id
stringlengths
14
28
title
stringclasses
18 values
content
stringlengths
2
999
contents
stringlengths
19
1.02k
Cell_Biology_Alberts_1210
Cell_Biology_Alberts
origins of replication telomere centromere telomere Figure 5–30 The origins of DNA replication on chromosome III of the yeast S. cerevisiae. This chromosome, one of the smallest eukaryotic chromosomes known, carries a total of 180 genes. As indicated, it contains 18 replication origins, although they are used with different frequencies. Those in red are typically used in less than 10% of cell divisions, while those in green are used about 90% of the time.
Cell_Biology_Alberts. origins of replication telomere centromere telomere Figure 5–30 The origins of DNA replication on chromosome III of the yeast S. cerevisiae. This chromosome, one of the smallest eukaryotic chromosomes known, carries a total of 180 genes. As indicated, it contains 18 replication origins, although they are used with different frequencies. Those in red are typically used in less than 10% of cell divisions, while those in green are used about 90% of the time.
Cell_Biology_Alberts_1211
Cell_Biology_Alberts
Figure 5–31 DNA replication initiation in eukaryotes. This mechanism ensures that each origin of replication is activated only once per cell cycle. An origin of replication can be used only if a prereplicative complex forms there in G1 phase. At the beginning of S phase, specialized kinases phosphorylate Mcm and ORC, activating the former and inactivating the latter. A new prereplicative complex cannot form at the origin until the cell progresses to the next G1 phase, when the bound ORC has been dephosphorylated. Note that the eukaryotic Mcm helicase moves along the leading-strand template, whereas the bacterial helicase moves along the lagging-strand template (see Figure 5–25). As the forks begin to move, ORC is displaced, and new ORCs rapidly bind to the newly replicated origins.
Cell_Biology_Alberts. Figure 5–31 DNA replication initiation in eukaryotes. This mechanism ensures that each origin of replication is activated only once per cell cycle. An origin of replication can be used only if a prereplicative complex forms there in G1 phase. At the beginning of S phase, specialized kinases phosphorylate Mcm and ORC, activating the former and inactivating the latter. A new prereplicative complex cannot form at the origin until the cell progresses to the next G1 phase, when the bound ORC has been dephosphorylated. Note that the eukaryotic Mcm helicase moves along the leading-strand template, whereas the bacterial helicase moves along the lagging-strand template (see Figure 5–25). As the forks begin to move, ORC is displaced, and new ORCs rapidly bind to the newly replicated origins.
Cell_Biology_Alberts_1212
Cell_Biology_Alberts
The protein kinases that trigger DNA replication simultaneously prevent assembly of new prereplicative complexes until the next M phase resets the entire cycle (for details, see pp. 974–975). They do this, in part, by phosphorylating ORC, rendering it unable to accept new helicases. This strategy provides a single window of opportunity for prereplicative complexes to form (G1 phase, when kinase activity is low) and a second window for them to be activated and subsequently disassembled (S phase, when kinase activity is high). Because these two phases of the cell cycle are mutually exclusive and occur in a prescribed order, each origin of replication can fire once and only once during each cell cycle. Features of the Human Genome That Specify Origins of Replication Remain to Be Discovered
Cell_Biology_Alberts. The protein kinases that trigger DNA replication simultaneously prevent assembly of new prereplicative complexes until the next M phase resets the entire cycle (for details, see pp. 974–975). They do this, in part, by phosphorylating ORC, rendering it unable to accept new helicases. This strategy provides a single window of opportunity for prereplicative complexes to form (G1 phase, when kinase activity is low) and a second window for them to be activated and subsequently disassembled (S phase, when kinase activity is high). Because these two phases of the cell cycle are mutually exclusive and occur in a prescribed order, each origin of replication can fire once and only once during each cell cycle. Features of the Human Genome That Specify Origins of Replication Remain to Be Discovered
Cell_Biology_Alberts_1213
Cell_Biology_Alberts
Features of the Human Genome That Specify Origins of Replication Remain to Be Discovered Compared with the situation in budding yeast, the determinants of replication origins in other eukaryotes have been difficult to discover. It has been possible to identify specific human DNA sequences, each several thousand nucleotide pairs in length, that are sufficient to serve as replication origins. These origins continue to function when moved to a different chromosomal region by recombinant DNA methods, as long as they are placed in a region where the chromatin is relatively uncondensed. However, comparisons of such DNA sequences have not revealed specific DNA sequences that mark origins of replication.
Cell_Biology_Alberts. Features of the Human Genome That Specify Origins of Replication Remain to Be Discovered Compared with the situation in budding yeast, the determinants of replication origins in other eukaryotes have been difficult to discover. It has been possible to identify specific human DNA sequences, each several thousand nucleotide pairs in length, that are sufficient to serve as replication origins. These origins continue to function when moved to a different chromosomal region by recombinant DNA methods, as long as they are placed in a region where the chromatin is relatively uncondensed. However, comparisons of such DNA sequences have not revealed specific DNA sequences that mark origins of replication.
Cell_Biology_Alberts_1214
Cell_Biology_Alberts
Despite this, a human ORC that is very similar to the yeast ORC binds to origins of replication and initiates DNA replication in humans. Many of the other proteins that function in the initiation process in yeast likewise have central roles in humans. It therefore seems likely that the yeast and human initiation mechanisms are similar in outline, but chromatin structure, transcriptional activity, or some property of the genome other than a specific DNA sequence has the central role in attracting ORC and specifying mammalian origins of replication. These ideas could also help to explain how a given mammalian cell chooses which of the many possible origins to use when it replicates its genome and how this choice could differ from cell to cell. Clearly, we have a great deal to discover about the fundamental process of DNA replication initiation. New Nucleosomes Are Assembled Behind the Replication Fork
Cell_Biology_Alberts. Despite this, a human ORC that is very similar to the yeast ORC binds to origins of replication and initiates DNA replication in humans. Many of the other proteins that function in the initiation process in yeast likewise have central roles in humans. It therefore seems likely that the yeast and human initiation mechanisms are similar in outline, but chromatin structure, transcriptional activity, or some property of the genome other than a specific DNA sequence has the central role in attracting ORC and specifying mammalian origins of replication. These ideas could also help to explain how a given mammalian cell chooses which of the many possible origins to use when it replicates its genome and how this choice could differ from cell to cell. Clearly, we have a great deal to discover about the fundamental process of DNA replication initiation. New Nucleosomes Are Assembled Behind the Replication Fork
Cell_Biology_Alberts_1215
Cell_Biology_Alberts
Several additional aspects of DNA replication are specific to eukaryotes. As discussed in Chapter 4, eukaryotic chromosomes are composed of roughly equal mixtures of DNA and protein. Chromosome duplication therefore requires not only the replication of DNA, but also the synthesis and assembly of new chromosomal proteins onto the DNA behind each replication fork. Although we are far from understanding this process in detail, we are beginning to learn how the fundamental unit of chromatin packaging, the nucleosome, is duplicated. The cell requires a large amount of new histone protein, approximately equal in mass to the newly synthesized DNA, to make the new nucleosomes in each cell cycle. For this reason, most eukaryotic organisms possess multiple copies of the gene for each histone. Vertebrate cells, for example, have about 20 repeated gene sets, most containing the genes that encode all five histones (H1, H2A, H2B, H3, and H4).
Cell_Biology_Alberts. Several additional aspects of DNA replication are specific to eukaryotes. As discussed in Chapter 4, eukaryotic chromosomes are composed of roughly equal mixtures of DNA and protein. Chromosome duplication therefore requires not only the replication of DNA, but also the synthesis and assembly of new chromosomal proteins onto the DNA behind each replication fork. Although we are far from understanding this process in detail, we are beginning to learn how the fundamental unit of chromatin packaging, the nucleosome, is duplicated. The cell requires a large amount of new histone protein, approximately equal in mass to the newly synthesized DNA, to make the new nucleosomes in each cell cycle. For this reason, most eukaryotic organisms possess multiple copies of the gene for each histone. Vertebrate cells, for example, have about 20 repeated gene sets, most containing the genes that encode all five histones (H1, H2A, H2B, H3, and H4).
Cell_Biology_Alberts_1216
Cell_Biology_Alberts
Unlike most proteins, which are made continuously, histones are synthesized mainly in S phase, when the level of histone mRNA increases about fiftyfold as a result of both increased transcription and decreased mRNA degradation. The major histone mRNAs are degraded within minutes when DNA synthesis stops at the end of S phase. The mechanism depends on special properties of the 3ʹends of these mRNAs, as discussed in Chapter 7. In contrast, the histone proteins themselves are remarkably stable and may survive for the entire life of a cell. The tight linkage between DNA synthesis and histone synthesis appears to reflect a feedback mechanism that monitors the level of free histone to ensure that the amount of histone made exactly matches the amount of new DNA synthesized.
Cell_Biology_Alberts. Unlike most proteins, which are made continuously, histones are synthesized mainly in S phase, when the level of histone mRNA increases about fiftyfold as a result of both increased transcription and decreased mRNA degradation. The major histone mRNAs are degraded within minutes when DNA synthesis stops at the end of S phase. The mechanism depends on special properties of the 3ʹends of these mRNAs, as discussed in Chapter 7. In contrast, the histone proteins themselves are remarkably stable and may survive for the entire life of a cell. The tight linkage between DNA synthesis and histone synthesis appears to reflect a feedback mechanism that monitors the level of free histone to ensure that the amount of histone made exactly matches the amount of new DNA synthesized.
Cell_Biology_Alberts_1217
Cell_Biology_Alberts
As a replication fork advances, it must pass through the parental nucleosomes. In the cell, efficient replication requires chromatin remodeling complexes (discussed in Chapter 4) to destabilize the DNA–histone interfaces. Aided by such complexes, replication forks can transit even highly condensed chromatin efficiently.
Cell_Biology_Alberts. As a replication fork advances, it must pass through the parental nucleosomes. In the cell, efficient replication requires chromatin remodeling complexes (discussed in Chapter 4) to destabilize the DNA–histone interfaces. Aided by such complexes, replication forks can transit even highly condensed chromatin efficiently.
Cell_Biology_Alberts_1218
Cell_Biology_Alberts
As a replication fork passes through chromatin, the histones are transiently displaced leaving about 600 nucleotide pairs of non-nucleosomal DNA in its wake. The reestablishment of nucleosomes behind a moving fork occurs in an intriguing way. When a nucleosome is traversed by a replication fork, the histone octamer appears to be broken into an H3-H4 tetramer and two H2A-H2B dimers (discussed in Chapter 4). The H3-H4 tetramer remains loosely associated with DNA and is distributed at random to one or the other daughter duplex, but the H2A-H2B dimers are released completely from DNA. Freshly made H3-H4 tetramers are added to the newly synthesized DNA to fill in the “spaces,” and H2A-H2B dimers—half of which are old and half new—are then added at random to complete the nucleosomes (Figure 5–32). The formation of new nucleosomes behind a replication fork has an important consequence for the process of DNA replication itself. As DNA polymerase δ discontinuously synthesizes the lagging
Cell_Biology_Alberts. As a replication fork passes through chromatin, the histones are transiently displaced leaving about 600 nucleotide pairs of non-nucleosomal DNA in its wake. The reestablishment of nucleosomes behind a moving fork occurs in an intriguing way. When a nucleosome is traversed by a replication fork, the histone octamer appears to be broken into an H3-H4 tetramer and two H2A-H2B dimers (discussed in Chapter 4). The H3-H4 tetramer remains loosely associated with DNA and is distributed at random to one or the other daughter duplex, but the H2A-H2B dimers are released completely from DNA. Freshly made H3-H4 tetramers are added to the newly synthesized DNA to fill in the “spaces,” and H2A-H2B dimers—half of which are old and half new—are then added at random to complete the nucleosomes (Figure 5–32). The formation of new nucleosomes behind a replication fork has an important consequence for the process of DNA replication itself. As DNA polymerase δ discontinuously synthesizes the lagging
Cell_Biology_Alberts_1219
Cell_Biology_Alberts
5–32). The formation of new nucleosomes behind a replication fork has an important consequence for the process of DNA replication itself. As DNA polymerase δ discontinuously synthesizes the lagging strand (see pp. 253–254), the length of each Okazaki fragment is determined by the point at which DNA polymerase δ is blocked by a newly formed nucleosome. This tight coupling between nucleosome duplication and DNA replication explains why the length of Okazaki fragments in eukaryotes (~200 nucleotides) is approximately the same as the nucleosome repeat length.
Cell_Biology_Alberts. 5–32). The formation of new nucleosomes behind a replication fork has an important consequence for the process of DNA replication itself. As DNA polymerase δ discontinuously synthesizes the lagging strand (see pp. 253–254), the length of each Okazaki fragment is determined by the point at which DNA polymerase δ is blocked by a newly formed nucleosome. This tight coupling between nucleosome duplication and DNA replication explains why the length of Okazaki fragments in eukaryotes (~200 nucleotides) is approximately the same as the nucleosome repeat length.
Cell_Biology_Alberts_1220
Cell_Biology_Alberts
262 Chapter 5: DNA Replication, Repair, and Recombination H2A-H2B dimer displaced in front of replication fork The orderly and rapid addition of new H3-H4 tetramers and H2A-H2B dimers behind a replication fork requires histone chaperones (also called chromatin assembly factors). These multisubunit complexes bind the highly basic histones and release them for assembly only in the appropriate context. The histone chaperones, along with their cargoes, are directed to newly replicated DNA through a specific interaction with the eukaryotic sliding clamp called PCNA (see Figure 5–32B). These clamps are left behind moving replication forks and remain on the DNA long enough for the histone chaperones to complete their tasks. Telomerase Replicates the Ends of Chromosomes
Cell_Biology_Alberts. 262 Chapter 5: DNA Replication, Repair, and Recombination H2A-H2B dimer displaced in front of replication fork The orderly and rapid addition of new H3-H4 tetramers and H2A-H2B dimers behind a replication fork requires histone chaperones (also called chromatin assembly factors). These multisubunit complexes bind the highly basic histones and release them for assembly only in the appropriate context. The histone chaperones, along with their cargoes, are directed to newly replicated DNA through a specific interaction with the eukaryotic sliding clamp called PCNA (see Figure 5–32B). These clamps are left behind moving replication forks and remain on the DNA long enough for the histone chaperones to complete their tasks. Telomerase Replicates the Ends of Chromosomes
Cell_Biology_Alberts_1221
Cell_Biology_Alberts
Telomerase Replicates the Ends of Chromosomes We saw earlier that synthesis of the lagging strand at a replication fork must occur discontinuously through a backstitching mechanism that produces short DNA fragments. This mechanism encounters a special problem when the replication fork reaches an end of a linear chromosome. The final RNA primer synthesized on the lagging-strand template cannot be replaced by DNA because there is no 3ʹ-OH end available for the repair polymerase. Without a mechanism to deal with this problem, DNA would be lost from the ends of all chromosomes each time a cell divides.
Cell_Biology_Alberts. Telomerase Replicates the Ends of Chromosomes We saw earlier that synthesis of the lagging strand at a replication fork must occur discontinuously through a backstitching mechanism that produces short DNA fragments. This mechanism encounters a special problem when the replication fork reaches an end of a linear chromosome. The final RNA primer synthesized on the lagging-strand template cannot be replaced by DNA because there is no 3ʹ-OH end available for the repair polymerase. Without a mechanism to deal with this problem, DNA would be lost from the ends of all chromosomes each time a cell divides.
Cell_Biology_Alberts_1222
Cell_Biology_Alberts
Bacteria solve this “end-replication” problem by having circular DNA molecules as chromosomes (see Figure 5–24). Eukaryotes solve it in a different way: they have specialized nucleotide sequences at the ends of their chromosomes that are incorporated into structures called telomeres (discussed in Chapter 4). Telomeres contain many tandem repeats of a short sequence that is similar in organisms as diverse as protozoa, fungi, plants, and mammals. In humans, the sequence of the repeat unit is GGGTTA, and it is repeated roughly a thousand times at each telomere.
Cell_Biology_Alberts. Bacteria solve this “end-replication” problem by having circular DNA molecules as chromosomes (see Figure 5–24). Eukaryotes solve it in a different way: they have specialized nucleotide sequences at the ends of their chromosomes that are incorporated into structures called telomeres (discussed in Chapter 4). Telomeres contain many tandem repeats of a short sequence that is similar in organisms as diverse as protozoa, fungi, plants, and mammals. In humans, the sequence of the repeat unit is GGGTTA, and it is repeated roughly a thousand times at each telomere.
Cell_Biology_Alberts_1223
Cell_Biology_Alberts
Telomere DNA sequences are recognized by sequence-specific DNA-binding proteins that attract an enzyme, called telomerase, that replenishes these sequences each time a cell divides. Telomerase recognizes the tip of an existing telomere DNA repeat sequence and elongates it in the 5ʹ-to-3ʹ direction, using an RNA template that is a component of the enzyme itself to synthesize new copies of the repeat (Figure 5–33). The enzymatic portion of telomerase resembles other reverse transcriptases, proteins that synthesize DNA using an RNA template, although, in this case, the telomerase RNA also contributes functional groups to make the catalysis more efficient. After extension of the parental DNA strand by telomerase, replication of the lagging strand at the chromosome end can be completed by the conventional DNA polymerases, using these extensions as a template to synthesize the complementary strand (Figure 5–34).
Cell_Biology_Alberts. Telomere DNA sequences are recognized by sequence-specific DNA-binding proteins that attract an enzyme, called telomerase, that replenishes these sequences each time a cell divides. Telomerase recognizes the tip of an existing telomere DNA repeat sequence and elongates it in the 5ʹ-to-3ʹ direction, using an RNA template that is a component of the enzyme itself to synthesize new copies of the repeat (Figure 5–33). The enzymatic portion of telomerase resembles other reverse transcriptases, proteins that synthesize DNA using an RNA template, although, in this case, the telomerase RNA also contributes functional groups to make the catalysis more efficient. After extension of the parental DNA strand by telomerase, replication of the lagging strand at the chromosome end can be completed by the conventional DNA polymerases, using these extensions as a template to synthesize the complementary strand (Figure 5–34).
Cell_Biology_Alberts_1224
Cell_Biology_Alberts
Figure 5–32 Formation of nucleosomes behind a replication fork. Parental H3-H4 tetramers are distributed at random to the daughter DNA molecules, with roughly equal numbers inherited by each daughter. In contrast, H2A-H2B dimers are released from the DNA as the replication fork passes. This release begins just in front of the replication fork and is facilitated by chromatin remodeling complexes that move with the fork. Histone chaperones (NAP1 and CAF1) restore the full complement of histones to daughter molecules using both parental and newly synthesized histones. Although some daughter nucleosomes contain only parental histones or only newly synthesized histones, most are hybrids of old and new. For simplicity, the DNA double helix shown as a single red line. (Adapted from J.D. Watson et al., Molecular Biology of the Gene, 5th ed. Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 2004.) remainder of telomerase proteintelomerase RNA region of telomerase RNA used as
Cell_Biology_Alberts. Figure 5–32 Formation of nucleosomes behind a replication fork. Parental H3-H4 tetramers are distributed at random to the daughter DNA molecules, with roughly equal numbers inherited by each daughter. In contrast, H2A-H2B dimers are released from the DNA as the replication fork passes. This release begins just in front of the replication fork and is facilitated by chromatin remodeling complexes that move with the fork. Histone chaperones (NAP1 and CAF1) restore the full complement of histones to daughter molecules using both parental and newly synthesized histones. Although some daughter nucleosomes contain only parental histones or only newly synthesized histones, most are hybrids of old and new. For simplicity, the DNA double helix shown as a single red line. (Adapted from J.D. Watson et al., Molecular Biology of the Gene, 5th ed. Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 2004.) remainder of telomerase proteintelomerase RNA region of telomerase RNA used as
Cell_Biology_Alberts_1225
Cell_Biology_Alberts
Watson et al., Molecular Biology of the Gene, 5th ed. Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 2004.) remainder of telomerase proteintelomerase RNA region of telomerase RNA used as template “palm“—active site of telomerase protein rest of synthesized chromosome telomere DNA
Cell_Biology_Alberts. Watson et al., Molecular Biology of the Gene, 5th ed. Cold Spring Harbor: Cold Spring Harbor Laboratory Press, 2004.) remainder of telomerase proteintelomerase RNA region of telomerase RNA used as template “palm“—active site of telomerase protein rest of synthesized chromosome telomere DNA
Cell_Biology_Alberts_1226
Cell_Biology_Alberts
Telomeres Are Packaged Into Specialized Structures That Protect the Ends of Chromosomes The ends of chromosomes present cells with an additional problem. As we will see in the next part of this chapter, when a chromosome is accidently broken, the break is rapidly repaired (see Figure 5–45). Telomeres must clearly be distinguished from these accidental breaks; otherwise the cell will attempt to “repair” telomeres, causing chromosome fusions and other genetic abnormalities. Telomeres have several features to prevent this from happening.
Cell_Biology_Alberts. Telomeres Are Packaged Into Specialized Structures That Protect the Ends of Chromosomes The ends of chromosomes present cells with an additional problem. As we will see in the next part of this chapter, when a chromosome is accidently broken, the break is rapidly repaired (see Figure 5–45). Telomeres must clearly be distinguished from these accidental breaks; otherwise the cell will attempt to “repair” telomeres, causing chromosome fusions and other genetic abnormalities. Telomeres have several features to prevent this from happening.
Cell_Biology_Alberts_1227
Cell_Biology_Alberts
A specialized nuclease chews back the 5ʹend of a telomere leaving a protruding single-strand end. This protruding end—in combination with the GGGTTA repeats in telomeres—attracts a group of proteins that form a protective chromosome cap known as shelterin. In particular, shelterin “hides” telomeres from the cell’s damage detectors that continually monitor DNA. When human telomeres are artificially cross-linked and viewed by electron microscopy, structures known as “t-loops” are observed in which the protruding end of the telomere loops back and tucks itself into the duplex DNA of the telomere repeat sequence (Figure 5–35). It is believed that t-loops are regulated by shelterin and provide additional protection for the ends of chromosomes. incomplete, newly synthesized lagging strand direction of AACCCC telomere 5˜ synthesisTELOMERASE EXTENDS 3˜ END telomerase with bound RNA template
Cell_Biology_Alberts. A specialized nuclease chews back the 5ʹend of a telomere leaving a protruding single-strand end. This protruding end—in combination with the GGGTTA repeats in telomeres—attracts a group of proteins that form a protective chromosome cap known as shelterin. In particular, shelterin “hides” telomeres from the cell’s damage detectors that continually monitor DNA. When human telomeres are artificially cross-linked and viewed by electron microscopy, structures known as “t-loops” are observed in which the protruding end of the telomere loops back and tucks itself into the duplex DNA of the telomere repeat sequence (Figure 5–35). It is believed that t-loops are regulated by shelterin and provide additional protection for the ends of chromosomes. incomplete, newly synthesized lagging strand direction of AACCCC telomere 5˜ synthesisTELOMERASE EXTENDS 3˜ END telomerase with bound RNA template
Cell_Biology_Alberts_1228
Cell_Biology_Alberts
incomplete, newly synthesized lagging strand direction of AACCCC telomere 5˜ synthesisTELOMERASE EXTENDS 3˜ END telomerase with bound RNA template Figure 5–33 Structure of a portion of telomerase. Telomerase is a large protein– RNA complex. The RNA (blue) contains a templating sequence for synthesizing new DNA telomere repeats. The synthesis reaction itself is carried out by the reverse transcriptase domain of the protein, shown in green. A reverse transcriptase is a special form of polymerase enzyme that uses an RNA template to make a DNA strand; telomerase is unique in carrying its own RNA template with it. Telomerase also has several additional protein domains (not shown) that are needed to assemble the enzyme at the ends of chromosomes. (Modified from J. Lingner and T.R. Cech, Curr. Opin. Genet. Dev. 8:226–232, 1998. With permission from Elsevier.)
Cell_Biology_Alberts. incomplete, newly synthesized lagging strand direction of AACCCC telomere 5˜ synthesisTELOMERASE EXTENDS 3˜ END telomerase with bound RNA template Figure 5–33 Structure of a portion of telomerase. Telomerase is a large protein– RNA complex. The RNA (blue) contains a templating sequence for synthesizing new DNA telomere repeats. The synthesis reaction itself is carried out by the reverse transcriptase domain of the protein, shown in green. A reverse transcriptase is a special form of polymerase enzyme that uses an RNA template to make a DNA strand; telomerase is unique in carrying its own RNA template with it. Telomerase also has several additional protein domains (not shown) that are needed to assemble the enzyme at the ends of chromosomes. (Modified from J. Lingner and T.R. Cech, Curr. Opin. Genet. Dev. 8:226–232, 1998. With permission from Elsevier.)
Cell_Biology_Alberts_1229
Cell_Biology_Alberts
Figure 5–34 Telomere replication. Shown here are the reactions that synthesize the repeating sequences that form the ends of the chromosomes (telomeres) of diverse eukaryotic organisms. The 3ʹ end of the parental DNA strand is extended by RNA-templated DNA synthesis; this allows the incomplete daughter DNA strand that is paired with it to be extended in its 5ʹ direction. This incomplete, lagging strand is presumed to be completed by DNA polymerase α, which carries a DNA primase as one of its subunits (Movie 5.6). The telomere sequence illustrated is that of the ciliate Tetrahymena, in which these reactions were first discovered. Because the processes that grow and shrink each telomere sequence are only approximately balanced, a chromosome end contains a variable number of telomeric repeats. Not surprisingly, many cells have homeostatic mechanisms that maintain the number of these repeats within a limited range (Figure 5–36).
Cell_Biology_Alberts. Figure 5–34 Telomere replication. Shown here are the reactions that synthesize the repeating sequences that form the ends of the chromosomes (telomeres) of diverse eukaryotic organisms. The 3ʹ end of the parental DNA strand is extended by RNA-templated DNA synthesis; this allows the incomplete daughter DNA strand that is paired with it to be extended in its 5ʹ direction. This incomplete, lagging strand is presumed to be completed by DNA polymerase α, which carries a DNA primase as one of its subunits (Movie 5.6). The telomere sequence illustrated is that of the ciliate Tetrahymena, in which these reactions were first discovered. Because the processes that grow and shrink each telomere sequence are only approximately balanced, a chromosome end contains a variable number of telomeric repeats. Not surprisingly, many cells have homeostatic mechanisms that maintain the number of these repeats within a limited range (Figure 5–36).
Cell_Biology_Alberts_1230
Cell_Biology_Alberts
In most of the dividing somatic cells of humans, however, telomeres gradually shorten, and it has been proposed that this provides a counting mechanism that helps prevent the unlimited proliferation of wayward cells in adult tissues. In its simplest form, this idea holds that our somatic cells start off in the embryo with a full complement of telomeric repeats. These are then eroded to different extents in different cell types. Some stem cells, notably those in tissues that must be replenished at a high rate throughout life—bone marrow or gut lining, for example— retain full telomerase activity. However, in many other types of cells, the level of telomerase is turned down so that the enzyme cannot quite keep up with chromosome duplication. Such cells lose 100–200 nucleotides from each telomere every time they divide. After many cell generations, the descendant cells will inherit chromosomes that lack telomere function, and, as a result of this defect, activate a DNA-damage response
Cell_Biology_Alberts. In most of the dividing somatic cells of humans, however, telomeres gradually shorten, and it has been proposed that this provides a counting mechanism that helps prevent the unlimited proliferation of wayward cells in adult tissues. In its simplest form, this idea holds that our somatic cells start off in the embryo with a full complement of telomeric repeats. These are then eroded to different extents in different cell types. Some stem cells, notably those in tissues that must be replenished at a high rate throughout life—bone marrow or gut lining, for example— retain full telomerase activity. However, in many other types of cells, the level of telomerase is turned down so that the enzyme cannot quite keep up with chromosome duplication. Such cells lose 100–200 nucleotides from each telomere every time they divide. After many cell generations, the descendant cells will inherit chromosomes that lack telomere function, and, as a result of this defect, activate a DNA-damage response
Cell_Biology_Alberts_1231
Cell_Biology_Alberts
every time they divide. After many cell generations, the descendant cells will inherit chromosomes that lack telomere function, and, as a result of this defect, activate a DNA-damage response causing them to withdraw permanently from the cell cycle and cease dividing—a process called replicative cell senescence (discussed in Chapter 17). In theory, such a mechanism could provide a safeguard against the uncontrolled cell proliferation of abnormal cells in somatic tissues, thereby helping to protect us from cancer.
Cell_Biology_Alberts. every time they divide. After many cell generations, the descendant cells will inherit chromosomes that lack telomere function, and, as a result of this defect, activate a DNA-damage response causing them to withdraw permanently from the cell cycle and cease dividing—a process called replicative cell senescence (discussed in Chapter 17). In theory, such a mechanism could provide a safeguard against the uncontrolled cell proliferation of abnormal cells in somatic tissues, thereby helping to protect us from cancer.
Cell_Biology_Alberts_1232
Cell_Biology_Alberts
fraction of chromosome ends Figure 5–35 A t-loop at the end of a mammalian chromosome. (A) Electron micrograph of the DNA at the end of an interphase human chromosome. The chromosome was fixed, deproteinated, and artificially thickened before viewing. The loop seen here is approximately 15,000 nucleotide pairs in length. (B) Structure of a t-loop. The insertion of the single-strand 3ʹ end into the duplex repeats is carried out, and the structure maintained, by specialized proteins. (From J.D. Griffith et al., Cell 97:503–514, 1999. With permission from Elsevier.)
Cell_Biology_Alberts. fraction of chromosome ends Figure 5–35 A t-loop at the end of a mammalian chromosome. (A) Electron micrograph of the DNA at the end of an interphase human chromosome. The chromosome was fixed, deproteinated, and artificially thickened before viewing. The loop seen here is approximately 15,000 nucleotide pairs in length. (B) Structure of a t-loop. The insertion of the single-strand 3ʹ end into the duplex repeats is carried out, and the structure maintained, by specialized proteins. (From J.D. Griffith et al., Cell 97:503–514, 1999. With permission from Elsevier.)
Cell_Biology_Alberts_1233
Cell_Biology_Alberts
Figure 5–36 A demonstration that yeast cells control the length of their telomeres. In this experiment, the telomere at one end of a particular chromosome is artificially made either longer (left) or shorter (right) than average. After many cell divisions, the chromosome recovers, showing an average telomere length and a length distribution that is typical of the other chromosomes in the yeast cell. A similar feedback mechanism for controlling telomere length has been proposed for the germ-line cells of animals.
Cell_Biology_Alberts. Figure 5–36 A demonstration that yeast cells control the length of their telomeres. In this experiment, the telomere at one end of a particular chromosome is artificially made either longer (left) or shorter (right) than average. After many cell divisions, the chromosome recovers, showing an average telomere length and a length distribution that is typical of the other chromosomes in the yeast cell. A similar feedback mechanism for controlling telomere length has been proposed for the germ-line cells of animals.
Cell_Biology_Alberts_1234
Cell_Biology_Alberts
The idea that telomere length acts as a “measuring stick” to count cell divisions and thereby regulate the lifetime of the cell lineage has been tested in several ways. For certain types of human cells grown in tissue culture, the experimental results support such a theory. Human fibroblasts normally proliferate for about 60 cell divisions in culture before undergoing replicative cell senescence. Like most other somatic cells in humans, fibroblasts produce only low levels of telomerase, and their telomeres gradually shorten each time they divide. When telomerase is provided to the fibroblasts by inserting an active telomerase gene, telomere length is maintained and many of the cells now continue to proliferate indefinitely.
Cell_Biology_Alberts. The idea that telomere length acts as a “measuring stick” to count cell divisions and thereby regulate the lifetime of the cell lineage has been tested in several ways. For certain types of human cells grown in tissue culture, the experimental results support such a theory. Human fibroblasts normally proliferate for about 60 cell divisions in culture before undergoing replicative cell senescence. Like most other somatic cells in humans, fibroblasts produce only low levels of telomerase, and their telomeres gradually shorten each time they divide. When telomerase is provided to the fibroblasts by inserting an active telomerase gene, telomere length is maintained and many of the cells now continue to proliferate indefinitely.
Cell_Biology_Alberts_1235
Cell_Biology_Alberts
It has been proposed that this type of control on cell proliferation may contribute to the aging of animals like ourselves. These ideas have been tested by producing transgenic mice that lack telomerase entirely. The telomeres in mouse chromosomes are about five times longer than human telomeres, and the mice must therefore be bred through three or more generations before their telomeres have shrunk to the normal human length. It is therefore perhaps not surprising that the first generations of mice develop normally. However, the mice in later generations develop progressively more defects in some of their highly proliferative tissues. In addition, these mice show signs of premature aging and have a pronounced tendency to develop tumors. In these and other respects these mice resemble humans with the genetic disease dyskeratosis congenita. Individuals afflicted with this disease carry one functional and one nonfunctional copy of the telomerase RNA gene; they have prematurely shortened
Cell_Biology_Alberts. It has been proposed that this type of control on cell proliferation may contribute to the aging of animals like ourselves. These ideas have been tested by producing transgenic mice that lack telomerase entirely. The telomeres in mouse chromosomes are about five times longer than human telomeres, and the mice must therefore be bred through three or more generations before their telomeres have shrunk to the normal human length. It is therefore perhaps not surprising that the first generations of mice develop normally. However, the mice in later generations develop progressively more defects in some of their highly proliferative tissues. In addition, these mice show signs of premature aging and have a pronounced tendency to develop tumors. In these and other respects these mice resemble humans with the genetic disease dyskeratosis congenita. Individuals afflicted with this disease carry one functional and one nonfunctional copy of the telomerase RNA gene; they have prematurely shortened
Cell_Biology_Alberts_1236
Cell_Biology_Alberts
with the genetic disease dyskeratosis congenita. Individuals afflicted with this disease carry one functional and one nonfunctional copy of the telomerase RNA gene; they have prematurely shortened telomeres and typically die of progressive bone marrow failure. They also develop lung scarring and liver cirrhosis and show abnormalities in various epidermal structures including skin, hair follicles, and nails.
Cell_Biology_Alberts. with the genetic disease dyskeratosis congenita. Individuals afflicted with this disease carry one functional and one nonfunctional copy of the telomerase RNA gene; they have prematurely shortened telomeres and typically die of progressive bone marrow failure. They also develop lung scarring and liver cirrhosis and show abnormalities in various epidermal structures including skin, hair follicles, and nails.
Cell_Biology_Alberts_1237
Cell_Biology_Alberts
The above observations demonstrate that controlling cell proliferation by telomere shortening poses a risk to an organism, because not all of the cells that begin losing the ends of their chromosomes will stop dividing. Some apparently become genetically unstable, but continue to divide, giving rise to variant cells that can lead to cancer. Clearly, the use of telomere shortening as a regulating mechanism is not foolproof and, like many mechanisms in the cell, seems to strike a balance between benefit and risk. The proteins that initiate DNA replication bind to DNA sequences at a replication origin to catalyze the formation of a replication bubble with two outward-moving replication forks. The process begins when an initiator protein–DNA complex is formed that subsequently loads a DNA helicase onto the DNA template. Other proteins are then added to form the multienzyme “replication machine” that catalyzes DNA synthesis at each replication fork.
Cell_Biology_Alberts. The above observations demonstrate that controlling cell proliferation by telomere shortening poses a risk to an organism, because not all of the cells that begin losing the ends of their chromosomes will stop dividing. Some apparently become genetically unstable, but continue to divide, giving rise to variant cells that can lead to cancer. Clearly, the use of telomere shortening as a regulating mechanism is not foolproof and, like many mechanisms in the cell, seems to strike a balance between benefit and risk. The proteins that initiate DNA replication bind to DNA sequences at a replication origin to catalyze the formation of a replication bubble with two outward-moving replication forks. The process begins when an initiator protein–DNA complex is formed that subsequently loads a DNA helicase onto the DNA template. Other proteins are then added to form the multienzyme “replication machine” that catalyzes DNA synthesis at each replication fork.
Cell_Biology_Alberts_1238
Cell_Biology_Alberts
In bacteria and some simple eukaryotes, replication origins are specified by specific DNA sequences that are only several hundred nucleotide pairs long. In other eukaryotes, such as humans, the sequences needed to specify an origin of DNA replication seem to be less well defined, and the origin can span several thousand nucleotide pairs.
Cell_Biology_Alberts. In bacteria and some simple eukaryotes, replication origins are specified by specific DNA sequences that are only several hundred nucleotide pairs long. In other eukaryotes, such as humans, the sequences needed to specify an origin of DNA replication seem to be less well defined, and the origin can span several thousand nucleotide pairs.
Cell_Biology_Alberts_1239
Cell_Biology_Alberts
Bacteria typically have a single origin of replication in a circular chromosome. With fork speeds of up to 1000 nucleotides per second, they can replicate their genome in less than an hour. Eukaryotic DNA replication takes place in only one part of the cell cycle, the S phase. The replication fork in eukaryotes moves about 10 times more slowly than the bacterial replication fork, and the much longer eukaryotic chromosomes each require many replication origins to complete their replication in an S phase, which typically lasts for 8 hours in human cells. The different replication origins in these eukaryotic chromosomes are activated in a sequence, determined in part by the structure of the chromatin, with the most condensed regions of chromatin typically beginning their replication last. After the replication fork has passed, chromatin structure is re-formed by the addition of new histones to the old histones that are directly inherited by each daughter DNA molecule.
Cell_Biology_Alberts. Bacteria typically have a single origin of replication in a circular chromosome. With fork speeds of up to 1000 nucleotides per second, they can replicate their genome in less than an hour. Eukaryotic DNA replication takes place in only one part of the cell cycle, the S phase. The replication fork in eukaryotes moves about 10 times more slowly than the bacterial replication fork, and the much longer eukaryotic chromosomes each require many replication origins to complete their replication in an S phase, which typically lasts for 8 hours in human cells. The different replication origins in these eukaryotic chromosomes are activated in a sequence, determined in part by the structure of the chromatin, with the most condensed regions of chromatin typically beginning their replication last. After the replication fork has passed, chromatin structure is re-formed by the addition of new histones to the old histones that are directly inherited by each daughter DNA molecule.
Cell_Biology_Alberts_1240
Cell_Biology_Alberts
Eukaryotes solve the problem of replicating the ends of their linear chromosomes with a specialized end structure, the telomere, maintained by a special nucleotide polymerizing enzyme called telomerase. Telomerase extends one of the DNA strands at the end of a chromosome by using an RNA template that is an integral part of the enzyme itself, producing a highly repeated DNA sequence that typically extends for thousands of nucleotide pairs at each chromosome end. Telomeres have specialized structures that distinguish them from broken ends of chromosomes, ensuring that they are not mistakenly repaired.
Cell_Biology_Alberts. Eukaryotes solve the problem of replicating the ends of their linear chromosomes with a specialized end structure, the telomere, maintained by a special nucleotide polymerizing enzyme called telomerase. Telomerase extends one of the DNA strands at the end of a chromosome by using an RNA template that is an integral part of the enzyme itself, producing a highly repeated DNA sequence that typically extends for thousands of nucleotide pairs at each chromosome end. Telomeres have specialized structures that distinguish them from broken ends of chromosomes, ensuring that they are not mistakenly repaired.
Cell_Biology_Alberts_1241
Cell_Biology_Alberts
Maintaining the genetic stability that an organism needs for its survival requires not only an extremely accurate mechanism for replicating DNA, but also mechanisms for repairing the many accidental lesions that DNA continually suffers. Most such spontaneous changes in DNA are temporary because they are immediately corrected by a set of processes that are collectively called DNA repair. Of the tens of thousands of random changes created every day in the DNA of a human cell by heat, metabolic accidents, radiation of various sorts, and exposure to substances in the environment, only a few (less than 0.02%) accumulate as permanent mutations in the DNA sequence. The rest are eliminated with remarkable efficiency by DNA repair.
Cell_Biology_Alberts. Maintaining the genetic stability that an organism needs for its survival requires not only an extremely accurate mechanism for replicating DNA, but also mechanisms for repairing the many accidental lesions that DNA continually suffers. Most such spontaneous changes in DNA are temporary because they are immediately corrected by a set of processes that are collectively called DNA repair. Of the tens of thousands of random changes created every day in the DNA of a human cell by heat, metabolic accidents, radiation of various sorts, and exposure to substances in the environment, only a few (less than 0.02%) accumulate as permanent mutations in the DNA sequence. The rest are eliminated with remarkable efficiency by DNA repair.
Cell_Biology_Alberts_1242
Cell_Biology_Alberts
The importance of DNA repair is evident from the large investment that cells make in the enzymes that carry it out: several percent of the coding capacity of most genomes is devoted solely to DNA repair functions. The importance of DNA repair is also demonstrated by the increased rate of mutation that follows the inactivation of a DNA repair gene. Many DNA repair proteins and the genes that encode them—which we now know operate in a wide range of organisms, including humans—were originally identified in bacteria by the isolation and characterization of mutants that displayed an increased mutation rate or an increased sensitivity to DNA-damaging agents.
Cell_Biology_Alberts. The importance of DNA repair is evident from the large investment that cells make in the enzymes that carry it out: several percent of the coding capacity of most genomes is devoted solely to DNA repair functions. The importance of DNA repair is also demonstrated by the increased rate of mutation that follows the inactivation of a DNA repair gene. Many DNA repair proteins and the genes that encode them—which we now know operate in a wide range of organisms, including humans—were originally identified in bacteria by the isolation and characterization of mutants that displayed an increased mutation rate or an increased sensitivity to DNA-damaging agents.
Cell_Biology_Alberts_1243
Cell_Biology_Alberts
Recent studies of the consequences of a diminished capacity for DNA repair in humans have linked many human diseases with decreased repair (Table 5–2). Thus, we saw previously that defects in a human gene whose product normally functions to repair the mismatched base pairs resulting from DNA replication errors can lead to an inherited predisposition to cancers of the colon and some other organs, reflecting an increased mutation rate. In another human disease, xeroderma pigmentosum (XP), the afflicted individuals have an extreme sensitivity to ultraviolet radiation because they are unable to repair certain DNA photo-products. This repair defect results in an increased mutation rate that leads to serious skin lesions and an increased susceptibility to skin cancers. Finally, mutations in the Brca1 and Brca2 genes compromise a type of DNA repair known as homologous recombination and are a cause of hereditary breast and ovarian cancer.
Cell_Biology_Alberts. Recent studies of the consequences of a diminished capacity for DNA repair in humans have linked many human diseases with decreased repair (Table 5–2). Thus, we saw previously that defects in a human gene whose product normally functions to repair the mismatched base pairs resulting from DNA replication errors can lead to an inherited predisposition to cancers of the colon and some other organs, reflecting an increased mutation rate. In another human disease, xeroderma pigmentosum (XP), the afflicted individuals have an extreme sensitivity to ultraviolet radiation because they are unable to repair certain DNA photo-products. This repair defect results in an increased mutation rate that leads to serious skin lesions and an increased susceptibility to skin cancers. Finally, mutations in the Brca1 and Brca2 genes compromise a type of DNA repair known as homologous recombination and are a cause of hereditary breast and ovarian cancer.
Cell_Biology_Alberts_1244
Cell_Biology_Alberts
Without DNA Repair, Spontaneous DNA Damage Would Rapidly Change DNA Sequences Although DNA is a highly stable material—as required for the storage of genetic information—it is a complex organic molecule that is susceptible, even under normal cell conditions, to spontaneous changes that would lead to mutations if left unrepaired (Figure 5–37 and see Table 5–3). For example, the DNA of each Figure 5–37 A summary of spontaneous alterations that require DNA repair.
Cell_Biology_Alberts. Without DNA Repair, Spontaneous DNA Damage Would Rapidly Change DNA Sequences Although DNA is a highly stable material—as required for the storage of genetic information—it is a complex organic molecule that is susceptible, even under normal cell conditions, to spontaneous changes that would lead to mutations if left unrepaired (Figure 5–37 and see Table 5–3). For example, the DNA of each Figure 5–37 A summary of spontaneous alterations that require DNA repair.
Cell_Biology_Alberts_1245
Cell_Biology_Alberts
The sites on each nucleotide modified by spontaneous oxidative damage (red arrows), hydrolytic attack (blue arrows), and methylation (green arrows) are shown, with the width of each arrow indicating the relative frequency of each event (see Table 5–3). (After T. Lindahl, Nature 362:709–715, 1993. With permission from Macmillan Publishers Ltd.) human cell loses about 18,000 purine bases (adenine and guanine) every day because their N-glycosyl linkages to deoxyribose hydrolyze, a spontaneous reaction called depurination. Similarly, a spontaneous deamination of cytosine to uracil in DNA occurs at a rate of about 100 bases per cell per day (Figure 5–38). DNA bases are also occasionally damaged by an encounter with reactive metabolites produced in the cell, including reactive forms of oxygen and the high-energy methyl donor S-adenosylmethionine, or by exposure to chemicals in the environment. Likewise, ultraviolet radiation from the sun can produce a covalent linkage between two adjacent
Cell_Biology_Alberts. The sites on each nucleotide modified by spontaneous oxidative damage (red arrows), hydrolytic attack (blue arrows), and methylation (green arrows) are shown, with the width of each arrow indicating the relative frequency of each event (see Table 5–3). (After T. Lindahl, Nature 362:709–715, 1993. With permission from Macmillan Publishers Ltd.) human cell loses about 18,000 purine bases (adenine and guanine) every day because their N-glycosyl linkages to deoxyribose hydrolyze, a spontaneous reaction called depurination. Similarly, a spontaneous deamination of cytosine to uracil in DNA occurs at a rate of about 100 bases per cell per day (Figure 5–38). DNA bases are also occasionally damaged by an encounter with reactive metabolites produced in the cell, including reactive forms of oxygen and the high-energy methyl donor S-adenosylmethionine, or by exposure to chemicals in the environment. Likewise, ultraviolet radiation from the sun can produce a covalent linkage between two adjacent
Cell_Biology_Alberts_1246
Cell_Biology_Alberts
and the high-energy methyl donor S-adenosylmethionine, or by exposure to chemicals in the environment. Likewise, ultraviolet radiation from the sun can produce a covalent linkage between two adjacent pyrimidine bases in DNA to form, for example, thymine dimers (Figure 5–39). If left uncorrected when the DNA is replicated, most of these changes would be expected to lead either to the deletion of one or more base pairs or to a base-pair substitution in the daughter DNA chain (Figure 5–40). The mutations would then be propagated throughout subsequent cell generations. Such a high rate of random changes in the DNA sequence would have disastrous consequences.
Cell_Biology_Alberts. and the high-energy methyl donor S-adenosylmethionine, or by exposure to chemicals in the environment. Likewise, ultraviolet radiation from the sun can produce a covalent linkage between two adjacent pyrimidine bases in DNA to form, for example, thymine dimers (Figure 5–39). If left uncorrected when the DNA is replicated, most of these changes would be expected to lead either to the deletion of one or more base pairs or to a base-pair substitution in the daughter DNA chain (Figure 5–40). The mutations would then be propagated throughout subsequent cell generations. Such a high rate of random changes in the DNA sequence would have disastrous consequences.
Cell_Biology_Alberts_1247
Cell_Biology_Alberts
The DNA Double Helix Is Readily Repaired The double-helical structure of DNA is ideally suited for repair because it carries two separate copies of all the genetic information—one in each of its two strands. Thus, when one strand is damaged, the complementary strand retains an intact copy of the same information, and this copy is generally used to restore the correct nucleotide sequences to the damaged strand.
Cell_Biology_Alberts. The DNA Double Helix Is Readily Repaired The double-helical structure of DNA is ideally suited for repair because it carries two separate copies of all the genetic information—one in each of its two strands. Thus, when one strand is damaged, the complementary strand retains an intact copy of the same information, and this copy is generally used to restore the correct nucleotide sequences to the damaged strand.
Cell_Biology_Alberts_1248
Cell_Biology_Alberts
An indication of the importance of a double-strand helix to the safe storage of genetic information is that all cells use it; only a few small viruses use single-strand DNA or RNA as their genetic material. The types of repair processes described in this section cannot operate on such nucleic acids, and once damaged, the chance of a permanent nucleotide change occurring in these single-strand genomes of viruses is thus very high. It seems that only organisms with tiny genomes (and therefore tiny targets for DNA damage) can afford to encode their genetic information in any molecule other than a DNA double helix.
Cell_Biology_Alberts. An indication of the importance of a double-strand helix to the safe storage of genetic information is that all cells use it; only a few small viruses use single-strand DNA or RNA as their genetic material. The types of repair processes described in this section cannot operate on such nucleic acids, and once damaged, the chance of a permanent nucleotide change occurring in these single-strand genomes of viruses is thus very high. It seems that only organisms with tiny genomes (and therefore tiny targets for DNA damage) can afford to encode their genetic information in any molecule other than a DNA double helix.
Cell_Biology_Alberts_1249
Cell_Biology_Alberts
Figure 5–38 Depurination and deamination. These reactions are two of the most frequent spontaneous chemical reactions that create serious DNA damage in cells. Depurination can release guanine (shown here), as well as adenine, from DNA. The major type of deamination reaction converts cytosine to an altered DNA base, uracil (shown here), but deamination occurs on other bases as well. These reactions normally take place in double-helical DNA; for convenience, only one strand is shown. Figure 5–39 The most common type of thymine dimer. This type of damage occurs in the DNA of cells exposed to ultraviolet irradiation (as in sunlight). A similar dimer will form between any two neighboring pyrimidine bases (C or T residues) in DNA.
Cell_Biology_Alberts. Figure 5–38 Depurination and deamination. These reactions are two of the most frequent spontaneous chemical reactions that create serious DNA damage in cells. Depurination can release guanine (shown here), as well as adenine, from DNA. The major type of deamination reaction converts cytosine to an altered DNA base, uracil (shown here), but deamination occurs on other bases as well. These reactions normally take place in double-helical DNA; for convenience, only one strand is shown. Figure 5–39 The most common type of thymine dimer. This type of damage occurs in the DNA of cells exposed to ultraviolet irradiation (as in sunlight). A similar dimer will form between any two neighboring pyrimidine bases (C or T residues) in DNA.
Cell_Biology_Alberts_1250
Cell_Biology_Alberts
Cells have multiple pathways to repair their DNA using different enzymes that act upon different kinds of lesions. Figure 5–41 shows two of the most common pathways. In both, the damage is excised, the original DNA sequence is restored by a DNA polymerase that uses the undamaged strand as its template, and a remaining break in the double helix is sealed by DNA ligase (see Figure 5–12).
Cell_Biology_Alberts. Cells have multiple pathways to repair their DNA using different enzymes that act upon different kinds of lesions. Figure 5–41 shows two of the most common pathways. In both, the damage is excised, the original DNA sequence is restored by a DNA polymerase that uses the undamaged strand as its template, and a remaining break in the double helix is sealed by DNA ligase (see Figure 5–12).
Cell_Biology_Alberts_1251
Cell_Biology_Alberts
The two pathways differ in the way in which they remove the damage from DNA. The first pathway, called base excision repair, involves a battery of enzymes called DNA glycosylases, each of which can recognize a specific type of altered base in DNA and catalyze its hydrolytic removal. There are at least six types of these enzymes, including those that remove deaminated Cs, deaminated As, different types of alkylated or oxidized bases, bases with opened rings, and bases in which a carbon–carbon double bond has been accidentally converted to a carbon–carbon single bond. How is an altered base detected within the context of the double helix? A key step is an enzyme-mediated “flipping-out” of the altered nucleotide from the helix, which allows the DNA glycosylase to probe all faces of the base for damage (Figure 5–42). It is thought that these enzymes travel along DNA using base-flipping to evaluate the status of each base. Once an enzyme finds the damaged base that it recognizes, it
Cell_Biology_Alberts. The two pathways differ in the way in which they remove the damage from DNA. The first pathway, called base excision repair, involves a battery of enzymes called DNA glycosylases, each of which can recognize a specific type of altered base in DNA and catalyze its hydrolytic removal. There are at least six types of these enzymes, including those that remove deaminated Cs, deaminated As, different types of alkylated or oxidized bases, bases with opened rings, and bases in which a carbon–carbon double bond has been accidentally converted to a carbon–carbon single bond. How is an altered base detected within the context of the double helix? A key step is an enzyme-mediated “flipping-out” of the altered nucleotide from the helix, which allows the DNA glycosylase to probe all faces of the base for damage (Figure 5–42). It is thought that these enzymes travel along DNA using base-flipping to evaluate the status of each base. Once an enzyme finds the damaged base that it recognizes, it
Cell_Biology_Alberts_1252
Cell_Biology_Alberts
base for damage (Figure 5–42). It is thought that these enzymes travel along DNA using base-flipping to evaluate the status of each base. Once an enzyme finds the damaged base that it recognizes, it removes that base from its sugar.
Cell_Biology_Alberts. base for damage (Figure 5–42). It is thought that these enzymes travel along DNA using base-flipping to evaluate the status of each base. Once an enzyme finds the damaged base that it recognizes, it removes that base from its sugar.
Cell_Biology_Alberts_1253
Cell_Biology_Alberts
The “missing tooth” created by DNA glycosylase action is recognized by an enzyme called AP endonuclease (AP for apurinic or apyrimidinic, endo to signify that the nuclease cleaves within the polynucleotide chain), which cuts the phosphodiester backbone, after which the resulting gap is repaired (see Figure 5–41A). Depurination, which is by far the most frequent type of damage suffered by DNA, also leaves a deoxyribose sugar with a missing base. Depurinations are directly repaired beginning with AP endonuclease, following the bottom half of the pathway in Figure 5–41A. a G has been changed to an A
Cell_Biology_Alberts. The “missing tooth” created by DNA glycosylase action is recognized by an enzyme called AP endonuclease (AP for apurinic or apyrimidinic, endo to signify that the nuclease cleaves within the polynucleotide chain), which cuts the phosphodiester backbone, after which the resulting gap is repaired (see Figure 5–41A). Depurination, which is by far the most frequent type of damage suffered by DNA, also leaves a deoxyribose sugar with a missing base. Depurinations are directly repaired beginning with AP endonuclease, following the bottom half of the pathway in Figure 5–41A. a G has been changed to an A
Cell_Biology_Alberts_1254
Cell_Biology_Alberts
a G has been changed to an A Figure 5–40 How chemical modifications of nucleotides produce mutations. (A) Deamination of cytosine, if uncorrected, results in the substitution of one base for another when the DNA is replicated. As shown in Figure 5–38, deamination of cytosine produces uracil. Uracil differs from cytosine in its base-pairing properties and preferentially base-pairs with adenine. The DNA replication machinery therefore adds an adenine when it encounters a uracil on the template strand. (B) Depurination can lead to the loss of a nucleotide pair. When the replication machinery encounters a missing purine on the template strand, it may skip to the next complete nucleotide as illustrated here, thus producing a nucleotide deletion in the newly synthesized strand. Many other types of DNA damage (see Figure 5–37), if left uncorrected, also produce mutations when the DNA is replicated. G C T A T C C DNA helix with missing base C G A G T A G G
Cell_Biology_Alberts. a G has been changed to an A Figure 5–40 How chemical modifications of nucleotides produce mutations. (A) Deamination of cytosine, if uncorrected, results in the substitution of one base for another when the DNA is replicated. As shown in Figure 5–38, deamination of cytosine produces uracil. Uracil differs from cytosine in its base-pairing properties and preferentially base-pairs with adenine. The DNA replication machinery therefore adds an adenine when it encounters a uracil on the template strand. (B) Depurination can lead to the loss of a nucleotide pair. When the replication machinery encounters a missing purine on the template strand, it may skip to the next complete nucleotide as illustrated here, thus producing a nucleotide deletion in the newly synthesized strand. Many other types of DNA damage (see Figure 5–37), if left uncorrected, also produce mutations when the DNA is replicated. G C T A T C C DNA helix with missing base C G A G T A G G
Cell_Biology_Alberts_1255
Cell_Biology_Alberts
G C T A T C C DNA helix with missing base C G A G T A G G DNA helix with single-nucleotide gap C G A G T A G G DNA POLYMERASE ADDS NEW NUCLEOTIDE, DNA LIGASE SEALS NICK G C T C A T C C DNA helix with 12nucleotide gap
Cell_Biology_Alberts. G C T A T C C DNA helix with missing base C G A G T A G G DNA helix with single-nucleotide gap C G A G T A G G DNA POLYMERASE ADDS NEW NUCLEOTIDE, DNA LIGASE SEALS NICK G C T C A T C C DNA helix with 12nucleotide gap
Cell_Biology_Alberts_1256
Cell_Biology_Alberts
Figure 5– 41 A comparison of two major DNA repair pathways. (A) Base excision repair. This pathway starts with a DNA glycosylase. Here, the enzyme uracil DNA glycosylase removes an accidentally deaminated cytosine in DNA. After the action of this glycosylase (or another DNA glycosylase that recognizes a different kind of damage), the sugar phosphate with the missing base is cut out by the sequential action of AP endonuclease and a phosphodiesterase. (These same enzymes begin the repair of depurinated sites directly.) The gap of a single nucleotide is then filled by DNA polymerase and DNA ligase. The net result is that the U that was created by accidental deamination is restored to a C. AP endonuclease is so-named because it recognizes any site in the DNA helix that contains a deoxyribose sugar with a missing base; such sites can arise either by the loss of a purine (apurinic sites) or by the loss of a pyrimidine (apyrimidinic sites). (B) Nucleotide excision repair. In bacteria, after
Cell_Biology_Alberts. Figure 5– 41 A comparison of two major DNA repair pathways. (A) Base excision repair. This pathway starts with a DNA glycosylase. Here, the enzyme uracil DNA glycosylase removes an accidentally deaminated cytosine in DNA. After the action of this glycosylase (or another DNA glycosylase that recognizes a different kind of damage), the sugar phosphate with the missing base is cut out by the sequential action of AP endonuclease and a phosphodiesterase. (These same enzymes begin the repair of depurinated sites directly.) The gap of a single nucleotide is then filled by DNA polymerase and DNA ligase. The net result is that the U that was created by accidental deamination is restored to a C. AP endonuclease is so-named because it recognizes any site in the DNA helix that contains a deoxyribose sugar with a missing base; such sites can arise either by the loss of a purine (apurinic sites) or by the loss of a pyrimidine (apyrimidinic sites). (B) Nucleotide excision repair. In bacteria, after
Cell_Biology_Alberts_1257
Cell_Biology_Alberts
sugar with a missing base; such sites can arise either by the loss of a purine (apurinic sites) or by the loss of a pyrimidine (apyrimidinic sites). (B) Nucleotide excision repair. In bacteria, after a multienzyme complex has recognized a lesion such as a pyrimidine dimer (see Figure 5–39), one cut is made on each side of the lesion, and an associated DNA helicase then removes the entire portion of the damaged strand. The excision repair machinery in bacteria leaves the gap of 12 nucleotides shown. In humans, once the damaged DNA is recognized, a helicase is recruited to unwind the DNA duplex locally. Next, the excision nuclease enters and cleaves on either side of the damage, leaving a gap of about 30 nucleotides. The nucleotide excision repair machinery in both bacteria and humans can recognize and repair many different types of DNA damage.
Cell_Biology_Alberts. sugar with a missing base; such sites can arise either by the loss of a purine (apurinic sites) or by the loss of a pyrimidine (apyrimidinic sites). (B) Nucleotide excision repair. In bacteria, after a multienzyme complex has recognized a lesion such as a pyrimidine dimer (see Figure 5–39), one cut is made on each side of the lesion, and an associated DNA helicase then removes the entire portion of the damaged strand. The excision repair machinery in bacteria leaves the gap of 12 nucleotides shown. In humans, once the damaged DNA is recognized, a helicase is recruited to unwind the DNA duplex locally. Next, the excision nuclease enters and cleaves on either side of the damage, leaving a gap of about 30 nucleotides. The nucleotide excision repair machinery in both bacteria and humans can recognize and repair many different types of DNA damage.
Cell_Biology_Alberts_1258
Cell_Biology_Alberts
The second major repair pathway is called nucleotide excision repair. This mechanism can repair the damage caused by almost any large change in the structure of the DNA double helix. Such “bulky lesions” include those created by the covalent reaction of DNA bases with large hydrocarbons (such as the carcinogen benzopyrene, found in tobacco smoke, coal tar, and diesel exhaust), as well as the various pyrimidine dimers (T-T, T-C, and C-C) caused by sunlight. In this pathway, a large multienzyme complex scans the DNA for a distortion in the double helix, rather than for a specific base change. Once it finds a lesion, it cleaves the phosphodiester backbone of the abnormal strand on both sides of the distortion, and a DNA helicase peels away the single-strand oligonucleotide containing the lesion. The large gap produced in the DNA helix is then repaired by DNA polymerase and DNA ligase (see Figure 5–41B).
Cell_Biology_Alberts. The second major repair pathway is called nucleotide excision repair. This mechanism can repair the damage caused by almost any large change in the structure of the DNA double helix. Such “bulky lesions” include those created by the covalent reaction of DNA bases with large hydrocarbons (such as the carcinogen benzopyrene, found in tobacco smoke, coal tar, and diesel exhaust), as well as the various pyrimidine dimers (T-T, T-C, and C-C) caused by sunlight. In this pathway, a large multienzyme complex scans the DNA for a distortion in the double helix, rather than for a specific base change. Once it finds a lesion, it cleaves the phosphodiester backbone of the abnormal strand on both sides of the distortion, and a DNA helicase peels away the single-strand oligonucleotide containing the lesion. The large gap produced in the DNA helix is then repaired by DNA polymerase and DNA ligase (see Figure 5–41B).
Cell_Biology_Alberts_1259
Cell_Biology_Alberts
An alternative to base and nucleotide excision repair processes is direct chemical reversal of DNA damage, and this strategy is selectively employed for the rapid removal of certain highly mutagenic or cytotoxic lesions. For example, the alkylation lesion O6-methylguanine has its methyl group removed by direct transfer to a cysteine residue in the repair protein itself, which is destroyed in the reaction. In another example, methyl groups in the alkylation lesions 1-methyladenine and 3-methylcytosine are “burnt off” by an iron-dependent demethylase, with release of formaldehyde from the methylated DNA and regeneration of the native base. Coupling Nucleotide Excision Repair to Transcription Ensures That the Cell’s Most Important DNA Is Efficiently Repaired
Cell_Biology_Alberts. An alternative to base and nucleotide excision repair processes is direct chemical reversal of DNA damage, and this strategy is selectively employed for the rapid removal of certain highly mutagenic or cytotoxic lesions. For example, the alkylation lesion O6-methylguanine has its methyl group removed by direct transfer to a cysteine residue in the repair protein itself, which is destroyed in the reaction. In another example, methyl groups in the alkylation lesions 1-methyladenine and 3-methylcytosine are “burnt off” by an iron-dependent demethylase, with release of formaldehyde from the methylated DNA and regeneration of the native base. Coupling Nucleotide Excision Repair to Transcription Ensures That the Cell’s Most Important DNA Is Efficiently Repaired
Cell_Biology_Alberts_1260
Cell_Biology_Alberts
All of a cell’s DNA is under constant surveillance for damage, and the repair mechanisms we have described act on all parts of the genome. However, cells have a way of directing DNA repair to the DNA sequences that are most urgently needed. They do this by linking RNA polymerase, the enzyme that transcribes DNA into RNA as the first step in gene expression, to the nucleotide excision repair pathway. As discussed above, this repair system can correct many different types of DNA damage. RNA polymerase stalls at DNA lesions and, through the use of coupling proteins, directs the excision repair machinery to these sites. In bacteria, where genes are relatively short, the stalled RNA polymerase can be dissociated from the DNA; the DNA is repaired, and the gene is transcribed again from the beginning. In eukaryotes, where genes can be enormously long, a more complex reaction is used to “back up” the RNA polymerase, repair the damage, and then restart the polymerase.
Cell_Biology_Alberts. All of a cell’s DNA is under constant surveillance for damage, and the repair mechanisms we have described act on all parts of the genome. However, cells have a way of directing DNA repair to the DNA sequences that are most urgently needed. They do this by linking RNA polymerase, the enzyme that transcribes DNA into RNA as the first step in gene expression, to the nucleotide excision repair pathway. As discussed above, this repair system can correct many different types of DNA damage. RNA polymerase stalls at DNA lesions and, through the use of coupling proteins, directs the excision repair machinery to these sites. In bacteria, where genes are relatively short, the stalled RNA polymerase can be dissociated from the DNA; the DNA is repaired, and the gene is transcribed again from the beginning. In eukaryotes, where genes can be enormously long, a more complex reaction is used to “back up” the RNA polymerase, repair the damage, and then restart the polymerase.
Cell_Biology_Alberts_1261
Cell_Biology_Alberts
The importance of transcription-coupled excision repair is seen in people with Cockayne syndrome, which is caused by a defect in this coupling. These individuals suffer from growth retardation, skeletal abnormalities, progressive neural retardation, and severe sensitivity to sunlight. Most of these problems are thought to arise from RNA polymerase molecules that become permanently stalled at sites of DNA damage that lie in important genes. The Chemistry of the DNA Bases Facilitates Damage Detection
Cell_Biology_Alberts. The importance of transcription-coupled excision repair is seen in people with Cockayne syndrome, which is caused by a defect in this coupling. These individuals suffer from growth retardation, skeletal abnormalities, progressive neural retardation, and severe sensitivity to sunlight. Most of these problems are thought to arise from RNA polymerase molecules that become permanently stalled at sites of DNA damage that lie in important genes. The Chemistry of the DNA Bases Facilitates Damage Detection
Cell_Biology_Alberts_1262
Cell_Biology_Alberts
The Chemistry of the DNA Bases Facilitates Damage Detection The DNA double helix seems optimal for repair. As noted above, it contains a backup copy of all genetic information. Equally importantly, the nature of the four bases in DNA makes the distinction between undamaged and damaged bases very clear. For example, every possible deamination event in DNA yields an “unnatural” base, which can be directly recognized and removed by a specific DNA glycosylase. Hypoxanthine, for example, is the simplest purine base capable of pairing specifically with C, but hypoxanthine is the direct deamination product of A (Figure 5–43A). The addition of a second amino group to hypoxanthine
Cell_Biology_Alberts. The Chemistry of the DNA Bases Facilitates Damage Detection The DNA double helix seems optimal for repair. As noted above, it contains a backup copy of all genetic information. Equally importantly, the nature of the four bases in DNA makes the distinction between undamaged and damaged bases very clear. For example, every possible deamination event in DNA yields an “unnatural” base, which can be directly recognized and removed by a specific DNA glycosylase. Hypoxanthine, for example, is the simplest purine base capable of pairing specifically with C, but hypoxanthine is the direct deamination product of A (Figure 5–43A). The addition of a second amino group to hypoxanthine
Cell_Biology_Alberts_1263
Cell_Biology_Alberts
Figure 5–42 The recognition of an unusual nucleotide in DNA by base-flipping. The DNA glycosylase family of enzymes recognizes specific inappropriate bases in the conformation shown. Each of these enzymes cleaves the glycosyl bond that connects a particular recognized base (yellow) to the backbone sugar, removing it from the DNA. (A) Stick model; (B) space-filling model.
Cell_Biology_Alberts. Figure 5–42 The recognition of an unusual nucleotide in DNA by base-flipping. The DNA glycosylase family of enzymes recognizes specific inappropriate bases in the conformation shown. Each of these enzymes cleaves the glycosyl bond that connects a particular recognized base (yellow) to the backbone sugar, removing it from the DNA. (A) Stick model; (B) space-filling model.
Cell_Biology_Alberts_1264
Cell_Biology_Alberts
Figure 5–43 The deamination of DNA nucleotides. In each case, the oxygen atom that is added in this reaction with water is colored red. (A) The spontaneous deamination products of A and G are recognizable as unnatural when they occur in DNA and thus are readily found and repaired. The deamination of C to U was also illustrated in Figure 5–38; T has no amino group to remove. (B) About 3% of the C nucleotides in vertebrate DNAs are methylated to help in controlling gene expression (discussed in Chapter 7). When these 5-methyl C nucleotides are accidentally deaminated, they form the natural nucleotide T. However, this T will be paired with a G on the opposite strand, forming a mismatched base pair. produces G, which cannot be formed from A by spontaneous deamination, and whose deamination product (xanthine) is likewise unique.
Cell_Biology_Alberts. Figure 5–43 The deamination of DNA nucleotides. In each case, the oxygen atom that is added in this reaction with water is colored red. (A) The spontaneous deamination products of A and G are recognizable as unnatural when they occur in DNA and thus are readily found and repaired. The deamination of C to U was also illustrated in Figure 5–38; T has no amino group to remove. (B) About 3% of the C nucleotides in vertebrate DNAs are methylated to help in controlling gene expression (discussed in Chapter 7). When these 5-methyl C nucleotides are accidentally deaminated, they form the natural nucleotide T. However, this T will be paired with a G on the opposite strand, forming a mismatched base pair. produces G, which cannot be formed from A by spontaneous deamination, and whose deamination product (xanthine) is likewise unique.
Cell_Biology_Alberts_1265
Cell_Biology_Alberts
produces G, which cannot be formed from A by spontaneous deamination, and whose deamination product (xanthine) is likewise unique. As discussed in Chapter 6, RNA is thought, on an evolutionary time scale, to have served as the genetic material before DNA, and it seems likely that the genetic code was initially carried in the four nucleotides A, C, G, and U. This raises the question of why the U in RNA was replaced in DNA by T (which is 5-methyl U). We have seen that the spontaneous deamination of C converts it to U, but that this event is rendered relatively harmless by uracil DNA glycosylase. However, if DNA contained U as a natural base, the repair system would not be able to distinguish a deaminated C from a naturally occurring U.
Cell_Biology_Alberts. produces G, which cannot be formed from A by spontaneous deamination, and whose deamination product (xanthine) is likewise unique. As discussed in Chapter 6, RNA is thought, on an evolutionary time scale, to have served as the genetic material before DNA, and it seems likely that the genetic code was initially carried in the four nucleotides A, C, G, and U. This raises the question of why the U in RNA was replaced in DNA by T (which is 5-methyl U). We have seen that the spontaneous deamination of C converts it to U, but that this event is rendered relatively harmless by uracil DNA glycosylase. However, if DNA contained U as a natural base, the repair system would not be able to distinguish a deaminated C from a naturally occurring U.
Cell_Biology_Alberts_1266
Cell_Biology_Alberts
A special situation occurs in vertebrate DNA, in which selected C nucleotides are methylated at specific CG sequences that are associated with inactive genes (discussed in Chapter 7). The accidental deamination of these methylated C nucleotides produces the natural nucleotide T (Figure 5–43B) in a mismatched base pair with a G on the opposite DNA strand. To help in repairing deaminated methylated C nucleotides, a special DNA glycosylase recognizes a mismatched base pair involving T in the sequence T-G and removes the T. This DNA repair mechanism must be relatively ineffective, however, because methylated C nucleotides are exceptionally common sites for mutations in vertebrate DNA. It is striking that, even though only about 3% of the C nucleotides in human DNA are methylated, mutations in these methylated nucleotides account for about one-third of the single-base mutations that have been observed in inherited human diseases.
Cell_Biology_Alberts. A special situation occurs in vertebrate DNA, in which selected C nucleotides are methylated at specific CG sequences that are associated with inactive genes (discussed in Chapter 7). The accidental deamination of these methylated C nucleotides produces the natural nucleotide T (Figure 5–43B) in a mismatched base pair with a G on the opposite DNA strand. To help in repairing deaminated methylated C nucleotides, a special DNA glycosylase recognizes a mismatched base pair involving T in the sequence T-G and removes the T. This DNA repair mechanism must be relatively ineffective, however, because methylated C nucleotides are exceptionally common sites for mutations in vertebrate DNA. It is striking that, even though only about 3% of the C nucleotides in human DNA are methylated, mutations in these methylated nucleotides account for about one-third of the single-base mutations that have been observed in inherited human diseases.
Cell_Biology_Alberts_1267
Cell_Biology_Alberts
If a cell’s DNA suffers heavy damage, the repair mechanisms that we have discussed are often insufficient to cope with it. In these cases, a different strategy is called into play, one that entails some risk to the cell. The highly accurate replicative DNA polymerases stall when they encounter damaged DNA, and in emergencies cells employ versatile, but less accurate, backup polymerases, known as translesion polymerases, to replicate through the DNA damage.
Cell_Biology_Alberts. If a cell’s DNA suffers heavy damage, the repair mechanisms that we have discussed are often insufficient to cope with it. In these cases, a different strategy is called into play, one that entails some risk to the cell. The highly accurate replicative DNA polymerases stall when they encounter damaged DNA, and in emergencies cells employ versatile, but less accurate, backup polymerases, known as translesion polymerases, to replicate through the DNA damage.
Cell_Biology_Alberts_1268
Cell_Biology_Alberts
Human cells have seven translesion polymerases, some of which can recognize a specific type of DNA damage and correctly add the nucleotide required to restore the initial sequence. Others make only “good guesses,” especially when the template base has been extensively damaged. These enzymes are not as accurate as the normal replicative polymerases when they copy a normal DNA sequence. For one thing, the translesion polymerases lack exonucleolytic proofreading activity; in addition, many are much less discriminating than the replicative polymerase in choosing which nucleotide to incorporate initially. Presumably for this reason, each such translesion polymerase is given a chance to add only one or a few nucleotides before the highly accurate replicative polymerase resumes DNA synthesis.
Cell_Biology_Alberts. Human cells have seven translesion polymerases, some of which can recognize a specific type of DNA damage and correctly add the nucleotide required to restore the initial sequence. Others make only “good guesses,” especially when the template base has been extensively damaged. These enzymes are not as accurate as the normal replicative polymerases when they copy a normal DNA sequence. For one thing, the translesion polymerases lack exonucleolytic proofreading activity; in addition, many are much less discriminating than the replicative polymerase in choosing which nucleotide to incorporate initially. Presumably for this reason, each such translesion polymerase is given a chance to add only one or a few nucleotides before the highly accurate replicative polymerase resumes DNA synthesis.
Cell_Biology_Alberts_1269
Cell_Biology_Alberts
Despite their usefulness in allowing heavily damaged DNA to be replicated, these translesion polymerases do, as noted above, pose risks to the cell. They are probably responsible for most of the base-substitution and single-nucleotide deletion mutations that accumulate in genomes; although they generally produce mutations when copying damaged DNA (see Figure 5–40), they probably also create mutations—at a low level—on undamaged DNA. Clearly, it is important for the cell to tightly regulate these polymerases, releasing them only at sites of DNA damage. Exactly how this happens for each translesion polymerase remains to be discovered, but a conceptual model is given in Figure 5–44. The principle of this model applies to many of the DNA repair processes discussed in this chapter: because the enzymes that carry out these reactions are potentially dangerous to the genome, they must be brought into play only at sites of damage.
Cell_Biology_Alberts. Despite their usefulness in allowing heavily damaged DNA to be replicated, these translesion polymerases do, as noted above, pose risks to the cell. They are probably responsible for most of the base-substitution and single-nucleotide deletion mutations that accumulate in genomes; although they generally produce mutations when copying damaged DNA (see Figure 5–40), they probably also create mutations—at a low level—on undamaged DNA. Clearly, it is important for the cell to tightly regulate these polymerases, releasing them only at sites of DNA damage. Exactly how this happens for each translesion polymerase remains to be discovered, but a conceptual model is given in Figure 5–44. The principle of this model applies to many of the DNA repair processes discussed in this chapter: because the enzymes that carry out these reactions are potentially dangerous to the genome, they must be brought into play only at sites of damage.
Cell_Biology_Alberts_1270
Cell_Biology_Alberts
An especially dangerous type of DNA damage occurs when both strands of the double helix are broken, leaving no intact template strand to enable accurate Figure 5–44 Translesion DNA polymerases can use damaged templates. According to this model, a replicative polymerase stalled at a site of DNA damage is recognized by the cell as needing rescue. Specialized enzymes covalently modify the sliding clamp (typically, it is ubiquitylated—see Figure 3–69) which releases the replicative DNA polymerase and, together with damaged DNA, attracts a translesion polymerase specific to that type of damage. Once the damaged DNA is bypassed, the covalent modification of the clamp is removed, the translesion polymerase dissociates, and the replicative polymerase is brought back into play.
Cell_Biology_Alberts. An especially dangerous type of DNA damage occurs when both strands of the double helix are broken, leaving no intact template strand to enable accurate Figure 5–44 Translesion DNA polymerases can use damaged templates. According to this model, a replicative polymerase stalled at a site of DNA damage is recognized by the cell as needing rescue. Specialized enzymes covalently modify the sliding clamp (typically, it is ubiquitylated—see Figure 3–69) which releases the replicative DNA polymerase and, together with damaged DNA, attracts a translesion polymerase specific to that type of damage. Once the damaged DNA is bypassed, the covalent modification of the clamp is removed, the translesion polymerase dissociates, and the replicative polymerase is brought back into play.
Cell_Biology_Alberts_1271
Cell_Biology_Alberts
removal of covalent modifcations, reloading of replicative DNA polymerase, DNA synthesis continues repair. Ionizing radiation, replication errors, oxidizing agents, and other metabolites produced in the cell cause breaks of this type. If these lesions were left unrepaired, they would quickly lead to the breakdown of chromosomes into smaller fragments and to loss of genes when the cell divides. However, two distinct mechanisms have evolved to deal with this type of damage (Figure 5–45). The simplest to understand is nonhomologous end joining, in which the broken ends are simply brought together and rejoined by DNA ligation, generally with the loss of nucleotides at the site of joining (Figure 5–46). This end-joining mechanism, which can be seen as a “quick and dirty” solution to the repair of double-strand breaks, is common in mammalian somatic cells. Although a change in the DNA sequence (a mutation) results at the site of breakage, so little of the mammalian genome is essential for
Cell_Biology_Alberts. removal of covalent modifcations, reloading of replicative DNA polymerase, DNA synthesis continues repair. Ionizing radiation, replication errors, oxidizing agents, and other metabolites produced in the cell cause breaks of this type. If these lesions were left unrepaired, they would quickly lead to the breakdown of chromosomes into smaller fragments and to loss of genes when the cell divides. However, two distinct mechanisms have evolved to deal with this type of damage (Figure 5–45). The simplest to understand is nonhomologous end joining, in which the broken ends are simply brought together and rejoined by DNA ligation, generally with the loss of nucleotides at the site of joining (Figure 5–46). This end-joining mechanism, which can be seen as a “quick and dirty” solution to the repair of double-strand breaks, is common in mammalian somatic cells. Although a change in the DNA sequence (a mutation) results at the site of breakage, so little of the mammalian genome is essential for
Cell_Biology_Alberts_1272
Cell_Biology_Alberts
of double-strand breaks, is common in mammalian somatic cells. Although a change in the DNA sequence (a mutation) results at the site of breakage, so little of the mammalian genome is essential for life that this mechanism is apparently an acceptable solution to the problem of rejoining broken chromosomes. By the time a human reaches the age of 70, the typical somatic cell contains over 2000 such “scars,” distributed throughout its genome, representing places where DNA has been inaccurately repaired by nonhomologous end joining. But nonhomologous end joining presents another danger: because there seems to be no mechanism to ensure that two ends being joined were originally next to each other in the genome, nonhomologous end joining can occasionally generate rearrangements in which one broken chromosome becomes covalently attached to another. This can result in chromosomes with two centromeres and chromosomes lacking centromeres altogether; both processing of processing of 5˜ endsDNA
Cell_Biology_Alberts. of double-strand breaks, is common in mammalian somatic cells. Although a change in the DNA sequence (a mutation) results at the site of breakage, so little of the mammalian genome is essential for life that this mechanism is apparently an acceptable solution to the problem of rejoining broken chromosomes. By the time a human reaches the age of 70, the typical somatic cell contains over 2000 such “scars,” distributed throughout its genome, representing places where DNA has been inaccurately repaired by nonhomologous end joining. But nonhomologous end joining presents another danger: because there seems to be no mechanism to ensure that two ends being joined were originally next to each other in the genome, nonhomologous end joining can occasionally generate rearrangements in which one broken chromosome becomes covalently attached to another. This can result in chromosomes with two centromeres and chromosomes lacking centromeres altogether; both processing of processing of 5˜ endsDNA
Cell_Biology_Alberts_1273
Cell_Biology_Alberts
chromosome becomes covalently attached to another. This can result in chromosomes with two centromeres and chromosomes lacking centromeres altogether; both processing of processing of 5˜ endsDNA ends by nuclease deletion of DNA sequence damage repaired accurately using sister chromatid as the template types of aberrant chromosomes are missegregated during cell division. As previously discussed, the specialized structure of telomeres prevents the natural ends of chromosomes from being mistaken for broken DNA and “repaired” in this way.
Cell_Biology_Alberts. chromosome becomes covalently attached to another. This can result in chromosomes with two centromeres and chromosomes lacking centromeres altogether; both processing of processing of 5˜ endsDNA ends by nuclease deletion of DNA sequence damage repaired accurately using sister chromatid as the template types of aberrant chromosomes are missegregated during cell division. As previously discussed, the specialized structure of telomeres prevents the natural ends of chromosomes from being mistaken for broken DNA and “repaired” in this way.
Cell_Biology_Alberts_1274
Cell_Biology_Alberts
A much more accurate type of double-strand break repair occurs in newly replicated DNA (Figure 5–45B). Here, the DNA is repaired using the sister chromatid as a template. This reaction is an example of homologous recombination, and we consider its mechanism later in this chapter. Most organisms employ both non-homologous end joining and homologous recombination to repair double-strand breaks in DNA. Nonhomologous end joining predominates in humans; homologous recombination is used only during and shortly after DNA replication (in S and G2 phases), when sister chromatids are available to serve as templates. repaired DNA has generally suffered a deletion of nucleotides (A)
Cell_Biology_Alberts. A much more accurate type of double-strand break repair occurs in newly replicated DNA (Figure 5–45B). Here, the DNA is repaired using the sister chromatid as a template. This reaction is an example of homologous recombination, and we consider its mechanism later in this chapter. Most organisms employ both non-homologous end joining and homologous recombination to repair double-strand breaks in DNA. Nonhomologous end joining predominates in humans; homologous recombination is used only during and shortly after DNA replication (in S and G2 phases), when sister chromatids are available to serve as templates. repaired DNA has generally suffered a deletion of nucleotides (A)
Cell_Biology_Alberts_1275
Cell_Biology_Alberts
repaired DNA has generally suffered a deletion of nucleotides (A) Figure 5–45 Two ways to repair double-strand breaks. (A) Nonhomologous end joining alters the original DNA sequence when repairing a broken chromosome. The initial degradation of the broken DNA ends is important because the nucleotides at the site of the initial break are often damaged and cannot be ligated. Nonhomologous end joining usually takes place when cells have not yet duplicated their DNA. (B) Repairing double-strand breaks by homologous recombination is more difficult to accomplish but restores the original DNA sequence. It typically takes place after the DNA has been duplicated (when a duplex template is available) but before the cell has divided. Details of the homologous recombination pathway are presented in the following section (see Figure 5–48).
Cell_Biology_Alberts. repaired DNA has generally suffered a deletion of nucleotides (A) Figure 5–45 Two ways to repair double-strand breaks. (A) Nonhomologous end joining alters the original DNA sequence when repairing a broken chromosome. The initial degradation of the broken DNA ends is important because the nucleotides at the site of the initial break are often damaged and cannot be ligated. Nonhomologous end joining usually takes place when cells have not yet duplicated their DNA. (B) Repairing double-strand breaks by homologous recombination is more difficult to accomplish but restores the original DNA sequence. It typically takes place after the DNA has been duplicated (when a duplex template is available) but before the cell has divided. Details of the homologous recombination pathway are presented in the following section (see Figure 5–48).
Cell_Biology_Alberts_1276
Cell_Biology_Alberts
Figure 5–46 Nonhomologous end joining. (A) A central role is played by the Ku protein, a heterodimer that grasps the broken chromosome ends. The additional proteins shown are needed to hold the broken ends together while they are processed and eventually joined covalently. (B) Three-dimensional structure of a Ku heterodimer bound to the end of a duplex DNA fragment. The Ku protein is also essential for V(D)J joining, a specific recombination process through which antibody and T cell receptor diversity is generated in developing B and T cells (discussed in Chapter 24). V(D)J joining and nonhomologous end joining show many similarities in mechanism but the former relies on specific double-strand breaks produced deliberately by the cell. (B, from J.R. Walker, R.A. Corpina, and J. Goldberg, Nature 412:607–614, 2001. With permission from Macmillan Publishers Ltd.) DNA Damage Delays Progression of the Cell Cycle
Cell_Biology_Alberts. Figure 5–46 Nonhomologous end joining. (A) A central role is played by the Ku protein, a heterodimer that grasps the broken chromosome ends. The additional proteins shown are needed to hold the broken ends together while they are processed and eventually joined covalently. (B) Three-dimensional structure of a Ku heterodimer bound to the end of a duplex DNA fragment. The Ku protein is also essential for V(D)J joining, a specific recombination process through which antibody and T cell receptor diversity is generated in developing B and T cells (discussed in Chapter 24). V(D)J joining and nonhomologous end joining show many similarities in mechanism but the former relies on specific double-strand breaks produced deliberately by the cell. (B, from J.R. Walker, R.A. Corpina, and J. Goldberg, Nature 412:607–614, 2001. With permission from Macmillan Publishers Ltd.) DNA Damage Delays Progression of the Cell Cycle
Cell_Biology_Alberts_1277
Cell_Biology_Alberts
DNA Damage Delays Progression of the Cell Cycle We have just seen that cells contain multiple enzyme systems that can recognize and repair many types of DNA damage (Movie 5.7). Because of the importance of maintaining intact, undamaged DNA from generation to generation, eukaryotic cells have an additional mechanism that maximizes the effectiveness of their DNA repair enzymes: they delay progression of the cell cycle until DNA repair is complete. As discussed in detail in Chapter 17, the orderly progression of the cell cycle is stopped if damaged DNA is detected, and it restarts when the damage has been repaired. Thus, in mammalian cells, the presence of DNA damage can block entry from G1 into S phase, it can slow S phase once it has begun, and it can block the transition from G2 phase to M phase. These delays facilitate DNA repair by providing the time needed for the repair to reach completion.
Cell_Biology_Alberts. DNA Damage Delays Progression of the Cell Cycle We have just seen that cells contain multiple enzyme systems that can recognize and repair many types of DNA damage (Movie 5.7). Because of the importance of maintaining intact, undamaged DNA from generation to generation, eukaryotic cells have an additional mechanism that maximizes the effectiveness of their DNA repair enzymes: they delay progression of the cell cycle until DNA repair is complete. As discussed in detail in Chapter 17, the orderly progression of the cell cycle is stopped if damaged DNA is detected, and it restarts when the damage has been repaired. Thus, in mammalian cells, the presence of DNA damage can block entry from G1 into S phase, it can slow S phase once it has begun, and it can block the transition from G2 phase to M phase. These delays facilitate DNA repair by providing the time needed for the repair to reach completion.
Cell_Biology_Alberts_1278
Cell_Biology_Alberts
DNA damage also results in an increased synthesis of some DNA repair enzymes. This response depends on special signaling proteins that sense DNA damage and up-regulate the appropriate DNA repair enzymes. The importance of this mechanism is revealed by the phenotype of humans who are born with defects in the gene that encodes the ATM protein. These individuals have the disease ataxia telangiectasia (AT ), the symptoms of which include neurodegeneration, a predisposition to cancer, and genome instability. The ATM protein is a large kinase needed to generate the intracellular signals that sound the alarm in response to many types of spontaneous DNA damage (see Figure 17–62), and individuals with defects in this protein therefore suffer from the effects of unrepaired DNA lesions.
Cell_Biology_Alberts. DNA damage also results in an increased synthesis of some DNA repair enzymes. This response depends on special signaling proteins that sense DNA damage and up-regulate the appropriate DNA repair enzymes. The importance of this mechanism is revealed by the phenotype of humans who are born with defects in the gene that encodes the ATM protein. These individuals have the disease ataxia telangiectasia (AT ), the symptoms of which include neurodegeneration, a predisposition to cancer, and genome instability. The ATM protein is a large kinase needed to generate the intracellular signals that sound the alarm in response to many types of spontaneous DNA damage (see Figure 17–62), and individuals with defects in this protein therefore suffer from the effects of unrepaired DNA lesions.
Cell_Biology_Alberts_1279
Cell_Biology_Alberts
Genetic information can be stored stably in DNA sequences only because a large set of DNA repair enzymes continuously scan the DNA and replace any damaged nucleotides. Most types of DNA repair depend on the presence of a separate copy of the genetic information in each of the two strands of the DNA double helix. An accidental lesion on one strand can therefore be cut out by a repair enzyme and a corrected strand resynthesized by reference to the information in the undamaged strand.
Cell_Biology_Alberts. Genetic information can be stored stably in DNA sequences only because a large set of DNA repair enzymes continuously scan the DNA and replace any damaged nucleotides. Most types of DNA repair depend on the presence of a separate copy of the genetic information in each of the two strands of the DNA double helix. An accidental lesion on one strand can therefore be cut out by a repair enzyme and a corrected strand resynthesized by reference to the information in the undamaged strand.
Cell_Biology_Alberts_1280
Cell_Biology_Alberts
Most of the damage to DNA bases is excised by one of two major DNA repair pathways. In base excision repair, the altered base is removed by a DNA glycosylase enzyme, followed by excision of the resulting sugar phosphate. In nucleotide excision repair, a small section of the DNA strand surrounding the damage is removed from the DNA double helix as an oligonucleotide. In both cases, the gap left in the DNA helix is filled in by the sequential action of DNA polymerase and DNA ligase, using the undamaged DNA strand as the template. Some types of DNA damage can be repaired by a different strategy—the direct chemical reversal of the damage— which is carried out by specialized repair proteins. When DNA damage is excessive, a special class of inaccurate DNA polymerases, called translesion polymerases, is used to bypass the damage, allowing the cell to survive but sometimes creating permanent mutations at the sites of damage.
Cell_Biology_Alberts. Most of the damage to DNA bases is excised by one of two major DNA repair pathways. In base excision repair, the altered base is removed by a DNA glycosylase enzyme, followed by excision of the resulting sugar phosphate. In nucleotide excision repair, a small section of the DNA strand surrounding the damage is removed from the DNA double helix as an oligonucleotide. In both cases, the gap left in the DNA helix is filled in by the sequential action of DNA polymerase and DNA ligase, using the undamaged DNA strand as the template. Some types of DNA damage can be repaired by a different strategy—the direct chemical reversal of the damage— which is carried out by specialized repair proteins. When DNA damage is excessive, a special class of inaccurate DNA polymerases, called translesion polymerases, is used to bypass the damage, allowing the cell to survive but sometimes creating permanent mutations at the sites of damage.
Cell_Biology_Alberts_1281
Cell_Biology_Alberts
Other critical repair systems—based on either nonhomologous end joining or homologous recombination—reseal the accidental double-strand breaks that occur in the DNA helix. In most cells, an elevated level of DNA damage causes a delay in the cell cycle, which ensures that DNA damage is repaired before a cell divides.
Cell_Biology_Alberts. Other critical repair systems—based on either nonhomologous end joining or homologous recombination—reseal the accidental double-strand breaks that occur in the DNA helix. In most cells, an elevated level of DNA damage causes a delay in the cell cycle, which ensures that DNA damage is repaired before a cell divides.
Cell_Biology_Alberts_1282
Cell_Biology_Alberts
In the two preceding sections, we discussed the mechanisms that allow the DNA sequences in cells to be maintained from generation to generation with very little change. In this section, we further explore one of the DNA repair mechanisms, a diverse set of reactions known collectively as homologous recombination. The key feature of homologous recombination (also known as general recombination) is an exchange of DNA strands between a pair of homologous duplex DNA sequences, that is, segments of double helix that are very similar or identical in nucleotide sequence. This exchange allows one stretch of duplex DNA to act as a template to restore lost or damaged information on a second stretch of duplex DNA. Because the template for repair is not limited to the strand complementary to that containing the damage, homologous recombination can repair many types of DNA damage. It is, for example, the main way to accurately repair double-strand breaks, as introduced in the previous section (see
Cell_Biology_Alberts. In the two preceding sections, we discussed the mechanisms that allow the DNA sequences in cells to be maintained from generation to generation with very little change. In this section, we further explore one of the DNA repair mechanisms, a diverse set of reactions known collectively as homologous recombination. The key feature of homologous recombination (also known as general recombination) is an exchange of DNA strands between a pair of homologous duplex DNA sequences, that is, segments of double helix that are very similar or identical in nucleotide sequence. This exchange allows one stretch of duplex DNA to act as a template to restore lost or damaged information on a second stretch of duplex DNA. Because the template for repair is not limited to the strand complementary to that containing the damage, homologous recombination can repair many types of DNA damage. It is, for example, the main way to accurately repair double-strand breaks, as introduced in the previous section (see
Cell_Biology_Alberts_1283
Cell_Biology_Alberts
the damage, homologous recombination can repair many types of DNA damage. It is, for example, the main way to accurately repair double-strand breaks, as introduced in the previous section (see Figure 5–45B). Double-strand breaks can result from radiation and reactive chemicals, but most of the time they arise from DNA replication forks that become stalled or broken independently of any such external cause. Homologous recombination accurately corrects these accidents and, because they occur during nearly every round of DNA replication, this repair mechanism is essential for every proliferating cell. Homologous recombination is perhaps the most versatile DNA repair mechanism available to the cell; the “all-purpose” nature of recombinational repair probably explains why its mechanism and the proteins that carry it out have been conserved in virtually all cells on Earth.
Cell_Biology_Alberts. the damage, homologous recombination can repair many types of DNA damage. It is, for example, the main way to accurately repair double-strand breaks, as introduced in the previous section (see Figure 5–45B). Double-strand breaks can result from radiation and reactive chemicals, but most of the time they arise from DNA replication forks that become stalled or broken independently of any such external cause. Homologous recombination accurately corrects these accidents and, because they occur during nearly every round of DNA replication, this repair mechanism is essential for every proliferating cell. Homologous recombination is perhaps the most versatile DNA repair mechanism available to the cell; the “all-purpose” nature of recombinational repair probably explains why its mechanism and the proteins that carry it out have been conserved in virtually all cells on Earth.
Cell_Biology_Alberts_1284
Cell_Biology_Alberts
Additionally, we shall see that homologous recombination plays a special role in sexually reproducing organisms. During meiosis, a key step in gamete (sperm and egg) production, it catalyzes the orderly exchange of bits of genetic information between corresponding (homologous) maternal and paternal chromosomes to create new combinations of DNA sequences in the chromosomes passed to the offspring.
Cell_Biology_Alberts. Additionally, we shall see that homologous recombination plays a special role in sexually reproducing organisms. During meiosis, a key step in gamete (sperm and egg) production, it catalyzes the orderly exchange of bits of genetic information between corresponding (homologous) maternal and paternal chromosomes to create new combinations of DNA sequences in the chromosomes passed to the offspring.
Cell_Biology_Alberts_1285
Cell_Biology_Alberts
The current view of homologous recombination as a critical DNA repair mechanism in all cells evolved slowly from its original discovery as a key component in the specialized process of meiosis in plants and animals. The subsequent recognition that homologous recombination also occurs in unicellular organisms made it much more amenable to molecular analyses. Thus, most of what we know about the biochemistry of genetic recombination was originally derived from studies of bacteria, especially of E. coli and its viruses, as well as from experiments with simple eukaryotes such as yeasts. For these organisms with short generation times and relatively small genomes, it was possible to isolate a large set of mutants with defects in their recombination processes. The protein altered in each mutant was then identified and, ultimately, studied biochemically. Close relatives of these proteins have been found in more complex eukaryotes including flies, mice, and humans, and more recently, it has
Cell_Biology_Alberts. The current view of homologous recombination as a critical DNA repair mechanism in all cells evolved slowly from its original discovery as a key component in the specialized process of meiosis in plants and animals. The subsequent recognition that homologous recombination also occurs in unicellular organisms made it much more amenable to molecular analyses. Thus, most of what we know about the biochemistry of genetic recombination was originally derived from studies of bacteria, especially of E. coli and its viruses, as well as from experiments with simple eukaryotes such as yeasts. For these organisms with short generation times and relatively small genomes, it was possible to isolate a large set of mutants with defects in their recombination processes. The protein altered in each mutant was then identified and, ultimately, studied biochemically. Close relatives of these proteins have been found in more complex eukaryotes including flies, mice, and humans, and more recently, it has
Cell_Biology_Alberts_1286
Cell_Biology_Alberts
was then identified and, ultimately, studied biochemically. Close relatives of these proteins have been found in more complex eukaryotes including flies, mice, and humans, and more recently, it has been possible to directly analyze homologous recombination in these species as well. These studies reveal that the fundamental processes that catalyze homologous recombination are common to all cells.
Cell_Biology_Alberts. was then identified and, ultimately, studied biochemically. Close relatives of these proteins have been found in more complex eukaryotes including flies, mice, and humans, and more recently, it has been possible to directly analyze homologous recombination in these species as well. These studies reveal that the fundamental processes that catalyze homologous recombination are common to all cells.
Cell_Biology_Alberts_1287
Cell_Biology_Alberts
The hallmark of homologous recombination is that it takes place only between DNA duplexes that have extensive regions of sequence similarity (homology). Not surprisingly, base-pairing underlies this requirement, and two DNA duplexes that are undergoing homologous recombination “sample” each other’s DNA sequence by engaging in extensive base-pairing between a single strand from one DNA duplex and the complementary single strand from the other. The match need not be perfect, but it must be very close for homologous recombination to succeed.
Cell_Biology_Alberts. The hallmark of homologous recombination is that it takes place only between DNA duplexes that have extensive regions of sequence similarity (homology). Not surprisingly, base-pairing underlies this requirement, and two DNA duplexes that are undergoing homologous recombination “sample” each other’s DNA sequence by engaging in extensive base-pairing between a single strand from one DNA duplex and the complementary single strand from the other. The match need not be perfect, but it must be very close for homologous recombination to succeed.
Cell_Biology_Alberts_1288
Cell_Biology_Alberts
In its simplest form, this type of base-pairing interaction can be mimicked in a test tube by allowing a DNA double helix to re-form from its separated single strands. This process, called DNA renaturation or hybridization, occurs when a rare random collision juxtaposes complementary nucleotide sequences on two matching DNA single strands, allowing the formation of a short stretch of double helix between them. This relatively slow helix-nucleation step is followed by a very rapid “zippering” step, as the region of double helix is extended to maximize the number of base-pairing interactions (Figure 5–47).
Cell_Biology_Alberts. In its simplest form, this type of base-pairing interaction can be mimicked in a test tube by allowing a DNA double helix to re-form from its separated single strands. This process, called DNA renaturation or hybridization, occurs when a rare random collision juxtaposes complementary nucleotide sequences on two matching DNA single strands, allowing the formation of a short stretch of double helix between them. This relatively slow helix-nucleation step is followed by a very rapid “zippering” step, as the region of double helix is extended to maximize the number of base-pairing interactions (Figure 5–47).
Cell_Biology_Alberts_1289
Cell_Biology_Alberts
DNA hybridization can create a region of DNA double helix consisting of strands that originate from two different duplex DNA molecules as long as they are complementary, or nearly so. As we will see shortly, the formation of such a hybrid molecule, known as a heteroduplex, is an essential feature of homologous recombination. DNA hybridization and heteroduplex formation is also the basis for many of the methods used to study cells, and we will discuss these uses in Chapter 8. The DNA in a living cell is almost all in the stable double-helical form, so the reaction depicted in Figure 5–47 rarely occurs in vivo. Instead, as we shall see, homologous recombination is brought about through a carefully controlled set of reactions that allow two DNA duplexes to sample each other’s sequences without fully dissociating into single strands.
Cell_Biology_Alberts. DNA hybridization can create a region of DNA double helix consisting of strands that originate from two different duplex DNA molecules as long as they are complementary, or nearly so. As we will see shortly, the formation of such a hybrid molecule, known as a heteroduplex, is an essential feature of homologous recombination. DNA hybridization and heteroduplex formation is also the basis for many of the methods used to study cells, and we will discuss these uses in Chapter 8. The DNA in a living cell is almost all in the stable double-helical form, so the reaction depicted in Figure 5–47 rarely occurs in vivo. Instead, as we shall see, homologous recombination is brought about through a carefully controlled set of reactions that allow two DNA duplexes to sample each other’s sequences without fully dissociating into single strands.
Cell_Biology_Alberts_1290
Cell_Biology_Alberts
We saw in the previous section that nonhomologous end-joining occurs without a template and usually leaves a mutation at the site at which a double-strand break is repaired. In contrast, homologous recombination can repair double-strand breaks accurately, without any loss or alteration of nucleotides at the site of repair. For homologous recombination to do this repair job, the broken DNA has to be brought into proximity with homologous but unbroken DNA, which can serve as a template for repair. For this reason, homologous recombination often occurs just after DNA replication, when the two daughter DNA molecules lie close together and one can serve as a template for repair of the other. As we shall see, the process of DNA replication itself creates a special risk of accidents requiring this sort of repair.
Cell_Biology_Alberts. We saw in the previous section that nonhomologous end-joining occurs without a template and usually leaves a mutation at the site at which a double-strand break is repaired. In contrast, homologous recombination can repair double-strand breaks accurately, without any loss or alteration of nucleotides at the site of repair. For homologous recombination to do this repair job, the broken DNA has to be brought into proximity with homologous but unbroken DNA, which can serve as a template for repair. For this reason, homologous recombination often occurs just after DNA replication, when the two daughter DNA molecules lie close together and one can serve as a template for repair of the other. As we shall see, the process of DNA replication itself creates a special risk of accidents requiring this sort of repair.
Cell_Biology_Alberts_1291
Cell_Biology_Alberts
The simplest pathway through which homologous recombination can repair double-strand breaks is shown in Figure 5–48. In essence, the broken DNA duplex and the template duplex carry out a “strand dance” so that one of the damaged strands can use the complementary strand of the intact DNA duplex as a template for repair. First, the ends of the broken DNA are chewed back, or “resected,” by specialized nucleases to produce overhanging, single-strand 3ʹ ends. The next step is strand exchange (also called strand invasion), during which one of the single-strand 3ʹ ends from the damaged DNA molecule worms its way into the template duplex and searches it for homologous sequences through base-pairing. We describe this remarkable reaction in detail in the next section. Once stable base-pairing is established (which completes the strand exchange step), an accurate DNA polymerase extends the invading strand by using the information provided by the undamaged template molecule, thus restoring the
Cell_Biology_Alberts. The simplest pathway through which homologous recombination can repair double-strand breaks is shown in Figure 5–48. In essence, the broken DNA duplex and the template duplex carry out a “strand dance” so that one of the damaged strands can use the complementary strand of the intact DNA duplex as a template for repair. First, the ends of the broken DNA are chewed back, or “resected,” by specialized nucleases to produce overhanging, single-strand 3ʹ ends. The next step is strand exchange (also called strand invasion), during which one of the single-strand 3ʹ ends from the damaged DNA molecule worms its way into the template duplex and searches it for homologous sequences through base-pairing. We describe this remarkable reaction in detail in the next section. Once stable base-pairing is established (which completes the strand exchange step), an accurate DNA polymerase extends the invading strand by using the information provided by the undamaged template molecule, thus restoring the
Cell_Biology_Alberts_1292
Cell_Biology_Alberts
established (which completes the strand exchange step), an accurate DNA polymerase extends the invading strand by using the information provided by the undamaged template molecule, thus restoring the damaged DNA. The last steps—strand displacement, further repair synthesis, and ligation—restore the two original DNA double helices and complete the repair process. Homologous recombination resembles other DNA repair reactions in that a
Cell_Biology_Alberts. established (which completes the strand exchange step), an accurate DNA polymerase extends the invading strand by using the information provided by the undamaged template molecule, thus restoring the damaged DNA. The last steps—strand displacement, further repair synthesis, and ligation—restore the two original DNA double helices and complete the repair process. Homologous recombination resembles other DNA repair reactions in that a
Cell_Biology_Alberts_1293
Cell_Biology_Alberts
Figure 5–47 DNA hybridization. DNA double helices can re-form from their separated strands in a reaction that depends on the random collision of two complementary DNA strands. The vast majority of such collisions are not productive, as shown on the left, but a few result in a short region where complementary base pairs have formed (helix nucleation). A rapid zippering then leads to the formation of a complete double helix. Through this trial-and-error process, a DNA strand will find its complementary partner even in the midst of millions of nonmatching DNA strands. DNA polymerase utilizes a pristine template to restore damaged DNA. However, instead of using the partner complementary strand as a template, as occurs in most DNA repair pathways, homologous recombination exploits a complementary strand from a separate DNA duplex. Strand Exchange Is Carried Out by the RecA/Rad51 Protein
Cell_Biology_Alberts. Figure 5–47 DNA hybridization. DNA double helices can re-form from their separated strands in a reaction that depends on the random collision of two complementary DNA strands. The vast majority of such collisions are not productive, as shown on the left, but a few result in a short region where complementary base pairs have formed (helix nucleation). A rapid zippering then leads to the formation of a complete double helix. Through this trial-and-error process, a DNA strand will find its complementary partner even in the midst of millions of nonmatching DNA strands. DNA polymerase utilizes a pristine template to restore damaged DNA. However, instead of using the partner complementary strand as a template, as occurs in most DNA repair pathways, homologous recombination exploits a complementary strand from a separate DNA duplex. Strand Exchange Is Carried Out by the RecA/Rad51 Protein
Cell_Biology_Alberts_1294
Cell_Biology_Alberts
Strand Exchange Is Carried Out by the RecA/Rad51 Protein Of all the steps of homologous recombination, strand exchange is the most difficult to imagine. How does the invading single strand rapidly sample a DNA duplex for homology? Once the homology is found, how does the exchange occur? How is the inherent stability of the template double helix overcome? The answers to these questions came from biochemical and structural studies of the protein that carries out these feats, called RecA in E. coli and Rad51 in virtually all eukaryotic organisms. To catalyze strand exchange, RecA first binds cooperatively to the invading single strand, forming a protein–DNA filament that forces the DNA into an unusual configuration: groups of three consecutive nucleotides are held as though they were in a conventional DNA double helix but, between adjacent triplets, the DNA backbone is untwisted and stretched out (Figure 5–49). This unusual protein–DNA filament then binds to duplex DNA
Cell_Biology_Alberts. Strand Exchange Is Carried Out by the RecA/Rad51 Protein Of all the steps of homologous recombination, strand exchange is the most difficult to imagine. How does the invading single strand rapidly sample a DNA duplex for homology? Once the homology is found, how does the exchange occur? How is the inherent stability of the template double helix overcome? The answers to these questions came from biochemical and structural studies of the protein that carries out these feats, called RecA in E. coli and Rad51 in virtually all eukaryotic organisms. To catalyze strand exchange, RecA first binds cooperatively to the invading single strand, forming a protein–DNA filament that forces the DNA into an unusual configuration: groups of three consecutive nucleotides are held as though they were in a conventional DNA double helix but, between adjacent triplets, the DNA backbone is untwisted and stretched out (Figure 5–49). This unusual protein–DNA filament then binds to duplex DNA
Cell_Biology_Alberts_1295
Cell_Biology_Alberts
Figure 5–48 Mechanism of double-strand break repair by homologous recombination. This is the preferred method for repairing DNA double-strand breaks that arise shortly after the DNA has been replicated, while the daughter DNA molecules are still held close together. In general, homologous recombination can be regarded as a flexible series of reactions, with the exact pathway differing from one case to the next. For example, the length of the repair “patch” can vary considerably depending on the extent of 5ʹ processing and new DNA synthesis, indicated in green.
Cell_Biology_Alberts. Figure 5–48 Mechanism of double-strand break repair by homologous recombination. This is the preferred method for repairing DNA double-strand breaks that arise shortly after the DNA has been replicated, while the daughter DNA molecules are still held close together. In general, homologous recombination can be regarded as a flexible series of reactions, with the exact pathway differing from one case to the next. For example, the length of the repair “patch” can vary considerably depending on the extent of 5ʹ processing and new DNA synthesis, indicated in green.
Cell_Biology_Alberts_1296
Cell_Biology_Alberts
in a way that stretches the duplex, destabilizing it and making it easy to pull the strands apart. The invading single strand then can sample the sequence of the duplex by conventional base-pairing. This sampling occurs in triplet nucleotide blocks: if a triplet match is found, the adjacent triplet is sampled, and so on. In this way, mismatches quickly lead to dissociation and only an extended stretch of base-pairing (at least 15 nucleotides) stabilizes the invading strand and leads to strand exchange.
Cell_Biology_Alberts. in a way that stretches the duplex, destabilizing it and making it easy to pull the strands apart. The invading single strand then can sample the sequence of the duplex by conventional base-pairing. This sampling occurs in triplet nucleotide blocks: if a triplet match is found, the adjacent triplet is sampled, and so on. In this way, mismatches quickly lead to dissociation and only an extended stretch of base-pairing (at least 15 nucleotides) stabilizes the invading strand and leads to strand exchange.
Cell_Biology_Alberts_1297
Cell_Biology_Alberts
RecA hydrolyzes ATP, and the steps described above require that each RecA monomer along the filament be in the ATP-bound state. However, the searching itself does not require ATP hydrolysis; instead, the process occurs by simple molecular collision, allowing many potential sequences to be rapidly sampled. Once the strand-exchange reaction is completed, however, ATP hydrolysis is necessary to disassemble RecA from the complex of DNA molecules. At this point, repair DNA polymerases and DNA ligase can complete the repair process, as shown in Figure 5–48.
Cell_Biology_Alberts. RecA hydrolyzes ATP, and the steps described above require that each RecA monomer along the filament be in the ATP-bound state. However, the searching itself does not require ATP hydrolysis; instead, the process occurs by simple molecular collision, allowing many potential sequences to be rapidly sampled. Once the strand-exchange reaction is completed, however, ATP hydrolysis is necessary to disassemble RecA from the complex of DNA molecules. At this point, repair DNA polymerases and DNA ligase can complete the repair process, as shown in Figure 5–48.
Cell_Biology_Alberts_1298
Cell_Biology_Alberts
Although accurately repairing double-strand breaks, which can arise from radiation or chemical reactions, is a crucial function of homologous recombination, perhaps its most important role is in rescuing stalled or broken DNA replication forks. Many types of events can cause a replication fork to break, and here we consider just one example: a single-strand nick or gap in the parental DNA helix just ahead of a replication fork. When the fork reaches this lesion, it falls apart—resulting in one broken and one intact daughter chromosome. The broken fork can be flawlessly repaired (Figure 5–50) using the same basic homologous recombination reactions we discussed above for the repair of double-strand breaks. With slight modifications, the set of reactions depicted in Figures 5–48 and 5–50— known collectively as homologous recombination—can accurately repair many different types of DNA damage. Cells Carefully Regulate the Use of Homologous Recombination in DNA Repair
Cell_Biology_Alberts. Although accurately repairing double-strand breaks, which can arise from radiation or chemical reactions, is a crucial function of homologous recombination, perhaps its most important role is in rescuing stalled or broken DNA replication forks. Many types of events can cause a replication fork to break, and here we consider just one example: a single-strand nick or gap in the parental DNA helix just ahead of a replication fork. When the fork reaches this lesion, it falls apart—resulting in one broken and one intact daughter chromosome. The broken fork can be flawlessly repaired (Figure 5–50) using the same basic homologous recombination reactions we discussed above for the repair of double-strand breaks. With slight modifications, the set of reactions depicted in Figures 5–48 and 5–50— known collectively as homologous recombination—can accurately repair many different types of DNA damage. Cells Carefully Regulate the Use of Homologous Recombination in DNA Repair
Cell_Biology_Alberts_1299
Cell_Biology_Alberts
Cells Carefully Regulate the Use of Homologous Recombination in DNA Repair Although homologous recombination neatly solves the problem of accurately repairing double-strand breaks and other types of DNA damage, it does present
Cell_Biology_Alberts. Cells Carefully Regulate the Use of Homologous Recombination in DNA Repair Although homologous recombination neatly solves the problem of accurately repairing double-strand breaks and other types of DNA damage, it does present
Cell_Biology_Alberts_1300
Cell_Biology_Alberts
Figure 5–49 Strand invasion catalyzed by the RecA protein. Our understanding of this reaction is based in part on structures determined by x-ray diffraction studies of RecA bound to singleand double-strand DNA. These DNA structures (shown without the RecA protein) are on the left side of the diagram. Starting at the top, ATP-bound RecA associates with single-strand DNA, holding it in an elongated form where groups of three bases are separated from each other by a stretched and twisted backbone. In the next step, the RecA-bound single strand then binds to duplex DNA, destabilizing it and allowing the single strand to sample its sequence through base-pairing, three bases at a time. If no match is found, the RecA-bound single strand of DNA rapidly dissociates and begins a new search. If an extensive match is found, the structure is disassembled through ATP hydrolysis, resulting in the dissociation of RecA and the exchange of one single strand of DNA for another, thereby forming a
Cell_Biology_Alberts. Figure 5–49 Strand invasion catalyzed by the RecA protein. Our understanding of this reaction is based in part on structures determined by x-ray diffraction studies of RecA bound to singleand double-strand DNA. These DNA structures (shown without the RecA protein) are on the left side of the diagram. Starting at the top, ATP-bound RecA associates with single-strand DNA, holding it in an elongated form where groups of three bases are separated from each other by a stretched and twisted backbone. In the next step, the RecA-bound single strand then binds to duplex DNA, destabilizing it and allowing the single strand to sample its sequence through base-pairing, three bases at a time. If no match is found, the RecA-bound single strand of DNA rapidly dissociates and begins a new search. If an extensive match is found, the structure is disassembled through ATP hydrolysis, resulting in the dissociation of RecA and the exchange of one single strand of DNA for another, thereby forming a
Cell_Biology_Alberts_1301
Cell_Biology_Alberts
If an extensive match is found, the structure is disassembled through ATP hydrolysis, resulting in the dissociation of RecA and the exchange of one single strand of DNA for another, thereby forming a heteroduplex. (PDB code: 3CMX.)
Cell_Biology_Alberts. If an extensive match is found, the structure is disassembled through ATP hydrolysis, resulting in the dissociation of RecA and the exchange of one single strand of DNA for another, thereby forming a heteroduplex. (PDB code: 3CMX.)
Cell_Biology_Alberts_1302
Cell_Biology_Alberts
Figure 5–50 Repair of a broken replication fork by homologous recombination. When a moving replication fork encounters a single-strand break, it will collapse, but can be repaired by homologous recombination. The process uses many of the same reactions shown in Figure 5–48 and proceeds through the same basic steps. Green strands represent the new DNA synthesis that takes place after the replication fork has broken. This pathway allows the fork to move past the site that was nicked on the original template by using the undamaged duplex as a template to synthesize DNA. (Adapted from M.M. Cox, Proc. Natl Acad. Sci. USA 98:8173–8180, 2001. With permission from National Academy of Sciences.) some dangers to the cell as it sometimes “repairs” damage using the wrong bit of the genome as the template. For example, sometimes a broken human chromosome is “repaired” using the homolog from the other parent instead of the sister chromatid as the template. Because maternal and paternal chromosomes
Cell_Biology_Alberts. Figure 5–50 Repair of a broken replication fork by homologous recombination. When a moving replication fork encounters a single-strand break, it will collapse, but can be repaired by homologous recombination. The process uses many of the same reactions shown in Figure 5–48 and proceeds through the same basic steps. Green strands represent the new DNA synthesis that takes place after the replication fork has broken. This pathway allows the fork to move past the site that was nicked on the original template by using the undamaged duplex as a template to synthesize DNA. (Adapted from M.M. Cox, Proc. Natl Acad. Sci. USA 98:8173–8180, 2001. With permission from National Academy of Sciences.) some dangers to the cell as it sometimes “repairs” damage using the wrong bit of the genome as the template. For example, sometimes a broken human chromosome is “repaired” using the homolog from the other parent instead of the sister chromatid as the template. Because maternal and paternal chromosomes
Cell_Biology_Alberts_1303
Cell_Biology_Alberts
For example, sometimes a broken human chromosome is “repaired” using the homolog from the other parent instead of the sister chromatid as the template. Because maternal and paternal chromosomes differ in DNA sequence at many positions along their lengths, this type of repair can convert the sequence of the repaired DNA from the maternal to the paternal sequence or vice versa. The result of this type of errant recombination is known as loss of heterozygosity. It can have severe consequences if the homolog used for repair contains a deleterious mutation, because the recombination event destroys the “good” copy. Loss of heterozygosity, although rare, is a critical step in the formation of many cancers (discussed in Chapter 20).
Cell_Biology_Alberts. For example, sometimes a broken human chromosome is “repaired” using the homolog from the other parent instead of the sister chromatid as the template. Because maternal and paternal chromosomes differ in DNA sequence at many positions along their lengths, this type of repair can convert the sequence of the repaired DNA from the maternal to the paternal sequence or vice versa. The result of this type of errant recombination is known as loss of heterozygosity. It can have severe consequences if the homolog used for repair contains a deleterious mutation, because the recombination event destroys the “good” copy. Loss of heterozygosity, although rare, is a critical step in the formation of many cancers (discussed in Chapter 20).
Cell_Biology_Alberts_1304
Cell_Biology_Alberts
Cells go to great lengths to minimize the risk of mishaps of these types; indeed, nearly every step of homologous recombination is carefully regulated. For example, the first step, processing of the broken ends, is coordinated with the cell cycle: the nuclease enzymes that carry out this process are activated (in part, by phosphorylation) only in the S and G2 phases of the cell cycle, when a daughter duplex (either as a partially replicated chromosome or a fully replicated sister chromatid) can serve as a template for repair (see Figure 5–50). The close proximity of the two daughter chromosomes disfavors the use of other genome sequences in the repair process.
Cell_Biology_Alberts. Cells go to great lengths to minimize the risk of mishaps of these types; indeed, nearly every step of homologous recombination is carefully regulated. For example, the first step, processing of the broken ends, is coordinated with the cell cycle: the nuclease enzymes that carry out this process are activated (in part, by phosphorylation) only in the S and G2 phases of the cell cycle, when a daughter duplex (either as a partially replicated chromosome or a fully replicated sister chromatid) can serve as a template for repair (see Figure 5–50). The close proximity of the two daughter chromosomes disfavors the use of other genome sequences in the repair process.
Cell_Biology_Alberts_1305
Cell_Biology_Alberts
The loading of RecA or Rad52 onto the processed DNA ends and the subsequent strand-exchange reaction are also tightly controlled. Although these proteins alone can carry out these steps in vitro, a series of accessory proteins, including Rad52, is needed in eukaryotic cells to ensure that homologous recombination is efficient and accurate (Figure 5–51). There are many such accessory proteins, and exactly how they coordinate and control homologous recombination remains a mystery. We do know that the enzymes that catalyze recombinational repair are made at relatively high levels in eukaryotes and are dispersed throughout the nucleus in an inactive form. In response to DNA damage, they rapidly converge on the sites of DNA damage, become activated, and form “repair factories” where many lesions are apparently brought together and repaired (Figure 5–52).
Cell_Biology_Alberts. The loading of RecA or Rad52 onto the processed DNA ends and the subsequent strand-exchange reaction are also tightly controlled. Although these proteins alone can carry out these steps in vitro, a series of accessory proteins, including Rad52, is needed in eukaryotic cells to ensure that homologous recombination is efficient and accurate (Figure 5–51). There are many such accessory proteins, and exactly how they coordinate and control homologous recombination remains a mystery. We do know that the enzymes that catalyze recombinational repair are made at relatively high levels in eukaryotes and are dispersed throughout the nucleus in an inactive form. In response to DNA damage, they rapidly converge on the sites of DNA damage, become activated, and form “repair factories” where many lesions are apparently brought together and repaired (Figure 5–52).
Cell_Biology_Alberts_1306
Cell_Biology_Alberts
In Chapter 20, we shall see that both too much and too little homologous recombination can lead to cancer in humans, the former through repair using the “wrong” template (as described above) and the latter through an increased mutation rate caused by inefficient DNA repair. Clearly, a delicate balance has evolved that keeps this process in check on undamaged DNA, while still allowing it to act efficiently and rapidly on DNA lesions as soon as they arise. Not surprisingly, mutations in the components that carry out and regulate homologous recombination are responsible for several inherited forms of cancer. Two of these, the Brca1 and Brca2 proteins, were first discovered because
Cell_Biology_Alberts. In Chapter 20, we shall see that both too much and too little homologous recombination can lead to cancer in humans, the former through repair using the “wrong” template (as described above) and the latter through an increased mutation rate caused by inefficient DNA repair. Clearly, a delicate balance has evolved that keeps this process in check on undamaged DNA, while still allowing it to act efficiently and rapidly on DNA lesions as soon as they arise. Not surprisingly, mutations in the components that carry out and regulate homologous recombination are responsible for several inherited forms of cancer. Two of these, the Brca1 and Brca2 proteins, were first discovered because
Cell_Biology_Alberts_1307
Cell_Biology_Alberts
Figure 5–51 Structure of a portion of the Rad52 protein. This doughnutshaped structure is composed of 11 subunits. Single-strand DNA has been modeled into the deep groove running along the protein surface. Rad52 helps load Rad51 onto single-strand DNA to form the nucleoprotein filament that carries out strand exchange. Rad52 also acts later to re-form the double helix and complete the homologous recombination reaction. (From
Cell_Biology_Alberts. Figure 5–51 Structure of a portion of the Rad52 protein. This doughnutshaped structure is composed of 11 subunits. Single-strand DNA has been modeled into the deep groove running along the protein surface. Rad52 helps load Rad51 onto single-strand DNA to form the nucleoprotein filament that carries out strand exchange. Rad52 also acts later to re-form the double helix and complete the homologous recombination reaction. (From
Cell_Biology_Alberts_1308
Cell_Biology_Alberts
M.R. Singleton et al., Proc. Natl Acad. Sci. USA 99:13492–13497, 2002. With permission from National Academy of Sciences.) mutations in their genes lead to a greatly increased frequency of breast cancer. Because these mutations cause inefficient repair by homologous recombination, accumulation of DNA damage can, in a small proportion of cells, give rise to a cancer. Brca1 regulates an early step in broken-end processing; without it, such ends are not processed correctly for homologous recombination and instead are repaired inaccurately by the nonhomologous end-joining pathway (see Figure 5–45). Brca2 binds to the Rad51 protein, preventing its polymerization on DNA, and thereby maintaining it in an inactive form until it is needed. Normally, upon DNA damage, Brca2 helps to bring Rad51 protein rapidly to sites of damage and, once in place, to release it in its active form onto single-strand DNA.
Cell_Biology_Alberts. M.R. Singleton et al., Proc. Natl Acad. Sci. USA 99:13492–13497, 2002. With permission from National Academy of Sciences.) mutations in their genes lead to a greatly increased frequency of breast cancer. Because these mutations cause inefficient repair by homologous recombination, accumulation of DNA damage can, in a small proportion of cells, give rise to a cancer. Brca1 regulates an early step in broken-end processing; without it, such ends are not processed correctly for homologous recombination and instead are repaired inaccurately by the nonhomologous end-joining pathway (see Figure 5–45). Brca2 binds to the Rad51 protein, preventing its polymerization on DNA, and thereby maintaining it in an inactive form until it is needed. Normally, upon DNA damage, Brca2 helps to bring Rad51 protein rapidly to sites of damage and, once in place, to release it in its active form onto single-strand DNA.
Cell_Biology_Alberts_1309
Cell_Biology_Alberts
We have seen that homologous recombination comprises a group of reactions— including broken-end processing, strand exchange, limited DNA synthesis, and ligation—to exchange DNA sequences between two double helices of similar nucleotide sequence. Having discussed its role in accurately repairing damaged DNA, we now turn to homologous recombination as a means to generate DNA molecules that carry novel combinations of genes as a result of the deliberate exchange of material between different chromosomes. Although this occasionally occurs by accident in mitotic cells (and is often detrimental), it is a frequent and necessary part of meiosis, which occurs in sexually reproducing organisms such as fungi, plants, and animals.
Cell_Biology_Alberts. We have seen that homologous recombination comprises a group of reactions— including broken-end processing, strand exchange, limited DNA synthesis, and ligation—to exchange DNA sequences between two double helices of similar nucleotide sequence. Having discussed its role in accurately repairing damaged DNA, we now turn to homologous recombination as a means to generate DNA molecules that carry novel combinations of genes as a result of the deliberate exchange of material between different chromosomes. Although this occasionally occurs by accident in mitotic cells (and is often detrimental), it is a frequent and necessary part of meiosis, which occurs in sexually reproducing organisms such as fungi, plants, and animals.