id
stringlengths 14
28
| title
stringclasses 18
values | content
stringlengths 2
999
| contents
stringlengths 19
1.02k
|
---|---|---|---|
Cell_Biology_Alberts_710 | Cell_Biology_Alberts | Humans have invented many different types of mechanical pumps, and it should not be surprising that cells also contain membrane-bound pumps that Figure 3–76 The abC (aTP-binding cassette) transporter, a protein machine that pumps molecules through a membrane. (A) How this large family of transporters pumps molecules into the cell in bacteria. As indicated, the binding of two molecules of ATP causes the two ATP-binding domains to clamp together tightly, opening a channel to the cell exterior. The binding of a substrate molecule to the extracellular face of the protein complex then triggers ATP hydrolysis followed by ADP release, which opens the cytoplasmic gate; the pump is then reset for another cycle. | Cell_Biology_Alberts. Humans have invented many different types of mechanical pumps, and it should not be surprising that cells also contain membrane-bound pumps that Figure 3–76 The abC (aTP-binding cassette) transporter, a protein machine that pumps molecules through a membrane. (A) How this large family of transporters pumps molecules into the cell in bacteria. As indicated, the binding of two molecules of ATP causes the two ATP-binding domains to clamp together tightly, opening a channel to the cell exterior. The binding of a substrate molecule to the extracellular face of the protein complex then triggers ATP hydrolysis followed by ADP release, which opens the cytoplasmic gate; the pump is then reset for another cycle. |
Cell_Biology_Alberts_711 | Cell_Biology_Alberts | (B) As discussed in Chapter 11, in eukaryotes an opposite process occurs, causing selected substrate molecules to be pumped out of the cell. (C) The structure of a bacterial ABC transporter (see Movie 11.5). (C, from R.J. Dawson and K.P. Locher, Nature 443:180–185, 2006. With permission from Macmillan Publishers Ltd; PDB code: 2HYD). function in other ways. Among the most notable are the rotary pumps that couple the hydrolysis of ATP to the transport of H+ ions (protons). These pumps resemble miniature turbines, and they are used to acidify the interior of lysosomes and other eukaryotic organelles. Like other ion pumps that create ion gradients, they can function in reverse to catalyze the reaction ADP + Pi → ATP, if the gradient across their membrane of the ion that they transport is steep enough. | Cell_Biology_Alberts. (B) As discussed in Chapter 11, in eukaryotes an opposite process occurs, causing selected substrate molecules to be pumped out of the cell. (C) The structure of a bacterial ABC transporter (see Movie 11.5). (C, from R.J. Dawson and K.P. Locher, Nature 443:180–185, 2006. With permission from Macmillan Publishers Ltd; PDB code: 2HYD). function in other ways. Among the most notable are the rotary pumps that couple the hydrolysis of ATP to the transport of H+ ions (protons). These pumps resemble miniature turbines, and they are used to acidify the interior of lysosomes and other eukaryotic organelles. Like other ion pumps that create ion gradients, they can function in reverse to catalyze the reaction ADP + Pi → ATP, if the gradient across their membrane of the ion that they transport is steep enough. |
Cell_Biology_Alberts_712 | Cell_Biology_Alberts | One such pump, the ATP synthase, harnesses a gradient of proton concentration produced by electron-transport processes to produce most of the ATP used in the living world. This ubiquitous pump has a central role in energy conversion, and we shall discuss its three-dimensional structure and mechanism in Chapter 14. Proteins Often Form Large Complexes That Function as Protein Machines | Cell_Biology_Alberts. One such pump, the ATP synthase, harnesses a gradient of proton concentration produced by electron-transport processes to produce most of the ATP used in the living world. This ubiquitous pump has a central role in energy conversion, and we shall discuss its three-dimensional structure and mechanism in Chapter 14. Proteins Often Form Large Complexes That Function as Protein Machines |
Cell_Biology_Alberts_713 | Cell_Biology_Alberts | Large proteins formed from many domains are able to perform more elaborate functions than small, single-domain proteins. But large protein assemblies formed from many protein molecules linked together by noncovalent bonds perform the most impressive tasks. Now that it is possible to reconstruct most biological processes in cell-free systems in the laboratory, it is clear that each of the central processes in a cell—such as DNA replication, protein synthesis, vesicle budding, or transmembrane signaling—is catalyzed by a highly coordinated, linked set of 10 or more proteins. In most such protein machines, an energetically favorable reaction such as the hydrolysis of bound nucleoside triphosphates (ATP or GTP) drives an ordered series of conformational changes in one or more of the individual protein subunits, enabling the ensemble of proteins to move coordinately. In this way, each enzyme can be moved directly into position, as the machine catalyzes successive reactions in a series | Cell_Biology_Alberts. Large proteins formed from many domains are able to perform more elaborate functions than small, single-domain proteins. But large protein assemblies formed from many protein molecules linked together by noncovalent bonds perform the most impressive tasks. Now that it is possible to reconstruct most biological processes in cell-free systems in the laboratory, it is clear that each of the central processes in a cell—such as DNA replication, protein synthesis, vesicle budding, or transmembrane signaling—is catalyzed by a highly coordinated, linked set of 10 or more proteins. In most such protein machines, an energetically favorable reaction such as the hydrolysis of bound nucleoside triphosphates (ATP or GTP) drives an ordered series of conformational changes in one or more of the individual protein subunits, enabling the ensemble of proteins to move coordinately. In this way, each enzyme can be moved directly into position, as the machine catalyzes successive reactions in a series |
Cell_Biology_Alberts_714 | Cell_Biology_Alberts | protein subunits, enabling the ensemble of proteins to move coordinately. In this way, each enzyme can be moved directly into position, as the machine catalyzes successive reactions in a series (Figure 3–77). This is what occurs, for example, in protein synthesis on a ribosome (discussed in Chapter 6)—or in DNA replication, where a large multiprotein complex moves rapidly along the DNA (discussed in Chapter 5). | Cell_Biology_Alberts. protein subunits, enabling the ensemble of proteins to move coordinately. In this way, each enzyme can be moved directly into position, as the machine catalyzes successive reactions in a series (Figure 3–77). This is what occurs, for example, in protein synthesis on a ribosome (discussed in Chapter 6)—or in DNA replication, where a large multiprotein complex moves rapidly along the DNA (discussed in Chapter 5). |
Cell_Biology_Alberts_715 | Cell_Biology_Alberts | Cells have evolved protein machines for the same reason that humans have invented mechanical and electronic machines. For accomplishing almost any task, manipulations that are spatially and temporally coordinated through linked processes are much more efficient than the use of many separate tools. Scaffolds Concentrate Sets of Interacting Proteins | Cell_Biology_Alberts. Cells have evolved protein machines for the same reason that humans have invented mechanical and electronic machines. For accomplishing almost any task, manipulations that are spatially and temporally coordinated through linked processes are much more efficient than the use of many separate tools. Scaffolds Concentrate Sets of Interacting Proteins |
Cell_Biology_Alberts_716 | Cell_Biology_Alberts | Scaffolds Concentrate Sets of Interacting Proteins As scientists have learned more of the details of cell biology, they have recognized an increasing degree of sophistication in cell chemistry. Thus, not only do we now know that protein machines play a predominant role, but it has also become clear that they are very often localized to specific sites in the cell, being assembled and activated only where and when they are needed. As one example, when extracellular signaling molecules bind to receptor proteins in the plasma membrane, the activated receptors often recruit a set of other proteins to the inside surface of the plasma membrane to form a large protein complex that passes the signal on (discussed in Chapter 15). | Cell_Biology_Alberts. Scaffolds Concentrate Sets of Interacting Proteins As scientists have learned more of the details of cell biology, they have recognized an increasing degree of sophistication in cell chemistry. Thus, not only do we now know that protein machines play a predominant role, but it has also become clear that they are very often localized to specific sites in the cell, being assembled and activated only where and when they are needed. As one example, when extracellular signaling molecules bind to receptor proteins in the plasma membrane, the activated receptors often recruit a set of other proteins to the inside surface of the plasma membrane to form a large protein complex that passes the signal on (discussed in Chapter 15). |
Cell_Biology_Alberts_717 | Cell_Biology_Alberts | The mechanisms frequently involve scaffold proteins. These are proteins with binding sites for multiple other proteins, and they serve both to link together specific sets of interacting proteins and to position them at specific locations inside a cell. At one extreme are rigid scaffolds, such as the cullin in SCF ubiquitin ligase (see Figure 3–71). At the other extreme are the large, flexible scaffold proteins that often underlie regions of specialized plasma membrane. These include the Figure 3–77 How “protein machines” carry out complex functions. | Cell_Biology_Alberts. The mechanisms frequently involve scaffold proteins. These are proteins with binding sites for multiple other proteins, and they serve both to link together specific sets of interacting proteins and to position them at specific locations inside a cell. At one extreme are rigid scaffolds, such as the cullin in SCF ubiquitin ligase (see Figure 3–71). At the other extreme are the large, flexible scaffold proteins that often underlie regions of specialized plasma membrane. These include the Figure 3–77 How “protein machines” carry out complex functions. |
Cell_Biology_Alberts_718 | Cell_Biology_Alberts | Figure 3–77 How “protein machines” carry out complex functions. These machines are made of individual proteins that collaborate to perform a specific task (Movie 3.13). The movement of these proteins is often coordinated by the hydrolysis of a bound nucleotide such as ATP or GTP. Directional allosteric conformational changes of proteins that are driven in this way often occur in a large protein assembly in which the activities of several different protein molecules are coordinated by such movements within the complex. | Cell_Biology_Alberts. Figure 3–77 How “protein machines” carry out complex functions. These machines are made of individual proteins that collaborate to perform a specific task (Movie 3.13). The movement of these proteins is often coordinated by the hydrolysis of a bound nucleotide such as ATP or GTP. Directional allosteric conformational changes of proteins that are driven in this way often occur in a large protein assembly in which the activities of several different protein molecules are coordinated by such movements within the complex. |
Cell_Biology_Alberts_719 | Cell_Biology_Alberts | Discs-large protein (Dlg), a protein of about 900 amino acids that is concentrated in special regions beneath the plasma membrane in epithelial cells and at synapses. Dlg contains binding sites for at least seven other proteins, interspersed with regions of more flexible polypeptide chain. An ancient protein, conserved in organisms as diverse as sponges, worms, flies, and humans, Dlg derives its name from the mutant phenotype of the organism in which it was first discovered; the cells in the imaginal discs of a Drosophila embryo with a mutation in the Dlg gene fail to stop proliferating when they should, and they produce unusually large discs whose epithelial cells can form tumors. | Cell_Biology_Alberts. Discs-large protein (Dlg), a protein of about 900 amino acids that is concentrated in special regions beneath the plasma membrane in epithelial cells and at synapses. Dlg contains binding sites for at least seven other proteins, interspersed with regions of more flexible polypeptide chain. An ancient protein, conserved in organisms as diverse as sponges, worms, flies, and humans, Dlg derives its name from the mutant phenotype of the organism in which it was first discovered; the cells in the imaginal discs of a Drosophila embryo with a mutation in the Dlg gene fail to stop proliferating when they should, and they produce unusually large discs whose epithelial cells can form tumors. |
Cell_Biology_Alberts_720 | Cell_Biology_Alberts | Although incompletely studied, Dlg and a large number of similar scaffold proteins are thought to function like the protein that is schematically illustrated in Figure 3–78. By binding a specific set of interacting proteins, these scaffolds can enhance the rate of critical reactions, while also confining them to the particular region of the cell that contains the scaffold. For similar reasons, cells also make extensive use of scaffold RNA molecules, as discussed in Chapter 7. Many Proteins Are Controlled by Covalent Modifications That Direct Them to Specific Sites Inside the Cell We have thus far described only a few ways in which proteins are post-translationally modified. A large number of other such modifications also occur, more than 200 distinct types being known. To give a sense of the variety, Table 3–3 presents | Cell_Biology_Alberts. Although incompletely studied, Dlg and a large number of similar scaffold proteins are thought to function like the protein that is schematically illustrated in Figure 3–78. By binding a specific set of interacting proteins, these scaffolds can enhance the rate of critical reactions, while also confining them to the particular region of the cell that contains the scaffold. For similar reasons, cells also make extensive use of scaffold RNA molecules, as discussed in Chapter 7. Many Proteins Are Controlled by Covalent Modifications That Direct Them to Specific Sites Inside the Cell We have thus far described only a few ways in which proteins are post-translationally modified. A large number of other such modifications also occur, more than 200 distinct types being known. To give a sense of the variety, Table 3–3 presents |
Cell_Biology_Alberts_721 | Cell_Biology_Alberts | Figure 3–78 How the proximity created by scaffold proteins can greatly speed reactions in a cell. In this example, long unstructured regions of polypeptide chain in a large scaffold protein connect a series of structured domains that bind a set of reacting proteins. The unstructured regions serve as flexible “tethers” that greatly speed reaction rates by causing a rapid, random collision of all of the proteins that are bound to the scaffold. (For specific examples of protein tethering, see Figure 3–54 and Figure 16–18; for scaffold RNA molecules, see Figure 7–49B.) a few of the modifying groups with known regulatory roles. As in phosphate and ubiquitin additions described previously, these groups are added and then removed from proteins according to the needs of the cell. | Cell_Biology_Alberts. Figure 3–78 How the proximity created by scaffold proteins can greatly speed reactions in a cell. In this example, long unstructured regions of polypeptide chain in a large scaffold protein connect a series of structured domains that bind a set of reacting proteins. The unstructured regions serve as flexible “tethers” that greatly speed reaction rates by causing a rapid, random collision of all of the proteins that are bound to the scaffold. (For specific examples of protein tethering, see Figure 3–54 and Figure 16–18; for scaffold RNA molecules, see Figure 7–49B.) a few of the modifying groups with known regulatory roles. As in phosphate and ubiquitin additions described previously, these groups are added and then removed from proteins according to the needs of the cell. |
Cell_Biology_Alberts_722 | Cell_Biology_Alberts | A large number of proteins are now known to be modified on more than one amino acid side chain, with different regulatory events producing a different pattern of such modifications. A striking example is the protein p53, which plays a central part in controlling a cell’s response to adverse circumstances (see Figure 17–62). Through one of four different types of molecular additions, this protein can be modified at 20 different sites. Because an enormous number of different combinations of these 20 modifications are possible, the protein’s behavior can in principle be altered in a huge number of ways. Such modifications will often create a site on the modified protein that binds it to a scaffold protein in a specific region of the cell, thereby connecting it—via the scaffold—to the other proteins required for a reaction at that site. | Cell_Biology_Alberts. A large number of proteins are now known to be modified on more than one amino acid side chain, with different regulatory events producing a different pattern of such modifications. A striking example is the protein p53, which plays a central part in controlling a cell’s response to adverse circumstances (see Figure 17–62). Through one of four different types of molecular additions, this protein can be modified at 20 different sites. Because an enormous number of different combinations of these 20 modifications are possible, the protein’s behavior can in principle be altered in a huge number of ways. Such modifications will often create a site on the modified protein that binds it to a scaffold protein in a specific region of the cell, thereby connecting it—via the scaffold—to the other proteins required for a reaction at that site. |
Cell_Biology_Alberts_723 | Cell_Biology_Alberts | One can view each protein’s set of covalent modifications as a combinatorial regulatory code. Specific modifying groups are added to or removed from a protein in response to signals, and the code then alters protein behavior—changing the activity or stability of the protein, its binding partners, and/or its specific location within the cell (Figure 3–79). As a result, the cell is able to respond rapidly and with great versatility to changes in its condition or environment. A Complex Network of Protein Interactions Underlies Cell Function | Cell_Biology_Alberts. One can view each protein’s set of covalent modifications as a combinatorial regulatory code. Specific modifying groups are added to or removed from a protein in response to signals, and the code then alters protein behavior—changing the activity or stability of the protein, its binding partners, and/or its specific location within the cell (Figure 3–79). As a result, the cell is able to respond rapidly and with great versatility to changes in its condition or environment. A Complex Network of Protein Interactions Underlies Cell Function |
Cell_Biology_Alberts_724 | Cell_Biology_Alberts | A Complex Network of Protein Interactions Underlies Cell Function There are many challenges facing cell biologists in this information-rich era when a large number of complete genome sequences are known. One is the need to dissect and reconstruct each one of the thousands of protein machines that exist in an organism such as ourselves. To understand these remarkable protein complexes, each will need to be reconstituted from its purified protein parts, so that we can study its detailed mode of operation under controlled conditions in a test tube, free from all other cell components. This alone is a massive task. But we now know that each of these subcomponents of a cell also interacts with other sets of macromolecules, creating a large network of protein–protein and protein–nucleic acid interactions throughout the cell. To understand the cell, therefore, we will need to analyze most of these other interactions as well. | Cell_Biology_Alberts. A Complex Network of Protein Interactions Underlies Cell Function There are many challenges facing cell biologists in this information-rich era when a large number of complete genome sequences are known. One is the need to dissect and reconstruct each one of the thousands of protein machines that exist in an organism such as ourselves. To understand these remarkable protein complexes, each will need to be reconstituted from its purified protein parts, so that we can study its detailed mode of operation under controlled conditions in a test tube, free from all other cell components. This alone is a massive task. But we now know that each of these subcomponents of a cell also interacts with other sets of macromolecules, creating a large network of protein–protein and protein–nucleic acid interactions throughout the cell. To understand the cell, therefore, we will need to analyze most of these other interactions as well. |
Cell_Biology_Alberts_725 | Cell_Biology_Alberts | Figure 3–79 Multisite protein modification and its effects. (A) A protein that carries a post-translational addition to more than one of its amino acid side chains can be considered to carry a combinatorial regulatory code. Multisite modifications are added to (and removed from) a protein through signaling networks, and the resulting combinatorial regulatory code on the protein is read to alter its behavior in the cell. (B) The pattern of some covalent modifications to the protein p53. We can gain some idea of the complexity of intracellular protein networks from a particularly well-studied example described in Chapter 16: the many dozens of proteins that interact with the actin cytoskeleton to control actin filament behavior (see Panel 16–3, p. 905). | Cell_Biology_Alberts. Figure 3–79 Multisite protein modification and its effects. (A) A protein that carries a post-translational addition to more than one of its amino acid side chains can be considered to carry a combinatorial regulatory code. Multisite modifications are added to (and removed from) a protein through signaling networks, and the resulting combinatorial regulatory code on the protein is read to alter its behavior in the cell. (B) The pattern of some covalent modifications to the protein p53. We can gain some idea of the complexity of intracellular protein networks from a particularly well-studied example described in Chapter 16: the many dozens of proteins that interact with the actin cytoskeleton to control actin filament behavior (see Panel 16–3, p. 905). |
Cell_Biology_Alberts_726 | Cell_Biology_Alberts | The extent of such protein–protein interactions can also be estimated more generally. An enormous amount of valuable information is now freely available in protein databases on the Internet: tens of thousands of three-dimensional protein structures plus tens of millions of protein sequences derived from the nucleotide sequences of genes. Scientists have been developing new methods for mining this great resource to increase our understanding of cells. In particular, computer-based bioinformatics tools are being combined with robotics and other technologies to allow thousands of proteins to be investigated in a single set of experiments. Proteomics is a term that is often used to describe such research focused on the analysis of large sets of proteins, analogous to the term genomics describing the large-scale analysis of DNA sequences and genes. | Cell_Biology_Alberts. The extent of such protein–protein interactions can also be estimated more generally. An enormous amount of valuable information is now freely available in protein databases on the Internet: tens of thousands of three-dimensional protein structures plus tens of millions of protein sequences derived from the nucleotide sequences of genes. Scientists have been developing new methods for mining this great resource to increase our understanding of cells. In particular, computer-based bioinformatics tools are being combined with robotics and other technologies to allow thousands of proteins to be investigated in a single set of experiments. Proteomics is a term that is often used to describe such research focused on the analysis of large sets of proteins, analogous to the term genomics describing the large-scale analysis of DNA sequences and genes. |
Cell_Biology_Alberts_727 | Cell_Biology_Alberts | A biochemical method based on affinity tagging and mass spectroscopy has proven especially powerful for determining the direct binding interactions between the many different proteins in a cell (discussed in Chapter 8). The results are being tabulated and organized in Internet databases. This allows a cell biologist studying a small set of proteins to readily discover which other proteins in the same cell are likely to bind to, and thus interact with, that set of proteins. When displayed graphically as a protein interaction map, each protein is represented by a box or dot in a two-dimensional network, with a straight line connecting those proteins that have been found to bind to each other. | Cell_Biology_Alberts. A biochemical method based on affinity tagging and mass spectroscopy has proven especially powerful for determining the direct binding interactions between the many different proteins in a cell (discussed in Chapter 8). The results are being tabulated and organized in Internet databases. This allows a cell biologist studying a small set of proteins to readily discover which other proteins in the same cell are likely to bind to, and thus interact with, that set of proteins. When displayed graphically as a protein interaction map, each protein is represented by a box or dot in a two-dimensional network, with a straight line connecting those proteins that have been found to bind to each other. |
Cell_Biology_Alberts_728 | Cell_Biology_Alberts | When hundreds or thousands of proteins are displayed on the same map, the network diagram becomes bewilderingly complicated, serving to illustrate the enormous challenges that face scientists attempting to understand the cell (Figure 3–80). Much more useful are small subsections of these maps, centered on a few proteins of interest. | Cell_Biology_Alberts. When hundreds or thousands of proteins are displayed on the same map, the network diagram becomes bewilderingly complicated, serving to illustrate the enormous challenges that face scientists attempting to understand the cell (Figure 3–80). Much more useful are small subsections of these maps, centered on a few proteins of interest. |
Cell_Biology_Alberts_729 | Cell_Biology_Alberts | We have previously described the structure and mode of action of the SCF ubiquitin ligase, using it to illustrate how protein complexes are constructed from interchangeable parts (see Figure 3–71). Figure 3–81 shows a network of protein– protein interactions for the five proteins that form this protein complex in a yeast cell. Four of the subunits of this ligase are located at the bottom right of this figure. The remaining subunit, the F-box protein that serves as its substrate-binding arm, appears as a set of 15 different gene products that bind to adaptor protein 2 (the Skp1 protein). Along the top and left of the figure are sets of additional protein interactions marked with yellow and green shading: as indicated, these protein sets function at the origin of DNA replication, in cell cycle regulation, in methionine synthesis, in the kinetochore, and in vacuolar H+-ATPase assembly. We shall use this figure to explain how such protein interaction maps are used, and what they do and do | Cell_Biology_Alberts. We have previously described the structure and mode of action of the SCF ubiquitin ligase, using it to illustrate how protein complexes are constructed from interchangeable parts (see Figure 3–71). Figure 3–81 shows a network of protein– protein interactions for the five proteins that form this protein complex in a yeast cell. Four of the subunits of this ligase are located at the bottom right of this figure. The remaining subunit, the F-box protein that serves as its substrate-binding arm, appears as a set of 15 different gene products that bind to adaptor protein 2 (the Skp1 protein). Along the top and left of the figure are sets of additional protein interactions marked with yellow and green shading: as indicated, these protein sets function at the origin of DNA replication, in cell cycle regulation, in methionine synthesis, in the kinetochore, and in vacuolar H+-ATPase assembly. We shall use this figure to explain how such protein interaction maps are used, and what they do and do |
Cell_Biology_Alberts_730 | Cell_Biology_Alberts | regulation, in methionine synthesis, in the kinetochore, and in vacuolar H+-ATPase assembly. We shall use this figure to explain how such protein interaction maps are used, and what they do and do not mean. | Cell_Biology_Alberts. regulation, in methionine synthesis, in the kinetochore, and in vacuolar H+-ATPase assembly. We shall use this figure to explain how such protein interaction maps are used, and what they do and do not mean. |
Cell_Biology_Alberts_731 | Cell_Biology_Alberts | 1. Protein interaction maps are useful for identifying the likely function of previously uncharacterized proteins. Examples are the products of the genes that have thus far only been inferred to exist from the yeast genome sequence, which are the three proteins in the figure that lack a simple three-letter abbreviation (white letters beginning with Y). The three in this diagram are F-box proteins that bind to Skp1; these are therefore likely to function as part of the ubiquitin ligase, serving as substrate-binding arms that recognize different target proteins. However, as we discuss next, neither assignment can be considered certain without additional data. 2. | Cell_Biology_Alberts. 1. Protein interaction maps are useful for identifying the likely function of previously uncharacterized proteins. Examples are the products of the genes that have thus far only been inferred to exist from the yeast genome sequence, which are the three proteins in the figure that lack a simple three-letter abbreviation (white letters beginning with Y). The three in this diagram are F-box proteins that bind to Skp1; these are therefore likely to function as part of the ubiquitin ligase, serving as substrate-binding arms that recognize different target proteins. However, as we discuss next, neither assignment can be considered certain without additional data. 2. |
Cell_Biology_Alberts_732 | Cell_Biology_Alberts | 2. Protein interaction networks need to be interpreted with caution because, as a result of evolution making efficient use of each organism’s genetic information, the same protein can be used as part of different protein complexes that have different types of functions. Thus, although protein A binds to protein B and protein B binds to protein C, proteins A and C need not function in the same process. For example, we know from detailed biochemical studies that the functions of Skp1 in the kinetochore and in Figure 3–80 a network of protein-binding interactions in a yeast cell. Each line connecting a pair of dots (proteins) indicates a protein–protein interaction. (From A. Guimerá and M. Sales–Pardo, Mol. Syst. Biol. 2:42, 2006. With permission from Macmillan Publishers Ltd.) Figure 3–81 a map of some protein–protein interactions of the SCF ubiquitin ligase and other proteins in the yeast | Cell_Biology_Alberts. 2. Protein interaction networks need to be interpreted with caution because, as a result of evolution making efficient use of each organism’s genetic information, the same protein can be used as part of different protein complexes that have different types of functions. Thus, although protein A binds to protein B and protein B binds to protein C, proteins A and C need not function in the same process. For example, we know from detailed biochemical studies that the functions of Skp1 in the kinetochore and in Figure 3–80 a network of protein-binding interactions in a yeast cell. Each line connecting a pair of dots (proteins) indicates a protein–protein interaction. (From A. Guimerá and M. Sales–Pardo, Mol. Syst. Biol. 2:42, 2006. With permission from Macmillan Publishers Ltd.) Figure 3–81 a map of some protein–protein interactions of the SCF ubiquitin ligase and other proteins in the yeast |
Cell_Biology_Alberts_733 | Cell_Biology_Alberts | Figure 3–81 a map of some protein–protein interactions of the SCF ubiquitin ligase and other proteins in the yeast S. cerevisiae. The symbols and/or colors used for the five proteins of the ligase are those in Figure 3–71. Note that 15 different F-box proteins are shown (purple); those with white lettering (beginning with Y) are known from the genome sequence as open reading frames. For additional details, see text. (Courtesy of Peter Bowers and David Eisenberg, UCLA-DOE Institute for Genomics and Proteomics, UCLA.) vacuolar H+-ATPase assembly (yellow shading) are separate from its function in the SCF ubiquitin ligase. In fact, only the remaining three functions of Skp1 illustrated in the diagram—methionine synthesis, cell cycle regulation, and origin of replication (green shading)—involve ubiquitylation. | Cell_Biology_Alberts. Figure 3–81 a map of some protein–protein interactions of the SCF ubiquitin ligase and other proteins in the yeast S. cerevisiae. The symbols and/or colors used for the five proteins of the ligase are those in Figure 3–71. Note that 15 different F-box proteins are shown (purple); those with white lettering (beginning with Y) are known from the genome sequence as open reading frames. For additional details, see text. (Courtesy of Peter Bowers and David Eisenberg, UCLA-DOE Institute for Genomics and Proteomics, UCLA.) vacuolar H+-ATPase assembly (yellow shading) are separate from its function in the SCF ubiquitin ligase. In fact, only the remaining three functions of Skp1 illustrated in the diagram—methionine synthesis, cell cycle regulation, and origin of replication (green shading)—involve ubiquitylation. |
Cell_Biology_Alberts_734 | Cell_Biology_Alberts | 3. In cross-species comparisons, those proteins displaying similar patterns of interactions in the two protein interaction maps are likely to have the same function in the cell. Thus, as scientists generate more and more highly detailed maps for multiple organisms, the results will become increasingly useful for inferring protein function. These map comparisons will be a particularly powerful tool for deciphering the functions of human proteins, because a vast amount of direct information about protein function can be obtained from genetic engineering, mutational, and genetic analyses in experimental organisms—such as yeast, worms, and flies—that are not feasible in humans. | Cell_Biology_Alberts. 3. In cross-species comparisons, those proteins displaying similar patterns of interactions in the two protein interaction maps are likely to have the same function in the cell. Thus, as scientists generate more and more highly detailed maps for multiple organisms, the results will become increasingly useful for inferring protein function. These map comparisons will be a particularly powerful tool for deciphering the functions of human proteins, because a vast amount of direct information about protein function can be obtained from genetic engineering, mutational, and genetic analyses in experimental organisms—such as yeast, worms, and flies—that are not feasible in humans. |
Cell_Biology_Alberts_735 | Cell_Biology_Alberts | What does the future hold? There are likely to be on the order of 10,000 different proteins in a typical human cell, each of which interacts with 5 to 10 different partners. Despite the enormous progress made in recent years, we cannot yet claim to understand even the simplest known cells, such as the small Mycoplasma bacterium formed from only about 500 gene products (see Figure 1–10). How then can we hope to understand a human? Clearly, a great deal of new biochemistry will be essential, in which each protein in a particular interacting set is purified so that its chemistry and interactions can be dissected in a test tube. But in addition, more powerful ways of analyzing networks will be needed based on mathematical and computational tools not yet invented, as we shall emphasize in Chapter 8. Clearly, there are many wonderful challenges that remain for future generations of cell biologists. | Cell_Biology_Alberts. What does the future hold? There are likely to be on the order of 10,000 different proteins in a typical human cell, each of which interacts with 5 to 10 different partners. Despite the enormous progress made in recent years, we cannot yet claim to understand even the simplest known cells, such as the small Mycoplasma bacterium formed from only about 500 gene products (see Figure 1–10). How then can we hope to understand a human? Clearly, a great deal of new biochemistry will be essential, in which each protein in a particular interacting set is purified so that its chemistry and interactions can be dissected in a test tube. But in addition, more powerful ways of analyzing networks will be needed based on mathematical and computational tools not yet invented, as we shall emphasize in Chapter 8. Clearly, there are many wonderful challenges that remain for future generations of cell biologists. |
Cell_Biology_Alberts_736 | Cell_Biology_Alberts | Proteins can form enormously sophisticated chemical devices, whose functions largely depend on the detailed chemical properties of their surfaces. Binding sites for ligands are formed as surface cavities in which precisely positioned amino acid side chains are brought together by protein folding. In this way, normally unreactive amino acid side chains can be activated to make and break covalent bonds. Enzymes are catalytic proteins that greatly speed up reaction rates by binding the high-energy transition states for a specific reaction path; they also can perform acid catalysis and base catalysis simultaneously. The rates of enzyme reactions are often so fast that they are limited only by diffusion. Rates can be further increased only if enzymes that act sequentially on a substrate are joined into a single multienzyme complex, or if the enzymes and their substrates are attached to protein scaffolds, or otherwise confined to the same part of the cell. | Cell_Biology_Alberts. Proteins can form enormously sophisticated chemical devices, whose functions largely depend on the detailed chemical properties of their surfaces. Binding sites for ligands are formed as surface cavities in which precisely positioned amino acid side chains are brought together by protein folding. In this way, normally unreactive amino acid side chains can be activated to make and break covalent bonds. Enzymes are catalytic proteins that greatly speed up reaction rates by binding the high-energy transition states for a specific reaction path; they also can perform acid catalysis and base catalysis simultaneously. The rates of enzyme reactions are often so fast that they are limited only by diffusion. Rates can be further increased only if enzymes that act sequentially on a substrate are joined into a single multienzyme complex, or if the enzymes and their substrates are attached to protein scaffolds, or otherwise confined to the same part of the cell. |
Cell_Biology_Alberts_737 | Cell_Biology_Alberts | Proteins reversibly change their shape when ligands bind to their surface. The allosteric changes in protein conformation produced by one ligand affect the binding of a second ligand, and this linkage between two ligand-binding sites provides a crucial mechanism for regulating cell processes. Metabolic pathways, for example, are controlled by feedback regulation: some small molecules inhibit and other small molecules activate enzymes early in a pathway. Enzymes controlled in this way generally form symmetric assemblies, allowing cooperative conformational changes to create a steep response to changes in the concentrations of the ligands that regulate them. | Cell_Biology_Alberts. Proteins reversibly change their shape when ligands bind to their surface. The allosteric changes in protein conformation produced by one ligand affect the binding of a second ligand, and this linkage between two ligand-binding sites provides a crucial mechanism for regulating cell processes. Metabolic pathways, for example, are controlled by feedback regulation: some small molecules inhibit and other small molecules activate enzymes early in a pathway. Enzymes controlled in this way generally form symmetric assemblies, allowing cooperative conformational changes to create a steep response to changes in the concentrations of the ligands that regulate them. |
Cell_Biology_Alberts_738 | Cell_Biology_Alberts | The expenditure of chemical energy can drive unidirectional changes in protein shape. By coupling allosteric shape changes to ATP hydrolysis, for example, proteins can do useful work, such as generating a mechanical force or moving for long distances in a single direction. The three-dimensional structures of proteins have revealed how a small local change caused by nucleoside triphosphate hydrolysis is amplified to create major changes elsewhere in the protein. By such means, these proteins can serve as input–output devices that transmit information, as assembly factors, as motors, or as membrane-bound pumps. Highly efficient protein machines are formed by incorporating many different protein molecules into larger assemblies that coordinate the allosteric movements of the individual components. Such machines perform most of the important reactions in cells. | Cell_Biology_Alberts. The expenditure of chemical energy can drive unidirectional changes in protein shape. By coupling allosteric shape changes to ATP hydrolysis, for example, proteins can do useful work, such as generating a mechanical force or moving for long distances in a single direction. The three-dimensional structures of proteins have revealed how a small local change caused by nucleoside triphosphate hydrolysis is amplified to create major changes elsewhere in the protein. By such means, these proteins can serve as input–output devices that transmit information, as assembly factors, as motors, or as membrane-bound pumps. Highly efficient protein machines are formed by incorporating many different protein molecules into larger assemblies that coordinate the allosteric movements of the individual components. Such machines perform most of the important reactions in cells. |
Cell_Biology_Alberts_739 | Cell_Biology_Alberts | Proteins are subjected to many reversible, post-translational modifications, such as the covalent addition of a phosphate or an acetyl group to a specific amino acid side chain. The addition of these modifying groups is used to regulate the activity of a protein, changing its conformation, its binding to other proteins, and its location inside the cell. A typical protein in a cell will interact with more than five different partners. Through proteomics, biologists can analyze thousands of proteins in one set of experiments. One important result is the production of detailed protein interaction maps, which aim at describing all of the binding interactions between the thousands of distinct proteins in a cell. However, understanding these networks will require new biochemistry, through which small sets of interacting proteins can be purified and their chemistry dissected in detail. In addition, new computational techniques will be required to deal with the enormous complexity. | Cell_Biology_Alberts. Proteins are subjected to many reversible, post-translational modifications, such as the covalent addition of a phosphate or an acetyl group to a specific amino acid side chain. The addition of these modifying groups is used to regulate the activity of a protein, changing its conformation, its binding to other proteins, and its location inside the cell. A typical protein in a cell will interact with more than five different partners. Through proteomics, biologists can analyze thousands of proteins in one set of experiments. One important result is the production of detailed protein interaction maps, which aim at describing all of the binding interactions between the thousands of distinct proteins in a cell. However, understanding these networks will require new biochemistry, through which small sets of interacting proteins can be purified and their chemistry dissected in detail. In addition, new computational techniques will be required to deal with the enormous complexity. |
Cell_Biology_Alberts_740 | Cell_Biology_Alberts | What are the functions of the surprisingly large amount of unfolded polypeptide chain found in proteins? How many types of protein functions remain to be discovered? What are the most promising approaches for discovering them? When will scientists be able to take any amino acid sequence and accurately predict both that protein’s three-dimensional conformations and its chemical properties? What breakthroughs will be needed to accomplish this important goal? Are there ways to reveal the detailed workings of a protein machine that do not require the purification of each of its component parts in large amounts, so that the machine’s functions can be reconstituted and dissected using chemical techniques in a test tube? What are the roles of the dozens of different types of covalent modifications of proteins that have been found in addition to those listed in Table 3–3? Which ones are critical for cell function and why? | Cell_Biology_Alberts. What are the functions of the surprisingly large amount of unfolded polypeptide chain found in proteins? How many types of protein functions remain to be discovered? What are the most promising approaches for discovering them? When will scientists be able to take any amino acid sequence and accurately predict both that protein’s three-dimensional conformations and its chemical properties? What breakthroughs will be needed to accomplish this important goal? Are there ways to reveal the detailed workings of a protein machine that do not require the purification of each of its component parts in large amounts, so that the machine’s functions can be reconstituted and dissected using chemical techniques in a test tube? What are the roles of the dozens of different types of covalent modifications of proteins that have been found in addition to those listed in Table 3–3? Which ones are critical for cell function and why? |
Cell_Biology_Alberts_741 | Cell_Biology_Alberts | Why is amyloid toxic to cells and how does it contribute to neurodegenerative diseases such as Alzheimer’s disease? Which statements are true? explain why or why not. 3–1 Each strand in a β sheet is a helix with two amino acids per turn. 3–2 Intrinsically disordered regions of proteins can be identified using bioinformatic methods to search genes for encoded amino acid sequences that possess high hydrophobicity and low net charge. 3–3 Loops of polypeptide that protrude from the surface of a protein often form the binding sites for other molecules. 3–4 An enzyme reaches a maximum rate at high substrate concentration because it has a fixed number of active sites where substrate binds. 3–5 Higher concentrations of enzyme give rise to a higher turnover number. 3–6 Enzymes that undergo cooperative allosteric transitions invariably consist of symmetric assemblies of multiple subunits. | Cell_Biology_Alberts. Why is amyloid toxic to cells and how does it contribute to neurodegenerative diseases such as Alzheimer’s disease? Which statements are true? explain why or why not. 3–1 Each strand in a β sheet is a helix with two amino acids per turn. 3–2 Intrinsically disordered regions of proteins can be identified using bioinformatic methods to search genes for encoded amino acid sequences that possess high hydrophobicity and low net charge. 3–3 Loops of polypeptide that protrude from the surface of a protein often form the binding sites for other molecules. 3–4 An enzyme reaches a maximum rate at high substrate concentration because it has a fixed number of active sites where substrate binds. 3–5 Higher concentrations of enzyme give rise to a higher turnover number. 3–6 Enzymes that undergo cooperative allosteric transitions invariably consist of symmetric assemblies of multiple subunits. |
Cell_Biology_Alberts_742 | Cell_Biology_Alberts | 3–6 Enzymes that undergo cooperative allosteric transitions invariably consist of symmetric assemblies of multiple subunits. 3–7 Continual addition and removal of phosphates by protein kinases and protein phosphatases is wasteful of energy—since their combined action consumes ATP— but it is a necessary consequence of effective regulation by phosphorylation. Discuss the following problems. | Cell_Biology_Alberts. 3–6 Enzymes that undergo cooperative allosteric transitions invariably consist of symmetric assemblies of multiple subunits. 3–7 Continual addition and removal of phosphates by protein kinases and protein phosphatases is wasteful of energy—since their combined action consumes ATP— but it is a necessary consequence of effective regulation by phosphorylation. Discuss the following problems. |
Cell_Biology_Alberts_743 | Cell_Biology_Alberts | Discuss the following problems. 3–8 Consider the following statement. “To produce one molecule of each possible kind of polypeptide chain, 300 amino acids in length, would require more atoms than exist in the universe.” Given the size of the universe, do you suppose this statement could possibly be correct? Since counting atoms is a tricky business, consider the problem from the standpoint of mass. The mass of the observable universe is estimated to be about 1080 grams, give or take an order of magnitude or so. Assuming that the average One such kelch repeat domain is shown in Figure Q3–1. Would you classify this domain as an “in-line” or “plug-in” type domain? ˜7 Figure Q3–1 The kelch repeat domain of D. dendroides (Problem 3–10). The seven individual β propellers are color coded and labeled. The Nand C-termini are indicated by N and C. | Cell_Biology_Alberts. Discuss the following problems. 3–8 Consider the following statement. “To produce one molecule of each possible kind of polypeptide chain, 300 amino acids in length, would require more atoms than exist in the universe.” Given the size of the universe, do you suppose this statement could possibly be correct? Since counting atoms is a tricky business, consider the problem from the standpoint of mass. The mass of the observable universe is estimated to be about 1080 grams, give or take an order of magnitude or so. Assuming that the average One such kelch repeat domain is shown in Figure Q3–1. Would you classify this domain as an “in-line” or “plug-in” type domain? ˜7 Figure Q3–1 The kelch repeat domain of D. dendroides (Problem 3–10). The seven individual β propellers are color coded and labeled. The Nand C-termini are indicated by N and C. |
Cell_Biology_Alberts_744 | Cell_Biology_Alberts | ˜7 Figure Q3–1 The kelch repeat domain of D. dendroides (Problem 3–10). The seven individual β propellers are color coded and labeled. The Nand C-termini are indicated by N and C. 3–11 Titin, which has a molecular weight of about 3 × 106, is the largest polypeptide yet described. Titin molecules extend from muscle thick filaments to the Z disc; they are thought to act as springs to keep the thick filaments centered in the sarcomere. Titin is composed of a large number of repeated immunoglobulin (Ig) sequences of 89 amino acids, each of which is folded into a domain about 4 nm in length (Figure Q3–2A). | Cell_Biology_Alberts. ˜7 Figure Q3–1 The kelch repeat domain of D. dendroides (Problem 3–10). The seven individual β propellers are color coded and labeled. The Nand C-termini are indicated by N and C. 3–11 Titin, which has a molecular weight of about 3 × 106, is the largest polypeptide yet described. Titin molecules extend from muscle thick filaments to the Z disc; they are thought to act as springs to keep the thick filaments centered in the sarcomere. Titin is composed of a large number of repeated immunoglobulin (Ig) sequences of 89 amino acids, each of which is folded into a domain about 4 nm in length (Figure Q3–2A). |
Cell_Biology_Alberts_745 | Cell_Biology_Alberts | You suspect that the springlike behavior of titin is caused by the sequential unfolding (and refolding) of individual Ig domains. You test this hypothesis using the atomic force microscope, which allows you to pick up one end of a protein molecule and pull with an accurately measured force. For a fragment of titin containing seven repeats of the Ig domain, this experiment gives the sawtooth force-versus-extension curve shown in Figure Q3–2B. If the experiment is repeated in a solution of 8 M urea (a protein denaturant), the peaks disappear and the measured extension becomes much longer for a given force. If the experiment is repeated after the protein has been cross-linked by treatment with glutaraldehyde, once again the peaks disappear but the extension becomes much smaller for a given force. | Cell_Biology_Alberts. You suspect that the springlike behavior of titin is caused by the sequential unfolding (and refolding) of individual Ig domains. You test this hypothesis using the atomic force microscope, which allows you to pick up one end of a protein molecule and pull with an accurately measured force. For a fragment of titin containing seven repeats of the Ig domain, this experiment gives the sawtooth force-versus-extension curve shown in Figure Q3–2B. If the experiment is repeated in a solution of 8 M urea (a protein denaturant), the peaks disappear and the measured extension becomes much longer for a given force. If the experiment is repeated after the protein has been cross-linked by treatment with glutaraldehyde, once again the peaks disappear but the extension becomes much smaller for a given force. |
Cell_Biology_Alberts_746 | Cell_Biology_Alberts | mass of an amino acid is 110 daltons, what would be the mass of one molecule of each possible kind of polypeptide chain 300 amino acids in length? Is this greater than the mass of the universe? homologous proteins is to search the database using a short signature sequence indicative of the particular protein function. Why is it better to search with a short sequence than with a long sequence? Do you not have more chances 0 for a “hit” in the database with a long sequence? 3–10 The so-called kelch motif consists of a four-stranded β sheet, which forms what is known as a β pro- Figure Q3–2 Springlike behavior of titin (Problem 3–11). (A) The peller. It is usually found to be repeated four to seven times, structure of an individual Ig domain. (B) Force in piconewtons versus forming a kelch repeat domain in a multidomain protein. extension in nanometers obtained by atomic force microscopy. a. | Cell_Biology_Alberts. mass of an amino acid is 110 daltons, what would be the mass of one molecule of each possible kind of polypeptide chain 300 amino acids in length? Is this greater than the mass of the universe? homologous proteins is to search the database using a short signature sequence indicative of the particular protein function. Why is it better to search with a short sequence than with a long sequence? Do you not have more chances 0 for a “hit” in the database with a long sequence? 3–10 The so-called kelch motif consists of a four-stranded β sheet, which forms what is known as a β pro- Figure Q3–2 Springlike behavior of titin (Problem 3–11). (A) The peller. It is usually found to be repeated four to seven times, structure of an individual Ig domain. (B) Force in piconewtons versus forming a kelch repeat domain in a multidomain protein. extension in nanometers obtained by atomic force microscopy. a. |
Cell_Biology_Alberts_747 | Cell_Biology_Alberts | a. Are the data consistent with your hypothesis that titin’s springlike behavior is due to the sequential unfolding of individual Ig domains? Explain your reasoning. b. Is the extension for each putative domain-unfolding event the magnitude you would expect? (In an extended polypeptide chain, amino acids are spaced at intervals of 0.34 nm.) C. Why is each successive peak in Figure Q3–2B a little higher than the one before? D. Why does the force collapse so abruptly after each peak? | Cell_Biology_Alberts. a. Are the data consistent with your hypothesis that titin’s springlike behavior is due to the sequential unfolding of individual Ig domains? Explain your reasoning. b. Is the extension for each putative domain-unfolding event the magnitude you would expect? (In an extended polypeptide chain, amino acids are spaced at intervals of 0.34 nm.) C. Why is each successive peak in Figure Q3–2B a little higher than the one before? D. Why does the force collapse so abruptly after each peak? |
Cell_Biology_Alberts_748 | Cell_Biology_Alberts | C. Why is each successive peak in Figure Q3–2B a little higher than the one before? D. Why does the force collapse so abruptly after each peak? 3–12 Rous sarcoma virus (RSV) carries an oncogene called Src, which encodes a continuously active protein tyrosine kinase that leads to unchecked cell proliferation. Normally, Src carries an attached fatty acid (myristoylate) group that allows it to bind to the cytoplasmic side of the plasma membrane. A mutant version of Src that does not allow attachment of myristoylate does not bind to the membrane. Infection of cells with RSV encoding either the normal or the mutant form of Src leads to the same high level of protein tyrosine kinase activity, but the mutant Src does not cause cell proliferation. a. | Cell_Biology_Alberts. C. Why is each successive peak in Figure Q3–2B a little higher than the one before? D. Why does the force collapse so abruptly after each peak? 3–12 Rous sarcoma virus (RSV) carries an oncogene called Src, which encodes a continuously active protein tyrosine kinase that leads to unchecked cell proliferation. Normally, Src carries an attached fatty acid (myristoylate) group that allows it to bind to the cytoplasmic side of the plasma membrane. A mutant version of Src that does not allow attachment of myristoylate does not bind to the membrane. Infection of cells with RSV encoding either the normal or the mutant form of Src leads to the same high level of protein tyrosine kinase activity, but the mutant Src does not cause cell proliferation. a. |
Cell_Biology_Alberts_749 | Cell_Biology_Alberts | a. Assuming that the normal Src is all bound to the plasma membrane and that the mutant Src is distributed throughout the cytoplasm, calculate their relative concentrations in the neighborhood of the plasma membrane. For the purposes of this calculation, assume that the cell is a sphere with a radius (r) of 10 μm and that the mutant Src is distributed throughout the cell, whereas the normal Src is confined to a 4-nm-thick layer immediately beneath the membrane. [For this problem, assume that the membrane has no thickness. The volume of a sphere is (4/3)πr3.] b. The target (X) for phosphorylation by Src resides in the membrane. Explain why the mutant Src does not cause cell proliferation. | Cell_Biology_Alberts. a. Assuming that the normal Src is all bound to the plasma membrane and that the mutant Src is distributed throughout the cytoplasm, calculate their relative concentrations in the neighborhood of the plasma membrane. For the purposes of this calculation, assume that the cell is a sphere with a radius (r) of 10 μm and that the mutant Src is distributed throughout the cell, whereas the normal Src is confined to a 4-nm-thick layer immediately beneath the membrane. [For this problem, assume that the membrane has no thickness. The volume of a sphere is (4/3)πr3.] b. The target (X) for phosphorylation by Src resides in the membrane. Explain why the mutant Src does not cause cell proliferation. |
Cell_Biology_Alberts_750 | Cell_Biology_Alberts | 3–13 An antibody binds to another protein with an equilibrium constant, K, of 5 × 109 M–1. When it binds to a second, related protein, it forms three fewer hydrogen bonds, reducing its binding affinity by 11.9 kJ/mole. What is the K for its binding to the second protein? (Free-energy change is related to the equilibrium constant by the equation ΔG° = –2.3 RT log K, where R is 8.3 × 10–3 kJ/(mole K) and T is 310 K.) 3–14 The protein SmpB binds to a special species of tRNA, tmRNA, to eliminate the incomplete proteins made from truncated mRNAs in bacteria. If the binding of SmpB to tmRNA is plotted as fraction tmRNA bound versus SmpB concentration, one obtains a symmetrical S-shaped curve as shown in Figure Q3–3. This curve is a visual display of a very useful relationship between Kd and concentration, which has broad applicability. The general expression for fraction of ligand bound is derived from the equation for Kd(Kd = [Pr][L]/[Pr–L]) by substituting ([L]TOT – [L]) for [Pr–L] and | Cell_Biology_Alberts. 3–13 An antibody binds to another protein with an equilibrium constant, K, of 5 × 109 M–1. When it binds to a second, related protein, it forms three fewer hydrogen bonds, reducing its binding affinity by 11.9 kJ/mole. What is the K for its binding to the second protein? (Free-energy change is related to the equilibrium constant by the equation ΔG° = –2.3 RT log K, where R is 8.3 × 10–3 kJ/(mole K) and T is 310 K.) 3–14 The protein SmpB binds to a special species of tRNA, tmRNA, to eliminate the incomplete proteins made from truncated mRNAs in bacteria. If the binding of SmpB to tmRNA is plotted as fraction tmRNA bound versus SmpB concentration, one obtains a symmetrical S-shaped curve as shown in Figure Q3–3. This curve is a visual display of a very useful relationship between Kd and concentration, which has broad applicability. The general expression for fraction of ligand bound is derived from the equation for Kd(Kd = [Pr][L]/[Pr–L]) by substituting ([L]TOT – [L]) for [Pr–L] and |
Cell_Biology_Alberts_751 | Cell_Biology_Alberts | which has broad applicability. The general expression for fraction of ligand bound is derived from the equation for Kd(Kd = [Pr][L]/[Pr–L]) by substituting ([L]TOT – [L]) for [Pr–L] and rearranging. Because the total concentration of ligand ([L]TOT) is equal to the free ligand ([L]) plus bound ligand ([Pr–L]), 1.0 | Cell_Biology_Alberts. which has broad applicability. The general expression for fraction of ligand bound is derived from the equation for Kd(Kd = [Pr][L]/[Pr–L]) by substituting ([L]TOT – [L]) for [Pr–L] and rearranging. Because the total concentration of ligand ([L]TOT) is equal to the free ligand ([L]) plus bound ligand ([Pr–L]), 1.0 |
Cell_Biology_Alberts_752 | Cell_Biology_Alberts | Figure Q3–3 Fraction of tmRNA bound versus SmpB concentration (Problem 3–14). concentration of SmpB (M) For SmpB and tmRNA, the fraction bound = [SmpB– tmRNA]/[tmRNA]TOT = [SmpB]/([SmpB] + Kd). Using this relationship, calculate the fraction of tmRNA bound for SmpB concentrations equal to 104 Kd, 103 Kd, 102 Kd, 101 Kd, Kd, 10–1 Kd, 10–2 Kd, 10–3 Kd, and 10–4 Kd. 3–15 Many enzymes obey simple Michaelis–Menten kinetics, which are summarized by the equation where Vmax = maximum velocity, [S] = concentration of substrate, and Km = the Michaelis constant. It is instructive to plug a few values of [S] into the equation to see how rate is affected. What are the rates for [S] equal to zero, equal to Km, and equal to infinite concentration? 3–16 The enzyme hexokinase adds a phosphate to D-glucose but ignores its mirror image, L-glucose. Suppose that you were able to synthesize hexokinase entirely from D-amino acids, which are the mirror image of the normal L-amino acids. a. | Cell_Biology_Alberts. Figure Q3–3 Fraction of tmRNA bound versus SmpB concentration (Problem 3–14). concentration of SmpB (M) For SmpB and tmRNA, the fraction bound = [SmpB– tmRNA]/[tmRNA]TOT = [SmpB]/([SmpB] + Kd). Using this relationship, calculate the fraction of tmRNA bound for SmpB concentrations equal to 104 Kd, 103 Kd, 102 Kd, 101 Kd, Kd, 10–1 Kd, 10–2 Kd, 10–3 Kd, and 10–4 Kd. 3–15 Many enzymes obey simple Michaelis–Menten kinetics, which are summarized by the equation where Vmax = maximum velocity, [S] = concentration of substrate, and Km = the Michaelis constant. It is instructive to plug a few values of [S] into the equation to see how rate is affected. What are the rates for [S] equal to zero, equal to Km, and equal to infinite concentration? 3–16 The enzyme hexokinase adds a phosphate to D-glucose but ignores its mirror image, L-glucose. Suppose that you were able to synthesize hexokinase entirely from D-amino acids, which are the mirror image of the normal L-amino acids. a. |
Cell_Biology_Alberts_753 | Cell_Biology_Alberts | a. Assuming that the “D” enzyme would fold to a stable conformation, what relationship would you expect it to bear to the normal “L” enzyme? b. Do you suppose the “D” enzyme would add a phosphate to L-glucose, and ignore D-glucose? 3–17 How do you suppose that a molecule of hemoglobin is able to bind oxygen efficiently in the lungs, and yet release it efficiently in the tissues? 3–18 Synthesis of the purine nucleotides AMP and GMP proceeds by a branched pathway starting with ribose 5-phosphate (R5P), as shown schematically in Figure Q3–4. Using the principles of feedback inhibition, propose a regulatory strategy for this pathway that ensures an adequate supply of both AMP and GMP and minimizes the buildup of the intermediates (A–I) when supplies of AMP and GMP are adequate. 0 0.25 0.5 0.75 Figure Q3–4 Schematic diagram of the metabolic pathway for synthesis of AMP and GMP from R5P (Problem 3–18). | Cell_Biology_Alberts. a. Assuming that the “D” enzyme would fold to a stable conformation, what relationship would you expect it to bear to the normal “L” enzyme? b. Do you suppose the “D” enzyme would add a phosphate to L-glucose, and ignore D-glucose? 3–17 How do you suppose that a molecule of hemoglobin is able to bind oxygen efficiently in the lungs, and yet release it efficiently in the tissues? 3–18 Synthesis of the purine nucleotides AMP and GMP proceeds by a branched pathway starting with ribose 5-phosphate (R5P), as shown schematically in Figure Q3–4. Using the principles of feedback inhibition, propose a regulatory strategy for this pathway that ensures an adequate supply of both AMP and GMP and minimizes the buildup of the intermediates (A–I) when supplies of AMP and GMP are adequate. 0 0.25 0.5 0.75 Figure Q3–4 Schematic diagram of the metabolic pathway for synthesis of AMP and GMP from R5P (Problem 3–18). |
Cell_Biology_Alberts_754 | Cell_Biology_Alberts | 0 0.25 0.5 0.75 Figure Q3–4 Schematic diagram of the metabolic pathway for synthesis of AMP and GMP from R5P (Problem 3–18). Berg JM, Tymoczko JL & Stryer L (2011) Biochemistry, 7th ed. New York: WH Freeman. Branden C & Tooze J (1999) Introduction to Protein Structure, 2nd ed. New York: Garland Science. Dickerson RE (2005) Present at the Flood: How Structural Molecular Biology Came About. Sunderland, MA: Sinauer. Kuriyan J, Konforti B & Wemmer D (2013) The Molecules of Life: Physical and Chemical Principles. New York: Garland Science. Perutz M (1992) Protein Structure: New Approaches to Disease and Therapy. New York: WH Freeman. Petsko GA & Ringe D (2004) Protein Structure and Function. London: New Science Press. Williamson M (2011) How Proteins Work. New York: Garland Science. The Shape and Structure of Proteins Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181, 223–230. | Cell_Biology_Alberts. 0 0.25 0.5 0.75 Figure Q3–4 Schematic diagram of the metabolic pathway for synthesis of AMP and GMP from R5P (Problem 3–18). Berg JM, Tymoczko JL & Stryer L (2011) Biochemistry, 7th ed. New York: WH Freeman. Branden C & Tooze J (1999) Introduction to Protein Structure, 2nd ed. New York: Garland Science. Dickerson RE (2005) Present at the Flood: How Structural Molecular Biology Came About. Sunderland, MA: Sinauer. Kuriyan J, Konforti B & Wemmer D (2013) The Molecules of Life: Physical and Chemical Principles. New York: Garland Science. Perutz M (1992) Protein Structure: New Approaches to Disease and Therapy. New York: WH Freeman. Petsko GA & Ringe D (2004) Protein Structure and Function. London: New Science Press. Williamson M (2011) How Proteins Work. New York: Garland Science. The Shape and Structure of Proteins Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181, 223–230. |
Cell_Biology_Alberts_755 | Cell_Biology_Alberts | Williamson M (2011) How Proteins Work. New York: Garland Science. The Shape and Structure of Proteins Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181, 223–230. Caspar DLD & Klug A (1962) Physical principles in the construction of regular viruses. Cold Spring Harb. Symp. Quant. Biol. 27, 1–24. Dill KA & MacCallum JL (2012) The protein-folding problem, 50 years on. Science 338, 1042–1046. Eisenberg D (2003) The discovery of the alpha-helix and beta-sheet, the principal structural features of proteins. Proc. Natl Acad. Sci. USA 100, 11207–11210. Fraenkel-Conrat H & Williams RC (1955) Reconstitution of active tobacco mosaic virus from its inactive protein and nucleic acid components. Proc. Natl Acad. Sci. USA 41, 690–698. Goodsell DS & Olson AJ (2000) Structural symmetry and protein function. Annu. Rev. Biophys. Biomol. Struct. 29, 105–153. | Cell_Biology_Alberts. Williamson M (2011) How Proteins Work. New York: Garland Science. The Shape and Structure of Proteins Anfinsen CB (1973) Principles that govern the folding of protein chains. Science 181, 223–230. Caspar DLD & Klug A (1962) Physical principles in the construction of regular viruses. Cold Spring Harb. Symp. Quant. Biol. 27, 1–24. Dill KA & MacCallum JL (2012) The protein-folding problem, 50 years on. Science 338, 1042–1046. Eisenberg D (2003) The discovery of the alpha-helix and beta-sheet, the principal structural features of proteins. Proc. Natl Acad. Sci. USA 100, 11207–11210. Fraenkel-Conrat H & Williams RC (1955) Reconstitution of active tobacco mosaic virus from its inactive protein and nucleic acid components. Proc. Natl Acad. Sci. USA 41, 690–698. Goodsell DS & Olson AJ (2000) Structural symmetry and protein function. Annu. Rev. Biophys. Biomol. Struct. 29, 105–153. |
Cell_Biology_Alberts_756 | Cell_Biology_Alberts | Goodsell DS & Olson AJ (2000) Structural symmetry and protein function. Annu. Rev. Biophys. Biomol. Struct. 29, 105–153. Greenwald J & Riek R (2010) Biology of amyloid: structure, function, and regulation. Structure 18, 1244–1260. Harrison SC (1992) Viruses. Curr. Opin. Struct. Biol. 2, 293–299. Hudder A, Nathanson L & Deutscher MP (2003) Organization of mammalian cytoplasm. Mol. Cell. Biol. 23, 9318–9326. Kato M, Han TW, Xie S et al. (2012) Cell-free formation of RNA granules: low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767. Koga N, Tatsumi-Koga R, Liu G et al. (2012) Principles for designing ideal protein structures. Nature 491, 222–227. Li P, Banjade S, Cheng H-C et al. (2012) Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340. | Cell_Biology_Alberts. Goodsell DS & Olson AJ (2000) Structural symmetry and protein function. Annu. Rev. Biophys. Biomol. Struct. 29, 105–153. Greenwald J & Riek R (2010) Biology of amyloid: structure, function, and regulation. Structure 18, 1244–1260. Harrison SC (1992) Viruses. Curr. Opin. Struct. Biol. 2, 293–299. Hudder A, Nathanson L & Deutscher MP (2003) Organization of mammalian cytoplasm. Mol. Cell. Biol. 23, 9318–9326. Kato M, Han TW, Xie S et al. (2012) Cell-free formation of RNA granules: low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767. Koga N, Tatsumi-Koga R, Liu G et al. (2012) Principles for designing ideal protein structures. Nature 491, 222–227. Li P, Banjade S, Cheng H-C et al. (2012) Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340. |
Cell_Biology_Alberts_757 | Cell_Biology_Alberts | Li P, Banjade S, Cheng H-C et al. (2012) Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340. Lindquist SL & Kelly JW (2011) Chemical and biological approaches for adapting proteostasis to ameliorate protein misfolding and aggregation diseases—progress and prognosis. Cold Spring Harb. Perspect. Biol. 3, a004507. Maji SK, Perrin MH, Sawaya MR et al. (2009) Functional amyloids as natural storage of peptide hormones in pituitary secretory granules. Science 325, 328–332. Nelson R, Sawaya MR, Balbirnie M et al. (2005) Structure of the cross-β spine of amyloid-like fibrils. Nature 435, 773–778. Nomura M (1973) Assembly of bacterial ribosomes. Science 179, 864–873. Oldfield CJ & Dunker AK (2014) Intrinsically disordered proteins and intrinsically disordered protein regions. Annu. Rev. Biochem. 83, 553–584. Orengo CA & Thornton JM (2005) Protein families and their evolution—a structural perspective. Annu. Rev. Biochem. 74, 867–900. | Cell_Biology_Alberts. Li P, Banjade S, Cheng H-C et al. (2012) Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340. Lindquist SL & Kelly JW (2011) Chemical and biological approaches for adapting proteostasis to ameliorate protein misfolding and aggregation diseases—progress and prognosis. Cold Spring Harb. Perspect. Biol. 3, a004507. Maji SK, Perrin MH, Sawaya MR et al. (2009) Functional amyloids as natural storage of peptide hormones in pituitary secretory granules. Science 325, 328–332. Nelson R, Sawaya MR, Balbirnie M et al. (2005) Structure of the cross-β spine of amyloid-like fibrils. Nature 435, 773–778. Nomura M (1973) Assembly of bacterial ribosomes. Science 179, 864–873. Oldfield CJ & Dunker AK (2014) Intrinsically disordered proteins and intrinsically disordered protein regions. Annu. Rev. Biochem. 83, 553–584. Orengo CA & Thornton JM (2005) Protein families and their evolution—a structural perspective. Annu. Rev. Biochem. 74, 867–900. |
Cell_Biology_Alberts_758 | Cell_Biology_Alberts | Orengo CA & Thornton JM (2005) Protein families and their evolution—a structural perspective. Annu. Rev. Biochem. 74, 867–900. Pauling L & Corey RB (1951) Configurations of polypeptide chains with favored orientations around single bonds: two new pleated sheets. Proc. Natl Acad. Sci. USA 37, 729–740. Pauling L, Corey RB & Branson HR (1951) The structure of proteins: two hydrogen-bonded helical configurations of the polypeptide chain. Proc. Natl Acad. Sci. USA 37, 205–211. Prusiner SB (1998) Prions. Proc. Natl Acad. Sci. USA 95, 13363–13383. Toyama BH & Weissman JS (2011) Amyloid structure: conformational diversity and consequences. Annu. Rev. Biochem. 80, 557–585. Zhang C & Kim SH (2003) Overview of structural genomics: from structure to function. Curr. Opin. Chem. Biol. 7, 28–32. Alberts B (1998) The cell as a collection of protein machines: preparing the next generation of molecular biologists. Cell 92, 291–294. | Cell_Biology_Alberts. Orengo CA & Thornton JM (2005) Protein families and their evolution—a structural perspective. Annu. Rev. Biochem. 74, 867–900. Pauling L & Corey RB (1951) Configurations of polypeptide chains with favored orientations around single bonds: two new pleated sheets. Proc. Natl Acad. Sci. USA 37, 729–740. Pauling L, Corey RB & Branson HR (1951) The structure of proteins: two hydrogen-bonded helical configurations of the polypeptide chain. Proc. Natl Acad. Sci. USA 37, 205–211. Prusiner SB (1998) Prions. Proc. Natl Acad. Sci. USA 95, 13363–13383. Toyama BH & Weissman JS (2011) Amyloid structure: conformational diversity and consequences. Annu. Rev. Biochem. 80, 557–585. Zhang C & Kim SH (2003) Overview of structural genomics: from structure to function. Curr. Opin. Chem. Biol. 7, 28–32. Alberts B (1998) The cell as a collection of protein machines: preparing the next generation of molecular biologists. Cell 92, 291–294. |
Cell_Biology_Alberts_759 | Cell_Biology_Alberts | Alberts B (1998) The cell as a collection of protein machines: preparing the next generation of molecular biologists. Cell 92, 291–294. Benkovic SJ (1992) Catalytic antibodies. Annu. Rev. Biochem. 61, 29–54. Berg OG & von Hippel PH (1985) Diffusion-controlled macromolecular interactions. Annu. Rev. Biophys. Biophys. Chem. 14, 131–160. Bourne HR (1995) GTPases: a family of molecular switches and clocks. Philos. Trans. R. Soc. Lond. B Biol. Sci. 349, 283–289. Costanzo M, Baryshnikova A, Bellay J et al. (2010) The genetic landscape of a cell. Science 327, 425–431. Deshaies RJ & Joazeiro CAP (2009) RING domain E3 ubiquitin ligases. Annu. Rev. Biochem. 78, 399–434. Dickerson RE & Geis I (1983) Hemoglobin: Structure, Function, Evolution, and Pathology. Menlo Park, CA: Benjamin Cummings. Fersht AR (1999) Structure and Mechanism in Protein Science: A Guide to Enzyme Catalysis and Protein Folding. New York: WH Freeman. | Cell_Biology_Alberts. Alberts B (1998) The cell as a collection of protein machines: preparing the next generation of molecular biologists. Cell 92, 291–294. Benkovic SJ (1992) Catalytic antibodies. Annu. Rev. Biochem. 61, 29–54. Berg OG & von Hippel PH (1985) Diffusion-controlled macromolecular interactions. Annu. Rev. Biophys. Biophys. Chem. 14, 131–160. Bourne HR (1995) GTPases: a family of molecular switches and clocks. Philos. Trans. R. Soc. Lond. B Biol. Sci. 349, 283–289. Costanzo M, Baryshnikova A, Bellay J et al. (2010) The genetic landscape of a cell. Science 327, 425–431. Deshaies RJ & Joazeiro CAP (2009) RING domain E3 ubiquitin ligases. Annu. Rev. Biochem. 78, 399–434. Dickerson RE & Geis I (1983) Hemoglobin: Structure, Function, Evolution, and Pathology. Menlo Park, CA: Benjamin Cummings. Fersht AR (1999) Structure and Mechanism in Protein Science: A Guide to Enzyme Catalysis and Protein Folding. New York: WH Freeman. |
Cell_Biology_Alberts_760 | Cell_Biology_Alberts | Fersht AR (1999) Structure and Mechanism in Protein Science: A Guide to Enzyme Catalysis and Protein Folding. New York: WH Freeman. Haucke V, Neher E & Sigrist SJ (2011) Protein scaffolds in the coupling of synaptic exocytosis and endocytosis. Nat. Rev. Neurosci. 12, 127–138. Hua Z & Vierstra RD (2011) The cullin-RING ubiquitin-protein ligases. Annu. Rev. Plant Biol. 62, 299–334. Hunter T (2012) Why nature chose phosphate to modify proteins. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 2513–2516. Johnson LN & Lewis RJ (2001) Structural basis for control by phosphorylation. Chem. Rev. 101, 2209–2242. Kantrowitz ER & Lipscomb WN (1988) Escherichia coli aspartate transcarbamylase: the relation between structure and function. Science 241, 669–674. Kerscher O, Felberbaum R & Hochstrasser M (2006) Modification of proteins by ubiquitin and ubiquitin-like proteins. Annu. Rev. Cell Dev. Biol. 22, 159–180. | Cell_Biology_Alberts. Fersht AR (1999) Structure and Mechanism in Protein Science: A Guide to Enzyme Catalysis and Protein Folding. New York: WH Freeman. Haucke V, Neher E & Sigrist SJ (2011) Protein scaffolds in the coupling of synaptic exocytosis and endocytosis. Nat. Rev. Neurosci. 12, 127–138. Hua Z & Vierstra RD (2011) The cullin-RING ubiquitin-protein ligases. Annu. Rev. Plant Biol. 62, 299–334. Hunter T (2012) Why nature chose phosphate to modify proteins. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 2513–2516. Johnson LN & Lewis RJ (2001) Structural basis for control by phosphorylation. Chem. Rev. 101, 2209–2242. Kantrowitz ER & Lipscomb WN (1988) Escherichia coli aspartate transcarbamylase: the relation between structure and function. Science 241, 669–674. Kerscher O, Felberbaum R & Hochstrasser M (2006) Modification of proteins by ubiquitin and ubiquitin-like proteins. Annu. Rev. Cell Dev. Biol. 22, 159–180. |
Cell_Biology_Alberts_761 | Cell_Biology_Alberts | Kerscher O, Felberbaum R & Hochstrasser M (2006) Modification of proteins by ubiquitin and ubiquitin-like proteins. Annu. Rev. Cell Dev. Biol. 22, 159–180. Kim E & Sheng M (2004) PDZ domain proteins of synapses. Nat. Rev. Neurosci. 5, 771–781. Koshland DE Jr (1984) Control of enzyme activity and metabolic pathways. Trends Biochem. Sci. 9, 155–159. Krogan NJ, Cagney G, Yu H et al. (2006) Global landscape of protein complexes in the yeast Saccharomyces cerevisiae. Nature 440, 637–643. Lichtarge O, Bourne HR & Cohen FE (1996) An evolutionary trace method defines binding surfaces common to protein families. J. Mol. Biol. 257, 342–358. Maier T, Leibundgut M & Ban N (2008) The crystal structure of a mammalian fatty acid synthase. Science 321, 1315–1322. Monod J, Changeux JP & Jacob F (1963) Allosteric proteins and cellular control systems. J. Mol. Biol. 6, 306–329. | Cell_Biology_Alberts. Kerscher O, Felberbaum R & Hochstrasser M (2006) Modification of proteins by ubiquitin and ubiquitin-like proteins. Annu. Rev. Cell Dev. Biol. 22, 159–180. Kim E & Sheng M (2004) PDZ domain proteins of synapses. Nat. Rev. Neurosci. 5, 771–781. Koshland DE Jr (1984) Control of enzyme activity and metabolic pathways. Trends Biochem. Sci. 9, 155–159. Krogan NJ, Cagney G, Yu H et al. (2006) Global landscape of protein complexes in the yeast Saccharomyces cerevisiae. Nature 440, 637–643. Lichtarge O, Bourne HR & Cohen FE (1996) An evolutionary trace method defines binding surfaces common to protein families. J. Mol. Biol. 257, 342–358. Maier T, Leibundgut M & Ban N (2008) The crystal structure of a mammalian fatty acid synthase. Science 321, 1315–1322. Monod J, Changeux JP & Jacob F (1963) Allosteric proteins and cellular control systems. J. Mol. Biol. 6, 306–329. |
Cell_Biology_Alberts_762 | Cell_Biology_Alberts | Monod J, Changeux JP & Jacob F (1963) Allosteric proteins and cellular control systems. J. Mol. Biol. 6, 306–329. Perutz M (1990) Mechanisms of Cooperativity and Allosteric Regulation in Proteins. Cambridge: Cambridge University Press. Radzicka A & Wolfenden R (1995) A proficient enzyme. Science 267, 90–93. Schramm VL (2011) Enzymatic transition states, transition-state analogs, dynamics, thermodynamics, and lifetimes. Annu. Rev. Biochem. 80, 703–732. Scott JD & Pawson T (2009) Cell signaling in space and time: where proteins come together and when they’re apart. Science 326, 1220–1224. Taylor SS, Keshwani MM, Steichen JM & Kornev AP (2012) Evolution of the eukaryotic protein kinases as dynamic molecular switches. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 2517–2528. Vale RD & Milligan RA (2000) The way things move: looking under the hood of molecular motor proteins. Science 288, 88–95. | Cell_Biology_Alberts. Monod J, Changeux JP & Jacob F (1963) Allosteric proteins and cellular control systems. J. Mol. Biol. 6, 306–329. Perutz M (1990) Mechanisms of Cooperativity and Allosteric Regulation in Proteins. Cambridge: Cambridge University Press. Radzicka A & Wolfenden R (1995) A proficient enzyme. Science 267, 90–93. Schramm VL (2011) Enzymatic transition states, transition-state analogs, dynamics, thermodynamics, and lifetimes. Annu. Rev. Biochem. 80, 703–732. Scott JD & Pawson T (2009) Cell signaling in space and time: where proteins come together and when they’re apart. Science 326, 1220–1224. Taylor SS, Keshwani MM, Steichen JM & Kornev AP (2012) Evolution of the eukaryotic protein kinases as dynamic molecular switches. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 2517–2528. Vale RD & Milligan RA (2000) The way things move: looking under the hood of molecular motor proteins. Science 288, 88–95. |
Cell_Biology_Alberts_763 | Cell_Biology_Alberts | Vale RD & Milligan RA (2000) The way things move: looking under the hood of molecular motor proteins. Science 288, 88–95. Wilson MZ & Gitai Z (2013) Beyond the cytoskeleton: mesoscale assemblies and their function in spatial organization. Curr. Opin. Microbiol. 16, 177–183. DNA, Chromosomes, and genomes Life depends on the ability of cells to store, retrieve, and translate the genetic instructions required to make and maintain a living organism. This hereditary information is passed on from a cell to its daughter cells at cell division, and from one generation of an organism to the next through the organism’s reproductive cells. The instructions are stored within every living cell as its genes, the information-containing elements that determine the characteristics of a species as a whole and of the individuals within it. | Cell_Biology_Alberts. Vale RD & Milligan RA (2000) The way things move: looking under the hood of molecular motor proteins. Science 288, 88–95. Wilson MZ & Gitai Z (2013) Beyond the cytoskeleton: mesoscale assemblies and their function in spatial organization. Curr. Opin. Microbiol. 16, 177–183. DNA, Chromosomes, and genomes Life depends on the ability of cells to store, retrieve, and translate the genetic instructions required to make and maintain a living organism. This hereditary information is passed on from a cell to its daughter cells at cell division, and from one generation of an organism to the next through the organism’s reproductive cells. The instructions are stored within every living cell as its genes, the information-containing elements that determine the characteristics of a species as a whole and of the individuals within it. |
Cell_Biology_Alberts_764 | Cell_Biology_Alberts | As soon as genetics emerged as a science at the beginning of the twentieth century, scientists became intrigued by the chemical structure of genes. The information in genes is copied and transmitted from cell to daughter cell millions of times during the life of a multicellular organism, and it survives the process essentially unchanged. What form of molecule could be capable of such accurate and almost unlimited replication and also be able to exert precise control, directing multicellular development as well as the daily life of every cell? What kind of instructions does the genetic information contain? And how can the enormous amount of information required for the development and maintenance of an organism fit within the tiny space of a cell? | Cell_Biology_Alberts. As soon as genetics emerged as a science at the beginning of the twentieth century, scientists became intrigued by the chemical structure of genes. The information in genes is copied and transmitted from cell to daughter cell millions of times during the life of a multicellular organism, and it survives the process essentially unchanged. What form of molecule could be capable of such accurate and almost unlimited replication and also be able to exert precise control, directing multicellular development as well as the daily life of every cell? What kind of instructions does the genetic information contain? And how can the enormous amount of information required for the development and maintenance of an organism fit within the tiny space of a cell? |
Cell_Biology_Alberts_765 | Cell_Biology_Alberts | The answers to several of these questions began to emerge in the 1940s. At this time researchers discovered, from studies in simple fungi, that genetic information consists largely of instructions for making proteins. Proteins are phenomenally versatile macromolecules that perform most cell functions. As we saw in Chapter 3, they serve as building blocks for cell structures and form the enzymes that catalyze most of the cell’s chemical reactions. They also regulate gene expression (Chapter 7), and they enable cells to communicate with each other (Chapter 15) and to move (Chapter 16). The properties and functions of cells and organisms are determined to a great extent by the proteins that they are able to make. | Cell_Biology_Alberts. The answers to several of these questions began to emerge in the 1940s. At this time researchers discovered, from studies in simple fungi, that genetic information consists largely of instructions for making proteins. Proteins are phenomenally versatile macromolecules that perform most cell functions. As we saw in Chapter 3, they serve as building blocks for cell structures and form the enzymes that catalyze most of the cell’s chemical reactions. They also regulate gene expression (Chapter 7), and they enable cells to communicate with each other (Chapter 15) and to move (Chapter 16). The properties and functions of cells and organisms are determined to a great extent by the proteins that they are able to make. |
Cell_Biology_Alberts_766 | Cell_Biology_Alberts | Painstaking observations of cells and embryos in the late nineteenth century had led to the recognition that the hereditary information is carried on chromosomes—threadlike structures in the nucleus of a eukaryotic cell that become visible by light microscopy as the cell begins to divide (Figure 4–1). Later, when biochemical analysis became possible, chromosomes were found to consist of deoxyribonucleic acid (DNA) and protein, with both being present in roughly the same amounts. For many decades, the DNA was thought to be merely a structural | Cell_Biology_Alberts. Painstaking observations of cells and embryos in the late nineteenth century had led to the recognition that the hereditary information is carried on chromosomes—threadlike structures in the nucleus of a eukaryotic cell that become visible by light microscopy as the cell begins to divide (Figure 4–1). Later, when biochemical analysis became possible, chromosomes were found to consist of deoxyribonucleic acid (DNA) and protein, with both being present in roughly the same amounts. For many decades, the DNA was thought to be merely a structural |
Cell_Biology_Alberts_767 | Cell_Biology_Alberts | Figure 4–1 Chromosomes in cells. (A) Two adjacent plant cells photographed through a light microscope. The DNA has been stained with a fluorescent dye (DAPI) that binds to it. The DNA is present in chromosomes, which become visible as distinct structures in the light microscope only when they become compact, sausage-shaped structures in preparation for cell division, as shown on the left. The cell on the right, which is not dividing, contains identical chromosomes, but they cannot be clearly distinguished at this phase in the cell’s life cycle, because they are in a more extended conformation. (b) Schematic diagram of the outlines of the two cells along with their chromosomes. (A, courtesy of Peter Shaw.) element. However, the other crucial advance made in the 1940s was the identification of DNA as the likely carrier of genetic information. This breakthrough in our understanding of cells came from studies of inheritance in bacteria (Figure 4–2). But still, as the 1950s began, both how | Cell_Biology_Alberts. Figure 4–1 Chromosomes in cells. (A) Two adjacent plant cells photographed through a light microscope. The DNA has been stained with a fluorescent dye (DAPI) that binds to it. The DNA is present in chromosomes, which become visible as distinct structures in the light microscope only when they become compact, sausage-shaped structures in preparation for cell division, as shown on the left. The cell on the right, which is not dividing, contains identical chromosomes, but they cannot be clearly distinguished at this phase in the cell’s life cycle, because they are in a more extended conformation. (b) Schematic diagram of the outlines of the two cells along with their chromosomes. (A, courtesy of Peter Shaw.) element. However, the other crucial advance made in the 1940s was the identification of DNA as the likely carrier of genetic information. This breakthrough in our understanding of cells came from studies of inheritance in bacteria (Figure 4–2). But still, as the 1950s began, both how |
Cell_Biology_Alberts_768 | Cell_Biology_Alberts | of DNA as the likely carrier of genetic information. This breakthrough in our understanding of cells came from studies of inheritance in bacteria (Figure 4–2). But still, as the 1950s began, both how proteins could be specified by instructions in the DNA and how this information might be copied for transmission from cell to cell seemed completely mysterious. The puzzle was suddenly solved in 1953, when James Watson and Francis Crick derived the mechanism from their model of DNA structure. As outlined in Chapter 1, the determination of the double-helical structure of DNA immediately solved the problem of how the information in this molecule might be copied, or replicated. It also provided the first clues as to how a molecule of DNA might use the sequence of its subunits to encode the instructions for making proteins. Today, the fact that DNA is the genetic material is so fundamental to biological thought that it is difficult to appreciate the enormous intellectual gap that was filled | Cell_Biology_Alberts. of DNA as the likely carrier of genetic information. This breakthrough in our understanding of cells came from studies of inheritance in bacteria (Figure 4–2). But still, as the 1950s began, both how proteins could be specified by instructions in the DNA and how this information might be copied for transmission from cell to cell seemed completely mysterious. The puzzle was suddenly solved in 1953, when James Watson and Francis Crick derived the mechanism from their model of DNA structure. As outlined in Chapter 1, the determination of the double-helical structure of DNA immediately solved the problem of how the information in this molecule might be copied, or replicated. It also provided the first clues as to how a molecule of DNA might use the sequence of its subunits to encode the instructions for making proteins. Today, the fact that DNA is the genetic material is so fundamental to biological thought that it is difficult to appreciate the enormous intellectual gap that was filled |
Cell_Biology_Alberts_769 | Cell_Biology_Alberts | for making proteins. Today, the fact that DNA is the genetic material is so fundamental to biological thought that it is difficult to appreciate the enormous intellectual gap that was filled by this breakthrough discovery. | Cell_Biology_Alberts. for making proteins. Today, the fact that DNA is the genetic material is so fundamental to biological thought that it is difficult to appreciate the enormous intellectual gap that was filled by this breakthrough discovery. |
Cell_Biology_Alberts_770 | Cell_Biology_Alberts | We begin this chapter by describing the structure of DNA. We see how, despite its chemical simplicity, the structure and chemical properties of DNA make it ideally suited as the raw material of genes. We then consider how the many proteins in chromosomes arrange and package this DNA. The packing has to be done in an orderly fashion so that the chromosomes can be replicated and apportioned correctly between the two daughter cells at each cell division. And it must also allow access to chromosomal DNA, both for the enzymes that repair DNA damage and for the specialized proteins that direct the expression of its many genes. In the past two decades, there has been a revolution in our ability to determine the exact order of subunits in DNA molecules. As a result, we now know the RNA protein DNA lipid carbohydratelive R strain cells grown in presence of either heat-killed S strain cells or cell-free molecules tested for transformation of R strain cellsextract of S strain cells | Cell_Biology_Alberts. We begin this chapter by describing the structure of DNA. We see how, despite its chemical simplicity, the structure and chemical properties of DNA make it ideally suited as the raw material of genes. We then consider how the many proteins in chromosomes arrange and package this DNA. The packing has to be done in an orderly fashion so that the chromosomes can be replicated and apportioned correctly between the two daughter cells at each cell division. And it must also allow access to chromosomal DNA, both for the enzymes that repair DNA damage and for the specialized proteins that direct the expression of its many genes. In the past two decades, there has been a revolution in our ability to determine the exact order of subunits in DNA molecules. As a result, we now know the RNA protein DNA lipid carbohydratelive R strain cells grown in presence of either heat-killed S strain cells or cell-free molecules tested for transformation of R strain cellsextract of S strain cells |
Cell_Biology_Alberts_771 | Cell_Biology_Alberts | CONCLUSION: Molecules that can CONCLUSION: The molecule that carry heritable information are carries the heritable information present in S strain cells. is DNA. Figure 4–2 The first experimental demonstration that DNA is the genetic material. These experiments, carried out in the 1920s (A) and 1940s (b), showed that adding purified DNA to a bacterium changed the bacterium’s properties and that this change was faithfully passed on to subsequent generations. Two closely related strains of the bacterium Streptococcus pneumoniae differ from each other in both their appearance under the microscope and their pathogenicity. One strain appears smooth (S) and causes death when injected into mice, and the other appears rough (R) and is nonlethal. An initial experiment shows that some substance present in the S strain can change (or transform) the R strain into the S strain and that this change is inherited by subsequent generations of bacteria. | Cell_Biology_Alberts. CONCLUSION: Molecules that can CONCLUSION: The molecule that carry heritable information are carries the heritable information present in S strain cells. is DNA. Figure 4–2 The first experimental demonstration that DNA is the genetic material. These experiments, carried out in the 1920s (A) and 1940s (b), showed that adding purified DNA to a bacterium changed the bacterium’s properties and that this change was faithfully passed on to subsequent generations. Two closely related strains of the bacterium Streptococcus pneumoniae differ from each other in both their appearance under the microscope and their pathogenicity. One strain appears smooth (S) and causes death when injected into mice, and the other appears rough (R) and is nonlethal. An initial experiment shows that some substance present in the S strain can change (or transform) the R strain into the S strain and that this change is inherited by subsequent generations of bacteria. |
Cell_Biology_Alberts_772 | Cell_Biology_Alberts | This experiment, in which the R strain has been incubated with various classes of biological molecules purified from the S strain, identifies the active substance as DNA. sequence of the 3.2 billion nucleotide pairs that provide the information for producing a human adult from a fertilized egg, as well as having the DNA sequences for thousands of other organisms. Detailed analyses of these sequences are providing exciting insights into the process of evolution, and it is with this subject that the chapter ends. | Cell_Biology_Alberts. This experiment, in which the R strain has been incubated with various classes of biological molecules purified from the S strain, identifies the active substance as DNA. sequence of the 3.2 billion nucleotide pairs that provide the information for producing a human adult from a fertilized egg, as well as having the DNA sequences for thousands of other organisms. Detailed analyses of these sequences are providing exciting insights into the process of evolution, and it is with this subject that the chapter ends. |
Cell_Biology_Alberts_773 | Cell_Biology_Alberts | This is the first of four chapters that deal with basic genetic mechanisms—the ways in which the cell maintains, replicates, and expresses the genetic information carried in its DNA. In the next chapter (Chapter 5), we shall discuss the mechanisms by which the cell accurately replicates and repairs DNA; we also describe how DNA sequences can be rearranged through the process of genetic recombination. Gene expression—the process through which the information encoded in DNA is interpreted by the cell to guide the synthesis of proteins—is the main topic of Chapter 6. In Chapter 7, we describe how this gene expression is controlled by the cell to ensure that each of the many thousands of proteins and RNA molecules encrypted in its DNA is manufactured only at the proper time and place in the life of a cell. | Cell_Biology_Alberts. This is the first of four chapters that deal with basic genetic mechanisms—the ways in which the cell maintains, replicates, and expresses the genetic information carried in its DNA. In the next chapter (Chapter 5), we shall discuss the mechanisms by which the cell accurately replicates and repairs DNA; we also describe how DNA sequences can be rearranged through the process of genetic recombination. Gene expression—the process through which the information encoded in DNA is interpreted by the cell to guide the synthesis of proteins—is the main topic of Chapter 6. In Chapter 7, we describe how this gene expression is controlled by the cell to ensure that each of the many thousands of proteins and RNA molecules encrypted in its DNA is manufactured only at the proper time and place in the life of a cell. |
Cell_Biology_Alberts_774 | Cell_Biology_Alberts | Biologists in the 1940s had difficulty in conceiving how DNA could be the genetic material. The molecule seemed too simple: a long polymer composed of only four types of nucleotide subunits, which resemble one another chemically. Early in the 1950s, DNA was examined by x-ray diffraction analysis, a technique for determining the three-dimensional atomic structure of a molecule (discussed in Chapter 8). The early x-ray diffraction results indicated that DNA was composed of two strands of the polymer wound into a helix. The observation that DNA was double-stranded provided one of the major clues that led to the Watson–Crick model for DNA structure that, as soon as it was proposed in 1953, made DNA’s potential for replication and information storage apparent. A DNA molecule Consists of Two Complementary Chains of Nucleotides | Cell_Biology_Alberts. Biologists in the 1940s had difficulty in conceiving how DNA could be the genetic material. The molecule seemed too simple: a long polymer composed of only four types of nucleotide subunits, which resemble one another chemically. Early in the 1950s, DNA was examined by x-ray diffraction analysis, a technique for determining the three-dimensional atomic structure of a molecule (discussed in Chapter 8). The early x-ray diffraction results indicated that DNA was composed of two strands of the polymer wound into a helix. The observation that DNA was double-stranded provided one of the major clues that led to the Watson–Crick model for DNA structure that, as soon as it was proposed in 1953, made DNA’s potential for replication and information storage apparent. A DNA molecule Consists of Two Complementary Chains of Nucleotides |
Cell_Biology_Alberts_775 | Cell_Biology_Alberts | A deoxyribonucleic acid (DNA) molecule consists of two long polynucleotide chains composed of four types of nucleotide subunits. Each of these chains is known as a DNA chain, or a DNA strand. The chains run antiparallel to each other, and hydrogen bonds between the base portions of the nucleotides hold the two chains together (Figure 4–3). As we saw in Chapter 2 (Panel 2–6, pp. 100–101), nucleotides are composed of a five-carbon sugar to which are attached one or more phosphate groups and a nitrogen-containing base. In the case of the nucleotides in DNA, the sugar is deoxyribose attached to a single phosphate group (hence the name deoxyribonucleic acid), and the base may be either adenine (A), cytosine (C), guanine (G), or thymine (T). The nucleotides are covalently linked together in a chain through the sugars and phosphates, which thus form a “backbone” of alternating sugar–phosphate–sugar–phosphate. Because only the base differs in each of the four types of nucleotide subunit, each | Cell_Biology_Alberts. A deoxyribonucleic acid (DNA) molecule consists of two long polynucleotide chains composed of four types of nucleotide subunits. Each of these chains is known as a DNA chain, or a DNA strand. The chains run antiparallel to each other, and hydrogen bonds between the base portions of the nucleotides hold the two chains together (Figure 4–3). As we saw in Chapter 2 (Panel 2–6, pp. 100–101), nucleotides are composed of a five-carbon sugar to which are attached one or more phosphate groups and a nitrogen-containing base. In the case of the nucleotides in DNA, the sugar is deoxyribose attached to a single phosphate group (hence the name deoxyribonucleic acid), and the base may be either adenine (A), cytosine (C), guanine (G), or thymine (T). The nucleotides are covalently linked together in a chain through the sugars and phosphates, which thus form a “backbone” of alternating sugar–phosphate–sugar–phosphate. Because only the base differs in each of the four types of nucleotide subunit, each |
Cell_Biology_Alberts_776 | Cell_Biology_Alberts | through the sugars and phosphates, which thus form a “backbone” of alternating sugar–phosphate–sugar–phosphate. Because only the base differs in each of the four types of nucleotide subunit, each polynucleotide chain in DNA is analogous to a sugar-phosphate necklace (the backbone), from which hang the four types of beads (the bases A, C, G, and T). These same symbols (A, C, G, and T) are commonly used to denote either the four bases or the four entire nucleotides—that is, the bases with their attached sugar and phosphate groups. | Cell_Biology_Alberts. through the sugars and phosphates, which thus form a “backbone” of alternating sugar–phosphate–sugar–phosphate. Because only the base differs in each of the four types of nucleotide subunit, each polynucleotide chain in DNA is analogous to a sugar-phosphate necklace (the backbone), from which hang the four types of beads (the bases A, C, G, and T). These same symbols (A, C, G, and T) are commonly used to denote either the four bases or the four entire nucleotides—that is, the bases with their attached sugar and phosphate groups. |
Cell_Biology_Alberts_777 | Cell_Biology_Alberts | The way in which the nucleotides are linked together gives a DNA strand a chemical polarity. If we think of each sugar as a block with a protruding knob (the 5ʹ phosphate) on one side and a hole (the 3ʹ hydroxyl) on the other (see Figure 4–3), each completed chain, formed by interlocking knobs with holes, will have all of its subunits lined up in the same orientation. Moreover, the two ends of the chain will be easily distinguishable, as one has a hole (the 3ʹ hydroxyl) and the other a knob (the 5ʹ phosphate) at its terminus. This polarity in a DNA chain is indicated by referring to one end as the 3ʹ end and the other as the 5ʹ end, names derived from the orientation of the deoxyribose sugar. With respect to DNA’s building blocks of DNA DNA strand information-carrying capacity, the chain of nucleotides in a DNA strand, being both directional and linear, can be read in much the same way as the letters on this page. | Cell_Biology_Alberts. The way in which the nucleotides are linked together gives a DNA strand a chemical polarity. If we think of each sugar as a block with a protruding knob (the 5ʹ phosphate) on one side and a hole (the 3ʹ hydroxyl) on the other (see Figure 4–3), each completed chain, formed by interlocking knobs with holes, will have all of its subunits lined up in the same orientation. Moreover, the two ends of the chain will be easily distinguishable, as one has a hole (the 3ʹ hydroxyl) and the other a knob (the 5ʹ phosphate) at its terminus. This polarity in a DNA chain is indicated by referring to one end as the 3ʹ end and the other as the 5ʹ end, names derived from the orientation of the deoxyribose sugar. With respect to DNA’s building blocks of DNA DNA strand information-carrying capacity, the chain of nucleotides in a DNA strand, being both directional and linear, can be read in much the same way as the letters on this page. |
Cell_Biology_Alberts_778 | Cell_Biology_Alberts | The three-dimensional structure of DNA—the DNA double helix—arises from the chemical and structural features of its two polynucleotide chains. Because these two chains are held together by hydrogen-bonding between the bases on the different strands, all the bases are on the inside of the double helix, and the sugar-phosphate backbones are on the outside (see Figure 4–3). In each case, a bulkier two-ring base (a purine; see Panel 2–6, pp. 100–101) is paired with a single-ring base (a pyrimidine): A always pairs with T, and G with C (Figure 4–4). This complementary base-pairing enables the base pairs to be packed in the energetically most favorable arrangement in the interior of the double helix. In this arrangement, each base pair is of similar width, thus holding the sugar-phosphate backbones a constant distance apart along the DNA molecule. To maximize the efficiency of base-pair packing, the two sugar-phosphate backbones wind around each other to form a right-handed double helix, | Cell_Biology_Alberts. The three-dimensional structure of DNA—the DNA double helix—arises from the chemical and structural features of its two polynucleotide chains. Because these two chains are held together by hydrogen-bonding between the bases on the different strands, all the bases are on the inside of the double helix, and the sugar-phosphate backbones are on the outside (see Figure 4–3). In each case, a bulkier two-ring base (a purine; see Panel 2–6, pp. 100–101) is paired with a single-ring base (a pyrimidine): A always pairs with T, and G with C (Figure 4–4). This complementary base-pairing enables the base pairs to be packed in the energetically most favorable arrangement in the interior of the double helix. In this arrangement, each base pair is of similar width, thus holding the sugar-phosphate backbones a constant distance apart along the DNA molecule. To maximize the efficiency of base-pair packing, the two sugar-phosphate backbones wind around each other to form a right-handed double helix, |
Cell_Biology_Alberts_779 | Cell_Biology_Alberts | a constant distance apart along the DNA molecule. To maximize the efficiency of base-pair packing, the two sugar-phosphate backbones wind around each other to form a right-handed double helix, with one complete turn every ten base pairs (Figure 4–5). | Cell_Biology_Alberts. a constant distance apart along the DNA molecule. To maximize the efficiency of base-pair packing, the two sugar-phosphate backbones wind around each other to form a right-handed double helix, with one complete turn every ten base pairs (Figure 4–5). |
Cell_Biology_Alberts_780 | Cell_Biology_Alberts | The members of each base pair can fit together within the double helix only if the two strands of the helix are antiparallel—that is, only if the polarity of one strand is oriented opposite to that of the other strand (see Figures 4–3 and 4–4). A consequence of DNA’s structure and base-pairing requirements is that each strand of a DNA molecule contains a sequence of nucleotides that is exactly complementary to the nucleotide sequence of its partner strand. Figure 4–3 DNA and its building blocks. | Cell_Biology_Alberts. The members of each base pair can fit together within the double helix only if the two strands of the helix are antiparallel—that is, only if the polarity of one strand is oriented opposite to that of the other strand (see Figures 4–3 and 4–4). A consequence of DNA’s structure and base-pairing requirements is that each strand of a DNA molecule contains a sequence of nucleotides that is exactly complementary to the nucleotide sequence of its partner strand. Figure 4–3 DNA and its building blocks. |
Cell_Biology_Alberts_781 | Cell_Biology_Alberts | Figure 4–3 DNA and its building blocks. DNA is made of four types of nucleotides, which are linked covalently into a polynucleotide chain (a DNA strand) with a sugar-phosphate backbone from which the bases (A, C, g, and T) extend. A DNA molecule is composed of two antiparallel DNA strands held together by hydrogen bonds between the paired bases. The arrowheads at the ends of the DNA strands indicate the polarities of the two strands. In the diagram at the bottom left of the figure, the DNA molecule is shown straightened out; in reality, it is twisted into a double helix, as shown on the right. For details, see Figure 4–5 and Movie 4.1. The Structure of DNA Provides a mechanism for Heredity | Cell_Biology_Alberts. Figure 4–3 DNA and its building blocks. DNA is made of four types of nucleotides, which are linked covalently into a polynucleotide chain (a DNA strand) with a sugar-phosphate backbone from which the bases (A, C, g, and T) extend. A DNA molecule is composed of two antiparallel DNA strands held together by hydrogen bonds between the paired bases. The arrowheads at the ends of the DNA strands indicate the polarities of the two strands. In the diagram at the bottom left of the figure, the DNA molecule is shown straightened out; in reality, it is twisted into a double helix, as shown on the right. For details, see Figure 4–5 and Movie 4.1. The Structure of DNA Provides a mechanism for Heredity |
Cell_Biology_Alberts_782 | Cell_Biology_Alberts | The Structure of DNA Provides a mechanism for Heredity The discovery of the structure of DNA immediately suggested answers to the two most fundamental questions about heredity. First, how could the information to specify an organism be carried in a chemical form? And second, how could this information be duplicated and copied from generation to generation? The answer to the first question came from the realization that DNA is a linear polymer of four different kinds of monomer, strung out in a defined sequence like the letters of a document written in an alphabetic script. The answer to the second question came from the double-stranded nature of the structure: because each strand of DNA contains a sequence of nucleotides that is exactly complementary to the nucleotide sequence of its partner strand, each strand can act as a template, or mold, for the synthesis of a new complementary strand. In other words, if we designate the two DNA strands as S and Sʹ, strand 0.34 nm | Cell_Biology_Alberts. The Structure of DNA Provides a mechanism for Heredity The discovery of the structure of DNA immediately suggested answers to the two most fundamental questions about heredity. First, how could the information to specify an organism be carried in a chemical form? And second, how could this information be duplicated and copied from generation to generation? The answer to the first question came from the realization that DNA is a linear polymer of four different kinds of monomer, strung out in a defined sequence like the letters of a document written in an alphabetic script. The answer to the second question came from the double-stranded nature of the structure: because each strand of DNA contains a sequence of nucleotides that is exactly complementary to the nucleotide sequence of its partner strand, each strand can act as a template, or mold, for the synthesis of a new complementary strand. In other words, if we designate the two DNA strands as S and Sʹ, strand 0.34 nm |
Cell_Biology_Alberts_783 | Cell_Biology_Alberts | Figure 4–4 Complementary base pairs in the DNA double helix. The shapes and chemical structures of the bases allow hydrogen bonds to form efficiently only between A and T and between g and C, because atoms that are able to form hydrogen bonds (see Panel 2–3, pp. 94–95) can then be brought close together without distorting the double helix. As indicated, two hydrogen bonds form between A and T, while three form between g and C. The bases can pair in this way only if the two polynucleotide chains that contain them are antiparallel to each other. Figure 4–5 The DNA double helix. A space-filling model of 1.5 turns of the DNA double helix. Each turn of DNA is made up of 10.4 nucleotide pairs, and the center-to-center distance between adjacent nucleotide pairs is 0.34 nm. The coiling of the two strands around each other creates two grooves in the double helix: the wider groove is called the major groove, and the smaller the minor groove, as indicated. | Cell_Biology_Alberts. Figure 4–4 Complementary base pairs in the DNA double helix. The shapes and chemical structures of the bases allow hydrogen bonds to form efficiently only between A and T and between g and C, because atoms that are able to form hydrogen bonds (see Panel 2–3, pp. 94–95) can then be brought close together without distorting the double helix. As indicated, two hydrogen bonds form between A and T, while three form between g and C. The bases can pair in this way only if the two polynucleotide chains that contain them are antiparallel to each other. Figure 4–5 The DNA double helix. A space-filling model of 1.5 turns of the DNA double helix. Each turn of DNA is made up of 10.4 nucleotide pairs, and the center-to-center distance between adjacent nucleotide pairs is 0.34 nm. The coiling of the two strands around each other creates two grooves in the double helix: the wider groove is called the major groove, and the smaller the minor groove, as indicated. |
Cell_Biology_Alberts_784 | Cell_Biology_Alberts | A short section of the double helix viewed from its side, showing four base pairs. The nucleotides are linked together covalently by phosphodiester bonds that join the 3ʹ-hydroxyl (–OH) group of one sugar to the 5ʹ-hydroxyl group of the next sugar. Thus, each polynucleotide strand has a chemical polarity; that is, its two ends are chemically different. The 5ʹ end of the DNA polymer is by convention often illustrated carrying a phosphate group, while the 3ʹ end is shown with a hydroxyl. S can serve as a template for making a new strand Sʹ, while strand Sʹ can serve as a template for making a new strand S (Figure 4–6). Thus, the genetic information in DNA can be accurately copied by the beautifully simple process in which strand S separates from strand Sʹ, and each separated strand then serves as a template for the production of a new complementary partner strand that is identical to its former partner. | Cell_Biology_Alberts. A short section of the double helix viewed from its side, showing four base pairs. The nucleotides are linked together covalently by phosphodiester bonds that join the 3ʹ-hydroxyl (–OH) group of one sugar to the 5ʹ-hydroxyl group of the next sugar. Thus, each polynucleotide strand has a chemical polarity; that is, its two ends are chemically different. The 5ʹ end of the DNA polymer is by convention often illustrated carrying a phosphate group, while the 3ʹ end is shown with a hydroxyl. S can serve as a template for making a new strand Sʹ, while strand Sʹ can serve as a template for making a new strand S (Figure 4–6). Thus, the genetic information in DNA can be accurately copied by the beautifully simple process in which strand S separates from strand Sʹ, and each separated strand then serves as a template for the production of a new complementary partner strand that is identical to its former partner. |
Cell_Biology_Alberts_785 | Cell_Biology_Alberts | The ability of each strand of a DNA molecule to act as a template for producing a complementary strand enables a cell to copy, or replicate, its genome before passing it on to its descendants. We shall describe the elegant machinery that the cell uses to perform this task in Chapter 5. Organisms differ from one another because their respective DNA molecules have different nucleotide sequences and, consequently, carry different biological messages. But how is the nucleotide alphabet used to make messages, and what do they spell out? | Cell_Biology_Alberts. The ability of each strand of a DNA molecule to act as a template for producing a complementary strand enables a cell to copy, or replicate, its genome before passing it on to its descendants. We shall describe the elegant machinery that the cell uses to perform this task in Chapter 5. Organisms differ from one another because their respective DNA molecules have different nucleotide sequences and, consequently, carry different biological messages. But how is the nucleotide alphabet used to make messages, and what do they spell out? |
Cell_Biology_Alberts_786 | Cell_Biology_Alberts | As discussed above, it was known well before the structure of DNA was determined that genes contain the instructions for producing proteins. If genes are made of DNA, the DNA must therefore somehow encode proteins (Figure 4–7). As discussed in Chapter 3, the properties of a protein, which are responsible for its biological function, are determined by its three-dimensional structure. This structure is determined in turn by the linear sequence of the amino acids of which it is composed. The linear sequence of nucleotides in a gene must therefore somehow spell out the linear sequence of amino acids in a protein. The exact correspondence between the four-letter nucleotide alphabet of DNA and the twenty-letter amino acid alphabet of proteins—the genetic code—is not at all obvious from the DNA structure, and it took over a decade after the discovery of the double helix before it was worked out. In Chapter 6, we will describe this code in detail in the course of elaborating the process of | Cell_Biology_Alberts. As discussed above, it was known well before the structure of DNA was determined that genes contain the instructions for producing proteins. If genes are made of DNA, the DNA must therefore somehow encode proteins (Figure 4–7). As discussed in Chapter 3, the properties of a protein, which are responsible for its biological function, are determined by its three-dimensional structure. This structure is determined in turn by the linear sequence of the amino acids of which it is composed. The linear sequence of nucleotides in a gene must therefore somehow spell out the linear sequence of amino acids in a protein. The exact correspondence between the four-letter nucleotide alphabet of DNA and the twenty-letter amino acid alphabet of proteins—the genetic code—is not at all obvious from the DNA structure, and it took over a decade after the discovery of the double helix before it was worked out. In Chapter 6, we will describe this code in detail in the course of elaborating the process of |
Cell_Biology_Alberts_787 | Cell_Biology_Alberts | structure, and it took over a decade after the discovery of the double helix before it was worked out. In Chapter 6, we will describe this code in detail in the course of elaborating the process of gene expression, through which a cell converts the nucleotide sequence of a gene first into the nucleotide sequence of an RNA molecule, and then into the amino acid sequence of a protein. | Cell_Biology_Alberts. structure, and it took over a decade after the discovery of the double helix before it was worked out. In Chapter 6, we will describe this code in detail in the course of elaborating the process of gene expression, through which a cell converts the nucleotide sequence of a gene first into the nucleotide sequence of an RNA molecule, and then into the amino acid sequence of a protein. |
Cell_Biology_Alberts_788 | Cell_Biology_Alberts | The complete store of information in an organism’s DNA is called its genome, and it specifies all the RNA molecules and proteins that the organism will ever synthesize. (The term genome is also used to describe the DNA that carries this information.) The amount of information contained in genomes is staggering. The nucleotide sequence of a very small human gene, written out in the four-letter nucleotide alphabet, occupies a quarter of a page of text (Figure 4–8), while the complete sequence of nucleotides in the human genome would fill more than a thousand books the size of this one. In addition to other critical information, it includes roughly 21,000 protein-coding genes, which (through alternative splicing; see p. 415) give rise to a much greater number of distinct proteins. In Eukaryotes, DNA Is Enclosed in a Cell Nucleus | Cell_Biology_Alberts. The complete store of information in an organism’s DNA is called its genome, and it specifies all the RNA molecules and proteins that the organism will ever synthesize. (The term genome is also used to describe the DNA that carries this information.) The amount of information contained in genomes is staggering. The nucleotide sequence of a very small human gene, written out in the four-letter nucleotide alphabet, occupies a quarter of a page of text (Figure 4–8), while the complete sequence of nucleotides in the human genome would fill more than a thousand books the size of this one. In addition to other critical information, it includes roughly 21,000 protein-coding genes, which (through alternative splicing; see p. 415) give rise to a much greater number of distinct proteins. In Eukaryotes, DNA Is Enclosed in a Cell Nucleus |
Cell_Biology_Alberts_789 | Cell_Biology_Alberts | In Eukaryotes, DNA Is Enclosed in a Cell Nucleus As described in Chapter 1, nearly all the DNA in a eukaryotic cell is sequestered in a nucleus, which in many cells occupies about 10% of the total cell volume. This compartment is delimited by a nuclear envelope formed by two concentric lipid Figure 4–6 DNA as a template for its own duplication. because the nucleotide A successfully pairs only with T, and g pairs with C, each strand of DNA can act as a template to specify the sequence of nucleotides in its complementary strand. In this way, double-helical DNA can be copied precisely, with each parental DNA helix producing two identical daughter DNA helices. Figure 4–7 The relationship between genetic information carried in DNA and proteins. (Discussed in Chapter 1.) | Cell_Biology_Alberts. In Eukaryotes, DNA Is Enclosed in a Cell Nucleus As described in Chapter 1, nearly all the DNA in a eukaryotic cell is sequestered in a nucleus, which in many cells occupies about 10% of the total cell volume. This compartment is delimited by a nuclear envelope formed by two concentric lipid Figure 4–6 DNA as a template for its own duplication. because the nucleotide A successfully pairs only with T, and g pairs with C, each strand of DNA can act as a template to specify the sequence of nucleotides in its complementary strand. In this way, double-helical DNA can be copied precisely, with each parental DNA helix producing two identical daughter DNA helices. Figure 4–7 The relationship between genetic information carried in DNA and proteins. (Discussed in Chapter 1.) |
Cell_Biology_Alberts_790 | Cell_Biology_Alberts | Figure 4–7 The nucleotide sequence of the human β-globin gene. by convention, a nucleotide sequence is written from its 5ʹ end to its 3ʹ end, and it should be read from left to right in successive lines down the page as though it were normal English text. This gene carries the information for the amino acid sequence of one of the two types of subunits of the hemoglobin molecule; a different gene, the α-globin gene, carries the information for the other. (Hemoglobin, the protein that carries oxygen in the blood, has four subunits, two of each type.) Only one of the two strands of the DNA double helix containing the β-globin gene is shown; the other strand has the exact complementary sequence. The DNA sequences highlighted in yellow show the three regions of the gene that specify the amino acid sequence for the β-globin protein. we shall see in Chapter 6 how the cell splices these three sequences together at the level of messenger RNA in order to synthesize a full-length β-globin | Cell_Biology_Alberts. Figure 4–7 The nucleotide sequence of the human β-globin gene. by convention, a nucleotide sequence is written from its 5ʹ end to its 3ʹ end, and it should be read from left to right in successive lines down the page as though it were normal English text. This gene carries the information for the amino acid sequence of one of the two types of subunits of the hemoglobin molecule; a different gene, the α-globin gene, carries the information for the other. (Hemoglobin, the protein that carries oxygen in the blood, has four subunits, two of each type.) Only one of the two strands of the DNA double helix containing the β-globin gene is shown; the other strand has the exact complementary sequence. The DNA sequences highlighted in yellow show the three regions of the gene that specify the amino acid sequence for the β-globin protein. we shall see in Chapter 6 how the cell splices these three sequences together at the level of messenger RNA in order to synthesize a full-length β-globin |
Cell_Biology_Alberts_791 | Cell_Biology_Alberts | amino acid sequence for the β-globin protein. we shall see in Chapter 6 how the cell splices these three sequences together at the level of messenger RNA in order to synthesize a full-length β-globin protein. | Cell_Biology_Alberts. amino acid sequence for the β-globin protein. we shall see in Chapter 6 how the cell splices these three sequences together at the level of messenger RNA in order to synthesize a full-length β-globin protein. |
Cell_Biology_Alberts_792 | Cell_Biology_Alberts | bilayer membranes (Figure 4–9). These membranes are punctured at intervals by large nuclear pores, through which molecules move between the nucleus and the cytosol. The nuclear envelope is directly connected to the extensive system of intracellular membranes called the endoplasmic reticulum, which extend out from it into the cytoplasm. And it is mechanically supported by a network of intermediate filaments called the nuclear lamina—a thin feltlike mesh just beneath the inner nuclear membrane (see Figure 4–9B). The nuclear envelope allows the many proteins that act on DNA to be concentrated where they are needed in the cell, and, as we see in subsequent chapters, it also keeps nuclear and cytosolic enzymes separate, a feature that is crucial for the proper functioning of eukaryotic cells. | Cell_Biology_Alberts. bilayer membranes (Figure 4–9). These membranes are punctured at intervals by large nuclear pores, through which molecules move between the nucleus and the cytosol. The nuclear envelope is directly connected to the extensive system of intracellular membranes called the endoplasmic reticulum, which extend out from it into the cytoplasm. And it is mechanically supported by a network of intermediate filaments called the nuclear lamina—a thin feltlike mesh just beneath the inner nuclear membrane (see Figure 4–9B). The nuclear envelope allows the many proteins that act on DNA to be concentrated where they are needed in the cell, and, as we see in subsequent chapters, it also keeps nuclear and cytosolic enzymes separate, a feature that is crucial for the proper functioning of eukaryotic cells. |
Cell_Biology_Alberts_793 | Cell_Biology_Alberts | Genetic information is carried in the linear sequence of nucleotides in DNA. Each molecule of DNA is a double helix formed from two complementary antiparallel strands of nucleotides held together by hydrogen bonds between G-C and A-T base pairs. Duplication of the genetic information occurs by the use of one DNA strand as a template for the formation of a complementary strand. The genetic information stored in an organism’s DNA contains the instructions for all the RNA molecules and proteins that the organism will ever synthesize and is said to comprise its genome. In eukaryotes, DNA is contained in the cell nucleus, a large membrane-bound compartment. | Cell_Biology_Alberts. Genetic information is carried in the linear sequence of nucleotides in DNA. Each molecule of DNA is a double helix formed from two complementary antiparallel strands of nucleotides held together by hydrogen bonds between G-C and A-T base pairs. Duplication of the genetic information occurs by the use of one DNA strand as a template for the formation of a complementary strand. The genetic information stored in an organism’s DNA contains the instructions for all the RNA molecules and proteins that the organism will ever synthesize and is said to comprise its genome. In eukaryotes, DNA is contained in the cell nucleus, a large membrane-bound compartment. |
Cell_Biology_Alberts_794 | Cell_Biology_Alberts | The most important function of DNA is to carry genes, the information that specifies all the RNA molecules and proteins that make up an organism—including information about when, in what types of cells, and in what quantity each RNA molecule and protein is to be made. The nuclear DNA of eukaryotes is divided up into chromosomes, and in this section we see how genes are typically arranged on each chromosome. In addition, we describe the specialized DNA sequences that are required for a chromosome to be accurately duplicated as a separate entity and passed on from one generation to the next. | Cell_Biology_Alberts. The most important function of DNA is to carry genes, the information that specifies all the RNA molecules and proteins that make up an organism—including information about when, in what types of cells, and in what quantity each RNA molecule and protein is to be made. The nuclear DNA of eukaryotes is divided up into chromosomes, and in this section we see how genes are typically arranged on each chromosome. In addition, we describe the specialized DNA sequences that are required for a chromosome to be accurately duplicated as a separate entity and passed on from one generation to the next. |
Cell_Biology_Alberts_795 | Cell_Biology_Alberts | We also confront the serious challenge of DNA packaging. If the double helices comprising all 46 chromosomes in a human cell could be laid end to end, they would reach approximately 2 meters; yet the nucleus, which contains the DNA, is only about 6 μm in diameter. This is geometrically equivalent to packing 40 km (24 miles) of extremely fine thread into a tennis ball. The complex task of packaging DNA is accomplished by specialized proteins that bind to the DNA and fold it, generating a series of organized coils and loops that provide increasingly higher levels of organization, and prevent the DNA from becoming an unmanageable tangle. Amazingly, although the DNA is very tightly compacted, it nevertheless remains accessible to the many enzymes in the cell that replicate it, repair it, and use its genes to produce RNA molecules and proteins. | Cell_Biology_Alberts. We also confront the serious challenge of DNA packaging. If the double helices comprising all 46 chromosomes in a human cell could be laid end to end, they would reach approximately 2 meters; yet the nucleus, which contains the DNA, is only about 6 μm in diameter. This is geometrically equivalent to packing 40 km (24 miles) of extremely fine thread into a tennis ball. The complex task of packaging DNA is accomplished by specialized proteins that bind to the DNA and fold it, generating a series of organized coils and loops that provide increasingly higher levels of organization, and prevent the DNA from becoming an unmanageable tangle. Amazingly, although the DNA is very tightly compacted, it nevertheless remains accessible to the many enzymes in the cell that replicate it, repair it, and use its genes to produce RNA molecules and proteins. |
Cell_Biology_Alberts_796 | Cell_Biology_Alberts | endoplasmic reticulum nuclear pore outer nuclear membrane inner nuclear membrane nuclear envelope nuclear lamina microtubule centrosome nucleolus DNA and associated proteins (chromatin), plus many RNA and protein molecules 2 µm (A) (B) nucleolus nuclear envelope heterochromatinheterochromatin | Cell_Biology_Alberts. endoplasmic reticulum nuclear pore outer nuclear membrane inner nuclear membrane nuclear envelope nuclear lamina microtubule centrosome nucleolus DNA and associated proteins (chromatin), plus many RNA and protein molecules 2 µm (A) (B) nucleolus nuclear envelope heterochromatinheterochromatin |
Cell_Biology_Alberts_797 | Cell_Biology_Alberts | Figure 4–9 A cross-sectional view of a typical cell nucleus. (A) Electron micrograph of a thin section through the nucleus of a human fibroblast. (b) Schematic drawing, showing that the nuclear envelope consists of two membranes, the outer one being continuous with the endoplasmic reticulum (ER) membrane (see also Figure 12–7). The space inside the endoplasmic reticulum (the ER lumen) is colored yellow; it is continuous with the space between the two nuclear membranes. The lipid bilayers of the inner and outer nuclear membranes are connected at each nuclear pore. A sheetlike network of intermediate filaments (brown) inside the nucleus forms the nuclear lamina (brown), providing mechanical support for the nuclear envelope (for details, see Chapter 12). The dark-staining heterochromatin contains specially condensed regions of DNA that will be discussed later. (A, courtesy of E.g. Jordan and J. mcgovern.) Eukaryotic DNA Is Packaged into a Set of Chromosomes | Cell_Biology_Alberts. Figure 4–9 A cross-sectional view of a typical cell nucleus. (A) Electron micrograph of a thin section through the nucleus of a human fibroblast. (b) Schematic drawing, showing that the nuclear envelope consists of two membranes, the outer one being continuous with the endoplasmic reticulum (ER) membrane (see also Figure 12–7). The space inside the endoplasmic reticulum (the ER lumen) is colored yellow; it is continuous with the space between the two nuclear membranes. The lipid bilayers of the inner and outer nuclear membranes are connected at each nuclear pore. A sheetlike network of intermediate filaments (brown) inside the nucleus forms the nuclear lamina (brown), providing mechanical support for the nuclear envelope (for details, see Chapter 12). The dark-staining heterochromatin contains specially condensed regions of DNA that will be discussed later. (A, courtesy of E.g. Jordan and J. mcgovern.) Eukaryotic DNA Is Packaged into a Set of Chromosomes |
Cell_Biology_Alberts_798 | Cell_Biology_Alberts | Eukaryotic DNA Is Packaged into a Set of Chromosomes Each chromosome in a eukaryotic cell consists of a single, enormously long linear DNA molecule along with the proteins that fold and pack the fine DNA thread into a more compact structure. In addition to the proteins involved in packaging, chromosomes are also associated with many other proteins (as well as numerous RNA molecules). These are required for the processes of gene expression, DNA replication, and DNA repair. The complex of DNA and tightly bound protein is called chromatin (from the Greek chroma, “color,” because of its staining properties). | Cell_Biology_Alberts. Eukaryotic DNA Is Packaged into a Set of Chromosomes Each chromosome in a eukaryotic cell consists of a single, enormously long linear DNA molecule along with the proteins that fold and pack the fine DNA thread into a more compact structure. In addition to the proteins involved in packaging, chromosomes are also associated with many other proteins (as well as numerous RNA molecules). These are required for the processes of gene expression, DNA replication, and DNA repair. The complex of DNA and tightly bound protein is called chromatin (from the Greek chroma, “color,” because of its staining properties). |
Cell_Biology_Alberts_799 | Cell_Biology_Alberts | Bacteria lack a special nuclear compartment, and they generally carry their genes on a single DNA molecule, which is often circular (see Figure 1–24). This DNA is also associated with proteins that package and condense it, but they are different from the proteins that perform these functions in eukaryotes. Although the bacterial DNA with its attendant proteins is often called the bacterial “chromosome,” it does not have the same structure as eukaryotic chromosomes, and less is known about how the bacterial DNA is packaged. Therefore, our discussion of chromosome structure will focus almost entirely on eukaryotic chromosomes. | Cell_Biology_Alberts. Bacteria lack a special nuclear compartment, and they generally carry their genes on a single DNA molecule, which is often circular (see Figure 1–24). This DNA is also associated with proteins that package and condense it, but they are different from the proteins that perform these functions in eukaryotes. Although the bacterial DNA with its attendant proteins is often called the bacterial “chromosome,” it does not have the same structure as eukaryotic chromosomes, and less is known about how the bacterial DNA is packaged. Therefore, our discussion of chromosome structure will focus almost entirely on eukaryotic chromosomes. |
Cell_Biology_Alberts_800 | Cell_Biology_Alberts | With the exception of the gametes (eggs and sperm) and a few highly specialized cell types that cannot multiply and either lack DNA altogether (for example, red blood cells) or have replicated their DNA without completing cell division (for example, megakaryocytes), each human cell nucleus contains two copies of each chromosome, one inherited from the mother and one from the father. The maternal and paternal chromosomes of a pair are called homologous chromosomes (homologs). The only nonhomologous chromosome pairs are the sex chromosomes in males, where a Y chromosome is inherited from the father and an X chromosome from the mother. Thus, each human cell contains a total of 46 chromosomes—22 pairs common to both males and females, plus two so-called sex chromosomes (X and Y in males, two Xs in females). These human chromosomes can be readily distinguished by “painting” each one a different color using a technique based on DNA hybridization (Figure 4–10). In this method (described in | Cell_Biology_Alberts. With the exception of the gametes (eggs and sperm) and a few highly specialized cell types that cannot multiply and either lack DNA altogether (for example, red blood cells) or have replicated their DNA without completing cell division (for example, megakaryocytes), each human cell nucleus contains two copies of each chromosome, one inherited from the mother and one from the father. The maternal and paternal chromosomes of a pair are called homologous chromosomes (homologs). The only nonhomologous chromosome pairs are the sex chromosomes in males, where a Y chromosome is inherited from the father and an X chromosome from the mother. Thus, each human cell contains a total of 46 chromosomes—22 pairs common to both males and females, plus two so-called sex chromosomes (X and Y in males, two Xs in females). These human chromosomes can be readily distinguished by “painting” each one a different color using a technique based on DNA hybridization (Figure 4–10). In this method (described in |
Cell_Biology_Alberts_801 | Cell_Biology_Alberts | Xs in females). These human chromosomes can be readily distinguished by “painting” each one a different color using a technique based on DNA hybridization (Figure 4–10). In this method (described in detail in Chapter 8), a short strand of nucleic acid tagged with a fluorescent dye serves as a “probe” that picks out its complementary DNA sequence, lighting up the target chromosome at any site where it binds. Chromosome painting is most frequently done at the stage in the cell cycle called mitosis, when chromosomes are especially compacted and easy to visualize (see below). | Cell_Biology_Alberts. Xs in females). These human chromosomes can be readily distinguished by “painting” each one a different color using a technique based on DNA hybridization (Figure 4–10). In this method (described in detail in Chapter 8), a short strand of nucleic acid tagged with a fluorescent dye serves as a “probe” that picks out its complementary DNA sequence, lighting up the target chromosome at any site where it binds. Chromosome painting is most frequently done at the stage in the cell cycle called mitosis, when chromosomes are especially compacted and easy to visualize (see below). |
Cell_Biology_Alberts_802 | Cell_Biology_Alberts | Another more traditional way to distinguish one chromosome from another is to stain them with dyes that reveal a striking and reproducible pattern of bands along each mitotic chromosome (Figure 4–11). These banding patterns presumably reflect variations in chromatin structure, but their basis is not well understood. Nevertheless, the pattern of bands on each type of chromosome is unique, and it provided the initial means to identify and number each human chromosome reliably. | Cell_Biology_Alberts. Another more traditional way to distinguish one chromosome from another is to stain them with dyes that reveal a striking and reproducible pattern of bands along each mitotic chromosome (Figure 4–11). These banding patterns presumably reflect variations in chromatin structure, but their basis is not well understood. Nevertheless, the pattern of bands on each type of chromosome is unique, and it provided the initial means to identify and number each human chromosome reliably. |
Cell_Biology_Alberts_803 | Cell_Biology_Alberts | Figure 4–10 The complete set of human chromosomes. These chromosomes, from a female, were isolated from a cell undergoing nuclear division (mitosis) and are therefore highly compacted. Each chromosome has been “painted” a different color to permit its unambiguous identification under the fluorescence microscope, using a technique called “spectral karyotyping.” Chromosome painting can be performed by exposing the chromosomes to a large collection of DNA molecules whose sequence matches known DNA sequences from the human genome. The set of sequences matching each chromosome is coupled to a different combination of fluorescent dyes. DNA molecules derived from chromosome 1 are labeled with one specific dye combination, those from chromosome 2 with another, and so on. because the labeled DNA can form base pairs, or hybridize, only to the chromosome from which it was derived, each chromosome becomes labeled with a different combination of dyes. For such experiments, the chromosomes are | Cell_Biology_Alberts. Figure 4–10 The complete set of human chromosomes. These chromosomes, from a female, were isolated from a cell undergoing nuclear division (mitosis) and are therefore highly compacted. Each chromosome has been “painted” a different color to permit its unambiguous identification under the fluorescence microscope, using a technique called “spectral karyotyping.” Chromosome painting can be performed by exposing the chromosomes to a large collection of DNA molecules whose sequence matches known DNA sequences from the human genome. The set of sequences matching each chromosome is coupled to a different combination of fluorescent dyes. DNA molecules derived from chromosome 1 are labeled with one specific dye combination, those from chromosome 2 with another, and so on. because the labeled DNA can form base pairs, or hybridize, only to the chromosome from which it was derived, each chromosome becomes labeled with a different combination of dyes. For such experiments, the chromosomes are |
Cell_Biology_Alberts_804 | Cell_Biology_Alberts | can form base pairs, or hybridize, only to the chromosome from which it was derived, each chromosome becomes labeled with a different combination of dyes. For such experiments, the chromosomes are subjected to treatments that separate the two strands of double-helical DNA in a way that permits base-pairing with the single-stranded labeled DNA, but keeps the overall chromosome structure relatively intact. (A) The chromosomes visualized as they originally spilled from the lysed cell. | Cell_Biology_Alberts. can form base pairs, or hybridize, only to the chromosome from which it was derived, each chromosome becomes labeled with a different combination of dyes. For such experiments, the chromosomes are subjected to treatments that separate the two strands of double-helical DNA in a way that permits base-pairing with the single-stranded labeled DNA, but keeps the overall chromosome structure relatively intact. (A) The chromosomes visualized as they originally spilled from the lysed cell. |
Cell_Biology_Alberts_805 | Cell_Biology_Alberts | The same chromosomes artificially lined up in their numerical order. This arrangement of the full chromosome set is called a karyotype. (Adapted from N. mcNeil and T. Ried, Expert Rev. Mol. Med. 2:1–14, 2000. with permission from Cambridge University Press.) | Cell_Biology_Alberts. The same chromosomes artificially lined up in their numerical order. This arrangement of the full chromosome set is called a karyotype. (Adapted from N. mcNeil and T. Ried, Expert Rev. Mol. Med. 2:1–14, 2000. with permission from Cambridge University Press.) |
Cell_Biology_Alberts_806 | Cell_Biology_Alberts | N. mcNeil and T. Ried, Expert Rev. Mol. Med. 2:1–14, 2000. with permission from Cambridge University Press.) Figure 4–11 The banding patterns of human chromosomes. Chromosomes 1–22 are numbered in approximate order of size. A typical human cell contains two of each of these chromosomes, plus two sex chromosomes—two X chromosomes in a female, one X and one Y chromosome in a male. The chromosomes used to make these maps were stained at an early stage in mitosis, when the chromosomes are incompletely compacted. The horizontal red line represents the position of the centromere (see Figure 4–19), which appears as a constriction on mitotic chromosomes. The red knobs on chromosomes 13, 14, 15, 21, and 22 indicate the positions of genes that code for the large ribosomal RNAs (discussed in Chapter 6). These banding patterns are obtained by staining chromosomes with giemsa stain, and they can be observed under the light microscope. (Adapted from | Cell_Biology_Alberts. N. mcNeil and T. Ried, Expert Rev. Mol. Med. 2:1–14, 2000. with permission from Cambridge University Press.) Figure 4–11 The banding patterns of human chromosomes. Chromosomes 1–22 are numbered in approximate order of size. A typical human cell contains two of each of these chromosomes, plus two sex chromosomes—two X chromosomes in a female, one X and one Y chromosome in a male. The chromosomes used to make these maps were stained at an early stage in mitosis, when the chromosomes are incompletely compacted. The horizontal red line represents the position of the centromere (see Figure 4–19), which appears as a constriction on mitotic chromosomes. The red knobs on chromosomes 13, 14, 15, 21, and 22 indicate the positions of genes that code for the large ribosomal RNAs (discussed in Chapter 6). These banding patterns are obtained by staining chromosomes with giemsa stain, and they can be observed under the light microscope. (Adapted from |
Cell_Biology_Alberts_807 | Cell_Biology_Alberts | U. Francke, Cytogenet. Cell Genet. 31:24– 32, 1981. with permission from the author.) Figure 4–12 Aberrant human chromosomes. (A) Two normal human chromosomes, 4 and 6. (b) In an individual carrying a balanced chromosomal translocation, the DNA double helix in one chromosome has crossed over with the DNA double helix in the other chromosome due to an abnormal recombination event. The chromosome painting technique used on the chromosomes in each of the sets allows the identification of even short pieces of chromosomes that have become translocated, a frequent event in cancer cells. (Courtesy of Zhenya Tang and the NIgmS Human genetic Cell Repository at the Coriell Institute for medical Research: gm21880.) | Cell_Biology_Alberts. U. Francke, Cytogenet. Cell Genet. 31:24– 32, 1981. with permission from the author.) Figure 4–12 Aberrant human chromosomes. (A) Two normal human chromosomes, 4 and 6. (b) In an individual carrying a balanced chromosomal translocation, the DNA double helix in one chromosome has crossed over with the DNA double helix in the other chromosome due to an abnormal recombination event. The chromosome painting technique used on the chromosomes in each of the sets allows the identification of even short pieces of chromosomes that have become translocated, a frequent event in cancer cells. (Courtesy of Zhenya Tang and the NIgmS Human genetic Cell Repository at the Coriell Institute for medical Research: gm21880.) |
Cell_Biology_Alberts_808 | Cell_Biology_Alberts | The display of the 46 human chromosomes at mitosis is called the human karyotype. If parts of chromosomes are lost or are switched between chromosomes, these changes can be detected either by changes in the banding patterns or—with greater sensitivity—by changes in the pattern of chromosome painting (Figure 4–12). Cytogeneticists use these alterations to detect inherited chromosome abnormalities and to reveal the chromosome rearrangements that occur in cancer cells as they progress to malignancy (discussed in Chapter 20). Chromosomes Contain long Strings of genes | Cell_Biology_Alberts. The display of the 46 human chromosomes at mitosis is called the human karyotype. If parts of chromosomes are lost or are switched between chromosomes, these changes can be detected either by changes in the banding patterns or—with greater sensitivity—by changes in the pattern of chromosome painting (Figure 4–12). Cytogeneticists use these alterations to detect inherited chromosome abnormalities and to reveal the chromosome rearrangements that occur in cancer cells as they progress to malignancy (discussed in Chapter 20). Chromosomes Contain long Strings of genes |
Cell_Biology_Alberts_809 | Cell_Biology_Alberts | Chromosomes Contain long Strings of genes Chromosomes carry genes—the functional units of heredity. A gene is often defined as a segment of DNA that contains the instructions for making a particular protein (or a set of closely related proteins), but this definition is too narrow. Genes that code for protein are indeed the majority, and most of the genes with clear-cut mutant phenotypes fall under this heading. In addition, however, there are many “RNA genes”—segments of DNA that generate a functionally significant RNA molecule, instead of a protein, as their final product. We shall say more about the RNA genes and their products later. | Cell_Biology_Alberts. Chromosomes Contain long Strings of genes Chromosomes carry genes—the functional units of heredity. A gene is often defined as a segment of DNA that contains the instructions for making a particular protein (or a set of closely related proteins), but this definition is too narrow. Genes that code for protein are indeed the majority, and most of the genes with clear-cut mutant phenotypes fall under this heading. In addition, however, there are many “RNA genes”—segments of DNA that generate a functionally significant RNA molecule, instead of a protein, as their final product. We shall say more about the RNA genes and their products later. |
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.