id
stringlengths
14
28
title
stringclasses
18 values
content
stringlengths
2
999
contents
stringlengths
19
1.02k
Biochemistry_Lippincott_1583
Biochemistry_Lippinco
A. Antiparallel binding between codon and anticodon Binding of the tRNA anticodon to the mRNA codon follows the rules of complementary and antiparallel binding, that is, the mRNA codon is read 5′→3′ by an anticodon pairing in the opposite (3′→5′) orientation (Fig. 32.9). [Note: Nucleotide sequences are always written in the 5′ to 3′ direction unless otherwise noted. Two nucleotide sequences orient in an antiparallel manner.] B. Wobble hypothesis
Biochemistry_Lippinco. A. Antiparallel binding between codon and anticodon Binding of the tRNA anticodon to the mRNA codon follows the rules of complementary and antiparallel binding, that is, the mRNA codon is read 5′→3′ by an anticodon pairing in the opposite (3′→5′) orientation (Fig. 32.9). [Note: Nucleotide sequences are always written in the 5′ to 3′ direction unless otherwise noted. Two nucleotide sequences orient in an antiparallel manner.] B. Wobble hypothesis
Biochemistry_Lippincott_1584
Biochemistry_Lippinco
B. Wobble hypothesis The mechanism by which a tRNA can recognize more than one codon for a specific amino acid is described by the wobble hypothesis, which states that codon–anticodon pairing follows the traditional Watson-Crick rules (G pairs with C and A pairs with U) for the first two bases of the codon but can be less stringent for the last base. The base at the 5′-end of the anticodon (the first base of the anticodon) is not as spatially defined as the other two bases. Movement of that first base allows nontraditional base-pairing with the 3′-base of the codon (the last base of the codon). This movement is called wobble and allows a single tRNA to recognize more than one codon. Examples of these flexible pairings are shown in Figure 32.9. The result of wobble is that 61 tRNA species are not required to read the 61 codons that code for amino acids. V. STEPS IN TRANSLATION
Biochemistry_Lippinco. B. Wobble hypothesis The mechanism by which a tRNA can recognize more than one codon for a specific amino acid is described by the wobble hypothesis, which states that codon–anticodon pairing follows the traditional Watson-Crick rules (G pairs with C and A pairs with U) for the first two bases of the codon but can be less stringent for the last base. The base at the 5′-end of the anticodon (the first base of the anticodon) is not as spatially defined as the other two bases. Movement of that first base allows nontraditional base-pairing with the 3′-base of the codon (the last base of the codon). This movement is called wobble and allows a single tRNA to recognize more than one codon. Examples of these flexible pairings are shown in Figure 32.9. The result of wobble is that 61 tRNA species are not required to read the 61 codons that code for amino acids. V. STEPS IN TRANSLATION
Biochemistry_Lippincott_1585
Biochemistry_Lippinco
V. STEPS IN TRANSLATION The process of protein synthesis translates the 3-letter alphabet of nucleotide sequences on mRNA into the 20-letter alphabet of amino acids that constitute proteins. The mRNA is translated from its 5′-end to its 3′-end, producing a protein synthesized from its amino (N)-terminal end to its carboxyl (C)-terminal end. Prokaryotic mRNA often have several coding regions (that is, they are polycistronic; see p. 434). Each coding region has its own initiation and termination codon and produces a separate species of polypeptide. In contrast, each eukaryotic mRNA has only one coding region (that is, it is monocistronic). The process of translation is divided into three separate steps: initiation, elongation, and termination. Eukaryotic translation resembles that of prokaryotes in most aspects. Individual differences are noted in the text.
Biochemistry_Lippinco. V. STEPS IN TRANSLATION The process of protein synthesis translates the 3-letter alphabet of nucleotide sequences on mRNA into the 20-letter alphabet of amino acids that constitute proteins. The mRNA is translated from its 5′-end to its 3′-end, producing a protein synthesized from its amino (N)-terminal end to its carboxyl (C)-terminal end. Prokaryotic mRNA often have several coding regions (that is, they are polycistronic; see p. 434). Each coding region has its own initiation and termination codon and produces a separate species of polypeptide. In contrast, each eukaryotic mRNA has only one coding region (that is, it is monocistronic). The process of translation is divided into three separate steps: initiation, elongation, and termination. Eukaryotic translation resembles that of prokaryotes in most aspects. Individual differences are noted in the text.
Biochemistry_Lippincott_1586
Biochemistry_Lippinco
One important difference is that translation and transcription are temporally linked in prokaryotes, with translation starting before transcription is completed as a consequence of the lack of a nuclear membrane in prokaryotes. A. Initiation Initiation of protein synthesis involves the assembly of the components of the translation system before peptide-bond formation occurs. These components include the two ribosomal subunits, the mRNA to be translated, the aminoacyl-tRNA specified by the first codon in the message, GTP, and initiation factors that facilitate the assembly of this initiation complex (see Fig. 32.13). [Note: In prokaryotes, three initiation factors are known (IF-1, IF-2, and IF-3), whereas in eukaryotes, there are many (designated eIF to indicate eukaryotic origin). Eukaryotes also require ATP for initiation.] The following are two mechanisms by which the ribosome recognizes the nucleotide sequence (AUG) that initiates translation.
Biochemistry_Lippinco. One important difference is that translation and transcription are temporally linked in prokaryotes, with translation starting before transcription is completed as a consequence of the lack of a nuclear membrane in prokaryotes. A. Initiation Initiation of protein synthesis involves the assembly of the components of the translation system before peptide-bond formation occurs. These components include the two ribosomal subunits, the mRNA to be translated, the aminoacyl-tRNA specified by the first codon in the message, GTP, and initiation factors that facilitate the assembly of this initiation complex (see Fig. 32.13). [Note: In prokaryotes, three initiation factors are known (IF-1, IF-2, and IF-3), whereas in eukaryotes, there are many (designated eIF to indicate eukaryotic origin). Eukaryotes also require ATP for initiation.] The following are two mechanisms by which the ribosome recognizes the nucleotide sequence (AUG) that initiates translation.
Biochemistry_Lippincott_1587
Biochemistry_Lippinco
1. Shine-Dalgarno sequence: In Escherichia coli (E. coli), a purine-rich sequence of nucleotide bases, known as the Shine-Dalgarno (SD) sequence, is located six to ten bases upstream of the initiating AUG codon on the mRNA molecule (that is, near its 5′-end). The 16S rRNA component of the small (30S) ribosomal subunit has a nucleotide sequence near its 3′-end that is complementary to all or part of the SD sequence. Therefore, the 5′-end of the mRNA and the 3′-end of the 16S rRNA can form complementary base pairs, facilitating the positioning of the 30S subunit on the mRNA in close proximity to the initiating AUG codon (Fig. 32.10). 2.
Biochemistry_Lippinco. 1. Shine-Dalgarno sequence: In Escherichia coli (E. coli), a purine-rich sequence of nucleotide bases, known as the Shine-Dalgarno (SD) sequence, is located six to ten bases upstream of the initiating AUG codon on the mRNA molecule (that is, near its 5′-end). The 16S rRNA component of the small (30S) ribosomal subunit has a nucleotide sequence near its 3′-end that is complementary to all or part of the SD sequence. Therefore, the 5′-end of the mRNA and the 3′-end of the 16S rRNA can form complementary base pairs, facilitating the positioning of the 30S subunit on the mRNA in close proximity to the initiating AUG codon (Fig. 32.10). 2.
Biochemistry_Lippincott_1588
Biochemistry_Lippinco
2. 5′-Cap: Eukaryotic mRNA do not have SD sequences. In eukaryotes, the small (40S) ribosomal subunit (aided by members of the eIF-4 family of proteins) binds close to the cap structure at the 5′-end of the mRNA and moves 5′→3′ along the mRNA until it encounters the initiator AUG. This scanning process requires ATP. Cap-independent initiation can occur if the 40S subunit binds to an internal ribosome entry site close to the start codon. [Note: Interactions between the cap-binding eIF-4 proteins and the poly-A tail–binding proteins on eukaryotic mRNA mediate circularization of the mRNA and likely prevent the use of incompletely processed mRNA in translation.] 3.
Biochemistry_Lippinco. 2. 5′-Cap: Eukaryotic mRNA do not have SD sequences. In eukaryotes, the small (40S) ribosomal subunit (aided by members of the eIF-4 family of proteins) binds close to the cap structure at the 5′-end of the mRNA and moves 5′→3′ along the mRNA until it encounters the initiator AUG. This scanning process requires ATP. Cap-independent initiation can occur if the 40S subunit binds to an internal ribosome entry site close to the start codon. [Note: Interactions between the cap-binding eIF-4 proteins and the poly-A tail–binding proteins on eukaryotic mRNA mediate circularization of the mRNA and likely prevent the use of incompletely processed mRNA in translation.] 3.
Biochemistry_Lippincott_1589
Biochemistry_Lippinco
Initiation codon: The initiating AUG is recognized by a special initiator tRNA (tRNAi). Recognition is facilitated by IF-2-GTP in prokaryotes and eIF-2-GTP (plus additional eIF) in eukaryotes. The charged tRNAi is the only tRNA recognized by (e)IF-2 and the only tRNA to go directly to the P site on the small subunit. [Note: Base modifications distinguish tRNAi from the tRNA used for internal AUG codons.] In bacteria and mitochondria, tRNAi carries an N-formylated methionine (fMet), as shown in Figure 32.11. After Met is attached to tRNAi, the formyl group is added by the enzyme transformylase, which uses N10-formyl tetrahydrofolate (see p. 267) as the carbon donor. In eukaryotes, tRNAi carries a Met that is not formylated. In both prokaryotic and eukaryotic cells, this N-terminal Met is usually removed before translation is completed. The large ribosomal subunit then joins the complex, and a functional ribosome is formed with the charged tRNAi in the P site. The A site is empty.
Biochemistry_Lippinco. Initiation codon: The initiating AUG is recognized by a special initiator tRNA (tRNAi). Recognition is facilitated by IF-2-GTP in prokaryotes and eIF-2-GTP (plus additional eIF) in eukaryotes. The charged tRNAi is the only tRNA recognized by (e)IF-2 and the only tRNA to go directly to the P site on the small subunit. [Note: Base modifications distinguish tRNAi from the tRNA used for internal AUG codons.] In bacteria and mitochondria, tRNAi carries an N-formylated methionine (fMet), as shown in Figure 32.11. After Met is attached to tRNAi, the formyl group is added by the enzyme transformylase, which uses N10-formyl tetrahydrofolate (see p. 267) as the carbon donor. In eukaryotes, tRNAi carries a Met that is not formylated. In both prokaryotic and eukaryotic cells, this N-terminal Met is usually removed before translation is completed. The large ribosomal subunit then joins the complex, and a functional ribosome is formed with the charged tRNAi in the P site. The A site is empty.
Biochemistry_Lippincott_1590
Biochemistry_Lippinco
is usually removed before translation is completed. The large ribosomal subunit then joins the complex, and a functional ribosome is formed with the charged tRNAi in the P site. The A site is empty. [Note: Specific (e)IF function as anti-association factors and prevent premature addition of the large subunit.] The GTP on (e)IF-2 gets hydrolyzed to GDP. In eukaryotes, the guanine nucleotide exchange factor eIF-2B facilitates the reactivation of eIF-2-GDP through replacement of GDP by GTP.
Biochemistry_Lippinco. is usually removed before translation is completed. The large ribosomal subunit then joins the complex, and a functional ribosome is formed with the charged tRNAi in the P site. The A site is empty. [Note: Specific (e)IF function as anti-association factors and prevent premature addition of the large subunit.] The GTP on (e)IF-2 gets hydrolyzed to GDP. In eukaryotes, the guanine nucleotide exchange factor eIF-2B facilitates the reactivation of eIF-2-GDP through replacement of GDP by GTP.
Biochemistry_Lippincott_1591
Biochemistry_Lippinco
B. Elongation
Biochemistry_Lippinco. B. Elongation
Biochemistry_Lippincott_1592
Biochemistry_Lippinco
Elongation of the polypeptide involves the addition of amino acids to the carboxyl end of the growing chain. Delivery of the aminoacyl-tRNA whose codon appears next on the mRNA template in the ribosomal A site (a process known as decoding) is facilitated in E. coli by elongation factors EF-Tu-GTP and EF-Ts and requires GTP hydrolysis. [Note: In eukaryotes, comparable elongation factors are EF-1α-GTP and EF-1βγ. Both EF-Ts and EF-1βγ function in guanine nucleotide exchange.] Peptide-bond formation between the α-carboxyl group of the amino acid in the P site and the αamino group of the amino acid in the A site is catalyzed by peptidyltransferase, an activity intrinsic to an rRNA of the large subunit (Fig. 32.12). [Note: Because this rRNA catalyzes the reaction, it is a ribozyme (see p. 54).] After the peptide bond has been formed, the peptide on the tRNA at the P site is transferred to the amino acid on the tRNA at the A site, a process known as transpeptidation. The ribosome then
Biochemistry_Lippinco. Elongation of the polypeptide involves the addition of amino acids to the carboxyl end of the growing chain. Delivery of the aminoacyl-tRNA whose codon appears next on the mRNA template in the ribosomal A site (a process known as decoding) is facilitated in E. coli by elongation factors EF-Tu-GTP and EF-Ts and requires GTP hydrolysis. [Note: In eukaryotes, comparable elongation factors are EF-1α-GTP and EF-1βγ. Both EF-Ts and EF-1βγ function in guanine nucleotide exchange.] Peptide-bond formation between the α-carboxyl group of the amino acid in the P site and the αamino group of the amino acid in the A site is catalyzed by peptidyltransferase, an activity intrinsic to an rRNA of the large subunit (Fig. 32.12). [Note: Because this rRNA catalyzes the reaction, it is a ribozyme (see p. 54).] After the peptide bond has been formed, the peptide on the tRNA at the P site is transferred to the amino acid on the tRNA at the A site, a process known as transpeptidation. The ribosome then
Biochemistry_Lippincott_1593
Biochemistry_Lippinco
54).] After the peptide bond has been formed, the peptide on the tRNA at the P site is transferred to the amino acid on the tRNA at the A site, a process known as transpeptidation. The ribosome then advances three nucleotides toward the 3′-end of the mRNA. This process is known as translocation and, in prokaryotes, requires the participation of EF-G-GTP (eukaryotes use EF-2-GTP) and GTP hydrolysis. Translocation causes movement of the uncharged tRNA from the P to the E site for release and movement of the peptidyl-tRNA from the A to the P site. The process is repeated until a termination codon is encountered. [Note: Because of the length of most mRNA, more than one ribosome at a time can translate a message. Such a complex of one mRNA and a number of ribosomes is called a polysome, or polyribosome.]
Biochemistry_Lippinco. 54).] After the peptide bond has been formed, the peptide on the tRNA at the P site is transferred to the amino acid on the tRNA at the A site, a process known as transpeptidation. The ribosome then advances three nucleotides toward the 3′-end of the mRNA. This process is known as translocation and, in prokaryotes, requires the participation of EF-G-GTP (eukaryotes use EF-2-GTP) and GTP hydrolysis. Translocation causes movement of the uncharged tRNA from the P to the E site for release and movement of the peptidyl-tRNA from the A to the P site. The process is repeated until a termination codon is encountered. [Note: Because of the length of most mRNA, more than one ribosome at a time can translate a message. Such a complex of one mRNA and a number of ribosomes is called a polysome, or polyribosome.]
Biochemistry_Lippincott_1594
Biochemistry_Lippinco
C. Termination
Biochemistry_Lippinco. C. Termination
Biochemistry_Lippincott_1595
Biochemistry_Lippinco
Termination occurs when one of the three termination codons moves into the A site. These codons are recognized in E. coli by release factors: RF-1, which recognizes UAA and UAG, and RF-2, which recognizes UGA and UAA. The binding of these release factors results in hydrolysis of the bond linking the peptide to the tRNA at the P site, causing the nascent protein to be released from the ribosome. A third release factor, RF-3-GTP, then causes the release of RF-1 or RF-2 as GTP is hydrolyzed (see Fig. 32.13). [Note: Eukaryotes have a single release factor, eRF, which recognizes all three termination codons. A second factor, eRF-3, functions like the prokaryotic RF-3. See Figure 32.14 for a summary of the factors used in translation.] The steps in prokaryotic protein synthesis, as well as some antibiotic inhibitors of the process, are summarized in Figure 32.13. The newly synthesized polypeptide may undergo further modification as described below, and the ribosomal subunits, mRNA, tRNA,
Biochemistry_Lippinco. Termination occurs when one of the three termination codons moves into the A site. These codons are recognized in E. coli by release factors: RF-1, which recognizes UAA and UAG, and RF-2, which recognizes UGA and UAA. The binding of these release factors results in hydrolysis of the bond linking the peptide to the tRNA at the P site, causing the nascent protein to be released from the ribosome. A third release factor, RF-3-GTP, then causes the release of RF-1 or RF-2 as GTP is hydrolyzed (see Fig. 32.13). [Note: Eukaryotes have a single release factor, eRF, which recognizes all three termination codons. A second factor, eRF-3, functions like the prokaryotic RF-3. See Figure 32.14 for a summary of the factors used in translation.] The steps in prokaryotic protein synthesis, as well as some antibiotic inhibitors of the process, are summarized in Figure 32.13. The newly synthesized polypeptide may undergo further modification as described below, and the ribosomal subunits, mRNA, tRNA,
Biochemistry_Lippincott_1596
Biochemistry_Lippinco
antibiotic inhibitors of the process, are summarized in Figure 32.13. The newly synthesized polypeptide may undergo further modification as described below, and the ribosomal subunits, mRNA, tRNA, and protein factors can be recycled and used to synthesize another polypeptide. [Note: In prokaryotes, ribosome recycling factors mediate separation of the subunits. In eukaryotes, eRF and ATP hydrolysis are required.]
Biochemistry_Lippinco. antibiotic inhibitors of the process, are summarized in Figure 32.13. The newly synthesized polypeptide may undergo further modification as described below, and the ribosomal subunits, mRNA, tRNA, and protein factors can be recycled and used to synthesize another polypeptide. [Note: In prokaryotes, ribosome recycling factors mediate separation of the subunits. In eukaryotes, eRF and ATP hydrolysis are required.]
Biochemistry_Lippincott_1597
Biochemistry_Lippinco
D. Translation regulation Gene expression is most commonly regulated at the transcriptional level, but translation may also be regulated. An important mechanism by which this is achieved in eukaryotes is by covalent modification of eIF-2: Phosphorylated eIF-2 is inactive (see p. 476). In both eukaryotes and prokaryotes, regulation can also be achieved through proteins that bind mRNA and inhibit its use by blocking translation. E. Protein folding Proteins must fold to assume their functional, native state. Folding can be spontaneous (as a result of the primary structure) or facilitated by proteins known as chaperones (see p. 20). F. Protein targeting
Biochemistry_Lippinco. D. Translation regulation Gene expression is most commonly regulated at the transcriptional level, but translation may also be regulated. An important mechanism by which this is achieved in eukaryotes is by covalent modification of eIF-2: Phosphorylated eIF-2 is inactive (see p. 476). In both eukaryotes and prokaryotes, regulation can also be achieved through proteins that bind mRNA and inhibit its use by blocking translation. E. Protein folding Proteins must fold to assume their functional, native state. Folding can be spontaneous (as a result of the primary structure) or facilitated by proteins known as chaperones (see p. 20). F. Protein targeting
Biochemistry_Lippincott_1598
Biochemistry_Lippinco
Although most protein synthesis in eukaryotes is initiated in the cytoplasm, many proteins perform their functions within subcellular organelles or outside of the cell. Such proteins normally contain amino acid sequences that direct the proteins to their final locations. For example, secreted proteins are targeted during synthesis (cotranslational targeting) to the RER by the presence of an N-terminal hydrophobic signal sequence. The sequence is recognized by the signal recognition particle (SRP), a ribonucleoprotein that binds the ribosome, halts elongation, and delivers the ribosome–peptide complex to an RER membrane channel (the translocon) via interaction with the SRP receptor. Translation resumes, the protein enters the RER lumen, and its signal sequence is cleaved (Fig. 32.15). The protein moves through the RER and the Golgi, is processed, packaged into vesicles, and secreted. Proteins targeted after synthesis (posttranslational) include nuclear proteins that contain an
Biochemistry_Lippinco. Although most protein synthesis in eukaryotes is initiated in the cytoplasm, many proteins perform their functions within subcellular organelles or outside of the cell. Such proteins normally contain amino acid sequences that direct the proteins to their final locations. For example, secreted proteins are targeted during synthesis (cotranslational targeting) to the RER by the presence of an N-terminal hydrophobic signal sequence. The sequence is recognized by the signal recognition particle (SRP), a ribonucleoprotein that binds the ribosome, halts elongation, and delivers the ribosome–peptide complex to an RER membrane channel (the translocon) via interaction with the SRP receptor. Translation resumes, the protein enters the RER lumen, and its signal sequence is cleaved (Fig. 32.15). The protein moves through the RER and the Golgi, is processed, packaged into vesicles, and secreted. Proteins targeted after synthesis (posttranslational) include nuclear proteins that contain an
Biochemistry_Lippincott_1599
Biochemistry_Lippinco
The protein moves through the RER and the Golgi, is processed, packaged into vesicles, and secreted. Proteins targeted after synthesis (posttranslational) include nuclear proteins that contain an internal, short, basic nuclear localization signal; mitochondrial matrix proteins that contain an N-terminal, amphipathic, α-helical mitochondrial entry sequence; and peroxisomal proteins that contain a C-terminal tripeptide signal.
Biochemistry_Lippinco. The protein moves through the RER and the Golgi, is processed, packaged into vesicles, and secreted. Proteins targeted after synthesis (posttranslational) include nuclear proteins that contain an internal, short, basic nuclear localization signal; mitochondrial matrix proteins that contain an N-terminal, amphipathic, α-helical mitochondrial entry sequence; and peroxisomal proteins that contain a C-terminal tripeptide signal.
Biochemistry_Lippincott_1600
Biochemistry_Lippinco
VI. CO-AND POSTTRANSLATIONAL MODIFICATIONS Many polypeptides are covalently modified, either while they are still attached to the ribosome (cotranslational) or after their synthesis has been completed (posttranslational). These modifications may include removal of part of the translated sequence or the covalent addition of one or more chemical groups required for protein activity. A. Trimming Many proteins destined for secretion are initially made as large, precursor molecules that are not functionally active. Portions of the protein must be removed by specialized endoproteases, resulting in the release of an active molecule. The cellular site of the cleavage reaction depends on the protein to be modified. Some precursor proteins are cleaved in the RER or the Golgi; others are cleaved in developing secretory vesicles (for example, insulin; see Fig. 23.4, p. 309); and still others, such as collagen (see p. 47), are cleaved after secretion. B. Covalent attachments
Biochemistry_Lippinco. VI. CO-AND POSTTRANSLATIONAL MODIFICATIONS Many polypeptides are covalently modified, either while they are still attached to the ribosome (cotranslational) or after their synthesis has been completed (posttranslational). These modifications may include removal of part of the translated sequence or the covalent addition of one or more chemical groups required for protein activity. A. Trimming Many proteins destined for secretion are initially made as large, precursor molecules that are not functionally active. Portions of the protein must be removed by specialized endoproteases, resulting in the release of an active molecule. The cellular site of the cleavage reaction depends on the protein to be modified. Some precursor proteins are cleaved in the RER or the Golgi; others are cleaved in developing secretory vesicles (for example, insulin; see Fig. 23.4, p. 309); and still others, such as collagen (see p. 47), are cleaved after secretion. B. Covalent attachments
Biochemistry_Lippincott_1601
Biochemistry_Lippinco
B. Covalent attachments Protein function can be affected by the covalent attachment of a variety of chemical groups (Fig. 32.16). Examples include the following. 1. Phosphorylation: Phosphorylation occurs on the hydroxyl groups of serine, threonine, or, less frequently, tyrosine residues in a protein. It is catalyzed by one of a family of protein kinases and may be reversed by the action of protein phosphatases. The phosphorylation may increase or decrease the functional activity of the protein. Several examples of phosphorylation reactions have been previously discussed (for example, see Chapter 11, p. 132, for the regulation of glycogen synthesis and degradation). 2.
Biochemistry_Lippinco. B. Covalent attachments Protein function can be affected by the covalent attachment of a variety of chemical groups (Fig. 32.16). Examples include the following. 1. Phosphorylation: Phosphorylation occurs on the hydroxyl groups of serine, threonine, or, less frequently, tyrosine residues in a protein. It is catalyzed by one of a family of protein kinases and may be reversed by the action of protein phosphatases. The phosphorylation may increase or decrease the functional activity of the protein. Several examples of phosphorylation reactions have been previously discussed (for example, see Chapter 11, p. 132, for the regulation of glycogen synthesis and degradation). 2.
Biochemistry_Lippincott_1602
Biochemistry_Lippinco
2. Glycosylation: Many of the proteins that are destined to become part of a membrane or to be secreted from a cell have carbohydrate chains added en bloc to the amide nitrogen of an asparagine (N-linked) or built sequentially on the hydroxyl groups of a serine, threonine, or hydroxylysine (O-linked). N-glycosylation occurs in the RER and Oglycosylation in the Golgi. (The process of producing such glycoproteins was discussed on p. 165.) N-glycosylated acid hydrolases are targeted to the matrix of lysosomes by the phosphorylation of mannose residues at carbon 6 (see p. 169). 3. Hydroxylation: Proline and lysine residues of the α chains of collagen are extensively hydroxylated by vitamin C–dependent hydroxylases in the RER (see p. 47). 4.
Biochemistry_Lippinco. 2. Glycosylation: Many of the proteins that are destined to become part of a membrane or to be secreted from a cell have carbohydrate chains added en bloc to the amide nitrogen of an asparagine (N-linked) or built sequentially on the hydroxyl groups of a serine, threonine, or hydroxylysine (O-linked). N-glycosylation occurs in the RER and Oglycosylation in the Golgi. (The process of producing such glycoproteins was discussed on p. 165.) N-glycosylated acid hydrolases are targeted to the matrix of lysosomes by the phosphorylation of mannose residues at carbon 6 (see p. 169). 3. Hydroxylation: Proline and lysine residues of the α chains of collagen are extensively hydroxylated by vitamin C–dependent hydroxylases in the RER (see p. 47). 4.
Biochemistry_Lippincott_1603
Biochemistry_Lippinco
3. Hydroxylation: Proline and lysine residues of the α chains of collagen are extensively hydroxylated by vitamin C–dependent hydroxylases in the RER (see p. 47). 4. Other covalent modifications: These may be required for the functional activity of a protein. For example, additional carboxyl groups can be added to glutamate residues by vitamin K–dependent carboxylation (see p. 393). The resulting γ-carboxyglutamate (Gla) residues are essential for the activity of several of the blood-clotting proteins. (See online Chapter 35.) Biotin is covalently bound to the ε-amino groups of lysine residues of biotin-dependent enzymes that catalyze carboxylation reactions such as pyruvate carboxylase (see Fig. 10.3 on p. 119). Attachment of lipids, such as farnesyl groups, can help anchor proteins to membranes (see p. 221). Many eukaryotic proteins are cotranslationally acetylated at the N-end. [Note: Reversible acetylation of histone proteins influences gene expression (see p. 476).]
Biochemistry_Lippinco. 3. Hydroxylation: Proline and lysine residues of the α chains of collagen are extensively hydroxylated by vitamin C–dependent hydroxylases in the RER (see p. 47). 4. Other covalent modifications: These may be required for the functional activity of a protein. For example, additional carboxyl groups can be added to glutamate residues by vitamin K–dependent carboxylation (see p. 393). The resulting γ-carboxyglutamate (Gla) residues are essential for the activity of several of the blood-clotting proteins. (See online Chapter 35.) Biotin is covalently bound to the ε-amino groups of lysine residues of biotin-dependent enzymes that catalyze carboxylation reactions such as pyruvate carboxylase (see Fig. 10.3 on p. 119). Attachment of lipids, such as farnesyl groups, can help anchor proteins to membranes (see p. 221). Many eukaryotic proteins are cotranslationally acetylated at the N-end. [Note: Reversible acetylation of histone proteins influences gene expression (see p. 476).]
Biochemistry_Lippincott_1604
Biochemistry_Lippinco
C. Protein degradation Proteins that are defective (for example, misfolded) or destined for rapid turnover are often marked for destruction by ubiquitination, the covalent attachment of chains of a small, highly conserved protein called ubiquitin (see Fig. 19.3 on p. 247). Proteins marked in this way are rapidly degraded by the proteasome, which is a macromolecular, ATP-dependent, proteolytic system located in the cytosol. For example, misfolding of the CFTR protein (see p. 450) results in its proteasomal degradation. [Note: If folding is impeded, unfolded proteins accumulate in the RER causing stress that triggers the unfolded protein response, in which the expression of chaperones is increased; global translation is decreased by eIF-2 phosphorylation; and the unfolded proteins are sent to the cytosol, ubiquitinated, and degraded in the proteasome by a process called ER-associated degradation.] VII. CHAPTER SUMMARY
Biochemistry_Lippinco. C. Protein degradation Proteins that are defective (for example, misfolded) or destined for rapid turnover are often marked for destruction by ubiquitination, the covalent attachment of chains of a small, highly conserved protein called ubiquitin (see Fig. 19.3 on p. 247). Proteins marked in this way are rapidly degraded by the proteasome, which is a macromolecular, ATP-dependent, proteolytic system located in the cytosol. For example, misfolding of the CFTR protein (see p. 450) results in its proteasomal degradation. [Note: If folding is impeded, unfolded proteins accumulate in the RER causing stress that triggers the unfolded protein response, in which the expression of chaperones is increased; global translation is decreased by eIF-2 phosphorylation; and the unfolded proteins are sent to the cytosol, ubiquitinated, and degraded in the proteasome by a process called ER-associated degradation.] VII. CHAPTER SUMMARY
Biochemistry_Lippincott_1605
Biochemistry_Lippinco
Codons are composed of three nucleotide bases presented in the messenger RNA (mRNA) language of adenine (A), guanine (G), cytosine (C), and uracil (U). They are always written 5′→3′. Of the 64 possible three-base combinations, 61 code for the 20 standard amino acids and 3 signal termination of protein synthesis (translation). Altering the nucleotide sequence in a codon can cause silent mutations (the altered codon codes for the original amino acid), missense mutations (the altered codon codes for a different amino acid), or nonsense mutations (the altered codon is a termination codon). Characteristics of the genetic code include specificity, universality, and degeneracy, and it is nonoverlapping and commaless (Fig. 32.17). Requirements for protein synthesis include all the amino acids that eventually appear in the finished protein; at least one specific type of transfer RNA (tRNA) for each amino acid; one aminoacyl-tRNA synthetase for each amino acid; the mRNA coding for the protein
Biochemistry_Lippinco. Codons are composed of three nucleotide bases presented in the messenger RNA (mRNA) language of adenine (A), guanine (G), cytosine (C), and uracil (U). They are always written 5′→3′. Of the 64 possible three-base combinations, 61 code for the 20 standard amino acids and 3 signal termination of protein synthesis (translation). Altering the nucleotide sequence in a codon can cause silent mutations (the altered codon codes for the original amino acid), missense mutations (the altered codon codes for a different amino acid), or nonsense mutations (the altered codon is a termination codon). Characteristics of the genetic code include specificity, universality, and degeneracy, and it is nonoverlapping and commaless (Fig. 32.17). Requirements for protein synthesis include all the amino acids that eventually appear in the finished protein; at least one specific type of transfer RNA (tRNA) for each amino acid; one aminoacyl-tRNA synthetase for each amino acid; the mRNA coding for the protein
Biochemistry_Lippincott_1606
Biochemistry_Lippinco
eventually appear in the finished protein; at least one specific type of transfer RNA (tRNA) for each amino acid; one aminoacyl-tRNA synthetase for each amino acid; the mRNA coding for the protein to be synthesized; fully competent ribosomes (70S in prokaryotes, 80S in eukaryotes); protein factors needed for initiation, elongation, and termination of protein synthesis; and ATP and guanosine triphosphate (GTP) as energy sources. tRNA has an attachment site for a specific amino acid at its 3′-end and an anticodon region that can recognize the codon specifying the amino acid the tRNA is carrying. Ribosomes are large complexes of protein and ribosomal RNA (rRNA). They consist of two subunits, 30S and 50S in prokaryotes and 40S and 60S in eukaryotes. Each ribosome has three binding sites for tRNA molecules: the A, P, and E sites that cover three neighboring codons. The A site binds an incoming aminoacyl-tRNA, the P site is occupied by peptidyl-tRNA, and the E site is occupied by the empty
Biochemistry_Lippinco. eventually appear in the finished protein; at least one specific type of transfer RNA (tRNA) for each amino acid; one aminoacyl-tRNA synthetase for each amino acid; the mRNA coding for the protein to be synthesized; fully competent ribosomes (70S in prokaryotes, 80S in eukaryotes); protein factors needed for initiation, elongation, and termination of protein synthesis; and ATP and guanosine triphosphate (GTP) as energy sources. tRNA has an attachment site for a specific amino acid at its 3′-end and an anticodon region that can recognize the codon specifying the amino acid the tRNA is carrying. Ribosomes are large complexes of protein and ribosomal RNA (rRNA). They consist of two subunits, 30S and 50S in prokaryotes and 40S and 60S in eukaryotes. Each ribosome has three binding sites for tRNA molecules: the A, P, and E sites that cover three neighboring codons. The A site binds an incoming aminoacyl-tRNA, the P site is occupied by peptidyl-tRNA, and the E site is occupied by the empty
Biochemistry_Lippincott_1607
Biochemistry_Lippinco
molecules: the A, P, and E sites that cover three neighboring codons. The A site binds an incoming aminoacyl-tRNA, the P site is occupied by peptidyl-tRNA, and the E site is occupied by the empty tRNA as it is about to exit the ribosome. Recognition of an mRNA codon is accomplished by the tRNA anticodon, which binds to the codon following the rules of complementarity and antiparallel binding. The wobble hypothesis states that the first (5′) base of the anticodon is not as spatially defined as the other two bases. Movement of that first base allows nontraditional base-pairing with the last (3′) base of the codon, thus allowing a single tRNA to recognize more than one codon for a specific amino acid. For initiation of protein synthesis, the components of the translation system are assembled, and mRNA associates with the small ribosomal subunit. The process requires initiation factors (IF). In prokaryotes, a purine-rich region of the mRNA (the Shine-Dalgarno sequence) base-pairs with a
Biochemistry_Lippinco. molecules: the A, P, and E sites that cover three neighboring codons. The A site binds an incoming aminoacyl-tRNA, the P site is occupied by peptidyl-tRNA, and the E site is occupied by the empty tRNA as it is about to exit the ribosome. Recognition of an mRNA codon is accomplished by the tRNA anticodon, which binds to the codon following the rules of complementarity and antiparallel binding. The wobble hypothesis states that the first (5′) base of the anticodon is not as spatially defined as the other two bases. Movement of that first base allows nontraditional base-pairing with the last (3′) base of the codon, thus allowing a single tRNA to recognize more than one codon for a specific amino acid. For initiation of protein synthesis, the components of the translation system are assembled, and mRNA associates with the small ribosomal subunit. The process requires initiation factors (IF). In prokaryotes, a purine-rich region of the mRNA (the Shine-Dalgarno sequence) base-pairs with a
Biochemistry_Lippincott_1608
Biochemistry_Lippinco
and mRNA associates with the small ribosomal subunit. The process requires initiation factors (IF). In prokaryotes, a purine-rich region of the mRNA (the Shine-Dalgarno sequence) base-pairs with a complementary sequence on 16S rRNA, resulting in the positioning of the small subunit on the mRNA so that translation can begin. The 5′-cap (bound by proteins of the eIF-4 family) on eukaryotic mRNA is used to position the small subunit on the mRNA. The initiation codon is AUG, and N-formylmethionine is the initiating amino acid in prokaryotes, whereas methionine is used in eukaryotes. The charged initiating tRNA (tRNAi) is brought to the P site by (e)IF-2. In elongation, the polypeptide chain is lengthened by the addition of amino acids to the carboxyl end of its growing chain. The process requires elongation factors that facilitate the binding of the aminoacyl-tRNA to the A site as well as the movement of the ribosome along the mRNA. The formation of the peptide bond is catalyzed by
Biochemistry_Lippinco. and mRNA associates with the small ribosomal subunit. The process requires initiation factors (IF). In prokaryotes, a purine-rich region of the mRNA (the Shine-Dalgarno sequence) base-pairs with a complementary sequence on 16S rRNA, resulting in the positioning of the small subunit on the mRNA so that translation can begin. The 5′-cap (bound by proteins of the eIF-4 family) on eukaryotic mRNA is used to position the small subunit on the mRNA. The initiation codon is AUG, and N-formylmethionine is the initiating amino acid in prokaryotes, whereas methionine is used in eukaryotes. The charged initiating tRNA (tRNAi) is brought to the P site by (e)IF-2. In elongation, the polypeptide chain is lengthened by the addition of amino acids to the carboxyl end of its growing chain. The process requires elongation factors that facilitate the binding of the aminoacyl-tRNA to the A site as well as the movement of the ribosome along the mRNA. The formation of the peptide bond is catalyzed by
Biochemistry_Lippincott_1609
Biochemistry_Lippinco
requires elongation factors that facilitate the binding of the aminoacyl-tRNA to the A site as well as the movement of the ribosome along the mRNA. The formation of the peptide bond is catalyzed by peptidyltransferase, which is an activity intrinsic to the rRNA of the large subunit and, therefore, is a ribozyme. Following peptide-bond formation, the ribosome advances along the mRNA in the 5′→3′ direction to the next codon (translocation). Because of the length of most mRNA, more than one ribosome at a time can translate a message, forming a polysome. Termination begins when one of the three termination codons moves into the A site. These codons are recognized by release factors. The newly synthesized protein is released from the ribosomal complex, and the ribosome is dissociated from the mRNA. Initiation, elongation, and termination are driven by the hydrolysis of GTP. Initiation in eukaryotes also requires ATP for scanning. Numerous antibiotics interfere with the process of protein
Biochemistry_Lippinco. requires elongation factors that facilitate the binding of the aminoacyl-tRNA to the A site as well as the movement of the ribosome along the mRNA. The formation of the peptide bond is catalyzed by peptidyltransferase, which is an activity intrinsic to the rRNA of the large subunit and, therefore, is a ribozyme. Following peptide-bond formation, the ribosome advances along the mRNA in the 5′→3′ direction to the next codon (translocation). Because of the length of most mRNA, more than one ribosome at a time can translate a message, forming a polysome. Termination begins when one of the three termination codons moves into the A site. These codons are recognized by release factors. The newly synthesized protein is released from the ribosomal complex, and the ribosome is dissociated from the mRNA. Initiation, elongation, and termination are driven by the hydrolysis of GTP. Initiation in eukaryotes also requires ATP for scanning. Numerous antibiotics interfere with the process of protein
Biochemistry_Lippincott_1610
Biochemistry_Lippinco
mRNA. Initiation, elongation, and termination are driven by the hydrolysis of GTP. Initiation in eukaryotes also requires ATP for scanning. Numerous antibiotics interfere with the process of protein synthesis. Many polypeptide chains are covalently modified during or after translation. Such modifications include amino acid removal; phosphorylation, which may activate or inactivate the protein; glycosylation, which plays a role in protein targeting; and hydroxylation such as that seen in collagen. Protein targeting can be either cotranslational (as with secreted proteins) or posttranslational (as with mitochondrial matrix proteins). Proteins must fold to achieve their functional form. Folding can be spontaneous or facilitated by chaperones. Proteins that are defective (for example, misfolded) or destined for rapid turnover are marked for destruction by the attachment of chains of a small, highly conserved protein called ubiquitin. Ubiquitinated proteins are rapidly degraded by a
Biochemistry_Lippinco. mRNA. Initiation, elongation, and termination are driven by the hydrolysis of GTP. Initiation in eukaryotes also requires ATP for scanning. Numerous antibiotics interfere with the process of protein synthesis. Many polypeptide chains are covalently modified during or after translation. Such modifications include amino acid removal; phosphorylation, which may activate or inactivate the protein; glycosylation, which plays a role in protein targeting; and hydroxylation such as that seen in collagen. Protein targeting can be either cotranslational (as with secreted proteins) or posttranslational (as with mitochondrial matrix proteins). Proteins must fold to achieve their functional form. Folding can be spontaneous or facilitated by chaperones. Proteins that are defective (for example, misfolded) or destined for rapid turnover are marked for destruction by the attachment of chains of a small, highly conserved protein called ubiquitin. Ubiquitinated proteins are rapidly degraded by a
Biochemistry_Lippincott_1611
Biochemistry_Lippinco
or destined for rapid turnover are marked for destruction by the attachment of chains of a small, highly conserved protein called ubiquitin. Ubiquitinated proteins are rapidly degraded by a cytosolic complex known as the proteasome.
Biochemistry_Lippinco. or destined for rapid turnover are marked for destruction by the attachment of chains of a small, highly conserved protein called ubiquitin. Ubiquitinated proteins are rapidly degraded by a cytosolic complex known as the proteasome.
Biochemistry_Lippincott_1612
Biochemistry_Lippinco
Choose the ONE best answer. 2.1. A 20-year-old man with a microcytic anemia is found to have an abnormal form of β-globin (Hemoglobin Constant Spring) that is 172 amino acids long, rather than the 141 found in the normal protein. Which of the following point mutations is consistent with this abnormality? Use Figure 32.2 to answer the question. A. CGA→UGA B. GAU→GAC C. GCA→GAA D. UAA→CAA D. UAA→UAG
Biochemistry_Lippinco. Choose the ONE best answer. 2.1. A 20-year-old man with a microcytic anemia is found to have an abnormal form of β-globin (Hemoglobin Constant Spring) that is 172 amino acids long, rather than the 141 found in the normal protein. Which of the following point mutations is consistent with this abnormality? Use Figure 32.2 to answer the question. A. CGA→UGA B. GAU→GAC C. GCA→GAA D. UAA→CAA D. UAA→UAG
Biochemistry_Lippincott_1613
Biochemistry_Lippinco
A. CGA→UGA B. GAU→GAC C. GCA→GAA D. UAA→CAA D. UAA→UAG Correct answer = D. Mutating the normal termination (stop) codon from UAA to CAA in β-globin messenger RNA causes the ribosome to insert a glutamine at that point. It will continue extending the protein chain until it comes upon the next stop codon farther down the message, resulting in an abnormally long protein. The replacement of CGA (arginine) with UGA (stop) would cause the protein to be too short. GAU and GAC both code for aspartate and would cause no change in the protein. Changing GCA (alanine) to GAA (glutamate) would not change the size of the protein product. A change from UAA to UAG would simply change one termination codon for another and would have no effect on the protein.
Biochemistry_Lippinco. A. CGA→UGA B. GAU→GAC C. GCA→GAA D. UAA→CAA D. UAA→UAG Correct answer = D. Mutating the normal termination (stop) codon from UAA to CAA in β-globin messenger RNA causes the ribosome to insert a glutamine at that point. It will continue extending the protein chain until it comes upon the next stop codon farther down the message, resulting in an abnormally long protein. The replacement of CGA (arginine) with UGA (stop) would cause the protein to be too short. GAU and GAC both code for aspartate and would cause no change in the protein. Changing GCA (alanine) to GAA (glutamate) would not change the size of the protein product. A change from UAA to UAG would simply change one termination codon for another and would have no effect on the protein.
Biochemistry_Lippincott_1614
Biochemistry_Lippinco
2.2. A pharmaceutical company is studying a new antibiotic that inhibits bacterial protein synthesis. When this antibiotic is added to an in vitro protein synthesis system that is translating the messenger RNA sequence AUGUUUUUUUAG, the only product formed is the dipeptide fMet-Phe. What step in protein synthesis is most likely inhibited by the antibiotic? A. Initiation B. Binding of a charged transfer RNA to the ribosomal A site C. Peptidyltransferase activity D. Ribosomal translocation E. Termination Correct answer = D. Because fMet-Phe (formylated methionyl-phenylalanine) is made, the ribosomes must be able to complete initiation, bind Phe-tRNA to the A site, and use peptidyltransferase activity to form the first peptide bond. Because the ribosome is not able to proceed any further, ribosomal movement (translocation) is most likely the inhibited step. Therefore, the ribosome is stopped before it reaches the termination codon of this message.
Biochemistry_Lippinco. 2.2. A pharmaceutical company is studying a new antibiotic that inhibits bacterial protein synthesis. When this antibiotic is added to an in vitro protein synthesis system that is translating the messenger RNA sequence AUGUUUUUUUAG, the only product formed is the dipeptide fMet-Phe. What step in protein synthesis is most likely inhibited by the antibiotic? A. Initiation B. Binding of a charged transfer RNA to the ribosomal A site C. Peptidyltransferase activity D. Ribosomal translocation E. Termination Correct answer = D. Because fMet-Phe (formylated methionyl-phenylalanine) is made, the ribosomes must be able to complete initiation, bind Phe-tRNA to the A site, and use peptidyltransferase activity to form the first peptide bond. Because the ribosome is not able to proceed any further, ribosomal movement (translocation) is most likely the inhibited step. Therefore, the ribosome is stopped before it reaches the termination codon of this message.
Biochemistry_Lippincott_1615
Biochemistry_Lippinco
2.3. A transfer RNA (tRNA) molecule that is supposed to carry cysteine (tRNAcys) is mischarged, so that it actually carries alanine (ala-tRNAcys). Assuming no correction occurs, what will be the fate of this alanine residue during protein synthesis? It will: A. be incorporated into a protein in response to a codon for alanine. B. be incorporated into a protein in response to a codon for cysteine. C. be incorporated randomly at any codon. D. remain attached to the tRNA because it cannot be used for protein synthesis. E. be chemically converted to cysteine by cellular enzymes. Correct answer = B. Once an amino acid is attached to a tRNA molecule, only the anticodon of that tRNA determines the specificity of incorporation. Therefore, the incorrectly activated alanine will be incorporated into the protein at a position determined by a cysteine codon.
Biochemistry_Lippinco. 2.3. A transfer RNA (tRNA) molecule that is supposed to carry cysteine (tRNAcys) is mischarged, so that it actually carries alanine (ala-tRNAcys). Assuming no correction occurs, what will be the fate of this alanine residue during protein synthesis? It will: A. be incorporated into a protein in response to a codon for alanine. B. be incorporated into a protein in response to a codon for cysteine. C. be incorporated randomly at any codon. D. remain attached to the tRNA because it cannot be used for protein synthesis. E. be chemically converted to cysteine by cellular enzymes. Correct answer = B. Once an amino acid is attached to a tRNA molecule, only the anticodon of that tRNA determines the specificity of incorporation. Therefore, the incorrectly activated alanine will be incorporated into the protein at a position determined by a cysteine codon.
Biochemistry_Lippincott_1616
Biochemistry_Lippinco
2.4. In a patient with cystic fibrosis (CF) caused by the ∆F508 mutation, the mutant CF transmembrane conductance regulator (CFTR) protein folds incorrectly. The patient’s cells modify this abnormal protein by attaching ubiquitin molecules to it. What is the fate of this modified CFTR protein? A. It performs its normal function because the ubiquitin largely corrects for the effect of the mutation. B. It is degraded by the proteasome. C. It is placed into storage vesicles. D. It is repaired by cellular enzymes. E. It is secreted from the cell. Correct answer = B. Ubiquitination usually marks old, damaged, or misfolded proteins for destruction by the cytosolic proteasome. There is no known cellular mechanism for repair of damaged proteins. 2.5. Many antimicrobials inhibit translation. Which of the following antimicrobials is correctly paired with its mechanism of action? A. Erythromycin binds to the 60S ribosomal subunit. B. Puromycin inactivates elongation factor-2.
Biochemistry_Lippinco. 2.4. In a patient with cystic fibrosis (CF) caused by the ∆F508 mutation, the mutant CF transmembrane conductance regulator (CFTR) protein folds incorrectly. The patient’s cells modify this abnormal protein by attaching ubiquitin molecules to it. What is the fate of this modified CFTR protein? A. It performs its normal function because the ubiquitin largely corrects for the effect of the mutation. B. It is degraded by the proteasome. C. It is placed into storage vesicles. D. It is repaired by cellular enzymes. E. It is secreted from the cell. Correct answer = B. Ubiquitination usually marks old, damaged, or misfolded proteins for destruction by the cytosolic proteasome. There is no known cellular mechanism for repair of damaged proteins. 2.5. Many antimicrobials inhibit translation. Which of the following antimicrobials is correctly paired with its mechanism of action? A. Erythromycin binds to the 60S ribosomal subunit. B. Puromycin inactivates elongation factor-2.
Biochemistry_Lippincott_1617
Biochemistry_Lippinco
A. Erythromycin binds to the 60S ribosomal subunit. B. Puromycin inactivates elongation factor-2. C. Streptomycin binds to the 30S ribosomal subunit. D. Tetracyclines inhibit peptidyltransferase. Correct answer = C. Streptomycin binds the 30S subunit and inhibits translation initiation. Erythromycin binds the 50S ribosomal subunit (60S denotes a eukaryote) and blocks the tunnel through which the peptide leaves the ribosome. Puromycin has structural similarity to aminoacyl-transfer RNA. It is incorporated into the growing chain, inhibits elongation, and results in premature termination in both prokaryotes and eukaryotes. Tetracyclines bind the 30S ribosomal subunit and block access to the A site, inhibiting elongation.
Biochemistry_Lippinco. A. Erythromycin binds to the 60S ribosomal subunit. B. Puromycin inactivates elongation factor-2. C. Streptomycin binds to the 30S ribosomal subunit. D. Tetracyclines inhibit peptidyltransferase. Correct answer = C. Streptomycin binds the 30S subunit and inhibits translation initiation. Erythromycin binds the 50S ribosomal subunit (60S denotes a eukaryote) and blocks the tunnel through which the peptide leaves the ribosome. Puromycin has structural similarity to aminoacyl-transfer RNA. It is incorporated into the growing chain, inhibits elongation, and results in premature termination in both prokaryotes and eukaryotes. Tetracyclines bind the 30S ribosomal subunit and block access to the A site, inhibiting elongation.
Biochemistry_Lippincott_1618
Biochemistry_Lippinco
2.6. Translation of a synthetic polyribonucleotide containing the repeating sequence CAA in a cell-free protein-synthesizing system produces three homopolypeptides: polyglutamine, polyasparagine, and polythreonine. If the codons for glutamine and asparagine are CAA and AAC, respectively, which of the following triplets is the codon for threonine? A. AAC B. ACA C. CAA D. CAC E. CCA Correct answer = B. The synthetic polynucleotide sequence of CAACAACAACAA … could be read by the in vitro protein-synthesizing system starting at the first C, the first A, or the second A (that is, in any one of three reading frames). In the first case, the first triplet codon would be CAA, which codes glutamine; in the second case, the first triplet codon would be AAC, which codes for asparagine; in the last case, the first triplet codon would be ACA, which codes for threonine. 2.7. Which of the following is required for both prokaryotic and eukaryotic protein synthesis?
Biochemistry_Lippinco. 2.6. Translation of a synthetic polyribonucleotide containing the repeating sequence CAA in a cell-free protein-synthesizing system produces three homopolypeptides: polyglutamine, polyasparagine, and polythreonine. If the codons for glutamine and asparagine are CAA and AAC, respectively, which of the following triplets is the codon for threonine? A. AAC B. ACA C. CAA D. CAC E. CCA Correct answer = B. The synthetic polynucleotide sequence of CAACAACAACAA … could be read by the in vitro protein-synthesizing system starting at the first C, the first A, or the second A (that is, in any one of three reading frames). In the first case, the first triplet codon would be CAA, which codes glutamine; in the second case, the first triplet codon would be AAC, which codes for asparagine; in the last case, the first triplet codon would be ACA, which codes for threonine. 2.7. Which of the following is required for both prokaryotic and eukaryotic protein synthesis?
Biochemistry_Lippincott_1619
Biochemistry_Lippinco
2.7. Which of the following is required for both prokaryotic and eukaryotic protein synthesis? A. Binding of the small ribosomal subunit to the Shine-Dalgarno sequence B. Formylated methionyl-transfer (t)RNA C. Movement of the messenger RNA out of the nucleus and into the cytoplasm D. Recognition of the 5′-cap by initiation factors E. Translocation of the peptidyl-tRNA from the A site to the P site Correct answer = E. In both prokaryotes and eukaryotes, continued translation (elongation) requires movement of the peptidyl-tRNA from the A to the P site to allow the next aminoacyl-tRNA to enter the A site. Only prokaryotes have a Shine-Dalgarno sequence and use formylated methionine and only eukaryotes have a nucleus and co-and posttranscriptionally process their mRNA.
Biochemistry_Lippinco. 2.7. Which of the following is required for both prokaryotic and eukaryotic protein synthesis? A. Binding of the small ribosomal subunit to the Shine-Dalgarno sequence B. Formylated methionyl-transfer (t)RNA C. Movement of the messenger RNA out of the nucleus and into the cytoplasm D. Recognition of the 5′-cap by initiation factors E. Translocation of the peptidyl-tRNA from the A site to the P site Correct answer = E. In both prokaryotes and eukaryotes, continued translation (elongation) requires movement of the peptidyl-tRNA from the A to the P site to allow the next aminoacyl-tRNA to enter the A site. Only prokaryotes have a Shine-Dalgarno sequence and use formylated methionine and only eukaryotes have a nucleus and co-and posttranscriptionally process their mRNA.
Biochemistry_Lippincott_1620
Biochemistry_Lippinco
2.8. α1-Antitrypsin (AAT) deficiency can result in emphysema, a lung pathology, because the action of elastase, a serine protease, is unopposed. Deficiency of AAT in the lungs is the consequence of impaired secretion from the liver, the site of its synthesis. Proteins such as AAT that are destined to be secreted are best characterized by which of the following statements? A. Their synthesis is initiated on the smooth endoplasmic reticulum. B. They contain a mannose 6-phosphate targeting signal. C. They always contain methionine as the N-terminal amino acid. D. They are produced from translation products that have an N-terminal hydrophobic signal sequence. E. They contain no sugars with O-glycosidic linkages because their synthesis does not involve the Golgi.
Biochemistry_Lippinco. 2.8. α1-Antitrypsin (AAT) deficiency can result in emphysema, a lung pathology, because the action of elastase, a serine protease, is unopposed. Deficiency of AAT in the lungs is the consequence of impaired secretion from the liver, the site of its synthesis. Proteins such as AAT that are destined to be secreted are best characterized by which of the following statements? A. Their synthesis is initiated on the smooth endoplasmic reticulum. B. They contain a mannose 6-phosphate targeting signal. C. They always contain methionine as the N-terminal amino acid. D. They are produced from translation products that have an N-terminal hydrophobic signal sequence. E. They contain no sugars with O-glycosidic linkages because their synthesis does not involve the Golgi.
Biochemistry_Lippincott_1621
Biochemistry_Lippinco
E. They contain no sugars with O-glycosidic linkages because their synthesis does not involve the Golgi. Correct answer = D. Synthesis of secreted proteins is begun on free (cytosolic) ribosomes. As the N-terminal signal sequence of the peptide emerges from the ribosome, it is bound by the signal recognition particle, taken to the rough endoplasmic reticulum (RER), threaded into the lumen, and cleaved as translation continues. The proteins move through the RER and the Golgi and undergo processing such as N-glycosylation (RER) and O-glycosylation (Golgi). In the Golgi, they are packaged in secretory vesicles and released from the cell. The smooth endoplasmic reticulum is associated with synthesis of lipids, not proteins, and has no ribosomes attached. Phosphorylation at carbon 6 of terminal mannose residues in glycoproteins targets these proteins (acid hydrolases) to lysosomes. The N-terminal methionine is removed from most proteins during processing.
Biochemistry_Lippinco. E. They contain no sugars with O-glycosidic linkages because their synthesis does not involve the Golgi. Correct answer = D. Synthesis of secreted proteins is begun on free (cytosolic) ribosomes. As the N-terminal signal sequence of the peptide emerges from the ribosome, it is bound by the signal recognition particle, taken to the rough endoplasmic reticulum (RER), threaded into the lumen, and cleaved as translation continues. The proteins move through the RER and the Golgi and undergo processing such as N-glycosylation (RER) and O-glycosylation (Golgi). In the Golgi, they are packaged in secretory vesicles and released from the cell. The smooth endoplasmic reticulum is associated with synthesis of lipids, not proteins, and has no ribosomes attached. Phosphorylation at carbon 6 of terminal mannose residues in glycoproteins targets these proteins (acid hydrolases) to lysosomes. The N-terminal methionine is removed from most proteins during processing.
Biochemistry_Lippincott_1622
Biochemistry_Lippinco
2.9. Why is the genetic code described as both degenerate and unambiguous? A given amino acid can be coded for by more than one codon (degenerate code), but a given codon codes for just one particular amino acid (unambiguous code). Regulation of Gene Expression 33 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW
Biochemistry_Lippinco. 2.9. Why is the genetic code described as both degenerate and unambiguous? A given amino acid can be coded for by more than one codon (degenerate code), but a given codon codes for just one particular amino acid (unambiguous code). Regulation of Gene Expression 33 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW
Biochemistry_Lippincott_1623
Biochemistry_Lippinco
Regulation of Gene Expression 33 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW Gene expression refers to the multistep process that ultimately results in the production of a functional gene product, either ribonucleic acid (RNA) or protein. The first step in gene expression, the use of deoxyribonucleic acid (DNA) for the synthesis of RNA (transcription), is the primary site of regulation in both prokaryotes and eukaryotes. In eukaryotes, however, gene expression also involves extensive posttranscriptional and posttranslational processes as well as actions that influence access to particular regions of the DNA. Each of these steps can be regulated to provide additional control over the kinds and amounts of functional products that are produced.
Biochemistry_Lippinco. Regulation of Gene Expression 33 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW Gene expression refers to the multistep process that ultimately results in the production of a functional gene product, either ribonucleic acid (RNA) or protein. The first step in gene expression, the use of deoxyribonucleic acid (DNA) for the synthesis of RNA (transcription), is the primary site of regulation in both prokaryotes and eukaryotes. In eukaryotes, however, gene expression also involves extensive posttranscriptional and posttranslational processes as well as actions that influence access to particular regions of the DNA. Each of these steps can be regulated to provide additional control over the kinds and amounts of functional products that are produced.
Biochemistry_Lippincott_1624
Biochemistry_Lippinco
Not all genes are tightly regulated. For example, genes described as constitutive encode products required for basic cellular functions and so are expressed at essentially a constant level. They are also known as “housekeeping” genes. Regulated genes, however, are expressed only under certain conditions. They may be expressed in all cells or in only a subset of cells, for example, hepatocytes. The ability to regulate gene expression (that is, to determine if, how much, and when particular gene products will be made) gives the cell control over structure and function. It is the basis for cellular differentiation, morphogenesis, and adaptability of any organism. Control of gene expression is best understood in prokaryotes, but many themes are repeated in eukaryotes. Figure 33.1 shows some of the sites where gene expression can be controlled. II. REGULATORY SEQUENCES AND MOLECULES
Biochemistry_Lippinco. Not all genes are tightly regulated. For example, genes described as constitutive encode products required for basic cellular functions and so are expressed at essentially a constant level. They are also known as “housekeeping” genes. Regulated genes, however, are expressed only under certain conditions. They may be expressed in all cells or in only a subset of cells, for example, hepatocytes. The ability to regulate gene expression (that is, to determine if, how much, and when particular gene products will be made) gives the cell control over structure and function. It is the basis for cellular differentiation, morphogenesis, and adaptability of any organism. Control of gene expression is best understood in prokaryotes, but many themes are repeated in eukaryotes. Figure 33.1 shows some of the sites where gene expression can be controlled. II. REGULATORY SEQUENCES AND MOLECULES
Biochemistry_Lippincott_1625
Biochemistry_Lippinco
Regulation of transcription, the initial step in all gene expression, is controlled by regulatory sequences of DNA that are usually embedded in the noncoding regions of the genome. The interaction between these DNA sequences and regulatory molecules, such as transcription factors, can induce or repress the transcriptional machinery, influencing the kinds and amounts of products that are produced. The regulatory DNA sequences are called cis-acting because they influence expression of genes on the same chromosome as the regulatory sequence (see p. 439). The regulatory molecules are called trans-acting because they can diffuse (transit) through the cell from their site of synthesis to their DNA-binding sites (Fig. 33.2). For example, a protein transcription factor (a trans-acting molecule) that regulates a gene on chromosome 6 might itself have been produced from a gene on chromosome 11. The binding of proteins to DNA is through structural motifs such as the zinc finger (Fig. 33.3),
Biochemistry_Lippinco. Regulation of transcription, the initial step in all gene expression, is controlled by regulatory sequences of DNA that are usually embedded in the noncoding regions of the genome. The interaction between these DNA sequences and regulatory molecules, such as transcription factors, can induce or repress the transcriptional machinery, influencing the kinds and amounts of products that are produced. The regulatory DNA sequences are called cis-acting because they influence expression of genes on the same chromosome as the regulatory sequence (see p. 439). The regulatory molecules are called trans-acting because they can diffuse (transit) through the cell from their site of synthesis to their DNA-binding sites (Fig. 33.2). For example, a protein transcription factor (a trans-acting molecule) that regulates a gene on chromosome 6 might itself have been produced from a gene on chromosome 11. The binding of proteins to DNA is through structural motifs such as the zinc finger (Fig. 33.3),
Biochemistry_Lippincott_1626
Biochemistry_Lippinco
that regulates a gene on chromosome 6 might itself have been produced from a gene on chromosome 11. The binding of proteins to DNA is through structural motifs such as the zinc finger (Fig. 33.3), leucine zipper, or helix-turn-helix in the protein.
Biochemistry_Lippinco. that regulates a gene on chromosome 6 might itself have been produced from a gene on chromosome 11. The binding of proteins to DNA is through structural motifs such as the zinc finger (Fig. 33.3), leucine zipper, or helix-turn-helix in the protein.
Biochemistry_Lippincott_1627
Biochemistry_Lippinco
III. REGULATION OF PROKARYOTIC GENE EXPRESSION In prokaryotes such as the bacterium Escherichia coli (E. coli), regulation of gene expression occurs primarily at the level of transcription and, in general, is mediated by the binding of trans-acting proteins to cis-acting regulatory elements on their single DNA molecule (chromosome). [Note: Regulating the first step in the expression of a gene is an efficient approach, insofar as energy is not wasted making unneeded gene products.] Transcriptional control in prokaryotes can involve the initiation or premature termination of transcription. A. Messenger RNA transcription from bacterial operons
Biochemistry_Lippinco. III. REGULATION OF PROKARYOTIC GENE EXPRESSION In prokaryotes such as the bacterium Escherichia coli (E. coli), regulation of gene expression occurs primarily at the level of transcription and, in general, is mediated by the binding of trans-acting proteins to cis-acting regulatory elements on their single DNA molecule (chromosome). [Note: Regulating the first step in the expression of a gene is an efficient approach, insofar as energy is not wasted making unneeded gene products.] Transcriptional control in prokaryotes can involve the initiation or premature termination of transcription. A. Messenger RNA transcription from bacterial operons
Biochemistry_Lippincott_1628
Biochemistry_Lippinco
A. Messenger RNA transcription from bacterial operons In bacteria, the structural genes that encode proteins involved in a particular metabolic pathway are often found sequentially grouped on the chromosome along with the cis-acting elements that regulate the transcription of these genes. The transcription product is a single polycistronic messenger RNA ([mRNA] see p. 434). The genes are, thus, coordinately regulated (that is, turned on or off as a unit). This entire package is referred to as an operon. B. Operators in bacterial operons
Biochemistry_Lippinco. A. Messenger RNA transcription from bacterial operons In bacteria, the structural genes that encode proteins involved in a particular metabolic pathway are often found sequentially grouped on the chromosome along with the cis-acting elements that regulate the transcription of these genes. The transcription product is a single polycistronic messenger RNA ([mRNA] see p. 434). The genes are, thus, coordinately regulated (that is, turned on or off as a unit). This entire package is referred to as an operon. B. Operators in bacterial operons
Biochemistry_Lippincott_1629
Biochemistry_Lippinco
B. Operators in bacterial operons Bacterial operons contain an operator, a segment of DNA that regulates the activity of the structural genes of the operon by reversibly binding a protein known as the repressor. If the operator is not bound by the repressor, RNA polymerase (RNA pol) binds the promoter, passes over the operator, and reaches the protein-coding genes that it transcribes to mRNA. If the repressor is bound to the operator, the polymerase is blocked and does not produce mRNA. As long as the repressor is bound to the operator, no mRNA (and, therefore, no proteins) are made. However, when an inducer molecule is present, it binds to the repressor, causing the repressor to change shape so that it no longer binds the operator. When this happens, RNA pol can initiate transcription. One of the best-understood examples is the inducible lactose (lac) operon of E. coli that illustrates both positive and negative regulation (Fig. 33.4). C. Lactose operon
Biochemistry_Lippinco. B. Operators in bacterial operons Bacterial operons contain an operator, a segment of DNA that regulates the activity of the structural genes of the operon by reversibly binding a protein known as the repressor. If the operator is not bound by the repressor, RNA polymerase (RNA pol) binds the promoter, passes over the operator, and reaches the protein-coding genes that it transcribes to mRNA. If the repressor is bound to the operator, the polymerase is blocked and does not produce mRNA. As long as the repressor is bound to the operator, no mRNA (and, therefore, no proteins) are made. However, when an inducer molecule is present, it binds to the repressor, causing the repressor to change shape so that it no longer binds the operator. When this happens, RNA pol can initiate transcription. One of the best-understood examples is the inducible lactose (lac) operon of E. coli that illustrates both positive and negative regulation (Fig. 33.4). C. Lactose operon
Biochemistry_Lippincott_1630
Biochemistry_Lippinco
The lac operon contains the genes that code for three proteins involved in the catabolism of the disaccharide lactose: the lacZ gene codes for βgalactosidase, which hydrolyzes lactose to galactose and glucose; the lacY gene codes for a permease, which facilitates the movement of lactose into the cell; and the lacA gene codes for thiogalactoside transacetylase, which acetylates lactose. [Note: The physiologic function of this acetylation is unknown.] All of these proteins are maximally produced only when lactose is available to the cell and glucose is not. [Note: Bacteria use glucose, if available, as a fuel in preference to any other sugar.] The regulatory portion of the operon is upstream of the three structural genes and consists of the promoter region where RNA pol binds and two additional sites, the operator (O) and the catabolite activator protein (CAP) sites, where regulatory proteins bind. The lacZ, lacY, and lacA genes are maximally expressed only when the O site is empty and
Biochemistry_Lippinco. The lac operon contains the genes that code for three proteins involved in the catabolism of the disaccharide lactose: the lacZ gene codes for βgalactosidase, which hydrolyzes lactose to galactose and glucose; the lacY gene codes for a permease, which facilitates the movement of lactose into the cell; and the lacA gene codes for thiogalactoside transacetylase, which acetylates lactose. [Note: The physiologic function of this acetylation is unknown.] All of these proteins are maximally produced only when lactose is available to the cell and glucose is not. [Note: Bacteria use glucose, if available, as a fuel in preference to any other sugar.] The regulatory portion of the operon is upstream of the three structural genes and consists of the promoter region where RNA pol binds and two additional sites, the operator (O) and the catabolite activator protein (CAP) sites, where regulatory proteins bind. The lacZ, lacY, and lacA genes are maximally expressed only when the O site is empty and
Biochemistry_Lippincott_1631
Biochemistry_Lippinco
sites, the operator (O) and the catabolite activator protein (CAP) sites, where regulatory proteins bind. The lacZ, lacY, and lacA genes are maximally expressed only when the O site is empty and the CAP site is bound by a complex of cyclic adenosine monophosphate ([cAMP] see p. 94) and the CAP, sometimes called the cAMP regulatory protein (CRP). A regulatory gene, the lacI gene, codes for the repressor protein (a trans-acting factor) that binds to the O site with high affinity. [Note: The lacI gene has its own promoter and is not part of the lac operon.] 1.
Biochemistry_Lippinco. sites, the operator (O) and the catabolite activator protein (CAP) sites, where regulatory proteins bind. The lacZ, lacY, and lacA genes are maximally expressed only when the O site is empty and the CAP site is bound by a complex of cyclic adenosine monophosphate ([cAMP] see p. 94) and the CAP, sometimes called the cAMP regulatory protein (CRP). A regulatory gene, the lacI gene, codes for the repressor protein (a trans-acting factor) that binds to the O site with high affinity. [Note: The lacI gene has its own promoter and is not part of the lac operon.] 1.
Biochemistry_Lippincott_1632
Biochemistry_Lippinco
When only glucose is available: In this case, the lac operon is repressed (turned off). Repression is mediated by the repressor protein binding via a helix-turn-helix motif (Fig. 33.5) to the O site, which is downstream of the promoter (see Fig. 33.4A). Binding of the repressor interferes with the binding of RNA pol to the promoter, thereby inhibiting transcription of the structural genes. This is an example of negative regulation. 2.
Biochemistry_Lippinco. When only glucose is available: In this case, the lac operon is repressed (turned off). Repression is mediated by the repressor protein binding via a helix-turn-helix motif (Fig. 33.5) to the O site, which is downstream of the promoter (see Fig. 33.4A). Binding of the repressor interferes with the binding of RNA pol to the promoter, thereby inhibiting transcription of the structural genes. This is an example of negative regulation. 2.
Biochemistry_Lippincott_1633
Biochemistry_Lippinco
When only lactose is available: In this case, the lac operon is induced (maximally expressed, or turned on). A small amount of lactose is converted to an isomer, allolactose. This compound is an inducer that binds to the repressor protein, changing its conformation so that it can no longer bind to the O site. In the absence of glucose, adenylyl cyclase is active, and cAMP is made and binds to the CAP. The cAMP–CAP transacting complex binds to the CAP site, causing RNA pol to initiate transcription with high efficiency at the promoter site (see Fig. 33.4B). This is an example of positive regulation. The transcript is a single polycistronic mRNA molecule that contains three sets of start and stop codons. Translation of the mRNA produces the three proteins that allow lactose to be used for energy production by the cell. [Note: In contrast to the inducible lacZ, lacY, and lacA genes, whose expression is regulated, the lacI gene is constitutive. Its gene product, the repressor protein, is
Biochemistry_Lippinco. When only lactose is available: In this case, the lac operon is induced (maximally expressed, or turned on). A small amount of lactose is converted to an isomer, allolactose. This compound is an inducer that binds to the repressor protein, changing its conformation so that it can no longer bind to the O site. In the absence of glucose, adenylyl cyclase is active, and cAMP is made and binds to the CAP. The cAMP–CAP transacting complex binds to the CAP site, causing RNA pol to initiate transcription with high efficiency at the promoter site (see Fig. 33.4B). This is an example of positive regulation. The transcript is a single polycistronic mRNA molecule that contains three sets of start and stop codons. Translation of the mRNA produces the three proteins that allow lactose to be used for energy production by the cell. [Note: In contrast to the inducible lacZ, lacY, and lacA genes, whose expression is regulated, the lacI gene is constitutive. Its gene product, the repressor protein, is
Biochemistry_Lippincott_1634
Biochemistry_Lippinco
production by the cell. [Note: In contrast to the inducible lacZ, lacY, and lacA genes, whose expression is regulated, the lacI gene is constitutive. Its gene product, the repressor protein, is always made and is active unless the inducer is present.] 3.
Biochemistry_Lippinco. production by the cell. [Note: In contrast to the inducible lacZ, lacY, and lacA genes, whose expression is regulated, the lacI gene is constitutive. Its gene product, the repressor protein, is always made and is active unless the inducer is present.] 3.
Biochemistry_Lippincott_1635
Biochemistry_Lippinco
When both glucose and lactose are available: In this case, the lac operon is uninduced, and transcription is negligible, even if lactose is present at a high concentration. Adenylyl cyclase is inhibited in the presence of glucose (a process known as catabolite repression) so no cAMP–CAP complex forms, and the CAP site remains empty. Therefore, the RNA pol is unable to effectively initiate transcription, even though the repressor is not bound to the O site. Consequently, the three structural genes of the operon are expressed only at a very low (basal) level (see Fig. 33.4C). [Note: Induction causes a 50-fold enhancement over basal expression.] D. Tryptophan operon
Biochemistry_Lippinco. When both glucose and lactose are available: In this case, the lac operon is uninduced, and transcription is negligible, even if lactose is present at a high concentration. Adenylyl cyclase is inhibited in the presence of glucose (a process known as catabolite repression) so no cAMP–CAP complex forms, and the CAP site remains empty. Therefore, the RNA pol is unable to effectively initiate transcription, even though the repressor is not bound to the O site. Consequently, the three structural genes of the operon are expressed only at a very low (basal) level (see Fig. 33.4C). [Note: Induction causes a 50-fold enhancement over basal expression.] D. Tryptophan operon
Biochemistry_Lippincott_1636
Biochemistry_Lippinco
The tryptophan (trp) operon contains five structural genes that code for enzymes required for the synthesis of the amino acid tryptophan. As with the lac operon, the trp operon is subject to negative control. However, for the repressible trp operon, negative control includes Trp itself binding to a repressor protein and facilitating the binding of the repressor to the operator: Trp is a corepressor. Because repression by Trp is not always complete, the trp operon, unlike the lac operon, is also regulated by a process known as attenuation. With attenuation, transcription is initiated but is terminated well before completion (Fig. 33.6). If Trp is plentiful, transcription initiation that escaped repression by Trp is attenuated (stopped) by the formation of an attenuator, a hairpin (stem-loop) structure in the mRNA similar to that seen in rho-independent termination (see p. 437). [Note: Because transcription and translation are temporally linked in prokaryotes (see p. 454), attenuation
Biochemistry_Lippinco. The tryptophan (trp) operon contains five structural genes that code for enzymes required for the synthesis of the amino acid tryptophan. As with the lac operon, the trp operon is subject to negative control. However, for the repressible trp operon, negative control includes Trp itself binding to a repressor protein and facilitating the binding of the repressor to the operator: Trp is a corepressor. Because repression by Trp is not always complete, the trp operon, unlike the lac operon, is also regulated by a process known as attenuation. With attenuation, transcription is initiated but is terminated well before completion (Fig. 33.6). If Trp is plentiful, transcription initiation that escaped repression by Trp is attenuated (stopped) by the formation of an attenuator, a hairpin (stem-loop) structure in the mRNA similar to that seen in rho-independent termination (see p. 437). [Note: Because transcription and translation are temporally linked in prokaryotes (see p. 454), attenuation
Biochemistry_Lippincott_1637
Biochemistry_Lippinco
structure in the mRNA similar to that seen in rho-independent termination (see p. 437). [Note: Because transcription and translation are temporally linked in prokaryotes (see p. 454), attenuation also results in the formation of a truncated, nonfunctional peptide product that is rapidly degraded.] If Trp becomes scarce, the operon is expressed. The 5′-end of the mRNA contains two adjacent codons for Trp. The lack of Trp causes ribosomes to stall at these codons, covering regions of the mRNA required for formation of the attenuation hairpin. This prevents attenuation and allows transcription to continue.
Biochemistry_Lippinco. structure in the mRNA similar to that seen in rho-independent termination (see p. 437). [Note: Because transcription and translation are temporally linked in prokaryotes (see p. 454), attenuation also results in the formation of a truncated, nonfunctional peptide product that is rapidly degraded.] If Trp becomes scarce, the operon is expressed. The 5′-end of the mRNA contains two adjacent codons for Trp. The lack of Trp causes ribosomes to stall at these codons, covering regions of the mRNA required for formation of the attenuation hairpin. This prevents attenuation and allows transcription to continue.
Biochemistry_Lippincott_1638
Biochemistry_Lippinco
Transcriptional attenuation can occur in prokaryotes because translation of an mRNA begins before its synthesis is complete. This does not occur in eukaryotes because the presence of a membrane-bound nucleus spatially and temporally separates transcription and translation. E. Coordination of transcription and translation Although transcriptional regulation of mRNA production is primary in bacteria, regulation of ribosomal RNA (rRNA) and protein synthesis plays important roles in adaptation to environmental stress. 1.
Biochemistry_Lippinco. Transcriptional attenuation can occur in prokaryotes because translation of an mRNA begins before its synthesis is complete. This does not occur in eukaryotes because the presence of a membrane-bound nucleus spatially and temporally separates transcription and translation. E. Coordination of transcription and translation Although transcriptional regulation of mRNA production is primary in bacteria, regulation of ribosomal RNA (rRNA) and protein synthesis plays important roles in adaptation to environmental stress. 1.
Biochemistry_Lippincott_1639
Biochemistry_Lippinco
1. Stringent response: E. coli has seven operons that synthesize the rRNA needed for ribosome assembly, and each is regulated in response to changes in environmental conditions. Regulation in response to amino acid starvation is known as the stringent response. The binding of an uncharged transfer RNA (tRNA) to the A site of a ribosome (see p. 452) triggers a series of events that leads to the production of the alarmone guanosine tetraphosphate (ppGpp). The synthesis of this unusual derivative of guanosine diphosphate (GDP) is catalyzed by stringent factor (RelA), an enzyme physically associated with ribosomes. Elevated levels of ppGpp result in inhibition of rRNA synthesis (Fig. 33.7). [Note: In addition to rRNA synthesis, tRNA synthesis and some mRNA synthesis (for example, for ribosomal proteins) are also inhibited. However, synthesis of mRNA for enzymes required for amino acid biosynthesis is not inhibited. ppGpp binds RNA pol and alters promoter 2.
Biochemistry_Lippinco. 1. Stringent response: E. coli has seven operons that synthesize the rRNA needed for ribosome assembly, and each is regulated in response to changes in environmental conditions. Regulation in response to amino acid starvation is known as the stringent response. The binding of an uncharged transfer RNA (tRNA) to the A site of a ribosome (see p. 452) triggers a series of events that leads to the production of the alarmone guanosine tetraphosphate (ppGpp). The synthesis of this unusual derivative of guanosine diphosphate (GDP) is catalyzed by stringent factor (RelA), an enzyme physically associated with ribosomes. Elevated levels of ppGpp result in inhibition of rRNA synthesis (Fig. 33.7). [Note: In addition to rRNA synthesis, tRNA synthesis and some mRNA synthesis (for example, for ribosomal proteins) are also inhibited. However, synthesis of mRNA for enzymes required for amino acid biosynthesis is not inhibited. ppGpp binds RNA pol and alters promoter 2.
Biochemistry_Lippincott_1640
Biochemistry_Lippinco
Regulatory ribosomal proteins: Operons for ribosomal proteins (rproteins) can be inhibited by an excess of their own protein products. For each operon, one specific r-protein functions in the repression of selection through use of different sigma factors for the polymerase (see p. 435).] translation of the polycistronic mRNA from that operon (Fig. 33.8). The r-protein does so by binding to the Shine-Dalgarno (SD) sequence located on the mRNA just upstream of the first initiating AUG codon (see p. 448) and acting as a physical impediment to the binding of the small ribosomal subunit to the SD sequence. Thus, one r-protein inhibits synthesis of all the r-proteins of the operon. This same r-protein also binds to rRNA and with a higher affinity than for mRNA. If the concentration of rRNA falls, the r-protein then is available to bind its own mRNA and inhibit its translation. This coordinated regulation keeps the synthesis of r-proteins in balance with the transcription of rRNA, so that
Biochemistry_Lippinco. Regulatory ribosomal proteins: Operons for ribosomal proteins (rproteins) can be inhibited by an excess of their own protein products. For each operon, one specific r-protein functions in the repression of selection through use of different sigma factors for the polymerase (see p. 435).] translation of the polycistronic mRNA from that operon (Fig. 33.8). The r-protein does so by binding to the Shine-Dalgarno (SD) sequence located on the mRNA just upstream of the first initiating AUG codon (see p. 448) and acting as a physical impediment to the binding of the small ribosomal subunit to the SD sequence. Thus, one r-protein inhibits synthesis of all the r-proteins of the operon. This same r-protein also binds to rRNA and with a higher affinity than for mRNA. If the concentration of rRNA falls, the r-protein then is available to bind its own mRNA and inhibit its translation. This coordinated regulation keeps the synthesis of r-proteins in balance with the transcription of rRNA, so that
Biochemistry_Lippincott_1641
Biochemistry_Lippinco
the r-protein then is available to bind its own mRNA and inhibit its translation. This coordinated regulation keeps the synthesis of r-proteins in balance with the transcription of rRNA, so that each is present in appropriate amounts for the formation of ribosomes.
Biochemistry_Lippinco. the r-protein then is available to bind its own mRNA and inhibit its translation. This coordinated regulation keeps the synthesis of r-proteins in balance with the transcription of rRNA, so that each is present in appropriate amounts for the formation of ribosomes.
Biochemistry_Lippincott_1642
Biochemistry_Lippinco
IV. REGULATION OF EUKARYOTIC GENE EXPRESSION The higher degree of complexity of eukaryotic genomes, as well as the presence of a nuclear membrane, necessitates a wider range of regulatory processes. As with the prokaryotes, transcription is the primary site of regulation. Again, the theme of trans-acting factors binding to cis-acting elements is seen. Operons, however, are not found in eukaryotes, which must use alternate strategies to solve the problem of how to coordinately regulate all the genes required for a specific response. In eukaryotes, gene expression is also regulated at multiple levels other than transcription. For example, the major modes of posttranscriptional regulation at the mRNA level are alternative mRNA splicing and polyadenylation, control of mRNA stability, and control of translational efficiency. Additional regulation at the protein level occurs by mechanisms that modulate stability, processing, or targeting of the protein. A. Coordinate regulation
Biochemistry_Lippinco. IV. REGULATION OF EUKARYOTIC GENE EXPRESSION The higher degree of complexity of eukaryotic genomes, as well as the presence of a nuclear membrane, necessitates a wider range of regulatory processes. As with the prokaryotes, transcription is the primary site of regulation. Again, the theme of trans-acting factors binding to cis-acting elements is seen. Operons, however, are not found in eukaryotes, which must use alternate strategies to solve the problem of how to coordinately regulate all the genes required for a specific response. In eukaryotes, gene expression is also regulated at multiple levels other than transcription. For example, the major modes of posttranscriptional regulation at the mRNA level are alternative mRNA splicing and polyadenylation, control of mRNA stability, and control of translational efficiency. Additional regulation at the protein level occurs by mechanisms that modulate stability, processing, or targeting of the protein. A. Coordinate regulation
Biochemistry_Lippincott_1643
Biochemistry_Lippinco
The need to coordinately regulate a group of genes to cause a particular response is of key importance in organisms with more than one chromosome. An underlying theme occurs repeatedly: A trans-acting protein functions as a specific transcription factor (STF) that binds to a cisacting regulatory consensus sequence (see p. 415) on each of the genes in the group even if they are on different chromosomes. [Note: The STF has a DNA-binding domain (DBD) and a transcription activation domain (TAD). The TAD recruits coactivators, such as histone acetyltransferases (see p. 438), and the general transcription factors (see p. 439) that, along with RNA pol, are required for formation of the transcription initiation complex at the promoter. Although the TAD recruits a variety of proteins, the specific effect of any one of them is dependent upon the protein composition of the complex. This is known as combinatorial control.] Examples of coordinate regulation in eukaryotes include the galactose
Biochemistry_Lippinco. The need to coordinately regulate a group of genes to cause a particular response is of key importance in organisms with more than one chromosome. An underlying theme occurs repeatedly: A trans-acting protein functions as a specific transcription factor (STF) that binds to a cisacting regulatory consensus sequence (see p. 415) on each of the genes in the group even if they are on different chromosomes. [Note: The STF has a DNA-binding domain (DBD) and a transcription activation domain (TAD). The TAD recruits coactivators, such as histone acetyltransferases (see p. 438), and the general transcription factors (see p. 439) that, along with RNA pol, are required for formation of the transcription initiation complex at the promoter. Although the TAD recruits a variety of proteins, the specific effect of any one of them is dependent upon the protein composition of the complex. This is known as combinatorial control.] Examples of coordinate regulation in eukaryotes include the galactose
Biochemistry_Lippincott_1644
Biochemistry_Lippinco
effect of any one of them is dependent upon the protein composition of the complex. This is known as combinatorial control.] Examples of coordinate regulation in eukaryotes include the galactose circuit and the hormone response system.
Biochemistry_Lippinco. effect of any one of them is dependent upon the protein composition of the complex. This is known as combinatorial control.] Examples of coordinate regulation in eukaryotes include the galactose circuit and the hormone response system.
Biochemistry_Lippincott_1645
Biochemistry_Lippinco
1. Galactose circuit: This regulatory scheme allows for the use of galactose when glucose is not available. In yeast, a unicellular organism, the genes required to metabolize galactose are on different chromosomes. Coordinated expression is mediated by the protein Gal4 (Gal = galactose), a STF that binds to a short regulatory DNA sequence upstream of each of the genes. The sequence is called the upstream activating sequence Gal (UASGal). Binding of Gal4 to UASGal through zinc fingers in its DBD occurs in both the absence and presence of galactose. When the sugar is absent, the regulatory protein Gal80 binds Gal4 at its TAD, thereby inhibiting gene transcription (Fig. 33.9A). When present, galactose activates the Gal3 protein. Gal3 binds Gal80, thereby allowing Gal4 to activate transcription (Fig. 33.9B). [Note: Glucose prevents the use of galactose by inhibting expression of Gal4 protein.]
Biochemistry_Lippinco. 1. Galactose circuit: This regulatory scheme allows for the use of galactose when glucose is not available. In yeast, a unicellular organism, the genes required to metabolize galactose are on different chromosomes. Coordinated expression is mediated by the protein Gal4 (Gal = galactose), a STF that binds to a short regulatory DNA sequence upstream of each of the genes. The sequence is called the upstream activating sequence Gal (UASGal). Binding of Gal4 to UASGal through zinc fingers in its DBD occurs in both the absence and presence of galactose. When the sugar is absent, the regulatory protein Gal80 binds Gal4 at its TAD, thereby inhibiting gene transcription (Fig. 33.9A). When present, galactose activates the Gal3 protein. Gal3 binds Gal80, thereby allowing Gal4 to activate transcription (Fig. 33.9B). [Note: Glucose prevents the use of galactose by inhibting expression of Gal4 protein.]
Biochemistry_Lippincott_1646
Biochemistry_Lippinco
B. presence of galactose. [Note: Target genes, whether on the same or a different chromosome, each have an upstream activating sequence galactose (UASGal).] TAD = transcription activation domain; DBD = DNA-binding domain; mRNA = messenger RNA. 2. Hormone response system: Hormone response elements (HRE) are DNA sequences that bind trans-acting proteins and regulate gene expression in response to hormonal signals in multicellular organisms. Hormones bind to either intracellular (nuclear) receptors (for example, steroid hormones; see p. 240) or cell-surface receptors (for example, the peptide hormone glucagon; see p. 314). a.
Biochemistry_Lippinco. B. presence of galactose. [Note: Target genes, whether on the same or a different chromosome, each have an upstream activating sequence galactose (UASGal).] TAD = transcription activation domain; DBD = DNA-binding domain; mRNA = messenger RNA. 2. Hormone response system: Hormone response elements (HRE) are DNA sequences that bind trans-acting proteins and regulate gene expression in response to hormonal signals in multicellular organisms. Hormones bind to either intracellular (nuclear) receptors (for example, steroid hormones; see p. 240) or cell-surface receptors (for example, the peptide hormone glucagon; see p. 314). a.
Biochemistry_Lippincott_1647
Biochemistry_Lippinco
Intracellular receptors: Members of the nuclear receptor superfamily, which includes the steroid hormone (glucocorticoids, mineralocorticoids, androgens, and estrogens), vitamin D, retinoic acid, and thyroid hormone receptors, function as STF. In addition to domains for DNA-binding and transcriptional activation, these receptors also contain a ligand-binding domain. For example, the steroid hormone cortisol (a glucocorticoid) binds intracellular receptors at the ligand-binding domain (Fig. 33.10). Binding causes a conformational change in the receptor that activates it. The receptor– hormone complex enters the nucleus, dimerizes, and binds via a zinc finger motif to DNA at a regulatory element, the glucocorticoid response element (GRE) that is an example of a HRE. Binding allows recruitment of coactivators to the TAD and results in expression of cortisol-responsive genes, each of which is under the control of its own GRE. Binding of the receptor–hormone complex to the GRE allows
Biochemistry_Lippinco. Intracellular receptors: Members of the nuclear receptor superfamily, which includes the steroid hormone (glucocorticoids, mineralocorticoids, androgens, and estrogens), vitamin D, retinoic acid, and thyroid hormone receptors, function as STF. In addition to domains for DNA-binding and transcriptional activation, these receptors also contain a ligand-binding domain. For example, the steroid hormone cortisol (a glucocorticoid) binds intracellular receptors at the ligand-binding domain (Fig. 33.10). Binding causes a conformational change in the receptor that activates it. The receptor– hormone complex enters the nucleus, dimerizes, and binds via a zinc finger motif to DNA at a regulatory element, the glucocorticoid response element (GRE) that is an example of a HRE. Binding allows recruitment of coactivators to the TAD and results in expression of cortisol-responsive genes, each of which is under the control of its own GRE. Binding of the receptor–hormone complex to the GRE allows
Biochemistry_Lippincott_1648
Biochemistry_Lippinco
of coactivators to the TAD and results in expression of cortisol-responsive genes, each of which is under the control of its own GRE. Binding of the receptor–hormone complex to the GRE allows coordinate expression of a group of target genes, even though these genes are on different chromosomes. The GRE can be located upstream or downstream of the genes it regulates and at great distances from them. The GRE, then, can function as a true enhancer (see p. 440). [Note: If associated with repressors, hormone–receptor complexes inhibit transcription.] b.
Biochemistry_Lippinco. of coactivators to the TAD and results in expression of cortisol-responsive genes, each of which is under the control of its own GRE. Binding of the receptor–hormone complex to the GRE allows coordinate expression of a group of target genes, even though these genes are on different chromosomes. The GRE can be located upstream or downstream of the genes it regulates and at great distances from them. The GRE, then, can function as a true enhancer (see p. 440). [Note: If associated with repressors, hormone–receptor complexes inhibit transcription.] b.
Biochemistry_Lippincott_1649
Biochemistry_Lippinco
Cell-surface receptors: These receptors include those for insulin, epinephrine, and glucagon. Glucagon, for example, is a peptide hormone that binds its G protein–coupled plasma membrane receptor on glucagon-responsive cells. This extracellular signal is then transduced to intracellular cAMP, a second messenger (Fig. 33.11; also see Fig. 8.7 on p. 95), which can affect protein expression (and activity) through protein kinase A–mediated phosphorylation. In response to a rise in cAMP, a trans-acting factor (cAMP response element–binding [CREB] protein) is phosphorylated and activated. Active CREB protein binds via a leucine zipper motif to a cis-acting regulatory element, the cAMP response element (CRE), resulting in transcription of target genes with CRE in their promoters. [Note: The genes for phosphoenolpyruvate carboxykinase and glucose 6phosphatase, key enzymes of gluconeogenesis (see p. 122), are examples of genes upregulated by the cAMP/CRE/CREB system.]
Biochemistry_Lippinco. Cell-surface receptors: These receptors include those for insulin, epinephrine, and glucagon. Glucagon, for example, is a peptide hormone that binds its G protein–coupled plasma membrane receptor on glucagon-responsive cells. This extracellular signal is then transduced to intracellular cAMP, a second messenger (Fig. 33.11; also see Fig. 8.7 on p. 95), which can affect protein expression (and activity) through protein kinase A–mediated phosphorylation. In response to a rise in cAMP, a trans-acting factor (cAMP response element–binding [CREB] protein) is phosphorylated and activated. Active CREB protein binds via a leucine zipper motif to a cis-acting regulatory element, the cAMP response element (CRE), resulting in transcription of target genes with CRE in their promoters. [Note: The genes for phosphoenolpyruvate carboxykinase and glucose 6phosphatase, key enzymes of gluconeogenesis (see p. 122), are examples of genes upregulated by the cAMP/CRE/CREB system.]
Biochemistry_Lippincott_1650
Biochemistry_Lippinco
B. Messenger RNA processing and use Eukaryotic mRNA undergoes several processing events before it is exported from the nucleus to the cytoplasm for use in protein synthesis. Capping at the 5′-end (see p. 441), polyadenylation at the 3′-end (see p. 442), and splicing (see p. 442) are essential for the production of a functional eukaryotic messenger from most pre-mRNA. Variations in splicing and polyadenylation can affect gene expression. In addition, messenger stability also affects gene expression. 1. Alternative splicing: Tissue-specific protein isoforms can be made from the same pre-mRNA through alternative splicing, which can involve exon skipping (loss), intron retention, and use of alternative splice-donor or -acceptor sites (Fig. 33.12). For example, the pre-mRNA for tropomyosin (TM) undergoes tissue-specific alternative splicing to yield a number of TM isoforms (see p. 443). [Note: Over 90% of all human genes undergo alternative splicing.] 2.
Biochemistry_Lippinco. B. Messenger RNA processing and use Eukaryotic mRNA undergoes several processing events before it is exported from the nucleus to the cytoplasm for use in protein synthesis. Capping at the 5′-end (see p. 441), polyadenylation at the 3′-end (see p. 442), and splicing (see p. 442) are essential for the production of a functional eukaryotic messenger from most pre-mRNA. Variations in splicing and polyadenylation can affect gene expression. In addition, messenger stability also affects gene expression. 1. Alternative splicing: Tissue-specific protein isoforms can be made from the same pre-mRNA through alternative splicing, which can involve exon skipping (loss), intron retention, and use of alternative splice-donor or -acceptor sites (Fig. 33.12). For example, the pre-mRNA for tropomyosin (TM) undergoes tissue-specific alternative splicing to yield a number of TM isoforms (see p. 443). [Note: Over 90% of all human genes undergo alternative splicing.] 2.
Biochemistry_Lippincott_1651
Biochemistry_Lippinco
Alternative polyadenylation: Some pre-mRNA transcripts have more than one site for cleavage and polyadenylation. Alternative polyadenylation (APA) generates mRNA with different 3′-ends, altering the untranslated region (UTR) or the coding (translated) sequence. [Note: APA is involved in the production of the membrane-bound and secreted forms of immunoglobulin M.] The use of alternative splicing and polyadenylation sites, as well as alternative transcription start sites explains, at least in part, how the ~20,000 to 25,000 genes in the human genome can give rise to well over 100,000 proteins. 3. Messenger RNA editing: Even after mRNA has been fully processed, it may undergo an additional posttranscriptional modification in which a base in the mRNA is altered. This is known as RNA editing. An important example in humans occurs with the transcript for apolipoprotein (apo) B, an essential component of chylomicrons (see p.
Biochemistry_Lippinco. Alternative polyadenylation: Some pre-mRNA transcripts have more than one site for cleavage and polyadenylation. Alternative polyadenylation (APA) generates mRNA with different 3′-ends, altering the untranslated region (UTR) or the coding (translated) sequence. [Note: APA is involved in the production of the membrane-bound and secreted forms of immunoglobulin M.] The use of alternative splicing and polyadenylation sites, as well as alternative transcription start sites explains, at least in part, how the ~20,000 to 25,000 genes in the human genome can give rise to well over 100,000 proteins. 3. Messenger RNA editing: Even after mRNA has been fully processed, it may undergo an additional posttranscriptional modification in which a base in the mRNA is altered. This is known as RNA editing. An important example in humans occurs with the transcript for apolipoprotein (apo) B, an essential component of chylomicrons (see p.
Biochemistry_Lippincott_1652
Biochemistry_Lippinco
228) and very-low-density lipoproteins ([VLDL] see p. 230). Apo B mRNA is made in the liver and the small intestine. However, in the intestine only, the cytosine (C) base in the CAA codon for glutamine is enzymatically deaminated to uracil (U), changing the sense codon to the nonsense or stop codon UAA, as shown in Figure 33.13. This results in a shorter protein (apo B-48, representing 48% of the message) being made in the intestine (and incorporated into chylomicrons) than is made in the liver (apo B-100, full-length, incorporated into VLDL). U = uracil. 4. Messenger RNA stability: How long an mRNA remains in the cytosol before it is degraded influences how much protein product can be produced from it. Regulation of iron metabolism and the gene-silencing process of RNA interference (RNAi) illustrate the importance of mRNA stability in the regulation of gene expression.
Biochemistry_Lippinco. 228) and very-low-density lipoproteins ([VLDL] see p. 230). Apo B mRNA is made in the liver and the small intestine. However, in the intestine only, the cytosine (C) base in the CAA codon for glutamine is enzymatically deaminated to uracil (U), changing the sense codon to the nonsense or stop codon UAA, as shown in Figure 33.13. This results in a shorter protein (apo B-48, representing 48% of the message) being made in the intestine (and incorporated into chylomicrons) than is made in the liver (apo B-100, full-length, incorporated into VLDL). U = uracil. 4. Messenger RNA stability: How long an mRNA remains in the cytosol before it is degraded influences how much protein product can be produced from it. Regulation of iron metabolism and the gene-silencing process of RNA interference (RNAi) illustrate the importance of mRNA stability in the regulation of gene expression.
Biochemistry_Lippincott_1653
Biochemistry_Lippinco
a. Iron metabolism: Transferrin (Tf) is a plasma protein that transports iron. Tf binds to cell-surface receptors (transferrin receptors [TfR]) that get internalized and provide cells, such as erythroblasts, with iron. The mRNA for the TfR has several cis-acting iron-responsive elements (IRE) in its 3′-UTR. IRE have a short stem-loop structure that can be bound by trans-acting iron regulatory proteins (IRP), as shown in Figure 33.14. When the iron concentration in the cell is low, the IRP bind to the 3′-IRE and stabilize the mRNA for TfR, allowing TfR synthesis. When intracellular iron levels are high, the IRP dissociate. The lack of IRP bound to the mRNA hastens its destruction, resulting in decreased TfR synthesis. [Note: The mRNA for ferritin, an intracellular protein of iron storage, has a single IRE in its 5′-UTR. When iron levels in the cell are low, IRP bind the 5′-IRE and prevent the use of the mRNA, and less ferritin is made. When iron accumulates in the cell, the IRP
Biochemistry_Lippinco. a. Iron metabolism: Transferrin (Tf) is a plasma protein that transports iron. Tf binds to cell-surface receptors (transferrin receptors [TfR]) that get internalized and provide cells, such as erythroblasts, with iron. The mRNA for the TfR has several cis-acting iron-responsive elements (IRE) in its 3′-UTR. IRE have a short stem-loop structure that can be bound by trans-acting iron regulatory proteins (IRP), as shown in Figure 33.14. When the iron concentration in the cell is low, the IRP bind to the 3′-IRE and stabilize the mRNA for TfR, allowing TfR synthesis. When intracellular iron levels are high, the IRP dissociate. The lack of IRP bound to the mRNA hastens its destruction, resulting in decreased TfR synthesis. [Note: The mRNA for ferritin, an intracellular protein of iron storage, has a single IRE in its 5′-UTR. When iron levels in the cell are low, IRP bind the 5′-IRE and prevent the use of the mRNA, and less ferritin is made. When iron accumulates in the cell, the IRP
Biochemistry_Lippincott_1654
Biochemistry_Lippinco
has a single IRE in its 5′-UTR. When iron levels in the cell are low, IRP bind the 5′-IRE and prevent the use of the mRNA, and less ferritin is made. When iron accumulates in the cell, the IRP dissociate, allowing synthesis of ferritin molecules to store the excess iron. Aminolevulinic acid synthase 2, the regulated enzyme of heme synthesis (see p. 278) in erythroblasts, also contains a 5′-IRE.] (See Chapter 29 for a discussion of iron metabolism.) b. RNA interference: RNAi is a mechanism of gene silencing through decreased expression of mRNA, either by repression of translation or by increased degradation. It plays a key role in such fundamental processes as cell proliferation, differentiation, and apoptosis. RNAi is mediated by short (~22 nucleotides), noncoding RNA called microRNA (miRNA). The miRNA arise from far longer, genomically encoded nuclear transcripts, primary miRNA (pri-miRNA), that are partially processed in the nucleus to pre-miRNA by an endonuclease (Drosha) then
Biochemistry_Lippinco. has a single IRE in its 5′-UTR. When iron levels in the cell are low, IRP bind the 5′-IRE and prevent the use of the mRNA, and less ferritin is made. When iron accumulates in the cell, the IRP dissociate, allowing synthesis of ferritin molecules to store the excess iron. Aminolevulinic acid synthase 2, the regulated enzyme of heme synthesis (see p. 278) in erythroblasts, also contains a 5′-IRE.] (See Chapter 29 for a discussion of iron metabolism.) b. RNA interference: RNAi is a mechanism of gene silencing through decreased expression of mRNA, either by repression of translation or by increased degradation. It plays a key role in such fundamental processes as cell proliferation, differentiation, and apoptosis. RNAi is mediated by short (~22 nucleotides), noncoding RNA called microRNA (miRNA). The miRNA arise from far longer, genomically encoded nuclear transcripts, primary miRNA (pri-miRNA), that are partially processed in the nucleus to pre-miRNA by an endonuclease (Drosha) then
Biochemistry_Lippincott_1655
Biochemistry_Lippinco
(miRNA). The miRNA arise from far longer, genomically encoded nuclear transcripts, primary miRNA (pri-miRNA), that are partially processed in the nucleus to pre-miRNA by an endonuclease (Drosha) then transported to the cytoplasm. There, an endonuclease (Dicer) completes the processing and generates short, double-stranded miRNA. A single strand (the guide or antisense strand) of the miRNA associates with a cytosolic protein complex known as the RNA-induced silencing complex (RISC). The guide strand hybridizes with a complementary sequence in the 3′-UTR of a full-length target mRNA, bringing RISC to the mRNA. This can result in repression of translation of the mRNA or its degradation by an endonuclease (Argonaute/Ago/Slicer) of the RISC. The extent of complementarity appears to be the determining factor (Fig. 33.15). RNAi can also be triggered by the introduction of exogenous double-stranded short interfering RNA (siRNA) into a cell, a process that has enormous therapeutic potential.
Biochemistry_Lippinco. (miRNA). The miRNA arise from far longer, genomically encoded nuclear transcripts, primary miRNA (pri-miRNA), that are partially processed in the nucleus to pre-miRNA by an endonuclease (Drosha) then transported to the cytoplasm. There, an endonuclease (Dicer) completes the processing and generates short, double-stranded miRNA. A single strand (the guide or antisense strand) of the miRNA associates with a cytosolic protein complex known as the RNA-induced silencing complex (RISC). The guide strand hybridizes with a complementary sequence in the 3′-UTR of a full-length target mRNA, bringing RISC to the mRNA. This can result in repression of translation of the mRNA or its degradation by an endonuclease (Argonaute/Ago/Slicer) of the RISC. The extent of complementarity appears to be the determining factor (Fig. 33.15). RNAi can also be triggered by the introduction of exogenous double-stranded short interfering RNA (siRNA) into a cell, a process that has enormous therapeutic potential.
Biochemistry_Lippincott_1656
Biochemistry_Lippinco
The first clinical trial of RNAi-based therapy involved the neovascular form of age-related macular degeneration (AMD), which is triggered by overproduction of vascular endothelial growth factor (VEGF), leading to the sprouting of excess blood vessels behind the retina. The vessels leak, clouding and often entirely destroying vision (therefore, neovascular AMD is also referred to as wet AMD). An siRNA was designed to target the mRNA of VEGF and promote its degradation. Although considerable effort and resources have been expended to develop RNAi-based therapeutics, especially for the treatment of cancer, no products have gone from trials to the market. The research applications of RNAi, however, have grown rapidly.
Biochemistry_Lippinco. The first clinical trial of RNAi-based therapy involved the neovascular form of age-related macular degeneration (AMD), which is triggered by overproduction of vascular endothelial growth factor (VEGF), leading to the sprouting of excess blood vessels behind the retina. The vessels leak, clouding and often entirely destroying vision (therefore, neovascular AMD is also referred to as wet AMD). An siRNA was designed to target the mRNA of VEGF and promote its degradation. Although considerable effort and resources have been expended to develop RNAi-based therapeutics, especially for the treatment of cancer, no products have gone from trials to the market. The research applications of RNAi, however, have grown rapidly.
Biochemistry_Lippincott_1657
Biochemistry_Lippinco
5. Messenger RNA translation: Regulation of gene expression can also occur at the level of mRNA translation. One mechanism by which translation is regulated is through phosphorylation of the eukaryotic translation initiation factor, eIF-2 (Fig. 33.16). Phosphorylation of eIF-2 inhibits its function and so inhibits translation at the initiation step (see p. 459). [Note: Phosphorylation of eIF-2 prevents its reactivation by inhibiting GDP-GTP exchange.] Phosphorylation is catalyzed by kinases that are activated in response to environmental conditions, such as amino acid starvation, heme deficiency in erythroblasts, the presence of double-stranded RNA (signaling viral infection), and the accumulation of misfolded proteins in the rough endoplasmic reticulum (see p. 460). phosphate; = phosphate. C. Regulation through variations in DNA
Biochemistry_Lippinco. 5. Messenger RNA translation: Regulation of gene expression can also occur at the level of mRNA translation. One mechanism by which translation is regulated is through phosphorylation of the eukaryotic translation initiation factor, eIF-2 (Fig. 33.16). Phosphorylation of eIF-2 inhibits its function and so inhibits translation at the initiation step (see p. 459). [Note: Phosphorylation of eIF-2 prevents its reactivation by inhibiting GDP-GTP exchange.] Phosphorylation is catalyzed by kinases that are activated in response to environmental conditions, such as amino acid starvation, heme deficiency in erythroblasts, the presence of double-stranded RNA (signaling viral infection), and the accumulation of misfolded proteins in the rough endoplasmic reticulum (see p. 460). phosphate; = phosphate. C. Regulation through variations in DNA
Biochemistry_Lippincott_1658
Biochemistry_Lippinco
phosphate; = phosphate. C. Regulation through variations in DNA Gene expression in eukaryotes is also influenced by the accessibility of DNA to the transcriptional apparatus, the amount of DNA, and the arrangement of DNA. [Note: Localized transitions between the B and Z forms of DNA (see p. 414) can also affect gene expression.] 1.
Biochemistry_Lippinco. phosphate; = phosphate. C. Regulation through variations in DNA Gene expression in eukaryotes is also influenced by the accessibility of DNA to the transcriptional apparatus, the amount of DNA, and the arrangement of DNA. [Note: Localized transitions between the B and Z forms of DNA (see p. 414) can also affect gene expression.] 1.
Biochemistry_Lippincott_1659
Biochemistry_Lippinco
Access to DNA: In eukaryotes, DNA is found complexed with histone and nonhistone proteins to form chromatin (see p. 425). Transcriptionally active, decondensed chromatin (euchromatin) differs from the more condensed, inactive form (heterochromatin) in a number of ways. Active chromatin contains histone proteins that have been covalently modified at their amino terminal ends by reversible methylation, acetylation, or phosphorylation (see p. 438 for a discussion of histone acetylation/deacetylation by histone acetyltransferase and histone deacetylase). Such modifications decrease the positive charge of these basic proteins, thereby decreasing the strength of their association with negatively charged DNA. This relaxes the nucleosome (see p. 425), allowing transcription factors access to specific regions on the DNA. Nucleosomes can also be repositioned, an ATP-requiring process that is part of chromatin remodeling. Another difference between transcriptionally active and inactive chromatin
Biochemistry_Lippinco. Access to DNA: In eukaryotes, DNA is found complexed with histone and nonhistone proteins to form chromatin (see p. 425). Transcriptionally active, decondensed chromatin (euchromatin) differs from the more condensed, inactive form (heterochromatin) in a number of ways. Active chromatin contains histone proteins that have been covalently modified at their amino terminal ends by reversible methylation, acetylation, or phosphorylation (see p. 438 for a discussion of histone acetylation/deacetylation by histone acetyltransferase and histone deacetylase). Such modifications decrease the positive charge of these basic proteins, thereby decreasing the strength of their association with negatively charged DNA. This relaxes the nucleosome (see p. 425), allowing transcription factors access to specific regions on the DNA. Nucleosomes can also be repositioned, an ATP-requiring process that is part of chromatin remodeling. Another difference between transcriptionally active and inactive chromatin
Biochemistry_Lippincott_1660
Biochemistry_Lippinco
regions on the DNA. Nucleosomes can also be repositioned, an ATP-requiring process that is part of chromatin remodeling. Another difference between transcriptionally active and inactive chromatin is the extent of methylation of cytosine bases in CG-rich regions (CpG islands) in the promoter region of many genes. Methylation is by methyltransferases that use S-adenosylmethionine as the methyl donor (Fig. 33.17). Transcriptionally active genes are less methylated (hypomethylated) than their inactive counterparts, suggesting that DNA hypermethylation silences gene expression. Modification of histones and methylation of DNA are epigenetic in that they are heritable changes in DNA that alter gene expression without altering the base sequence.
Biochemistry_Lippinco. regions on the DNA. Nucleosomes can also be repositioned, an ATP-requiring process that is part of chromatin remodeling. Another difference between transcriptionally active and inactive chromatin is the extent of methylation of cytosine bases in CG-rich regions (CpG islands) in the promoter region of many genes. Methylation is by methyltransferases that use S-adenosylmethionine as the methyl donor (Fig. 33.17). Transcriptionally active genes are less methylated (hypomethylated) than their inactive counterparts, suggesting that DNA hypermethylation silences gene expression. Modification of histones and methylation of DNA are epigenetic in that they are heritable changes in DNA that alter gene expression without altering the base sequence.
Biochemistry_Lippincott_1661
Biochemistry_Lippinco
2. Amount of DNA: A change up or down in the number of copies of a gene can affect the amount of gene product produced. An increase in copy number (gene amplification) has contributed to increased genomic complexity and is still a normal developmental process in certain nonmammalian species. In mammals, however, gene amplification is seen with some diseases and in response to particular chemotherapeutic drugs such as methotrexate, an inhibitor of the enzyme dihydrofolate reductase (DHFR), required for the synthesis of thymidine triphosphate (TTP) in the pyrimidine biosynthetic pathway (see p. 303). TTP is essential for DNA synthesis. Gene amplification results in an increase in the number of DHFR genes and resistance to the drug, allowing TTP to be made. 3.
Biochemistry_Lippinco. 2. Amount of DNA: A change up or down in the number of copies of a gene can affect the amount of gene product produced. An increase in copy number (gene amplification) has contributed to increased genomic complexity and is still a normal developmental process in certain nonmammalian species. In mammals, however, gene amplification is seen with some diseases and in response to particular chemotherapeutic drugs such as methotrexate, an inhibitor of the enzyme dihydrofolate reductase (DHFR), required for the synthesis of thymidine triphosphate (TTP) in the pyrimidine biosynthetic pathway (see p. 303). TTP is essential for DNA synthesis. Gene amplification results in an increase in the number of DHFR genes and resistance to the drug, allowing TTP to be made. 3.
Biochemistry_Lippincott_1662
Biochemistry_Lippinco
3. Arrangement of DNA: The process by which immunoglobulins (antibodies) are produced by B lymphocytes involves permanent rearrangements of the DNA in these cells. The immunoglobulins (for example, IgG) consist of two light and two heavy chains, with each chain containing regions of variable and constant amino acid sequence. The variable region is the result of somatic recombination of segments within both the light-and the heavy-chain genes. During B-lymphocyte development, single variable (V), diversity (D), and joining (J) gene segments are brought together through gene rearrangement to form a unique variable region (Fig. 33.18). This process allows the generation of 109−1011 different immunoglobulins from a single gene, providing the diversity needed for the recognition of an enormous number of antigens. [Note: Pathologic DNA rearrangement is seen with translocation, a process by which two different chromosomes exchange DNA segments.] 4.
Biochemistry_Lippinco. 3. Arrangement of DNA: The process by which immunoglobulins (antibodies) are produced by B lymphocytes involves permanent rearrangements of the DNA in these cells. The immunoglobulins (for example, IgG) consist of two light and two heavy chains, with each chain containing regions of variable and constant amino acid sequence. The variable region is the result of somatic recombination of segments within both the light-and the heavy-chain genes. During B-lymphocyte development, single variable (V), diversity (D), and joining (J) gene segments are brought together through gene rearrangement to form a unique variable region (Fig. 33.18). This process allows the generation of 109−1011 different immunoglobulins from a single gene, providing the diversity needed for the recognition of an enormous number of antigens. [Note: Pathologic DNA rearrangement is seen with translocation, a process by which two different chromosomes exchange DNA segments.] 4.
Biochemistry_Lippincott_1663
Biochemistry_Lippinco
Mobile DNA elements: Transposons (Tn) are mobile segments of DNA that move in an essentially random manner from one site to another on the same or a different chromosome. Movement is mediated by transposase, an enzyme encoded by the Tn itself. Movement can be direct, in which transposase cuts out and then inserts the Tn at a new site, or replicative, in which the Tn is copied and the copy inserted elsewhere while the original remains in place. In eukaryotes, including humans, replicative transposition frequently involves an RNA intermediate made by a reverse transcriptase (see p. 424), in which case the Tn is called a retrotransposon. Transposition has contributed to structural variation in the genome but also has the potential to alter gene expression and even to cause disease. Tn comprise ~50% of the human genome, with retrotransposons accounting for 90% of Tn. Although the vast majority of these retrotransposons have lost the ability to move, some are still active. Their
Biochemistry_Lippinco. Mobile DNA elements: Transposons (Tn) are mobile segments of DNA that move in an essentially random manner from one site to another on the same or a different chromosome. Movement is mediated by transposase, an enzyme encoded by the Tn itself. Movement can be direct, in which transposase cuts out and then inserts the Tn at a new site, or replicative, in which the Tn is copied and the copy inserted elsewhere while the original remains in place. In eukaryotes, including humans, replicative transposition frequently involves an RNA intermediate made by a reverse transcriptase (see p. 424), in which case the Tn is called a retrotransposon. Transposition has contributed to structural variation in the genome but also has the potential to alter gene expression and even to cause disease. Tn comprise ~50% of the human genome, with retrotransposons accounting for 90% of Tn. Although the vast majority of these retrotransposons have lost the ability to move, some are still active. Their
Biochemistry_Lippincott_1664
Biochemistry_Lippinco
Tn comprise ~50% of the human genome, with retrotransposons accounting for 90% of Tn. Although the vast majority of these retrotransposons have lost the ability to move, some are still active. Their transposition is thought to be the basis for some rare cases of hemophilia A and Duchenne muscular dystrophy. [Note: The growing problem of antibiotic-resistant bacteria is a consequence, at least in part, of the exchange of plasmids among bacterial cells. If the plasmids contain Tn-carrying antibiotic resistance genes, the recipient bacteria gain resistance to one or more antimicrobial drugs.]
Biochemistry_Lippinco. Tn comprise ~50% of the human genome, with retrotransposons accounting for 90% of Tn. Although the vast majority of these retrotransposons have lost the ability to move, some are still active. Their transposition is thought to be the basis for some rare cases of hemophilia A and Duchenne muscular dystrophy. [Note: The growing problem of antibiotic-resistant bacteria is a consequence, at least in part, of the exchange of plasmids among bacterial cells. If the plasmids contain Tn-carrying antibiotic resistance genes, the recipient bacteria gain resistance to one or more antimicrobial drugs.]
Biochemistry_Lippincott_1665
Biochemistry_Lippinco
V. CHAPTER SUMMARY
Biochemistry_Lippinco. V. CHAPTER SUMMARY
Biochemistry_Lippincott_1666
Biochemistry_Lippinco
Gene expression results in the production of a functional gene product (either RNA or protein) through the processes of transcription and translation (Fig. 33.19). Genes can be either constitutive (always expressed, housekeeping genes) or regulated (expressed only under certain conditions in all cells or in a subset of cells). The ability to appropriately induce (positively regulate) or repress (negatively regulate) genes is essential in all organisms. Regulation of gene expression occurs primarily at transcription in both prokaryotes and eukaryotes and is mediated through trans-acting proteins binding to cis-acting regulatory DNA elements. In eukaryotes, regulation also occurs through DNA modifications and through posttranscriptional and posttranslational processing. In prokaryotes, such as Escherichia coli, the coordinate regulation of genes whose protein products are required for a particular process is achieved through operons (groups of genes sequentially arranged on the
Biochemistry_Lippinco. Gene expression results in the production of a functional gene product (either RNA or protein) through the processes of transcription and translation (Fig. 33.19). Genes can be either constitutive (always expressed, housekeeping genes) or regulated (expressed only under certain conditions in all cells or in a subset of cells). The ability to appropriately induce (positively regulate) or repress (negatively regulate) genes is essential in all organisms. Regulation of gene expression occurs primarily at transcription in both prokaryotes and eukaryotes and is mediated through trans-acting proteins binding to cis-acting regulatory DNA elements. In eukaryotes, regulation also occurs through DNA modifications and through posttranscriptional and posttranslational processing. In prokaryotes, such as Escherichia coli, the coordinate regulation of genes whose protein products are required for a particular process is achieved through operons (groups of genes sequentially arranged on the
Biochemistry_Lippincott_1667
Biochemistry_Lippinco
such as Escherichia coli, the coordinate regulation of genes whose protein products are required for a particular process is achieved through operons (groups of genes sequentially arranged on the chromosome along with the regulatory elements that determine their transcription). The lac operon contains the Z, Y, and A structural genes, the protein products of which are needed for the catabolism of lactose. It is subject to negative and positive regulation. When glucose is available, the operon is repressed by the binding of the repressor protein (the product of the lacI gene) to the operator, thus preventing transcription. When only lactose is present, the operon is induced by an isomer of lactose (allolactose) that binds the repressor protein, preventing it from binding to the operator. In addition, cyclic adenosine monophosphate (cAMP) binds the catabolite activator protein (CAP), and the complex binds the DNA at the CAP site. This increases promoter efficiency and results in the
Biochemistry_Lippinco. such as Escherichia coli, the coordinate regulation of genes whose protein products are required for a particular process is achieved through operons (groups of genes sequentially arranged on the chromosome along with the regulatory elements that determine their transcription). The lac operon contains the Z, Y, and A structural genes, the protein products of which are needed for the catabolism of lactose. It is subject to negative and positive regulation. When glucose is available, the operon is repressed by the binding of the repressor protein (the product of the lacI gene) to the operator, thus preventing transcription. When only lactose is present, the operon is induced by an isomer of lactose (allolactose) that binds the repressor protein, preventing it from binding to the operator. In addition, cyclic adenosine monophosphate (cAMP) binds the catabolite activator protein (CAP), and the complex binds the DNA at the CAP site. This increases promoter efficiency and results in the
Biochemistry_Lippincott_1668
Biochemistry_Lippinco
In addition, cyclic adenosine monophosphate (cAMP) binds the catabolite activator protein (CAP), and the complex binds the DNA at the CAP site. This increases promoter efficiency and results in the expression of the structural genes through the production of a polycistronic messenger RNA (mRNA). When both glucose and lactose are present, glucose prevents formation of cAMP, and transcription of these genes is negligible. The trp operon contains genes needed for the synthesis of tryptophan (Trp), and, like the lac operon, it is regulated by negative control. Unlike the lac operon, it is also regulated by attenuation, in which mRNA synthesis that escaped repression by Trp is terminated before completion. Transcription of ribosomal RNA and transfer RNA is selectively inhibited in prokaryotes by the stringent response to amino acid starvation. Translation is also a site of prokaryotic gene regulation: Excess ribosomal proteins bind the Shine-Dalgarno sequence on their own polycistronic
Biochemistry_Lippinco. In addition, cyclic adenosine monophosphate (cAMP) binds the catabolite activator protein (CAP), and the complex binds the DNA at the CAP site. This increases promoter efficiency and results in the expression of the structural genes through the production of a polycistronic messenger RNA (mRNA). When both glucose and lactose are present, glucose prevents formation of cAMP, and transcription of these genes is negligible. The trp operon contains genes needed for the synthesis of tryptophan (Trp), and, like the lac operon, it is regulated by negative control. Unlike the lac operon, it is also regulated by attenuation, in which mRNA synthesis that escaped repression by Trp is terminated before completion. Transcription of ribosomal RNA and transfer RNA is selectively inhibited in prokaryotes by the stringent response to amino acid starvation. Translation is also a site of prokaryotic gene regulation: Excess ribosomal proteins bind the Shine-Dalgarno sequence on their own polycistronic
Biochemistry_Lippincott_1669
Biochemistry_Lippinco
by the stringent response to amino acid starvation. Translation is also a site of prokaryotic gene regulation: Excess ribosomal proteins bind the Shine-Dalgarno sequence on their own polycistronic mRNA, preventing ribosomes from binding. Gene regulation is more complex in eukaryotes. Operons are not present, but coordinate regulation of the transcription of genes located on different chromosomes can be achieved through the binding of trans-acting proteins to cis-acting elements as seen in the galactose circuit in unicellular yeast. In multicellular organisms, hormones can cause coordinated regulation, either through the binding of the hormone receptor–hormone complex to the DNA (as with steroid hormones) or through the binding of a protein that is activated in response to a second messenger (as with glucagon). In each case, binding to DNA is mediated through structural motifs such as the zinc finger. Co-and posttranscriptional regulation is also seen in eukaryotes and includes
Biochemistry_Lippinco. by the stringent response to amino acid starvation. Translation is also a site of prokaryotic gene regulation: Excess ribosomal proteins bind the Shine-Dalgarno sequence on their own polycistronic mRNA, preventing ribosomes from binding. Gene regulation is more complex in eukaryotes. Operons are not present, but coordinate regulation of the transcription of genes located on different chromosomes can be achieved through the binding of trans-acting proteins to cis-acting elements as seen in the galactose circuit in unicellular yeast. In multicellular organisms, hormones can cause coordinated regulation, either through the binding of the hormone receptor–hormone complex to the DNA (as with steroid hormones) or through the binding of a protein that is activated in response to a second messenger (as with glucagon). In each case, binding to DNA is mediated through structural motifs such as the zinc finger. Co-and posttranscriptional regulation is also seen in eukaryotes and includes
Biochemistry_Lippincott_1670
Biochemistry_Lippinco
messenger (as with glucagon). In each case, binding to DNA is mediated through structural motifs such as the zinc finger. Co-and posttranscriptional regulation is also seen in eukaryotes and includes alternative mRNA splicing and polyadenylation, mRNA editing, and variations in mRNA stability as seen with transferrin receptor synthesis and with RNA interference. Regulation at the translational level can be caused by the phosphorylation and inhibition of eukaryotic initiation factor 2. Gene expression in eukaryotes is also influenced by accessibility of DNA to the transcriptional apparatus (as seen with epigenetic changes to histone proteins), the amount of DNA, and the arrangement of the DNA.
Biochemistry_Lippinco. messenger (as with glucagon). In each case, binding to DNA is mediated through structural motifs such as the zinc finger. Co-and posttranscriptional regulation is also seen in eukaryotes and includes alternative mRNA splicing and polyadenylation, mRNA editing, and variations in mRNA stability as seen with transferrin receptor synthesis and with RNA interference. Regulation at the translational level can be caused by the phosphorylation and inhibition of eukaryotic initiation factor 2. Gene expression in eukaryotes is also influenced by accessibility of DNA to the transcriptional apparatus (as seen with epigenetic changes to histone proteins), the amount of DNA, and the arrangement of the DNA.
Biochemistry_Lippincott_1671
Biochemistry_Lippinco
reticulum. Choose the ONE best answer. 3.1. Which of the following mutations is most likely to result in reduced expression of the lac operon? A. cya− (no adenylyl cyclase made) B. i− (no repressor protein made) C. Oc (operator cannot bind repressor protein) D. One resulting in impaired glucose uptake Correct answer = A. In the absence of glucose, adenylyl cyclase makes cyclic adenosine monophosphate (cAMP), which forms a complex with the catabolite activator protein (CAP). The cAMP–CAP complex binds the CAP site on the DNA, causing RNA polymerase to bind more efficiently to the lac operon promoter, thereby increasing expression of the operon. With cya− mutations, adenylyl cyclase is not made, and so the operon is unable to be maximally expressed even when glucose is absent and lactose is present. The absence of a repressor protein or decreased ability of the repressor to bind the operator results in constitutive (essentially constant) expression of the lac operon.
Biochemistry_Lippinco. reticulum. Choose the ONE best answer. 3.1. Which of the following mutations is most likely to result in reduced expression of the lac operon? A. cya− (no adenylyl cyclase made) B. i− (no repressor protein made) C. Oc (operator cannot bind repressor protein) D. One resulting in impaired glucose uptake Correct answer = A. In the absence of glucose, adenylyl cyclase makes cyclic adenosine monophosphate (cAMP), which forms a complex with the catabolite activator protein (CAP). The cAMP–CAP complex binds the CAP site on the DNA, causing RNA polymerase to bind more efficiently to the lac operon promoter, thereby increasing expression of the operon. With cya− mutations, adenylyl cyclase is not made, and so the operon is unable to be maximally expressed even when glucose is absent and lactose is present. The absence of a repressor protein or decreased ability of the repressor to bind the operator results in constitutive (essentially constant) expression of the lac operon.
Biochemistry_Lippincott_1672
Biochemistry_Lippinco
3.2. Which of the following is best described as cis-acting? A. Cyclic adenosine monophosphate response element–binding protein B. Operator C. Repressor protein D. Thyroid hormone nuclear receptor Correct answer = B. The operator is part of the DNA itself, and so is cis-acting. The cyclic adenosine monophosphate response element–binding protein, repressor protein, and thyroid hormone nuclear receptor protein are molecules that diffuse (transit) to the DNA, bind, and affect the expression of that DNA and so are trans-acting. 3.3. Which of the following is the basis for the intestine-specific expression of apolipoprotein B-48? A. DNA rearrangement and loss B. DNA transposition C. RNA alternative splicing D. RNA editing E. RNA interference
Biochemistry_Lippinco. 3.2. Which of the following is best described as cis-acting? A. Cyclic adenosine monophosphate response element–binding protein B. Operator C. Repressor protein D. Thyroid hormone nuclear receptor Correct answer = B. The operator is part of the DNA itself, and so is cis-acting. The cyclic adenosine monophosphate response element–binding protein, repressor protein, and thyroid hormone nuclear receptor protein are molecules that diffuse (transit) to the DNA, bind, and affect the expression of that DNA and so are trans-acting. 3.3. Which of the following is the basis for the intestine-specific expression of apolipoprotein B-48? A. DNA rearrangement and loss B. DNA transposition C. RNA alternative splicing D. RNA editing E. RNA interference
Biochemistry_Lippincott_1673
Biochemistry_Lippinco
A. DNA rearrangement and loss B. DNA transposition C. RNA alternative splicing D. RNA editing E. RNA interference Correct answer = D. The production of apolipoprotein (apo) B-48 in the intestine and apo B-100 in liver is the result of RNA editing in the intestine, where a sense codon is changed to a nonsense codon by posttranscriptional deamination of cytosine to uracil. DNA rearrangement and transposition, as well as RNA interference and alternative splicing, do alter gene expression but are not the basis of apo B-48 tissue-specific production. 3.4. Which of the following is most likely to be true in hemochromatosis, a disease of iron accumulation? A. The messenger RNA for the transferrin receptor is stabilized by the binding of iron regulatory proteins to its 3′-iron-responsive elements. B. The messenger RNA for the transferrin receptor is not bound by iron regulatory proteins and is degraded.
Biochemistry_Lippinco. A. DNA rearrangement and loss B. DNA transposition C. RNA alternative splicing D. RNA editing E. RNA interference Correct answer = D. The production of apolipoprotein (apo) B-48 in the intestine and apo B-100 in liver is the result of RNA editing in the intestine, where a sense codon is changed to a nonsense codon by posttranscriptional deamination of cytosine to uracil. DNA rearrangement and transposition, as well as RNA interference and alternative splicing, do alter gene expression but are not the basis of apo B-48 tissue-specific production. 3.4. Which of the following is most likely to be true in hemochromatosis, a disease of iron accumulation? A. The messenger RNA for the transferrin receptor is stabilized by the binding of iron regulatory proteins to its 3′-iron-responsive elements. B. The messenger RNA for the transferrin receptor is not bound by iron regulatory proteins and is degraded.
Biochemistry_Lippincott_1674
Biochemistry_Lippinco
B. The messenger RNA for the transferrin receptor is not bound by iron regulatory proteins and is degraded. C. The messenger RNA for ferritin is not bound by iron regulatory proteins at its 5′-iron-responsive element and is translated. D. The messenger RNA for ferritin is bound by iron regulatory proteins and is not translated. E. Both B and C are correct. Correct answer = E. When iron levels in the body are high, as is seen with hemochromatosis, there is increased synthesis of the iron-storage molecule, ferritin, and decreased synthesis of the transferrin receptor (TfR) that mediates iron uptake by cells. These effects are the result of cis-acting iron-responsive elements not being bound by trans-acting iron regulatory proteins, resulting in degradation of the messenger RNA (mRNA) for TfR and increased translation of the mRNA for ferritin.
Biochemistry_Lippinco. B. The messenger RNA for the transferrin receptor is not bound by iron regulatory proteins and is degraded. C. The messenger RNA for ferritin is not bound by iron regulatory proteins at its 5′-iron-responsive element and is translated. D. The messenger RNA for ferritin is bound by iron regulatory proteins and is not translated. E. Both B and C are correct. Correct answer = E. When iron levels in the body are high, as is seen with hemochromatosis, there is increased synthesis of the iron-storage molecule, ferritin, and decreased synthesis of the transferrin receptor (TfR) that mediates iron uptake by cells. These effects are the result of cis-acting iron-responsive elements not being bound by trans-acting iron regulatory proteins, resulting in degradation of the messenger RNA (mRNA) for TfR and increased translation of the mRNA for ferritin.
Biochemistry_Lippincott_1675
Biochemistry_Lippinco
3.5. Patients with estrogen receptor–positive (hormone responsive) breast cancer may be treated with the drug tamoxifen, which binds the estrogen nuclear receptor without activating it. Which of the following is the most logical outcome of tamoxifen use? A. Increased acetylation of estrogen-responsive genes B. Increased growth of estrogen receptor–positive breast cancer cells C. Increased production of cyclic adenosine monophosphate D. Inhibition of the estrogen operon E. Inhibition of transcription of estrogen-responsive genes
Biochemistry_Lippinco. 3.5. Patients with estrogen receptor–positive (hormone responsive) breast cancer may be treated with the drug tamoxifen, which binds the estrogen nuclear receptor without activating it. Which of the following is the most logical outcome of tamoxifen use? A. Increased acetylation of estrogen-responsive genes B. Increased growth of estrogen receptor–positive breast cancer cells C. Increased production of cyclic adenosine monophosphate D. Inhibition of the estrogen operon E. Inhibition of transcription of estrogen-responsive genes
Biochemistry_Lippincott_1676
Biochemistry_Lippinco
C. Increased production of cyclic adenosine monophosphate D. Inhibition of the estrogen operon E. Inhibition of transcription of estrogen-responsive genes Correct answer = E. Tamoxifen competes with estrogen for binding to the estrogen nuclear receptor. Tamoxifen fails to activate the receptor, preventing its binding to DNA sequences that upregulate expression of estrogen-responsive genes. Tamoxifen, then, blocks the growth-promoting effects of these genes and results in growth inhibition of estrogen-dependent breast cancer cells. Acetylation increases transcription by relaxing the nucleosome. Cyclic adenosine monophosphate is a regulatory signal mediated by cell-surface rather than nuclear receptors. Mammalian cells do not have operons. 3.6. The ZYA region of the lac operon will be maximally expressed if: A. cyclic adenosine monophosphate levels are low. B. glucose and lactose are both available. C. the attenuation stem-loop is able to form. D. the CAP site is occupied.
Biochemistry_Lippinco. C. Increased production of cyclic adenosine monophosphate D. Inhibition of the estrogen operon E. Inhibition of transcription of estrogen-responsive genes Correct answer = E. Tamoxifen competes with estrogen for binding to the estrogen nuclear receptor. Tamoxifen fails to activate the receptor, preventing its binding to DNA sequences that upregulate expression of estrogen-responsive genes. Tamoxifen, then, blocks the growth-promoting effects of these genes and results in growth inhibition of estrogen-dependent breast cancer cells. Acetylation increases transcription by relaxing the nucleosome. Cyclic adenosine monophosphate is a regulatory signal mediated by cell-surface rather than nuclear receptors. Mammalian cells do not have operons. 3.6. The ZYA region of the lac operon will be maximally expressed if: A. cyclic adenosine monophosphate levels are low. B. glucose and lactose are both available. C. the attenuation stem-loop is able to form. D. the CAP site is occupied.
Biochemistry_Lippincott_1677
Biochemistry_Lippinco
A. cyclic adenosine monophosphate levels are low. B. glucose and lactose are both available. C. the attenuation stem-loop is able to form. D. the CAP site is occupied. Correct answer = D. It is only when glucose is gone, cyclic adenosine monophosphate (cAMP) levels are increased, the cAMP–catabolite activator protein (CAP) complex is bound to the CAP site, and lactose is available that the operon is maximally expressed (induced). If glucose is present, the operon is off as a result of catabolite repression. The lac operon is not regulated by attenuation, a mechanism for stopping transcription in some operons such as the trp operon. 3.7. X chromosome inactivation is a process by which one of two X chromosomes in mammalian females is condensed and inactivated to prevent overexpression of X-linked genes. What would most likely be true about the degree of DNA methylation and histone acetylation on the inactivated X chromosome?
Biochemistry_Lippinco. A. cyclic adenosine monophosphate levels are low. B. glucose and lactose are both available. C. the attenuation stem-loop is able to form. D. the CAP site is occupied. Correct answer = D. It is only when glucose is gone, cyclic adenosine monophosphate (cAMP) levels are increased, the cAMP–catabolite activator protein (CAP) complex is bound to the CAP site, and lactose is available that the operon is maximally expressed (induced). If glucose is present, the operon is off as a result of catabolite repression. The lac operon is not regulated by attenuation, a mechanism for stopping transcription in some operons such as the trp operon. 3.7. X chromosome inactivation is a process by which one of two X chromosomes in mammalian females is condensed and inactivated to prevent overexpression of X-linked genes. What would most likely be true about the degree of DNA methylation and histone acetylation on the inactivated X chromosome?
Biochemistry_Lippincott_1678
Biochemistry_Lippinco
Cytosines in CpG islands would be hypermethylated, and histone proteins would be deacetylated. Both conditions are associated with decreased gene expression, and both are important in maintaining X inactivation. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW
Biochemistry_Lippinco. Cytosines in CpG islands would be hypermethylated, and histone proteins would be deacetylated. Both conditions are associated with decreased gene expression, and both are important in maintaining X inactivation. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW
Biochemistry_Lippincott_1679
Biochemistry_Lippinco
In the past, efforts to understand genes and their expression have been confounded by the immense size and complexity of human deoxyribonucleic acid (DNA). The human genome contains ~3 billion (109) base pairs (bp) that encode 20,000–25,000 protein-coding genes located on 23 chromosomes in the haploid genome. It is now possible to determine the nucleotide sequence of long stretches of DNA, and the entire human genome has been sequenced. This effort (called the Human Genome Project and completed in 2003) was made possible by several tools that have already contributed to our understanding of many genetic diseases (Fig. 34.1). These include 1) the discovery of restriction endonucleases that permit the cleavage of huge DNA molecules into defined fragments, 2) the development of cloning techniques that provide a mechanism for amplification of specific nucleotide sequences, and 3) the ability to synthesize specific probes, which has allowed the identification and manipulation of nucleotide
Biochemistry_Lippinco. In the past, efforts to understand genes and their expression have been confounded by the immense size and complexity of human deoxyribonucleic acid (DNA). The human genome contains ~3 billion (109) base pairs (bp) that encode 20,000–25,000 protein-coding genes located on 23 chromosomes in the haploid genome. It is now possible to determine the nucleotide sequence of long stretches of DNA, and the entire human genome has been sequenced. This effort (called the Human Genome Project and completed in 2003) was made possible by several tools that have already contributed to our understanding of many genetic diseases (Fig. 34.1). These include 1) the discovery of restriction endonucleases that permit the cleavage of huge DNA molecules into defined fragments, 2) the development of cloning techniques that provide a mechanism for amplification of specific nucleotide sequences, and 3) the ability to synthesize specific probes, which has allowed the identification and manipulation of nucleotide
Biochemistry_Lippincott_1680
Biochemistry_Lippinco
that provide a mechanism for amplification of specific nucleotide sequences, and 3) the ability to synthesize specific probes, which has allowed the identification and manipulation of nucleotide sequences of interest. These and other experimental approaches have permitted the identification of both normal and mutant nucleotide sequences in DNA. This knowledge has led to the development of methods for the diagnosis of genetic diseases and some successes in the treatment of patients by gene therapy. [Note: The genomes of several viruses, prokaryotes, and nonhuman eukaryotes have also been sequenced.]
Biochemistry_Lippinco. that provide a mechanism for amplification of specific nucleotide sequences, and 3) the ability to synthesize specific probes, which has allowed the identification and manipulation of nucleotide sequences of interest. These and other experimental approaches have permitted the identification of both normal and mutant nucleotide sequences in DNA. This knowledge has led to the development of methods for the diagnosis of genetic diseases and some successes in the treatment of patients by gene therapy. [Note: The genomes of several viruses, prokaryotes, and nonhuman eukaryotes have also been sequenced.]
Biochemistry_Lippincott_1681
Biochemistry_Lippinco
II. RESTRICTION ENDONUCLEASES One of the major obstacles to molecular analysis of genomic DNA is the immense size of the molecules involved. The discovery of a special group of bacterial enzymes, called restriction endonucleases (restriction enzymes), which cleave double-stranded DNA (dsDNA) into smaller, more manageable fragments, opened the way for DNA analysis. Because each enzyme cleaves dsDNA at a specific nucleotide sequence (restriction site), restriction enzymes are used experimentally to obtain precisely defined DNA segments called restriction fragments. A. Specificity
Biochemistry_Lippinco. II. RESTRICTION ENDONUCLEASES One of the major obstacles to molecular analysis of genomic DNA is the immense size of the molecules involved. The discovery of a special group of bacterial enzymes, called restriction endonucleases (restriction enzymes), which cleave double-stranded DNA (dsDNA) into smaller, more manageable fragments, opened the way for DNA analysis. Because each enzyme cleaves dsDNA at a specific nucleotide sequence (restriction site), restriction enzymes are used experimentally to obtain precisely defined DNA segments called restriction fragments. A. Specificity
Biochemistry_Lippincott_1682
Biochemistry_Lippinco
A. Specificity Restriction endonucleases recognize short stretches of dsDNA (4–8 bp) that contain specific nucleotide sequences. These sequences, which differ for each restriction enzyme, are palindromes, that is, they exhibit twofold rotational symmetry (Fig. 34.2). This means that, within a short region of the dsDNA, the nucleotide sequence on the two strands is identical if each is read in the 5′→3′ direction. Therefore, if you turn the page upside down (that is, rotate it 180° around its axis of symmetry) the sequence remains the same. In bacteria, restriction endonucleases limit (restrict) the expression of nonbacterial (foreign) DNA through cleavage. Bacterial DNA is protected from cleavage by methylation of adenine at the restriction site. B. Nomenclature
Biochemistry_Lippinco. A. Specificity Restriction endonucleases recognize short stretches of dsDNA (4–8 bp) that contain specific nucleotide sequences. These sequences, which differ for each restriction enzyme, are palindromes, that is, they exhibit twofold rotational symmetry (Fig. 34.2). This means that, within a short region of the dsDNA, the nucleotide sequence on the two strands is identical if each is read in the 5′→3′ direction. Therefore, if you turn the page upside down (that is, rotate it 180° around its axis of symmetry) the sequence remains the same. In bacteria, restriction endonucleases limit (restrict) the expression of nonbacterial (foreign) DNA through cleavage. Bacterial DNA is protected from cleavage by methylation of adenine at the restriction site. B. Nomenclature