id
stringlengths 14
28
| title
stringclasses 18
values | content
stringlengths 2
999
| contents
stringlengths 19
1.02k
|
---|---|---|---|
Biochemistry_Lippincott_183 | Biochemistry_Lippinco | A. Active site Enzyme molecules contain a special pocket or cleft called the active site. The active site, formed by folding of the protein, contains amino acid side chains that participate in substrate binding and catalysis (Fig. 5.2). The substrate binds the enzyme noncovalently, forming an enzyme–substrate (ES) complex. Binding is thought to cause a conformational change in the enzyme (induced fit model) that allows catalysis. ES is converted to an enzyme–product (EP) complex that subsequently dissociates to enzyme and product. B. Efficiency Enzyme-catalyzed reactions are highly efficient, proceeding from 103 to 108 times faster than uncatalyzed reactions. The number of substrate molecules converted to product per enzyme molecule per second is called the turnover number, or kcat, and typically is 102–104 s−1 . [Note: kcat is the rate constant for the conversion of ES to E + P (see p. 58).] C. Specificity | Biochemistry_Lippinco. A. Active site Enzyme molecules contain a special pocket or cleft called the active site. The active site, formed by folding of the protein, contains amino acid side chains that participate in substrate binding and catalysis (Fig. 5.2). The substrate binds the enzyme noncovalently, forming an enzyme–substrate (ES) complex. Binding is thought to cause a conformational change in the enzyme (induced fit model) that allows catalysis. ES is converted to an enzyme–product (EP) complex that subsequently dissociates to enzyme and product. B. Efficiency Enzyme-catalyzed reactions are highly efficient, proceeding from 103 to 108 times faster than uncatalyzed reactions. The number of substrate molecules converted to product per enzyme molecule per second is called the turnover number, or kcat, and typically is 102–104 s−1 . [Note: kcat is the rate constant for the conversion of ES to E + P (see p. 58).] C. Specificity |
Biochemistry_Lippincott_184 | Biochemistry_Lippinco | C. Specificity Enzymes are highly specific, interacting with one or a few substrates and catalyzing only one type of chemical reaction. The set of enzymes made in a cell determines which reactions occur in that cell. D. Holoenzymes, apoenzymes, cofactors, and coenzymes | Biochemistry_Lippinco. C. Specificity Enzymes are highly specific, interacting with one or a few substrates and catalyzing only one type of chemical reaction. The set of enzymes made in a cell determines which reactions occur in that cell. D. Holoenzymes, apoenzymes, cofactors, and coenzymes |
Biochemistry_Lippincott_185 | Biochemistry_Lippinco | D. Holoenzymes, apoenzymes, cofactors, and coenzymes Some enzymes require nonproteins for enzymic activity. The term holoenzyme refers to the active enzyme with its nonprotein component, whereas the enzyme without its nonprotein moiety is termed an apoenzyme and is inactive. If the nonprotein moiety is a metal ion, such as zinc (Zn2+) or iron (Fe2+), it is called a cofactor (see Chapter 29). If it is a small organic molecule, it is termed a coenzyme. Coenzymes that only transiently associate with the enzyme are called cosubstrates. Cosubstrates dissociate from the enzyme in an altered state (NAD+ is an example; see p. 101). If the coenzyme is permanently associated with the enzyme and returned to its original form, it is called a prosthetic group (FAD is an example; see p. 110). Coenzymes commonly are derived from vitamins. For example, NAD+ contains niacin, and FAD contains riboflavin (see Chapter 28). E. Regulation | Biochemistry_Lippinco. D. Holoenzymes, apoenzymes, cofactors, and coenzymes Some enzymes require nonproteins for enzymic activity. The term holoenzyme refers to the active enzyme with its nonprotein component, whereas the enzyme without its nonprotein moiety is termed an apoenzyme and is inactive. If the nonprotein moiety is a metal ion, such as zinc (Zn2+) or iron (Fe2+), it is called a cofactor (see Chapter 29). If it is a small organic molecule, it is termed a coenzyme. Coenzymes that only transiently associate with the enzyme are called cosubstrates. Cosubstrates dissociate from the enzyme in an altered state (NAD+ is an example; see p. 101). If the coenzyme is permanently associated with the enzyme and returned to its original form, it is called a prosthetic group (FAD is an example; see p. 110). Coenzymes commonly are derived from vitamins. For example, NAD+ contains niacin, and FAD contains riboflavin (see Chapter 28). E. Regulation |
Biochemistry_Lippincott_186 | Biochemistry_Lippinco | E. Regulation Enzyme activity can be regulated, that is, increased or decreased, so that the rate of product formation responds to cellular need. F. Location within the cell Many enzymes are localized in specific organelles within the cell (Fig. 5.3). Such compartmentalization serves to isolate the reaction substrate or product from other competing reactions. This provides a favorable environment for the reaction and organizes the thousands of enzymes present in the cell into purposeful pathways. IV. MECHANISM OF ENZYME ACTION The mechanism of enzyme action can be viewed from two different perspectives. The first treats catalysis in terms of energy changes that occur during the reaction. That is, enzymes provide an alternate, energetically favorable reaction pathway different from the uncatalyzed reaction. The second perspective describes how the active site chemically facilitates catalysis. A. Energy changes occurring during the reaction | Biochemistry_Lippinco. E. Regulation Enzyme activity can be regulated, that is, increased or decreased, so that the rate of product formation responds to cellular need. F. Location within the cell Many enzymes are localized in specific organelles within the cell (Fig. 5.3). Such compartmentalization serves to isolate the reaction substrate or product from other competing reactions. This provides a favorable environment for the reaction and organizes the thousands of enzymes present in the cell into purposeful pathways. IV. MECHANISM OF ENZYME ACTION The mechanism of enzyme action can be viewed from two different perspectives. The first treats catalysis in terms of energy changes that occur during the reaction. That is, enzymes provide an alternate, energetically favorable reaction pathway different from the uncatalyzed reaction. The second perspective describes how the active site chemically facilitates catalysis. A. Energy changes occurring during the reaction |
Biochemistry_Lippincott_187 | Biochemistry_Lippinco | A. Energy changes occurring during the reaction Virtually all chemical reactions have an energy barrier separating the reactants and the products. This barrier, called the activation energy (Ea), is the energy difference between that of the reactants and a high-energy intermediate, the transition state (T*), which is formed during the conversion of reactant to product. Figure 5.4 shows the changes in energy during the conversion of a molecule of reactant A to product B as it proceeds through the transition state. 1. Activation energy: The peak of energy in Figure 5.4 is the difference in free energy between the reactant and T*, in which the high-energy, short-lived intermediate is formed during the conversion of reactant to product. Because of the high Ea, the rates of uncatalyzed chemical reactions are often slow. 2. | Biochemistry_Lippinco. A. Energy changes occurring during the reaction Virtually all chemical reactions have an energy barrier separating the reactants and the products. This barrier, called the activation energy (Ea), is the energy difference between that of the reactants and a high-energy intermediate, the transition state (T*), which is formed during the conversion of reactant to product. Figure 5.4 shows the changes in energy during the conversion of a molecule of reactant A to product B as it proceeds through the transition state. 1. Activation energy: The peak of energy in Figure 5.4 is the difference in free energy between the reactant and T*, in which the high-energy, short-lived intermediate is formed during the conversion of reactant to product. Because of the high Ea, the rates of uncatalyzed chemical reactions are often slow. 2. |
Biochemistry_Lippincott_188 | Biochemistry_Lippinco | 2. Rate of reaction: For molecules to react, they must contain sufficient energy to overcome the energy barrier of the transition state. In the absence of an enzyme, only a small proportion of a population of molecules may possess enough energy to achieve the transition state between reactant and product. The rate of reaction is determined by the number of such energized molecules. In general, the lower the Ea, the more molecules have sufficient energy to pass through the transition state and, therefore, the faster the rate of the reaction. 3. Alternate reaction pathway: An enzyme allows a reaction to proceed rapidly under conditions prevailing in the cell by providing an alternate reaction pathway with a lower Ea (see Fig. 5.4). The enzyme does not change the free energies of the reactants (substrates) or products and, therefore, does not change the equilibrium of the reaction (see p. 70). It does, however, accelerate the rate by which equilibrium is reached. | Biochemistry_Lippinco. 2. Rate of reaction: For molecules to react, they must contain sufficient energy to overcome the energy barrier of the transition state. In the absence of an enzyme, only a small proportion of a population of molecules may possess enough energy to achieve the transition state between reactant and product. The rate of reaction is determined by the number of such energized molecules. In general, the lower the Ea, the more molecules have sufficient energy to pass through the transition state and, therefore, the faster the rate of the reaction. 3. Alternate reaction pathway: An enzyme allows a reaction to proceed rapidly under conditions prevailing in the cell by providing an alternate reaction pathway with a lower Ea (see Fig. 5.4). The enzyme does not change the free energies of the reactants (substrates) or products and, therefore, does not change the equilibrium of the reaction (see p. 70). It does, however, accelerate the rate by which equilibrium is reached. |
Biochemistry_Lippincott_189 | Biochemistry_Lippinco | B. Active site chemistry The active site is not a passive receptacle for binding the substrate but, rather, is a complex molecular machine employing a diversity of chemical mechanisms to facilitate the conversion of substrate to product. A number of factors are responsible for the catalytic efficiency of enzymes, including the following examples. 1. Transition-state stabilization: The active site often acts as a flexible molecular template that binds the substrate and initiates its conversion to the transition state, a structure in which the bonds are not like those in the substrate or the product (see T* at the top of the curve in Fig. 5.4). By stabilizing the transition state, the enzyme greatly increases the concentration of the reactive intermediate that can be converted to product and, thus, accelerates the reaction. [Note: The transition state cannot be isolated.] 2. | Biochemistry_Lippinco. B. Active site chemistry The active site is not a passive receptacle for binding the substrate but, rather, is a complex molecular machine employing a diversity of chemical mechanisms to facilitate the conversion of substrate to product. A number of factors are responsible for the catalytic efficiency of enzymes, including the following examples. 1. Transition-state stabilization: The active site often acts as a flexible molecular template that binds the substrate and initiates its conversion to the transition state, a structure in which the bonds are not like those in the substrate or the product (see T* at the top of the curve in Fig. 5.4). By stabilizing the transition state, the enzyme greatly increases the concentration of the reactive intermediate that can be converted to product and, thus, accelerates the reaction. [Note: The transition state cannot be isolated.] 2. |
Biochemistry_Lippincott_190 | Biochemistry_Lippinco | Catalysis: The active site can provide catalytic groups that enhance the probability that the transition state is formed. In some enzymes, these groups can participate in general acid–base catalysis in which amino acid residues provide or accept protons. In other enzymes, catalysis may involve the transient formation of a covalent ES complex. [Note: The mechanism of action of chymotrypsin, an enzyme of protein digestion in the intestine, includes general base, general acid, and covalent catalysis. A histidine at the active site of the enzyme gains (general base) and loses (general acid) protons, mediated by the pK of histidine in proteins being close to physiologic pH. Serine at the active site forms a transient covalent bond with the substrate.] 3. | Biochemistry_Lippinco. Catalysis: The active site can provide catalytic groups that enhance the probability that the transition state is formed. In some enzymes, these groups can participate in general acid–base catalysis in which amino acid residues provide or accept protons. In other enzymes, catalysis may involve the transient formation of a covalent ES complex. [Note: The mechanism of action of chymotrypsin, an enzyme of protein digestion in the intestine, includes general base, general acid, and covalent catalysis. A histidine at the active site of the enzyme gains (general base) and loses (general acid) protons, mediated by the pK of histidine in proteins being close to physiologic pH. Serine at the active site forms a transient covalent bond with the substrate.] 3. |
Biochemistry_Lippincott_191 | Biochemistry_Lippinco | Transition-state visualization: The enzyme-catalyzed conversion of substrate to product can be visualized as being similar to removing a sweater from an uncooperative infant (Fig. 5.5). The process has a high Ea because the only reasonable strategy for removing the garment (short of ripping it off) requires that the random flailing of the baby results in both arms being fully extended over the head, an unlikely posture. However, we can envision a parent acting as an enzyme, first coming in contact with the baby (forming ES) and then guiding the baby’s arms into an extended, vertical position, analogous to the transition state. This posture (conformation) of the baby facilitates the removal of the sweater, forming the disrobed baby, which here represents product. [Note: The substrate bound to the enzyme (ES) is at a slightly lower energy than unbound substrate (S) and explains the small dip in the curve at ES.] V. FACTORS AFFECTING REACTION VELOCITY | Biochemistry_Lippinco. Transition-state visualization: The enzyme-catalyzed conversion of substrate to product can be visualized as being similar to removing a sweater from an uncooperative infant (Fig. 5.5). The process has a high Ea because the only reasonable strategy for removing the garment (short of ripping it off) requires that the random flailing of the baby results in both arms being fully extended over the head, an unlikely posture. However, we can envision a parent acting as an enzyme, first coming in contact with the baby (forming ES) and then guiding the baby’s arms into an extended, vertical position, analogous to the transition state. This posture (conformation) of the baby facilitates the removal of the sweater, forming the disrobed baby, which here represents product. [Note: The substrate bound to the enzyme (ES) is at a slightly lower energy than unbound substrate (S) and explains the small dip in the curve at ES.] V. FACTORS AFFECTING REACTION VELOCITY |
Biochemistry_Lippincott_192 | Biochemistry_Lippinco | V. FACTORS AFFECTING REACTION VELOCITY Enzymes can be isolated from cells and their properties studied in a test tube (that is, in vitro). Different enzymes show different responses to changes in substrate concentration, temperature, and pH. This section describes factors that influence the reaction velocity of enzymes. Enzymic responses to these factors give us valuable clues as to how enzymes function in living cells (that is, in vivo). A. Substrate concentration 1. Maximal velocity: The rate or velocity of a reaction (v) is the number of substrate molecules converted to product per unit time. Velocity is usually expressed as µmol of product formed per minute. The rate of an enzyme-catalyzed reaction increases with substrate concentration until a maximal velocity (Vmax) is reached (Fig. 5.6). The leveling off of the reaction rate at high substrate concentrations reflects the saturation with substrate of all available binding sites on the enzyme molecules present. 2. | Biochemistry_Lippinco. V. FACTORS AFFECTING REACTION VELOCITY Enzymes can be isolated from cells and their properties studied in a test tube (that is, in vitro). Different enzymes show different responses to changes in substrate concentration, temperature, and pH. This section describes factors that influence the reaction velocity of enzymes. Enzymic responses to these factors give us valuable clues as to how enzymes function in living cells (that is, in vivo). A. Substrate concentration 1. Maximal velocity: The rate or velocity of a reaction (v) is the number of substrate molecules converted to product per unit time. Velocity is usually expressed as µmol of product formed per minute. The rate of an enzyme-catalyzed reaction increases with substrate concentration until a maximal velocity (Vmax) is reached (Fig. 5.6). The leveling off of the reaction rate at high substrate concentrations reflects the saturation with substrate of all available binding sites on the enzyme molecules present. 2. |
Biochemistry_Lippincott_193 | Biochemistry_Lippinco | 2. Shape of the enzyme kinetics curve: Most enzymes show Michaelis-Menten kinetics (see p. 58), in which the plot of initial reaction velocity (vo) against substrate concentration is hyperbolic (similar in shape to that of the oxygen-dissociation curve of myoglobin; see p. 29). In contrast, allosteric enzymes do not follow Michaelis-Menten kinetics and show a sigmoidal curve (see Fig. 5.6) that is similar in shape to the oxygen-dissociation curve of hemoglobin (see p. 29). B. Temperature 1. Velocity increase with temperature: The reaction velocity increases with temperature until a peak velocity is reached (Fig. 5.7). This increase is the result of the increased number of substrate molecules having sufficient energy to pass over the energy barrier and form the products of the reaction. 2. | Biochemistry_Lippinco. 2. Shape of the enzyme kinetics curve: Most enzymes show Michaelis-Menten kinetics (see p. 58), in which the plot of initial reaction velocity (vo) against substrate concentration is hyperbolic (similar in shape to that of the oxygen-dissociation curve of myoglobin; see p. 29). In contrast, allosteric enzymes do not follow Michaelis-Menten kinetics and show a sigmoidal curve (see Fig. 5.6) that is similar in shape to the oxygen-dissociation curve of hemoglobin (see p. 29). B. Temperature 1. Velocity increase with temperature: The reaction velocity increases with temperature until a peak velocity is reached (Fig. 5.7). This increase is the result of the increased number of substrate molecules having sufficient energy to pass over the energy barrier and form the products of the reaction. 2. |
Biochemistry_Lippincott_194 | Biochemistry_Lippinco | 2. Velocity decrease with higher temperature: Further elevation of the temperature causes a decrease in reaction velocity as a result of temperature-induced denaturation of the enzyme (see Fig. 5.7). The optimum temperature for most human enzymes is between 35°C and 40°C. Human enzymes start to denature (see p. 20) at temperatures above 40°C, but thermophilic bacteria found in hot springs have optimum temperatures of 70°C. C. pH 1. pH effect on active site ionization: The concentration of protons ([H+]) affects reaction velocity in several ways. First, the catalytic process usually requires that the enzyme and substrate have specific chemical groups in either an ionized or unionized state in order to interact. For example, catalytic activity may require that an amino group of the enzyme be in the protonated form (−NH3+). Because this group is deprotonated at alkaline pH, the rate of the reaction declines. 2. | Biochemistry_Lippinco. 2. Velocity decrease with higher temperature: Further elevation of the temperature causes a decrease in reaction velocity as a result of temperature-induced denaturation of the enzyme (see Fig. 5.7). The optimum temperature for most human enzymes is between 35°C and 40°C. Human enzymes start to denature (see p. 20) at temperatures above 40°C, but thermophilic bacteria found in hot springs have optimum temperatures of 70°C. C. pH 1. pH effect on active site ionization: The concentration of protons ([H+]) affects reaction velocity in several ways. First, the catalytic process usually requires that the enzyme and substrate have specific chemical groups in either an ionized or unionized state in order to interact. For example, catalytic activity may require that an amino group of the enzyme be in the protonated form (−NH3+). Because this group is deprotonated at alkaline pH, the rate of the reaction declines. 2. |
Biochemistry_Lippincott_195 | Biochemistry_Lippinco | 2. pH effect on enzyme denaturation: Extremes of pH can also lead to denaturation of the enzyme, because the structure of the catalytically active protein molecule depends on the ionic character of the amino acid side chains. 3. Variable pH optimum: The pH at which maximal enzyme activity is achieved is different for different enzymes and often reflects the [H+] at which the enzyme functions in the body. For example, pepsin, a digestive enzyme in the stomach, is maximally active at pH 2, whereas other enzymes, designed to work at neutral pH, are denatured by such an acidic environment (Fig. 5.8). VI. MICHAELIS-MENTEN KINETICS | Biochemistry_Lippinco. 2. pH effect on enzyme denaturation: Extremes of pH can also lead to denaturation of the enzyme, because the structure of the catalytically active protein molecule depends on the ionic character of the amino acid side chains. 3. Variable pH optimum: The pH at which maximal enzyme activity is achieved is different for different enzymes and often reflects the [H+] at which the enzyme functions in the body. For example, pepsin, a digestive enzyme in the stomach, is maximally active at pH 2, whereas other enzymes, designed to work at neutral pH, are denatured by such an acidic environment (Fig. 5.8). VI. MICHAELIS-MENTEN KINETICS |
Biochemistry_Lippincott_196 | Biochemistry_Lippinco | VI. MICHAELIS-MENTEN KINETICS Leonor Michaelis and Maude Menten proposed a simple model that accounts for most of the features of many enzyme-catalyzed reactions. In this model, the enzyme reversibly combines with its substrate to form an ES complex that subsequently yields product, regenerating the free enzyme. The reaction model, involving one substrate molecule, is represented below: where S is the substrate. E is the enzyme. ES is the enzyme–substrate complex. P is the product. k1, k−1, and k2 (or, kcat) are rate constants. A. Michaelis-Menten equation The Michaelis-Menten equation describes how reaction velocity varies with substrate concentration: The following assumptions are made in deriving the Michaelis-Menten rate equation. 1. Enzyme and substrate relative concentrations: The substrate concentration ([S]) is much greater than the concentration of enzyme so that the percentage of total substrate bound by the enzyme at any one time is small. 2. | Biochemistry_Lippinco. VI. MICHAELIS-MENTEN KINETICS Leonor Michaelis and Maude Menten proposed a simple model that accounts for most of the features of many enzyme-catalyzed reactions. In this model, the enzyme reversibly combines with its substrate to form an ES complex that subsequently yields product, regenerating the free enzyme. The reaction model, involving one substrate molecule, is represented below: where S is the substrate. E is the enzyme. ES is the enzyme–substrate complex. P is the product. k1, k−1, and k2 (or, kcat) are rate constants. A. Michaelis-Menten equation The Michaelis-Menten equation describes how reaction velocity varies with substrate concentration: The following assumptions are made in deriving the Michaelis-Menten rate equation. 1. Enzyme and substrate relative concentrations: The substrate concentration ([S]) is much greater than the concentration of enzyme so that the percentage of total substrate bound by the enzyme at any one time is small. 2. |
Biochemistry_Lippincott_197 | Biochemistry_Lippinco | 2. Steady-state assumption: The concentration of the ES complex does not change with time (the steady-state assumption), that is, the rate of formation of ES is equal to that of the breakdown of ES (to E + S and to E + P). In general, an intermediate in a series of reactions is said to be in steady state when its rate of synthesis is equal to its rate of degradation. 3. Initial velocity: Initial reaction velocities (vo) are used in the analysis of enzyme reactions. This means that the rate of the reaction is measured as soon as enzyme and substrate are mixed. At that time, the concentration of product is very small, and therefore, the rate of the back reaction from product to substrate can be ignored. | Biochemistry_Lippinco. 2. Steady-state assumption: The concentration of the ES complex does not change with time (the steady-state assumption), that is, the rate of formation of ES is equal to that of the breakdown of ES (to E + S and to E + P). In general, an intermediate in a series of reactions is said to be in steady state when its rate of synthesis is equal to its rate of degradation. 3. Initial velocity: Initial reaction velocities (vo) are used in the analysis of enzyme reactions. This means that the rate of the reaction is measured as soon as enzyme and substrate are mixed. At that time, the concentration of product is very small, and therefore, the rate of the back reaction from product to substrate can be ignored. |
Biochemistry_Lippincott_198 | Biochemistry_Lippinco | B. Important conclusions 1. Km characteristics: Km, the Michaelis constant, is characteristic of an enzyme and its particular substrate and reflects the affinity of the enzyme for that substrate. Km is numerically equal to the substrate concentration at which the reaction velocity is equal to one half Vmax. Km does not vary with enzyme concentration. a. Small Km: A numerically small (low) Km reflects a high affinity of the enzyme for substrate, because a low concentration of substrate is needed to half-saturate the enzyme—that is, to reach a velocity that is one half Vmax (Fig. 5.9). b. Large Km: A numerically large (high) Km reflects a low affinity of enzyme for substrate because a high concentration of substrate is needed to half-saturate the enzyme. 2. | Biochemistry_Lippinco. B. Important conclusions 1. Km characteristics: Km, the Michaelis constant, is characteristic of an enzyme and its particular substrate and reflects the affinity of the enzyme for that substrate. Km is numerically equal to the substrate concentration at which the reaction velocity is equal to one half Vmax. Km does not vary with enzyme concentration. a. Small Km: A numerically small (low) Km reflects a high affinity of the enzyme for substrate, because a low concentration of substrate is needed to half-saturate the enzyme—that is, to reach a velocity that is one half Vmax (Fig. 5.9). b. Large Km: A numerically large (high) Km reflects a low affinity of enzyme for substrate because a high concentration of substrate is needed to half-saturate the enzyme. 2. |
Biochemistry_Lippincott_199 | Biochemistry_Lippinco | b. Large Km: A numerically large (high) Km reflects a low affinity of enzyme for substrate because a high concentration of substrate is needed to half-saturate the enzyme. 2. Velocity relationship to enzyme concentration: The rate of the reaction is directly proportional to the enzyme concentration because [S] is not limiting. For example, if the enzyme concentration is halved, the initial rates of the reaction (vo) and that of Vmax are reduced to half that of the original. 3. Reaction order: When [S] is much less (<<) than Km, the velocity of the reaction is approximately proportional to the substrate concentration (Fig. 5.10). The rate of reaction is then said to be first order with respect to substrate. When [S] is much greater (>>) than Km, the velocity is constant and equal to Vmax. The rate of reaction is then independent of substrate concentration (the enzyme is saturated with substrate) and is said to be zero order with respect to substrate concentration (see Fig. 5.10). | Biochemistry_Lippinco. b. Large Km: A numerically large (high) Km reflects a low affinity of enzyme for substrate because a high concentration of substrate is needed to half-saturate the enzyme. 2. Velocity relationship to enzyme concentration: The rate of the reaction is directly proportional to the enzyme concentration because [S] is not limiting. For example, if the enzyme concentration is halved, the initial rates of the reaction (vo) and that of Vmax are reduced to half that of the original. 3. Reaction order: When [S] is much less (<<) than Km, the velocity of the reaction is approximately proportional to the substrate concentration (Fig. 5.10). The rate of reaction is then said to be first order with respect to substrate. When [S] is much greater (>>) than Km, the velocity is constant and equal to Vmax. The rate of reaction is then independent of substrate concentration (the enzyme is saturated with substrate) and is said to be zero order with respect to substrate concentration (see Fig. 5.10). |
Biochemistry_Lippincott_200 | Biochemistry_Lippinco | D. Lineweaver-Burk plot When vo is plotted against [S], it is not always possible to determine when Vmax has been achieved because of the gradual upward slope of the hyperbolic curve at high substrate concentrations. However, if 1/vo is plotted versus 1/[S], a straight line is obtained (Fig. 5.11). This plot, the Lineweaver-Burk plot (also called a double-reciprocal plot) can be used to calculate Km and Vmax as well as to determine the mechanism of action of enzyme inhibitors. The equation describing the Lineweaver-Burk plot is: where the intercept on the x axis is equal to − 1/Km, and the intercept on the y axis is equal to 1/Vmax. [Note: The slope = Km/Vmax.] VII. ENZYME INHIBITION | Biochemistry_Lippinco. D. Lineweaver-Burk plot When vo is plotted against [S], it is not always possible to determine when Vmax has been achieved because of the gradual upward slope of the hyperbolic curve at high substrate concentrations. However, if 1/vo is plotted versus 1/[S], a straight line is obtained (Fig. 5.11). This plot, the Lineweaver-Burk plot (also called a double-reciprocal plot) can be used to calculate Km and Vmax as well as to determine the mechanism of action of enzyme inhibitors. The equation describing the Lineweaver-Burk plot is: where the intercept on the x axis is equal to − 1/Km, and the intercept on the y axis is equal to 1/Vmax. [Note: The slope = Km/Vmax.] VII. ENZYME INHIBITION |
Biochemistry_Lippincott_201 | Biochemistry_Lippinco | VII. ENZYME INHIBITION Any substance that can decrease the velocity of an enzyme-catalyzed reaction is called an inhibitor. Inhibitors can be reversible or irreversible. Irreversible inhibitors bind to enzymes through covalent bonds. Lead, for example, forms covalent bonds with the sulfhydryl side chain of cysteine in proteins. Ferrochelatase, an enzyme involved in heme synthesis (see p. 279), is irreversibly inhibited by lead. [Note: An important group of irreversible inhibitors are the mechanism-based inhibitors that are converted by the enzyme itself to a form that covalently links to the enzyme, thereby inhibiting it. They also are referred to as “suicide” inhibitors.] Reversible inhibitors bind to enzymes through noncovalent bonds and, thus, dilution of the enzyme–inhibitor complex results in dissociation of the reversibly bound inhibitor and recovery of enzyme activity. The two most commonly encountered types of reversible inhibition are competitive and noncompetitive. | Biochemistry_Lippinco. VII. ENZYME INHIBITION Any substance that can decrease the velocity of an enzyme-catalyzed reaction is called an inhibitor. Inhibitors can be reversible or irreversible. Irreversible inhibitors bind to enzymes through covalent bonds. Lead, for example, forms covalent bonds with the sulfhydryl side chain of cysteine in proteins. Ferrochelatase, an enzyme involved in heme synthesis (see p. 279), is irreversibly inhibited by lead. [Note: An important group of irreversible inhibitors are the mechanism-based inhibitors that are converted by the enzyme itself to a form that covalently links to the enzyme, thereby inhibiting it. They also are referred to as “suicide” inhibitors.] Reversible inhibitors bind to enzymes through noncovalent bonds and, thus, dilution of the enzyme–inhibitor complex results in dissociation of the reversibly bound inhibitor and recovery of enzyme activity. The two most commonly encountered types of reversible inhibition are competitive and noncompetitive. |
Biochemistry_Lippincott_202 | Biochemistry_Lippinco | A. Competitive inhibition This type of inhibition occurs when the inhibitor binds reversibly to the same site that the substrate would normally occupy and, therefore, competes with the substrate for that site. 1. Effect on Vmax: The effect of a competitive inhibitor is reversed by increasing the concentration of substrate. At a sufficiently high [S], the 2. Effect on Km: A competitive inhibitor increases the apparent Km for a given substrate. This means that, in the presence of a competitive inhibitor, more substrate is needed to achieve one half Vmax. 3. Effect on the Lineweaver-Burk plot: Competitive inhibition shows a characteristic Lineweaver-Burk plot in which the plots of the inhibited and uninhibited reactions intersect on the y axis at 1/Vmax (Vmax is reaction velocity reaches the Vmax observed in the absence of inhibitor, that is, Vmax is unchanged (Fig. 5.12). | Biochemistry_Lippinco. A. Competitive inhibition This type of inhibition occurs when the inhibitor binds reversibly to the same site that the substrate would normally occupy and, therefore, competes with the substrate for that site. 1. Effect on Vmax: The effect of a competitive inhibitor is reversed by increasing the concentration of substrate. At a sufficiently high [S], the 2. Effect on Km: A competitive inhibitor increases the apparent Km for a given substrate. This means that, in the presence of a competitive inhibitor, more substrate is needed to achieve one half Vmax. 3. Effect on the Lineweaver-Burk plot: Competitive inhibition shows a characteristic Lineweaver-Burk plot in which the plots of the inhibited and uninhibited reactions intersect on the y axis at 1/Vmax (Vmax is reaction velocity reaches the Vmax observed in the absence of inhibitor, that is, Vmax is unchanged (Fig. 5.12). |
Biochemistry_Lippincott_203 | Biochemistry_Lippinco | unchanged). The inhibited and uninhibited reactions show different x-axis intercepts, indicating that the apparent Km is increased in the presence of the competitive inhibitor because − 1/Km moves closer to zero from a negative value (see Fig. 5.12). [Note: An important group of competitive inhibitors are the transition state analogs, stable molecules that approximate the structure of the transition state, and, therefore, bind the enzyme more tightly than does the substrate.] 4. Statin drugs as examples of competitive inhibitors: This group of antihyperlipidemic agents competitively inhibits the rate-limiting (slowest) step in cholesterol biosynthesis. This reaction is catalyzed by hydroxymethylglutaryl coenzyme A reductase (HMG CoA reductase; see p. 221). Statins, such as atorvastatin (Lipitor) and pravastatin (Pravachol), are structural analogs of the natural substrate for this enzyme and compete effectively to inhibit HMG CoA reductase. By doing so, they inhibit de novo cholesterol | Biochemistry_Lippinco. unchanged). The inhibited and uninhibited reactions show different x-axis intercepts, indicating that the apparent Km is increased in the presence of the competitive inhibitor because − 1/Km moves closer to zero from a negative value (see Fig. 5.12). [Note: An important group of competitive inhibitors are the transition state analogs, stable molecules that approximate the structure of the transition state, and, therefore, bind the enzyme more tightly than does the substrate.] 4. Statin drugs as examples of competitive inhibitors: This group of antihyperlipidemic agents competitively inhibits the rate-limiting (slowest) step in cholesterol biosynthesis. This reaction is catalyzed by hydroxymethylglutaryl coenzyme A reductase (HMG CoA reductase; see p. 221). Statins, such as atorvastatin (Lipitor) and pravastatin (Pravachol), are structural analogs of the natural substrate for this enzyme and compete effectively to inhibit HMG CoA reductase. By doing so, they inhibit de novo cholesterol |
Biochemistry_Lippincott_204 | Biochemistry_Lippinco | and pravastatin (Pravachol), are structural analogs of the natural substrate for this enzyme and compete effectively to inhibit HMG CoA reductase. By doing so, they inhibit de novo cholesterol synthesis, thereby lowering plasma cholesterol levels (Fig. 5.13). | Biochemistry_Lippinco. and pravastatin (Pravachol), are structural analogs of the natural substrate for this enzyme and compete effectively to inhibit HMG CoA reductase. By doing so, they inhibit de novo cholesterol synthesis, thereby lowering plasma cholesterol levels (Fig. 5.13). |
Biochemistry_Lippincott_205 | Biochemistry_Lippinco | B. Noncompetitive inhibition This type of inhibition is recognized by its characteristic effect on Vmax (Fig. 5.14). Noncompetitive inhibition occurs when the inhibitor and substrate bind at different sites on the enzyme. The noncompetitive inhibitor can bind either free enzyme or the enzyme–substrate complex, thereby preventing the reaction from occurring (Fig. 5.15). 1. Effect on Vmax: Noncompetitive inhibition cannot be overcome by increasing the concentration of substrate. Therefore, noncompetitive inhibitors decrease the apparent Vmax of the reaction. 2. Effect on Km: Noncompetitive inhibitors do not interfere with the binding of substrate to enzyme. Therefore, the enzyme shows the same Km in the presence or absence of the noncompetitive inhibitor, that is, Km is unchanged. 3. | Biochemistry_Lippinco. B. Noncompetitive inhibition This type of inhibition is recognized by its characteristic effect on Vmax (Fig. 5.14). Noncompetitive inhibition occurs when the inhibitor and substrate bind at different sites on the enzyme. The noncompetitive inhibitor can bind either free enzyme or the enzyme–substrate complex, thereby preventing the reaction from occurring (Fig. 5.15). 1. Effect on Vmax: Noncompetitive inhibition cannot be overcome by increasing the concentration of substrate. Therefore, noncompetitive inhibitors decrease the apparent Vmax of the reaction. 2. Effect on Km: Noncompetitive inhibitors do not interfere with the binding of substrate to enzyme. Therefore, the enzyme shows the same Km in the presence or absence of the noncompetitive inhibitor, that is, Km is unchanged. 3. |
Biochemistry_Lippincott_206 | Biochemistry_Lippinco | 3. Effect on Lineweaver-Burk plot: Noncompetitive inhibition is readily differentiated from competitive inhibition by plotting 1/vo versus 1/[S] and noting that the apparent Vmax decreases in the presence of a noncompetitive inhibitor, whereas Km is unchanged (see Fig. 5.14). [Note: Oxypurinol, a metabolite of the prodrug allopurinol, is a noncompetitive inhibitor of xanthine oxidase, an enzyme of purine degradation (see p. 301).] C. Enzyme inhibitors as drugs | Biochemistry_Lippinco. 3. Effect on Lineweaver-Burk plot: Noncompetitive inhibition is readily differentiated from competitive inhibition by plotting 1/vo versus 1/[S] and noting that the apparent Vmax decreases in the presence of a noncompetitive inhibitor, whereas Km is unchanged (see Fig. 5.14). [Note: Oxypurinol, a metabolite of the prodrug allopurinol, is a noncompetitive inhibitor of xanthine oxidase, an enzyme of purine degradation (see p. 301).] C. Enzyme inhibitors as drugs |
Biochemistry_Lippincott_207 | Biochemistry_Lippinco | C. Enzyme inhibitors as drugs At least half of the ten most commonly prescribed drugs in the United States act as enzyme inhibitors. For example, the widely prescribed βlactam antibiotics, such as penicillin and amoxicillin, act by inhibiting enzymes involved in bacterial cell wall synthesis. Drugs may also act by inhibiting extracellular reactions. This is illustrated by angiotensinconverting enzyme (ACE) inhibitors. They lower blood pressure by blocking plasma ACE that cleaves angiotensin I to form the potent vasoconstrictor, angiotensin II. These drugs, which include captopril, enalapril, and lisinopril, cause vasodilation and, therefore, a reduction in blood pressure. Aspirin, a nonprescription drug, irreversibly inhibits prostaglandin and thromboxane synthesis by inhibiting cyclooxygenase (see p. 214). VIII. ENZYME REGULATION | Biochemistry_Lippinco. C. Enzyme inhibitors as drugs At least half of the ten most commonly prescribed drugs in the United States act as enzyme inhibitors. For example, the widely prescribed βlactam antibiotics, such as penicillin and amoxicillin, act by inhibiting enzymes involved in bacterial cell wall synthesis. Drugs may also act by inhibiting extracellular reactions. This is illustrated by angiotensinconverting enzyme (ACE) inhibitors. They lower blood pressure by blocking plasma ACE that cleaves angiotensin I to form the potent vasoconstrictor, angiotensin II. These drugs, which include captopril, enalapril, and lisinopril, cause vasodilation and, therefore, a reduction in blood pressure. Aspirin, a nonprescription drug, irreversibly inhibits prostaglandin and thromboxane synthesis by inhibiting cyclooxygenase (see p. 214). VIII. ENZYME REGULATION |
Biochemistry_Lippincott_208 | Biochemistry_Lippinco | VIII. ENZYME REGULATION The regulation of the reaction velocity of enzymes is essential if an organism is to coordinate its numerous metabolic processes. The rates of most enzymes are responsive to changes in substrate concentration, because the intracellular level of many substrates is in the range of the Km. Thus, an increase in substrate concentration prompts an increase in reaction rate, which tends to return the concentration of substrate toward normal. In addition, some enzymes with specialized regulatory functions respond to allosteric effectors and/or covalent modification or they show altered rates of enzyme synthesis (or degradation) when physiologic conditions are changed. A. Allosteric enzymes | Biochemistry_Lippinco. VIII. ENZYME REGULATION The regulation of the reaction velocity of enzymes is essential if an organism is to coordinate its numerous metabolic processes. The rates of most enzymes are responsive to changes in substrate concentration, because the intracellular level of many substrates is in the range of the Km. Thus, an increase in substrate concentration prompts an increase in reaction rate, which tends to return the concentration of substrate toward normal. In addition, some enzymes with specialized regulatory functions respond to allosteric effectors and/or covalent modification or they show altered rates of enzyme synthesis (or degradation) when physiologic conditions are changed. A. Allosteric enzymes |
Biochemistry_Lippincott_209 | Biochemistry_Lippinco | A. Allosteric enzymes Allosteric enzymes are regulated by molecules called effectors that bind noncovalently at a site other than the active site. These enzymes are almost always composed of multiple subunits, and the regulatory (allosteric) site that binds the effector is distinct from the substrate-binding site and may be located on a subunit that is not itself catalytic. Effectors that inhibit enzyme activity are termed negative effectors, whereas those that increase enzyme activity are called positive effectors. Positive and negative effectors can affect the affinity of the enzyme for its substrate (K0.5), modify the maximal catalytic activity of the enzyme (Vmax), or both (Fig. 5.16). [Note: Allosteric enzymes frequently catalyze the committed step, often the rate-limiting step, early in a pathway.] 1. | Biochemistry_Lippinco. A. Allosteric enzymes Allosteric enzymes are regulated by molecules called effectors that bind noncovalently at a site other than the active site. These enzymes are almost always composed of multiple subunits, and the regulatory (allosteric) site that binds the effector is distinct from the substrate-binding site and may be located on a subunit that is not itself catalytic. Effectors that inhibit enzyme activity are termed negative effectors, whereas those that increase enzyme activity are called positive effectors. Positive and negative effectors can affect the affinity of the enzyme for its substrate (K0.5), modify the maximal catalytic activity of the enzyme (Vmax), or both (Fig. 5.16). [Note: Allosteric enzymes frequently catalyze the committed step, often the rate-limiting step, early in a pathway.] 1. |
Biochemistry_Lippincott_210 | Biochemistry_Lippinco | Homotropic effectors: When the substrate itself serves as an effector, the effect is said to be homotropic. Most often, an allosteric substrate functions as a positive effector. In such a case, the presence of a substrate molecule at one site on the enzyme enhances the catalytic properties of the other substrate-binding sites. That is, their binding sites exhibit cooperativity. These enzymes show a sigmoidal curve when vo is plotted against substrate concentration, as shown in Figure 5.16. This contrasts with the hyperbolic curve characteristic of enzymes following Michaelis-Menten kinetics, as previously discussed. [Note: The concept of cooperativity of substrate binding is analogous to the binding of oxygen to hemoglobin (see p. 29).] 2. | Biochemistry_Lippinco. Homotropic effectors: When the substrate itself serves as an effector, the effect is said to be homotropic. Most often, an allosteric substrate functions as a positive effector. In such a case, the presence of a substrate molecule at one site on the enzyme enhances the catalytic properties of the other substrate-binding sites. That is, their binding sites exhibit cooperativity. These enzymes show a sigmoidal curve when vo is plotted against substrate concentration, as shown in Figure 5.16. This contrasts with the hyperbolic curve characteristic of enzymes following Michaelis-Menten kinetics, as previously discussed. [Note: The concept of cooperativity of substrate binding is analogous to the binding of oxygen to hemoglobin (see p. 29).] 2. |
Biochemistry_Lippincott_211 | Biochemistry_Lippinco | Heterotropic effectors: The effector may be different from the substrate, in which case the effect is said to be heterotropic. For example, consider the feedback inhibition shown in Figure 5.17. The enzyme that converts D to E has an allosteric site that binds the end product, G. If the concentration of G increases (for example, because it is not used as rapidly as it is synthesized), the first irreversible step unique to the pathway is typically inhibited. Feedback inhibition provides the cell with appropriate amounts of a product it needs by regulating the flow of substrate molecules through the pathway that synthesizes that product. Heterotropic effectors are commonly encountered. For example, the glycolytic enzyme phosphofructokinase-1 is allosterically inhibited by citrate, which is not a substrate for the enzyme (see p. 99). Figure5.17Feedbackinhibitionofametabolicpathway. B. Covalent modification | Biochemistry_Lippinco. Heterotropic effectors: The effector may be different from the substrate, in which case the effect is said to be heterotropic. For example, consider the feedback inhibition shown in Figure 5.17. The enzyme that converts D to E has an allosteric site that binds the end product, G. If the concentration of G increases (for example, because it is not used as rapidly as it is synthesized), the first irreversible step unique to the pathway is typically inhibited. Feedback inhibition provides the cell with appropriate amounts of a product it needs by regulating the flow of substrate molecules through the pathway that synthesizes that product. Heterotropic effectors are commonly encountered. For example, the glycolytic enzyme phosphofructokinase-1 is allosterically inhibited by citrate, which is not a substrate for the enzyme (see p. 99). Figure5.17Feedbackinhibitionofametabolicpathway. B. Covalent modification |
Biochemistry_Lippincott_212 | Biochemistry_Lippinco | Figure5.17Feedbackinhibitionofametabolicpathway. B. Covalent modification Many enzymes are regulated by covalent modification, most often by the addition or removal of phosphate groups from specific serine, threonine, or tyrosine residues of the enzyme. Protein phosphorylation is recognized as one of the primary ways in which cellular processes are regulated. 1. Phosphorylation and dephosphorylation: Phosphorylation reactions are catalyzed by a family of enzymes called protein kinases that use ATP as the phosphate donor. Phosphate groups are cleaved from phosphorylated enzymes by the action of phosphoprotein phosphatases (Fig. 5.18). 2. | Biochemistry_Lippinco. Figure5.17Feedbackinhibitionofametabolicpathway. B. Covalent modification Many enzymes are regulated by covalent modification, most often by the addition or removal of phosphate groups from specific serine, threonine, or tyrosine residues of the enzyme. Protein phosphorylation is recognized as one of the primary ways in which cellular processes are regulated. 1. Phosphorylation and dephosphorylation: Phosphorylation reactions are catalyzed by a family of enzymes called protein kinases that use ATP as the phosphate donor. Phosphate groups are cleaved from phosphorylated enzymes by the action of phosphoprotein phosphatases (Fig. 5.18). 2. |
Biochemistry_Lippincott_213 | Biochemistry_Lippinco | 2. Enzyme response to phosphorylation: Depending on the specific enzyme, the phosphorylated form may be more or less active than the unphosphorylated enzyme. For example, hormone-mediated phosphorylation of glycogen phosphorylase (an enzyme that degrades glycogen) increases activity, whereas phosphorylation of glycogen synthase (an enzyme that synthesizes glycogen) decreases activity (see p. 132). C. Enzyme synthesis | Biochemistry_Lippinco. 2. Enzyme response to phosphorylation: Depending on the specific enzyme, the phosphorylated form may be more or less active than the unphosphorylated enzyme. For example, hormone-mediated phosphorylation of glycogen phosphorylase (an enzyme that degrades glycogen) increases activity, whereas phosphorylation of glycogen synthase (an enzyme that synthesizes glycogen) decreases activity (see p. 132). C. Enzyme synthesis |
Biochemistry_Lippincott_214 | Biochemistry_Lippinco | The regulatory mechanisms described above modify the activity of existing enzyme molecules. However, cells can also regulate the amount of enzyme present by altering the rate of enzyme degradation or, more typically, the rate of enzyme synthesis. The increase (induction) or decrease (repression) of enzyme synthesis leads to an alteration in the total population of active sites. Enzymes subject to regulation of synthesis are often those that are needed at only one stage of development or under selected physiologic conditions. For example, elevated levels of insulin as a result of high blood glucose levels cause an increase in the synthesis of key enzymes involved in glucose metabolism (see p. 105). In contrast, enzymes that are in constant use are usually not regulated by altering the rate of enzyme synthesis. Alterations in enzyme levels as a result of induction or repression of protein synthesis are slow (hours to days), compared with allosterically or covalently regulated changes in | Biochemistry_Lippinco. The regulatory mechanisms described above modify the activity of existing enzyme molecules. However, cells can also regulate the amount of enzyme present by altering the rate of enzyme degradation or, more typically, the rate of enzyme synthesis. The increase (induction) or decrease (repression) of enzyme synthesis leads to an alteration in the total population of active sites. Enzymes subject to regulation of synthesis are often those that are needed at only one stage of development or under selected physiologic conditions. For example, elevated levels of insulin as a result of high blood glucose levels cause an increase in the synthesis of key enzymes involved in glucose metabolism (see p. 105). In contrast, enzymes that are in constant use are usually not regulated by altering the rate of enzyme synthesis. Alterations in enzyme levels as a result of induction or repression of protein synthesis are slow (hours to days), compared with allosterically or covalently regulated changes in |
Biochemistry_Lippincott_215 | Biochemistry_Lippinco | of enzyme synthesis. Alterations in enzyme levels as a result of induction or repression of protein synthesis are slow (hours to days), compared with allosterically or covalently regulated changes in enzyme activity, which occur in seconds to minutes. Figure 5.19 summarizes the common ways that enzyme activity is regulated. | Biochemistry_Lippinco. of enzyme synthesis. Alterations in enzyme levels as a result of induction or repression of protein synthesis are slow (hours to days), compared with allosterically or covalently regulated changes in enzyme activity, which occur in seconds to minutes. Figure 5.19 summarizes the common ways that enzyme activity is regulated. |
Biochemistry_Lippincott_216 | Biochemistry_Lippinco | IX. Enzymes in Clinical Diagnosis Plasma enzymes can be classified into two major groups. First, a relatively small group of enzymes are actively secreted into the blood by certain cell types. For example, the liver secretes zymogens (inactive precursors) of the enzymes involved in blood coagulation. Second, a large number of enzyme species are released from cells during normal cell turnover. These enzymes almost always function intracellularly and have no physiologic use in the plasma. In healthy individuals, the levels of these enzymes are fairly constant and represent a steady state in which the rate of release from damaged cells into the plasma is balanced by an equal rate of removal from the plasma. Increased plasma levels of these enzymes may indicate tissue damage (Fig. 5.20). (B) cells. | Biochemistry_Lippinco. IX. Enzymes in Clinical Diagnosis Plasma enzymes can be classified into two major groups. First, a relatively small group of enzymes are actively secreted into the blood by certain cell types. For example, the liver secretes zymogens (inactive precursors) of the enzymes involved in blood coagulation. Second, a large number of enzyme species are released from cells during normal cell turnover. These enzymes almost always function intracellularly and have no physiologic use in the plasma. In healthy individuals, the levels of these enzymes are fairly constant and represent a steady state in which the rate of release from damaged cells into the plasma is balanced by an equal rate of removal from the plasma. Increased plasma levels of these enzymes may indicate tissue damage (Fig. 5.20). (B) cells. |
Biochemistry_Lippincott_217 | Biochemistry_Lippinco | (B) cells. Plasma is the fluid, noncellular part of blood. Laboratory assays of enzyme activity most often use serum, which is obtained by centrifugation of whole blood after it has been allowed to coagulate. Plasma is a physiologic fluid, whereas serum is prepared in the laboratory. A. Plasma enzyme levels in disease states Many diseases that cause tissue damage result in an increased release of intracellular enzymes into the plasma. The activities of many of these enzymes are routinely determined for diagnostic purposes in diseases of the heart, liver, skeletal muscle, and other tissues. The level of specific enzyme activity in the plasma frequently correlates with the extent of tissue damage. Therefore, determining the degree of elevation of a particular enzyme activity in the plasma is often useful in evaluating the prognosis for the patient. B. Plasma enzymes as diagnostic tools | Biochemistry_Lippinco. (B) cells. Plasma is the fluid, noncellular part of blood. Laboratory assays of enzyme activity most often use serum, which is obtained by centrifugation of whole blood after it has been allowed to coagulate. Plasma is a physiologic fluid, whereas serum is prepared in the laboratory. A. Plasma enzyme levels in disease states Many diseases that cause tissue damage result in an increased release of intracellular enzymes into the plasma. The activities of many of these enzymes are routinely determined for diagnostic purposes in diseases of the heart, liver, skeletal muscle, and other tissues. The level of specific enzyme activity in the plasma frequently correlates with the extent of tissue damage. Therefore, determining the degree of elevation of a particular enzyme activity in the plasma is often useful in evaluating the prognosis for the patient. B. Plasma enzymes as diagnostic tools |
Biochemistry_Lippincott_218 | Biochemistry_Lippinco | B. Plasma enzymes as diagnostic tools Some enzymes show relatively high activity in only one or a few tissues. Therefore, the presence of increased levels of these enzymes in plasma reflects damage to the corresponding tissue. For example, the enzyme alanine aminotransferase (ALT; see p. 251) is abundant in the liver. The appearance of elevated levels of ALT in plasma signals possible damage to hepatic tissue. [Note: Measurement of ALT is part of the liver function test panel.] Increases in plasma levels of enzymes with a wide tissue distribution provide a less specific indication of the site of cellular injury and limits their diagnostic value. C. Isoenzymes and heart disease | Biochemistry_Lippinco. B. Plasma enzymes as diagnostic tools Some enzymes show relatively high activity in only one or a few tissues. Therefore, the presence of increased levels of these enzymes in plasma reflects damage to the corresponding tissue. For example, the enzyme alanine aminotransferase (ALT; see p. 251) is abundant in the liver. The appearance of elevated levels of ALT in plasma signals possible damage to hepatic tissue. [Note: Measurement of ALT is part of the liver function test panel.] Increases in plasma levels of enzymes with a wide tissue distribution provide a less specific indication of the site of cellular injury and limits their diagnostic value. C. Isoenzymes and heart disease |
Biochemistry_Lippincott_219 | Biochemistry_Lippinco | C. Isoenzymes and heart disease Isoenzymes (also called isozymes) are enzymes that catalyze the same reaction. However, they do not necessarily have the same physical properties because of genetically determined differences in amino acid sequence. For this reason, isoenzymes may contain different numbers of charged amino acids, which allows electrophoresis (the movement of charged particles in an electric field) to separate them (Fig. 5.21). Different organs commonly contain characteristic proportions of different isoenzymes. The pattern of isoenzymes found in the plasma may, therefore, serve as a means of identifying the site of tissue damage. For example, the plasma levels of creatine kinase (CK) are commonly determined in the diagnosis of myocardial infarction (MI). They are particularly useful when the electrocardiogram (ECG) is difficult to interpret such as when there have been previous episodes of heart disease. 1. | Biochemistry_Lippinco. C. Isoenzymes and heart disease Isoenzymes (also called isozymes) are enzymes that catalyze the same reaction. However, they do not necessarily have the same physical properties because of genetically determined differences in amino acid sequence. For this reason, isoenzymes may contain different numbers of charged amino acids, which allows electrophoresis (the movement of charged particles in an electric field) to separate them (Fig. 5.21). Different organs commonly contain characteristic proportions of different isoenzymes. The pattern of isoenzymes found in the plasma may, therefore, serve as a means of identifying the site of tissue damage. For example, the plasma levels of creatine kinase (CK) are commonly determined in the diagnosis of myocardial infarction (MI). They are particularly useful when the electrocardiogram (ECG) is difficult to interpret such as when there have been previous episodes of heart disease. 1. |
Biochemistry_Lippincott_220 | Biochemistry_Lippinco | 1. Isoenzyme quaternary structure: Many isoenzymes contain different subunits in various combinations. For example, CK occurs as three isoenzymes. Each isoenzyme is a dimer composed of two polypeptides (called B and M subunits) associated in one of three combinations: CK1 = BB, CK2 = MB, and CK3 = MM. Each CK isoenzyme shows a characteristic electrophoretic mobility (see Fig. 5.21). [Note: Virtually all CK in the brain is the BB isoform, whereas it is MM in skeletal muscle. In cardiac muscle, about one third is MB with the rest as MM.] 2. | Biochemistry_Lippinco. 1. Isoenzyme quaternary structure: Many isoenzymes contain different subunits in various combinations. For example, CK occurs as three isoenzymes. Each isoenzyme is a dimer composed of two polypeptides (called B and M subunits) associated in one of three combinations: CK1 = BB, CK2 = MB, and CK3 = MM. Each CK isoenzyme shows a characteristic electrophoretic mobility (see Fig. 5.21). [Note: Virtually all CK in the brain is the BB isoform, whereas it is MM in skeletal muscle. In cardiac muscle, about one third is MB with the rest as MM.] 2. |
Biochemistry_Lippincott_221 | Biochemistry_Lippinco | Diagnosis of myocardial infarction: Measurement of blood levels of proteins with cardiac specificity (biomarkers) is used in the diagnosis of MI. Myocardial muscle is the only tissue that contains >5% of the total CK activity as the CK2 (MB) isoenzyme. Appearance of this hybrid isoenzyme in plasma is virtually specific for infarction of the myocardium. Following an acute MI, CK2 appears in plasma within 4–8 hours following onset of chest pain, reaches a peak of activity at ~24 hours, and returns to baseline after 48–72 hours (Fig. 5.22). Troponins T (TnT) and I (TnI) are regulatory proteins involved in muscle contractility. Cardiac-specific isoforms (cTn) are released into the plasma in response to cardiac damage. They are highly sensitive and specific for damage to cardiac tissue. cTn appear in plasma within 4–6 hours after an MI, peak in 24–36 hours, and remain elevated for 3–10 days. Elevated cTn, in combination with the clinical presentation and characteristic changes in the ECG, | Biochemistry_Lippinco. Diagnosis of myocardial infarction: Measurement of blood levels of proteins with cardiac specificity (biomarkers) is used in the diagnosis of MI. Myocardial muscle is the only tissue that contains >5% of the total CK activity as the CK2 (MB) isoenzyme. Appearance of this hybrid isoenzyme in plasma is virtually specific for infarction of the myocardium. Following an acute MI, CK2 appears in plasma within 4–8 hours following onset of chest pain, reaches a peak of activity at ~24 hours, and returns to baseline after 48–72 hours (Fig. 5.22). Troponins T (TnT) and I (TnI) are regulatory proteins involved in muscle contractility. Cardiac-specific isoforms (cTn) are released into the plasma in response to cardiac damage. They are highly sensitive and specific for damage to cardiac tissue. cTn appear in plasma within 4–6 hours after an MI, peak in 24–36 hours, and remain elevated for 3–10 days. Elevated cTn, in combination with the clinical presentation and characteristic changes in the ECG, |
Biochemistry_Lippincott_222 | Biochemistry_Lippinco | in plasma within 4–6 hours after an MI, peak in 24–36 hours, and remain elevated for 3–10 days. Elevated cTn, in combination with the clinical presentation and characteristic changes in the ECG, are currently considered the “gold standard” in the diagnosis of an MI. | Biochemistry_Lippinco. in plasma within 4–6 hours after an MI, peak in 24–36 hours, and remain elevated for 3–10 days. Elevated cTn, in combination with the clinical presentation and characteristic changes in the ECG, are currently considered the “gold standard” in the diagnosis of an MI. |
Biochemistry_Lippincott_223 | Biochemistry_Lippinco | X. CHAPTER SUMMARY | Biochemistry_Lippinco. X. CHAPTER SUMMARY |
Biochemistry_Lippincott_224 | Biochemistry_Lippinco | Enzymes are protein catalysts that increase the velocity of a chemical reaction by lowering the energy of the transition state (Fig. 5.23). They are not consumed during the reaction. Enzyme molecules contain a special cleft called the active site, which contains amino acid side chains that participate in substrate binding and catalysis. The active site binds the substrate, forming an enzyme–substrate (ES) complex. Binding is thought to cause a conformational change in the enzyme (induced fit) that allows catalysis. ES is converted to enzyme and product. An enzyme allows a reaction to proceed rapidly under conditions prevailing in the cell by providing an alternate reaction pathway with a lower activation energy (Ea). Because the enzyme does not change the free energies of the reactants or products, it does not change the equilibrium of the reaction. Most enzymes show Michaelis-Menten kinetics, and a plot of the initial reaction velocity (vo) against substrate concentration ([S]) has a | Biochemistry_Lippinco. Enzymes are protein catalysts that increase the velocity of a chemical reaction by lowering the energy of the transition state (Fig. 5.23). They are not consumed during the reaction. Enzyme molecules contain a special cleft called the active site, which contains amino acid side chains that participate in substrate binding and catalysis. The active site binds the substrate, forming an enzyme–substrate (ES) complex. Binding is thought to cause a conformational change in the enzyme (induced fit) that allows catalysis. ES is converted to enzyme and product. An enzyme allows a reaction to proceed rapidly under conditions prevailing in the cell by providing an alternate reaction pathway with a lower activation energy (Ea). Because the enzyme does not change the free energies of the reactants or products, it does not change the equilibrium of the reaction. Most enzymes show Michaelis-Menten kinetics, and a plot of the initial reaction velocity (vo) against substrate concentration ([S]) has a |
Biochemistry_Lippincott_225 | Biochemistry_Lippinco | products, it does not change the equilibrium of the reaction. Most enzymes show Michaelis-Menten kinetics, and a plot of the initial reaction velocity (vo) against substrate concentration ([S]) has a hyperbolic shape similar to the oxygen-dissociation curve of myoglobin. A Lineweaver-Burk plot of 1/v and 1/[S] allows determination of Vmax (maximal velocity) and Km (Michaelis constant, which reflects affinity for substrate). Any substance that can decrease the velocity of an enzyme-catalyzed reaction is called an inhibitor. The two most common types of reversible inhibition are competitive (which increases the apparent Km) and noncompetitive (which decreases the apparent Vmax). In contrast, the multisubunit allosteric enzymes show a sigmoidal curve similar in shape to the oxygen-dissociation curve of hemoglobin. They typically catalyze the committed step of a pathway. Allosteric enzymes are regulated by molecules called effectors that bind noncovalently at a site other than the active | Biochemistry_Lippinco. products, it does not change the equilibrium of the reaction. Most enzymes show Michaelis-Menten kinetics, and a plot of the initial reaction velocity (vo) against substrate concentration ([S]) has a hyperbolic shape similar to the oxygen-dissociation curve of myoglobin. A Lineweaver-Burk plot of 1/v and 1/[S] allows determination of Vmax (maximal velocity) and Km (Michaelis constant, which reflects affinity for substrate). Any substance that can decrease the velocity of an enzyme-catalyzed reaction is called an inhibitor. The two most common types of reversible inhibition are competitive (which increases the apparent Km) and noncompetitive (which decreases the apparent Vmax). In contrast, the multisubunit allosteric enzymes show a sigmoidal curve similar in shape to the oxygen-dissociation curve of hemoglobin. They typically catalyze the committed step of a pathway. Allosteric enzymes are regulated by molecules called effectors that bind noncovalently at a site other than the active |
Biochemistry_Lippincott_226 | Biochemistry_Lippinco | curve of hemoglobin. They typically catalyze the committed step of a pathway. Allosteric enzymes are regulated by molecules called effectors that bind noncovalently at a site other than the active site. Effectors can be either positive (increase enzyme activity) or negative (decrease enzyme activity). An allosteric effector can alter the affinity of the enzyme for its substrate (K0.5), the maximal catalytic activity of the enzyme (Vmax), or both. Enzymes can also be regulated by covalent modification and by changes in the rate of synthesis or degradation. | Biochemistry_Lippinco. curve of hemoglobin. They typically catalyze the committed step of a pathway. Allosteric enzymes are regulated by molecules called effectors that bind noncovalently at a site other than the active site. Effectors can be either positive (increase enzyme activity) or negative (decrease enzyme activity). An allosteric effector can alter the affinity of the enzyme for its substrate (K0.5), the maximal catalytic activity of the enzyme (Vmax), or both. Enzymes can also be regulated by covalent modification and by changes in the rate of synthesis or degradation. |
Biochemistry_Lippincott_227 | Biochemistry_Lippinco | Enzymes have diagnostic and therapeutic value in medicine. Choose the ONE best answer. .1. In cases of ethylene glycol poisoning and its characteristic metabolic acidosis, treatment involves correction of the acidosis, removal of any remaining ethylene glycol, and administration of an inhibitor of alcohol dehydrogenase (ADH), the enzyme that oxidizes ethylene glycol to the organic acids that cause the acidosis. Ethanol (grain alcohol) frequently is the inhibitor given to treat ethylene glycol poisoning. Results of experiments using ADH with and without ethanol are shown to the right. Based on these data, what type of inhibition is caused by the ethanol? A. Competitive B. Feedback C. Irreversible D. Noncompetitive | Biochemistry_Lippinco. Enzymes have diagnostic and therapeutic value in medicine. Choose the ONE best answer. .1. In cases of ethylene glycol poisoning and its characteristic metabolic acidosis, treatment involves correction of the acidosis, removal of any remaining ethylene glycol, and administration of an inhibitor of alcohol dehydrogenase (ADH), the enzyme that oxidizes ethylene glycol to the organic acids that cause the acidosis. Ethanol (grain alcohol) frequently is the inhibitor given to treat ethylene glycol poisoning. Results of experiments using ADH with and without ethanol are shown to the right. Based on these data, what type of inhibition is caused by the ethanol? A. Competitive B. Feedback C. Irreversible D. Noncompetitive |
Biochemistry_Lippincott_228 | Biochemistry_Lippinco | A. Competitive B. Feedback C. Irreversible D. Noncompetitive Correct answer = A. A competitive inhibitor increases the apparent Km for a given substrate. This means that, in the presence of a competitive inhibitor, more substrate is needed to achieve one half Vmax. The effect of a competitive inhibitor is reversed by increasing substrate concentration ([S]). At a sufficiently high [S], the reaction velocity reaches the Vmax observed in the absence of inhibitor. .2. Alcohol dehydrogenase (ADH) requires oxidized nicotinamide adenine dinucleotide (NAD+) for catalytic activity. In the reaction catalyzed by ADH, an alcohol is oxidized to an aldehyde as NAD+ is reduced to NADH and dissociates from the enzyme. The NAD+ is functioning as a/an: A. apoenzyme. B. coenzyme–cosubstrate. C. coenzyme–prosthetic group. D. cofactor. E. heterotropic effector. | Biochemistry_Lippinco. A. Competitive B. Feedback C. Irreversible D. Noncompetitive Correct answer = A. A competitive inhibitor increases the apparent Km for a given substrate. This means that, in the presence of a competitive inhibitor, more substrate is needed to achieve one half Vmax. The effect of a competitive inhibitor is reversed by increasing substrate concentration ([S]). At a sufficiently high [S], the reaction velocity reaches the Vmax observed in the absence of inhibitor. .2. Alcohol dehydrogenase (ADH) requires oxidized nicotinamide adenine dinucleotide (NAD+) for catalytic activity. In the reaction catalyzed by ADH, an alcohol is oxidized to an aldehyde as NAD+ is reduced to NADH and dissociates from the enzyme. The NAD+ is functioning as a/an: A. apoenzyme. B. coenzyme–cosubstrate. C. coenzyme–prosthetic group. D. cofactor. E. heterotropic effector. |
Biochemistry_Lippincott_229 | Biochemistry_Lippinco | A. apoenzyme. B. coenzyme–cosubstrate. C. coenzyme–prosthetic group. D. cofactor. E. heterotropic effector. Correct answer = B. A Coenzymes–cosubstrates are small organic molecules that associate transiently with an enzyme and leave the enzyme in a changed form. Coenzyme–prosthetic groups are small organic molecules that associate permanently with an enzyme and are returned to their original form on the enzyme. Cofactors are metal ions. Heterotropic effectors are not substrates. For Questions 5.3 and 5.4, use the graph below that shows the changes in free energy when a reactant is converted to a product in the presence and absence of an enzyme. Select the letter that best represents: .3. the activation energy of the catalyzed forward reaction. .4. the free energy of the reaction. | Biochemistry_Lippinco. A. apoenzyme. B. coenzyme–cosubstrate. C. coenzyme–prosthetic group. D. cofactor. E. heterotropic effector. Correct answer = B. A Coenzymes–cosubstrates are small organic molecules that associate transiently with an enzyme and leave the enzyme in a changed form. Coenzyme–prosthetic groups are small organic molecules that associate permanently with an enzyme and are returned to their original form on the enzyme. Cofactors are metal ions. Heterotropic effectors are not substrates. For Questions 5.3 and 5.4, use the graph below that shows the changes in free energy when a reactant is converted to a product in the presence and absence of an enzyme. Select the letter that best represents: .3. the activation energy of the catalyzed forward reaction. .4. the free energy of the reaction. |
Biochemistry_Lippincott_230 | Biochemistry_Lippinco | Correct answers = B; D. Enzymes (protein catalysts) provide an alternate reaction pathway with a lower activation energy. However, they do not change the free energy of the reactant or product. A is the activation energy of the uncatalyzed reaction. C is the activation energy of the catalyzed reverse reaction. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW | Biochemistry_Lippinco. Correct answers = B; D. Enzymes (protein catalysts) provide an alternate reaction pathway with a lower activation energy. However, they do not change the free energy of the reactant or product. A is the activation energy of the uncatalyzed reaction. C is the activation energy of the catalyzed reverse reaction. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW |
Biochemistry_Lippincott_231 | Biochemistry_Lippinco | For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW Bioenergetics describes the transfer and utilization of energy in biologic systems. It concerns the initial and final energy states of the reaction components, not the reaction mechanism or how much time it takes for the chemical change to occur. Bioenergetics makes use of a few basic ideas from the field of thermodynamics, particularly the concept of free energy. Because changes in free energy provide a measure of the energetic feasibility of a chemical reaction, they allow prediction of whether a reaction or process can take place. In short, bioenergetics predicts if a process is possible, whereas kinetics measures the reaction rate (see p. 54). II. FREE ENERGY | Biochemistry_Lippinco. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW Bioenergetics describes the transfer and utilization of energy in biologic systems. It concerns the initial and final energy states of the reaction components, not the reaction mechanism or how much time it takes for the chemical change to occur. Bioenergetics makes use of a few basic ideas from the field of thermodynamics, particularly the concept of free energy. Because changes in free energy provide a measure of the energetic feasibility of a chemical reaction, they allow prediction of whether a reaction or process can take place. In short, bioenergetics predicts if a process is possible, whereas kinetics measures the reaction rate (see p. 54). II. FREE ENERGY |
Biochemistry_Lippincott_232 | Biochemistry_Lippinco | II. FREE ENERGY The direction and extent to which a chemical reaction proceeds are determined by the degree to which two factors change during the reaction. These are enthalpy (∆H, a measure of the change [∆] in heat content of the reactants and products) and entropy (∆S, a measure of the change in randomness or disorder of the reactants and products), as shown in Figure 6.1. Neither of these thermodynamic quantities by itself is sufficient to determine whether a chemical reaction will proceed spontaneously in the direction it is written. However, when combined mathematically (see Fig. 6.1), enthalpy and entropy can be used to define a third quantity, free energy (G), which predicts the direction in which a reaction will spontaneously proceed. III. FREE ENERGY CHANGE | Biochemistry_Lippinco. II. FREE ENERGY The direction and extent to which a chemical reaction proceeds are determined by the degree to which two factors change during the reaction. These are enthalpy (∆H, a measure of the change [∆] in heat content of the reactants and products) and entropy (∆S, a measure of the change in randomness or disorder of the reactants and products), as shown in Figure 6.1. Neither of these thermodynamic quantities by itself is sufficient to determine whether a chemical reaction will proceed spontaneously in the direction it is written. However, when combined mathematically (see Fig. 6.1), enthalpy and entropy can be used to define a third quantity, free energy (G), which predicts the direction in which a reaction will spontaneously proceed. III. FREE ENERGY CHANGE |
Biochemistry_Lippincott_233 | Biochemistry_Lippinco | III. FREE ENERGY CHANGE The change in free energy is represented in two ways, ∆G and ∆G0. The first, ∆G (without the superscript “0”), represents the change in free energy and, thus, the direction of a reaction at any specified concentration of products and reactants. ∆G, then, is a variable. This contrasts with the standard free energy change, ∆G0 (with the superscript “0”), which is the energy change when reactants and products are at a concentration of 1 mol/l. [Note: The concentration of protons (H+) is assumed to be 10−7 mol/l (that is, pH = 7). This may be shown by a prime sign (ʹ ), for example, ∆G0ʹ.] Although ∆G0, a constant, represents energy changes at these nonphysiologic concentrations of reactants and products, it is nonetheless useful in comparing the energy changes of different reactions. Furthermore, ∆G0 can readily be determined from measurement of the equilibrium constant (see p. 71). [Note: This section outlines the uses of ∆G, and ∆G0 is described in D. below.] | Biochemistry_Lippinco. III. FREE ENERGY CHANGE The change in free energy is represented in two ways, ∆G and ∆G0. The first, ∆G (without the superscript “0”), represents the change in free energy and, thus, the direction of a reaction at any specified concentration of products and reactants. ∆G, then, is a variable. This contrasts with the standard free energy change, ∆G0 (with the superscript “0”), which is the energy change when reactants and products are at a concentration of 1 mol/l. [Note: The concentration of protons (H+) is assumed to be 10−7 mol/l (that is, pH = 7). This may be shown by a prime sign (ʹ ), for example, ∆G0ʹ.] Although ∆G0, a constant, represents energy changes at these nonphysiologic concentrations of reactants and products, it is nonetheless useful in comparing the energy changes of different reactions. Furthermore, ∆G0 can readily be determined from measurement of the equilibrium constant (see p. 71). [Note: This section outlines the uses of ∆G, and ∆G0 is described in D. below.] |
Biochemistry_Lippincott_234 | Biochemistry_Lippinco | A. ∆G and reaction direction The sign of ∆G can be used to predict the direction of a reaction at constant temperature and pressure. Consider the reaction: 1. Negative ∆G: If ∆G is negative, then there is a net loss of energy, and the reaction goes spontaneously as written (that is, A is converted into B) as shown in Figure 6.2A. The reaction is said to be exergonic. 2. Positive ∆G: If ∆G is positive, then there is a net gain of energy, and the reaction does not go spontaneously from B to A (Fig. 6.2B). Energy must be added to the system to make the reaction go from B to A. The reaction is said to be endergonic. 3. Zero ∆G: If ∆G = 0, then the reaction is in equilibrium. [Note: When a reaction is proceeding spontaneously (that is, ∆G is negative), the reaction continues until ∆G reaches zero and equilibrium is established.] B. ∆G of the forward and back reactions | Biochemistry_Lippinco. A. ∆G and reaction direction The sign of ∆G can be used to predict the direction of a reaction at constant temperature and pressure. Consider the reaction: 1. Negative ∆G: If ∆G is negative, then there is a net loss of energy, and the reaction goes spontaneously as written (that is, A is converted into B) as shown in Figure 6.2A. The reaction is said to be exergonic. 2. Positive ∆G: If ∆G is positive, then there is a net gain of energy, and the reaction does not go spontaneously from B to A (Fig. 6.2B). Energy must be added to the system to make the reaction go from B to A. The reaction is said to be endergonic. 3. Zero ∆G: If ∆G = 0, then the reaction is in equilibrium. [Note: When a reaction is proceeding spontaneously (that is, ∆G is negative), the reaction continues until ∆G reaches zero and equilibrium is established.] B. ∆G of the forward and back reactions |
Biochemistry_Lippincott_235 | Biochemistry_Lippinco | B. ∆G of the forward and back reactions The free energy of the forward reaction (A → B) is equal in magnitude but opposite in sign to that of the back reaction (B → A). For example, if ∆G of the forward reaction is −5 kcal/mol, then that of the back reaction is +5 kcal/mol. [Note: ∆G can also be expressed in kilojoules per mole or kJ/mol (1 kcal = 4.2 kJ).] C. ∆G and reactant and product concentrations The ∆G of the reaction A → B depends on the concentration of the reactant and product. At constant temperature and pressure, the following relationship can be derived: where ∆G0 is the standard free energy change (see D. below) R is the gas constant (1.987 cal/mol K) T is the absolute temperature (K) [A] and [B] are the actual concentrations of the reactant and product ln represents the natural logarithm. | Biochemistry_Lippinco. B. ∆G of the forward and back reactions The free energy of the forward reaction (A → B) is equal in magnitude but opposite in sign to that of the back reaction (B → A). For example, if ∆G of the forward reaction is −5 kcal/mol, then that of the back reaction is +5 kcal/mol. [Note: ∆G can also be expressed in kilojoules per mole or kJ/mol (1 kcal = 4.2 kJ).] C. ∆G and reactant and product concentrations The ∆G of the reaction A → B depends on the concentration of the reactant and product. At constant temperature and pressure, the following relationship can be derived: where ∆G0 is the standard free energy change (see D. below) R is the gas constant (1.987 cal/mol K) T is the absolute temperature (K) [A] and [B] are the actual concentrations of the reactant and product ln represents the natural logarithm. |
Biochemistry_Lippincott_236 | Biochemistry_Lippinco | A reaction with a positive ∆G0 can proceed in the forward direction if the ratio of products to reactants ([B]/[A]) is sufficiently small (that is, the ratio of reactants to products is large) to make ∆G negative. For example, consider the reaction: D. Standard free energy change The standard free energy change, ∆G0, is so called because it is equal to the free energy change, ∆G, under standard conditions (that is, when reactants and products are at 1 mol/l concentrations; Fig. 6.3B). Under these conditions, the natural logarithm of the ratio of products to reactants is zero (ln1 = 0), and, therefore, the equation shown at the bottom of the previous page becomes: 1. | Biochemistry_Lippinco. A reaction with a positive ∆G0 can proceed in the forward direction if the ratio of products to reactants ([B]/[A]) is sufficiently small (that is, the ratio of reactants to products is large) to make ∆G negative. For example, consider the reaction: D. Standard free energy change The standard free energy change, ∆G0, is so called because it is equal to the free energy change, ∆G, under standard conditions (that is, when reactants and products are at 1 mol/l concentrations; Fig. 6.3B). Under these conditions, the natural logarithm of the ratio of products to reactants is zero (ln1 = 0), and, therefore, the equation shown at the bottom of the previous page becomes: 1. |
Biochemistry_Lippincott_237 | Biochemistry_Lippinco | ∆G0 and reaction direction: Under standard conditions, ∆G0 can be used to predict the direction a reaction proceeds because, under these conditions, ∆G0 is equal to ∆G. However, ∆G0 cannot predict the direction of a reaction under physiologic conditions because it is composed solely of constants (R, T, and Keq [see 2. below]) and is not, therefore, altered by changes in product or substrate concentrations. 2. | Biochemistry_Lippinco. ∆G0 and reaction direction: Under standard conditions, ∆G0 can be used to predict the direction a reaction proceeds because, under these conditions, ∆G0 is equal to ∆G. However, ∆G0 cannot predict the direction of a reaction under physiologic conditions because it is composed solely of constants (R, T, and Keq [see 2. below]) and is not, therefore, altered by changes in product or substrate concentrations. 2. |
Biochemistry_Lippincott_238 | Biochemistry_Lippinco | 2. Relationship between ∆G0 and Keq: In a reaction A ⇄ B, a point of equilibrium is reached at which no further net chemical change takes place (that is, when A is being converted to B as fast as B is being converted to A). In this state, the ratio of [B] to [A] is constant, regardless of the actual concentrations of the two compounds: where Keq is the equilibrium constant, and [A]eq and [B]eq are the concentrations of A and B at equilibrium. If the reaction A ⇄ B is allowed to go to equilibrium at constant temperature and pressure, then, at equilibrium, the overall ∆G is zero (Fig. 6.3C). Therefore, where the actual concentrations of A and B are equal to the equilibrium concentrations of reactant and product ([A]eq and [B]eq), and their ratio is equal to the Keq. Thus, This equation allows some simple predictions: 3. ∆G0s of two consecutive reactions: The ∆G0s are additive in any sequence of consecutive reactions, as are the ∆Gs. For example: 4. | Biochemistry_Lippinco. 2. Relationship between ∆G0 and Keq: In a reaction A ⇄ B, a point of equilibrium is reached at which no further net chemical change takes place (that is, when A is being converted to B as fast as B is being converted to A). In this state, the ratio of [B] to [A] is constant, regardless of the actual concentrations of the two compounds: where Keq is the equilibrium constant, and [A]eq and [B]eq are the concentrations of A and B at equilibrium. If the reaction A ⇄ B is allowed to go to equilibrium at constant temperature and pressure, then, at equilibrium, the overall ∆G is zero (Fig. 6.3C). Therefore, where the actual concentrations of A and B are equal to the equilibrium concentrations of reactant and product ([A]eq and [B]eq), and their ratio is equal to the Keq. Thus, This equation allows some simple predictions: 3. ∆G0s of two consecutive reactions: The ∆G0s are additive in any sequence of consecutive reactions, as are the ∆Gs. For example: 4. |
Biochemistry_Lippincott_239 | Biochemistry_Lippinco | This equation allows some simple predictions: 3. ∆G0s of two consecutive reactions: The ∆G0s are additive in any sequence of consecutive reactions, as are the ∆Gs. For example: 4. ∆Gs of a pathway: The additive property of ∆G is very important in biochemical pathways through which substrates (reactants) must pass in a particular direction (for example, A → B → C → D → …). As long as the sum of the ∆Gs of the individual reactions is negative, the pathway can proceed as written, even if some of the individual reactions of the pathway have a positive ∆G. However, the actual rates of the reactions depend on the lowering of activation energies (Ea) by the enzymes that catalyze the reactions (see p. 55). IV. ATP: AN ENERGY CARRIER | Biochemistry_Lippinco. This equation allows some simple predictions: 3. ∆G0s of two consecutive reactions: The ∆G0s are additive in any sequence of consecutive reactions, as are the ∆Gs. For example: 4. ∆Gs of a pathway: The additive property of ∆G is very important in biochemical pathways through which substrates (reactants) must pass in a particular direction (for example, A → B → C → D → …). As long as the sum of the ∆Gs of the individual reactions is negative, the pathway can proceed as written, even if some of the individual reactions of the pathway have a positive ∆G. However, the actual rates of the reactions depend on the lowering of activation energies (Ea) by the enzymes that catalyze the reactions (see p. 55). IV. ATP: AN ENERGY CARRIER |
Biochemistry_Lippincott_240 | Biochemistry_Lippinco | IV. ATP: AN ENERGY CARRIER Reactions or processes that have a large positive ∆G, such as moving ions against a concentration gradient across a cell membrane, are made possible by coupling the endergonic movement of ions with a second, spontaneous process with a large negative ∆G such as the exergonic hydrolysis of ATP (see p. 87). [Note: In the absence of enzymes, ATP is a stable molecule because its hydrolysis has a high Ea.] Figure 6.4 shows a mechanical model of energy coupling. The simplest example of energy coupling in biologic reactions occurs when the energy-requiring and the energy-yielding reactions share a common intermediate. A. Common intermediates Two chemical reactions have a common intermediate when they occur sequentially in that the product of the first reaction is a substrate for the second. For example, given the reactions | Biochemistry_Lippinco. IV. ATP: AN ENERGY CARRIER Reactions or processes that have a large positive ∆G, such as moving ions against a concentration gradient across a cell membrane, are made possible by coupling the endergonic movement of ions with a second, spontaneous process with a large negative ∆G such as the exergonic hydrolysis of ATP (see p. 87). [Note: In the absence of enzymes, ATP is a stable molecule because its hydrolysis has a high Ea.] Figure 6.4 shows a mechanical model of energy coupling. The simplest example of energy coupling in biologic reactions occurs when the energy-requiring and the energy-yielding reactions share a common intermediate. A. Common intermediates Two chemical reactions have a common intermediate when they occur sequentially in that the product of the first reaction is a substrate for the second. For example, given the reactions |
Biochemistry_Lippincott_241 | Biochemistry_Lippinco | Two chemical reactions have a common intermediate when they occur sequentially in that the product of the first reaction is a substrate for the second. For example, given the reactions D is the common intermediate and can serve as a carrier of chemical energy between the two reactions. [Note: The intermediate may be linked to an enzyme.] Many coupled reactions use ATP to generate a common intermediate. These reactions may involve the transfer of a phosphate group from ATP to another molecule. Other reactions involve the transfer of phosphate from an energy-rich intermediate to adenosine diphosphate (ADP), forming ATP. B. Energy carried by ATP | Biochemistry_Lippinco. Two chemical reactions have a common intermediate when they occur sequentially in that the product of the first reaction is a substrate for the second. For example, given the reactions D is the common intermediate and can serve as a carrier of chemical energy between the two reactions. [Note: The intermediate may be linked to an enzyme.] Many coupled reactions use ATP to generate a common intermediate. These reactions may involve the transfer of a phosphate group from ATP to another molecule. Other reactions involve the transfer of phosphate from an energy-rich intermediate to adenosine diphosphate (ADP), forming ATP. B. Energy carried by ATP |
Biochemistry_Lippincott_242 | Biochemistry_Lippinco | B. Energy carried by ATP ATP consists of a molecule of adenosine (adenine + ribose) to which three phosphate groups are attached (Fig. 6.5). Removal of one phosphate produces ADP, and removal of two phosphates produces adenosine monophosphate (AMP). For ATP, the ∆G0 of hydrolysis is approximately – 7.3 kcal/mol for each of the two terminal phosphate groups. Because of this large negative ∆G0 of hydrolysis, ATP is called a high-energy phosphate compound. [Note: Adenine nucleotides are interconverted (2 ADP ⇄ ATP + AMP) by adenylate kinase.] V. ELECTRON TRANSPORT CHAIN | Biochemistry_Lippinco. B. Energy carried by ATP ATP consists of a molecule of adenosine (adenine + ribose) to which three phosphate groups are attached (Fig. 6.5). Removal of one phosphate produces ADP, and removal of two phosphates produces adenosine monophosphate (AMP). For ATP, the ∆G0 of hydrolysis is approximately – 7.3 kcal/mol for each of the two terminal phosphate groups. Because of this large negative ∆G0 of hydrolysis, ATP is called a high-energy phosphate compound. [Note: Adenine nucleotides are interconverted (2 ADP ⇄ ATP + AMP) by adenylate kinase.] V. ELECTRON TRANSPORT CHAIN |
Biochemistry_Lippincott_243 | Biochemistry_Lippinco | Energy-rich molecules, such as glucose, are metabolized by a series of oxidation reactions ultimately yielding carbon dioxide and water (H2O), as shown in Figure 6.6. The metabolic intermediates of these reactions donate electrons to specific coenzymes, nicotinamide adenine dinucleotide (NAD+) and flavin adenine dinucleotide (FAD), to form the energy-rich reduced forms, NADH and FADH2. These reduced coenzymes can, in turn, each donate a pair of electrons to a specialized set of electron carriers, collectively called the electron transport chain (ETC), described in this section. As electrons are passed down the ETC, they lose much of their free energy. This energy is used to move H+ across the inner mitochondrial membrane, creating a H+ gradient that drives the production of ATP from ADP and inorganic phosphate (Pi), described on p. 77. The coupling of electron transport with ATP synthesis is called oxidative phosphorylation, sometimes denoted as OXPHOS. It proceeds continuously in all | Biochemistry_Lippinco. Energy-rich molecules, such as glucose, are metabolized by a series of oxidation reactions ultimately yielding carbon dioxide and water (H2O), as shown in Figure 6.6. The metabolic intermediates of these reactions donate electrons to specific coenzymes, nicotinamide adenine dinucleotide (NAD+) and flavin adenine dinucleotide (FAD), to form the energy-rich reduced forms, NADH and FADH2. These reduced coenzymes can, in turn, each donate a pair of electrons to a specialized set of electron carriers, collectively called the electron transport chain (ETC), described in this section. As electrons are passed down the ETC, they lose much of their free energy. This energy is used to move H+ across the inner mitochondrial membrane, creating a H+ gradient that drives the production of ATP from ADP and inorganic phosphate (Pi), described on p. 77. The coupling of electron transport with ATP synthesis is called oxidative phosphorylation, sometimes denoted as OXPHOS. It proceeds continuously in all |
Biochemistry_Lippincott_244 | Biochemistry_Lippinco | inorganic phosphate (Pi), described on p. 77. The coupling of electron transport with ATP synthesis is called oxidative phosphorylation, sometimes denoted as OXPHOS. It proceeds continuously in all tissues that contain mitochondria. [Note: The free energy not trapped as ATP is used to drive ancillary reactions such as transport of calcium ions into mitochondria and to generate heat.] | Biochemistry_Lippinco. inorganic phosphate (Pi), described on p. 77. The coupling of electron transport with ATP synthesis is called oxidative phosphorylation, sometimes denoted as OXPHOS. It proceeds continuously in all tissues that contain mitochondria. [Note: The free energy not trapped as ATP is used to drive ancillary reactions such as transport of calcium ions into mitochondria and to generate heat.] |
Biochemistry_Lippincott_245 | Biochemistry_Lippinco | A. Mitochondrial electron transport chain The ETC (except for cytochrome c, see p. 75) is located in the inner mitochondrial membrane and is the final common pathway by which electrons derived from different fuels of the body flow to oxygen (O2), reducing it to H2O (see Fig. 6.6). | Biochemistry_Lippinco. A. Mitochondrial electron transport chain The ETC (except for cytochrome c, see p. 75) is located in the inner mitochondrial membrane and is the final common pathway by which electrons derived from different fuels of the body flow to oxygen (O2), reducing it to H2O (see Fig. 6.6). |
Biochemistry_Lippincott_246 | Biochemistry_Lippinco | 1. Mitochondrial membranes: The mitochondrion contains an outer and an inner membrane separated by the intermembrane space. Although the outer membrane contains special channels (formed by the protein porin), making it freely permeable to most ions and small molecules, the inner membrane is a specialized structure that is impermeable to most small ions, including H+, and small molecules such as ATP, ADP, pyruvate, and other metabolites important to mitochondrial function (Fig. 6.7). Specialized carriers or transport systems are required to move ions or molecules across this membrane. The inner mitochondrial membrane is unusually rich in proteins, over half of which are directly involved in oxidative phosphorylation. It also contains convolutions, called cristae, which greatly increase its surface area. | Biochemistry_Lippinco. 1. Mitochondrial membranes: The mitochondrion contains an outer and an inner membrane separated by the intermembrane space. Although the outer membrane contains special channels (formed by the protein porin), making it freely permeable to most ions and small molecules, the inner membrane is a specialized structure that is impermeable to most small ions, including H+, and small molecules such as ATP, ADP, pyruvate, and other metabolites important to mitochondrial function (Fig. 6.7). Specialized carriers or transport systems are required to move ions or molecules across this membrane. The inner mitochondrial membrane is unusually rich in proteins, over half of which are directly involved in oxidative phosphorylation. It also contains convolutions, called cristae, which greatly increase its surface area. |
Biochemistry_Lippincott_247 | Biochemistry_Lippinco | 2. Mitochondrial matrix: The gel-like solution of the matrix (interior) of mitochondria is also rich in proteins. These include the enzymes responsible for the oxidation of pyruvate, amino acids, and fatty acids (by β-oxidation) as well as those of the tricarboxylic acid (TCA) cycle. The synthesis of glucose, urea, and heme occurs partially in the matrix of mitochondria. In addition, the matrix contains NAD+ and FAD (the oxidized forms of the two coenzymes that are required as electron acceptors), and ADP and Pi, which are used to produce ATP. [Note: The matrix also contains mitochondrial deoxyribonucleic acid (mtDNA), ribonucleic acid (mtRNA), and ribosomes.] B. Organization | Biochemistry_Lippinco. 2. Mitochondrial matrix: The gel-like solution of the matrix (interior) of mitochondria is also rich in proteins. These include the enzymes responsible for the oxidation of pyruvate, amino acids, and fatty acids (by β-oxidation) as well as those of the tricarboxylic acid (TCA) cycle. The synthesis of glucose, urea, and heme occurs partially in the matrix of mitochondria. In addition, the matrix contains NAD+ and FAD (the oxidized forms of the two coenzymes that are required as electron acceptors), and ADP and Pi, which are used to produce ATP. [Note: The matrix also contains mitochondrial deoxyribonucleic acid (mtDNA), ribonucleic acid (mtRNA), and ribosomes.] B. Organization |
Biochemistry_Lippincott_248 | Biochemistry_Lippinco | B. Organization The inner mitochondrial membrane contains four separate protein complexes, called Complexes I, II, III, and IV that each contain part of the ETC (Fig. 6.8). These complexes accept or donate electrons to the relatively mobile electron carrier coenzyme Q (CoQ) and cytochrome c. Each carrier in the ETC can receive electrons from an electron donor and can subsequently donate electrons to the next acceptor in the chain. The electrons ultimately combine with O2 and H+ to form H2O. This requirement for O2 makes the electron transport process the respiratory chain, which accounts for the greatest portion of the body’s use of O2. C. Reactions | Biochemistry_Lippinco. B. Organization The inner mitochondrial membrane contains four separate protein complexes, called Complexes I, II, III, and IV that each contain part of the ETC (Fig. 6.8). These complexes accept or donate electrons to the relatively mobile electron carrier coenzyme Q (CoQ) and cytochrome c. Each carrier in the ETC can receive electrons from an electron donor and can subsequently donate electrons to the next acceptor in the chain. The electrons ultimately combine with O2 and H+ to form H2O. This requirement for O2 makes the electron transport process the respiratory chain, which accounts for the greatest portion of the body’s use of O2. C. Reactions |
Biochemistry_Lippincott_249 | Biochemistry_Lippinco | C. Reactions With the exception of CoQ, which is a lipid-soluble quinone, all members of the ETC are proteins. These may function as enzymes as is the case with the flavin-containing dehydrogenases, may contain iron as part of an iron-sulfur (Fe-S) center, may contain iron as part of the porphyrin prosthetic group of heme as in the cytochromes, or may contain copper (Cu) as does the cytochrome a + a3 complex. 1. NADH formation: NAD+ is reduced to NADH by dehydrogenases that remove two hydrogen atoms from their substrate. [Note: For examples of these reactions, see the discussion of the dehydrogenases of the TCA cycle, p. 112.] Both electrons but only one H+ (that is, a hydride ion [:H−]) are transferred to the NAD+, forming NADH plus a free H+. 2. | Biochemistry_Lippinco. C. Reactions With the exception of CoQ, which is a lipid-soluble quinone, all members of the ETC are proteins. These may function as enzymes as is the case with the flavin-containing dehydrogenases, may contain iron as part of an iron-sulfur (Fe-S) center, may contain iron as part of the porphyrin prosthetic group of heme as in the cytochromes, or may contain copper (Cu) as does the cytochrome a + a3 complex. 1. NADH formation: NAD+ is reduced to NADH by dehydrogenases that remove two hydrogen atoms from their substrate. [Note: For examples of these reactions, see the discussion of the dehydrogenases of the TCA cycle, p. 112.] Both electrons but only one H+ (that is, a hydride ion [:H−]) are transferred to the NAD+, forming NADH plus a free H+. 2. |
Biochemistry_Lippincott_250 | Biochemistry_Lippinco | 2. NADH dehydrogenase: The free H+ plus the hydride ion carried by NADH are transferred to NADH dehydrogenase, a protein complex (Complex I) embedded in the inner mitochondrial membrane. Complex I has a tightly bound molecule of flavin mononucleotide (FMN), a coenzyme structurally related to FAD (see Fig. 28.15, p. 384) that accepts the two hydrogen atoms (2 electrons + 2 H+), becoming FMNH2. NADH dehydrogenase also contains peptide subunits with Fe-S centers (Fig. 6.9). At Complex I, electrons move from NADH to FMN to the iron of the Fe-S centers and then to CoQ. As electrons flow, they lose energy. This energy is used to pump four H+ across the inner mitochondrial membrane, from the matrix to the intermembrane space. 3. Succinate dehydrogenase: At Complex II, electrons from the succinate dehydrogenase–catalyzed oxidation of succinate to fumarate move from the coenzyme, FADH2, to an Fe-S protein, and then to CoQ. [Note: | Biochemistry_Lippinco. 2. NADH dehydrogenase: The free H+ plus the hydride ion carried by NADH are transferred to NADH dehydrogenase, a protein complex (Complex I) embedded in the inner mitochondrial membrane. Complex I has a tightly bound molecule of flavin mononucleotide (FMN), a coenzyme structurally related to FAD (see Fig. 28.15, p. 384) that accepts the two hydrogen atoms (2 electrons + 2 H+), becoming FMNH2. NADH dehydrogenase also contains peptide subunits with Fe-S centers (Fig. 6.9). At Complex I, electrons move from NADH to FMN to the iron of the Fe-S centers and then to CoQ. As electrons flow, they lose energy. This energy is used to pump four H+ across the inner mitochondrial membrane, from the matrix to the intermembrane space. 3. Succinate dehydrogenase: At Complex II, electrons from the succinate dehydrogenase–catalyzed oxidation of succinate to fumarate move from the coenzyme, FADH2, to an Fe-S protein, and then to CoQ. [Note: |
Biochemistry_Lippincott_251 | Biochemistry_Lippinco | Because no energy is lost in this process, no H+ are pumped at Complex II.] 4. Coenzyme Q: CoQ is a quinone derivative with a long, hydrophobic isoprenoid tail. It is made from an intermediate of cholesterol synthesis (see p. 221). [Note: It is also called ubiquinone because it is ubiquitous in biologic systems.] CoQ is a mobile electron carrier and can accept electrons from NADH dehydrogenase (Complex I), from succinate dehydrogenase (Complex II) and from other mitochondrial dehydrogenases, such as glycerol 3-phosphate dehydrogenase (see p. 80) and acyl CoA dehydrogenases (see p. 192). CoQ transfers electrons to Complex III (cytochrome bc1). Thus, a function of CoQ is to link the flavoprotein dehydrogenases to the cytochromes. 5. | Biochemistry_Lippinco. Because no energy is lost in this process, no H+ are pumped at Complex II.] 4. Coenzyme Q: CoQ is a quinone derivative with a long, hydrophobic isoprenoid tail. It is made from an intermediate of cholesterol synthesis (see p. 221). [Note: It is also called ubiquinone because it is ubiquitous in biologic systems.] CoQ is a mobile electron carrier and can accept electrons from NADH dehydrogenase (Complex I), from succinate dehydrogenase (Complex II) and from other mitochondrial dehydrogenases, such as glycerol 3-phosphate dehydrogenase (see p. 80) and acyl CoA dehydrogenases (see p. 192). CoQ transfers electrons to Complex III (cytochrome bc1). Thus, a function of CoQ is to link the flavoprotein dehydrogenases to the cytochromes. 5. |
Biochemistry_Lippincott_252 | Biochemistry_Lippinco | 80) and acyl CoA dehydrogenases (see p. 192). CoQ transfers electrons to Complex III (cytochrome bc1). Thus, a function of CoQ is to link the flavoprotein dehydrogenases to the cytochromes. 5. Cytochromes: The remaining members of the ETC are cytochrome proteins. Each contains a heme group (a porphyrin ring plus iron). Unlike the heme groups of hemoglobin, the cytochrome iron is reversibly converted from its ferric (Fe3+) to its ferrous (Fe2+) form as a normal part of its function as an acceptor and donor of electrons. Electrons are passed along the chain from cytochrome bc1 (Complex III), to cytochrome c, and then to cytochromes a + a3 ([Complex IV] see Fig. 6.8). As electrons flow, four H+ are pumped across the inner mitochondrial membrane at Complex III and two at Complex IV. [Note: Cytochrome c is located in the intermembrane space, loosely associated with the outer face of the inner membrane. As seen with CoQ, cytochrome c is a mobile electron carrier.] 6. | Biochemistry_Lippinco. 80) and acyl CoA dehydrogenases (see p. 192). CoQ transfers electrons to Complex III (cytochrome bc1). Thus, a function of CoQ is to link the flavoprotein dehydrogenases to the cytochromes. 5. Cytochromes: The remaining members of the ETC are cytochrome proteins. Each contains a heme group (a porphyrin ring plus iron). Unlike the heme groups of hemoglobin, the cytochrome iron is reversibly converted from its ferric (Fe3+) to its ferrous (Fe2+) form as a normal part of its function as an acceptor and donor of electrons. Electrons are passed along the chain from cytochrome bc1 (Complex III), to cytochrome c, and then to cytochromes a + a3 ([Complex IV] see Fig. 6.8). As electrons flow, four H+ are pumped across the inner mitochondrial membrane at Complex III and two at Complex IV. [Note: Cytochrome c is located in the intermembrane space, loosely associated with the outer face of the inner membrane. As seen with CoQ, cytochrome c is a mobile electron carrier.] 6. |
Biochemistry_Lippincott_253 | Biochemistry_Lippinco | Cytochrome a + a3: Because this cytochrome complex (Complex IV) is the only electron carrier in which the heme iron has an available coordination site that can react directly with O2, it also is called cytochrome c oxidase. At Complex IV, the transported electrons, O2, and free H+ are brought together, and O2 is reduced to H2O (see Fig. 6.8). [Note: Four electrons are required to reduce one molecule of O2 to two molecules of H2O.] Cytochrome c oxidase contains Cu atoms that are required for this complicated reaction to occur. Electrons move from CuA to cytochrome a to cytochrome a3 (in association with CuB) to O2. 7. | Biochemistry_Lippinco. Cytochrome a + a3: Because this cytochrome complex (Complex IV) is the only electron carrier in which the heme iron has an available coordination site that can react directly with O2, it also is called cytochrome c oxidase. At Complex IV, the transported electrons, O2, and free H+ are brought together, and O2 is reduced to H2O (see Fig. 6.8). [Note: Four electrons are required to reduce one molecule of O2 to two molecules of H2O.] Cytochrome c oxidase contains Cu atoms that are required for this complicated reaction to occur. Electrons move from CuA to cytochrome a to cytochrome a3 (in association with CuB) to O2. 7. |
Biochemistry_Lippincott_254 | Biochemistry_Lippinco | 7. Site-specific inhibitors: Inhibitors of specific sites in the ETC have been identified and are illustrated in Figure 6.10. These respiratory inhibitors prevent the passage of electrons by binding to a component of the chain, blocking the oxidation-reduction reaction. Therefore, all electron carriers before the block are fully reduced, whereas those located after the block are oxidized. [Note: Inhibition of the ETC inhibits ATP synthesis because these processes are tightly coupled (see p. 78).] NaN3 = sodium azide. Leakage of electrons from the ETC produces reactive oxygen species (ROS), such as superoxide (O2−·), hydrogen peroxide (H2O2), and hydroxyl radicals (OH·). ROS damage DNA and proteins and cause lipid peroxidation. Enzymes such as superoxide dismutase (SOD), catalase, and glutathione peroxidase are cellular defenses against ROS (see p. 148). D. Free energy release during electron transport | Biochemistry_Lippinco. 7. Site-specific inhibitors: Inhibitors of specific sites in the ETC have been identified and are illustrated in Figure 6.10. These respiratory inhibitors prevent the passage of electrons by binding to a component of the chain, blocking the oxidation-reduction reaction. Therefore, all electron carriers before the block are fully reduced, whereas those located after the block are oxidized. [Note: Inhibition of the ETC inhibits ATP synthesis because these processes are tightly coupled (see p. 78).] NaN3 = sodium azide. Leakage of electrons from the ETC produces reactive oxygen species (ROS), such as superoxide (O2−·), hydrogen peroxide (H2O2), and hydroxyl radicals (OH·). ROS damage DNA and proteins and cause lipid peroxidation. Enzymes such as superoxide dismutase (SOD), catalase, and glutathione peroxidase are cellular defenses against ROS (see p. 148). D. Free energy release during electron transport |
Biochemistry_Lippincott_255 | Biochemistry_Lippinco | D. Free energy release during electron transport The free energy released as electrons are transferred along the ETC from an electron donor (reducing agent or reductant) to an electron acceptor (oxidizing agent or oxidant) is used to pump H+ at Complexes I, III, and IV. [Note: The electrons can be transferred as hydride ions to NAD+; as hydrogen atoms to FMN, CoQ, and FAD; or as electrons to cytochromes.] 1. | Biochemistry_Lippinco. D. Free energy release during electron transport The free energy released as electrons are transferred along the ETC from an electron donor (reducing agent or reductant) to an electron acceptor (oxidizing agent or oxidant) is used to pump H+ at Complexes I, III, and IV. [Note: The electrons can be transferred as hydride ions to NAD+; as hydrogen atoms to FMN, CoQ, and FAD; or as electrons to cytochromes.] 1. |
Biochemistry_Lippincott_256 | Biochemistry_Lippinco | Redox pairs: Oxidation (loss of electrons) of one substance is always accompanied by reduction (gain of electrons) of a second. For example, Figure 6.11 shows the oxidation of NADH to NAD+ by NADH dehydrogenase at Complex I, accompanied by the reduction of FMN, the prosthetic group, to FMNH2. Such redox reactions can be written as the sum of two separate half reactions, one an oxidation and the other a reduction (see Fig. 6.11). NAD+ and NADH form a redox pair, as do FMN and FMNH2. Redox pairs differ in their tendency to lose electrons. This tendency is a characteristic of a particular redox pair and can be quantitatively specified by a constant, E0 (the standard reduction potential), with units in volts. 2. | Biochemistry_Lippinco. Redox pairs: Oxidation (loss of electrons) of one substance is always accompanied by reduction (gain of electrons) of a second. For example, Figure 6.11 shows the oxidation of NADH to NAD+ by NADH dehydrogenase at Complex I, accompanied by the reduction of FMN, the prosthetic group, to FMNH2. Such redox reactions can be written as the sum of two separate half reactions, one an oxidation and the other a reduction (see Fig. 6.11). NAD+ and NADH form a redox pair, as do FMN and FMNH2. Redox pairs differ in their tendency to lose electrons. This tendency is a characteristic of a particular redox pair and can be quantitatively specified by a constant, E0 (the standard reduction potential), with units in volts. 2. |
Biochemistry_Lippincott_257 | Biochemistry_Lippinco | 2. Standard reduction potential: The E0 of various redox pairs can be ordered from the most negative E0 to the most positive. The more negative the E0 of a redox pair, the greater the tendency of the reductant member of that pair to lose electrons. The more positive the E0, the greater the tendency of the oxidant member of that pair to accept electrons. Therefore, electrons flow from the pair with the more negative E0 to that with the more positive E0. The E0 values for some members of the ETC are shown in Figure 6.12. [Note: The components of the chain are arranged in order of increasingly positive E0 values.] 3. Relationship of ∆G0 to ∆E0: The ∆G0 is related directly to the magnitude of the change in E0: where n = number of electrons transferred (1 for a cytochrome, 2 for NADH, FADH2, and CoQ) | Biochemistry_Lippinco. 2. Standard reduction potential: The E0 of various redox pairs can be ordered from the most negative E0 to the most positive. The more negative the E0 of a redox pair, the greater the tendency of the reductant member of that pair to lose electrons. The more positive the E0, the greater the tendency of the oxidant member of that pair to accept electrons. Therefore, electrons flow from the pair with the more negative E0 to that with the more positive E0. The E0 values for some members of the ETC are shown in Figure 6.12. [Note: The components of the chain are arranged in order of increasingly positive E0 values.] 3. Relationship of ∆G0 to ∆E0: The ∆G0 is related directly to the magnitude of the change in E0: where n = number of electrons transferred (1 for a cytochrome, 2 for NADH, FADH2, and CoQ) |
Biochemistry_Lippincott_258 | Biochemistry_Lippinco | F = Faraday constant (23.1 kcal/volt mol) ∆E0 = E0 of the electron-accepting pair minus the E0 of the electron-donating pair ∆G0 = change in the standard free energy 4. ∆G0 of ATP: The ∆G0 for the phosphorylation of ADP to ATP is +7.3 kcal/mol. The transport of a pair of electrons from NADH to O2 through the ETC releases 52.6 kcal. Therefore, more than sufficient energy is available to produce three ATP from three ADP and three Pi (3 × 7.3 = 21.9 kcal/mol), sometimes expressed as a P/O ratio (ATP made per O atom reduced) of 3:1. The remaining calories are used for ancillary reactions or released as heat. [Note: The P:O for FADH2 is 2:1 because Complex I is bypassed.] VI. PHOSPHORYLATION OF ADP TO ATP The transfer of electrons down the ETC is energetically favored because NADH is a strong electron donor and O2 is an avid electron acceptor. However, the flow of electrons does not directly result in ATP synthesis. A. Chemiosmotic hypothesis | Biochemistry_Lippinco. F = Faraday constant (23.1 kcal/volt mol) ∆E0 = E0 of the electron-accepting pair minus the E0 of the electron-donating pair ∆G0 = change in the standard free energy 4. ∆G0 of ATP: The ∆G0 for the phosphorylation of ADP to ATP is +7.3 kcal/mol. The transport of a pair of electrons from NADH to O2 through the ETC releases 52.6 kcal. Therefore, more than sufficient energy is available to produce three ATP from three ADP and three Pi (3 × 7.3 = 21.9 kcal/mol), sometimes expressed as a P/O ratio (ATP made per O atom reduced) of 3:1. The remaining calories are used for ancillary reactions or released as heat. [Note: The P:O for FADH2 is 2:1 because Complex I is bypassed.] VI. PHOSPHORYLATION OF ADP TO ATP The transfer of electrons down the ETC is energetically favored because NADH is a strong electron donor and O2 is an avid electron acceptor. However, the flow of electrons does not directly result in ATP synthesis. A. Chemiosmotic hypothesis |
Biochemistry_Lippincott_259 | Biochemistry_Lippinco | A. Chemiosmotic hypothesis The chemiosmotic hypothesis (also known as the Mitchell hypothesis) explains how the free energy generated by the transport of electrons by the ETC is used to produce ATP from ADP + Pi. 1. Proton pump: Electron transport is coupled to ADP phosphorylation by the pumping of H+ across the inner mitochondrial membrane, from the matrix to the intermembrane space, at Complexes I, III, and IV. For each pair of electrons transferred from NADH to O2, 10 H+ are pumped. This creates an electrical gradient (with more positive charges on the cytosolic side of the membrane than on the matrix side) and a pH (chemical) gradient (the cytosolic side of the membrane is at a lower pH than the matrix side), as shown in Figure 6.13. The energy (proton-motive force) generated by these gradients is sufficient to drive ATP synthesis. Thus, the H+ gradient serves as the common intermediate that couples oxidation to phosphorylation. 2. | Biochemistry_Lippinco. A. Chemiosmotic hypothesis The chemiosmotic hypothesis (also known as the Mitchell hypothesis) explains how the free energy generated by the transport of electrons by the ETC is used to produce ATP from ADP + Pi. 1. Proton pump: Electron transport is coupled to ADP phosphorylation by the pumping of H+ across the inner mitochondrial membrane, from the matrix to the intermembrane space, at Complexes I, III, and IV. For each pair of electrons transferred from NADH to O2, 10 H+ are pumped. This creates an electrical gradient (with more positive charges on the cytosolic side of the membrane than on the matrix side) and a pH (chemical) gradient (the cytosolic side of the membrane is at a lower pH than the matrix side), as shown in Figure 6.13. The energy (proton-motive force) generated by these gradients is sufficient to drive ATP synthesis. Thus, the H+ gradient serves as the common intermediate that couples oxidation to phosphorylation. 2. |
Biochemistry_Lippincott_260 | Biochemistry_Lippinco | ATP synthase: The multisubunit enzyme ATP synthase ([Complex V] Fig. 6.14) synthesizes ATP using the energy of the H+ gradient. It contains a membrane domain (Fo) that spans the inner mitochondrial membrane and an extramembranous domain (F1) that appears as a sphere that protrudes into the mitochondrial matrix (see Fig. 6.13). The chemiosmotic hypothesis proposes that after H+ have been pumped to the cytosolic side of the inner mitochondrial membrane, they reenter the matrix by passing through a H+ channel in the Fo domain, driving rotation of the c ring of Fo and, at the same time, dissipating the pH and electrical gradients. Rotation in Fo causes conformational changes in the three β subunits of F1 that allow them to bind ADP + Pi, phosphorylate ADP to ATP, and release ATP. One complete rotation of the c ring produces three ATP. [Note: ATP synthase is also called F1/Fo-ATPase because the enzyme can also catalyze the hydrolysis of ATP to ADP and Pi.] contains eight subunits. One | Biochemistry_Lippinco. ATP synthase: The multisubunit enzyme ATP synthase ([Complex V] Fig. 6.14) synthesizes ATP using the energy of the H+ gradient. It contains a membrane domain (Fo) that spans the inner mitochondrial membrane and an extramembranous domain (F1) that appears as a sphere that protrudes into the mitochondrial matrix (see Fig. 6.13). The chemiosmotic hypothesis proposes that after H+ have been pumped to the cytosolic side of the inner mitochondrial membrane, they reenter the matrix by passing through a H+ channel in the Fo domain, driving rotation of the c ring of Fo and, at the same time, dissipating the pH and electrical gradients. Rotation in Fo causes conformational changes in the three β subunits of F1 that allow them to bind ADP + Pi, phosphorylate ADP to ATP, and release ATP. One complete rotation of the c ring produces three ATP. [Note: ATP synthase is also called F1/Fo-ATPase because the enzyme can also catalyze the hydrolysis of ATP to ADP and Pi.] contains eight subunits. One |
Biochemistry_Lippincott_261 | Biochemistry_Lippinco | rotation of the c ring produces three ATP. [Note: ATP synthase is also called F1/Fo-ATPase because the enzyme can also catalyze the hydrolysis of ATP to ADP and Pi.] contains eight subunits. One complete turn of the ring is driven by eight H+ (protons) moving through the Fo domain. The resulting conformational changes in the three β subunits of the F1 domain allow phosphorylation of three adenosine diphosphates (ADP) to three ATP.] Pi = inorganic phosphate. | Biochemistry_Lippinco. rotation of the c ring produces three ATP. [Note: ATP synthase is also called F1/Fo-ATPase because the enzyme can also catalyze the hydrolysis of ATP to ADP and Pi.] contains eight subunits. One complete turn of the ring is driven by eight H+ (protons) moving through the Fo domain. The resulting conformational changes in the three β subunits of the F1 domain allow phosphorylation of three adenosine diphosphates (ADP) to three ATP.] Pi = inorganic phosphate. |
Biochemistry_Lippincott_262 | Biochemistry_Lippinco | a. Coupling in oxidative phosphorylation: In normal mitochondria, ATP synthesis is coupled to electron transport through the H+ gradient. Increasing (or decreasing) one process has the same effect on the other. For example, hydrolysis of ATP to ADP and Pi in energy-requiring reactions increases the availability of substrates for ATP synthase and, thus, increases H+ flow through the enzyme. Electron transport and H+ pumping by the ETC increase to maintain the H+ gradient and allow ATP synthesis. | Biochemistry_Lippinco. a. Coupling in oxidative phosphorylation: In normal mitochondria, ATP synthesis is coupled to electron transport through the H+ gradient. Increasing (or decreasing) one process has the same effect on the other. For example, hydrolysis of ATP to ADP and Pi in energy-requiring reactions increases the availability of substrates for ATP synthase and, thus, increases H+ flow through the enzyme. Electron transport and H+ pumping by the ETC increase to maintain the H+ gradient and allow ATP synthesis. |
Biochemistry_Lippincott_263 | Biochemistry_Lippinco | b. Oligomycin: This drug binds to the Fo (hence the letter “o”) domain of ATP synthase, closing the H+ channel and preventing reentry of H+ into the matrix, thereby inhibiting phosphorylation of ADP to ATP. Because the pH and electrical gradients cannot be dissipated in the presence of this phosphorylation inhibitor, electron transport stops because of the difficulty of pumping any more H+ against the steep gradient. This dependency of cellular respiration on the ability to phosphorylate ADP to ATP is known as respiratory control and is the consequence of the tight coupling of these processes. c. | Biochemistry_Lippinco. b. Oligomycin: This drug binds to the Fo (hence the letter “o”) domain of ATP synthase, closing the H+ channel and preventing reentry of H+ into the matrix, thereby inhibiting phosphorylation of ADP to ATP. Because the pH and electrical gradients cannot be dissipated in the presence of this phosphorylation inhibitor, electron transport stops because of the difficulty of pumping any more H+ against the steep gradient. This dependency of cellular respiration on the ability to phosphorylate ADP to ATP is known as respiratory control and is the consequence of the tight coupling of these processes. c. |
Biochemistry_Lippincott_264 | Biochemistry_Lippinco | c. Uncoupling proteins: Uncoupling proteins (UCP) occur in the inner mitochondrial membrane of mammals, including humans. These proteins form channels that allow H+ to reenter the mitochondrial matrix without energy being captured as ATP (Fig. 6.15). The energy is released as heat, and the process is called nonshivering thermogenesis. UCP1, also called thermogenin, is responsible for heat production in the mitochondria-rich brown adipocytes of mammals. [Note: Cold causes catecholamine-dependent activation of UCP1 expression.] In brown fat, unlike the more abundant white fat, ~90% of its respiratory energy is used for thermogenesis in infants in response to cold. Thus, brown fat is involved in energy expenditure, whereas white fat is involved in energy storage. [Note: Brown fat depots have recently been shown to be present in adults.] d. | Biochemistry_Lippinco. c. Uncoupling proteins: Uncoupling proteins (UCP) occur in the inner mitochondrial membrane of mammals, including humans. These proteins form channels that allow H+ to reenter the mitochondrial matrix without energy being captured as ATP (Fig. 6.15). The energy is released as heat, and the process is called nonshivering thermogenesis. UCP1, also called thermogenin, is responsible for heat production in the mitochondria-rich brown adipocytes of mammals. [Note: Cold causes catecholamine-dependent activation of UCP1 expression.] In brown fat, unlike the more abundant white fat, ~90% of its respiratory energy is used for thermogenesis in infants in response to cold. Thus, brown fat is involved in energy expenditure, whereas white fat is involved in energy storage. [Note: Brown fat depots have recently been shown to be present in adults.] d. |
Biochemistry_Lippincott_265 | Biochemistry_Lippinco | Synthetic uncouplers: Electron transport and phosphorylation of ADP can also be uncoupled by compounds that shuttle H+ across the inner mitochondrial membrane, dissipating the gradient. The classic example is 2,4-dinitrophenol, a lipophilic H+ carrier (ionophore) that readily diffuses through the mitochondrial membrane (Fig. 6.16). This uncoupler causes electron transport to proceed at a rapid rate without establishing a H+ gradient, much as do the UCP. Again, energy is released as heat rather than being used to synthesize ATP. [Note: In high doses, aspirin and other salicylates uncouple oxidative phosphorylation. This explains the fever that accompanies toxic overdoses of these drugs.] B. Membrane transport systems The inner mitochondrial membrane is impermeable to most charged or hydrophilic substances. However, it contains numerous transport proteins that permit passage of certain molecules from the cytosol to the mitochondrial matrix. 1. | Biochemistry_Lippinco. Synthetic uncouplers: Electron transport and phosphorylation of ADP can also be uncoupled by compounds that shuttle H+ across the inner mitochondrial membrane, dissipating the gradient. The classic example is 2,4-dinitrophenol, a lipophilic H+ carrier (ionophore) that readily diffuses through the mitochondrial membrane (Fig. 6.16). This uncoupler causes electron transport to proceed at a rapid rate without establishing a H+ gradient, much as do the UCP. Again, energy is released as heat rather than being used to synthesize ATP. [Note: In high doses, aspirin and other salicylates uncouple oxidative phosphorylation. This explains the fever that accompanies toxic overdoses of these drugs.] B. Membrane transport systems The inner mitochondrial membrane is impermeable to most charged or hydrophilic substances. However, it contains numerous transport proteins that permit passage of certain molecules from the cytosol to the mitochondrial matrix. 1. |
Biochemistry_Lippincott_266 | Biochemistry_Lippinco | 1. ATP and ADP transport: The inner membrane requires specialized carriers to transport ADP and Pi from the cytosol (where ATP is hydrolyzed to ADP in many energy-requiring reactions) into mitochondria, where ATP can be resynthesized. An adenine nucleotide antiporter imports one ADP from the cytosol into the matrix, while exporting one ATP from the matrix into the cytosol (see Fig. 6.13). A symporter cotransports Pi and H+ from the cytosol into the matrix. 2. | Biochemistry_Lippinco. 1. ATP and ADP transport: The inner membrane requires specialized carriers to transport ADP and Pi from the cytosol (where ATP is hydrolyzed to ADP in many energy-requiring reactions) into mitochondria, where ATP can be resynthesized. An adenine nucleotide antiporter imports one ADP from the cytosol into the matrix, while exporting one ATP from the matrix into the cytosol (see Fig. 6.13). A symporter cotransports Pi and H+ from the cytosol into the matrix. 2. |
Biochemistry_Lippincott_267 | Biochemistry_Lippinco | Reducing equivalent transport: The inner mitochondrial membrane lacks an NADH transporter, and NADH produced in the cytosol (for example, in glycolysis; see p. 101) cannot directly enter the mitochondrial matrix. However, reducing equivalents of NADH are transported from the cytosol into the matrix using substrate shuttles. In the glycerol 3phosphate shuttle (Fig. 6.17A), two electrons are transferred from NADH to dihydroxyacetone phosphate by cytosolic glycerol 3-phosphate dehydrogenase. The glycerol 3-phosphate produced is oxidized by the mitochondrial isozyme as FAD is reduced to FADH2. CoQ of the ETC oxidizes the FADH2. Therefore, the glycerol 3-phosphate shuttle results in the synthesis of two ATP for each cytosolic NADH oxidized. This contrasts with the malate-aspartate shuttle (Fig. 6.17B), which produces NADH (rather than FADH2) in the mitochondrial matrix, thereby yielding three ATP for each cytosolic NADH oxidized by malate dehydrogenase as oxaloacetate is reduced to malate. | Biochemistry_Lippinco. Reducing equivalent transport: The inner mitochondrial membrane lacks an NADH transporter, and NADH produced in the cytosol (for example, in glycolysis; see p. 101) cannot directly enter the mitochondrial matrix. However, reducing equivalents of NADH are transported from the cytosol into the matrix using substrate shuttles. In the glycerol 3phosphate shuttle (Fig. 6.17A), two electrons are transferred from NADH to dihydroxyacetone phosphate by cytosolic glycerol 3-phosphate dehydrogenase. The glycerol 3-phosphate produced is oxidized by the mitochondrial isozyme as FAD is reduced to FADH2. CoQ of the ETC oxidizes the FADH2. Therefore, the glycerol 3-phosphate shuttle results in the synthesis of two ATP for each cytosolic NADH oxidized. This contrasts with the malate-aspartate shuttle (Fig. 6.17B), which produces NADH (rather than FADH2) in the mitochondrial matrix, thereby yielding three ATP for each cytosolic NADH oxidized by malate dehydrogenase as oxaloacetate is reduced to malate. |
Biochemistry_Lippincott_268 | Biochemistry_Lippinco | 6.17B), which produces NADH (rather than FADH2) in the mitochondrial matrix, thereby yielding three ATP for each cytosolic NADH oxidized by malate dehydrogenase as oxaloacetate is reduced to malate. A transport protein moves malate into the mitochondrial matrix. | Biochemistry_Lippinco. 6.17B), which produces NADH (rather than FADH2) in the mitochondrial matrix, thereby yielding three ATP for each cytosolic NADH oxidized by malate dehydrogenase as oxaloacetate is reduced to malate. A transport protein moves malate into the mitochondrial matrix. |
Biochemistry_Lippincott_269 | Biochemistry_Lippinco | inner mitochondrial membrane. A. Glycerol 3-phosphate shuttle. B. Malate aspartate shuttle. DHAP = dihydroxyacetone phosphate; NAD(H) = dinucleotide; CoQ = coenzyme Q. C. Inherited defects in oxidative phosphorylation | Biochemistry_Lippinco. inner mitochondrial membrane. A. Glycerol 3-phosphate shuttle. B. Malate aspartate shuttle. DHAP = dihydroxyacetone phosphate; NAD(H) = dinucleotide; CoQ = coenzyme Q. C. Inherited defects in oxidative phosphorylation |
Biochemistry_Lippincott_270 | Biochemistry_Lippinco | Thirteen of the ~90 polypeptides required for oxidative phosphorylation are encoded by mtDNA and synthesized in mitochondria, whereas the remaining proteins are encoded by nuclear DNA, synthesized in the cytosol, and then transported into mitochondria. Defects in oxidative phosphorylation are more likely a result of alterations in mtDNA, which has a mutation rate about 10 times greater than that of nuclear DNA. Tissues with the greatest ATP requirement (for example, the central nervous system, skeletal and heart muscle, and the liver) are most affected by defects in oxidative phosphorylation. Mutations in mtDNA are responsible for several diseases, including some cases of mitochondrial myopathies, and Leber hereditary optic neuropathy, a disease in which bilateral loss of central vision occurs as a result of neuroretinal degeneration, including damage to the optic nerve. [Note: mtDNA is maternally inherited because mitochondria from the sperm cell do not enter the fertilized egg.] | Biochemistry_Lippinco. Thirteen of the ~90 polypeptides required for oxidative phosphorylation are encoded by mtDNA and synthesized in mitochondria, whereas the remaining proteins are encoded by nuclear DNA, synthesized in the cytosol, and then transported into mitochondria. Defects in oxidative phosphorylation are more likely a result of alterations in mtDNA, which has a mutation rate about 10 times greater than that of nuclear DNA. Tissues with the greatest ATP requirement (for example, the central nervous system, skeletal and heart muscle, and the liver) are most affected by defects in oxidative phosphorylation. Mutations in mtDNA are responsible for several diseases, including some cases of mitochondrial myopathies, and Leber hereditary optic neuropathy, a disease in which bilateral loss of central vision occurs as a result of neuroretinal degeneration, including damage to the optic nerve. [Note: mtDNA is maternally inherited because mitochondria from the sperm cell do not enter the fertilized egg.] |
Biochemistry_Lippincott_271 | Biochemistry_Lippinco | D. Mitochondria and apoptosis The process of apoptosis (programmed cell death) may be initiated through the intrinsic (mitochondrial-mediated) pathway by the formation of pores in the outer mitochondrial membrane. These pores allow cytochrome c to leave the intermembrane space and enter the cytosol. There, cytochrome c, in association with proapoptotic factors, activates a family of proteolytic enzymes (the caspases), causing cleavage of key proteins and resulting in the morphologic and biochemical changes characteristic of apoptosis. VII. CHAPTER SUMMARY | Biochemistry_Lippinco. D. Mitochondria and apoptosis The process of apoptosis (programmed cell death) may be initiated through the intrinsic (mitochondrial-mediated) pathway by the formation of pores in the outer mitochondrial membrane. These pores allow cytochrome c to leave the intermembrane space and enter the cytosol. There, cytochrome c, in association with proapoptotic factors, activates a family of proteolytic enzymes (the caspases), causing cleavage of key proteins and resulting in the morphologic and biochemical changes characteristic of apoptosis. VII. CHAPTER SUMMARY |
Biochemistry_Lippincott_272 | Biochemistry_Lippinco | The change in free energy (∆G) occurring during a reaction predicts the direction in which that reaction will spontaneously proceed. If ∆G is negative (that is, the product has a lower free energy than the substrate), then the reaction is spontaneous as written. If ∆G is positive, then the reaction is not spontaneous. If ∆G = 0, then the reaction is in equilibrium. The ∆G of the forward reaction is equal in magnitude but opposite in sign to that of the back reaction. The ∆G are additive in any sequence of consecutive reactions, as are the standard free energy changes (∆G0). Therefore, reactions or processes that have a large, positive ∆G are made possible by coupling with those that have a large, negative ∆G such as ATP hydrolysis. The reduced coenzymes nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2) each donate a pair of electrons to a specialized set of electron carriers, consisting of flavin mononucleotide (FMN), iron-sulfur centers, coenzyme Q, and | Biochemistry_Lippinco. The change in free energy (∆G) occurring during a reaction predicts the direction in which that reaction will spontaneously proceed. If ∆G is negative (that is, the product has a lower free energy than the substrate), then the reaction is spontaneous as written. If ∆G is positive, then the reaction is not spontaneous. If ∆G = 0, then the reaction is in equilibrium. The ∆G of the forward reaction is equal in magnitude but opposite in sign to that of the back reaction. The ∆G are additive in any sequence of consecutive reactions, as are the standard free energy changes (∆G0). Therefore, reactions or processes that have a large, positive ∆G are made possible by coupling with those that have a large, negative ∆G such as ATP hydrolysis. The reduced coenzymes nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2) each donate a pair of electrons to a specialized set of electron carriers, consisting of flavin mononucleotide (FMN), iron-sulfur centers, coenzyme Q, and |
Biochemistry_Lippincott_273 | Biochemistry_Lippinco | and flavin adenine dinucleotide (FADH2) each donate a pair of electrons to a specialized set of electron carriers, consisting of flavin mononucleotide (FMN), iron-sulfur centers, coenzyme Q, and a series of heme-containing cytochromes, collectively called the electron transport chain. This pathway is present in the inner mitochondrial membrane (impermeable to most substances) and is the final common pathway by which electrons derived from different fuels of the body flow to oxygen (O2), which has a large, positive reduction potential (E0), reducing it to water. The terminal cytochrome, cytochrome c oxidase, is the only cytochrome able to bind O2. Electron transport results in the pumping of protons (H+) across the inner mitochondrial membrane from the matrix to the intermembrane space, 10 H+ per NADH oxidized. This process creates electrical and pH gradients across the inner mitochondrial membrane. After H+ have been transferred to the cytosolic side of the membrane, they reenter the | Biochemistry_Lippinco. and flavin adenine dinucleotide (FADH2) each donate a pair of electrons to a specialized set of electron carriers, consisting of flavin mononucleotide (FMN), iron-sulfur centers, coenzyme Q, and a series of heme-containing cytochromes, collectively called the electron transport chain. This pathway is present in the inner mitochondrial membrane (impermeable to most substances) and is the final common pathway by which electrons derived from different fuels of the body flow to oxygen (O2), which has a large, positive reduction potential (E0), reducing it to water. The terminal cytochrome, cytochrome c oxidase, is the only cytochrome able to bind O2. Electron transport results in the pumping of protons (H+) across the inner mitochondrial membrane from the matrix to the intermembrane space, 10 H+ per NADH oxidized. This process creates electrical and pH gradients across the inner mitochondrial membrane. After H+ have been transferred to the cytosolic side of the membrane, they reenter the |
Biochemistry_Lippincott_274 | Biochemistry_Lippinco | H+ per NADH oxidized. This process creates electrical and pH gradients across the inner mitochondrial membrane. After H+ have been transferred to the cytosolic side of the membrane, they reenter the matrix by passing through the Fo H+ channel in ATP synthase (Complex V), dissipating the pH and electrical gradients and causing conformational changes in the F1 β subunits of the synthase that result in the synthesis of ATP from ADP + inorganic phosphate. Electron transport and phosphorylation are tightly coupled in oxidative phosphorylation ([OXPHOS] Fig. 6.18). Inhibition of one process inhibits the other. These processes can be uncoupled by uncoupling protein-1 of the inner mitochondrial membrane of brown adipocytes and by synthetic compounds such as 2,4-dinitrophenol and aspirin, all of which dissipate the H+ gradient. In uncoupled mitochondria, the energy produced by electron transport is released as heat rather than being used to synthesize ATP. Mutations in mitochondrial DNA, which | Biochemistry_Lippinco. H+ per NADH oxidized. This process creates electrical and pH gradients across the inner mitochondrial membrane. After H+ have been transferred to the cytosolic side of the membrane, they reenter the matrix by passing through the Fo H+ channel in ATP synthase (Complex V), dissipating the pH and electrical gradients and causing conformational changes in the F1 β subunits of the synthase that result in the synthesis of ATP from ADP + inorganic phosphate. Electron transport and phosphorylation are tightly coupled in oxidative phosphorylation ([OXPHOS] Fig. 6.18). Inhibition of one process inhibits the other. These processes can be uncoupled by uncoupling protein-1 of the inner mitochondrial membrane of brown adipocytes and by synthetic compounds such as 2,4-dinitrophenol and aspirin, all of which dissipate the H+ gradient. In uncoupled mitochondria, the energy produced by electron transport is released as heat rather than being used to synthesize ATP. Mutations in mitochondrial DNA, which |
Biochemistry_Lippincott_275 | Biochemistry_Lippinco | dissipate the H+ gradient. In uncoupled mitochondria, the energy produced by electron transport is released as heat rather than being used to synthesize ATP. Mutations in mitochondrial DNA, which is maternally inherited, are responsible for some cases of mitochondrial diseases such as Leber hereditary optic neuropathy. The release of cytochrome c into the cytoplasm and subsequent activation of proteolytic caspases results in apoptotic cell death. | Biochemistry_Lippinco. dissipate the H+ gradient. In uncoupled mitochondria, the energy produced by electron transport is released as heat rather than being used to synthesize ATP. Mutations in mitochondrial DNA, which is maternally inherited, are responsible for some cases of mitochondrial diseases such as Leber hereditary optic neuropathy. The release of cytochrome c into the cytoplasm and subsequent activation of proteolytic caspases results in apoptotic cell death. |
Biochemistry_Lippincott_276 | Biochemistry_Lippinco | Choose the ONE best answer. .1. 2,4-Dinitrophenol (DNP), an uncoupler of oxidative phosphorylation, was used as a weight-loss agent in the 1930s. Reports of fatal overdoses led to its discontinuation in 1939. Which of the following would most likely be true concerning individuals taking 2,4-DNP? A. ATP levels in the mitochondria are greater than normal. B. Body temperature is elevated as a result of hypermetabolism. C. Cyanide has no effect on electron flow. D. The proton gradient across the inner mitochondrial membrane is greater than normal. E. The rate of electron transport is abnormally low. | Biochemistry_Lippinco. Choose the ONE best answer. .1. 2,4-Dinitrophenol (DNP), an uncoupler of oxidative phosphorylation, was used as a weight-loss agent in the 1930s. Reports of fatal overdoses led to its discontinuation in 1939. Which of the following would most likely be true concerning individuals taking 2,4-DNP? A. ATP levels in the mitochondria are greater than normal. B. Body temperature is elevated as a result of hypermetabolism. C. Cyanide has no effect on electron flow. D. The proton gradient across the inner mitochondrial membrane is greater than normal. E. The rate of electron transport is abnormally low. |
Biochemistry_Lippincott_277 | Biochemistry_Lippinco | C. Cyanide has no effect on electron flow. D. The proton gradient across the inner mitochondrial membrane is greater than normal. E. The rate of electron transport is abnormally low. Correct answer = B. When phosphorylation is uncoupled from electron flow, a decrease in the proton gradient across the inner mitochondrial membrane and, therefore, impaired ATP synthesis are expected. In an attempt to compensate for this defect in energy capture, metabolism and electron flow to oxygen are increased. This hypermetabolism will be accompanied by elevated body temperature because the energy in fuels is largely wasted, appearing as heat. The electron transport chain will still be inhibited by cyanide. .2. Which of the following has the strongest tendency to gain electrons? A. Coenzyme Q B. Cytochrome c C. Flavin adenine dinucleotide D. Nicotinamide adenine dinucleotide E. Oxygen | Biochemistry_Lippinco. C. Cyanide has no effect on electron flow. D. The proton gradient across the inner mitochondrial membrane is greater than normal. E. The rate of electron transport is abnormally low. Correct answer = B. When phosphorylation is uncoupled from electron flow, a decrease in the proton gradient across the inner mitochondrial membrane and, therefore, impaired ATP synthesis are expected. In an attempt to compensate for this defect in energy capture, metabolism and electron flow to oxygen are increased. This hypermetabolism will be accompanied by elevated body temperature because the energy in fuels is largely wasted, appearing as heat. The electron transport chain will still be inhibited by cyanide. .2. Which of the following has the strongest tendency to gain electrons? A. Coenzyme Q B. Cytochrome c C. Flavin adenine dinucleotide D. Nicotinamide adenine dinucleotide E. Oxygen |
Biochemistry_Lippincott_278 | Biochemistry_Lippinco | .2. Which of the following has the strongest tendency to gain electrons? A. Coenzyme Q B. Cytochrome c C. Flavin adenine dinucleotide D. Nicotinamide adenine dinucleotide E. Oxygen Correct answer = E. Oxygen is the terminal acceptor of electrons in the electron transport chain (ETC). Electrons flow down the ETC to oxygen because it has the highest (most positive) reduction potential (E0). The other choices precede oxygen in the ETC and have lower E0 values. .3. Explain why and how the malate-aspartate shuttle moves nicotinamide adenine dinucleotide reducing equivalents from the cytosol to the mitochondrial matrix. | Biochemistry_Lippinco. .2. Which of the following has the strongest tendency to gain electrons? A. Coenzyme Q B. Cytochrome c C. Flavin adenine dinucleotide D. Nicotinamide adenine dinucleotide E. Oxygen Correct answer = E. Oxygen is the terminal acceptor of electrons in the electron transport chain (ETC). Electrons flow down the ETC to oxygen because it has the highest (most positive) reduction potential (E0). The other choices precede oxygen in the ETC and have lower E0 values. .3. Explain why and how the malate-aspartate shuttle moves nicotinamide adenine dinucleotide reducing equivalents from the cytosol to the mitochondrial matrix. |
Biochemistry_Lippincott_279 | Biochemistry_Lippinco | .3. Explain why and how the malate-aspartate shuttle moves nicotinamide adenine dinucleotide reducing equivalents from the cytosol to the mitochondrial matrix. There is no transporter for nicotinamide adenine dinucleotide (NADH) in the inner mitochondrial membrane. However, cytoplasmic NADH can be oxidized to NAD+ by malate dehydrogenase as oxaloacetate (OAA) is reduced to malate. The malate is transported across the inner membrane to the matrix where the mitochondrial isozyme of malate dehydrogenase oxidizes it to OAA as mitochondrial NAD+ is reduced to NADH. This NADH can be oxidized by Complex I of the electron transport chain, generating three ATP through the coupled processes of oxidative phosphorylation. .4. Carbon monoxide (CO) binds to and inhibits Complex IV of the electron transport chain. What effect, if any, should this respiratory inhibitor have on phosphorylation of adenosine diphosphate (ADP) to ATP? | Biochemistry_Lippinco. .3. Explain why and how the malate-aspartate shuttle moves nicotinamide adenine dinucleotide reducing equivalents from the cytosol to the mitochondrial matrix. There is no transporter for nicotinamide adenine dinucleotide (NADH) in the inner mitochondrial membrane. However, cytoplasmic NADH can be oxidized to NAD+ by malate dehydrogenase as oxaloacetate (OAA) is reduced to malate. The malate is transported across the inner membrane to the matrix where the mitochondrial isozyme of malate dehydrogenase oxidizes it to OAA as mitochondrial NAD+ is reduced to NADH. This NADH can be oxidized by Complex I of the electron transport chain, generating three ATP through the coupled processes of oxidative phosphorylation. .4. Carbon monoxide (CO) binds to and inhibits Complex IV of the electron transport chain. What effect, if any, should this respiratory inhibitor have on phosphorylation of adenosine diphosphate (ADP) to ATP? |
Biochemistry_Lippincott_280 | Biochemistry_Lippinco | Inhibition of electron transport by respiratory inhibitors such as CO results in an inability to maintain the proton (H+) gradient. Therefore, phosphorylation of ADP to ATP is inhibited, as are ancillary reactions such as calcium uptake by mitochondria, because they also require the H+ gradient. Introduction to Carbohydrates 7 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW | Biochemistry_Lippinco. Inhibition of electron transport by respiratory inhibitors such as CO results in an inability to maintain the proton (H+) gradient. Therefore, phosphorylation of ADP to ATP is inhibited, as are ancillary reactions such as calcium uptake by mitochondria, because they also require the H+ gradient. Introduction to Carbohydrates 7 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW |
Biochemistry_Lippincott_281 | Biochemistry_Lippinco | Introduction to Carbohydrates 7 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW Carbohydrates (saccharides) are the most abundant organic molecules in nature. They have a wide range of functions, including providing a significant fraction of the dietary calories for most organisms, acting as a storage form of energy in the body, and serving as cell membrane components that mediate some forms of intercellular communication. Carbohydrates also serve as a structural component of many organisms, including the cell walls of bacteria, the exoskeleton of insects, and the fibrous cellulose of plants. [Note: The full set of carbohydrates produced by an organism is its glycome.] The empiric formula for many of the simpler carbohydrates is (CH2O)n, where n ≥3, hence the name “hydrate of carbon.” II. CLASSIFICATION AND STRUCTURE | Biochemistry_Lippinco. Introduction to Carbohydrates 7 For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW Carbohydrates (saccharides) are the most abundant organic molecules in nature. They have a wide range of functions, including providing a significant fraction of the dietary calories for most organisms, acting as a storage form of energy in the body, and serving as cell membrane components that mediate some forms of intercellular communication. Carbohydrates also serve as a structural component of many organisms, including the cell walls of bacteria, the exoskeleton of insects, and the fibrous cellulose of plants. [Note: The full set of carbohydrates produced by an organism is its glycome.] The empiric formula for many of the simpler carbohydrates is (CH2O)n, where n ≥3, hence the name “hydrate of carbon.” II. CLASSIFICATION AND STRUCTURE |
Biochemistry_Lippincott_282 | Biochemistry_Lippinco | Monosaccharides (simple sugars) can be classified according to the number of carbon atoms they contain. Examples of some monosaccharides commonly found in humans are listed in Figure 7.1. They can also be classified by the type of carbonyl group they contain. Carbohydrates with an aldehyde as their carbonyl group are called aldoses, whereas those with a keto as their carbonyl group are called ketoses (Fig. 7.2). For example, glyceraldehyde is an aldose, whereas dihydroxyacetone is a ketose. Carbohydrates that have a free carbonyl group have the suffix -ose. [Note: Ketoses have an additional “ul” in their suffix such as xylulose. There are exceptions, such as fructose, to this rule.] Monosaccharides can be linked by glycosidic bonds to create larger structures (Fig. 7.3). Disaccharides contain two monosaccharide units, oligosaccharides contain three to ten monosaccharide units, and polysaccharides contain more than ten monosaccharide units and can be hundreds of sugar units in length. | Biochemistry_Lippinco. Monosaccharides (simple sugars) can be classified according to the number of carbon atoms they contain. Examples of some monosaccharides commonly found in humans are listed in Figure 7.1. They can also be classified by the type of carbonyl group they contain. Carbohydrates with an aldehyde as their carbonyl group are called aldoses, whereas those with a keto as their carbonyl group are called ketoses (Fig. 7.2). For example, glyceraldehyde is an aldose, whereas dihydroxyacetone is a ketose. Carbohydrates that have a free carbonyl group have the suffix -ose. [Note: Ketoses have an additional “ul” in their suffix such as xylulose. There are exceptions, such as fructose, to this rule.] Monosaccharides can be linked by glycosidic bonds to create larger structures (Fig. 7.3). Disaccharides contain two monosaccharide units, oligosaccharides contain three to ten monosaccharide units, and polysaccharides contain more than ten monosaccharide units and can be hundreds of sugar units in length. |
Subsets and Splits