id
stringlengths
14
28
title
stringclasses
18 values
content
stringlengths
2
999
contents
stringlengths
19
1.02k
Cell_Biology_Alberts_1110
Cell_Biology_Alberts
The cells of a sexually reproducing animal or plant are of two types: germ cells and somatic cells. The germ cells transmit genetic information from parent to offspring; the somatic cells form the body of the organism (Figure 5–1). We have seen that germ cells must be protected against high rates of mutation to maintain the species. However, the somatic cells of multicellular organisms must also be protected from genetic change to properly maintain the organized structure of the body. Nucleotide changes in somatic cells can give rise to variant cells, some of which, through “local” natural selection, proliferate rapidly at the expense of the rest of the organism. In an extreme case, the result is the uncontrolled cell proliferation that we know as cancer, a disease that causes (in Europe and North America) more than 20% of human deaths each year. These deaths are due largely to an accumulation of changes in the DNA sequences of somatic cells, as discussed in Chapter 20. A significant
Cell_Biology_Alberts. The cells of a sexually reproducing animal or plant are of two types: germ cells and somatic cells. The germ cells transmit genetic information from parent to offspring; the somatic cells form the body of the organism (Figure 5–1). We have seen that germ cells must be protected against high rates of mutation to maintain the species. However, the somatic cells of multicellular organisms must also be protected from genetic change to properly maintain the organized structure of the body. Nucleotide changes in somatic cells can give rise to variant cells, some of which, through “local” natural selection, proliferate rapidly at the expense of the rest of the organism. In an extreme case, the result is the uncontrolled cell proliferation that we know as cancer, a disease that causes (in Europe and North America) more than 20% of human deaths each year. These deaths are due largely to an accumulation of changes in the DNA sequences of somatic cells, as discussed in Chapter 20. A significant
Cell_Biology_Alberts_1111
Cell_Biology_Alberts
and North America) more than 20% of human deaths each year. These deaths are due largely to an accumulation of changes in the DNA sequences of somatic cells, as discussed in Chapter 20. A significant increase in the mutation frequency would presumably cause a disastrous increase in the incidence of cancer by accelerating the rate at which somatic-cell variants arise. Thus, both for the perpetuation of a species with a large number of genes (germ-cell stability) and for the prevention of cancer resulting from mutations in somatic cells (somatic-cell stability), multicellular organisms like ourselves depend on the remarkably high fidelity with which their DNA sequences are replicated and maintained.
Cell_Biology_Alberts. and North America) more than 20% of human deaths each year. These deaths are due largely to an accumulation of changes in the DNA sequences of somatic cells, as discussed in Chapter 20. A significant increase in the mutation frequency would presumably cause a disastrous increase in the incidence of cancer by accelerating the rate at which somatic-cell variants arise. Thus, both for the perpetuation of a species with a large number of genes (germ-cell stability) and for the prevention of cancer resulting from mutations in somatic cells (somatic-cell stability), multicellular organisms like ourselves depend on the remarkably high fidelity with which their DNA sequences are replicated and maintained.
Cell_Biology_Alberts_1112
Cell_Biology_Alberts
In all cells, DNA sequences are maintained and replicated with high fidelity. The mutation rate, approximately one nucleotide change per 1010 nucleotides each time the DNA is replicated, is roughly the same for organisms as different as bacteria and humans. Because of this remarkable accuracy, the sequence of the human genome (approximately 3.2 × 109 nucleotide pairs) is unchanged or changed by only a few nucleotides each time a typical human cell divides. This allows most humans to pass accurate genetic instructions from one generation to the next, and also to avoid the changes in somatic cells that lead to cancer. All organisms duplicate their DNA with extraordinary accuracy before each cell division. In this section, we explore how an elaborate “replication machine” achieves this accuracy, while duplicating DNA at rates as high as 1000 nucleotides per second.
Cell_Biology_Alberts. In all cells, DNA sequences are maintained and replicated with high fidelity. The mutation rate, approximately one nucleotide change per 1010 nucleotides each time the DNA is replicated, is roughly the same for organisms as different as bacteria and humans. Because of this remarkable accuracy, the sequence of the human genome (approximately 3.2 × 109 nucleotide pairs) is unchanged or changed by only a few nucleotides each time a typical human cell divides. This allows most humans to pass accurate genetic instructions from one generation to the next, and also to avoid the changes in somatic cells that lead to cancer. All organisms duplicate their DNA with extraordinary accuracy before each cell division. In this section, we explore how an elaborate “replication machine” achieves this accuracy, while duplicating DNA at rates as high as 1000 nucleotides per second.
Cell_Biology_Alberts_1113
Cell_Biology_Alberts
As introduced in Chapter 1, DNA templating is the mechanism the cell uses to copy the nucleotide sequence of one DNA strand into a complementary DNA sequence (Figure 5–2). This process requires the separation of the DNA helix into two template strands, and entails the recognition of each nucleotide in the DNA template strands by a free (unpolymerized) complementary nucleotide. The separation of Figure 5–1 Germ-line cells and somatic cells carry out fundamentally different functions. In sexually reproducing organisms, the germ-line cells (red) propagate genetic information into the next generation. Somatic cells (blue), which form the body of the organism, are necessary for the survival of germ-line cells but do not themselves leave any progeny. Figure 5–2 The DNA double helix acts as a template for its own duplication.
Cell_Biology_Alberts. As introduced in Chapter 1, DNA templating is the mechanism the cell uses to copy the nucleotide sequence of one DNA strand into a complementary DNA sequence (Figure 5–2). This process requires the separation of the DNA helix into two template strands, and entails the recognition of each nucleotide in the DNA template strands by a free (unpolymerized) complementary nucleotide. The separation of Figure 5–1 Germ-line cells and somatic cells carry out fundamentally different functions. In sexually reproducing organisms, the germ-line cells (red) propagate genetic information into the next generation. Somatic cells (blue), which form the body of the organism, are necessary for the survival of germ-line cells but do not themselves leave any progeny. Figure 5–2 The DNA double helix acts as a template for its own duplication.
Cell_Biology_Alberts_1114
Cell_Biology_Alberts
Figure 5–2 The DNA double helix acts as a template for its own duplication. Because the nucleotide A will pair successfully only with T, and G only with C, each strand of DNA can serve as a template to specify the sequence of nucleotides in its complementary strand by DNA base-pairing. In this way, a double-helical DNA molecule can be copied precisely. the DNA helix exposes the hydrogen-bond donor and acceptor groups on each DNA base for base-pairing with the appropriate incoming free nucleotide, aligning it for its enzyme-catalyzed polymerization into a new DNA chain. The first nucleotide-polymerizing enzyme, DNA polymerase, was discovered in 1957. The free nucleotides that serve as substrates for this enzyme were found to be deoxyribonucleoside triphosphates, and their polymerization into DNA required a single-strand DNA template. Figure 5–3 and Figure 5–4 illustrate the stepwise mechanism of this reaction. The DNA Replication Fork Is Asymmetrical
Cell_Biology_Alberts. Figure 5–2 The DNA double helix acts as a template for its own duplication. Because the nucleotide A will pair successfully only with T, and G only with C, each strand of DNA can serve as a template to specify the sequence of nucleotides in its complementary strand by DNA base-pairing. In this way, a double-helical DNA molecule can be copied precisely. the DNA helix exposes the hydrogen-bond donor and acceptor groups on each DNA base for base-pairing with the appropriate incoming free nucleotide, aligning it for its enzyme-catalyzed polymerization into a new DNA chain. The first nucleotide-polymerizing enzyme, DNA polymerase, was discovered in 1957. The free nucleotides that serve as substrates for this enzyme were found to be deoxyribonucleoside triphosphates, and their polymerization into DNA required a single-strand DNA template. Figure 5–3 and Figure 5–4 illustrate the stepwise mechanism of this reaction. The DNA Replication Fork Is Asymmetrical
Cell_Biology_Alberts_1115
Cell_Biology_Alberts
The DNA Replication Fork Is Asymmetrical During DNA replication inside a cell, each of the two original DNA strands serves as a template for the formation of an entire new strand. Because each of the two daughters of a dividing cell inherits a new DNA double helix containing one original and one new strand (Figure 5–5), the DNA double helix is said to be replicated “semiconservatively.” How is this feat accomplished? 3˜ end of strand 5˜ end of strand Figure 5–3 The chemistry of DNA synthesis. The addition of a deoxyribonucleotide to the 3ʹ end of a polynucleotide chain (the primer strand) is the fundamental reaction by which DNA is synthesized. As shown, base-pairing between an incoming deoxyribonucleoside triphosphate and an existing strand of DNA (the template strand) guides the formation of the new strand of DNA and causes it to have a complementary nucleotide sequence. The way in which complementary nucleotides base-pair is shown in Figure 4–4. direction of chain growth
Cell_Biology_Alberts. The DNA Replication Fork Is Asymmetrical During DNA replication inside a cell, each of the two original DNA strands serves as a template for the formation of an entire new strand. Because each of the two daughters of a dividing cell inherits a new DNA double helix containing one original and one new strand (Figure 5–5), the DNA double helix is said to be replicated “semiconservatively.” How is this feat accomplished? 3˜ end of strand 5˜ end of strand Figure 5–3 The chemistry of DNA synthesis. The addition of a deoxyribonucleotide to the 3ʹ end of a polynucleotide chain (the primer strand) is the fundamental reaction by which DNA is synthesized. As shown, base-pairing between an incoming deoxyribonucleoside triphosphate and an existing strand of DNA (the template strand) guides the formation of the new strand of DNA and causes it to have a complementary nucleotide sequence. The way in which complementary nucleotides base-pair is shown in Figure 4–4. direction of chain growth
Cell_Biology_Alberts_1116
Cell_Biology_Alberts
Figure 5–4 DNA synthesis catalyzed by DNA polymerase. (A) DNA polymerase catalyzes the stepwise addition of a deoxyribonucleotide to the 3ʹ-OH end of a polynucleotide chain, the growing primer strand that is paired to an existing template strand. The newly synthesized DNA strand therefore polymerizes in the 5ʹ-to-3ʹ direction as shown also in the previous figure. Because each incoming deoxyribonucleoside triphosphate must pair with the template strand to be recognized by the DNA polymerase, this strand determines which of the four possible deoxyribonucleotides (A, C, G, or T) will be added. The reaction is driven by a large, favorable free-energy change, caused by the release of pyrophosphate and its subsequent hydrolysis to two molecules of inorganic phosphate. (B) Structure of DNA polymerase complexed wth DNA (orange), as determined by x-ray crystallography (Movie 5.1). The template DNA strand is the longer strand and the newly synthesized DNA is the shorter. (C) Schematic diagram
Cell_Biology_Alberts. Figure 5–4 DNA synthesis catalyzed by DNA polymerase. (A) DNA polymerase catalyzes the stepwise addition of a deoxyribonucleotide to the 3ʹ-OH end of a polynucleotide chain, the growing primer strand that is paired to an existing template strand. The newly synthesized DNA strand therefore polymerizes in the 5ʹ-to-3ʹ direction as shown also in the previous figure. Because each incoming deoxyribonucleoside triphosphate must pair with the template strand to be recognized by the DNA polymerase, this strand determines which of the four possible deoxyribonucleotides (A, C, G, or T) will be added. The reaction is driven by a large, favorable free-energy change, caused by the release of pyrophosphate and its subsequent hydrolysis to two molecules of inorganic phosphate. (B) Structure of DNA polymerase complexed wth DNA (orange), as determined by x-ray crystallography (Movie 5.1). The template DNA strand is the longer strand and the newly synthesized DNA is the shorter. (C) Schematic diagram
Cell_Biology_Alberts_1117
Cell_Biology_Alberts
complexed wth DNA (orange), as determined by x-ray crystallography (Movie 5.1). The template DNA strand is the longer strand and the newly synthesized DNA is the shorter. (C) Schematic diagram of DNA polymerase, based on the structure in (B). The proper base-pair geometry of a correct incoming deoxyribonucleoside triphosphate causes the polymerase to tighten around the base pair, thereby initiating the nucleotide addition reaction (middle diagram (C)). Dissociation of pyrophosphate relaxes the polymerase, allowing translocation of the DNA by one nucleotide so the active site of the polymerase is ready to receive the next deoxyribonucleoside triphosphate.
Cell_Biology_Alberts. complexed wth DNA (orange), as determined by x-ray crystallography (Movie 5.1). The template DNA strand is the longer strand and the newly synthesized DNA is the shorter. (C) Schematic diagram of DNA polymerase, based on the structure in (B). The proper base-pair geometry of a correct incoming deoxyribonucleoside triphosphate causes the polymerase to tighten around the base pair, thereby initiating the nucleotide addition reaction (middle diagram (C)). Dissociation of pyrophosphate relaxes the polymerase, allowing translocation of the DNA by one nucleotide so the active site of the polymerase is ready to receive the next deoxyribonucleoside triphosphate.
Cell_Biology_Alberts_1118
Cell_Biology_Alberts
Analyses carried out in the early 1960s on whole replicating chromosomes revealed a localized region of replication that moves progressively along the parental DNA double helix. Because of its Y-shaped structure, this active region is called a replication fork (Figure 5–6). At the replication fork, a multienzyme complex that contains the DNA polymerase synthesizes the DNA of both new daughter strands.
Cell_Biology_Alberts. Analyses carried out in the early 1960s on whole replicating chromosomes revealed a localized region of replication that moves progressively along the parental DNA double helix. Because of its Y-shaped structure, this active region is called a replication fork (Figure 5–6). At the replication fork, a multienzyme complex that contains the DNA polymerase synthesizes the DNA of both new daughter strands.
Cell_Biology_Alberts_1119
Cell_Biology_Alberts
Initially, the simplest mechanism of DNA replication seemed to be the continuous growth of both new strands, nucleotide by nucleotide, at the replication fork as it moves from one end of a DNA molecule to the other. But because of the antiparallel orientation of the two DNA strands in the DNA double helix (see Figure 5–2), this mechanism would require one daughter strand to polymerize in the 5ʹ-to-3ʹ direction and the other in the 3ʹ-to-5ʹ direction. Such a replication fork would require two distinct types of DNA polymerase enzymes. However, as attractive as this model might be, the DNA polymerases at replication forks can synthesize only in the 5ʹ-to-3ʹ direction.
Cell_Biology_Alberts. Initially, the simplest mechanism of DNA replication seemed to be the continuous growth of both new strands, nucleotide by nucleotide, at the replication fork as it moves from one end of a DNA molecule to the other. But because of the antiparallel orientation of the two DNA strands in the DNA double helix (see Figure 5–2), this mechanism would require one daughter strand to polymerize in the 5ʹ-to-3ʹ direction and the other in the 3ʹ-to-5ʹ direction. Such a replication fork would require two distinct types of DNA polymerase enzymes. However, as attractive as this model might be, the DNA polymerases at replication forks can synthesize only in the 5ʹ-to-3ʹ direction.
Cell_Biology_Alberts_1120
Cell_Biology_Alberts
How, then, can a DNA strand grow in the 3ʹ-to-5ʹ direction? The answer was first suggested by the results of an experiment performed in the late 1960s. Researchers added highly radioactive 3H-thymidine to dividing bacteria for a few seconds, so that only the most recently replicated DNA—that just behind the replication fork—became radiolabeled. This experiment revealed the transient existence of pieces of DNA that were 1000–2000 nucleotides long, now commonly known as Okazaki fragments, at the growing replication fork. (Similar replication Figure 5–5 The semiconservative nature of DNA replication. In a round of replication, each of the two strands of DNA is used as a template for the formation of a complementary DNA strand. The original strands therefore remain intact through many cell generations.
Cell_Biology_Alberts. How, then, can a DNA strand grow in the 3ʹ-to-5ʹ direction? The answer was first suggested by the results of an experiment performed in the late 1960s. Researchers added highly radioactive 3H-thymidine to dividing bacteria for a few seconds, so that only the most recently replicated DNA—that just behind the replication fork—became radiolabeled. This experiment revealed the transient existence of pieces of DNA that were 1000–2000 nucleotides long, now commonly known as Okazaki fragments, at the growing replication fork. (Similar replication Figure 5–5 The semiconservative nature of DNA replication. In a round of replication, each of the two strands of DNA is used as a template for the formation of a complementary DNA strand. The original strands therefore remain intact through many cell generations.
Cell_Biology_Alberts_1121
Cell_Biology_Alberts
intermediates were later found in eukaryotes, where they are only 100–200 nucleotides long.) The Okazaki fragments were shown to be polymerized only in the 5ʹ-to-3ʹ chain direction and to be joined together after their synthesis to create long DNA chains. A replication fork therefore has an asymmetric structure (Figure 5–7). The DNA daughter strand that is synthesized continuously is known as the leading strand. Its synthesis slightly precedes the synthesis of the daughter strand that is synthesized discontinuously, known as the lagging strand. For the lagging strand, the direction of nucleotide polymerization is opposite to the overall direction of DNA chain growth. The synthesis of this strand by a discontinuous “backstitching” mechanism means that DNA replication requires only the 5ʹ-to-3ʹ type of DNA polymerase. The High Fidelity of DNA Replication Requires Several Proofreading Mechanisms
Cell_Biology_Alberts. intermediates were later found in eukaryotes, where they are only 100–200 nucleotides long.) The Okazaki fragments were shown to be polymerized only in the 5ʹ-to-3ʹ chain direction and to be joined together after their synthesis to create long DNA chains. A replication fork therefore has an asymmetric structure (Figure 5–7). The DNA daughter strand that is synthesized continuously is known as the leading strand. Its synthesis slightly precedes the synthesis of the daughter strand that is synthesized discontinuously, known as the lagging strand. For the lagging strand, the direction of nucleotide polymerization is opposite to the overall direction of DNA chain growth. The synthesis of this strand by a discontinuous “backstitching” mechanism means that DNA replication requires only the 5ʹ-to-3ʹ type of DNA polymerase. The High Fidelity of DNA Replication Requires Several Proofreading Mechanisms
Cell_Biology_Alberts_1122
Cell_Biology_Alberts
The High Fidelity of DNA Replication Requires Several Proofreading Mechanisms As discussed above, the fidelity of copying DNA during replication is such that only about one mistake occurs for every 1010 nucleotides copied. This fidelity is much higher than one would expect from the accuracy of complementary base-pairing. The standard complementary base pairs (see Figure 4–4) are not the only ones possible. For example, with small changes in helix geometry, two hydrogen bonds can form between G and T in DNA. In addition, rare tautomeric forms of the four DNA bases occur transiently in ratios of 1 part to 104 or 105. These forms mispair without a change in helix geometry: the rare tautomeric form of C pairs with A instead of G, for example.
Cell_Biology_Alberts. The High Fidelity of DNA Replication Requires Several Proofreading Mechanisms As discussed above, the fidelity of copying DNA during replication is such that only about one mistake occurs for every 1010 nucleotides copied. This fidelity is much higher than one would expect from the accuracy of complementary base-pairing. The standard complementary base pairs (see Figure 4–4) are not the only ones possible. For example, with small changes in helix geometry, two hydrogen bonds can form between G and T in DNA. In addition, rare tautomeric forms of the four DNA bases occur transiently in ratios of 1 part to 104 or 105. These forms mispair without a change in helix geometry: the rare tautomeric form of C pairs with A instead of G, for example.
Cell_Biology_Alberts_1123
Cell_Biology_Alberts
If the DNA polymerase did nothing special when a mispairing occurred between an incoming deoxyribonucleoside triphosphate and the DNA template, the wrong nucleotide would often be incorporated into the new DNA chain, producing frequent mutations. The high fidelity of DNA replication, however, depends not only on the initial base-pairing but also on several “proofreading” mechanisms that act sequentially to correct any initial mispairings that might have occurred. Figure 5–6 Two replication forks moving in opposite directions on a circular chromosome. An active zone of DNA replication moves progressively along a replicating DNA molecule, creating a Y-shaped DNA structure known as a replication fork: the two arms of each Y are the two daughter DNA molecules, and the stem of the Y is the parental DNA helix. In this diagram, parental strands are orange; newly synthesized strands are red. 1 µm (Micrograph courtesy of Jerome Vinograd.)
Cell_Biology_Alberts. If the DNA polymerase did nothing special when a mispairing occurred between an incoming deoxyribonucleoside triphosphate and the DNA template, the wrong nucleotide would often be incorporated into the new DNA chain, producing frequent mutations. The high fidelity of DNA replication, however, depends not only on the initial base-pairing but also on several “proofreading” mechanisms that act sequentially to correct any initial mispairings that might have occurred. Figure 5–6 Two replication forks moving in opposite directions on a circular chromosome. An active zone of DNA replication moves progressively along a replicating DNA molecule, creating a Y-shaped DNA structure known as a replication fork: the two arms of each Y are the two daughter DNA molecules, and the stem of the Y is the parental DNA helix. In this diagram, parental strands are orange; newly synthesized strands are red. 1 µm (Micrograph courtesy of Jerome Vinograd.)
Cell_Biology_Alberts_1124
Cell_Biology_Alberts
Figure 5–7 The structure of a DNA replication fork. Left, replication fork with newly synthesized DNA in red and arrows indicating the 5ʹ-to-3ʹ direction of DNA synthesis. Because both daughter DNA strands are polymerized in the 5ʹ-to-3ʹ direction, the DNA synthesized on the lagging strand must be made initially as a series of short DNA molecules, called Okazaki fragments, named after the scientist who discovered them. Right, the same fork a short time later. On the lagging strand, the Okazaki fragments are synthesized sequentially, with those nearest the fork being the most recently made.
Cell_Biology_Alberts. Figure 5–7 The structure of a DNA replication fork. Left, replication fork with newly synthesized DNA in red and arrows indicating the 5ʹ-to-3ʹ direction of DNA synthesis. Because both daughter DNA strands are polymerized in the 5ʹ-to-3ʹ direction, the DNA synthesized on the lagging strand must be made initially as a series of short DNA molecules, called Okazaki fragments, named after the scientist who discovered them. Right, the same fork a short time later. On the lagging strand, the Okazaki fragments are synthesized sequentially, with those nearest the fork being the most recently made.
Cell_Biology_Alberts_1125
Cell_Biology_Alberts
DNA polymerase performs the first proofreading step just before a new nucleotide is covalently added to the growing chain. Our knowledge of this mechanism comes from studies of several different DNA polymerases, including one produced by a bacterial virus, T7, that replicates inside E. coli. The correct nucleotide has a higher affinity for the moving polymerase than does the incorrect nucleotide, because the correct pairing is more energetically favorable. Moreover, after nucleotide binding, but before the nucleotide is covalently added to the grow-AAAAAAAAA ing chain, the enzyme must undergo a conformational change in which its “grip” tightens around the active site (see Figure 5–4). Because this change occurs more readily with correct than incorrect base-pairing, it allows the polymerase to “double-check” the exact base-pair geometry before it catalyzes the addition of the nucleotide. Incorrectly paired nucleotides are harder to add and therefore more likely to diffuse away before
Cell_Biology_Alberts. DNA polymerase performs the first proofreading step just before a new nucleotide is covalently added to the growing chain. Our knowledge of this mechanism comes from studies of several different DNA polymerases, including one produced by a bacterial virus, T7, that replicates inside E. coli. The correct nucleotide has a higher affinity for the moving polymerase than does the incorrect nucleotide, because the correct pairing is more energetically favorable. Moreover, after nucleotide binding, but before the nucleotide is covalently added to the grow-AAAAAAAAA ing chain, the enzyme must undergo a conformational change in which its “grip” tightens around the active site (see Figure 5–4). Because this change occurs more readily with correct than incorrect base-pairing, it allows the polymerase to “double-check” the exact base-pair geometry before it catalyzes the addition of the nucleotide. Incorrectly paired nucleotides are harder to add and therefore more likely to diffuse away before
Cell_Biology_Alberts_1126
Cell_Biology_Alberts
to “double-check” the exact base-pair geometry before it catalyzes the addition of the nucleotide. Incorrectly paired nucleotides are harder to add and therefore more likely to diffuse away before the polymerase can mistakenly add them.
Cell_Biology_Alberts. to “double-check” the exact base-pair geometry before it catalyzes the addition of the nucleotide. Incorrectly paired nucleotides are harder to add and therefore more likely to diffuse away before the polymerase can mistakenly add them.
Cell_Biology_Alberts_1127
Cell_Biology_Alberts
The next error-correcting reaction, known as exonucleolytic proofreading, takes place immediately after those rare instances in which an incorrect nucle otide is covalently added to the growing chain. DNA polymerase enzymes are unpaired or mispaired residues at the primer terminus, continuing until enough nucleotides have been removed to regenerate a correctly base-paired 3ʹ-OH ter minus that can prime DNA synthesis. In this way, DNA polymerase functions as a attached to DNA polymerase “self-correcting” enzyme that removes its own polymerization errors as it moves chews back to create a base- along the DNA (Figure 5–8 and Figure 5–9). paired 3˜-OH end on the primer
Cell_Biology_Alberts. The next error-correcting reaction, known as exonucleolytic proofreading, takes place immediately after those rare instances in which an incorrect nucle otide is covalently added to the growing chain. DNA polymerase enzymes are unpaired or mispaired residues at the primer terminus, continuing until enough nucleotides have been removed to regenerate a correctly base-paired 3ʹ-OH ter minus that can prime DNA synthesis. In this way, DNA polymerase functions as a attached to DNA polymerase “self-correcting” enzyme that removes its own polymerization errors as it moves chews back to create a base- along the DNA (Figure 5–8 and Figure 5–9). paired 3˜-OH end on the primer
Cell_Biology_Alberts_1128
Cell_Biology_Alberts
paired 3˜-OH end on the primer The self-correcting properties of the DNA polymerase depend on its require ment for a perfectly base-paired primer terminus, and it is apparently not possible for such an enzyme to start synthesis de novo, without an existing primer. By contrast, the RNA polymerase enzymes involved in gene transcription do not need such an efficient exonucleolytic proofreading mechanism: errors in making RNA are not passed on to the next generation, and the occasional defective RNA molecule that is produced has no long-term significance. RNA polymerases are thus able to start new polynucleotide chains without a primer. DNA polymerase resumes the process of adding nucleotides to the base-paired 3˜-OH end of the primer strand
Cell_Biology_Alberts. paired 3˜-OH end on the primer The self-correcting properties of the DNA polymerase depend on its require ment for a perfectly base-paired primer terminus, and it is apparently not possible for such an enzyme to start synthesis de novo, without an existing primer. By contrast, the RNA polymerase enzymes involved in gene transcription do not need such an efficient exonucleolytic proofreading mechanism: errors in making RNA are not passed on to the next generation, and the occasional defective RNA molecule that is produced has no long-term significance. RNA polymerases are thus able to start new polynucleotide chains without a primer. DNA polymerase resumes the process of adding nucleotides to the base-paired 3˜-OH end of the primer strand
Cell_Biology_Alberts_1129
Cell_Biology_Alberts
DNA polymerase resumes the process of adding nucleotides to the base-paired 3˜-OH end of the primer strand Figure 5–8 Exonucleolytic proofreading by DNA polymerase during DNA replication. In this example, a C is accidentally incorporated at the growing 3ʹ-OH end of a DNA chain. The part of DNA polymerase that removes the misincorporated nucleotide is a specialized member of a large class of enzymes, known as exonucleases, that cleave nucleotides one at a time from the ends of polynucleotides. Figure 5–9 Editing by DNA polymerase. DNA polymerase complexed with the DNA template in the polymerizing mode (left) and the editing mode (right). The catalytic sites for the exonucleolytic (E) and the polymerization (P) reactions are indicated. In the editing mode, the newly synthesized DNA transiently unpairs from the template and enters the editing site where the most recently added nucleotide is catalytically removed.
Cell_Biology_Alberts. DNA polymerase resumes the process of adding nucleotides to the base-paired 3˜-OH end of the primer strand Figure 5–8 Exonucleolytic proofreading by DNA polymerase during DNA replication. In this example, a C is accidentally incorporated at the growing 3ʹ-OH end of a DNA chain. The part of DNA polymerase that removes the misincorporated nucleotide is a specialized member of a large class of enzymes, known as exonucleases, that cleave nucleotides one at a time from the ends of polynucleotides. Figure 5–9 Editing by DNA polymerase. DNA polymerase complexed with the DNA template in the polymerizing mode (left) and the editing mode (right). The catalytic sites for the exonucleolytic (E) and the polymerization (P) reactions are indicated. In the editing mode, the newly synthesized DNA transiently unpairs from the template and enters the editing site where the most recently added nucleotide is catalytically removed.
Cell_Biology_Alberts_1130
Cell_Biology_Alberts
There is an error frequency of about one mistake for every 104 polymerization events both in RNA synthesis and in the separate process of translating mRNA sequences into protein sequences. This error rate is over 100,000 times greater than that in DNA replication, where, as we have seen, a series of proofreading processes makes the process unusually accurate (Table 5–1). Only DNA Replication in the 5ʹ-to-3ʹ Direction Allows Efficient Error Correction
Cell_Biology_Alberts. There is an error frequency of about one mistake for every 104 polymerization events both in RNA synthesis and in the separate process of translating mRNA sequences into protein sequences. This error rate is over 100,000 times greater than that in DNA replication, where, as we have seen, a series of proofreading processes makes the process unusually accurate (Table 5–1). Only DNA Replication in the 5ʹ-to-3ʹ Direction Allows Efficient Error Correction
Cell_Biology_Alberts_1131
Cell_Biology_Alberts
Only DNA Replication in the 5ʹ-to-3ʹ Direction Allows Efficient Error Correction The need for accuracy probably explains why DNA replication occurs only in the 5ʹ-to-3ʹ direction. If there were a DNA polymerase that added deoxyribonucleoside triphosphates in the 3ʹ-to-5ʹ direction, the growing 5ʹ end of the chain, rather than the incoming mononucleotide, would have to provide the activating triphosphate needed for the covalent linkage. In this case, the mistakes in polymerization could not be simply hydrolyzed away, because the bare 5ʹ end of the chain thus created would immediately terminate DNA synthesis (see Figure 5–3). It is therefore possible to correct a mismatched base only if it has been added to the 3ʹ end of a DNA chain. Although the backstitching mechanism for DNA replication seems complex, it preserves the 5ʹ-to-3ʹ direction of polymerization that is required for exonucleolytic proofreading.
Cell_Biology_Alberts. Only DNA Replication in the 5ʹ-to-3ʹ Direction Allows Efficient Error Correction The need for accuracy probably explains why DNA replication occurs only in the 5ʹ-to-3ʹ direction. If there were a DNA polymerase that added deoxyribonucleoside triphosphates in the 3ʹ-to-5ʹ direction, the growing 5ʹ end of the chain, rather than the incoming mononucleotide, would have to provide the activating triphosphate needed for the covalent linkage. In this case, the mistakes in polymerization could not be simply hydrolyzed away, because the bare 5ʹ end of the chain thus created would immediately terminate DNA synthesis (see Figure 5–3). It is therefore possible to correct a mismatched base only if it has been added to the 3ʹ end of a DNA chain. Although the backstitching mechanism for DNA replication seems complex, it preserves the 5ʹ-to-3ʹ direction of polymerization that is required for exonucleolytic proofreading.
Cell_Biology_Alberts_1132
Cell_Biology_Alberts
Despite these safeguards against DNA replication errors, DNA polymerases occasionally make mistakes. However, as we shall see later, cells have yet another Figure 5–10 RNA primer synthesis. A schematic view of the reaction catalyzed by DNA primase, the enzyme that synthesizes the short RNA primers made on the lagging strand using DNA as a template. Unlike DNA polymerase, this enzyme can start a new polynucleotide chain by joining two nucleoside triphosphates together. The primase synthesizes a short polynucleotide in the 5ʹ-to-3ʹ direction and then stops, making the 3ʹ end of this primer available for the DNA polymerase. chance to correct these errors by a process called strand-directed mismatch repair. Before discussing this mechanism, however, we describe the other types of proteins that function at the replication fork. A Special Nucleotide-Polymerizing Enzyme Synthesizes Short RNA Primer Molecules on the Lagging Strand
Cell_Biology_Alberts. Despite these safeguards against DNA replication errors, DNA polymerases occasionally make mistakes. However, as we shall see later, cells have yet another Figure 5–10 RNA primer synthesis. A schematic view of the reaction catalyzed by DNA primase, the enzyme that synthesizes the short RNA primers made on the lagging strand using DNA as a template. Unlike DNA polymerase, this enzyme can start a new polynucleotide chain by joining two nucleoside triphosphates together. The primase synthesizes a short polynucleotide in the 5ʹ-to-3ʹ direction and then stops, making the 3ʹ end of this primer available for the DNA polymerase. chance to correct these errors by a process called strand-directed mismatch repair. Before discussing this mechanism, however, we describe the other types of proteins that function at the replication fork. A Special Nucleotide-Polymerizing Enzyme Synthesizes Short RNA Primer Molecules on the Lagging Strand
Cell_Biology_Alberts_1133
Cell_Biology_Alberts
A Special Nucleotide-Polymerizing Enzyme Synthesizes Short RNA Primer Molecules on the Lagging Strand For the leading strand, a primer is needed only at the start of replication: once a replication fork is established, the DNA polymerase is continuously presented with a base-paired chain end on which to add new nucleotides. On the lagging side of the fork, however, each time the DNA polymerase completes a short DNA Okazaki fragment (which takes a few seconds), it must start synthesizing a completely new fragment at a site further along the template strand (see Figure 5–7). A special mechanism produces the base-paired primer strand required by the DNA polymerase molecules. The mechanism depends on an enzyme called DNA primase, which uses ribonucleoside triphosphates to synthesize short RNA primers on the lagging strand (Figure 5–10). In eukaryotes, these primers are about 10 nucleotides long and are made at intervals of 100–200 nucleotides on the lagging strand.
Cell_Biology_Alberts. A Special Nucleotide-Polymerizing Enzyme Synthesizes Short RNA Primer Molecules on the Lagging Strand For the leading strand, a primer is needed only at the start of replication: once a replication fork is established, the DNA polymerase is continuously presented with a base-paired chain end on which to add new nucleotides. On the lagging side of the fork, however, each time the DNA polymerase completes a short DNA Okazaki fragment (which takes a few seconds), it must start synthesizing a completely new fragment at a site further along the template strand (see Figure 5–7). A special mechanism produces the base-paired primer strand required by the DNA polymerase molecules. The mechanism depends on an enzyme called DNA primase, which uses ribonucleoside triphosphates to synthesize short RNA primers on the lagging strand (Figure 5–10). In eukaryotes, these primers are about 10 nucleotides long and are made at intervals of 100–200 nucleotides on the lagging strand.
Cell_Biology_Alberts_1134
Cell_Biology_Alberts
The chemical structure of RNA was introduced in Chapter 1 and is described in detail in Chapter 6. Here, we note only that RNA is very similar in structure to DNA. A strand of RNA can form base pairs with a strand of DNA, generating a DNA–RNA hybrid double helix if the two nucleotide sequences are complementary. Thus, the same templating principle used for DNA synthesis guides the synthesis of RNA primers. Because an RNA primer contains a properly base-paired nucleotide with a 3ʹ-OH group at one end, it can be elongated by the DNA polymerase at this end to begin an Okazaki fragment. The synthesis of each Okazaki fragment ends when this DNA polymerase runs into the RNA primer attached to the 5ʹ end of the previous fragment. To produce a continuous DNA chain from the many DNA fragments made on the lagging strand, a special DNA repair system acts quickly to erase the old RNA primer and replace it with DNA. An enzyme called DNA ligase then joins the 3ʹend of the new DNA fragment to the
Cell_Biology_Alberts. The chemical structure of RNA was introduced in Chapter 1 and is described in detail in Chapter 6. Here, we note only that RNA is very similar in structure to DNA. A strand of RNA can form base pairs with a strand of DNA, generating a DNA–RNA hybrid double helix if the two nucleotide sequences are complementary. Thus, the same templating principle used for DNA synthesis guides the synthesis of RNA primers. Because an RNA primer contains a properly base-paired nucleotide with a 3ʹ-OH group at one end, it can be elongated by the DNA polymerase at this end to begin an Okazaki fragment. The synthesis of each Okazaki fragment ends when this DNA polymerase runs into the RNA primer attached to the 5ʹ end of the previous fragment. To produce a continuous DNA chain from the many DNA fragments made on the lagging strand, a special DNA repair system acts quickly to erase the old RNA primer and replace it with DNA. An enzyme called DNA ligase then joins the 3ʹend of the new DNA fragment to the
Cell_Biology_Alberts_1135
Cell_Biology_Alberts
on the lagging strand, a special DNA repair system acts quickly to erase the old RNA primer and replace it with DNA. An enzyme called DNA ligase then joins the 3ʹend of the new DNA fragment to the 5ʹend of the previous one to complete the process (Figure 5–11 and Figure 5–12).
Cell_Biology_Alberts. on the lagging strand, a special DNA repair system acts quickly to erase the old RNA primer and replace it with DNA. An enzyme called DNA ligase then joins the 3ʹend of the new DNA fragment to the 5ʹend of the previous one to complete the process (Figure 5–11 and Figure 5–12).
Cell_Biology_Alberts_1136
Cell_Biology_Alberts
Why might an erasable RNA primer be preferred to a DNA primer that would not need to be erased? The argument that a self-correcting polymerase cannot start chains de novo also implies the converse: an enzyme that starts chains anew cannot be efficient at self-correction. Thus, any enzyme that primes the synthesis of Okazaki fragments will of necessity make a relatively inaccurate copy (at least one error in 105). Even if the copies retained in the final product constituted as little as 5% of the total genome (for example, 10 nucleotides per 200-nucleotide DNA fragment), the resulting increase in the overall mutation rate would be enormous. It therefore seems likely that the use of RNA rather than DNA for priming brings a powerful advantage to the cell: the ribonucleotides in the primer automatically mark these sequences as “suspect copy” to be efficiently removed and replaced.
Cell_Biology_Alberts. Why might an erasable RNA primer be preferred to a DNA primer that would not need to be erased? The argument that a self-correcting polymerase cannot start chains de novo also implies the converse: an enzyme that starts chains anew cannot be efficient at self-correction. Thus, any enzyme that primes the synthesis of Okazaki fragments will of necessity make a relatively inaccurate copy (at least one error in 105). Even if the copies retained in the final product constituted as little as 5% of the total genome (for example, 10 nucleotides per 200-nucleotide DNA fragment), the resulting increase in the overall mutation rate would be enormous. It therefore seems likely that the use of RNA rather than DNA for priming brings a powerful advantage to the cell: the ribonucleotides in the primer automatically mark these sequences as “suspect copy” to be efficiently removed and replaced.
Cell_Biology_Alberts_1137
Cell_Biology_Alberts
Figure 5–11 The synthesis of one of many DNA fragments on the lagging strand. In eukaryotes, RNA primers are made at intervals spaced by about 200 nucleotides on the lagging strand, and each RNA primer is approximately 10 nucleotides long. This primer is erased by a special DNA repair enzyme (an RNAse H) that recognizes an RNA strand in an RNA/DNA helix and fragments it; this leaves gaps that are filled in by DNA polymerase and DNA ligase. DNA polymerase adds to new RNA primer to start new Okazaki fragment sealing by DNA ligase joins new Okazaki fragment to the growing chain Special Proteins Help to Open Up the DNA Double Helix in Front of the Replication Fork
Cell_Biology_Alberts. Figure 5–11 The synthesis of one of many DNA fragments on the lagging strand. In eukaryotes, RNA primers are made at intervals spaced by about 200 nucleotides on the lagging strand, and each RNA primer is approximately 10 nucleotides long. This primer is erased by a special DNA repair enzyme (an RNAse H) that recognizes an RNA strand in an RNA/DNA helix and fragments it; this leaves gaps that are filled in by DNA polymerase and DNA ligase. DNA polymerase adds to new RNA primer to start new Okazaki fragment sealing by DNA ligase joins new Okazaki fragment to the growing chain Special Proteins Help to Open Up the DNA Double Helix in Front of the Replication Fork
Cell_Biology_Alberts_1138
Cell_Biology_Alberts
Special Proteins Help to Open Up the DNA Double Helix in Front of the Replication Fork For DNA synthesis to proceed, the DNA double helix must be opened up (“melted”) ahead of the replication fork so that the incoming deoxyribonucleoside triphosphates can form base pairs with the template strands. However, the DNA double helix is very stable under physiological conditions; the base pairs are locked in place so strongly that it requires temperatures approaching that of boiling water to separate the two strands in a test tube. For this reason, two additional types of replication proteins—DNA helicases and single-strand DNA-binding proteins—are needed to open the double helix and provide the appropriate single-strand DNA templates for the DNA polymerase to copy.
Cell_Biology_Alberts. Special Proteins Help to Open Up the DNA Double Helix in Front of the Replication Fork For DNA synthesis to proceed, the DNA double helix must be opened up (“melted”) ahead of the replication fork so that the incoming deoxyribonucleoside triphosphates can form base pairs with the template strands. However, the DNA double helix is very stable under physiological conditions; the base pairs are locked in place so strongly that it requires temperatures approaching that of boiling water to separate the two strands in a test tube. For this reason, two additional types of replication proteins—DNA helicases and single-strand DNA-binding proteins—are needed to open the double helix and provide the appropriate single-strand DNA templates for the DNA polymerase to copy.
Cell_Biology_Alberts_1139
Cell_Biology_Alberts
DNA helicases were first isolated as proteins that hydrolyze ATP when they are bound to single strands of DNA. As described in Chapter 3, the hydrolysis of ATP can change the shape of a protein molecule in a cyclical manner that allows the protein to perform mechanical work. DNA helicases use this principle to propel themselves rapidly along a DNA single strand. When they encounter a region of double helix, they continue to move along their strand, thereby prying apart the helix at rates of up to 1000 nucleotide pairs per second (Figure 5–13 and Figure 5–14). The two strands of DNA have opposite polarities, and, in principle, a helicase
Cell_Biology_Alberts. DNA helicases were first isolated as proteins that hydrolyze ATP when they are bound to single strands of DNA. As described in Chapter 3, the hydrolysis of ATP can change the shape of a protein molecule in a cyclical manner that allows the protein to perform mechanical work. DNA helicases use this principle to propel themselves rapidly along a DNA single strand. When they encounter a region of double helix, they continue to move along their strand, thereby prying apart the helix at rates of up to 1000 nucleotide pairs per second (Figure 5–13 and Figure 5–14). The two strands of DNA have opposite polarities, and, in principle, a helicase
Cell_Biology_Alberts_1140
Cell_Biology_Alberts
The two strands of DNA have opposite polarities, and, in principle, a helicase Figure 5–12 The reaction catalyzed by DNA ligase. This enzyme seals a broken phosphodiester bond. As shown, DNA ligase uses a molecule of ATP to activate the 5ʹ end at the nick (step 1) before forming the new bond (step 2). In this way, the energetically unfavorable nick-sealing reaction is driven by being coupled to the energetically favorable process of ATP hydrolysis. could unwind the DNA double helix by moving in the 5ʹ-to-3ʹ direction along one 5˜ 3˜ strand or in the 3ʹ-to-5ʹ direction along the other. In fact, both types of DNA helicase exist. In the best-understood replication systems in bacteria, a helicase moving 5ʹ to 3ʹ along the lagging-strand template appears to have the predominant role, for reasons that will become clear shortly.
Cell_Biology_Alberts. The two strands of DNA have opposite polarities, and, in principle, a helicase Figure 5–12 The reaction catalyzed by DNA ligase. This enzyme seals a broken phosphodiester bond. As shown, DNA ligase uses a molecule of ATP to activate the 5ʹ end at the nick (step 1) before forming the new bond (step 2). In this way, the energetically unfavorable nick-sealing reaction is driven by being coupled to the energetically favorable process of ATP hydrolysis. could unwind the DNA double helix by moving in the 5ʹ-to-3ʹ direction along one 5˜ 3˜ strand or in the 3ʹ-to-5ʹ direction along the other. In fact, both types of DNA helicase exist. In the best-understood replication systems in bacteria, a helicase moving 5ʹ to 3ʹ along the lagging-strand template appears to have the predominant role, for reasons that will become clear shortly.
Cell_Biology_Alberts_1141
Cell_Biology_Alberts
Single-strand DNA-binding (SSB) proteins, also called helix-destabilizing proteins, bind tightly and cooperatively to exposed single-strand DNA without covering the bases, which therefore remain available as templates. These proteins are unable to open a long DNA helix directly, but they aid helicases by stabilizing the unwound, single-strand conformation. In addition, through cooperative binding, they coat and straighten out the regions of single-strand DNA, which occur routinely in the lagging-strand template, thereby preventing the formation of the short hairpin helices that readily form in single-strand DNA (Figure 5–15 and Figure 5–16). If not removed, these hairpin helices can impede the DNA synthesis catalyzed by DNA polymerase. A Sliding Ring Holds a Moving DNA Polymerase Onto the DNA
Cell_Biology_Alberts. Single-strand DNA-binding (SSB) proteins, also called helix-destabilizing proteins, bind tightly and cooperatively to exposed single-strand DNA without covering the bases, which therefore remain available as templates. These proteins are unable to open a long DNA helix directly, but they aid helicases by stabilizing the unwound, single-strand conformation. In addition, through cooperative binding, they coat and straighten out the regions of single-strand DNA, which occur routinely in the lagging-strand template, thereby preventing the formation of the short hairpin helices that readily form in single-strand DNA (Figure 5–15 and Figure 5–16). If not removed, these hairpin helices can impede the DNA synthesis catalyzed by DNA polymerase. A Sliding Ring Holds a Moving DNA Polymerase Onto the DNA
Cell_Biology_Alberts_1142
Cell_Biology_Alberts
A Sliding Ring Holds a Moving DNA Polymerase Onto the DNA On their own, most DNA polymerase molecules will synthesize only a short string of nucleotides before falling off the DNA template. The tendency to dissociate quickly from a DNA molecule allows a DNA polymerase molecule that has just Figure 5–13 An assay for DNA helicase enzymes. A short DNA fragment is annealed to a long DNA single strand to form a region of DNA double helix. The double helix is melted as the helicase runs along the DNA single strand, releasing the short DNA fragment in a reaction that requires the presence of both the helicase protein and ATP. The rapid stepwise movement of the helicase is powered by its ATP hydrolysis (shown schematically in Figure 3–75A). As indicated, many DNA helicases are composed of six subunits.
Cell_Biology_Alberts. A Sliding Ring Holds a Moving DNA Polymerase Onto the DNA On their own, most DNA polymerase molecules will synthesize only a short string of nucleotides before falling off the DNA template. The tendency to dissociate quickly from a DNA molecule allows a DNA polymerase molecule that has just Figure 5–13 An assay for DNA helicase enzymes. A short DNA fragment is annealed to a long DNA single strand to form a region of DNA double helix. The double helix is melted as the helicase runs along the DNA single strand, releasing the short DNA fragment in a reaction that requires the presence of both the helicase protein and ATP. The rapid stepwise movement of the helicase is powered by its ATP hydrolysis (shown schematically in Figure 3–75A). As indicated, many DNA helicases are composed of six subunits.
Cell_Biology_Alberts_1143
Cell_Biology_Alberts
Figure 5–14 The structure of a DNA helicase. (A) Diagram of the protein 5˜ as a hexameric ring drawn to scale with a replication fork. (B) Detailed structure of the bacteriophage T7 replicative helicase, as determined by x-ray diffraction. Six identical subunits bind and hydrolyze ATP in an ordered fashion to propel this molecule, like a rotary engine, along a DNA single strand that passes through the central hole. Red indicates bound ATP molecules in the structure (Movie 5.2). (PDB code: 1E0J.) finished synthesizing one Okazaki fragment on the lagging strand to be recycled quickly, so as to begin the synthesis of the next Okazaki fragment on the same strand. This rapid dissociation, however, would make it difficult for the polymerase to synthesize the long DNA strands produced at a replication fork were it not for an accessory protein (called PCNA in eukaryotes) that functions as a regulated sliding clamp. This clamp keeps the polymerase firmly on the DNA when it is moving, but
Cell_Biology_Alberts. Figure 5–14 The structure of a DNA helicase. (A) Diagram of the protein 5˜ as a hexameric ring drawn to scale with a replication fork. (B) Detailed structure of the bacteriophage T7 replicative helicase, as determined by x-ray diffraction. Six identical subunits bind and hydrolyze ATP in an ordered fashion to propel this molecule, like a rotary engine, along a DNA single strand that passes through the central hole. Red indicates bound ATP molecules in the structure (Movie 5.2). (PDB code: 1E0J.) finished synthesizing one Okazaki fragment on the lagging strand to be recycled quickly, so as to begin the synthesis of the next Okazaki fragment on the same strand. This rapid dissociation, however, would make it difficult for the polymerase to synthesize the long DNA strands produced at a replication fork were it not for an accessory protein (called PCNA in eukaryotes) that functions as a regulated sliding clamp. This clamp keeps the polymerase firmly on the DNA when it is moving, but
Cell_Biology_Alberts_1144
Cell_Biology_Alberts
replication fork were it not for an accessory protein (called PCNA in eukaryotes) that functions as a regulated sliding clamp. This clamp keeps the polymerase firmly on the DNA when it is moving, but releases it as soon as the polymerase runs into a double-strand region of DNA.
Cell_Biology_Alberts. replication fork were it not for an accessory protein (called PCNA in eukaryotes) that functions as a regulated sliding clamp. This clamp keeps the polymerase firmly on the DNA when it is moving, but releases it as soon as the polymerase runs into a double-strand region of DNA.
Cell_Biology_Alberts_1145
Cell_Biology_Alberts
How can a sliding clamp prevent the polymerase from dissociating without at the same time impeding the polymerase’s rapid movement along the DNA molecule? The three-dimensional structure of the clamp protein, determined by x-ray diffraction, revealed it to be a large ring around the DNA double helix. One face of the ring binds to the back of the DNA polymerase, and the whole ring slides freely along the DNA as the polymerase moves. The assembly of the clamp around the DNA requires ATP hydrolysis by a special protein complex, the clamp loader, which hydrolyzes ATP as it loads the clamp on to a primer–template junction (Figure 5–17).
Cell_Biology_Alberts. How can a sliding clamp prevent the polymerase from dissociating without at the same time impeding the polymerase’s rapid movement along the DNA molecule? The three-dimensional structure of the clamp protein, determined by x-ray diffraction, revealed it to be a large ring around the DNA double helix. One face of the ring binds to the back of the DNA polymerase, and the whole ring slides freely along the DNA as the polymerase moves. The assembly of the clamp around the DNA requires ATP hydrolysis by a special protein complex, the clamp loader, which hydrolyzes ATP as it loads the clamp on to a primer–template junction (Figure 5–17).
Cell_Biology_Alberts_1146
Cell_Biology_Alberts
On the leading-strand template, the moving DNA polymerase is tightly bound to the clamp, and the two remain associated for a very long time. The DNA polymerase on the lagging-strand template also makes use of the clamp, but each time the polymerase reaches the 5ʹ end of the preceding Okazaki fragment, the polymerase releases itself from the clamp and dissociates from the template. This polymerase molecule then associates with a new clamp that is assembled on the RNA primer of the next Okazaki fragment. single-stranded region of DNA template with short regions of base-paired “hairpins”
Cell_Biology_Alberts. On the leading-strand template, the moving DNA polymerase is tightly bound to the clamp, and the two remain associated for a very long time. The DNA polymerase on the lagging-strand template also makes use of the clamp, but each time the polymerase reaches the 5ʹ end of the preceding Okazaki fragment, the polymerase releases itself from the clamp and dissociates from the template. This polymerase molecule then associates with a new clamp that is assembled on the RNA primer of the next Okazaki fragment. single-stranded region of DNA template with short regions of base-paired “hairpins”
Cell_Biology_Alberts_1147
Cell_Biology_Alberts
single-stranded region of DNA template with short regions of base-paired “hairpins” Figure 5–15 The effect of single-strand DNA-binding proteins (SSb proteins) on the structure of single-strand DNA. Because each protein molecule prefers to bind next to a previously bound molecule, long rows of this protein form on a DNA single strand. This cooperative binding straightens out the DNA template and facilitates the DNA polymerization process. The “hairpin helices” shown in the bare, single-strand DNA result from a chance matching of short regions of complementary nucleotide sequence; they are similar to the short helices that typically form in RNA molecules (see Figure 1–6).
Cell_Biology_Alberts. single-stranded region of DNA template with short regions of base-paired “hairpins” Figure 5–15 The effect of single-strand DNA-binding proteins (SSb proteins) on the structure of single-strand DNA. Because each protein molecule prefers to bind next to a previously bound molecule, long rows of this protein form on a DNA single strand. This cooperative binding straightens out the DNA template and facilitates the DNA polymerization process. The “hairpin helices” shown in the bare, single-strand DNA result from a chance matching of short regions of complementary nucleotide sequence; they are similar to the short helices that typically form in RNA molecules (see Figure 1–6).
Cell_Biology_Alberts_1148
Cell_Biology_Alberts
Figure 5–16 Human single-strand binding protein bound to DNA. (A) Front view of the two DNA-binding domains of the protein (called RPA) which cover a total of eight nucleotides. Note that the DNA bases remain exposed in this protein–DNA complex. (B) Diagram showing the three-dimensional structure, with the DNA strand (orange) viewed end-on. (PDB code: 1JMC.)
Cell_Biology_Alberts. Figure 5–16 Human single-strand binding protein bound to DNA. (A) Front view of the two DNA-binding domains of the protein (called RPA) which cover a total of eight nucleotides. Note that the DNA bases remain exposed in this protein–DNA complex. (B) Diagram showing the three-dimensional structure, with the DNA strand (orange) viewed end-on. (PDB code: 1JMC.)
Cell_Biology_Alberts_1149
Cell_Biology_Alberts
Figure 5–17 The regulated sliding clamp that holds DNA polymerase on the DNA. (A) The structure of the clamp protein from E. coli, as determined by x-ray crystallography, with a DNA helix added to indicate how the protein fits around DNA (Movie 5.3). (B) Schematic illustration showing how the clamp (with red and yellow subunits) is loaded onto DNA to serve as a tether for a moving DNA polymerase molecule. The structure of the clamp loader (dark green) resembles a screw nut, (A) (B) sliding clamp clamp loader 5˜ATPADPPi ATPATP+ DNA + DNA polymerase 3˜5˜5˜3˜3˜RECYCLING OF RELEASED CLAMP LOADER ATP BINDING TO CLAMP LOADER OPENS SLIDING CLAMP DNA ENGAGED IN CLAMP ATP HYDROLYSIS LOCKS SLIDING CLAMP AROUND DNA AND RELEASES CLAMP LOADER DNA POLYMERASE BINDS TO SLIDING CLAMP + with its threads matching the grooves of double-stranded DNA. The loader binds to a free clamp molecule, forcing a gap in its ring of subunits so that this ring is able to slip around DNA. The clamp loader, thanks to
Cell_Biology_Alberts. Figure 5–17 The regulated sliding clamp that holds DNA polymerase on the DNA. (A) The structure of the clamp protein from E. coli, as determined by x-ray crystallography, with a DNA helix added to indicate how the protein fits around DNA (Movie 5.3). (B) Schematic illustration showing how the clamp (with red and yellow subunits) is loaded onto DNA to serve as a tether for a moving DNA polymerase molecule. The structure of the clamp loader (dark green) resembles a screw nut, (A) (B) sliding clamp clamp loader 5˜ATPADPPi ATPATP+ DNA + DNA polymerase 3˜5˜5˜3˜3˜RECYCLING OF RELEASED CLAMP LOADER ATP BINDING TO CLAMP LOADER OPENS SLIDING CLAMP DNA ENGAGED IN CLAMP ATP HYDROLYSIS LOCKS SLIDING CLAMP AROUND DNA AND RELEASES CLAMP LOADER DNA POLYMERASE BINDS TO SLIDING CLAMP + with its threads matching the grooves of double-stranded DNA. The loader binds to a free clamp molecule, forcing a gap in its ring of subunits so that this ring is able to slip around DNA. The clamp loader, thanks to
Cell_Biology_Alberts_1150
Cell_Biology_Alberts
matching the grooves of double-stranded DNA. The loader binds to a free clamp molecule, forcing a gap in its ring of subunits so that this ring is able to slip around DNA. The clamp loader, thanks to its screw-nut structure, recognises the region of DNA that is double-stranded and latches onto it, tightening around the complex of a template strand with a freshly synthesized elongating (primer) strand. It carries the clamp along this double-stranded region until it encounters the 3ʹ end of the primer, at which point the loader hydrolyzes ATP and releases the clamp, allowing it to close around the DNA and bind to DNA polymerase. In the simplified reaction shown here, the clamp loader dissociates into solution once the clamp has been assembled. At a true replication fork, the clamp loader remains close to the polymerase so that, on the lagging strand, it is ready to assemble a new clamp at the start of each new Okazaki fragment (see Figure 5–18). (A, from X.P. Kong et al., Cell
Cell_Biology_Alberts. matching the grooves of double-stranded DNA. The loader binds to a free clamp molecule, forcing a gap in its ring of subunits so that this ring is able to slip around DNA. The clamp loader, thanks to its screw-nut structure, recognises the region of DNA that is double-stranded and latches onto it, tightening around the complex of a template strand with a freshly synthesized elongating (primer) strand. It carries the clamp along this double-stranded region until it encounters the 3ʹ end of the primer, at which point the loader hydrolyzes ATP and releases the clamp, allowing it to close around the DNA and bind to DNA polymerase. In the simplified reaction shown here, the clamp loader dissociates into solution once the clamp has been assembled. At a true replication fork, the clamp loader remains close to the polymerase so that, on the lagging strand, it is ready to assemble a new clamp at the start of each new Okazaki fragment (see Figure 5–18). (A, from X.P. Kong et al., Cell
Cell_Biology_Alberts_1151
Cell_Biology_Alberts
remains close to the polymerase so that, on the lagging strand, it is ready to assemble a new clamp at the start of each new Okazaki fragment (see Figure 5–18). (A, from X.P. Kong et al., Cell 69:425–437, 1992. With permission from Elsevier; B, adapted from B.A. Kelch et al., Science 334:1675–1680, 2011. With permission from AAAS. PDB code: 3BEP.)
Cell_Biology_Alberts. remains close to the polymerase so that, on the lagging strand, it is ready to assemble a new clamp at the start of each new Okazaki fragment (see Figure 5–18). (A, from X.P. Kong et al., Cell 69:425–437, 1992. With permission from Elsevier; B, adapted from B.A. Kelch et al., Science 334:1675–1680, 2011. With permission from AAAS. PDB code: 3BEP.)
Cell_Biology_Alberts_1152
Cell_Biology_Alberts
The Proteins at a Replication Fork Cooperate to Form a Replication Machine Although we have discussed DNA replication as though it were performed by a mix of proteins all acting independently, in reality most of the proteins are held together in a large and orderly multienzyme complex that rapidly synthesizes DNA. This complex can be likened to a tiny sewing machine composed of protein parts and powered by nucleoside triphosphate hydrolysis. Like a sewing machine, the replication complex probably remains stationary with respect to its immediate surroundings; the DNA can be thought of as a long strip of cloth being rapidly threaded through it. Although the replication complex has been most intensively studied in E. coli and several of its viruses, a very similar complex also operates in eukaryotes, as we see below.
Cell_Biology_Alberts. The Proteins at a Replication Fork Cooperate to Form a Replication Machine Although we have discussed DNA replication as though it were performed by a mix of proteins all acting independently, in reality most of the proteins are held together in a large and orderly multienzyme complex that rapidly synthesizes DNA. This complex can be likened to a tiny sewing machine composed of protein parts and powered by nucleoside triphosphate hydrolysis. Like a sewing machine, the replication complex probably remains stationary with respect to its immediate surroundings; the DNA can be thought of as a long strip of cloth being rapidly threaded through it. Although the replication complex has been most intensively studied in E. coli and several of its viruses, a very similar complex also operates in eukaryotes, as we see below.
Cell_Biology_Alberts_1153
Cell_Biology_Alberts
Figure 5–18 summarizes the functions of the subunits of the replication machine. At the front of the replication fork, DNA helicase opens the DNA helix. Two DNA polymerase molecules work at the fork, one on the leading strand and one on the lagging strand. Whereas the DNA polymerase molecule on the leading strand can operate in a continuous fashion, the DNA polymerase molecule on the lagging strand must restart at short intervals, using a short RNA primer made by a DNA primase molecule. The close association of all these protein components increases the efficiency of replication and is made possible by a folding back of the lagging strand as shown in Figure 5–18A. This arrangement also facilitates the loading of the polymerase clamp each time that an Okazaki fragment is synthesized: the clamp loader and the lagging-strand DNA polymerase molecule are kept in place as a part of the protein machine even when they detach from their DNA template. The replication proteins are thus linked
Cell_Biology_Alberts. Figure 5–18 summarizes the functions of the subunits of the replication machine. At the front of the replication fork, DNA helicase opens the DNA helix. Two DNA polymerase molecules work at the fork, one on the leading strand and one on the lagging strand. Whereas the DNA polymerase molecule on the leading strand can operate in a continuous fashion, the DNA polymerase molecule on the lagging strand must restart at short intervals, using a short RNA primer made by a DNA primase molecule. The close association of all these protein components increases the efficiency of replication and is made possible by a folding back of the lagging strand as shown in Figure 5–18A. This arrangement also facilitates the loading of the polymerase clamp each time that an Okazaki fragment is synthesized: the clamp loader and the lagging-strand DNA polymerase molecule are kept in place as a part of the protein machine even when they detach from their DNA template. The replication proteins are thus linked
Cell_Biology_Alberts_1154
Cell_Biology_Alberts
clamp loader and the lagging-strand DNA polymerase molecule are kept in place as a part of the protein machine even when they detach from their DNA template. The replication proteins are thus linked together into a single large
Cell_Biology_Alberts. clamp loader and the lagging-strand DNA polymerase molecule are kept in place as a part of the protein machine even when they detach from their DNA template. The replication proteins are thus linked together into a single large
Cell_Biology_Alberts_1155
Cell_Biology_Alberts
Figure 5–18 A bacterial replication fork. (A) This schematic diagram shows a current view of the arrangement of replication proteins at a replication fork when DNA is being synthesized. The lagging-strand DNA is folded to bring the lagging-strand DNA polymerase molecule into a complex with the leading-strand DNA polymerase molecule. This folding also brings the 3ʹ end of each completed Okazaki fragment close to the start site for the next Okazaki fragment. Because the lagging-strand DNA polymerase molecule remains bound to the rest of the replication proteins, it can be reused to synthesize successive Okazaki fragments. In this diagram, it is about to let go of its completed DNA fragment and move to the RNA primer that is just being synthesized. Additional proteins (not shown) help to hold the different protein components of the fork together, enabling them to function as a well-coordinated protein machine (Movie 5.4 and Movie 5.5). (B) An electron micrograph showing the replication
Cell_Biology_Alberts. Figure 5–18 A bacterial replication fork. (A) This schematic diagram shows a current view of the arrangement of replication proteins at a replication fork when DNA is being synthesized. The lagging-strand DNA is folded to bring the lagging-strand DNA polymerase molecule into a complex with the leading-strand DNA polymerase molecule. This folding also brings the 3ʹ end of each completed Okazaki fragment close to the start site for the next Okazaki fragment. Because the lagging-strand DNA polymerase molecule remains bound to the rest of the replication proteins, it can be reused to synthesize successive Okazaki fragments. In this diagram, it is about to let go of its completed DNA fragment and move to the RNA primer that is just being synthesized. Additional proteins (not shown) help to hold the different protein components of the fork together, enabling them to function as a well-coordinated protein machine (Movie 5.4 and Movie 5.5). (B) An electron micrograph showing the replication
Cell_Biology_Alberts_1156
Cell_Biology_Alberts
the different protein components of the fork together, enabling them to function as a well-coordinated protein machine (Movie 5.4 and Movie 5.5). (B) An electron micrograph showing the replication machine from the bacteriophage T4 as it moves along a template synthesizing DNA behind it. (C) An interpretation of the micrograph is given in the sketch: note especially the DNA loop on the lagging strand. Apparently, the replication proteins became partly detached from the very front of the replication fork during the preparation of this sample for electron microscopy. (B, courtesy of Jack Griffith; see P.D. Chastain et al., J. Biol. Chem. 278:21276–21285, 2003.) unit (total molecular mass >106 daltons), enabling DNA to be synthesized on both sides of the replication fork in a coordinated and efficient manner.
Cell_Biology_Alberts. the different protein components of the fork together, enabling them to function as a well-coordinated protein machine (Movie 5.4 and Movie 5.5). (B) An electron micrograph showing the replication machine from the bacteriophage T4 as it moves along a template synthesizing DNA behind it. (C) An interpretation of the micrograph is given in the sketch: note especially the DNA loop on the lagging strand. Apparently, the replication proteins became partly detached from the very front of the replication fork during the preparation of this sample for electron microscopy. (B, courtesy of Jack Griffith; see P.D. Chastain et al., J. Biol. Chem. 278:21276–21285, 2003.) unit (total molecular mass >106 daltons), enabling DNA to be synthesized on both sides of the replication fork in a coordinated and efficient manner.
Cell_Biology_Alberts_1157
Cell_Biology_Alberts
On the lagging strand, the DNA replication machine leaves behind a series of unsealed Okazaki fragments, which still contain the RNA that primed their synthesis at their 5ʹ ends. As discussed earlier, this RNA is removed and the resulting gap is filled in by DNA repair enzymes that operate behind the replication fork (see Figure 5–11). A Strand-Directed Mismatch Repair System Removes Replication Errors That Escape from the Replication Machine
Cell_Biology_Alberts. On the lagging strand, the DNA replication machine leaves behind a series of unsealed Okazaki fragments, which still contain the RNA that primed their synthesis at their 5ʹ ends. As discussed earlier, this RNA is removed and the resulting gap is filled in by DNA repair enzymes that operate behind the replication fork (see Figure 5–11). A Strand-Directed Mismatch Repair System Removes Replication Errors That Escape from the Replication Machine
Cell_Biology_Alberts_1158
Cell_Biology_Alberts
A Strand-Directed Mismatch Repair System Removes Replication Errors That Escape from the Replication Machine As stated previously, bacteria such as E. coli are capable of dividing once every 30 minutes, making it relatively easy to screen large populations to find a rare mutant cell that is altered in a specific process. One interesting class of mutants consists of those with alterations in so-called mutator genes, which greatly increase the rate of spontaneous mutation. Not surprisingly, one such mutant makes a defective form of the 3ʹ-to-5ʹ proofreading exonuclease that is a part of the DNA polymerase enzyme (see Figures 5–8 and 5–9). The mutant DNA polymerase no longer proofreads effectively, and many replication errors that would otherwise have been removed accumulate in the DNA.
Cell_Biology_Alberts. A Strand-Directed Mismatch Repair System Removes Replication Errors That Escape from the Replication Machine As stated previously, bacteria such as E. coli are capable of dividing once every 30 minutes, making it relatively easy to screen large populations to find a rare mutant cell that is altered in a specific process. One interesting class of mutants consists of those with alterations in so-called mutator genes, which greatly increase the rate of spontaneous mutation. Not surprisingly, one such mutant makes a defective form of the 3ʹ-to-5ʹ proofreading exonuclease that is a part of the DNA polymerase enzyme (see Figures 5–8 and 5–9). The mutant DNA polymerase no longer proofreads effectively, and many replication errors that would otherwise have been removed accumulate in the DNA.
Cell_Biology_Alberts_1159
Cell_Biology_Alberts
The study of other E. coli mutants exhibiting abnormally high mutation rates has uncovered a proofreading system that removes replication errors made by the polymerase that have been missed by the proofreading exonuclease. This strand-directed mismatch repair system detects the potential for distortion in the DNA helix from the misfit between noncomplementary base pairs. If the proofreading system simply recognized a mismatch in newly replicated DNA and randomly corrected one of the two mismatched nucleotides, it would mistakenly “correct” the original template strand to match the error exactly half the time, thereby failing to lower the overall error rate. To be effective, such a proofreading system must be able to distinguish and remove the mismatched nucleotide only on the newly synthesized strand, where the replication error occurred.
Cell_Biology_Alberts. The study of other E. coli mutants exhibiting abnormally high mutation rates has uncovered a proofreading system that removes replication errors made by the polymerase that have been missed by the proofreading exonuclease. This strand-directed mismatch repair system detects the potential for distortion in the DNA helix from the misfit between noncomplementary base pairs. If the proofreading system simply recognized a mismatch in newly replicated DNA and randomly corrected one of the two mismatched nucleotides, it would mistakenly “correct” the original template strand to match the error exactly half the time, thereby failing to lower the overall error rate. To be effective, such a proofreading system must be able to distinguish and remove the mismatched nucleotide only on the newly synthesized strand, where the replication error occurred.
Cell_Biology_Alberts_1160
Cell_Biology_Alberts
The strand-distinction mechanism used by the mismatch proofreading system in E. coli depends on the methylation of selected A residues in the DNA. Methyl groups are added to all A residues in the sequence GATC, but not until some time after the A has been incorporated into a newly synthesized DNA chain. As a result, the only GATC sequences that have not yet been methylated are in the new strands just behind a replication fork. The recognition of these unmethylated GATCs allows the new DNA strands to be transiently distinguished from old ones, as required if their mismatches are to be selectively removed. The three-step process involves recognition of a newly synthesized strand, excision of the portion containing the mismatch, and resynthesis of the excised segment using the old strand as a template. This strand-directed mismatch repair system reduces the number of errors made during DNA replication by an additional factor of 100 to 1000 (see Table 5–1, p. 244).
Cell_Biology_Alberts. The strand-distinction mechanism used by the mismatch proofreading system in E. coli depends on the methylation of selected A residues in the DNA. Methyl groups are added to all A residues in the sequence GATC, but not until some time after the A has been incorporated into a newly synthesized DNA chain. As a result, the only GATC sequences that have not yet been methylated are in the new strands just behind a replication fork. The recognition of these unmethylated GATCs allows the new DNA strands to be transiently distinguished from old ones, as required if their mismatches are to be selectively removed. The three-step process involves recognition of a newly synthesized strand, excision of the portion containing the mismatch, and resynthesis of the excised segment using the old strand as a template. This strand-directed mismatch repair system reduces the number of errors made during DNA replication by an additional factor of 100 to 1000 (see Table 5–1, p. 244).
Cell_Biology_Alberts_1161
Cell_Biology_Alberts
A similar mismatch proofreading system functions in eukaryotic cells but uses a different strategy to distinguish the new strand from the old (Figure 5–19). Newly synthesized lagging-strand DNA transiently contains nicks (before they are sealed by DNA ligase) and such nicks (also called single-strand breaks) provide the signal that directs the mismatch proofreading system to the appropriate strand. This strategy also requires that the newly synthesized DNA on the leading strand be transiently nicked; how this occurs is uncertain.
Cell_Biology_Alberts. A similar mismatch proofreading system functions in eukaryotic cells but uses a different strategy to distinguish the new strand from the old (Figure 5–19). Newly synthesized lagging-strand DNA transiently contains nicks (before they are sealed by DNA ligase) and such nicks (also called single-strand breaks) provide the signal that directs the mismatch proofreading system to the appropriate strand. This strategy also requires that the newly synthesized DNA on the leading strand be transiently nicked; how this occurs is uncertain.
Cell_Biology_Alberts_1162
Cell_Biology_Alberts
The importance of mismatch proofreading in humans is seen in individuals who inherit one defective copy of a mismatch repair gene (along with a functional gene on the other copy of the chromosome). These people have a marked predisposition for certain types of cancers. For example, in a type of colon cancer called hereditary nonpolyposis colon cancer (HNPCC), spontaneous mutation of the one functional gene produces a clone of somatic cells that, because they are deficient in mismatch proofreading, accumulate mutations unusually rapidly. Most cancers arise in cells that have accumulated multiple mutations (see pp. 1096–1097), and cells deficient in mismatch proofreading therefore have a greatly enhanced chance of becoming cancerous. Fortunately, most of us inherit two good copies of each gene that encodes a mismatch proofreading protein; this protects us, because it is highly unlikely for both copies to become mutated in the same cell.
Cell_Biology_Alberts. The importance of mismatch proofreading in humans is seen in individuals who inherit one defective copy of a mismatch repair gene (along with a functional gene on the other copy of the chromosome). These people have a marked predisposition for certain types of cancers. For example, in a type of colon cancer called hereditary nonpolyposis colon cancer (HNPCC), spontaneous mutation of the one functional gene produces a clone of somatic cells that, because they are deficient in mismatch proofreading, accumulate mutations unusually rapidly. Most cancers arise in cells that have accumulated multiple mutations (see pp. 1096–1097), and cells deficient in mismatch proofreading therefore have a greatly enhanced chance of becoming cancerous. Fortunately, most of us inherit two good copies of each gene that encodes a mismatch proofreading protein; this protects us, because it is highly unlikely for both copies to become mutated in the same cell.
Cell_Biology_Alberts_1163
Cell_Biology_Alberts
As a replication fork moves along double-strand DNA, it creates what has been called the “winding problem.” The two parental strands, which are wound around each other, must be unwound and separated for replication to occur. For every 10 nucleotide pairs replicated at the fork, one complete turn of the parental double helix must be unwound. In principle, this unwinding can be achieved by rapidly rotating the entire chromosome ahead of a moving fork; however, this is energetically highly unfavorable (particularly for long chromosomes) and, instead, the DNA in front of a replication fork becomes overwound (Figure 5–20). The over-winding, in turn, is continually relieved by proteins known as DNA topoisomerases. A DNA topoisomerase can be viewed as a reversible nuclease that adds itself covalently to a DNA backbone phosphate, thereby breaking a phosphodiester bond in a DNA strand. This reaction is reversible, and the phosphodiester bond re-forms as the protein leaves.
Cell_Biology_Alberts. As a replication fork moves along double-strand DNA, it creates what has been called the “winding problem.” The two parental strands, which are wound around each other, must be unwound and separated for replication to occur. For every 10 nucleotide pairs replicated at the fork, one complete turn of the parental double helix must be unwound. In principle, this unwinding can be achieved by rapidly rotating the entire chromosome ahead of a moving fork; however, this is energetically highly unfavorable (particularly for long chromosomes) and, instead, the DNA in front of a replication fork becomes overwound (Figure 5–20). The over-winding, in turn, is continually relieved by proteins known as DNA topoisomerases. A DNA topoisomerase can be viewed as a reversible nuclease that adds itself covalently to a DNA backbone phosphate, thereby breaking a phosphodiester bond in a DNA strand. This reaction is reversible, and the phosphodiester bond re-forms as the protein leaves.
Cell_Biology_Alberts_1164
Cell_Biology_Alberts
One type of topoisomerase, called topoisomerase I, produces a transient sin-gle-strand break; this break in the phosphodiester backbone allows the two sections of DNA helix on either side of the nick to rotate freely relative to each other, using the phosphodiester bond in the strand opposite the nick as a swivel point (Figure 5–21). Any tension in the DNA helix will drive this rotation in the direction that relieves the tension. As a result, DNA replication can occur with the rotation of only a short length of helix—the part just ahead of the fork. Because the covalent linkage that joins the DNA topoisomerase protein to a DNA phosphate retains rotate, torsional stress will build up
Cell_Biology_Alberts. One type of topoisomerase, called topoisomerase I, produces a transient sin-gle-strand break; this break in the phosphodiester backbone allows the two sections of DNA helix on either side of the nick to rotate freely relative to each other, using the phosphodiester bond in the strand opposite the nick as a swivel point (Figure 5–21). Any tension in the DNA helix will drive this rotation in the direction that relieves the tension. As a result, DNA replication can occur with the rotation of only a short length of helix—the part just ahead of the fork. Because the covalent linkage that joins the DNA topoisomerase protein to a DNA phosphate retains rotate, torsional stress will build up
Cell_Biology_Alberts_1165
Cell_Biology_Alberts
Figure 5–19 Strand-directed mismatch repair. (A) The two proteins shown are present in both bacteria and eukaryotic cells: MutS binds specifically to a mismatched base pair, while MutL scans the nearby DNA for a nick. Once MutL finds a nick, it triggers the degradation of the nicked strand all the way back through the mismatch. Because nicks are largely confined to newly replicated strands in eukaryotes, replication errors are selectively removed. In bacteria, an additional protein in the complex (MutH) nicks unmethylated (and therefore newly replicated) GATC sequences, thereby beginning the process illustrated here. In eukaryotes, MutL contains a DNA nicking activity that aids in the removal of the damaged strand.
Cell_Biology_Alberts. Figure 5–19 Strand-directed mismatch repair. (A) The two proteins shown are present in both bacteria and eukaryotic cells: MutS binds specifically to a mismatched base pair, while MutL scans the nearby DNA for a nick. Once MutL finds a nick, it triggers the degradation of the nicked strand all the way back through the mismatch. Because nicks are largely confined to newly replicated strands in eukaryotes, replication errors are selectively removed. In bacteria, an additional protein in the complex (MutH) nicks unmethylated (and therefore newly replicated) GATC sequences, thereby beginning the process illustrated here. In eukaryotes, MutL contains a DNA nicking activity that aids in the removal of the damaged strand.
Cell_Biology_Alberts_1166
Cell_Biology_Alberts
(B) The structure of the MutS protein bound to a DNA mismatch. This protein is a dimer, which grips the DNA double helix as shown, kinking the DNA at the mismatched base pair. It seems that the MutS protein scans the DNA for mismatches by testing for sites that can be readily kinked, which are those with an abnormal base pair. (PDB code: 1EWQ.) Figure 5–20 The “winding problem” that arises during DNA replication.
Cell_Biology_Alberts. (B) The structure of the MutS protein bound to a DNA mismatch. This protein is a dimer, which grips the DNA double helix as shown, kinking the DNA at the mismatched base pair. It seems that the MutS protein scans the DNA for mismatches by testing for sites that can be readily kinked, which are those with an abnormal base pair. (PDB code: 1EWQ.) Figure 5–20 The “winding problem” that arises during DNA replication.
Cell_Biology_Alberts_1167
Cell_Biology_Alberts
Figure 5–20 The “winding problem” that arises during DNA replication. (A) For a bacterial replication fork moving at 500 nucleotides per second, the parental DNA helix ahead of the fork must rotate at 50 revolutions per second. (B) If the ends of the DNA double helix remain fixed (or difficult to rotate), tension builds up in front of the replication fork as it becomes overwound. Some of this tension can be taken up by supercoiling, whereby the DNA double helix twists around itself (see Figure 6–19). However, if the tension continues to build up, the replication fork will eventually stop because further unwinding requires more energy than the helicase can provide. Note that in (A), the dotted line represents about 20 turns of DNA. one end of the DNA double helix cannot rotate relative to the other end topoisomerase with tyrosine at the active site
Cell_Biology_Alberts. Figure 5–20 The “winding problem” that arises during DNA replication. (A) For a bacterial replication fork moving at 500 nucleotides per second, the parental DNA helix ahead of the fork must rotate at 50 revolutions per second. (B) If the ends of the DNA double helix remain fixed (or difficult to rotate), tension builds up in front of the replication fork as it becomes overwound. Some of this tension can be taken up by supercoiling, whereby the DNA double helix twists around itself (see Figure 6–19). However, if the tension continues to build up, the replication fork will eventually stop because further unwinding requires more energy than the helicase can provide. Note that in (A), the dotted line represents about 20 turns of DNA. one end of the DNA double helix cannot rotate relative to the other end topoisomerase with tyrosine at the active site
Cell_Biology_Alberts_1168
Cell_Biology_Alberts
one end of the DNA double helix cannot rotate relative to the other end topoisomerase with tyrosine at the active site CH2 OH CH2 OH DNA topoisomerase covalently attaches to a DNA phosphate, thereby breaking a phosphodiester linkage in one DNA strand the two ends of the DNA double helix can now rotate relative to each other, relieving accumulated strain Figure 5–21 The reversible DNA nicking reaction catalyzed by a eukaryotic DNA topoisomerase I enzyme. As indicated, these enzymes transiently form a single covalent bond with DNA; this allows free rotation of the DNA around the covalent backbone bonds linked to the blue phosphate.
Cell_Biology_Alberts. one end of the DNA double helix cannot rotate relative to the other end topoisomerase with tyrosine at the active site CH2 OH CH2 OH DNA topoisomerase covalently attaches to a DNA phosphate, thereby breaking a phosphodiester linkage in one DNA strand the two ends of the DNA double helix can now rotate relative to each other, relieving accumulated strain Figure 5–21 The reversible DNA nicking reaction catalyzed by a eukaryotic DNA topoisomerase I enzyme. As indicated, these enzymes transiently form a single covalent bond with DNA; this allows free rotation of the DNA around the covalent backbone bonds linked to the blue phosphate.
Cell_Biology_Alberts_1169
Cell_Biology_Alberts
CH2HO the original phosphodiester bond energy is stored in the phosphotyrosine linkage, making the reaction reversible CH2 OH spontaneous re-formation of the phosphodiester bond regenerates both the DNA helix and the DNA topoisomerase the energy of the cleaved phosphodiester bond, resealing is rapid and does not require additional energy input. In this respect, the rejoining mechanism differs from that catalyzed by the enzyme DNA ligase, discussed previously (see Figure 5–12). A second type of DNA topoisomerase, topoisomerase II, forms a covalent linkage to both strands of the DNA helix at the same time, making a transient
Cell_Biology_Alberts. CH2HO the original phosphodiester bond energy is stored in the phosphotyrosine linkage, making the reaction reversible CH2 OH spontaneous re-formation of the phosphodiester bond regenerates both the DNA helix and the DNA topoisomerase the energy of the cleaved phosphodiester bond, resealing is rapid and does not require additional energy input. In this respect, the rejoining mechanism differs from that catalyzed by the enzyme DNA ligase, discussed previously (see Figure 5–12). A second type of DNA topoisomerase, topoisomerase II, forms a covalent linkage to both strands of the DNA helix at the same time, making a transient
Cell_Biology_Alberts_1170
Cell_Biology_Alberts
A second type of DNA topoisomerase, topoisomerase II, forms a covalent linkage to both strands of the DNA helix at the same time, making a transient Figure 5–22 The DNA-helix-passing reaction catalyzed by DNA topoisomerase II. Unlike type I topoisomerases, type II enzymes hydrolyze ATP (red), which is needed to release and reset the enzyme after each cycle. Type II topoisomerases are largely confined to proliferating cells in eukaryotes; partly for that reason, they have been effective targets for anticancer drugs. Some of these drugs inhibit topoisomerase II at the third step in the figure and thereby produce high levels of double-strand breaks that kill rapidly dividing cells. The small yellow circles represent the phosphates in the DNA backbone that become covalently bonded to the topoisomerase (see Figure 5–21).
Cell_Biology_Alberts. A second type of DNA topoisomerase, topoisomerase II, forms a covalent linkage to both strands of the DNA helix at the same time, making a transient Figure 5–22 The DNA-helix-passing reaction catalyzed by DNA topoisomerase II. Unlike type I topoisomerases, type II enzymes hydrolyze ATP (red), which is needed to release and reset the enzyme after each cycle. Type II topoisomerases are largely confined to proliferating cells in eukaryotes; partly for that reason, they have been effective targets for anticancer drugs. Some of these drugs inhibit topoisomerase II at the third step in the figure and thereby produce high levels of double-strand breaks that kill rapidly dividing cells. The small yellow circles represent the phosphates in the DNA backbone that become covalently bonded to the topoisomerase (see Figure 5–21).
Cell_Biology_Alberts_1171
Cell_Biology_Alberts
double-strand break in the helix. These enzymes are activated by sites on chromosomes where two double helices cross over each other such as those generated by supercoiling in front of a replication fork (see Figure 5–20). Once a topoisomerase II molecule binds to such a crossing site, the protein uses ATP hydrolysis to perform the following set of reactions efficiently: (1) it breaks one double helix reversibly to create a DNA “gate”; (2) it causes the second, nearby double helix to pass through this opening; and (3) it then reseals the break and dissociates from the DNA. At crossover points generated by supercoiling, passage of the double helix through the gate occurs in the direction that will reduce supercoiling. In this way, type II topoisomerases can relieve the overwinding tension generated in front of a replication fork. Their reaction mechanism also allows type II DNA topoisomerases to efficiently separate two interlocked DNA circles (Figure 5–22).
Cell_Biology_Alberts. double-strand break in the helix. These enzymes are activated by sites on chromosomes where two double helices cross over each other such as those generated by supercoiling in front of a replication fork (see Figure 5–20). Once a topoisomerase II molecule binds to such a crossing site, the protein uses ATP hydrolysis to perform the following set of reactions efficiently: (1) it breaks one double helix reversibly to create a DNA “gate”; (2) it causes the second, nearby double helix to pass through this opening; and (3) it then reseals the break and dissociates from the DNA. At crossover points generated by supercoiling, passage of the double helix through the gate occurs in the direction that will reduce supercoiling. In this way, type II topoisomerases can relieve the overwinding tension generated in front of a replication fork. Their reaction mechanism also allows type II DNA topoisomerases to efficiently separate two interlocked DNA circles (Figure 5–22).
Cell_Biology_Alberts_1172
Cell_Biology_Alberts
Topoisomerase II also prevents the severe DNA tangling problems that would otherwise arise during DNA replication. This role is nicely illustrated by mutant yeast cells that produce, in place of the normal topoisomerase II, a version that is inactive above 37°C. When the mutant cells are warmed to this temperature, their daughter chromosomes remain intertwined after DNA replication and are unable to separate. The enormous usefulness of topoisomerase II for untangling chromosomes can readily be appreciated by anyone who has struggled to remove a tangle from a fishing line without the aid of scissors.
Cell_Biology_Alberts. Topoisomerase II also prevents the severe DNA tangling problems that would otherwise arise during DNA replication. This role is nicely illustrated by mutant yeast cells that produce, in place of the normal topoisomerase II, a version that is inactive above 37°C. When the mutant cells are warmed to this temperature, their daughter chromosomes remain intertwined after DNA replication and are unable to separate. The enormous usefulness of topoisomerase II for untangling chromosomes can readily be appreciated by anyone who has struggled to remove a tangle from a fishing line without the aid of scissors.
Cell_Biology_Alberts_1173
Cell_Biology_Alberts
Much of what we know about DNA replication was first derived from studies of purified bacterial and bacteriophage multienzyme systems capable of DNA replication in vitro. The development of these systems in the 1970s was greatly facilitated by the prior isolation of mutants in a variety of replication genes; these mutants were exploited to identify and purify the corresponding replication proteins. The first mammalian replication system that accurately replicated DNA in vitro was described in the mid-1980s, and mutations in genes encoding nearly all of the replication components have now been isolated and analyzed in the yeast Saccharomyces cerevisiae. As a result, much is known about the detailed enzymology of DNA replication in eukaryotes, and it is clear that the fundamental features of DNA replication—including replication-fork geometry and the use of a multi-protein replication machine—have been conserved during the long evolutionary process that separated bacteria from
Cell_Biology_Alberts. Much of what we know about DNA replication was first derived from studies of purified bacterial and bacteriophage multienzyme systems capable of DNA replication in vitro. The development of these systems in the 1970s was greatly facilitated by the prior isolation of mutants in a variety of replication genes; these mutants were exploited to identify and purify the corresponding replication proteins. The first mammalian replication system that accurately replicated DNA in vitro was described in the mid-1980s, and mutations in genes encoding nearly all of the replication components have now been isolated and analyzed in the yeast Saccharomyces cerevisiae. As a result, much is known about the detailed enzymology of DNA replication in eukaryotes, and it is clear that the fundamental features of DNA replication—including replication-fork geometry and the use of a multi-protein replication machine—have been conserved during the long evolutionary process that separated bacteria from
Cell_Biology_Alberts_1174
Cell_Biology_Alberts
of DNA replication—including replication-fork geometry and the use of a multi-protein replication machine—have been conserved during the long evolutionary process that separated bacteria from eukaryotes.
Cell_Biology_Alberts. of DNA replication—including replication-fork geometry and the use of a multi-protein replication machine—have been conserved during the long evolutionary process that separated bacteria from eukaryotes.
Cell_Biology_Alberts_1175
Cell_Biology_Alberts
There are more protein components in eukaryotic replication machines than there are in the bacterial analogs, even though the basic functions are the same. Thus, for example, the eukaryotic single-strand binding (SSB) protein is formed from three subunits, whereas only a single subunit is found in bacteria. Similarly, the eukaryotic DNA primase is incorporated into a multisubunit enzyme that also contains a polymerase called DNA polymerase α-primase. This protein complex begins each Okazaki fragment on the lagging strand with RNA and then extends the RNA primer with a short length of DNA. At this point, the two main eukaryotic replicative DNA polymerases, Polδand Polε, come into play: Polδcompletes each Okazaki fragment on the lagging strand and Polε extends the leading strand. The increased complexity of eukaryotic replication machinery probably reflects reversal of the covalent attachment of the topoisomerase restores an intact orange double helix more elaborate controls. For
Cell_Biology_Alberts. There are more protein components in eukaryotic replication machines than there are in the bacterial analogs, even though the basic functions are the same. Thus, for example, the eukaryotic single-strand binding (SSB) protein is formed from three subunits, whereas only a single subunit is found in bacteria. Similarly, the eukaryotic DNA primase is incorporated into a multisubunit enzyme that also contains a polymerase called DNA polymerase α-primase. This protein complex begins each Okazaki fragment on the lagging strand with RNA and then extends the RNA primer with a short length of DNA. At this point, the two main eukaryotic replicative DNA polymerases, Polδand Polε, come into play: Polδcompletes each Okazaki fragment on the lagging strand and Polε extends the leading strand. The increased complexity of eukaryotic replication machinery probably reflects reversal of the covalent attachment of the topoisomerase restores an intact orange double helix more elaborate controls. For
Cell_Biology_Alberts_1176
Cell_Biology_Alberts
increased complexity of eukaryotic replication machinery probably reflects reversal of the covalent attachment of the topoisomerase restores an intact orange double helix more elaborate controls. For example, the orderly maintenance of different cell types and tissues in animals and plants requires that DNA replication be tightly regulated. Moreover, eukaryotic DNA replication must be coordinated with the elaborate process of mitosis, as we discuss in Chapter 17.
Cell_Biology_Alberts. increased complexity of eukaryotic replication machinery probably reflects reversal of the covalent attachment of the topoisomerase restores an intact orange double helix more elaborate controls. For example, the orderly maintenance of different cell types and tissues in animals and plants requires that DNA replication be tightly regulated. Moreover, eukaryotic DNA replication must be coordinated with the elaborate process of mitosis, as we discuss in Chapter 17.
Cell_Biology_Alberts_1177
Cell_Biology_Alberts
As we see in the next section, the eukaryotic replication machinery has the added complication of having to replicate through nucleosomes, the repeating structural unit of chromosomes discussed in Chapter 4. Nucleosomes are spaced at intervals of about 200 nucleotide pairs along the DNA, which, as we will see, explains why new Okazaki fragments are synthesized on the lagging strand at intervals of 100–200 nucleotides in eukaryotes, instead of 1000–2000 nucleotides as in bacteria. Nucleosomes may also act as barriers that slow down the movement of DNA polymerase molecules, which may be why eukaryotic replication forks move only about one-tenth as fast as bacterial replication forks.
Cell_Biology_Alberts. As we see in the next section, the eukaryotic replication machinery has the added complication of having to replicate through nucleosomes, the repeating structural unit of chromosomes discussed in Chapter 4. Nucleosomes are spaced at intervals of about 200 nucleotide pairs along the DNA, which, as we will see, explains why new Okazaki fragments are synthesized on the lagging strand at intervals of 100–200 nucleotides in eukaryotes, instead of 1000–2000 nucleotides as in bacteria. Nucleosomes may also act as barriers that slow down the movement of DNA polymerase molecules, which may be why eukaryotic replication forks move only about one-tenth as fast as bacterial replication forks.
Cell_Biology_Alberts_1178
Cell_Biology_Alberts
DNA replication takes place at a Y-shaped structure called a replication fork. A self-correcting DNA polymerase enzyme catalyzes nucleotide polymerization in a 5ʹ-to-3ʹ direction, copying a DNA template strand with remarkable fidelity. Since the two strands of a DNA double helix are antiparallel, this 5ʹ-to-3ʹ DNA synthesis can take place continuously on only one of the strands at a replication fork (the leading strand). On the lagging strand, short DNA fragments must be made by a “backstitching” process. Because the self-correcting DNA polymerase cannot start a new chain, these lagging-strand DNA fragments are primed by short RNA primer molecules that are subsequently erased and replaced with DNA.
Cell_Biology_Alberts. DNA replication takes place at a Y-shaped structure called a replication fork. A self-correcting DNA polymerase enzyme catalyzes nucleotide polymerization in a 5ʹ-to-3ʹ direction, copying a DNA template strand with remarkable fidelity. Since the two strands of a DNA double helix are antiparallel, this 5ʹ-to-3ʹ DNA synthesis can take place continuously on only one of the strands at a replication fork (the leading strand). On the lagging strand, short DNA fragments must be made by a “backstitching” process. Because the self-correcting DNA polymerase cannot start a new chain, these lagging-strand DNA fragments are primed by short RNA primer molecules that are subsequently erased and replaced with DNA.
Cell_Biology_Alberts_1179
Cell_Biology_Alberts
DNA replication requires the cooperation of many proteins. These include (1) DNA polymerase and DNA primase to catalyze nucleoside triphosphate polymerization; (2) DNA helicases and single-strand DNA-binding (SSB) proteins to help in opening up the DNA helix so that it can be copied; (3) DNA ligase and an enzyme that degrades RNA primers to seal together the discontinuously synthesized lagging-strand DNA fragments; and (4) DNA topoisomerases to help to relieve helical winding and DNA tangling problems. Many of these proteins associate with each other at a replication fork to form a highly efficient “replication machine,” through which the activities and spatial movements of the individual components are coordinated.
Cell_Biology_Alberts. DNA replication requires the cooperation of many proteins. These include (1) DNA polymerase and DNA primase to catalyze nucleoside triphosphate polymerization; (2) DNA helicases and single-strand DNA-binding (SSB) proteins to help in opening up the DNA helix so that it can be copied; (3) DNA ligase and an enzyme that degrades RNA primers to seal together the discontinuously synthesized lagging-strand DNA fragments; and (4) DNA topoisomerases to help to relieve helical winding and DNA tangling problems. Many of these proteins associate with each other at a replication fork to form a highly efficient “replication machine,” through which the activities and spatial movements of the individual components are coordinated.
Cell_Biology_Alberts_1180
Cell_Biology_Alberts
We have seen how a set of replication proteins rapidly and accurately generates two daughter DNA double helices behind a replication fork. But how is this replication machinery assembled in the first place, and how are replication forks created on an intact, double-strand DNA molecule? In this section, we discuss how cells initiate DNA replication and how they carefully regulate this process to ensure that it takes place not only at the proper positions on the chromosome but also at the appropriate time in the life of the cell. We also discuss a few of the special problems that the replication machinery in eukaryotic cells must overcome. These include the need to replicate the enormously long DNA molecules found in eukaryotic chromosomes, as well as the difficulty of copying DNA molecules that are tightly complexed with histones in nucleosomes.
Cell_Biology_Alberts. We have seen how a set of replication proteins rapidly and accurately generates two daughter DNA double helices behind a replication fork. But how is this replication machinery assembled in the first place, and how are replication forks created on an intact, double-strand DNA molecule? In this section, we discuss how cells initiate DNA replication and how they carefully regulate this process to ensure that it takes place not only at the proper positions on the chromosome but also at the appropriate time in the life of the cell. We also discuss a few of the special problems that the replication machinery in eukaryotic cells must overcome. These include the need to replicate the enormously long DNA molecules found in eukaryotic chromosomes, as well as the difficulty of copying DNA molecules that are tightly complexed with histones in nucleosomes.
Cell_Biology_Alberts_1181
Cell_Biology_Alberts
As discussed previously, the DNA double helix is normally very stable: the two DNA strands are locked together firmly by many hydrogen bonds formed between the bases on each strand. To begin DNA replication, the double helix must first be opened up and the two strands separated to expose unpaired bases. As we shall see, the process of DNA replication is begun by special initiator proteins that bind to double-strand DNA and pry the two strands apart, breaking the hydrogen bonds between the bases. leading strand of fork 1 of fork 2 leading strand lagging strand of fork 1 of fork 2 Figure 5–23 A replication bubble formed by replication-fork initiation. This diagram outlines the major steps in the initiation of replication forks at replication origins. The structure formed at the last step, in which both strands of the parental DNA helix have been separated from each other and serve as templates for DNA synthesis, is called a replication bubble.
Cell_Biology_Alberts. As discussed previously, the DNA double helix is normally very stable: the two DNA strands are locked together firmly by many hydrogen bonds formed between the bases on each strand. To begin DNA replication, the double helix must first be opened up and the two strands separated to expose unpaired bases. As we shall see, the process of DNA replication is begun by special initiator proteins that bind to double-strand DNA and pry the two strands apart, breaking the hydrogen bonds between the bases. leading strand of fork 1 of fork 2 leading strand lagging strand of fork 1 of fork 2 Figure 5–23 A replication bubble formed by replication-fork initiation. This diagram outlines the major steps in the initiation of replication forks at replication origins. The structure formed at the last step, in which both strands of the parental DNA helix have been separated from each other and serve as templates for DNA synthesis, is called a replication bubble.
Cell_Biology_Alberts_1182
Cell_Biology_Alberts
Figure 5–24 DNA replication of a bacterial genome. It takes E. coli about 30 minutes to duplicate its genome of 4.6 × 106 nucleotide pairs. For simplicity, no Okazaki fragments are shown on the lagging strand. What happens as the two replication forks approach each other and collide at the end of the replication cycle is not well understood, although the replication machines are disassembled as part of the process.
Cell_Biology_Alberts. Figure 5–24 DNA replication of a bacterial genome. It takes E. coli about 30 minutes to duplicate its genome of 4.6 × 106 nucleotide pairs. For simplicity, no Okazaki fragments are shown on the lagging strand. What happens as the two replication forks approach each other and collide at the end of the replication cycle is not well understood, although the replication machines are disassembled as part of the process.
Cell_Biology_Alberts_1183
Cell_Biology_Alberts
The positions at which the DNA helix is first opened are called replication origins (Figure 5–23). In simple cells like those of bacteria or yeast, origins are specified by DNA sequences several hundred nucleotide pairs in length. This DNA contains both short sequences that attract initiator proteins and stretches of DNA that are especially easy to open. We saw in Figure 4–4 that an A-T base pair is held together by fewer hydrogen bonds than a G-C base pair. Therefore, DNA rich in A-T base pairs is relatively easy to pull apart, and regions of DNA enriched in A-T base pairs are typically found at replication origins.
Cell_Biology_Alberts. The positions at which the DNA helix is first opened are called replication origins (Figure 5–23). In simple cells like those of bacteria or yeast, origins are specified by DNA sequences several hundred nucleotide pairs in length. This DNA contains both short sequences that attract initiator proteins and stretches of DNA that are especially easy to open. We saw in Figure 4–4 that an A-T base pair is held together by fewer hydrogen bonds than a G-C base pair. Therefore, DNA rich in A-T base pairs is relatively easy to pull apart, and regions of DNA enriched in A-T base pairs are typically found at replication origins.
Cell_Biology_Alberts_1184
Cell_Biology_Alberts
Although the basic process of replication-fork initiation depicted in Figure 5–23 is fundamentally the same for bacteria and eukaryotes, the detailed way in which this process is performed and regulated differs between these two groups of organisms. We first consider the simpler and better-understood case in bacteria and then turn to the more complex situation found in yeasts, mammals, and other eukaryotes. Bacterial Chromosomes Typically Have a Single Origin of DNA Replication
Cell_Biology_Alberts. Although the basic process of replication-fork initiation depicted in Figure 5–23 is fundamentally the same for bacteria and eukaryotes, the detailed way in which this process is performed and regulated differs between these two groups of organisms. We first consider the simpler and better-understood case in bacteria and then turn to the more complex situation found in yeasts, mammals, and other eukaryotes. Bacterial Chromosomes Typically Have a Single Origin of DNA Replication
Cell_Biology_Alberts_1185
Cell_Biology_Alberts
The genome of E. coli is contained in a single circular DNA molecule of 4.6 × 106 nucleotide pairs. DNA replication begins at a single origin of replication, and the two replication forks assembled there proceed (at approximately 1000 nucleotides per second) in opposite directions until they meet up roughly halfway around the chromosome (Figure 5–24). The only point at which E. coli can control DNA replication is initiation: once the forks have been assembled at the origin, they synthesize DNA at relatively constant speed until replication is finished. Therefore, it is not surprising that the initiation of DNA replication is highly regulated. The process begins when initiator proteins (in their ATP-bound state) bind in multiple copies to specific DNA sites located at the replication origin, wrapping the DNA around the proteins to form a large protein–DNA complex that destabilizes the adjacent double helix. This complex then attracts two DNA helicases, each bound to a helicase loader,
Cell_Biology_Alberts. The genome of E. coli is contained in a single circular DNA molecule of 4.6 × 106 nucleotide pairs. DNA replication begins at a single origin of replication, and the two replication forks assembled there proceed (at approximately 1000 nucleotides per second) in opposite directions until they meet up roughly halfway around the chromosome (Figure 5–24). The only point at which E. coli can control DNA replication is initiation: once the forks have been assembled at the origin, they synthesize DNA at relatively constant speed until replication is finished. Therefore, it is not surprising that the initiation of DNA replication is highly regulated. The process begins when initiator proteins (in their ATP-bound state) bind in multiple copies to specific DNA sites located at the replication origin, wrapping the DNA around the proteins to form a large protein–DNA complex that destabilizes the adjacent double helix. This complex then attracts two DNA helicases, each bound to a helicase loader,
Cell_Biology_Alberts_1186
Cell_Biology_Alberts
wrapping the DNA around the proteins to form a large protein–DNA complex that destabilizes the adjacent double helix. This complex then attracts two DNA helicases, each bound to a helicase loader, and these are placed around adjacent DNA single strands whose bases have been exposed by the assembly of the initiator protein–DNA complex. The helicase loader is analogous to the clamp loader we encountered above; it has the additional job of keeping the helicase in an inactive form until it is properly loaded onto a nascent replication fork. Once the helicases are loaded, the loaders dissociate and the helicases begin to unwind DNA, exposing enough single-strand DNA for DNA primase to synthesize the first RNA primers (Figure 5–25). This quickly leads to the assembly of remaining proteins to create two replication forks, with replication machines that move, with respect to the replication origin, in opposite directions. They continue to synthesize DNA until all of the DNA template
Cell_Biology_Alberts. wrapping the DNA around the proteins to form a large protein–DNA complex that destabilizes the adjacent double helix. This complex then attracts two DNA helicases, each bound to a helicase loader, and these are placed around adjacent DNA single strands whose bases have been exposed by the assembly of the initiator protein–DNA complex. The helicase loader is analogous to the clamp loader we encountered above; it has the additional job of keeping the helicase in an inactive form until it is properly loaded onto a nascent replication fork. Once the helicases are loaded, the loaders dissociate and the helicases begin to unwind DNA, exposing enough single-strand DNA for DNA primase to synthesize the first RNA primers (Figure 5–25). This quickly leads to the assembly of remaining proteins to create two replication forks, with replication machines that move, with respect to the replication origin, in opposite directions. They continue to synthesize DNA until all of the DNA template
Cell_Biology_Alberts_1187
Cell_Biology_Alberts
to create two replication forks, with replication machines that move, with respect to the replication origin, in opposite directions. They continue to synthesize DNA until all of the DNA template downstream of each fork has been replicated.
Cell_Biology_Alberts. to create two replication forks, with replication machines that move, with respect to the replication origin, in opposite directions. They continue to synthesize DNA until all of the DNA template downstream of each fork has been replicated.
Cell_Biology_Alberts_1188
Cell_Biology_Alberts
In E. coli, the interaction of the initiator protein with the replication origin is carefully regulated, with initiation occurring only when sufficient nutrients are available for the bacterium to complete an entire round of replication. Initiation is also controlled to ensure that only one round of DNA replication occurs for each cell division. After replication is initiated, the initiator protein is inactivated by hydrolysis of its bound ATP molecule, and the origin of replication experiences a “refractory period.” The refractory period is caused by a delay in the methylation of newly incorporated A nucleotides in the origin (Figure 5–26). Initiation cannot occur again until the A’s are methylated and the initiator protein is restored to its ATP-bound state. Eukaryotic Chromosomes Contain Multiple Origins of Replication
Cell_Biology_Alberts. In E. coli, the interaction of the initiator protein with the replication origin is carefully regulated, with initiation occurring only when sufficient nutrients are available for the bacterium to complete an entire round of replication. Initiation is also controlled to ensure that only one round of DNA replication occurs for each cell division. After replication is initiated, the initiator protein is inactivated by hydrolysis of its bound ATP molecule, and the origin of replication experiences a “refractory period.” The refractory period is caused by a delay in the methylation of newly incorporated A nucleotides in the origin (Figure 5–26). Initiation cannot occur again until the A’s are methylated and the initiator protein is restored to its ATP-bound state. Eukaryotic Chromosomes Contain Multiple Origins of Replication
Cell_Biology_Alberts_1189
Cell_Biology_Alberts
Eukaryotic Chromosomes Contain Multiple Origins of Replication We have seen how two replication forks begin at a single replication origin in bacteria and proceed in opposite directions, moving away from the origin until all of the DNA in the single circular chromosome is replicated. The bacterial genome is sufficiently small for these two replication forks to duplicate the genome in about 30 minutes. Because of the much greater size of most eukaryotic chromosomes, a different strategy is required to allow their replication in a timely manner.
Cell_Biology_Alberts. Eukaryotic Chromosomes Contain Multiple Origins of Replication We have seen how two replication forks begin at a single replication origin in bacteria and proceed in opposite directions, moving away from the origin until all of the DNA in the single circular chromosome is replicated. The bacterial genome is sufficiently small for these two replication forks to duplicate the genome in about 30 minutes. Because of the much greater size of most eukaryotic chromosomes, a different strategy is required to allow their replication in a timely manner.
Cell_Biology_Alberts_1190
Cell_Biology_Alberts
A method for determining the general pattern of eukaryotic chromosome replication was developed in the early 1960s. Human cells growing in culture are labeled for a short time with 3H-thymidine so that the DNA synthesized during this period becomes highly radioactive. The cells are then gently lysed, and the DNA is stretched on the surface of a glass slide coated with a photographic emulsion. Development of the emulsion reveals the pattern of labeled DNA through a technique known as autoradiography. The time allotted for radioactive labeling is chosen to allow each replication fork to move several micrometers along the DNA, so that the replicated DNA can be detected in the light microscope as lines of silver grains, even though the DNA molecule itself is too thin to be visible.
Cell_Biology_Alberts. A method for determining the general pattern of eukaryotic chromosome replication was developed in the early 1960s. Human cells growing in culture are labeled for a short time with 3H-thymidine so that the DNA synthesized during this period becomes highly radioactive. The cells are then gently lysed, and the DNA is stretched on the surface of a glass slide coated with a photographic emulsion. Development of the emulsion reveals the pattern of labeled DNA through a technique known as autoradiography. The time allotted for radioactive labeling is chosen to allow each replication fork to move several micrometers along the DNA, so that the replicated DNA can be detected in the light microscope as lines of silver grains, even though the DNA molecule itself is too thin to be visible.
Cell_Biology_Alberts_1191
Cell_Biology_Alberts
Figure 5–25 The proteins that initiate DNA replication in bacteria. The mechanism shown was established by studies in vitro with mixtures of highly purified proteins. For E. coli DNA replication, the major initiator protein, the helicase, and the primase are the dnaA, dnaB, and dnaG proteins, respectively. In the first step, several molecules of the initiator protein bind to specific DNA sequences at the replication origin and destabilize the double helix by forming a compact structure in which the DNA is tightly wrapped around the protein. Next, two helicases are brought in by helicaseloading proteins (the dnaC proteins), which inhibit the helicases until they are properly loaded at the replication origin. Helicase-loading proteins prevent the replicative DNA helices from inappropriately entering other single-strand stretches of DNA in the bacterial genome. Aided by single-strand binding protein (not shown), the loaded helicases open up the DNA, thereby enabling primases to enter and
Cell_Biology_Alberts. Figure 5–25 The proteins that initiate DNA replication in bacteria. The mechanism shown was established by studies in vitro with mixtures of highly purified proteins. For E. coli DNA replication, the major initiator protein, the helicase, and the primase are the dnaA, dnaB, and dnaG proteins, respectively. In the first step, several molecules of the initiator protein bind to specific DNA sequences at the replication origin and destabilize the double helix by forming a compact structure in which the DNA is tightly wrapped around the protein. Next, two helicases are brought in by helicaseloading proteins (the dnaC proteins), which inhibit the helicases until they are properly loaded at the replication origin. Helicase-loading proteins prevent the replicative DNA helices from inappropriately entering other single-strand stretches of DNA in the bacterial genome. Aided by single-strand binding protein (not shown), the loaded helicases open up the DNA, thereby enabling primases to enter and
Cell_Biology_Alberts_1192
Cell_Biology_Alberts
entering other single-strand stretches of DNA in the bacterial genome. Aided by single-strand binding protein (not shown), the loaded helicases open up the DNA, thereby enabling primases to enter and synthesize initial primers. In subsequent steps, two complete replication forks are assembled at the origin and move off in opposite directions. The initiator proteins are displaced as the left-hand fork moves through them (not shown).
Cell_Biology_Alberts. entering other single-strand stretches of DNA in the bacterial genome. Aided by single-strand binding protein (not shown), the loaded helicases open up the DNA, thereby enabling primases to enter and synthesize initial primers. In subsequent steps, two complete replication forks are assembled at the origin and move off in opposite directions. The initiator proteins are displaced as the left-hand fork moves through them (not shown).
Cell_Biology_Alberts_1193
Cell_Biology_Alberts
fully methylated hemimethylated origins are origin resistant to initiation initiation occurs if suffcient origins become fully resources are available to complete methylated, making them a round of DNA replication again competent for initiation In this way, both the rate and the direction of replication-fork movement can be determined (Figure 5–27). From the rate at which tracks of replicated DNA increase in length with increasing labeling time, the eukaryotic replication forks are estimated to travel at about 50 nucleotides per second. This is approximately twentyfold slower than the rate at which bacterial replication forks move, possibly reflecting the increased difficulty of replicating DNA that is packaged tightly in chromatin.
Cell_Biology_Alberts. fully methylated hemimethylated origins are origin resistant to initiation initiation occurs if suffcient origins become fully resources are available to complete methylated, making them a round of DNA replication again competent for initiation In this way, both the rate and the direction of replication-fork movement can be determined (Figure 5–27). From the rate at which tracks of replicated DNA increase in length with increasing labeling time, the eukaryotic replication forks are estimated to travel at about 50 nucleotides per second. This is approximately twentyfold slower than the rate at which bacterial replication forks move, possibly reflecting the increased difficulty of replicating DNA that is packaged tightly in chromatin.
Cell_Biology_Alberts_1194
Cell_Biology_Alberts
An average-size human chromosome contains a single linear DNA molecule of about 150 million nucleotide pairs. It would take 0.02 seconds/nucleotide × 150 × 106 nucleotides = 3.0 × 106 seconds (about 35 days) to replicate such a DNA molecule from end to end with a single replication fork moving at a rate of 50 nucleotides per second. As expected, therefore, the autoradiographic experiments just described reveal that many forks, belonging to separate replication bubbles, are moving simultaneously on each eukaryotic chromosome.
Cell_Biology_Alberts. An average-size human chromosome contains a single linear DNA molecule of about 150 million nucleotide pairs. It would take 0.02 seconds/nucleotide × 150 × 106 nucleotides = 3.0 × 106 seconds (about 35 days) to replicate such a DNA molecule from end to end with a single replication fork moving at a rate of 50 nucleotides per second. As expected, therefore, the autoradiographic experiments just described reveal that many forks, belonging to separate replication bubbles, are moving simultaneously on each eukaryotic chromosome.
Cell_Biology_Alberts_1195
Cell_Biology_Alberts
Much faster and more sophisticated methods now exist for monitoring DNA replication initiation and tracking the movement of DNA replication forks across whole genomes. One approach uses DNA microarrays—grids the size of a postage stamp studded with hundreds of thousands of fragments of known DNA sequence. As we will see in detail in Chapter 8, each different DNA fragment is placed at a unique position on the microarray, and whole genomes can thereby be represented in an orderly manner. If a DNA sample from a group of replicating cells is broken up and hybridized to a microarray representing that organism’s genome, the amount of each DNA sequence can be determined. Because a segment of a genome that has been replicated will contain twice as much DNA as an unreplicated segment, replication-fork initiation and fork movement can be accurately monitored across an entire genome (Figure 5–28).
Cell_Biology_Alberts. Much faster and more sophisticated methods now exist for monitoring DNA replication initiation and tracking the movement of DNA replication forks across whole genomes. One approach uses DNA microarrays—grids the size of a postage stamp studded with hundreds of thousands of fragments of known DNA sequence. As we will see in detail in Chapter 8, each different DNA fragment is placed at a unique position on the microarray, and whole genomes can thereby be represented in an orderly manner. If a DNA sample from a group of replicating cells is broken up and hybridized to a microarray representing that organism’s genome, the amount of each DNA sequence can be determined. Because a segment of a genome that has been replicated will contain twice as much DNA as an unreplicated segment, replication-fork initiation and fork movement can be accurately monitored across an entire genome (Figure 5–28).
Cell_Biology_Alberts_1196
Cell_Biology_Alberts
Experiments of this type have shown the following: (1) Approximately 30,000– 50,000 origins of replication are used each time a human cell divides. (2) The human genome has many more (perhaps tenfold more) potential origins than this, and different cell types use different sets of origins. This may allow a cell to coordinate its active origins with other features of its chromosomes such as which Figure 5–26 Methylation of the E. coli replication origin creates a refractory period for DNA initiation. DNA methylation occurs at GATC sequences, 11 of which are found in the origin of replication (spanning approximately 250 nucleotide pairs). In its hemimethylated state, the origin of replication is bound by an inhibitor protein (Seq A, not shown), which blocks the ability of the initiator proteins to unwind the origin DNA. Eventually (about 15 minutes after replication is initiated), the hemimethylated origins become fully methylated by a DNA methylase enzyme; Seq A then dissociates.
Cell_Biology_Alberts. Experiments of this type have shown the following: (1) Approximately 30,000– 50,000 origins of replication are used each time a human cell divides. (2) The human genome has many more (perhaps tenfold more) potential origins than this, and different cell types use different sets of origins. This may allow a cell to coordinate its active origins with other features of its chromosomes such as which Figure 5–26 Methylation of the E. coli replication origin creates a refractory period for DNA initiation. DNA methylation occurs at GATC sequences, 11 of which are found in the origin of replication (spanning approximately 250 nucleotide pairs). In its hemimethylated state, the origin of replication is bound by an inhibitor protein (Seq A, not shown), which blocks the ability of the initiator proteins to unwind the origin DNA. Eventually (about 15 minutes after replication is initiated), the hemimethylated origins become fully methylated by a DNA methylase enzyme; Seq A then dissociates.
Cell_Biology_Alberts_1197
Cell_Biology_Alberts
A single enzyme, the Dam methylase, is responsible for methylating all E. coli GATC sequences. A lag in methylation after the replication of GATC sequences is also used by the E. coli mismatch proofreading system to distinguish the newly synthesized DNA strand from the parental DNA strand; in that case, the relevant GATC sequences are scattered throughout the chromosome, and they are not bound by Seq A. Figure 5–27 The experiments that demonstrated the pattern in which replication forks are formed and move on eukaryotic chromosomes. The new DNA made in human cells in culture was labeled briefly with a pulse of highly radioactive thymidine (3H-thymidine).
Cell_Biology_Alberts. A single enzyme, the Dam methylase, is responsible for methylating all E. coli GATC sequences. A lag in methylation after the replication of GATC sequences is also used by the E. coli mismatch proofreading system to distinguish the newly synthesized DNA strand from the parental DNA strand; in that case, the relevant GATC sequences are scattered throughout the chromosome, and they are not bound by Seq A. Figure 5–27 The experiments that demonstrated the pattern in which replication forks are formed and move on eukaryotic chromosomes. The new DNA made in human cells in culture was labeled briefly with a pulse of highly radioactive thymidine (3H-thymidine).
Cell_Biology_Alberts_1198
Cell_Biology_Alberts
(A) In this experiment, the cells were lysed, and the DNA was stretched out on a glass slide that was subsequently covered with a photographic emulsion. After several months, the emulsion was developed, revealing a line of silver grains over the radioactive DNA. The brown DNA in this figure is shown only to help with the interpretation of the autoradiograph; the unlabeled DNA is invisible in such experiments. (B) This experiment was the same except that a further incubation in unlabeled medium allowed additional DNA, with a lower level of radioactivity, to be replicated. The pairs of dark tracks in (B) were found to have silver grains tapering off in opposite directions, demonstrating bidirectional fork movement from a central replication origin where a replication bubble forms (see Figure 5–23). A replication fork is thought to stop only when it encounters a replication fork moving in the opposite direction or when it reaches the end of the chromosome; in this way, all the DNA is
Cell_Biology_Alberts. (A) In this experiment, the cells were lysed, and the DNA was stretched out on a glass slide that was subsequently covered with a photographic emulsion. After several months, the emulsion was developed, revealing a line of silver grains over the radioactive DNA. The brown DNA in this figure is shown only to help with the interpretation of the autoradiograph; the unlabeled DNA is invisible in such experiments. (B) This experiment was the same except that a further incubation in unlabeled medium allowed additional DNA, with a lower level of radioactivity, to be replicated. The pairs of dark tracks in (B) were found to have silver grains tapering off in opposite directions, demonstrating bidirectional fork movement from a central replication origin where a replication bubble forms (see Figure 5–23). A replication fork is thought to stop only when it encounters a replication fork moving in the opposite direction or when it reaches the end of the chromosome; in this way, all the DNA is
Cell_Biology_Alberts_1199
Cell_Biology_Alberts
5–23). A replication fork is thought to stop only when it encounters a replication fork moving in the opposite direction or when it reaches the end of the chromosome; in this way, all the DNA is eventually replicated.
Cell_Biology_Alberts. 5–23). A replication fork is thought to stop only when it encounters a replication fork moving in the opposite direction or when it reaches the end of the chromosome; in this way, all the DNA is eventually replicated.
Cell_Biology_Alberts_1200
Cell_Biology_Alberts
258 Chapter 5: DNA Replication, Repair, and Recombination culture of cells arrested before DNA replication begins allow replication to begin fragment DNA, separate strands, and fuorescently label genes are being expressed. The excess origins also provide “backups” in case a primary origin fails. (3) As in bacteria, replication forks are formed in pairs and create a replication bubble as they move in opposite directions away from a common point of origin, stopping only when they collide head-on with a replication fork moving in the opposite direction or when they reach a chromosome end. In this way, many replication forks operate independently on each chromosome and yet form two complete daughter DNA helices. In Eukaryotes, DNA Replication Takes Place During Only One Part of the Cell Cycle
Cell_Biology_Alberts. 258 Chapter 5: DNA Replication, Repair, and Recombination culture of cells arrested before DNA replication begins allow replication to begin fragment DNA, separate strands, and fuorescently label genes are being expressed. The excess origins also provide “backups” in case a primary origin fails. (3) As in bacteria, replication forks are formed in pairs and create a replication bubble as they move in opposite directions away from a common point of origin, stopping only when they collide head-on with a replication fork moving in the opposite direction or when they reach a chromosome end. In this way, many replication forks operate independently on each chromosome and yet form two complete daughter DNA helices. In Eukaryotes, DNA Replication Takes Place During Only One Part of the Cell Cycle
Cell_Biology_Alberts_1201
Cell_Biology_Alberts
In Eukaryotes, DNA Replication Takes Place During Only One Part of the Cell Cycle When growing rapidly, bacteria replicate their DNA nearly continuously. In contrast, DNA replication in most eukaryotic cells occurs only during a specific part of the cell-division cycle, called the DNA synthesis phase or S phase (Figure 5–29). In a mammalian cell, the S phase typically lasts for about 8 hours; in simpler eukaryotic cells such as yeasts, the S phase can be as short as 40 minutes. By its end, each chromosome has been replicated to produce two complete copies, which remain joined together at their centromeres until the M phase (M for mitosis), which soon follows. In Chapter 17, we describe the control system that runs the cell cycle, and we explain why entry into each phase of the cycle requires the cell to have successfully completed the previous phase. In the following sections, we explore how chromosome replication is coordinated within the S phase of the cell cycle.
Cell_Biology_Alberts. In Eukaryotes, DNA Replication Takes Place During Only One Part of the Cell Cycle When growing rapidly, bacteria replicate their DNA nearly continuously. In contrast, DNA replication in most eukaryotic cells occurs only during a specific part of the cell-division cycle, called the DNA synthesis phase or S phase (Figure 5–29). In a mammalian cell, the S phase typically lasts for about 8 hours; in simpler eukaryotic cells such as yeasts, the S phase can be as short as 40 minutes. By its end, each chromosome has been replicated to produce two complete copies, which remain joined together at their centromeres until the M phase (M for mitosis), which soon follows. In Chapter 17, we describe the control system that runs the cell cycle, and we explain why entry into each phase of the cycle requires the cell to have successfully completed the previous phase. In the following sections, we explore how chromosome replication is coordinated within the S phase of the cell cycle.
Cell_Biology_Alberts_1202
Cell_Biology_Alberts
In the following sections, we explore how chromosome replication is coordinated within the S phase of the cell cycle. Different Regions on the Same Chromosome Replicate at Distinct Times in S Phase In mammalian cells, the replication of DNA in the region between one replication origin and the next should normally require only about an hour to complete, given the rate at which a replication fork moves and the largest distances measured between replication origins. Yet S phase usually lasts for about 8 hours in a mammalian cell. This implies that the replication origins are not all activated simultaneously; indeed, replication origins are activated in clusters of about 50 adjacent replication origins, each of which is replicated during only a small part of the total S-phase interval.
Cell_Biology_Alberts. In the following sections, we explore how chromosome replication is coordinated within the S phase of the cell cycle. Different Regions on the Same Chromosome Replicate at Distinct Times in S Phase In mammalian cells, the replication of DNA in the region between one replication origin and the next should normally require only about an hour to complete, given the rate at which a replication fork moves and the largest distances measured between replication origins. Yet S phase usually lasts for about 8 hours in a mammalian cell. This implies that the replication origins are not all activated simultaneously; indeed, replication origins are activated in clusters of about 50 adjacent replication origins, each of which is replicated during only a small part of the total S-phase interval.
Cell_Biology_Alberts_1203
Cell_Biology_Alberts
Figure 5–28 Use of DNA microarrays to monitor the formation and progress of replication forks. For this experiment, a population of cells is synchronized so that they all begin replication at the same time. DNA is collected and hybridized to the microarray; DNA that has been replicated once gives a hybridization signal (dark green squares) twice as high as that of unreplicated DNA (light green squares). The spots on these microarrays represent consecutive sequences along a segment of a chromosome arranged left to right, top to bottom. Only 81 spots are shown here, but the actual arrays contain hundreds of thousands of sequences that span an entire genome. As can be seen, replication begins at an origin and proceeds bidirectionally. For simplicity, only one origin is shown here. In human cells, replication begins at 30,000–50,000 origins located throughout the genome. Using this approach it is possible to observe the formation and progress of every replication fork across a genome.
Cell_Biology_Alberts. Figure 5–28 Use of DNA microarrays to monitor the formation and progress of replication forks. For this experiment, a population of cells is synchronized so that they all begin replication at the same time. DNA is collected and hybridized to the microarray; DNA that has been replicated once gives a hybridization signal (dark green squares) twice as high as that of unreplicated DNA (light green squares). The spots on these microarrays represent consecutive sequences along a segment of a chromosome arranged left to right, top to bottom. Only 81 spots are shown here, but the actual arrays contain hundreds of thousands of sequences that span an entire genome. As can be seen, replication begins at an origin and proceeds bidirectionally. For simplicity, only one origin is shown here. In human cells, replication begins at 30,000–50,000 origins located throughout the genome. Using this approach it is possible to observe the formation and progress of every replication fork across a genome.
Cell_Biology_Alberts_1204
Cell_Biology_Alberts
Figure 5–29 The four successive phases of a standard eukaryotic cell cycle. During the G1, S, and G2 phases, the cell grows continuously. During M phase growth stops, the nucleus divides, and the cell divides in two. DNA replication is confined to the part of the cell cycle known as S phase. G1 is the gap between M phase and S phase; G2 is the gap between S phase and M phase. It seems that the order in which replication origins are activated depends, in part, on the chromatin structure in which the origins reside. We saw in Chapter 4 that heterochromatin is a particularly condensed state of chromatin, while euchromatin, where most transcription occurs, has a less condensed conformation. Heterochromatin tends to be replicated very late in S phase, suggesting that the timing of replication is related to the packing of the DNA in chromatin.
Cell_Biology_Alberts. Figure 5–29 The four successive phases of a standard eukaryotic cell cycle. During the G1, S, and G2 phases, the cell grows continuously. During M phase growth stops, the nucleus divides, and the cell divides in two. DNA replication is confined to the part of the cell cycle known as S phase. G1 is the gap between M phase and S phase; G2 is the gap between S phase and M phase. It seems that the order in which replication origins are activated depends, in part, on the chromatin structure in which the origins reside. We saw in Chapter 4 that heterochromatin is a particularly condensed state of chromatin, while euchromatin, where most transcription occurs, has a less condensed conformation. Heterochromatin tends to be replicated very late in S phase, suggesting that the timing of replication is related to the packing of the DNA in chromatin.
Cell_Biology_Alberts_1205
Cell_Biology_Alberts
Once initiated, however, replication forks seem to move at comparable rates throughout S phase, so the extent of chromosome condensation seems to influence the time at which replication forks are initiated, rather than their speed once formed. A Large Multisubunit Complex Binds to Eukaryotic Origins of Replication Having seen that a eukaryotic chromosome is replicated using many origins of replication, each of which “fires” at a characteristic time in S phase of the cell cycle, we turn to the nature of these origins of replication. We saw earlier in this chapter that replication origins have been precisely defined in bacteria as specific DNA sequences that attract initiator proteins, which then assemble the DNA replication machinery. We shall see that this is the case for the single-cell budding yeast S. cerevisiae, but it appears not to be strictly true for most other eukaryotes.
Cell_Biology_Alberts. Once initiated, however, replication forks seem to move at comparable rates throughout S phase, so the extent of chromosome condensation seems to influence the time at which replication forks are initiated, rather than their speed once formed. A Large Multisubunit Complex Binds to Eukaryotic Origins of Replication Having seen that a eukaryotic chromosome is replicated using many origins of replication, each of which “fires” at a characteristic time in S phase of the cell cycle, we turn to the nature of these origins of replication. We saw earlier in this chapter that replication origins have been precisely defined in bacteria as specific DNA sequences that attract initiator proteins, which then assemble the DNA replication machinery. We shall see that this is the case for the single-cell budding yeast S. cerevisiae, but it appears not to be strictly true for most other eukaryotes.
Cell_Biology_Alberts_1206
Cell_Biology_Alberts
For budding yeast, the location of every origin of replication on each chromosome has been determined. The particular chromosome shown in Figure 5–30— chromosome III from S. cerevisiae—is one of the smallest chromosomes known, with a length less than 1/100 that of a typical human chromosome. Its major origins are spaced an average of 30,000 nucleotide pairs apart, but only a subset of these origins is used by a given cell. Nonetheless, this chromosome can be replicated in about 15 minutes.
Cell_Biology_Alberts. For budding yeast, the location of every origin of replication on each chromosome has been determined. The particular chromosome shown in Figure 5–30— chromosome III from S. cerevisiae—is one of the smallest chromosomes known, with a length less than 1/100 that of a typical human chromosome. Its major origins are spaced an average of 30,000 nucleotide pairs apart, but only a subset of these origins is used by a given cell. Nonetheless, this chromosome can be replicated in about 15 minutes.
Cell_Biology_Alberts_1207
Cell_Biology_Alberts
The minimal DNA sequence required for directing DNA replication initiation in S. cerevisiae has been determined by taking a segment of DNA that spans an origin of replication and testing smaller and smaller DNA fragments for their ability to function as origins. Most DNA sequences that can serve as an origin of replication are found to contain (1) a binding site for a large, multisubunit initiator protein called ORC, for origin recognition complex; (2) a stretch of DNA that is rich in As and Ts and therefore easy to melt; and (3) at least one binding site for proteins that facilitate ORC binding, probably by adjusting chromatin structure.
Cell_Biology_Alberts. The minimal DNA sequence required for directing DNA replication initiation in S. cerevisiae has been determined by taking a segment of DNA that spans an origin of replication and testing smaller and smaller DNA fragments for their ability to function as origins. Most DNA sequences that can serve as an origin of replication are found to contain (1) a binding site for a large, multisubunit initiator protein called ORC, for origin recognition complex; (2) a stretch of DNA that is rich in As and Ts and therefore easy to melt; and (3) at least one binding site for proteins that facilitate ORC binding, probably by adjusting chromatin structure.
Cell_Biology_Alberts_1208
Cell_Biology_Alberts
In bacteria, once the initiator protein is properly bound to the single origin of replication, the assembly of the replication forks seems to follow more or less automatically. In eukaryotes, the situation is significantly different because of a profound problem eukaryotes have in replicating chromosomes: with so many places to begin replication, how is the process regulated to ensure that all the DNA is copied once and only once?
Cell_Biology_Alberts. In bacteria, once the initiator protein is properly bound to the single origin of replication, the assembly of the replication forks seems to follow more or less automatically. In eukaryotes, the situation is significantly different because of a profound problem eukaryotes have in replicating chromosomes: with so many places to begin replication, how is the process regulated to ensure that all the DNA is copied once and only once?
Cell_Biology_Alberts_1209
Cell_Biology_Alberts
The answer lies in the sequential manner in which the replicative helicase is first loaded onto origins and is then activated to initiate DNA replication. This matter is discussed in detail in Chapter 17, where we consider the machinery that underlies the cell-division cycle. In brief, during G1 phase, the replicative helicases are loaded onto DNA next to ORC to create a prereplicative complex. Then, upon passage from G1 phase to S phase, specialized protein kinases come into play to activate the helicases. The resulting opening of the double helix allows the loading of the remaining replication proteins, including the DNA polymerases. origins of replication telomere centromere telomere
Cell_Biology_Alberts. The answer lies in the sequential manner in which the replicative helicase is first loaded onto origins and is then activated to initiate DNA replication. This matter is discussed in detail in Chapter 17, where we consider the machinery that underlies the cell-division cycle. In brief, during G1 phase, the replicative helicases are loaded onto DNA next to ORC to create a prereplicative complex. Then, upon passage from G1 phase to S phase, specialized protein kinases come into play to activate the helicases. The resulting opening of the double helix allows the loading of the remaining replication proteins, including the DNA polymerases. origins of replication telomere centromere telomere