Transcription factor

In the field of molecular biology, a transcription factor (sometimes called a sequence-specific DNA binding factor) is a protein that binds to specific sequences of DNA and thereby controls the transfer (or transcription) of genetic information from DNA to RNA.[1][2] Transcription factors perform this function alone, or with other proteins in a complex, by promoting (as an activator), or blocking (as a repressor) the recruitment of RNA polymerase (the enzyme which activates the transcription of genetic information from DNA to RNA) to specific genes.[3][4][5]

A defining feature of transcription factors is that they contain one or more DNA binding domains (DBDs) which attach to specific sequences of DNA adjacent to the genes that they regulate.[6][7] Additional proteins such as coactivators, chromatin remodelers, histone acetylases, deacetylases, kinases, and methylases, while also playing crucial roles in gene regulation, lack DNA binding domains, and therefore are not classified as transcription factors.[8]

Transcription factor glossary
transcription – copying of DNA by RNA polymerase into messenger RNA
factor – a substance, such as a protein, that contributes to the cause of a specific biochemical reaction or bodily process
transcriptional regulationcontrolling the rate of gene transcription for example by helping or hindering RNA polymerase binding to DNA
upregulation, activation, or promotionincrease the rate of gene transcription
downregulation, repression, or suppressiondecrease the rate of gene transcription
coactivator – a protein which works with transcription factors to increase the rate of gene transcription
corepressor – a protein which works with transcription factors to decrease the rate of gene transcription
edit

Contents

Conservation in different organisms

Transcription factors are essential for the regulation of gene expression and consequently are found in all living organisms. The number of transcription factors found within an organism increases with the genome size and the larger genomes tend to have more transcription factors per gene.[9]

There are approximately 2600 proteins in the human genome that contain DNA-binding domains and most of these are presumed to function as transcription factors.[10] Therefore approximately 10% of genes in the genome code for transcription factors which makes this family the single largest family of human proteins. Furthermore genes are often flanked by several binding sites for distinct transcription factors and efficient expression of each of these genes requires the cooperative action of several different transcription factors (see for example hepatocyte nuclear factors). Hence the combinatorial use of a subset of the approximately 2000 human transcription factors easily accounts for the unique regulation of each gene in the human genome during development.[8]

Function

The transcription factor TATA binding protein (blue) bound to DNA (red). Image by David S. Goodsell based on the crystal structure 1cdw from the Protein Data Bank.

Transcription factors are one of the groups of proteins that read and interpret the genetic "blueprint" in the DNA. They bind DNA and help initiate a program of increased or decreased gene transcription. As such, they are vital for many important cellular processes. Below are some of the important functions and biological roles transcription factors are involved in:

Basal transcription regulation

In eukaryotes, an important class of transcription factors called general transcription factors (GTFs) are necessary for transcription to occur.[11][12] Many of these GTFs don't actually bind DNA but are part of the large transcription preinitiation complex that interacts with RNA polymerase directly. The most common GTFs are TFIIA, TFIIB, TFIID (see also TATA binding protein), TFIIE, TFIIF, and TFIIH.[13]

Development

Many transcription factors in multicellular organisms are involved in development.[14] Responding to cues (stimuli), these transcription factors turn on/off the transcription of the appropriate genes which in turn allows for changes in cell morphology or activities needed for cell fate determination and cellular differentiation. The Hox transcription factor family, for example, is important for proper body pattern formation in organisms as diverse as fruit flies to humans.[15][16] Another example is the transcription factor encoded by the Sex-determining Region Y (SRY) gene which plays a major role in determining gender in humans.[17]

Response to intercellular signals

Cells can communicate with each other by releasing molecules that produce signaling cascades within another receptive cell. If the signal requires upregulation or downregulation of genes in the recipient cell, often transcription factors will be downstream in the signaling cascade.[18] Estrogen signaling is an example of a fairly short signaling cascade that involves the estrogen receptor transcription factor: estrogen is secreted by tissues such as the ovaries and placenta, crosses the cell membrane of the recipient cell, and is bound by the estrogen receptor in the cell's cytoplasm. The estrogen receptor then goes to the cell's nucleus and binds to its DNA binding sites, changing the transcriptional regulation of the associated genes.[19]

Response to environment

Not only do transcription factors act downstream of signaling cascades related to biological stimuli, but they can also be downstream of signaling cascades involved in environmental stimuli. Examples include heat shock factor (HSF) which upregulates genes necessary for survival at higher temperatures,[20] hypoxia inducible factor (HIF) which upregulates genes necessary for cell survival in low oxygen environments,[21] and sterol regulatory element binding protein (SREBP) which helps maintain proper lipid levels in the cell.[22]

Cell cycle control

Many transcription factors, especially some that are oncogenes or tumor suppressors, help regulate the cell cycle and as such determine how large a cell will get and when it can divide into two daughter cells. One example is the Myc oncogene, which has important roles in cell growth and apoptosis.

Regulation

It is common in biology for important processes to have multiple layers of regulation and control. This is also true with transcription factors: not only do transcription factors control the rates of transcription to regulate the amounts of gene products (RNA and protein) available to the cell, but transcription factors themselves are regulated (often by other transcription factors). Below is a brief synopsis of some of the ways that the activity of transcription factors can be regulated:

Synthesis

Transcription factors (like all proteins) are transcribed from a gene on a chromosome into RNA, and then the RNA is translated into protein. Any of these steps can be regulated to affect the production (and thus activity) of a transcription factor. One interesting implication of this is that transcription factors can regulate themselves. For example, in a negative feedback loop, the transcription factor acts as its own repressor: if the transcription factor protein binds the DNA of its own gene, it will down-regulate the production of more of itself. This is one mechanism to maintain low levels of a transcription factor in a cell.

Nuclear localization

In eukaryotes, transcription factors (like most proteins) are transcribed in the nucleus but are then translated in the cell's cytoplasm. Many proteins that are active in the nucleus contain nuclear localization signals that direct them to the nucleus. But for many transcription factors this is a key point in their regulation.[23] Important classes of transcription factors such as some nuclear receptors must first bind a ligand while in the cytoplasm before they can relocate to the nucleus.[23]

Activation

Transcription factors may be activated (or deactivated) through their signal sensing domain by a number of mechanisms including:

Accessibility of DNA binding site

In eukaryotes, genes that are not being actively transcribed are often located in heterochromatin. Heterochromatin are regions of chromosomes that are heavily compacted by tightly bundling the DNA onto histones and then organizing the histones into compact chromatin fibers. DNA within heterochromatin is inaccessible to many transcription factors. For the transcription factor to bind to its DNA binding site the heterochromatin must be first converted to euchromatin, usually via histone modifications. A transcription factor's DNA binding site may also be inaccessible if the site is already occupied by another transcription factor. Pairs of transcription factors can play antagonistic roles (activator versus repressor) in the regulation of the same gene.

Availability of other cofactors/transcription factors

Most transcription factors don't work alone. Often for gene transcription to occur, a number of transcription factors must bind to DNA regulatory sequences. This collection of transcription factors in turn recruit intermediary proteins such as cofactors that allow efficient recruitment of the preinitiation complex and RNA polymerase. Thus, for a single transcription factor to initiate transcription, all of these other proteins must also be present and the transcription factor must be in a state where it can bind to them if necessary.

Structure

Schematic diagram of the amino acid sequence (amino terminus to the left and carboxylic acid terminus to the right) of a prototypical transcription factor which contains (1) a DNA-binding domain (DBD), (2) signal sensing domain (SSD), and a transactivation domain (TAD). The order of placement and the number of domains may differ in various types of transcription factors. In addition, the transactivation and signal sensing functions are frequently contained within the same domain.

Transcription factors are modular in structure and contain the following domains:[1]

DNA binding domain

Main article: DNA binding domain

The portion (domain) of the transcription factor that binds DNA is called its DNA binding domain. Below is a partial list of some of the major families of DNA-binding domains/transcription factors:

Family SCOP InterPro
basic-helix-loop-helix[27] SCOP 47460 IPR001092
basic-leucine zipper (bZIP)[28] SCOP 57959 IPR004827
C-terminal effector domain of the bipartite response regulators SCOP 46894 IPR001789
GCC box SCOP 54175
helix-turn-helix[29]
homeodomain proteins - bind to homeobox DNA sequences which in turn encode other transcription factors. Homeodomain proteins play critical roles in the regulation of development.[30] SCOP 46689 IPR009057
lambda repressor-like SCOP 47413 IPR010982
srf-like (serum response factor) SCOP 55455 IPR002100
paired box[31]
winged helix SCOP 46785 IPR011991
zinc fingers[32]
* multi-domain Cys2His2 zinc fingers[33] SCOP 57667 IPR007087
* Zn2/Cys6 SCOP 57701
* Zn2/Cys8 nuclear receptor zinc finger SCOP 57716 IPR001628

Response elements

The DNA sequence that a transcription factor binds to is called a transcription factor binding site or response element.

Chemically, transcription factors interact with their binding sites using a combination of electrostatic (of which hydrogen bonds are a special case) and Van der Waals forces. Due to the nature of these chemical interactions, most transcription factors bind DNA in a sequence specific manner. However, not all bases in the transcription factor binding site may actually interact with the transcription factor. In addition some of these interactions may be weaker than others. Thus, transcription factors don't bind just one sequence but are capable of binding a subset of closely related sequences, each with a different strength of interaction.

For example, although the consensus binding site for the TATA binding protein (TBP) is TATAAAA. The TBP transcription factor can also bind similar sequences such as TATATAT or TATATAA.

Because transcription factors can bind a set of related sequences and these sequences tend to be short, potential transcription factor binding sites can occur by chance if the DNA sequence is long enough. It is unlikely, however, that a transcription factor binds all compatible sequences in the genome of the cell. Other constraints, such as DNA accessibility in the cell or availability of cofactors may also help dictate where a transcription factor will actually bind. Thus, given the genome sequence it is still difficult to predict where a transcription factor will actually bind in a living cell.

Additional recognition specificity however may be obtained through the use of more than one DNA binding domain (for example tandem DBDs in the same transcription factor or through dimerization of two transcription factors) which bind to two or more adjacent sequences of DNA.

Clinical significance

Due to their important roles in development, intercellular signaling, and cell cycle, some human diseases have been associated with mutations in transcription factors. Below are a few of the more well-studied examples:

Classes

As described in more detail below, transcription factors may be classified by their (1) mechanism of action, (2) regulatory function, or (3) sequence homology in their DNA binding domains.

Mechanistic

There are three mechanistic classes of transcription factors:

Functional

Transcription factors have been classified according to their regulatory function:[8]

Structural

Transcription factors are often classified based on the sequence similarity and hence the tertiary structure of their DNA binding domains:[35][36][37]

See also

  • DNA-binding protein
  • Inhibitor of DNA-binding protein
  • Nuclear receptor, a class of ligand activated transcription factors
  • Phylogenetic footprinting

References

  1. 1.0 1.1 Latchman DS (1997). "Transcription factors: an overview". Int. J. Biochem. Cell Biol. 29 (12): 1305–12. doi:10.1016/S1357-2725(97)00085-X. PMID 9570129. 
  2. Karin M (1990). "Too many transcription factors: positive and negative interactions". New Biol. 2 (2): 126–31. PMID 2128034. 
  3. Roeder RG (1996). "The role of general initiation factors in transcription by RNA polymerase II". Trends Biochem. Sci. 21 (9): 327–35. doi:10.1016/0968-0004(96)10050-5. PMID 8870495. 
  4. Nikolov DB, Burley SK (1997). "RNA polymerase II transcription initiation: a structural view". Proc. Natl. Acad. Sci. U.S.A. 94 (1): 15–22. doi:10.1073/pnas.94.1.15. PMID 8990153. 
  5. Lee TI, Young RA (2000). "Transcription of eukaryotic protein-coding genes". Annu. Rev. Genet. 34: 77–137. doi:10.1146/annurev.genet.34.1.77. PMID 11092823. 
  6. Mitchell PJ, Tjian R (1989). "Transcriptional regulation in mammalian cells by sequence-specific DNA binding proteins". Science 245 (4916): 371–8. doi:10.1126/science.2667136. PMID 2667136. 
  7. Ptashne M, Gann A (1997). "Transcriptional activation by recruitment". Nature 386 (6625): 569–77. doi:10.1038/386569a0. PMID 9121580. 
  8. 8.0 8.1 8.2 Brivanlou AH, Darnell JE (2002). "Signal transduction and the control of gene expression". Science 295 (5556): 813–8. doi:10.1126/science.1066355. PMID 11823631. 
  9. van Nimwegen E (2003). "Scaling laws in the functional content of genomes". Trends Genet. 19 (9): 479–84. doi:10.1016/S0168-9525(03)00203-8. PMID 12957540. 
  10. Babu MM, Luscombe NM, Aravind L, Gerstein M, Teichmann SA (2004). "Structure and evolution of transcriptional regulatory networks". Curr. Opin. Struct. Biol. 14 (3): 283–91. doi:10.1016/j.sbi.2004.05.004. PMID 15193307. 
  11. Reese JC (April 2003). "Basal transcription factors". Current opinion in genetics & development 13 (2): 114–8. doi:10.1016/S0959-437X(03)00013-3. PMID 12672487. 
  12. Shilatifard A, Conaway RC, Conaway JW (2003). "The RNA polymerase II elongation complex". Annual review of biochemistry 72: 693–715. doi:10.1146/annurev.biochem.72.121801.161551. PMID 12676794. 
  13. Thomas MC, Chiang CM (2006). "The general transcription machinery and general cofactors". Critical reviews in biochemistry and molecular biology 41 (3): 105–78. PMID 16858867. 
  14. Lobe CG (1992). "Transcription factors and mammalian development". Current topics in developmental biology 27: 351–83. PMID 1424766. 
  15. Lemons D, McGinnis W (September 2006). "Genomic evolution of Hox gene clusters". Science (New York, N.Y.) 313 (5795): 1918–22. doi:10.1126/science.1132040. PMID 17008523. 
  16. Moens CB, Selleri L (March 2006). "Hox cofactors in vertebrate development". Developmental biology 291 (2): 193–206. doi:10.1016/j.ydbio.2005.10.032. PMID 16515781. 
  17. Ottolenghi C, Uda M, Crisponi L, Omari S, Cao A, Forabosco A, Schlessinger D (January 2007). "Determination and stability of sex". BioEssays : news and reviews in molecular, cellular and developmental biology 29 (1): 15–25. doi:10.1002/bies.20515. PMID 17187356. 
  18. Pawson T (1993). "Signal transduction--a conserved pathway from the membrane to the nucleus". Developmental genetics 14 (5): 333–8. doi:10.1002/dvg.1020140502. PMID 8293575. 
  19. Osborne CK, Schiff R, Fuqua SA, Shou J (December 2001). "Estrogen receptor: current understanding of its activation and modulation". Clin. Cancer Res. 7 (12 Suppl): 4338s–4342s; discussion 4411s–4412s. PMID 11916222. 
  20. Shamovsky I, Nudler E (March 2008). "New insights into the mechanism of heat shock response activation". Cell. Mol. Life Sci. 65 (6): 855–61. doi:10.1007/s00018-008-7458-y. PMID 18239856. 
  21. Benizri E, Ginouvès A, Berra E (April 2008). "The magic of the hypoxia-signaling cascade". Cell. Mol. Life Sci. 65 (7-8): 1133–49. doi:10.1007/s00018-008-7472-0. PMID 18202826. 
  22. Weber LW, Boll M, Stampfl A (November 2004). "Maintaining cholesterol homeostasis: sterol regulatory element-binding proteins". World J. Gastroenterol. 10 (21): 3081–7. PMID 15457548. http://www.wjgnet.com/1007-9327/10/3081.asp. 
  23. 23.0 23.1 Whiteside ST, Goodbourn S (April 1993). "Signal transduction and nuclear targeting: regulation of transcription factor activity by subcellular localisation". Journal of cell science 104 ( Pt 4): 949–55. PMID 8314906. http://jcs.biologists.org/cgi/pmidlookup?view=long&pmid=8314906. 
  24. Bohmann D (November 1990). "Transcription factor phosphorylation: a link between signal transduction and the regulation of gene expression". Cancer cells (Cold Spring Harbor, N.Y. : 1989) 2 (11): 337–44. PMID 2149275. 
  25. Weigel NL, Moore NL (2007). Steroid Receptor Phosphorylation: A Key Modulator of Multiple Receptor Functions. doi:10.1210/me.2007-0101. PMID 17536004. 
  26. Wärnmark A, Treuter E, Wright AP, Gustafsson J-Å (2003). "Activation functions 1 and 2 of nuclear receptors: molecular strategies for transcriptional activation". Mol. Endocrinol. 17 (10): 1901–9. doi:10.1210/me.2002-0384. PMID 12893880. 
  27. Littlewood TD, Evan GI (1995). "Transcription factors 2: helix-loop-helix". Protein profile 2 (6): 621–702. PMID 7553065. 
  28. Vinson C, Myakishev M, Acharya A, Mir AA, Moll JR, Bonovich M (September 2002). "Classification of human B-ZIP proteins based on dimerization properties". Molecular and cellular biology 22 (18): 6321–35. doi:10.1128/MCB.22.18.6321-6335.2002. PMID 12192032. 
  29. Wintjens R, Rooman M (September 1996). "Structural classification of HTH DNA-binding domains and protein-DNA interaction modes". Journal of molecular biology 262 (2): 294–313. doi:10.1006/jmbi.1996.0514. PMID 8831795. 
  30. Gehring WJ, Affolter M, Bürglin T (1994). "Homeodomain proteins". Annual review of biochemistry 63: 487–526. doi:10.1146/annurev.bi.63.070194.002415. PMID 7979246. 
  31. Dahl E, Koseki H, Balling R (September 1997). "Pax genes and organogenesis". BioEssays : news and reviews in molecular, cellular and developmental biology 19 (9): 755–65. doi:10.1002/bies.950190905. PMID 9297966. 
  32. Laity JH, Lee BM, Wright PE (February 2001). "Zinc finger proteins: new insights into structural and functional diversity". Current opinion in structural biology 11 (1): 39–46. doi:10.1016/S0959-440X(00)00167-6. PMID 11179890. 
  33. Wolfe SA, Nekludova L, Pabo CO (2000). "DNA recognition by Cys2His2 zinc finger proteins". Annual review of biophysics and biomolecular structure 29: 183–212. doi:10.1146/annurev.biophys.29.1.183. PMID 10940247. 
  34. Orphanides G, Lagrange T, Reinberg D (1996). "The general transcription factors of RNA polymerase II". Genes Dev. 10 (21): 2657–83. doi:10.1101/gad.10.21.2657. PMID 8946909. 
  35. Stegmaier P, Kel AE, Wingender E (2004). "Systematic DNA-binding domain classification of transcription factors". Genome informatics. International Conference on Genome Informatics 15 (2): 276–86. PMID 15706513. http://www.jsbi.org/journal/GIW04/GIW04F028.html. 
  36. Matys V, Kel-Margoulis OV, Fricke E, Liebich I, Land S, Barre-Dirrie A, Reuter I, Chekmenev D, Krull M, Hornischer K, Voss N, Stegmaier P, Lewicki-Potapov B, Saxel H, Kel AE, Wingender E (2006). "TRANSFAC and its module TRANSCompel: transcriptional gene regulation in eukaryotes". Nucleic Acids Res. 34 (Database issue): D108–10. doi:10.1093/nar/gkj143. PMID 16381825. 
  37. "TRANSFAC® database". Retrieved on 2007-08-05.

Further reading

  • Wilson D, Charoensawan V, Kummerfeld SK, Teichmann SA (2008). "DBD--taxonomically broad transcription factor predictions: new content and functionality". Nucleic Acids Res. 36 (Database issue): D88–92. doi:10.1093/nar/gkm964. PMID 18073188. 
  • Singer, Susan R.; Gilbert, Scott F. (2006). Developmental Biology. Sunderland, Mass: Sinauer Associates. ISBN 0-87893-250-X. 
  • Bruce Alberts, Dennis Bray, Karen Hopkin, Alexander Johnson, Julian Lewis, Martin Raff, Keith Roberts, Peter Walter (2004). Essential cell biology. New York: Garland Science. pp. 896 pages. ISBN 0-8153-3480-X. 

External links