed into RNA or translated into a functional protein. These regions primarily function in the regulation of gene transcription. 5′ flanking regions are categorized between prokaryotes and eukaryotes.

References

Open Access logo PLoS transparent.svg This article was adapted from the following source under a CC BY 4.0 license (2017) (reviewer reports): Thomas Shafee; Rohan Lowe (17 January 2017). "Eukaryotic and prokaryotic gene structure". WikiJournal of Medicine. 4 (1). doi:10.15347/WJM/2017.002. ISSN   2002-4436. Wikidata   Q28867140.

  1. Alberts, Bruce; Johnson, Alexander; Lewis, Julian; Raff, Martin; Roberts, Keith; Walter, Peter (2002). "How Genetic Switches Work". Molecular Biology of the Cell (4 ed.).
  2. 1 2 3 Polyak, Kornelia; Meyerson, Matthew (2003). "Overview: Gene Structure". Cancer Medicine (6 ed.). BC Decker.
  3. Werner, Finn; Grohmann, Dina (2011). "Evolution of multisubunit RNA polymerases in the three domains of life". Nature Reviews Microbiology. 9 (2): 85–98. doi:10.1038/nrmicro2507. ISSN   1740-1526. PMID   21233849. S2CID   30004345.
  4. Kozak, Marilyn (1999). "Initiation of translation in prokaryotes and eukaryotes". Gene. 234 (2): 187–208. doi:10.1016/S0378-1119(99)00210-3. ISSN   0378-1119. PMID   10395892.
  5. 1 2 3 4 Struhl, Kevin (1999). "Fundamentally Different Logic of Gene Regulation in Eukaryotes and Prokaryotes". Cell. 98 (1): 1–4. doi: 10.1016/S0092-8674(00)80599-1 . ISSN   0092-8674. PMID   10412974. S2CID   12411218.
  6. Alberts B, Johnson A, Lewis J, Raff M, Roberts K, Walter P (2002). Molecular Biology of the Cell (Fourth ed.). New York: Garland Science. ISBN   978-0-8153-3218-3.
  7. Lu, G. (2004). "Vector NTI, a balanced all-in-one sequence analysis suite". Briefings in Bioinformatics. 5 (4): 378–88. doi: 10.1093/bib/5.4.378 . ISSN   1467-5463. PMID   15606974.
  8. Wiper-Bergeron, Nadine; Skerjanc, Ilona S. (2009). Transcription and the Control of Gene Expression. Humana Press. pp. 33–49. doi:10.1007/978-1-59745-440-7_2. ISBN   978-1-59745-440-7.
  9. Thomas, Mary C.; Chiang, Cheng-Ming (2008). "The General Transcription Machinery and General Cofactors". Critical Reviews in Biochemistry and Molecular Biology. 41 (3): 105–78. CiteSeerX   10.1.1.376.5724 . doi:10.1080/10409230600648736. ISSN   1040-9238. PMID   16858867. S2CID   13073440.
  10. Juven-Gershon, Tamar; Hsu, Jer-Yuan; Theisen, Joshua WM; Kadonaga, James T (2008). "The RNA polymerase II core promoter – the gateway to transcription". Current Opinion in Cell Biology. 20 (3): 253–59. doi:10.1016/j.ceb.2008.03.003. ISSN   0955-0674. PMC   2586601 . PMID   18436437.
  11. 1 2 Maston, Glenn A.; Evans, Sara K.; Green, Michael R. (2006). "Transcriptional Regulatory Elements in the Human Genome". Annual Review of Genomics and Human Genetics. 7 (1): 29–59. doi:10.1146/annurev.genom.7.080505.115623. ISSN   1527-8204. PMID   16719718. S2CID   12346247.
  12. Pennacchio, L. A.; Bickmore, W.; Dean, A.; Nobrega, M. A.; Bejerano, G. (2013). "Enhancers: Five essential questions". Nature Reviews Genetics. 14 (4): 288–95. doi:10.1038/nrg3458. PMC   4445073 . PMID   23503198.
  13. Maston, G. A.; Evans, S. K.; Green, M. R. (2006). "Transcriptional Regulatory Elements in the Human Genome". Annual Review of Genomics and Human Genetics. 7: 29–59. doi:10.1146/annurev.genom.7.080505.115623. PMID   16719718. S2CID   12346247.
  14. Ogbourne, Steven; Antalis, Toni M. (1998). "Transcriptional control and the role of silencers in transcriptional regulation in eukaryotes". Biochemical Journal. 331 (1): 1–14. doi:10.1042/bj3310001. ISSN   0264-6021. PMC   1219314 . PMID   9512455.
  15. Buchler, N. E.; Gerland, U.; Hwa, T. (2003). "On schemes of combinatorial transcription logic". Proceedings of the National Academy of Sciences. 100 (9): 5136–41. Bibcode:2003PNAS..100.5136B. doi: 10.1073/pnas.0930314100 . ISSN   0027-8424. PMC   404558 . PMID   12702751.
  16. Kazemian, M.; Pham, H.; Wolfe, S. A.; Brodsky, M. H.; Sinha, S. (11 July 2013). "Widespread evidence of cooperative DNA binding by transcription factors in Drosophila development". Nucleic Acids Research. 41 (17): 8237–52. doi:10.1093/nar/gkt598. PMC   3783179 . PMID   23847101.
  17. 1 2 Shafee, Thomas; Lowe, Rohan (2017). "Eukaryotic and prokaryotic gene structure". WikiJournal of Medicine. 4 (1). doi:10.15347/wjm/2017.002. ISSN   2002-4436.
  18. 1 2 Guhaniyogi, Jayita; Brewer, Gary (2001). "Regulation of mRNA stability in mammalian cells". Gene. 265 (1–2): 11–23. doi:10.1016/S0378-1119(01)00350-X. ISSN   0378-1119. PMID   11255003.
  19. Kuehner, Jason N.; Pearson, Erika L.; Moore, Claire (2011). "Unravelling the means to an end: RNA polymerase II transcription termination". Nature Reviews Molecular Cell Biology. 12 (5): 283–94. doi:10.1038/nrm3098. ISSN   1471-0072. PMC   6995273 . PMID   21487437.
  20. Mattick, J. S. (2006). "Non-coding RNA". Human Molecular Genetics. 15 (90001): R17–R29. doi: 10.1093/hmg/ddl046 . ISSN   0964-6906. PMID   16651366.
  21. Palazzo, Alexander F.; Lee, Eliza S. (2015). "Non-coding RNA: what is functional and what is junk?". Frontiers in Genetics. 6: 2. doi: 10.3389/fgene.2015.00002 . ISSN   1664-8021. PMC   4306305 . PMID   25674102.
  22. Matera, A. Gregory; Wang, Zefeng (2014). "A day in the life of the spliceosome". Nature Reviews Molecular Cell Biology. 15 (2): 108–21. doi:10.1038/nrm3742. ISSN   1471-0072. PMC   4060434 . PMID   24452469.
  23. 1 2 Salgado, H.; Moreno-Hagelsieb, G.; Smith, T.; Collado-Vides, J. (2000). "Operons in Escherichia coli: Genomic analyses and predictions". Proceedings of the National Academy of Sciences. 97 (12): 6652–57. Bibcode:2000PNAS...97.6652S. doi: 10.1073/pnas.110147297 . PMC   18690 . PMID   10823905.
  24. Jacob, F.; Monod, J. (1961-06-01). "Genetic regulatory mechanisms in the synthesis of proteins". Journal of Molecular Biology. 3 (3): 318–56. doi:10.1016/s0022-2836(61)80072-7. ISSN   0022-2836. PMID   13718526.
  25. Tian, Tian; Salis, Howard M. (2015). "A predictive biophysical model of translational coupling to coordinate and control protein expression in bacterial operons". Nucleic Acids Research. 43 (14): 7137–51. doi:10.1093/nar/gkv635. ISSN   0305-1048. PMC   4538824 . PMID   26117546.
  26. Schümperli, Daniel; McKenney, Keith; Sobieski, Donna A.; Rosenberg, Martin (1982). "Translational coupling at an intercistronic boundary of the Escherichia coli galactose operon". Cell. 30 (3): 865–71. doi:10.1016/0092-8674(82)90291-4. ISSN   0092-8674. PMID   6754091. S2CID   31496240.
  27. Levin-Karp, Ayelet; Barenholz, Uri; Bareia, Tasneem; Dayagi, Michal; Zelcbuch, Lior; Antonovsky, Niv; Noor, Elad; Milo, Ron (2013). "Quantifying Translational Coupling in E. coli Synthetic Operons Using RBS Modulation and Fluorescent Reporters". ACS Synthetic Biology. 2 (6): 327–36. doi:10.1021/sb400002n. ISSN   2161-5063. PMID   23654261. S2CID   63692.
  28. Lewis, Mitchell (June 2005). "The lac repressor". Comptes Rendus Biologies. 328 (6): 521–48. doi:10.1016/j.crvi.2005.04.004. PMID   15950160.
  29. McClure, W R (1985). "Mechanism and Control of Transcription Initiation in Prokaryotes". Annual Review of Biochemistry. 54 (1): 171–204. doi:10.1146/annurev.bi.54.070185.001131. ISSN   0066-4154. PMID   3896120.
  30. Bell, Charles E; Lewis, Mitchell (2001). "The Lac repressor: a second generation of structural and functional studies". Current Opinion in Structural Biology. 11 (1): 19–25. doi:10.1016/S0959-440X(00)00180-9. ISSN   0959-440X. PMID   11179887.
  31. Rodríguez-Trelles, Francisco; Tarrío, Rosa; Ayala, Francisco J. (2006). "Origins and Evolution of Spliceosomal Introns". Annual Review of Genetics. 40 (1): 47–76. doi:10.1146/annurev.genet.40.110405.090625. ISSN   0066-4197. PMID   17094737.