id
stringlengths 14
28
| title
stringclasses 18
values | content
stringlengths 2
999
| contents
stringlengths 19
1.02k
|
---|---|---|---|
Anatomy_Gray_3000 | Anatomy_Gray | The tuberal region in the medial zone contains three nuclei: ventromedial, dorsomedial, and arcuate. The largest and best defined is the ventromedial nucleus, which functions as a satiety center to decrease feeding behavior. Posterior to the ventromedial nucleus is the dorsomedial nucleus, which functions in the behavioral expression of rage or aggressive behavior. Finally, the arcuate nucleus serves as a center for releasing hormones, which are transmitted by the tuberoinfundibular tract and hypophysial portal system to the anterior pituitary (eFig. 9.89). | Anatomy_Gray. The tuberal region in the medial zone contains three nuclei: ventromedial, dorsomedial, and arcuate. The largest and best defined is the ventromedial nucleus, which functions as a satiety center to decrease feeding behavior. Posterior to the ventromedial nucleus is the dorsomedial nucleus, which functions in the behavioral expression of rage or aggressive behavior. Finally, the arcuate nucleus serves as a center for releasing hormones, which are transmitted by the tuberoinfundibular tract and hypophysial portal system to the anterior pituitary (eFig. 9.89). |
Anatomy_Gray_3001 | Anatomy_Gray | The mammillary region or mammillary body is the final group of nuclei in the medial zone. Four nuclei comprise this region: medial, intermediate, lateral mammillary, and posterior hypothalamic. The best defined is the medial mammillary nucleus, as it is the primary site for the termination of axons from the postcommissural fornix. This pathway originates from the subiculum of the hippocampal complex and plays a key role in memory. The medial mammillary nuclei also connects to structures of the limbic system. The periventricular zone resides medial to the medial zone and adjacent to the ependymal cells of the third ventricle. Neurons from this zone predominantly synthesize releasing hormones. Axons from these neuronal cells project through the tuberoinfundibular tract to the hypophysial portal system to influence release of hormones from the anterior pituitary. | Anatomy_Gray. The mammillary region or mammillary body is the final group of nuclei in the medial zone. Four nuclei comprise this region: medial, intermediate, lateral mammillary, and posterior hypothalamic. The best defined is the medial mammillary nucleus, as it is the primary site for the termination of axons from the postcommissural fornix. This pathway originates from the subiculum of the hippocampal complex and plays a key role in memory. The medial mammillary nuclei also connects to structures of the limbic system. The periventricular zone resides medial to the medial zone and adjacent to the ependymal cells of the third ventricle. Neurons from this zone predominantly synthesize releasing hormones. Axons from these neuronal cells project through the tuberoinfundibular tract to the hypophysial portal system to influence release of hormones from the anterior pituitary. |
Anatomy_Gray_3002 | Anatomy_Gray | Through the review of the hypothalamus thus far, it is apparent that this small, 4-gram structure has a significant role in regulating visceral, endocrine, and behavioral system functions through multiple pathways. It is important to remember the majority of the pathways mentioned represent input–output relationships between the hypothalamus and other structures. For a review of neural and non-neural inputs and outputs, refer to the summary figures (eFig. 9.92A and B). Part XI: Olfactory and | Anatomy_Gray. Through the review of the hypothalamus thus far, it is apparent that this small, 4-gram structure has a significant role in regulating visceral, endocrine, and behavioral system functions through multiple pathways. It is important to remember the majority of the pathways mentioned represent input–output relationships between the hypothalamus and other structures. For a review of neural and non-neural inputs and outputs, refer to the summary figures (eFig. 9.92A and B). Part XI: Olfactory and |
Anatomy_Gray_3003 | Anatomy_Gray | Part XI: Olfactory and The sense of olfaction has a role in both pleasurable experiences and survival. The same receptors that allow us to enjoy the food we consume or experience odorants in the environment also help us avoid spoiled food or potentially hazardous situations like a fire. Unlike the other special sensory system pathways, the olfactory sensory pathway is unique in that it does not have a thalamic relay before reaching the primary olfactory cortex. In this section we will review the course of the neurons in the olfactory system and their connection to the limbic system. | Anatomy_Gray. Part XI: Olfactory and The sense of olfaction has a role in both pleasurable experiences and survival. The same receptors that allow us to enjoy the food we consume or experience odorants in the environment also help us avoid spoiled food or potentially hazardous situations like a fire. Unlike the other special sensory system pathways, the olfactory sensory pathway is unique in that it does not have a thalamic relay before reaching the primary olfactory cortex. In this section we will review the course of the neurons in the olfactory system and their connection to the limbic system. |
Anatomy_Gray_3004 | Anatomy_Gray | Three types of olfactory receptors make up the olfactory epithelium along the lateral and septal walls of the nasal cavity. These cells allow for regeneration (basal stem cells), support (sustentacular cells), and transmission of information (olfactory receptor neurons). Each olfactory receptor neuron has an olfactory vesicle with cilia that contain receptors for odorant molecules and an unmyelinated axon that passes through the cribriform plate to terminate in the olfactory bulb (eFig. 9.93). As the olfactory receptor neurons originate embryologically from the CNS, they are considered part of the CNS and not the PNS. | Anatomy_Gray. Three types of olfactory receptors make up the olfactory epithelium along the lateral and septal walls of the nasal cavity. These cells allow for regeneration (basal stem cells), support (sustentacular cells), and transmission of information (olfactory receptor neurons). Each olfactory receptor neuron has an olfactory vesicle with cilia that contain receptors for odorant molecules and an unmyelinated axon that passes through the cribriform plate to terminate in the olfactory bulb (eFig. 9.93). As the olfactory receptor neurons originate embryologically from the CNS, they are considered part of the CNS and not the PNS. |
Anatomy_Gray_3005 | Anatomy_Gray | After synapsing with the mitral cells in the glomeruli of the olfactory bulb, mitral cell axons converge to form the olfactory tract. The olfactory tract then divides into medial and lateral olfactory striae to reach different synaptic targets (eFig. 9.94). Some of the axons in the medial olfactory striae travel through the diagonal band to reach the septal area, whereas others cross the midline in the anterior commissure and inhibit mitral cell activity in the contralateral olfactory bulb to enhance localization of the olfactory stimulant. Axons in the lateral olfactory stria primarily terminate in the piriform cortex/primary olfactory cortex of the uncus and in the amygdala (eFig. 9.94). The medial forebrain bundle, traveling through the lateral hypothalamus, connects the olfactory cortex with both the hypothalamus and the brainstem to regulate autonomic responses such as arousal through the reticular formation, salivation, and gastric contraction. | Anatomy_Gray. After synapsing with the mitral cells in the glomeruli of the olfactory bulb, mitral cell axons converge to form the olfactory tract. The olfactory tract then divides into medial and lateral olfactory striae to reach different synaptic targets (eFig. 9.94). Some of the axons in the medial olfactory striae travel through the diagonal band to reach the septal area, whereas others cross the midline in the anterior commissure and inhibit mitral cell activity in the contralateral olfactory bulb to enhance localization of the olfactory stimulant. Axons in the lateral olfactory stria primarily terminate in the piriform cortex/primary olfactory cortex of the uncus and in the amygdala (eFig. 9.94). The medial forebrain bundle, traveling through the lateral hypothalamus, connects the olfactory cortex with both the hypothalamus and the brainstem to regulate autonomic responses such as arousal through the reticular formation, salivation, and gastric contraction. |
Anatomy_Gray_3006 | Anatomy_Gray | The limbic system is composed of several cortical and subcortical structures that participate in an intricate network of connections to regulate complicated behaviors such as memory, emotions, homeostatic functions, and motivational state. In this section we will review the major structures and pathways that form the limbic system. Grossly, the limbic lobe includes a ring-shaped area of cortical structures that border the brainstem. These cortical areas include the cingulate gyrus, parahippocampal gyrus, and subcallosal area (eFig. 9.12). Laterally, the insular cortex also participates in limbic system function (eFig. 9.10). Nuclear structures of the limbic system include the amygdala, hippocampal formation, anterior and mediodorsal thalamic nuclei, septal nuclei in the forebrain, and nucleus accumbens (eFig. 9.95). | Anatomy_Gray. The limbic system is composed of several cortical and subcortical structures that participate in an intricate network of connections to regulate complicated behaviors such as memory, emotions, homeostatic functions, and motivational state. In this section we will review the major structures and pathways that form the limbic system. Grossly, the limbic lobe includes a ring-shaped area of cortical structures that border the brainstem. These cortical areas include the cingulate gyrus, parahippocampal gyrus, and subcallosal area (eFig. 9.12). Laterally, the insular cortex also participates in limbic system function (eFig. 9.10). Nuclear structures of the limbic system include the amygdala, hippocampal formation, anterior and mediodorsal thalamic nuclei, septal nuclei in the forebrain, and nucleus accumbens (eFig. 9.95). |
Anatomy_Gray_3007 | Anatomy_Gray | The amygdaloid nucleus is an almond-shaped structure located anterior to the inferior horn of the lateral ventricle and tail of the caudate within the temporal lobe (eFig. 9.96). Structurally, the amygdala consists of three nuclear regions: a large basolateral group and a smaller corticomedial group, which includes the central nucleus. Functionally, the amygdala is primarily associated with the emotion of fear, but it also has an important role in autonomic and neuroendocrine pathways. Connections of the amygdala are predominantly bidirectional and follow three different pathways: the uncinate fasciculus, stria terminalis, and ventral amygdalofugal pathway (eFig. 9.97). Connections to cortical areas pass through the uncinate fasciculus, which progresses anterior to the amygdala. Projections to the septal area and hypothalamus follow the stria terminalis (eFig. 9.98A–D). Fibers forming the ventral amygdalofugal pathway project to thalamic nuclei and several brainstem and forebrain | Anatomy_Gray. The amygdaloid nucleus is an almond-shaped structure located anterior to the inferior horn of the lateral ventricle and tail of the caudate within the temporal lobe (eFig. 9.96). Structurally, the amygdala consists of three nuclear regions: a large basolateral group and a smaller corticomedial group, which includes the central nucleus. Functionally, the amygdala is primarily associated with the emotion of fear, but it also has an important role in autonomic and neuroendocrine pathways. Connections of the amygdala are predominantly bidirectional and follow three different pathways: the uncinate fasciculus, stria terminalis, and ventral amygdalofugal pathway (eFig. 9.97). Connections to cortical areas pass through the uncinate fasciculus, which progresses anterior to the amygdala. Projections to the septal area and hypothalamus follow the stria terminalis (eFig. 9.98A–D). Fibers forming the ventral amygdalofugal pathway project to thalamic nuclei and several brainstem and forebrain |
Anatomy_Gray_3008 | Anatomy_Gray | to the septal area and hypothalamus follow the stria terminalis (eFig. 9.98A–D). Fibers forming the ventral amygdalofugal pathway project to thalamic nuclei and several brainstem and forebrain structures (eFig. 9.98A–D). | Anatomy_Gray. to the septal area and hypothalamus follow the stria terminalis (eFig. 9.98A–D). Fibers forming the ventral amygdalofugal pathway project to thalamic nuclei and several brainstem and forebrain structures (eFig. 9.98A–D). |
Anatomy_Gray_3009 | Anatomy_Gray | The nucleus accumbens resides with the ventral forebrain adjacent to where the putamen and head of the caudate become continuous with one another (eFig. 9.99). Afferent axons to the nucleus accumbens come from the amygdala through the amygdalofugal pathway, hippocampal formation by way of the fornix, basal forebrain area from the stria terminalis, and ventral tegmentum through the medial forebrain bundle (eFig. 9.98A–D). Efferent axons leaving the nucleus accumbens project directly to the hypothalamus and globus pallidus and reach nuclei in the brainstem through the medial forebrain bundle. Its connections to the globus pallidus represent an important connection of the limbic system to the motor system. The overall function of the nucleus accumbens is recognized as a gratification center and has been shown to play a role in behaviors related to addiction. | Anatomy_Gray. The nucleus accumbens resides with the ventral forebrain adjacent to where the putamen and head of the caudate become continuous with one another (eFig. 9.99). Afferent axons to the nucleus accumbens come from the amygdala through the amygdalofugal pathway, hippocampal formation by way of the fornix, basal forebrain area from the stria terminalis, and ventral tegmentum through the medial forebrain bundle (eFig. 9.98A–D). Efferent axons leaving the nucleus accumbens project directly to the hypothalamus and globus pallidus and reach nuclei in the brainstem through the medial forebrain bundle. Its connections to the globus pallidus represent an important connection of the limbic system to the motor system. The overall function of the nucleus accumbens is recognized as a gratification center and has been shown to play a role in behaviors related to addiction. |
Anatomy_Gray_3010 | Anatomy_Gray | The septal region is located rostral to the anterior commissure along the medial aspect of the cerebral hemispheres (eFig. 9.99). This region appears to play a role in pleasurable behaviors. Conversely, lesion studies indicate that damage to this area evokes behaviors of extreme displeasure or rage. Afferent axons to the septal area arise from the amygdala, hippocampus, olfactory tract, and monoaminergic nuclei in the brainstem (eFigs. 9.100 and 9.101). The septal area also connects to a collection of cholinergic neurons along the wall and roof of the third ventricle known as the habenular nuclei. Axons from the habenular nuclei project to the interpeduncular nucleus of the reticular formation, which is believed to play a role in the sleep–wake cycle (eFig. 9.100). | Anatomy_Gray. The septal region is located rostral to the anterior commissure along the medial aspect of the cerebral hemispheres (eFig. 9.99). This region appears to play a role in pleasurable behaviors. Conversely, lesion studies indicate that damage to this area evokes behaviors of extreme displeasure or rage. Afferent axons to the septal area arise from the amygdala, hippocampus, olfactory tract, and monoaminergic nuclei in the brainstem (eFigs. 9.100 and 9.101). The septal area also connects to a collection of cholinergic neurons along the wall and roof of the third ventricle known as the habenular nuclei. Axons from the habenular nuclei project to the interpeduncular nucleus of the reticular formation, which is believed to play a role in the sleep–wake cycle (eFig. 9.100). |
Anatomy_Gray_3011 | Anatomy_Gray | The hippocampal formation is located in the medial ventral temporal lobe (eFig. 9.102). It consists of the hippocampus, dentate gyrus, and subiculum (eFig. 9.103A and B). The hippocampal formation plays a role in memory processes such as episodic memory, short-term memory, working memory, and consolidation of memories. Input to the hippocampal formation is primarily received by the entorhinal cortex from association cortices. Because of this, it is believed that the “storage” of memories is in the association and primary cortices, not in the medial temporal lobe. Neurons from the entorhinal cortex project to the hippocampal formation by two pathways: the perforant pathway and alvear pathway. The perforant courses directly through the hippocampal sulcus to reach the dentate gyrus (eFig. 9.104B). As the hippocampus resembles the appearance of a ram’s horn, early on it was named cornu ammonis (horn of Ammon). Based on its cytoarchitecture, it was subdivided into four regions termed the | Anatomy_Gray. The hippocampal formation is located in the medial ventral temporal lobe (eFig. 9.102). It consists of the hippocampus, dentate gyrus, and subiculum (eFig. 9.103A and B). The hippocampal formation plays a role in memory processes such as episodic memory, short-term memory, working memory, and consolidation of memories. Input to the hippocampal formation is primarily received by the entorhinal cortex from association cortices. Because of this, it is believed that the “storage” of memories is in the association and primary cortices, not in the medial temporal lobe. Neurons from the entorhinal cortex project to the hippocampal formation by two pathways: the perforant pathway and alvear pathway. The perforant courses directly through the hippocampal sulcus to reach the dentate gyrus (eFig. 9.104B). As the hippocampus resembles the appearance of a ram’s horn, early on it was named cornu ammonis (horn of Ammon). Based on its cytoarchitecture, it was subdivided into four regions termed the |
Anatomy_Gray_3012 | Anatomy_Gray | As the hippocampus resembles the appearance of a ram’s horn, early on it was named cornu ammonis (horn of Ammon). Based on its cytoarchitecture, it was subdivided into four regions termed the cornu ammonis 1 to 4 (eFig. 9.104A). From the dentate gyrus, axons project to CA3 of the hippocampus. Axons from the hippocampus leave via the fornix or as Shaffer collaterals to reach CA1. Axons from CA1 may enter the fornix or project to the subiculum. Finally, axons from the subiculum enter the fornix or go back to the entorhinal cortex. | Anatomy_Gray. As the hippocampus resembles the appearance of a ram’s horn, early on it was named cornu ammonis (horn of Ammon). Based on its cytoarchitecture, it was subdivided into four regions termed the cornu ammonis 1 to 4 (eFig. 9.104A). From the dentate gyrus, axons project to CA3 of the hippocampus. Axons from the hippocampus leave via the fornix or as Shaffer collaterals to reach CA1. Axons from CA1 may enter the fornix or project to the subiculum. Finally, axons from the subiculum enter the fornix or go back to the entorhinal cortex. |
Anatomy_Gray_3013 | Anatomy_Gray | A second afferent pathway from the entorhinal cortex to the hippocampal formation is through the alvear pathway. Axons in the alvear pathway project directly on to CA1 and CA3 of the hippocampus (eFig. 9.104B). Similar to the perforant pathway, axons leaving the alvear pathway primarily originate from CA1 and CA3, which then project to the subiculum. | Anatomy_Gray. A second afferent pathway from the entorhinal cortex to the hippocampal formation is through the alvear pathway. Axons in the alvear pathway project directly on to CA1 and CA3 of the hippocampus (eFig. 9.104B). Similar to the perforant pathway, axons leaving the alvear pathway primarily originate from CA1 and CA3, which then project to the subiculum. |
Anatomy_Gray_3014 | Anatomy_Gray | Efferent axons leaving the hippocampal formation primarily exit from the subiculum and form the fornix (Latin for “arch”), a white matter structure that arches over the ventricular system (eFig. 9.95). The fornix begins with axons exiting the hippocampus to form the alveus along the ventricular surface of the hippocampus. As the axons come together medially, they form a bundle referred to as the fimbria of the fornix. The fornix then emerges from the hippocampal formation and curves under the corpus callosum before coursing medially to run adjacent to the midline (eFig. 9.105). At the anterior commissure, the fornix divides into a precommissural fornix and postcommissural fornix to reach the nucleus accumbens, septal nuclei, medial frontal cortex, mammillary nucleus, ventromedial nucleus of the hypothalamus, and anterior nucleus of the dorsal thalamus (eFig. 9.95). | Anatomy_Gray. Efferent axons leaving the hippocampal formation primarily exit from the subiculum and form the fornix (Latin for “arch”), a white matter structure that arches over the ventricular system (eFig. 9.95). The fornix begins with axons exiting the hippocampus to form the alveus along the ventricular surface of the hippocampus. As the axons come together medially, they form a bundle referred to as the fimbria of the fornix. The fornix then emerges from the hippocampal formation and curves under the corpus callosum before coursing medially to run adjacent to the midline (eFig. 9.105). At the anterior commissure, the fornix divides into a precommissural fornix and postcommissural fornix to reach the nucleus accumbens, septal nuclei, medial frontal cortex, mammillary nucleus, ventromedial nucleus of the hypothalamus, and anterior nucleus of the dorsal thalamus (eFig. 9.95). |
Anatomy_Gray_3015 | Anatomy_Gray | Through this section we have described a collection of anatomical structures and defined their connections with other areas of the brain and brainstem without exploring how these individual structures are interconnected with one another. In the 1930s James Papez, an American neurologist, described a circuit that links these structures and cortical areas together in a way that was thought to be involved in the experience and expression of emotion. This is referred to as the Papez circuit (eFig. 9.106). The circuit begins with fibers from the subiculum, which then enter the fornix to reach the mammillary nuclei. These axons then project through the mammillothalamic tract to the anterior nucleus of the thalamus. Next, the axons from the anterior nucleus of the thalamus project through the internal capsule to the cingulate gyrus. Lastly, the cingulum fibers from the cingulate cortex project to the parahippocampal gyrus and then the entorhinal cortex and hippocampal formation (eFig. | Anatomy_Gray. Through this section we have described a collection of anatomical structures and defined their connections with other areas of the brain and brainstem without exploring how these individual structures are interconnected with one another. In the 1930s James Papez, an American neurologist, described a circuit that links these structures and cortical areas together in a way that was thought to be involved in the experience and expression of emotion. This is referred to as the Papez circuit (eFig. 9.106). The circuit begins with fibers from the subiculum, which then enter the fornix to reach the mammillary nuclei. These axons then project through the mammillothalamic tract to the anterior nucleus of the thalamus. Next, the axons from the anterior nucleus of the thalamus project through the internal capsule to the cingulate gyrus. Lastly, the cingulum fibers from the cingulate cortex project to the parahippocampal gyrus and then the entorhinal cortex and hippocampal formation (eFig. |
Anatomy_Gray_3016 | Anatomy_Gray | the internal capsule to the cingulate gyrus. Lastly, the cingulum fibers from the cingulate cortex project to the parahippocampal gyrus and then the entorhinal cortex and hippocampal formation (eFig. 9.106). Papez’s description of this circuit is useful for reviewing the major limbic system pathways; however, the role of some of the structures in the pathway has been shown to play little or no role in the expression of emotion. In addition, many of the structures that do play a role in the expression of emotion also have a role in other functions. | Anatomy_Gray. the internal capsule to the cingulate gyrus. Lastly, the cingulum fibers from the cingulate cortex project to the parahippocampal gyrus and then the entorhinal cortex and hippocampal formation (eFig. 9.106). Papez’s description of this circuit is useful for reviewing the major limbic system pathways; however, the role of some of the structures in the pathway has been shown to play little or no role in the expression of emotion. In addition, many of the structures that do play a role in the expression of emotion also have a role in other functions. |
Biochemistry_Lippincott_0 | Biochemistry_Lippinco | For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW | Biochemistry_Lippinco. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW |
Biochemistry_Lippincott_1 | Biochemistry_Lippinco | Proteins are the most abundant and functionally diverse molecules in living systems. Virtually every life process depends on this class of macromolecules. For example, enzymes and polypeptide hormones direct and regulate metabolism in the body, whereas contractile proteins in muscle permit movement. In bone, the protein collagen forms a framework for the deposition of calcium phosphate crystals, acting like the steel cables in reinforced concrete. In the bloodstream, proteins, such as hemoglobin and albumin, transport molecules essential to life, whereas immunoglobulins fight infectious bacteria and viruses. In short, proteins display an incredible diversity of functions, yet all share the common structural feature of being linear polymers of amino acids. This chapter describes the properties of amino acids. Chapter 2 explores how these simple building blocks are joined to form proteins that have unique three-dimensional structures, making them capable of performing specific biologic | Biochemistry_Lippinco. Proteins are the most abundant and functionally diverse molecules in living systems. Virtually every life process depends on this class of macromolecules. For example, enzymes and polypeptide hormones direct and regulate metabolism in the body, whereas contractile proteins in muscle permit movement. In bone, the protein collagen forms a framework for the deposition of calcium phosphate crystals, acting like the steel cables in reinforced concrete. In the bloodstream, proteins, such as hemoglobin and albumin, transport molecules essential to life, whereas immunoglobulins fight infectious bacteria and viruses. In short, proteins display an incredible diversity of functions, yet all share the common structural feature of being linear polymers of amino acids. This chapter describes the properties of amino acids. Chapter 2 explores how these simple building blocks are joined to form proteins that have unique three-dimensional structures, making them capable of performing specific biologic |
Biochemistry_Lippincott_2 | Biochemistry_Lippinco | of amino acids. Chapter 2 explores how these simple building blocks are joined to form proteins that have unique three-dimensional structures, making them capable of performing specific biologic functions. | Biochemistry_Lippinco. of amino acids. Chapter 2 explores how these simple building blocks are joined to form proteins that have unique three-dimensional structures, making them capable of performing specific biologic functions. |
Biochemistry_Lippincott_3 | Biochemistry_Lippinco | II. STRUCTURE | Biochemistry_Lippinco. II. STRUCTURE |
Biochemistry_Lippincott_4 | Biochemistry_Lippinco | Although >300 different amino acids have been described in nature, only 20 are commonly found as constituents of mammalian proteins. [Note: These standard amino acids are the only amino acids that are encoded by DNA, the genetic material in the cell (see p. 411). Nonstandard amino acids are produced by chemical modification of standard amino acids (see p. 45).] Each amino acid has a carboxyl group, a primary amino group (except for proline, which has a secondary amino group), and a distinctive side chain (R group) bonded to the α carbon atom. At physiologic pH (~7.4), the carboxyl group is dissociated, forming the negatively charged carboxylate ion (−COO−), and the amino group is protonated (−NH3+) (Fig. 1.1A). In proteins, almost all of these carboxyl and amino groups are combined through peptide linkage and, in general, are not available for chemical reaction except for hydrogen bond formation (Fig. 1.1B). Thus, it is the nature of the side chains that ultimately dictates the role | Biochemistry_Lippinco. Although >300 different amino acids have been described in nature, only 20 are commonly found as constituents of mammalian proteins. [Note: These standard amino acids are the only amino acids that are encoded by DNA, the genetic material in the cell (see p. 411). Nonstandard amino acids are produced by chemical modification of standard amino acids (see p. 45).] Each amino acid has a carboxyl group, a primary amino group (except for proline, which has a secondary amino group), and a distinctive side chain (R group) bonded to the α carbon atom. At physiologic pH (~7.4), the carboxyl group is dissociated, forming the negatively charged carboxylate ion (−COO−), and the amino group is protonated (−NH3+) (Fig. 1.1A). In proteins, almost all of these carboxyl and amino groups are combined through peptide linkage and, in general, are not available for chemical reaction except for hydrogen bond formation (Fig. 1.1B). Thus, it is the nature of the side chains that ultimately dictates the role |
Biochemistry_Lippincott_5 | Biochemistry_Lippinco | peptide linkage and, in general, are not available for chemical reaction except for hydrogen bond formation (Fig. 1.1B). Thus, it is the nature of the side chains that ultimately dictates the role an amino acid plays in a protein. Therefore, it is useful to classify the amino acids according to the properties of their side chains, that is, whether they are nonpolar (have an even distribution of electrons) or polar (have an uneven distribution of electrons, such as acids and bases) as shown in Figures 1.2 and 1.3. | Biochemistry_Lippinco. peptide linkage and, in general, are not available for chemical reaction except for hydrogen bond formation (Fig. 1.1B). Thus, it is the nature of the side chains that ultimately dictates the role an amino acid plays in a protein. Therefore, it is useful to classify the amino acids according to the properties of their side chains, that is, whether they are nonpolar (have an even distribution of electrons) or polar (have an uneven distribution of electrons, such as acids and bases) as shown in Figures 1.2 and 1.3. |
Biochemistry_Lippincott_6 | Biochemistry_Lippinco | and polarity of their side chains at acidic pH (continued from Fig. 1.2). [Note: At physiologic pH (7.35 to 7.45), the α-carboxyl groups, the acidic side chains, and the side chain of free histidine are deprotonated.] A. Amino acids with nonpolar side chains Each of these amino acids has a nonpolar side chain that does not gain or lose protons or participate in hydrogen or ionic bonds (see Fig. 1.2). The side chains of these amino acids can be thought of as “oily” or lipid-like, a property that promotes hydrophobic interactions (see Fig. 2.10, p. 19). | Biochemistry_Lippinco. and polarity of their side chains at acidic pH (continued from Fig. 1.2). [Note: At physiologic pH (7.35 to 7.45), the α-carboxyl groups, the acidic side chains, and the side chain of free histidine are deprotonated.] A. Amino acids with nonpolar side chains Each of these amino acids has a nonpolar side chain that does not gain or lose protons or participate in hydrogen or ionic bonds (see Fig. 1.2). The side chains of these amino acids can be thought of as “oily” or lipid-like, a property that promotes hydrophobic interactions (see Fig. 2.10, p. 19). |
Biochemistry_Lippincott_7 | Biochemistry_Lippinco | 1. Location in proteins: In proteins found in aqueous solutions (a polar environment), the side chains of the nonpolar amino acids tend to cluster together in the interior of the protein (Fig. 1.4). This phenomenon, known as the hydrophobic effect, is the result of the hydrophobicity of the nonpolar R groups, which act much like droplets of oil that coalesce in an aqueous environment. By filling up the interior of the folded protein, these nonpolar R groups help give the protein its three-dimensional shape. However, for proteins that are located in a hydrophobic environment, such as a membrane, the nonpolar R groups are found on the outside surface of the protein, interacting with the lipid environment (see Fig. 1.4). The importance of these hydrophobic interactions in stabilizing protein structure is discussed on p. 19. | Biochemistry_Lippinco. 1. Location in proteins: In proteins found in aqueous solutions (a polar environment), the side chains of the nonpolar amino acids tend to cluster together in the interior of the protein (Fig. 1.4). This phenomenon, known as the hydrophobic effect, is the result of the hydrophobicity of the nonpolar R groups, which act much like droplets of oil that coalesce in an aqueous environment. By filling up the interior of the folded protein, these nonpolar R groups help give the protein its three-dimensional shape. However, for proteins that are located in a hydrophobic environment, such as a membrane, the nonpolar R groups are found on the outside surface of the protein, interacting with the lipid environment (see Fig. 1.4). The importance of these hydrophobic interactions in stabilizing protein structure is discussed on p. 19. |
Biochemistry_Lippincott_8 | Biochemistry_Lippinco | Sickle cell anemia, a disease of red blood cells that causes them to become sickle shaped rather than disc shaped, results from the replacement of polar glutamate with nonpolar valine at the sixth position in the β subunit of hemoglobin A (see p. 36). 2. Proline: Proline differs from other amino acids in that its side chain and α-amino nitrogen form a rigid, five-membered ring structure (Fig. 1.5). Proline, then, has a secondary (rather than a primary) amino group. It is frequently referred to as an “imino acid.” The unique geometry of proline contributes to the formation of the fibrous structure of collagen (see p. 45), but it interrupts the α-helices found in globular proteins (see p. 16). B. Amino acids with uncharged polar side chains | Biochemistry_Lippinco. Sickle cell anemia, a disease of red blood cells that causes them to become sickle shaped rather than disc shaped, results from the replacement of polar glutamate with nonpolar valine at the sixth position in the β subunit of hemoglobin A (see p. 36). 2. Proline: Proline differs from other amino acids in that its side chain and α-amino nitrogen form a rigid, five-membered ring structure (Fig. 1.5). Proline, then, has a secondary (rather than a primary) amino group. It is frequently referred to as an “imino acid.” The unique geometry of proline contributes to the formation of the fibrous structure of collagen (see p. 45), but it interrupts the α-helices found in globular proteins (see p. 16). B. Amino acids with uncharged polar side chains |
Biochemistry_Lippincott_9 | Biochemistry_Lippinco | B. Amino acids with uncharged polar side chains These amino acids have zero net charge at physiologic pH, although the side chains of cysteine and tyrosine can lose a proton at an alkaline pH (see Fig. 1.3). Serine, threonine, and tyrosine each contain a polar hydroxyl group that can participate in hydrogen bond formation (Fig. 1.6). The side chains of asparagine and glutamine each contain a carbonyl group and an amide group, both of which can also participate in hydrogen bonds. 1. Disulfide bond: The side chain of cysteine contains a sulfhydryl (thiol) group (−SH), which is an important component of the active site of many enzymes. In proteins, the –SH groups of two cysteines can be oxidized to form a covalent cross-link called a disulfide bond (−S–S–). Two disulfide-linked cysteines are referred to as cystine. (See p. 19 for a further discussion of disulfide bond formation.) | Biochemistry_Lippinco. B. Amino acids with uncharged polar side chains These amino acids have zero net charge at physiologic pH, although the side chains of cysteine and tyrosine can lose a proton at an alkaline pH (see Fig. 1.3). Serine, threonine, and tyrosine each contain a polar hydroxyl group that can participate in hydrogen bond formation (Fig. 1.6). The side chains of asparagine and glutamine each contain a carbonyl group and an amide group, both of which can also participate in hydrogen bonds. 1. Disulfide bond: The side chain of cysteine contains a sulfhydryl (thiol) group (−SH), which is an important component of the active site of many enzymes. In proteins, the –SH groups of two cysteines can be oxidized to form a covalent cross-link called a disulfide bond (−S–S–). Two disulfide-linked cysteines are referred to as cystine. (See p. 19 for a further discussion of disulfide bond formation.) |
Biochemistry_Lippincott_10 | Biochemistry_Lippinco | Many extracellular proteins are stabilized by disulfide bonds. Albumin, a blood protein that functions as a transporter for a variety of molecules, is an example. 2. Side chains as attachment sites for other compounds: The polar hydroxyl group of serine, threonine, and (rarely) tyrosine can serve as a site of attachment for structures such as a phosphate group. In addition, the amide group of asparagine, as well as the hydroxyl group of serine or threonine, can serve as a site of attachment for oligosaccharide chains in glycoproteins (see p. 165). C. Amino acids with acidic side chains The amino acids aspartic acid and glutamic acid are proton donors. At physiologic pH, the side chains of these amino acids are fully ionized, containing a negatively charged carboxylate group (−COO−). The fully ionized forms are called aspartate and glutamate. D. Amino acids with basic side chains | Biochemistry_Lippinco. Many extracellular proteins are stabilized by disulfide bonds. Albumin, a blood protein that functions as a transporter for a variety of molecules, is an example. 2. Side chains as attachment sites for other compounds: The polar hydroxyl group of serine, threonine, and (rarely) tyrosine can serve as a site of attachment for structures such as a phosphate group. In addition, the amide group of asparagine, as well as the hydroxyl group of serine or threonine, can serve as a site of attachment for oligosaccharide chains in glycoproteins (see p. 165). C. Amino acids with acidic side chains The amino acids aspartic acid and glutamic acid are proton donors. At physiologic pH, the side chains of these amino acids are fully ionized, containing a negatively charged carboxylate group (−COO−). The fully ionized forms are called aspartate and glutamate. D. Amino acids with basic side chains |
Biochemistry_Lippincott_11 | Biochemistry_Lippinco | D. Amino acids with basic side chains The side chains of the basic amino acids accept protons (see Fig. 1.3). At physiologic pH, the R groups of lysine and arginine are fully ionized and positively charged. In contrast, the free amino acid histidine is weakly basic and largely uncharged at physiologic pH. However, when histidine is incorporated into a protein, its R group can be either positively charged (protonated) or neutral, depending on the ionic environment provided by the protein. This important property of histidine contributes to the buffering role it plays in the functioning of such proteins as hemoglobin (see p. 30). [Note: Histidine is the only amino acid with a side chain that can ionize within the physiologic pH range.] E. Abbreviations and symbols for commonly occurring amino acids letter symbol (Fig. 1.7). The one-letter codes are determined by the following rules. 1. | Biochemistry_Lippinco. D. Amino acids with basic side chains The side chains of the basic amino acids accept protons (see Fig. 1.3). At physiologic pH, the R groups of lysine and arginine are fully ionized and positively charged. In contrast, the free amino acid histidine is weakly basic and largely uncharged at physiologic pH. However, when histidine is incorporated into a protein, its R group can be either positively charged (protonated) or neutral, depending on the ionic environment provided by the protein. This important property of histidine contributes to the buffering role it plays in the functioning of such proteins as hemoglobin (see p. 30). [Note: Histidine is the only amino acid with a side chain that can ionize within the physiologic pH range.] E. Abbreviations and symbols for commonly occurring amino acids letter symbol (Fig. 1.7). The one-letter codes are determined by the following rules. 1. |
Biochemistry_Lippincott_12 | Biochemistry_Lippinco | E. Abbreviations and symbols for commonly occurring amino acids letter symbol (Fig. 1.7). The one-letter codes are determined by the following rules. 1. Unique first letter: If only one amino acid begins with a given letter, then that letter is used as its symbol. For example, V = valine. 2. Most commonly occurring amino acids have priority: If more than one amino acid begins with a particular letter, the most common of these amino acids receives this letter as its symbol. For example, glycine is more common than glutamate, so G = glycine. 3. Similar sounding names: Some one-letter symbols sound like the amino acid they represent. For example, F = phenylalanine, or W = tryptophan (“twyptophan” as Elmer Fudd would say). 4. | Biochemistry_Lippinco. E. Abbreviations and symbols for commonly occurring amino acids letter symbol (Fig. 1.7). The one-letter codes are determined by the following rules. 1. Unique first letter: If only one amino acid begins with a given letter, then that letter is used as its symbol. For example, V = valine. 2. Most commonly occurring amino acids have priority: If more than one amino acid begins with a particular letter, the most common of these amino acids receives this letter as its symbol. For example, glycine is more common than glutamate, so G = glycine. 3. Similar sounding names: Some one-letter symbols sound like the amino acid they represent. For example, F = phenylalanine, or W = tryptophan (“twyptophan” as Elmer Fudd would say). 4. |
Biochemistry_Lippincott_13 | Biochemistry_Lippinco | 3. Similar sounding names: Some one-letter symbols sound like the amino acid they represent. For example, F = phenylalanine, or W = tryptophan (“twyptophan” as Elmer Fudd would say). 4. Letter close to initial letter: For the remaining amino acids, a one-letter symbol is assigned that is as close in the alphabet as possible to the initial letter of the amino acid, for example, K = lysine. Furthermore, B is assigned to Asx, signifying either aspartic acid or asparagine; Z is assigned to Glx, signifying either glutamic acid or glutamine; and X is assigned to an unidentified amino acid. F. Amino acid isomers | Biochemistry_Lippinco. 3. Similar sounding names: Some one-letter symbols sound like the amino acid they represent. For example, F = phenylalanine, or W = tryptophan (“twyptophan” as Elmer Fudd would say). 4. Letter close to initial letter: For the remaining amino acids, a one-letter symbol is assigned that is as close in the alphabet as possible to the initial letter of the amino acid, for example, K = lysine. Furthermore, B is assigned to Asx, signifying either aspartic acid or asparagine; Z is assigned to Glx, signifying either glutamic acid or glutamine; and X is assigned to an unidentified amino acid. F. Amino acid isomers |
Biochemistry_Lippincott_14 | Biochemistry_Lippinco | F. Amino acid isomers Because the α-carbon of an amino acid is attached to four different chemical groups, it is an asymmetric (chiral) atom. Glycine is the exception because its α-carbon has two hydrogen substituents. Amino acids with a chiral αcarbon exist in two different isomeric forms, designated D and L, which are enantiomers, or mirror images (Fig. 1.8). [Note: Enantiomers are optically active. If an isomer, either D or L, causes the plane of polarized light to rotate clockwise, it is designated the (+) form.] All amino acids found in mammalian proteins are of the L configuration. However, D-amino acids are found in some antibiotics and in bacterial cell walls (see p. 252). [Note: Racemases enzymatically interconvert the D-and L-isomers of free amino acids.] III. ACIDIC AND BASIC PROPERTIES | Biochemistry_Lippinco. F. Amino acid isomers Because the α-carbon of an amino acid is attached to four different chemical groups, it is an asymmetric (chiral) atom. Glycine is the exception because its α-carbon has two hydrogen substituents. Amino acids with a chiral αcarbon exist in two different isomeric forms, designated D and L, which are enantiomers, or mirror images (Fig. 1.8). [Note: Enantiomers are optically active. If an isomer, either D or L, causes the plane of polarized light to rotate clockwise, it is designated the (+) form.] All amino acids found in mammalian proteins are of the L configuration. However, D-amino acids are found in some antibiotics and in bacterial cell walls (see p. 252). [Note: Racemases enzymatically interconvert the D-and L-isomers of free amino acids.] III. ACIDIC AND BASIC PROPERTIES |
Biochemistry_Lippincott_15 | Biochemistry_Lippinco | III. ACIDIC AND BASIC PROPERTIES Amino acids in aqueous solution contain weakly acidic α-carboxyl groups and weakly basic α-amino groups. In addition, each of the acidic and basic amino acids contains an ionizable group in its side chain. Thus, both free amino acids and some amino acids combined in peptide linkages can act as buffers. Acids may be defined as proton donors and bases as proton acceptors. Acids (or bases) described as weak ionize to only a limited extent. The concentration of protons ([H+]) in aqueous solution is expressed as pH, where pH = log 1/[H+] or –log [H+]. The quantitative relationship between the pH of the solution and concentration of a weak acid (HA) and its conjugate base (A−) is described by the Henderson-Hasselbalch equation. A. Equation derivation | Biochemistry_Lippinco. III. ACIDIC AND BASIC PROPERTIES Amino acids in aqueous solution contain weakly acidic α-carboxyl groups and weakly basic α-amino groups. In addition, each of the acidic and basic amino acids contains an ionizable group in its side chain. Thus, both free amino acids and some amino acids combined in peptide linkages can act as buffers. Acids may be defined as proton donors and bases as proton acceptors. Acids (or bases) described as weak ionize to only a limited extent. The concentration of protons ([H+]) in aqueous solution is expressed as pH, where pH = log 1/[H+] or –log [H+]. The quantitative relationship between the pH of the solution and concentration of a weak acid (HA) and its conjugate base (A−) is described by the Henderson-Hasselbalch equation. A. Equation derivation |
Biochemistry_Lippincott_16 | Biochemistry_Lippinco | A. Equation derivation Consider the release of a proton by a weak acid represented by HA: The salt or conjugate base, A−, is the ionized form of a weak acid. By definition, the dissociation constant of the acid, Ka, is: [Note: The larger the Ka, the stronger the acid, because most of the HA has dissociated into H+ and A−. Conversely, the smaller the Ka, the less acid has dissociated and, therefore, the weaker the acid.] By solving for the [H+] in the above equation, taking the logarithm of both sides of the equation, multiplying both sides of the equation by −1, and substituting pH = −log [H+] and pKa = −log Ka, we obtain the Henderson-Hasselbalch equation: B. Buffers | Biochemistry_Lippinco. A. Equation derivation Consider the release of a proton by a weak acid represented by HA: The salt or conjugate base, A−, is the ionized form of a weak acid. By definition, the dissociation constant of the acid, Ka, is: [Note: The larger the Ka, the stronger the acid, because most of the HA has dissociated into H+ and A−. Conversely, the smaller the Ka, the less acid has dissociated and, therefore, the weaker the acid.] By solving for the [H+] in the above equation, taking the logarithm of both sides of the equation, multiplying both sides of the equation by −1, and substituting pH = −log [H+] and pKa = −log Ka, we obtain the Henderson-Hasselbalch equation: B. Buffers |
Biochemistry_Lippincott_17 | Biochemistry_Lippinco | B. Buffers A buffer is a solution that resists change in pH following the addition of an acid or base. A buffer can be created by mixing a weak acid (HA) with its conjugate base (A−). If an acid such as HCl is added to a buffer, A− can neutralize it, being converted to HA in the process. If a base is added, HA can likewise neutralize it, being converted to A− in the process. Maximum buffering capacity occurs at a pH equal to the pKa, but a conjugate acid-base pair can still serve as an effective buffer when the pH of a solution is within approximately ±1 pH unit of the pKa. If the amounts of HA and A− are equal, the pH is equal to the pKa. As shown in Figure 1.9, a solution containing acetic acid (HA = CH3 – COOH) and acetate (A− = CH3 – COO−) with a pKa of 4.8 resists a change in pH from pH 3.8 to 5.8, with maximum buffering at pH 4.8. At pH values less than the pKa, the protonated acid form (CH3 – COOH) is the predominant species in solution. | Biochemistry_Lippinco. B. Buffers A buffer is a solution that resists change in pH following the addition of an acid or base. A buffer can be created by mixing a weak acid (HA) with its conjugate base (A−). If an acid such as HCl is added to a buffer, A− can neutralize it, being converted to HA in the process. If a base is added, HA can likewise neutralize it, being converted to A− in the process. Maximum buffering capacity occurs at a pH equal to the pKa, but a conjugate acid-base pair can still serve as an effective buffer when the pH of a solution is within approximately ±1 pH unit of the pKa. If the amounts of HA and A− are equal, the pH is equal to the pKa. As shown in Figure 1.9, a solution containing acetic acid (HA = CH3 – COOH) and acetate (A− = CH3 – COO−) with a pKa of 4.8 resists a change in pH from pH 3.8 to 5.8, with maximum buffering at pH 4.8. At pH values less than the pKa, the protonated acid form (CH3 – COOH) is the predominant species in solution. |
Biochemistry_Lippincott_18 | Biochemistry_Lippinco | At pH values greater than the pKa, the deprotonated base form (CH3 – COO−) is the predominant species. C. Amino acid titration The titration curve of an amino acid can be analyzed in the same way as described for acetic acid. 1. Carboxyl group dissociation: Consider alanine, for example, which contains an ionizable α-carboxyl and α-amino group. [Note: Its –CH3 R group is nonionizable.] At a low (acidic) pH, both of these groups are protonated (Fig. 1.10). As the pH of the solution is raised, the −COOH group of form I can dissociate by donating a H+ to the medium. The release of a H+ results in the formation of the carboxylate group, −COO−. This structure is shown as form II, which is the dipolar form of the molecule (see Fig. 1.10). This form, also called a zwitterion (from the German word for “hybrid”), is the isoelectric form of alanine, that is, it has an overall (net) charge of zero. 2. | Biochemistry_Lippinco. At pH values greater than the pKa, the deprotonated base form (CH3 – COO−) is the predominant species. C. Amino acid titration The titration curve of an amino acid can be analyzed in the same way as described for acetic acid. 1. Carboxyl group dissociation: Consider alanine, for example, which contains an ionizable α-carboxyl and α-amino group. [Note: Its –CH3 R group is nonionizable.] At a low (acidic) pH, both of these groups are protonated (Fig. 1.10). As the pH of the solution is raised, the −COOH group of form I can dissociate by donating a H+ to the medium. The release of a H+ results in the formation of the carboxylate group, −COO−. This structure is shown as form II, which is the dipolar form of the molecule (see Fig. 1.10). This form, also called a zwitterion (from the German word for “hybrid”), is the isoelectric form of alanine, that is, it has an overall (net) charge of zero. 2. |
Biochemistry_Lippincott_19 | Biochemistry_Lippinco | 2. Application of the Henderson-Hasselbalch equation: The dissociation constant of the carboxyl group of an amino acid is called K1, rather than Ka, because the molecule contains a second titratable group. The Henderson-Hasselbalch equation can be used to analyze the dissociation of the carboxyl group of alanine in the same way as described for acetic acid: where I is the fully protonated form of alanine and II is the isoelectric form of alanine (see Fig. 1.10). This equation can be rearranged and converted to its logarithmic form to yield: 3. Amino group dissociation: The second titratable group of alanine is the amino (−NH3+) group shown in Figure 1.10. Because this is a much weaker acid than the –COOH group, it has a much smaller dissociation constant, K2. [Note: Its pKa is, therefore, larger.] Release of a H+ from the protonated amino group of form II results in the fully deprotonated form of alanine, form III (see Fig. 1.10). 4. | Biochemistry_Lippinco. 2. Application of the Henderson-Hasselbalch equation: The dissociation constant of the carboxyl group of an amino acid is called K1, rather than Ka, because the molecule contains a second titratable group. The Henderson-Hasselbalch equation can be used to analyze the dissociation of the carboxyl group of alanine in the same way as described for acetic acid: where I is the fully protonated form of alanine and II is the isoelectric form of alanine (see Fig. 1.10). This equation can be rearranged and converted to its logarithmic form to yield: 3. Amino group dissociation: The second titratable group of alanine is the amino (−NH3+) group shown in Figure 1.10. Because this is a much weaker acid than the –COOH group, it has a much smaller dissociation constant, K2. [Note: Its pKa is, therefore, larger.] Release of a H+ from the protonated amino group of form II results in the fully deprotonated form of alanine, form III (see Fig. 1.10). 4. |
Biochemistry_Lippincott_20 | Biochemistry_Lippinco | 4. Alanine pKs: The sequential dissociation of H+ from the carboxyl and amino groups of alanine is summarized in Figure 1.10. Each titratable group has a pKa that is numerically equal to the pH at which exactly one half of the H+ have been removed from that group. The pKa for the most acidic group (−COOH) is pK1, whereas the pKa for the next most acidic group (−NH3+) is pK2. [Note: The pKa of the α-carboxyl group of amino acids is ~2, whereas that of the α-amino group is ~9.] 5. Alanine titration curve: By applying the Henderson-Hasselbalch equation to each dissociable acidic group, it is possible to calculate the complete titration curve of a weak acid. Figure 1.11 shows the change in pH that occurs during the addition of base to the fully protonated form of alanine (I) to produce the completely deprotonated form (III). Note the following: a. | Biochemistry_Lippinco. 4. Alanine pKs: The sequential dissociation of H+ from the carboxyl and amino groups of alanine is summarized in Figure 1.10. Each titratable group has a pKa that is numerically equal to the pH at which exactly one half of the H+ have been removed from that group. The pKa for the most acidic group (−COOH) is pK1, whereas the pKa for the next most acidic group (−NH3+) is pK2. [Note: The pKa of the α-carboxyl group of amino acids is ~2, whereas that of the α-amino group is ~9.] 5. Alanine titration curve: By applying the Henderson-Hasselbalch equation to each dissociable acidic group, it is possible to calculate the complete titration curve of a weak acid. Figure 1.11 shows the change in pH that occurs during the addition of base to the fully protonated form of alanine (I) to produce the completely deprotonated form (III). Note the following: a. |
Biochemistry_Lippincott_21 | Biochemistry_Lippinco | Buffer pairs: The –COOH/–COO− pair can serve as a buffer in the pH region around pK1, and the –NH3+/–NH2 pair can buffer in the region around pK2. b. When pH = pK: When the pH is equal to pK1 (2.3), equal amounts of forms I and II of alanine exist in solution. When the pH is equal to pK2 (9.1), equal amounts of forms II and III are present in solution. c. | Biochemistry_Lippinco. Buffer pairs: The –COOH/–COO− pair can serve as a buffer in the pH region around pK1, and the –NH3+/–NH2 pair can buffer in the region around pK2. b. When pH = pK: When the pH is equal to pK1 (2.3), equal amounts of forms I and II of alanine exist in solution. When the pH is equal to pK2 (9.1), equal amounts of forms II and III are present in solution. c. |
Biochemistry_Lippincott_22 | Biochemistry_Lippinco | c. Isoelectric point: At neutral pH, alanine exists predominantly as the dipolar form II in which the amino and carboxyl groups are ionized, but the net charge is zero. The isoelectric point (pI) is the pH at which an amino acid is electrically neutral, that is, in which the sum of the positive charges equals the sum of the negative charges. For an amino acid, such as alanine, that has only two dissociable hydrogens (one from the α-carboxyl and one from the α-amino group), the pI is the average of pK1 and pK2 (pI = [2.3 + 9.1]/2 = 5.7) as shown in Figure 1.11. The pI is, thus, midway between pK1 (2.3) and pK2 (9.1). pI corresponds to the pH at which the form II (with a net charge of zero) predominates and at which there are also equal amounts of forms I (net charge of +1) and III (net charge of −1). | Biochemistry_Lippinco. c. Isoelectric point: At neutral pH, alanine exists predominantly as the dipolar form II in which the amino and carboxyl groups are ionized, but the net charge is zero. The isoelectric point (pI) is the pH at which an amino acid is electrically neutral, that is, in which the sum of the positive charges equals the sum of the negative charges. For an amino acid, such as alanine, that has only two dissociable hydrogens (one from the α-carboxyl and one from the α-amino group), the pI is the average of pK1 and pK2 (pI = [2.3 + 9.1]/2 = 5.7) as shown in Figure 1.11. The pI is, thus, midway between pK1 (2.3) and pK2 (9.1). pI corresponds to the pH at which the form II (with a net charge of zero) predominates and at which there are also equal amounts of forms I (net charge of +1) and III (net charge of −1). |
Biochemistry_Lippincott_23 | Biochemistry_Lippinco | Separation of plasma proteins by charge typically is done at a pH above the pI of the major proteins. Therefore, the charge on the proteins is negative. In an electric field, the proteins will move toward the positive electrode at a rate determined by their net negative charge. Variations in the mobility pattern are suggestive of certain diseases. 6. Net charge at neutral pH: At physiologic pH, amino acids have a negatively charged group (−COO−) and a positively charged group (−NH3+), both attached to the α-carbon. [Note: Glutamate, aspartate, histidine, arginine, and lysine have additional potentially charged groups in their side chains.] Substances such as amino acids that can act either as an acid or a base are defined as amphoteric and are referred to as ampholytes (amphoteric electrolytes). D. Other applications of the Henderson-Hasselbalch equation | Biochemistry_Lippinco. Separation of plasma proteins by charge typically is done at a pH above the pI of the major proteins. Therefore, the charge on the proteins is negative. In an electric field, the proteins will move toward the positive electrode at a rate determined by their net negative charge. Variations in the mobility pattern are suggestive of certain diseases. 6. Net charge at neutral pH: At physiologic pH, amino acids have a negatively charged group (−COO−) and a positively charged group (−NH3+), both attached to the α-carbon. [Note: Glutamate, aspartate, histidine, arginine, and lysine have additional potentially charged groups in their side chains.] Substances such as amino acids that can act either as an acid or a base are defined as amphoteric and are referred to as ampholytes (amphoteric electrolytes). D. Other applications of the Henderson-Hasselbalch equation |
Biochemistry_Lippincott_24 | Biochemistry_Lippinco | D. Other applications of the Henderson-Hasselbalch equation The Henderson-Hasselbalch equation can be used to calculate how the pH of a physiologic solution responds to changes in the concentration of a weak acid and/or its corresponding salt form. For example, in the bicarbonate buffer system, the Henderson-Hasselbalch equation predicts how shifts in the bicarbonate ion concentration, [HCO3−], and the carbon dioxide concentration [CO2] influence pH (Fig. 1.12A). The equation is also useful for calculating the abundance of ionic forms of acidic and basic drugs. For example, most drugs are either weak acids or weak bases (Fig. 1.12B). Acidic drugs (HA) release a H+, causing a charged anion (A−) to form. Weak bases (BH+) can also release a H+. However, the protonated form of basic drugs is usually charged, and the loss of a proton produces the uncharged base (B). | Biochemistry_Lippinco. D. Other applications of the Henderson-Hasselbalch equation The Henderson-Hasselbalch equation can be used to calculate how the pH of a physiologic solution responds to changes in the concentration of a weak acid and/or its corresponding salt form. For example, in the bicarbonate buffer system, the Henderson-Hasselbalch equation predicts how shifts in the bicarbonate ion concentration, [HCO3−], and the carbon dioxide concentration [CO2] influence pH (Fig. 1.12A). The equation is also useful for calculating the abundance of ionic forms of acidic and basic drugs. For example, most drugs are either weak acids or weak bases (Fig. 1.12B). Acidic drugs (HA) release a H+, causing a charged anion (A−) to form. Weak bases (BH+) can also release a H+. However, the protonated form of basic drugs is usually charged, and the loss of a proton produces the uncharged base (B). |
Biochemistry_Lippincott_25 | Biochemistry_Lippinco | A drug passes through membranes more readily if it is uncharged. Thus, for a weak acid, such as aspirin, the uncharged HA can permeate through membranes, but A− cannot. Likewise, for a weak base, such as morphine, the uncharged B form permeates through the cell membrane, but BH+ does not. Therefore, the effective concentration of the permeable form of each drug at its absorption site is determined by the relative concentrations of the charged (impermeant) and uncharged (permeant) forms. The ratio between the two forms is determined by the pH at the site of absorption and by the strength of the weak acid or base, which is represented by the pKa of the ionizable group. The Henderson-Hasselbalch equation is useful in determining how much drug is found on either side of a membrane that separates two compartments that differ in pH, for example, the stomach (pH 1.0–1.5) and blood plasma (pH 7.4). IV. CONCEPT MAPS | Biochemistry_Lippinco. A drug passes through membranes more readily if it is uncharged. Thus, for a weak acid, such as aspirin, the uncharged HA can permeate through membranes, but A− cannot. Likewise, for a weak base, such as morphine, the uncharged B form permeates through the cell membrane, but BH+ does not. Therefore, the effective concentration of the permeable form of each drug at its absorption site is determined by the relative concentrations of the charged (impermeant) and uncharged (permeant) forms. The ratio between the two forms is determined by the pH at the site of absorption and by the strength of the weak acid or base, which is represented by the pKa of the ionizable group. The Henderson-Hasselbalch equation is useful in determining how much drug is found on either side of a membrane that separates two compartments that differ in pH, for example, the stomach (pH 1.0–1.5) and blood plasma (pH 7.4). IV. CONCEPT MAPS |
Biochemistry_Lippincott_26 | Biochemistry_Lippinco | Students sometimes view biochemistry as a list of facts or equations to be memorized, rather than a body of concepts to be understood. Details provided to enrich understanding of these concepts inadvertently turn into distractions. What seems to be missing is a road map—a guide that provides the student with an understanding of how various topics fit together to “tell a story.” Therefore, in this text, a series of biochemical concept maps have been created to graphically illustrate relationships between ideas presented in a chapter and to show how the information can be grouped or organized. A concept map is, thus, a tool for visualizing the connections between concepts. Material is represented in a hierarchic fashion, with the most inclusive, most general concepts at the top of the map, and the more specific, less general concepts arranged beneath. The concept maps ideally function as templates or guides for organizing information, so the student can readily find the best ways to | Biochemistry_Lippinco. Students sometimes view biochemistry as a list of facts or equations to be memorized, rather than a body of concepts to be understood. Details provided to enrich understanding of these concepts inadvertently turn into distractions. What seems to be missing is a road map—a guide that provides the student with an understanding of how various topics fit together to “tell a story.” Therefore, in this text, a series of biochemical concept maps have been created to graphically illustrate relationships between ideas presented in a chapter and to show how the information can be grouped or organized. A concept map is, thus, a tool for visualizing the connections between concepts. Material is represented in a hierarchic fashion, with the most inclusive, most general concepts at the top of the map, and the more specific, less general concepts arranged beneath. The concept maps ideally function as templates or guides for organizing information, so the student can readily find the best ways to |
Biochemistry_Lippincott_27 | Biochemistry_Lippinco | and the more specific, less general concepts arranged beneath. The concept maps ideally function as templates or guides for organizing information, so the student can readily find the best ways to integrate new information into knowledge they already possess. Concept map construction is described below. | Biochemistry_Lippinco. and the more specific, less general concepts arranged beneath. The concept maps ideally function as templates or guides for organizing information, so the student can readily find the best ways to integrate new information into knowledge they already possess. Concept map construction is described below. |
Biochemistry_Lippincott_28 | Biochemistry_Lippinco | A. Concept boxes and links Educators define concepts as “perceived regularities in events or objects.” In the biochemical maps, concepts include abstractions (for example, free energy), processes (for example, oxidative phosphorylation), and compounds (for example, glucose 6-phosphate). These broadly defined concepts are prioritized with the central idea positioned at the top of the page. The concepts that follow from this central idea are then drawn in boxes (Fig. 1.13A). The size of the type indicates the relative importance of each idea. Lines are drawn between concept boxes to show which are related. The label on the line defines the relationship between two concepts, so that it reads as a valid statement (that is, the connection creates meaning). The lines with arrowheads indicate in which direction the connection should be read (Fig. 1.14). B. Cross-links | Biochemistry_Lippinco. A. Concept boxes and links Educators define concepts as “perceived regularities in events or objects.” In the biochemical maps, concepts include abstractions (for example, free energy), processes (for example, oxidative phosphorylation), and compounds (for example, glucose 6-phosphate). These broadly defined concepts are prioritized with the central idea positioned at the top of the page. The concepts that follow from this central idea are then drawn in boxes (Fig. 1.13A). The size of the type indicates the relative importance of each idea. Lines are drawn between concept boxes to show which are related. The label on the line defines the relationship between two concepts, so that it reads as a valid statement (that is, the connection creates meaning). The lines with arrowheads indicate in which direction the connection should be read (Fig. 1.14). B. Cross-links |
Biochemistry_Lippincott_29 | Biochemistry_Lippinco | B. Cross-links Unlike linear flow charts or outlines, concept maps may contain cross-links that allow the reader to visualize complex relationships between ideas represented in different parts of the map (Fig. 1.13B) or between the map and other chapters in this book (Fig. 1.13C). Cross-links can, thus, identify concepts that are central to more than one topic in biochemistry, empowering students to be effective in clinical situations and on the United States Medical Licensure Examination (USMLE) or other examinations that require integration of material. Students learn to visually perceive nonlinear relationships between facts, in contrast to cross-referencing within linear text. V. CHAPTER SUMMARY Each amino acid has an α-carboxyl group and a primary α-amino group (except for proline, which has a secondary amino group). At physiologic pH, the α-carboxyl group is dissociated, forming the negatively charged carboxylate ion (−COO−), and the α-amino group is protonated (−NH3+). | Biochemistry_Lippinco. B. Cross-links Unlike linear flow charts or outlines, concept maps may contain cross-links that allow the reader to visualize complex relationships between ideas represented in different parts of the map (Fig. 1.13B) or between the map and other chapters in this book (Fig. 1.13C). Cross-links can, thus, identify concepts that are central to more than one topic in biochemistry, empowering students to be effective in clinical situations and on the United States Medical Licensure Examination (USMLE) or other examinations that require integration of material. Students learn to visually perceive nonlinear relationships between facts, in contrast to cross-referencing within linear text. V. CHAPTER SUMMARY Each amino acid has an α-carboxyl group and a primary α-amino group (except for proline, which has a secondary amino group). At physiologic pH, the α-carboxyl group is dissociated, forming the negatively charged carboxylate ion (−COO−), and the α-amino group is protonated (−NH3+). |
Biochemistry_Lippincott_30 | Biochemistry_Lippinco | Each amino acid also contains one of 20 distinctive side chains attached to the α-carbon atom. The chemical nature of this R group determines the function of an amino acid in a protein and provides the basis for classification of the amino acids as nonpolar, uncharged polar, acidic (polar negative), or basic (polar positive). All free amino acids, plus charged amino acids in peptide chains, can serve as buffers. The quantitative relationship between the pH of a solution and the concentration of a weak acid (HA) and its conjugate base (A−) is described by the Henderson-Hasselbalch equation. Buffering occurs within ±1 pH unit of the pKa and is maximal when pH = pKa, at which [A−] = [HA]. Because the α-carbon of each amino acid (except glycine) is attached to four different chemical groups, it is asymmetric (chiral), and amino acids exist in D-and L-isomeric forms that are optically active mirror images (enantiomers). The L-form of amino acids is found in proteins synthesized by the | Biochemistry_Lippinco. Each amino acid also contains one of 20 distinctive side chains attached to the α-carbon atom. The chemical nature of this R group determines the function of an amino acid in a protein and provides the basis for classification of the amino acids as nonpolar, uncharged polar, acidic (polar negative), or basic (polar positive). All free amino acids, plus charged amino acids in peptide chains, can serve as buffers. The quantitative relationship between the pH of a solution and the concentration of a weak acid (HA) and its conjugate base (A−) is described by the Henderson-Hasselbalch equation. Buffering occurs within ±1 pH unit of the pKa and is maximal when pH = pKa, at which [A−] = [HA]. Because the α-carbon of each amino acid (except glycine) is attached to four different chemical groups, it is asymmetric (chiral), and amino acids exist in D-and L-isomeric forms that are optically active mirror images (enantiomers). The L-form of amino acids is found in proteins synthesized by the |
Biochemistry_Lippincott_31 | Biochemistry_Lippinco | it is asymmetric (chiral), and amino acids exist in D-and L-isomeric forms that are optically active mirror images (enantiomers). The L-form of amino acids is found in proteins synthesized by the human body. | Biochemistry_Lippinco. it is asymmetric (chiral), and amino acids exist in D-and L-isomeric forms that are optically active mirror images (enantiomers). The L-form of amino acids is found in proteins synthesized by the human body. |
Biochemistry_Lippincott_32 | Biochemistry_Lippinco | Choose the ONE best answer. .1. Which one of the following statements concerning the titration curve for a nonpolar amino acid is correct? The letters A through D designate certain regions on the curve below. A. Point A represents the region where the amino acid is deprotonated. B. Point B represents a region of minimal buffering. C. Point C represents the region where the net charge on the amino acid is zero. D. Point D represents the pK of the amino acid’s carboxyl group. E. The amino acid could be lysine. Correct answer = C. Point C represents the isoelectric point, or pI, and as such is midway between pK1 and pK2 for a nonpolar amino acid. The amino acid is fully protonated at Point A. Point B represents a region of maximum buffering, as does Point D. Lysine is a basic amino acid, and free lysine has an ionizable side chain in addition to the ionizable α-amino and α-carboxyl groups. | Biochemistry_Lippinco. Choose the ONE best answer. .1. Which one of the following statements concerning the titration curve for a nonpolar amino acid is correct? The letters A through D designate certain regions on the curve below. A. Point A represents the region where the amino acid is deprotonated. B. Point B represents a region of minimal buffering. C. Point C represents the region where the net charge on the amino acid is zero. D. Point D represents the pK of the amino acid’s carboxyl group. E. The amino acid could be lysine. Correct answer = C. Point C represents the isoelectric point, or pI, and as such is midway between pK1 and pK2 for a nonpolar amino acid. The amino acid is fully protonated at Point A. Point B represents a region of maximum buffering, as does Point D. Lysine is a basic amino acid, and free lysine has an ionizable side chain in addition to the ionizable α-amino and α-carboxyl groups. |
Biochemistry_Lippincott_33 | Biochemistry_Lippinco | .2. Which one of the following statements concerning the peptide shown below is correct?Val-Cys-Glu-Ser-Asp-Arg-Cys A. The peptide contains asparagine. B. The peptide contains a side chain with a secondary amino group. C. The peptide contains a side chain that can be phosphorylated. D. The peptide cannot form an internal disulfide bond. E. The peptide would move to the cathode (negative electrode) during electrophoresis at pH 5. Correct answer = C. The hydroxyl group of serine can accept a phosphate group. Asp is aspartate. Proline contains a secondary amino group. The two cysteine residues can, under oxidizing conditions, form a disulfide (covalent) bond. The net charge on the peptide at pH 5 is negative, and it would move to the anode. | Biochemistry_Lippinco. .2. Which one of the following statements concerning the peptide shown below is correct?Val-Cys-Glu-Ser-Asp-Arg-Cys A. The peptide contains asparagine. B. The peptide contains a side chain with a secondary amino group. C. The peptide contains a side chain that can be phosphorylated. D. The peptide cannot form an internal disulfide bond. E. The peptide would move to the cathode (negative electrode) during electrophoresis at pH 5. Correct answer = C. The hydroxyl group of serine can accept a phosphate group. Asp is aspartate. Proline contains a secondary amino group. The two cysteine residues can, under oxidizing conditions, form a disulfide (covalent) bond. The net charge on the peptide at pH 5 is negative, and it would move to the anode. |
Biochemistry_Lippincott_34 | Biochemistry_Lippinco | .3. A 2-year-old child presents with metabolic acidosis after ingesting an unknown number of flavored aspirin tablets. At presentation, her blood pH was 7.0. Given that the pKa of aspirin (salicylic acid) is 3, calculate the ratio of its ionized to unionized forms at pH 7.0. Correct answer = 10,000 to 1. pH = pKa + log [A−]/[HA]. Therefore, 7 = 3 + × and × = 4. The ratio of A− (ionized) to HA (unionized), then, is 10,000 to 1 because the log of 10,000 is 4. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW | Biochemistry_Lippinco. .3. A 2-year-old child presents with metabolic acidosis after ingesting an unknown number of flavored aspirin tablets. At presentation, her blood pH was 7.0. Given that the pKa of aspirin (salicylic acid) is 3, calculate the ratio of its ionized to unionized forms at pH 7.0. Correct answer = 10,000 to 1. pH = pKa + log [A−]/[HA]. Therefore, 7 = 3 + × and × = 4. The ratio of A− (ionized) to HA (unionized), then, is 10,000 to 1 because the log of 10,000 is 4. For additional ancillary materials related to this chapter, please visit thePoint. I. OVERVIEW |
Biochemistry_Lippincott_35 | Biochemistry_Lippinco | I. OVERVIEW The 20 amino acids commonly found in proteins are joined together by peptide bonds. The linear sequence of the linked amino acids contains the information necessary to generate a protein molecule with a unique three-dimensional shape that determines function. The complexity of protein structure is best analyzed by considering the molecule in terms of four organizational levels: primary, secondary, tertiary, and quaternary (Fig. 2.1). An examination of these hierarchies of increasing complexity has revealed that certain structural elements are repeated in a wide variety of proteins, suggesting that there are general rules regarding the ways in which proteins achieve their native, functional form. These repeated structural elements range from simple combinations of α-helices and β-sheets forming small motifs to the complex folding of polypeptide domains of multifunctional proteins (see p. 19). II. PRIMARY STRUCTURE | Biochemistry_Lippinco. I. OVERVIEW The 20 amino acids commonly found in proteins are joined together by peptide bonds. The linear sequence of the linked amino acids contains the information necessary to generate a protein molecule with a unique three-dimensional shape that determines function. The complexity of protein structure is best analyzed by considering the molecule in terms of four organizational levels: primary, secondary, tertiary, and quaternary (Fig. 2.1). An examination of these hierarchies of increasing complexity has revealed that certain structural elements are repeated in a wide variety of proteins, suggesting that there are general rules regarding the ways in which proteins achieve their native, functional form. These repeated structural elements range from simple combinations of α-helices and β-sheets forming small motifs to the complex folding of polypeptide domains of multifunctional proteins (see p. 19). II. PRIMARY STRUCTURE |
Biochemistry_Lippincott_36 | Biochemistry_Lippinco | II. PRIMARY STRUCTURE The sequence of amino acids in a protein is called the primary structure of the protein. Understanding the primary structure of proteins is important because many genetic diseases result in proteins with abnormal amino acid sequences, which cause improper folding and loss or impairment of normal function. If the primary structures of the normal and the mutated proteins are known, this information may be used to diagnose or study the disease. A. Peptide bond | Biochemistry_Lippinco. II. PRIMARY STRUCTURE The sequence of amino acids in a protein is called the primary structure of the protein. Understanding the primary structure of proteins is important because many genetic diseases result in proteins with abnormal amino acid sequences, which cause improper folding and loss or impairment of normal function. If the primary structures of the normal and the mutated proteins are known, this information may be used to diagnose or study the disease. A. Peptide bond |
Biochemistry_Lippincott_37 | Biochemistry_Lippinco | A. Peptide bond In proteins, amino acids are joined covalently by peptide bonds, which are amide linkages between the α-carboxyl group of one amino acid and the αamino group of another. For example, valine and alanine can form the dipeptide valylalanine through the formation of a peptide bond (Fig. 2.2). Peptide bonds are resistant to conditions that denature proteins, such as heating and high concentrations of urea (see p. 20). Prolonged exposure to a strong acid or base at elevated temperatures is required to break these bonds nonenzymically (see p. 14). 1. | Biochemistry_Lippinco. A. Peptide bond In proteins, amino acids are joined covalently by peptide bonds, which are amide linkages between the α-carboxyl group of one amino acid and the αamino group of another. For example, valine and alanine can form the dipeptide valylalanine through the formation of a peptide bond (Fig. 2.2). Peptide bonds are resistant to conditions that denature proteins, such as heating and high concentrations of urea (see p. 20). Prolonged exposure to a strong acid or base at elevated temperatures is required to break these bonds nonenzymically (see p. 14). 1. |
Biochemistry_Lippincott_38 | Biochemistry_Lippinco | 1. Naming the peptide: By convention, the free amino end (N-terminal) of the peptide chain is written to the left and the free carboxyl end (C-terminal) to the right. Therefore, all amino acid sequences are read from the N-to the C-terminal end. For example, in Figure 2.2A, the order of the amino acids in the dipeptide is valine, alanine. Linkage of ≥50 amino acids through peptide bonds results in an unbranched chain called a polypeptide, or protein. Each component amino acid is called a residue because it is the portion of the amino acid remaining after the atoms of water are lost in the formation of the peptide bond. When a peptide is named, all amino acid residues have their suffixes (-ine, -an, -ic, or -ate) changed to -yl, with the exception of the C-terminal amino acid. For example, a tripeptide composed of an N-terminal valine, a glycine, and a C-terminal leucine is called valylglycylleucine. 2. | Biochemistry_Lippinco. 1. Naming the peptide: By convention, the free amino end (N-terminal) of the peptide chain is written to the left and the free carboxyl end (C-terminal) to the right. Therefore, all amino acid sequences are read from the N-to the C-terminal end. For example, in Figure 2.2A, the order of the amino acids in the dipeptide is valine, alanine. Linkage of ≥50 amino acids through peptide bonds results in an unbranched chain called a polypeptide, or protein. Each component amino acid is called a residue because it is the portion of the amino acid remaining after the atoms of water are lost in the formation of the peptide bond. When a peptide is named, all amino acid residues have their suffixes (-ine, -an, -ic, or -ate) changed to -yl, with the exception of the C-terminal amino acid. For example, a tripeptide composed of an N-terminal valine, a glycine, and a C-terminal leucine is called valylglycylleucine. 2. |
Biochemistry_Lippincott_39 | Biochemistry_Lippinco | 2. Peptide bond characteristics: The peptide bond has a partial double-bond character, that is, it is shorter than a single bond and is rigid and planar (Fig. 2.2B). This prevents free rotation around the bond between the carbonyl carbon and the nitrogen of the peptide bond. However, the bonds between the α-carbons and the α-amino or α-carboxyl groups can be freely rotated (although they are limited by the size and character of the R groups). This allows the polypeptide chain to assume a variety of possible conformations. The peptide bond is almost always in the trans configuration (instead of the cis; see Fig. 2.2B), in large part because of steric interference of the R groups (side chains) when in the cis position. 3. | Biochemistry_Lippinco. 2. Peptide bond characteristics: The peptide bond has a partial double-bond character, that is, it is shorter than a single bond and is rigid and planar (Fig. 2.2B). This prevents free rotation around the bond between the carbonyl carbon and the nitrogen of the peptide bond. However, the bonds between the α-carbons and the α-amino or α-carboxyl groups can be freely rotated (although they are limited by the size and character of the R groups). This allows the polypeptide chain to assume a variety of possible conformations. The peptide bond is almost always in the trans configuration (instead of the cis; see Fig. 2.2B), in large part because of steric interference of the R groups (side chains) when in the cis position. 3. |
Biochemistry_Lippincott_40 | Biochemistry_Lippinco | 3. Peptide bond polarity: Like all amide linkages, the −C = O and −NH groups of the peptide bond are uncharged, and neither accept nor release protons over the pH range of 2–12. Thus, the charged groups present in polypeptides consist solely of the N-terminal (α-amino) group, the C-terminal (α-carboxyl) group, and any ionized groups present in the side chains of the constituent amino acids. The −C = O and −NH groups of the peptide bond are polar, however, and are involved in hydrogen bonds (for example, in α-helices and β-sheets), as described on pp. 16–17. B. Determining the amino acid composition of a polypeptide | Biochemistry_Lippinco. 3. Peptide bond polarity: Like all amide linkages, the −C = O and −NH groups of the peptide bond are uncharged, and neither accept nor release protons over the pH range of 2–12. Thus, the charged groups present in polypeptides consist solely of the N-terminal (α-amino) group, the C-terminal (α-carboxyl) group, and any ionized groups present in the side chains of the constituent amino acids. The −C = O and −NH groups of the peptide bond are polar, however, and are involved in hydrogen bonds (for example, in α-helices and β-sheets), as described on pp. 16–17. B. Determining the amino acid composition of a polypeptide |
Biochemistry_Lippincott_41 | Biochemistry_Lippinco | The first step in determining the primary structure of a polypeptide is to identify and quantitate its constituent amino acids. A purified sample of the polypeptide to be analyzed is first hydrolyzed by strong acid at 110°C for 24 hours. This treatment cleaves the peptide bonds and releases the individual amino acids, which can be separated by cation-exchange chromatography. In this technique, a mixture of amino acids is applied to a column that contains a resin to which a negatively charged group is tightly attached. [Note: If the attached group is positively charged, the column becomes an anion-exchange column.] The amino acids bind to the column with different affinities, depending on their charges, hydrophobicity, and other characteristics. Each amino acid is sequentially released from the chromatography column by eluting with solutions of increasing ionic strength and pH (Fig. 2.3). The separated amino acids contained in the eluate from the column are quantitated by heating them | Biochemistry_Lippinco. The first step in determining the primary structure of a polypeptide is to identify and quantitate its constituent amino acids. A purified sample of the polypeptide to be analyzed is first hydrolyzed by strong acid at 110°C for 24 hours. This treatment cleaves the peptide bonds and releases the individual amino acids, which can be separated by cation-exchange chromatography. In this technique, a mixture of amino acids is applied to a column that contains a resin to which a negatively charged group is tightly attached. [Note: If the attached group is positively charged, the column becomes an anion-exchange column.] The amino acids bind to the column with different affinities, depending on their charges, hydrophobicity, and other characteristics. Each amino acid is sequentially released from the chromatography column by eluting with solutions of increasing ionic strength and pH (Fig. 2.3). The separated amino acids contained in the eluate from the column are quantitated by heating them |
Biochemistry_Lippincott_42 | Biochemistry_Lippinco | the chromatography column by eluting with solutions of increasing ionic strength and pH (Fig. 2.3). The separated amino acids contained in the eluate from the column are quantitated by heating them with ninhydrin (a reagent that forms a purple compound with most amino acids, ammonia, and amines). The amount of each amino acid is determined spectrophotometrically by measuring the amount of light absorbed by the ninhydrin derivative. The analysis described above is performed using an amino acid analyzer, an automated machine whose components are depicted in Figure 2.3. | Biochemistry_Lippinco. the chromatography column by eluting with solutions of increasing ionic strength and pH (Fig. 2.3). The separated amino acids contained in the eluate from the column are quantitated by heating them with ninhydrin (a reagent that forms a purple compound with most amino acids, ammonia, and amines). The amount of each amino acid is determined spectrophotometrically by measuring the amount of light absorbed by the ninhydrin derivative. The analysis described above is performed using an amino acid analyzer, an automated machine whose components are depicted in Figure 2.3. |
Biochemistry_Lippincott_43 | Biochemistry_Lippinco | C. Sequencing the peptide from its N-terminal end Sequencing is a stepwise process of identifying the specific amino acid at each position in the peptide chain, beginning at the N-terminal end. Phenylisothiocyanate, known as Edman reagent, is used to label the amino-terminal residue under mildly alkaline conditions (Fig. 2.4). The resulting phenylthiohydantoin (PTH) derivative introduces an instability in the N-terminal peptide bond such that it can be hydrolyzed without cleaving the other peptide bonds. The identity of the amino acid derivative can then be determined. Edman reagent can be applied repeatedly to the shortened peptide obtained in each previous cycle. Automated sequencers are now used. D. Cleaving the polypeptide into smaller fragments | Biochemistry_Lippinco. C. Sequencing the peptide from its N-terminal end Sequencing is a stepwise process of identifying the specific amino acid at each position in the peptide chain, beginning at the N-terminal end. Phenylisothiocyanate, known as Edman reagent, is used to label the amino-terminal residue under mildly alkaline conditions (Fig. 2.4). The resulting phenylthiohydantoin (PTH) derivative introduces an instability in the N-terminal peptide bond such that it can be hydrolyzed without cleaving the other peptide bonds. The identity of the amino acid derivative can then be determined. Edman reagent can be applied repeatedly to the shortened peptide obtained in each previous cycle. Automated sequencers are now used. D. Cleaving the polypeptide into smaller fragments |
Biochemistry_Lippincott_44 | Biochemistry_Lippinco | D. Cleaving the polypeptide into smaller fragments Many polypeptides have a primary structure composed of >100 amino acids. Such molecules cannot be sequenced directly from end to end. However, these large molecules can be cleaved at specific sites and the resulting fragments sequenced. By using more than one cleaving agent (enzymes and/or chemicals) on separate samples of the purified polypeptide, overlapping fragments can be generated that permit the proper ordering of the sequenced fragments, thereby providing a complete amino acid sequence of the large polypeptide (Fig. 2.5). Enzymes that hydrolyze peptide bonds are termed peptidases (proteases). [Note: Exopeptidases cut at the ends of proteins and are divided into aminopeptidases and carboxypeptidases. Carboxypeptidases are used in determining the C-terminal amino acid. Endopeptidases cleave within a protein.] E. Determining a protein’s primary structure by DNA sequencing | Biochemistry_Lippinco. D. Cleaving the polypeptide into smaller fragments Many polypeptides have a primary structure composed of >100 amino acids. Such molecules cannot be sequenced directly from end to end. However, these large molecules can be cleaved at specific sites and the resulting fragments sequenced. By using more than one cleaving agent (enzymes and/or chemicals) on separate samples of the purified polypeptide, overlapping fragments can be generated that permit the proper ordering of the sequenced fragments, thereby providing a complete amino acid sequence of the large polypeptide (Fig. 2.5). Enzymes that hydrolyze peptide bonds are termed peptidases (proteases). [Note: Exopeptidases cut at the ends of proteins and are divided into aminopeptidases and carboxypeptidases. Carboxypeptidases are used in determining the C-terminal amino acid. Endopeptidases cleave within a protein.] E. Determining a protein’s primary structure by DNA sequencing |
Biochemistry_Lippincott_45 | Biochemistry_Lippinco | E. Determining a protein’s primary structure by DNA sequencing The sequence of nucleotides in a protein-coding region of the DNA specifies the amino acid sequence of a polypeptide. Therefore, if the nucleotide sequence can be determined, knowledge of the genetic code (see p. 447) allows the sequence of nucleotides to be translated into the corresponding amino acid sequence of that polypeptide. This indirect process, although routinely used to obtain the amino acid sequences of proteins, has the limitations of not being able to predict the positions of disulfide bonds in the folded chain and of not identifying any amino acids that are modified after their incorporation into the polypeptide (posttranslational modification; see p. 459). Therefore, direct protein sequencing is an extremely important tool for determining the true character of the primary sequence of many polypeptides. III. SECONDARY STRUCTURE | Biochemistry_Lippinco. E. Determining a protein’s primary structure by DNA sequencing The sequence of nucleotides in a protein-coding region of the DNA specifies the amino acid sequence of a polypeptide. Therefore, if the nucleotide sequence can be determined, knowledge of the genetic code (see p. 447) allows the sequence of nucleotides to be translated into the corresponding amino acid sequence of that polypeptide. This indirect process, although routinely used to obtain the amino acid sequences of proteins, has the limitations of not being able to predict the positions of disulfide bonds in the folded chain and of not identifying any amino acids that are modified after their incorporation into the polypeptide (posttranslational modification; see p. 459). Therefore, direct protein sequencing is an extremely important tool for determining the true character of the primary sequence of many polypeptides. III. SECONDARY STRUCTURE |
Biochemistry_Lippincott_46 | Biochemistry_Lippinco | III. SECONDARY STRUCTURE The polypeptide backbone does not assume a random three-dimensional structure but, instead, generally forms regular arrangements of amino acids that are located near each other in the linear sequence. These arrangements are termed the secondary structure of the polypeptide. The α-helix, β-sheet, and βbend (or, β-turn) are examples of secondary structures commonly encountered in proteins. Each is stabilized by hydrogen bonds between atoms of the peptide backbone. [Note: The collagen α-chain helix, another example of secondary structure, is discussed on p. 45.] A. α-Helix | Biochemistry_Lippinco. III. SECONDARY STRUCTURE The polypeptide backbone does not assume a random three-dimensional structure but, instead, generally forms regular arrangements of amino acids that are located near each other in the linear sequence. These arrangements are termed the secondary structure of the polypeptide. The α-helix, β-sheet, and βbend (or, β-turn) are examples of secondary structures commonly encountered in proteins. Each is stabilized by hydrogen bonds between atoms of the peptide backbone. [Note: The collagen α-chain helix, another example of secondary structure, is discussed on p. 45.] A. α-Helix |
Biochemistry_Lippincott_47 | Biochemistry_Lippinco | A. α-Helix Several different polypeptide helices are found in nature, but the α-helix is the most common. It is a rigid, right-handed spiral structure, consisting of a tightly packed, coiled polypeptide backbone core, with the side chains of the component L-amino acids extending outward from the central axis to avoid interfering sterically with each other (Fig. 2.6). A very diverse group of proteins contains α-helices. For example, the keratins are a family of closely related, rigid, fibrous proteins whose structure is nearly entirely αhelical. They are a major component of tissues such as hair and skin. In contrast to keratin, myoglobin, whose structure is also highly α-helical, is a globular, flexible molecule (see p. 26) found in muscles. 1. | Biochemistry_Lippinco. A. α-Helix Several different polypeptide helices are found in nature, but the α-helix is the most common. It is a rigid, right-handed spiral structure, consisting of a tightly packed, coiled polypeptide backbone core, with the side chains of the component L-amino acids extending outward from the central axis to avoid interfering sterically with each other (Fig. 2.6). A very diverse group of proteins contains α-helices. For example, the keratins are a family of closely related, rigid, fibrous proteins whose structure is nearly entirely αhelical. They are a major component of tissues such as hair and skin. In contrast to keratin, myoglobin, whose structure is also highly α-helical, is a globular, flexible molecule (see p. 26) found in muscles. 1. |
Biochemistry_Lippincott_48 | Biochemistry_Lippinco | 1. Hydrogen bonds: An α-helix is stabilized by extensive hydrogen bonding between the peptide bond carbonyl oxygens and amide hydrogens that are part of the polypeptide backbone (see Fig. 2.6). The hydrogen bonds extend up and are parallel to the spiral from the carbonyl oxygen of one peptide bond to the –NH group of a peptide linkage four residues ahead in the polypeptide. This insures that all but the first and last peptide bond components are linked to each other through intrachain hydrogen bonds. Hydrogen bonds are individually weak, but they collectively serve to stabilize the helix. 2. Amino acids per turn: Each turn of an α-helix contains 3.6 amino acids. Thus, amino acids spaced three or four residues apart in the primary sequence are spatially close together when folded in the α-helix. 3. | Biochemistry_Lippinco. 1. Hydrogen bonds: An α-helix is stabilized by extensive hydrogen bonding between the peptide bond carbonyl oxygens and amide hydrogens that are part of the polypeptide backbone (see Fig. 2.6). The hydrogen bonds extend up and are parallel to the spiral from the carbonyl oxygen of one peptide bond to the –NH group of a peptide linkage four residues ahead in the polypeptide. This insures that all but the first and last peptide bond components are linked to each other through intrachain hydrogen bonds. Hydrogen bonds are individually weak, but they collectively serve to stabilize the helix. 2. Amino acids per turn: Each turn of an α-helix contains 3.6 amino acids. Thus, amino acids spaced three or four residues apart in the primary sequence are spatially close together when folded in the α-helix. 3. |
Biochemistry_Lippincott_49 | Biochemistry_Lippinco | 3. Amino acids that disrupt an α-helix: The R group of an amino acid determines its propensity to be in an α-helix. Proline disrupts an α-helix because its rigid secondary amino group is not geometrically compatible with the right-handed spiral of the α-helix. Instead, it inserts a kink in the chain, which interferes with the smooth, helical structure. Glycine is also a “helix breaker” because its R group (a hydrogen) confers high flexibility. Additionally, amino acids with charged or bulky R groups (such as glutamate and tryptophan, respectively) and those with a branch at the β-carbon, the first carbon in the R group (for example, valine), have low α-helix propensity. B. β-Sheet | Biochemistry_Lippinco. 3. Amino acids that disrupt an α-helix: The R group of an amino acid determines its propensity to be in an α-helix. Proline disrupts an α-helix because its rigid secondary amino group is not geometrically compatible with the right-handed spiral of the α-helix. Instead, it inserts a kink in the chain, which interferes with the smooth, helical structure. Glycine is also a “helix breaker” because its R group (a hydrogen) confers high flexibility. Additionally, amino acids with charged or bulky R groups (such as glutamate and tryptophan, respectively) and those with a branch at the β-carbon, the first carbon in the R group (for example, valine), have low α-helix propensity. B. β-Sheet |
Biochemistry_Lippincott_50 | Biochemistry_Lippinco | B. β-Sheet The β-sheet is another form of secondary structure in which all of the peptide bond components are involved in hydrogen bonding (Fig. 2.7A). Because the surfaces of β-sheets appear “pleated,” they are often called βpleated sheets. [Note: Pleating results from successive α-carbons being slightly above or below the plane of the sheet.] Illustrations of protein structure often show β-strands as broad arrows (Fig. 2.7B). 1. | Biochemistry_Lippinco. B. β-Sheet The β-sheet is another form of secondary structure in which all of the peptide bond components are involved in hydrogen bonding (Fig. 2.7A). Because the surfaces of β-sheets appear “pleated,” they are often called βpleated sheets. [Note: Pleating results from successive α-carbons being slightly above or below the plane of the sheet.] Illustrations of protein structure often show β-strands as broad arrows (Fig. 2.7B). 1. |
Biochemistry_Lippincott_51 | Biochemistry_Lippinco | 1. Formation: A β-sheet is formed by two or more peptide chains (βstrands) aligned laterally and stabilized by hydrogen bonds between the carboxyl and amino groups of amino acids that either are far apart in a single polypeptide (intrachain bonds) or are in different polypeptide chains (interchain bonds). The adjacent β-strands are arranged either antiparallel to each other (with the N-termini alternating as shown in Fig. 2.7B) or parallel to each other (with the N-termini together as shown in Fig. 2.7C). On each β-strand, the R groups of adjacent amino acids extend in opposite directions, above and below the plane of the β-sheet. [Note: β-sheets are not flat and have a right-handed curl (twist) when viewed along the polypeptide backbone.] 2. | Biochemistry_Lippinco. 1. Formation: A β-sheet is formed by two or more peptide chains (βstrands) aligned laterally and stabilized by hydrogen bonds between the carboxyl and amino groups of amino acids that either are far apart in a single polypeptide (intrachain bonds) or are in different polypeptide chains (interchain bonds). The adjacent β-strands are arranged either antiparallel to each other (with the N-termini alternating as shown in Fig. 2.7B) or parallel to each other (with the N-termini together as shown in Fig. 2.7C). On each β-strand, the R groups of adjacent amino acids extend in opposite directions, above and below the plane of the β-sheet. [Note: β-sheets are not flat and have a right-handed curl (twist) when viewed along the polypeptide backbone.] 2. |
Biochemistry_Lippincott_52 | Biochemistry_Lippinco | Comparing α-helices and β-sheets: In β-sheets, the β-strands are almost fully extended and the hydrogen bonds between the strands are perpendicular to the polypeptide backbone (see Fig. 2.7A). In contrast, in α-helices, the polypeptide is coiled and the hydrogen bonds are parallel to the backbone (see Fig. 2.6). The orientation of the R groups of the amino acid residues in both the αhelix and the β-sheet can result in formation of polar and nonpolar sides in these secondary structures, thereby making them amphipathic. | Biochemistry_Lippinco. Comparing α-helices and β-sheets: In β-sheets, the β-strands are almost fully extended and the hydrogen bonds between the strands are perpendicular to the polypeptide backbone (see Fig. 2.7A). In contrast, in α-helices, the polypeptide is coiled and the hydrogen bonds are parallel to the backbone (see Fig. 2.6). The orientation of the R groups of the amino acid residues in both the αhelix and the β-sheet can result in formation of polar and nonpolar sides in these secondary structures, thereby making them amphipathic. |
Biochemistry_Lippincott_53 | Biochemistry_Lippinco | C. β-Bends (reverse turns, β-turns) β-Bends reverse the direction of a polypeptide chain, helping it form a compact, globular shape. They are usually found on the surface of protein molecules and often include charged residues. [Note: β-Bends were given this name because they often connect successive strands of antiparallel βsheets.] β-Bends are generally composed of four amino acids, one of which may be proline, the amino acid that causes a kink in the polypeptide chain. Glycine, the amino acid with the smallest R group, is also frequently found in β-bends. β-Bends are stabilized by the formation of hydrogen bonds between the first and last residues in the bend. D. Nonrepetitive secondary structure | Biochemistry_Lippinco. C. β-Bends (reverse turns, β-turns) β-Bends reverse the direction of a polypeptide chain, helping it form a compact, globular shape. They are usually found on the surface of protein molecules and often include charged residues. [Note: β-Bends were given this name because they often connect successive strands of antiparallel βsheets.] β-Bends are generally composed of four amino acids, one of which may be proline, the amino acid that causes a kink in the polypeptide chain. Glycine, the amino acid with the smallest R group, is also frequently found in β-bends. β-Bends are stabilized by the formation of hydrogen bonds between the first and last residues in the bend. D. Nonrepetitive secondary structure |
Biochemistry_Lippincott_54 | Biochemistry_Lippinco | D. Nonrepetitive secondary structure Approximately one half of an average globular protein is organized into repetitive structures, such as the α-helix and β-sheet. The remainder of the polypeptide chain is described as having a loop or coil conformation. These nonrepetitive secondary structures are not random but rather simply have a less regular structure than those described above. [Note: The term “random coil” refers to the disordered structure obtained when proteins are denatured (see p. 20).] E. Supersecondary structures (motifs) | Biochemistry_Lippinco. D. Nonrepetitive secondary structure Approximately one half of an average globular protein is organized into repetitive structures, such as the α-helix and β-sheet. The remainder of the polypeptide chain is described as having a loop or coil conformation. These nonrepetitive secondary structures are not random but rather simply have a less regular structure than those described above. [Note: The term “random coil” refers to the disordered structure obtained when proteins are denatured (see p. 20).] E. Supersecondary structures (motifs) |
Biochemistry_Lippincott_55 | Biochemistry_Lippinco | E. Supersecondary structures (motifs) Globular proteins are constructed by combining secondary structural elements (that is, α-helices, β-sheets, and coils), producing specific geometric patterns, or motifs. These form primarily the core (interior) region of the molecule. They are connected by loop regions (for example, β-bends) at the surface of the protein. Supersecondary structures are usually produced by the close packing of side chains from adjacent secondary structural elements. For example, α-helices and β-sheets that are adjacent in the amino acid sequence are also usually (but not always) adjacent in the final, folded protein. Some of the more common motifs are illustrated in Figure 2.8. Motifs may be associated with particular functions. Proteins that bind to DNA contain a limited number of motifs. The helix–loop–helix motif is an example found in a number of proteins that function as transcription factors (see p. 438). IV. TERTIARY STRUCTURE | Biochemistry_Lippinco. E. Supersecondary structures (motifs) Globular proteins are constructed by combining secondary structural elements (that is, α-helices, β-sheets, and coils), producing specific geometric patterns, or motifs. These form primarily the core (interior) region of the molecule. They are connected by loop regions (for example, β-bends) at the surface of the protein. Supersecondary structures are usually produced by the close packing of side chains from adjacent secondary structural elements. For example, α-helices and β-sheets that are adjacent in the amino acid sequence are also usually (but not always) adjacent in the final, folded protein. Some of the more common motifs are illustrated in Figure 2.8. Motifs may be associated with particular functions. Proteins that bind to DNA contain a limited number of motifs. The helix–loop–helix motif is an example found in a number of proteins that function as transcription factors (see p. 438). IV. TERTIARY STRUCTURE |
Biochemistry_Lippincott_56 | Biochemistry_Lippinco | IV. TERTIARY STRUCTURE The primary structure of a polypeptide chain determines its tertiary structure. “Tertiary” refers both to the folding of domains (the basic units of structure and function; see A. below) and to the final arrangement of domains in the polypeptide. The tertiary structure of globular proteins in aqueous solution is compact, with a high density (close packing) of the atoms in the core of the molecule. Hydrophobic side chains are buried in the interior, whereas hydrophilic groups are generally found on the surface of the molecule. A. Domains | Biochemistry_Lippinco. IV. TERTIARY STRUCTURE The primary structure of a polypeptide chain determines its tertiary structure. “Tertiary” refers both to the folding of domains (the basic units of structure and function; see A. below) and to the final arrangement of domains in the polypeptide. The tertiary structure of globular proteins in aqueous solution is compact, with a high density (close packing) of the atoms in the core of the molecule. Hydrophobic side chains are buried in the interior, whereas hydrophilic groups are generally found on the surface of the molecule. A. Domains |
Biochemistry_Lippincott_57 | Biochemistry_Lippinco | A. Domains Domains are the fundamental functional and three-dimensional structural units of polypeptides. Polypeptide chains that are >200 amino acids in length generally consist of two or more domains. The core of a domain is built from combinations of supersecondary structural elements (motifs). Folding of the peptide chain within a domain usually occurs independently of folding in other domains. Therefore, each domain has the characteristics of a small, compact globular protein that is structurally independent of the other domains in the polypeptide chain. B. Stabilizing interactions The unique three-dimensional structure of each polypeptide is determined by its amino acid sequence. Interactions between the amino acid side chains guide the folding of the polypeptide to form a compact structure. The following four types of interactions cooperate in stabilizing the tertiary structures of globular proteins. | Biochemistry_Lippinco. A. Domains Domains are the fundamental functional and three-dimensional structural units of polypeptides. Polypeptide chains that are >200 amino acids in length generally consist of two or more domains. The core of a domain is built from combinations of supersecondary structural elements (motifs). Folding of the peptide chain within a domain usually occurs independently of folding in other domains. Therefore, each domain has the characteristics of a small, compact globular protein that is structurally independent of the other domains in the polypeptide chain. B. Stabilizing interactions The unique three-dimensional structure of each polypeptide is determined by its amino acid sequence. Interactions between the amino acid side chains guide the folding of the polypeptide to form a compact structure. The following four types of interactions cooperate in stabilizing the tertiary structures of globular proteins. |
Biochemistry_Lippincott_58 | Biochemistry_Lippinco | 1. Disulfide bonds: A disulfide bond (–S–S–) is a covalent linkage formed from the sulfhydryl group (−SH) of each of two cysteine residues to produce a cystine residue (Fig. 2.9). The two cysteines may be separated from each other by many amino acids in the primary sequence of a polypeptide or may even be located on two different polypeptides. The folding of the polypeptide(s) brings the cysteine residues into proximity and permits covalent bonding of their side chains. A disulfide bond contributes to the stability of the three-dimensional shape of the protein molecule and prevents it from becoming denatured in the extracellular environment. For example, many disulfide bonds are found in proteins such as immunoglobulins that are secreted by cells. [Note: Protein disulfide isomerase breaks and reforms disulfide bonds during folding.] 2. | Biochemistry_Lippinco. 1. Disulfide bonds: A disulfide bond (–S–S–) is a covalent linkage formed from the sulfhydryl group (−SH) of each of two cysteine residues to produce a cystine residue (Fig. 2.9). The two cysteines may be separated from each other by many amino acids in the primary sequence of a polypeptide or may even be located on two different polypeptides. The folding of the polypeptide(s) brings the cysteine residues into proximity and permits covalent bonding of their side chains. A disulfide bond contributes to the stability of the three-dimensional shape of the protein molecule and prevents it from becoming denatured in the extracellular environment. For example, many disulfide bonds are found in proteins such as immunoglobulins that are secreted by cells. [Note: Protein disulfide isomerase breaks and reforms disulfide bonds during folding.] 2. |
Biochemistry_Lippincott_59 | Biochemistry_Lippinco | Hydrophobic interactions: Amino acids with nonpolar side chains tend to be located in the interior of the polypeptide molecule, where they associate with other hydrophobic amino acids (Fig. 2.10). In contrast, amino acids with polar or charged side chains tend to be located on the surface of the molecule in contact with the polar solvent. [Note: Recall that proteins located in nonpolar (lipid) environments, such as a membrane, exhibit the reverse arrangement (see Fig. 1.4, p. 4).] In each case, a segregation of R groups occurs that is energetically most favorable. 3. | Biochemistry_Lippinco. Hydrophobic interactions: Amino acids with nonpolar side chains tend to be located in the interior of the polypeptide molecule, where they associate with other hydrophobic amino acids (Fig. 2.10). In contrast, amino acids with polar or charged side chains tend to be located on the surface of the molecule in contact with the polar solvent. [Note: Recall that proteins located in nonpolar (lipid) environments, such as a membrane, exhibit the reverse arrangement (see Fig. 1.4, p. 4).] In each case, a segregation of R groups occurs that is energetically most favorable. 3. |
Biochemistry_Lippincott_60 | Biochemistry_Lippinco | 3. Hydrogen bonds: Amino acid side chains containing oxygen-or nitrogen-bound hydrogen, such as in the alcohol groups of serine and threonine, can form hydrogen bonds with electron-rich atoms, such as the oxygen of a carboxyl group or carbonyl group of a peptide bond (Fig. 2.11; see also Fig. 1.6, p. 4). Formation of hydrogen bonds between polar groups on the surface of proteins and the aqueous solvent enhances the solubility of the protein. 4. Ionic interactions: Negatively charged groups, such as the carboxylate group (−COO−) in the side chain of aspartate or glutamate, can interact with positively charged groups such as the amino group (−NH3+) in the side chain of lysine (see Fig. 2.11). C. Protein folding | Biochemistry_Lippinco. 3. Hydrogen bonds: Amino acid side chains containing oxygen-or nitrogen-bound hydrogen, such as in the alcohol groups of serine and threonine, can form hydrogen bonds with electron-rich atoms, such as the oxygen of a carboxyl group or carbonyl group of a peptide bond (Fig. 2.11; see also Fig. 1.6, p. 4). Formation of hydrogen bonds between polar groups on the surface of proteins and the aqueous solvent enhances the solubility of the protein. 4. Ionic interactions: Negatively charged groups, such as the carboxylate group (−COO−) in the side chain of aspartate or glutamate, can interact with positively charged groups such as the amino group (−NH3+) in the side chain of lysine (see Fig. 2.11). C. Protein folding |
Biochemistry_Lippincott_61 | Biochemistry_Lippinco | C. Protein folding Interactions between the side chains of amino acids determine how a linear polypeptide chain folds into the intricate three-dimensional shape of the functional protein. Protein folding, which occurs within the cell in seconds to minutes, involves nonrandom, ordered pathways. As a peptide folds, secondary structures form, driven by the hydrophobic effect (that is, hydrophobic groups come together as water is released). These small structures combine to form larger structures. Additional events stabilize secondary structure and initiate formation of tertiary structure. In the last stage, the peptide achieves its fully folded, native (functional) form characterized by a low-energy state (Fig. 2.12). [Note: Some biologically active proteins or segments thereof lack a stable tertiary structure. They are referred to as intrinsically disordered proteins.] D. Protein denaturation | Biochemistry_Lippinco. C. Protein folding Interactions between the side chains of amino acids determine how a linear polypeptide chain folds into the intricate three-dimensional shape of the functional protein. Protein folding, which occurs within the cell in seconds to minutes, involves nonrandom, ordered pathways. As a peptide folds, secondary structures form, driven by the hydrophobic effect (that is, hydrophobic groups come together as water is released). These small structures combine to form larger structures. Additional events stabilize secondary structure and initiate formation of tertiary structure. In the last stage, the peptide achieves its fully folded, native (functional) form characterized by a low-energy state (Fig. 2.12). [Note: Some biologically active proteins or segments thereof lack a stable tertiary structure. They are referred to as intrinsically disordered proteins.] D. Protein denaturation |
Biochemistry_Lippincott_62 | Biochemistry_Lippinco | D. Protein denaturation Denaturation results in the unfolding and disorganization of a protein’s secondary and tertiary structures without the hydrolysis of peptide bonds. Denaturing agents include heat, urea, organic solvents, strong acids or bases, detergents, and ions of heavy metals such as lead. Denaturation may, under ideal conditions, be reversible, such that the protein refolds into its original native structure when the denaturing agent is removed. However, most proteins remain permanently disordered once denatured. Denatured proteins are often insoluble and precipitate from solution. E. Chaperones in protein folding | Biochemistry_Lippinco. D. Protein denaturation Denaturation results in the unfolding and disorganization of a protein’s secondary and tertiary structures without the hydrolysis of peptide bonds. Denaturing agents include heat, urea, organic solvents, strong acids or bases, detergents, and ions of heavy metals such as lead. Denaturation may, under ideal conditions, be reversible, such that the protein refolds into its original native structure when the denaturing agent is removed. However, most proteins remain permanently disordered once denatured. Denatured proteins are often insoluble and precipitate from solution. E. Chaperones in protein folding |
Biochemistry_Lippincott_63 | Biochemistry_Lippinco | The information needed for correct protein folding is contained in the primary structure of the polypeptide. However, most denatured proteins do not resume their native conformations even under favorable environmental conditions. This is because, for many proteins, folding is a facilitated process that requires a specialized group of proteins, referred to as molecular chaperones, and ATP hydrolysis. The chaperones, also known as heat shock proteins (HSP), interact with a polypeptide at various stages during the folding process. Some chaperones bind hydrophobic regions of an extended polypeptide and are important in keeping the protein unfolded until its synthesis is completed (for example, Hsp70). Others form cage-like macromolecular structures composed of two stacked rings. The partially folded protein enters the cage, binds the central cavity through hydrophobic interactions, folds, and is released (for example, mitochondrial Hsp60). [Note: Cage-like chaperones are sometimes | Biochemistry_Lippinco. The information needed for correct protein folding is contained in the primary structure of the polypeptide. However, most denatured proteins do not resume their native conformations even under favorable environmental conditions. This is because, for many proteins, folding is a facilitated process that requires a specialized group of proteins, referred to as molecular chaperones, and ATP hydrolysis. The chaperones, also known as heat shock proteins (HSP), interact with a polypeptide at various stages during the folding process. Some chaperones bind hydrophobic regions of an extended polypeptide and are important in keeping the protein unfolded until its synthesis is completed (for example, Hsp70). Others form cage-like macromolecular structures composed of two stacked rings. The partially folded protein enters the cage, binds the central cavity through hydrophobic interactions, folds, and is released (for example, mitochondrial Hsp60). [Note: Cage-like chaperones are sometimes |
Biochemistry_Lippincott_64 | Biochemistry_Lippinco | folded protein enters the cage, binds the central cavity through hydrophobic interactions, folds, and is released (for example, mitochondrial Hsp60). [Note: Cage-like chaperones are sometimes referred to as chaperonins.] Chaperones, then, facilitate correct protein folding by binding to and stabilizing exposed, aggregation-prone hydrophobic regions in nascent (and denatured) polypeptides, preventing premature folding. | Biochemistry_Lippinco. folded protein enters the cage, binds the central cavity through hydrophobic interactions, folds, and is released (for example, mitochondrial Hsp60). [Note: Cage-like chaperones are sometimes referred to as chaperonins.] Chaperones, then, facilitate correct protein folding by binding to and stabilizing exposed, aggregation-prone hydrophobic regions in nascent (and denatured) polypeptides, preventing premature folding. |
Biochemistry_Lippincott_65 | Biochemistry_Lippinco | V. QUATERNARY STRUCTURE Many proteins consist of a single polypeptide chain and are defined as monomeric proteins. However, others may consist of two or more polypeptide chains that may be structurally identical or totally unrelated. The arrangement of these polypeptide subunits is called the quaternary structure of the protein. Subunits are held together primarily by noncovalent interactions (for example, hydrogen bonds, ionic bonds, and hydrophobic interactions). Subunits either may function independently of each other or may work cooperatively, as in hemoglobin, in which the binding of oxygen to one subunit of the tetramer increases the affinity of the other subunits for oxygen (see p. 29). Isoforms are proteins that perform the same function but have different primary structures. They can arise from different genes or from tissue-specific processing of the product of a single gene. If the proteins function as enzymes, they are referred to as isozymes (see p. 65). | Biochemistry_Lippinco. V. QUATERNARY STRUCTURE Many proteins consist of a single polypeptide chain and are defined as monomeric proteins. However, others may consist of two or more polypeptide chains that may be structurally identical or totally unrelated. The arrangement of these polypeptide subunits is called the quaternary structure of the protein. Subunits are held together primarily by noncovalent interactions (for example, hydrogen bonds, ionic bonds, and hydrophobic interactions). Subunits either may function independently of each other or may work cooperatively, as in hemoglobin, in which the binding of oxygen to one subunit of the tetramer increases the affinity of the other subunits for oxygen (see p. 29). Isoforms are proteins that perform the same function but have different primary structures. They can arise from different genes or from tissue-specific processing of the product of a single gene. If the proteins function as enzymes, they are referred to as isozymes (see p. 65). |
Biochemistry_Lippincott_66 | Biochemistry_Lippinco | VI. PROTEIN MISFOLDING Protein folding is a complex process that can sometimes result in improperly folded molecules. These misfolded proteins are usually tagged and degraded within the cell (see p. 247). However, this quality control system is not perfect, and intracellular or extracellular aggregates of misfolded proteins can accumulate, particularly as individuals age. Deposits of misfolded proteins are associated with a number of diseases. A. Amyloid diseases | Biochemistry_Lippinco. VI. PROTEIN MISFOLDING Protein folding is a complex process that can sometimes result in improperly folded molecules. These misfolded proteins are usually tagged and degraded within the cell (see p. 247). However, this quality control system is not perfect, and intracellular or extracellular aggregates of misfolded proteins can accumulate, particularly as individuals age. Deposits of misfolded proteins are associated with a number of diseases. A. Amyloid diseases |
Biochemistry_Lippincott_67 | Biochemistry_Lippinco | Misfolding of proteins may occur spontaneously or be caused by a mutation in a particular gene, which then produces an altered protein. In addition, some apparently normal proteins can, after abnormal proteolytic cleavage, take on a unique conformation that leads to the spontaneous formation of long, fibrillar protein assemblies consisting of β-pleated sheets. Accumulation of these insoluble fibrous protein aggregates, called amyloids, has been implicated in neurodegenerative disorders such as Parkinson disease and Alzheimer disease (AD). The dominant component of the amyloid plaque that accumulates in AD is amyloid β (Aβ), an extracellular peptide containing 40–42 amino acid residues. X-ray crystallography and infrared spectroscopy demonstrate a characteristic βpleated sheet secondary structure in nonbranching fibrils. This peptide, when aggregated in a β-pleated sheet conformation, is neurotoxic and is the central pathogenic event leading to the cognitive impairment characteristic | Biochemistry_Lippinco. Misfolding of proteins may occur spontaneously or be caused by a mutation in a particular gene, which then produces an altered protein. In addition, some apparently normal proteins can, after abnormal proteolytic cleavage, take on a unique conformation that leads to the spontaneous formation of long, fibrillar protein assemblies consisting of β-pleated sheets. Accumulation of these insoluble fibrous protein aggregates, called amyloids, has been implicated in neurodegenerative disorders such as Parkinson disease and Alzheimer disease (AD). The dominant component of the amyloid plaque that accumulates in AD is amyloid β (Aβ), an extracellular peptide containing 40–42 amino acid residues. X-ray crystallography and infrared spectroscopy demonstrate a characteristic βpleated sheet secondary structure in nonbranching fibrils. This peptide, when aggregated in a β-pleated sheet conformation, is neurotoxic and is the central pathogenic event leading to the cognitive impairment characteristic |
Biochemistry_Lippincott_68 | Biochemistry_Lippinco | in nonbranching fibrils. This peptide, when aggregated in a β-pleated sheet conformation, is neurotoxic and is the central pathogenic event leading to the cognitive impairment characteristic of the disease. The Aβ that is deposited in the brain in AD is derived by enzymic cleavages (by secretases) from the larger amyloid precursor protein, a single transmembrane protein expressed on the cell surface in the brain and other tissues (Fig. 2.13). The Aβ peptides aggregate, generating the amyloid that is found in the brain parenchyma and around blood vessels. Most cases of AD are not genetically based, although at least 5% of cases are familial. A second biologic factor involved in the development of AD is the accumulation of neurofibrillary tangles inside neurons. A key component of these tangled fibers is an abnormal form (hyperphosphorylated and insoluble) of the tau (τ) protein, which, in its healthy version, helps in the assembly of the microtubular structure. The defective τ appears | Biochemistry_Lippinco. in nonbranching fibrils. This peptide, when aggregated in a β-pleated sheet conformation, is neurotoxic and is the central pathogenic event leading to the cognitive impairment characteristic of the disease. The Aβ that is deposited in the brain in AD is derived by enzymic cleavages (by secretases) from the larger amyloid precursor protein, a single transmembrane protein expressed on the cell surface in the brain and other tissues (Fig. 2.13). The Aβ peptides aggregate, generating the amyloid that is found in the brain parenchyma and around blood vessels. Most cases of AD are not genetically based, although at least 5% of cases are familial. A second biologic factor involved in the development of AD is the accumulation of neurofibrillary tangles inside neurons. A key component of these tangled fibers is an abnormal form (hyperphosphorylated and insoluble) of the tau (τ) protein, which, in its healthy version, helps in the assembly of the microtubular structure. The defective τ appears |
Biochemistry_Lippincott_69 | Biochemistry_Lippinco | fibers is an abnormal form (hyperphosphorylated and insoluble) of the tau (τ) protein, which, in its healthy version, helps in the assembly of the microtubular structure. The defective τ appears to block the actions of its normal counterpart. [Note: In Parkinson disease, amyloid is formed from α-synuclein protein.] | Biochemistry_Lippinco. fibers is an abnormal form (hyperphosphorylated and insoluble) of the tau (τ) protein, which, in its healthy version, helps in the assembly of the microtubular structure. The defective τ appears to block the actions of its normal counterpart. [Note: In Parkinson disease, amyloid is formed from α-synuclein protein.] |
Biochemistry_Lippincott_70 | Biochemistry_Lippinco | B. Prion (proteinaceous infectious particle) diseases | Biochemistry_Lippinco. B. Prion (proteinaceous infectious particle) diseases |
Biochemistry_Lippincott_71 | Biochemistry_Lippinco | The prion protein (PrP) is the causative agent of transmissible spongiform encephalopathies (TSE), including Creutzfeldt-Jakob disease in humans, scrapie in sheep, and bovine spongiform encephalopathy in cattle (popularly called “mad cow” disease). After an extensive series of purification procedures, scientists were surprised to find that the infectivity of the agent causing scrapie in sheep was associated with a single protein species that was not complexed with detectable nucleic acid. This infectious protein is designated PrPSc (Sc = scrapie). It is highly resistant to proteolytic degradation and tends to form insoluble aggregates of fibrils, similar to the amyloid found in some other diseases of the brain. A noninfectious form of PrPC (C = cellular), encoded by the same gene as the infectious agent, is present in normal mammalian brains on the surface of neurons and glial cells. Thus, PrPC is a host protein. No primary structure differences or alternate posttranslational | Biochemistry_Lippinco. The prion protein (PrP) is the causative agent of transmissible spongiform encephalopathies (TSE), including Creutzfeldt-Jakob disease in humans, scrapie in sheep, and bovine spongiform encephalopathy in cattle (popularly called “mad cow” disease). After an extensive series of purification procedures, scientists were surprised to find that the infectivity of the agent causing scrapie in sheep was associated with a single protein species that was not complexed with detectable nucleic acid. This infectious protein is designated PrPSc (Sc = scrapie). It is highly resistant to proteolytic degradation and tends to form insoluble aggregates of fibrils, similar to the amyloid found in some other diseases of the brain. A noninfectious form of PrPC (C = cellular), encoded by the same gene as the infectious agent, is present in normal mammalian brains on the surface of neurons and glial cells. Thus, PrPC is a host protein. No primary structure differences or alternate posttranslational |
Biochemistry_Lippincott_72 | Biochemistry_Lippinco | as the infectious agent, is present in normal mammalian brains on the surface of neurons and glial cells. Thus, PrPC is a host protein. No primary structure differences or alternate posttranslational modifications have been found between the normal and the infectious forms of the protein. The key to becoming infectious apparently lies in changes in the three-dimensional conformation of PrPC. Research has demonstrated that a number of αhelices present in noninfectious PrPC are replaced by β-sheets in the infectious form (Fig. 2.14). This conformational difference is presumably what confers relative resistance to proteolytic degradation of infectious prions and permits them to be distinguished from the normal PrPC in infected tissue. The infective agent is, thus, an altered version of a normal protein, which acts as a template for converting the normal protein to the pathogenic conformation. The TSE are invariably fatal, and no treatment is currently available that can alter this | Biochemistry_Lippinco. as the infectious agent, is present in normal mammalian brains on the surface of neurons and glial cells. Thus, PrPC is a host protein. No primary structure differences or alternate posttranslational modifications have been found between the normal and the infectious forms of the protein. The key to becoming infectious apparently lies in changes in the three-dimensional conformation of PrPC. Research has demonstrated that a number of αhelices present in noninfectious PrPC are replaced by β-sheets in the infectious form (Fig. 2.14). This conformational difference is presumably what confers relative resistance to proteolytic degradation of infectious prions and permits them to be distinguished from the normal PrPC in infected tissue. The infective agent is, thus, an altered version of a normal protein, which acts as a template for converting the normal protein to the pathogenic conformation. The TSE are invariably fatal, and no treatment is currently available that can alter this |
Biochemistry_Lippincott_73 | Biochemistry_Lippinco | a normal protein, which acts as a template for converting the normal protein to the pathogenic conformation. The TSE are invariably fatal, and no treatment is currently available that can alter this outcome. | Biochemistry_Lippinco. a normal protein, which acts as a template for converting the normal protein to the pathogenic conformation. The TSE are invariably fatal, and no treatment is currently available that can alter this outcome. |
Biochemistry_Lippincott_74 | Biochemistry_Lippinco | VII. CHAPTER SUMMARY | Biochemistry_Lippinco. VII. CHAPTER SUMMARY |
Biochemistry_Lippincott_75 | Biochemistry_Lippinco | Central to understanding protein structure is the concept of the native conformation (Fig. 2.15), which is the functional, fully folded protein structure (for example, an active enzyme or structural protein). The unique three-dimensional structure of the native conformation is determined by its primary structure, that is, its amino acid sequence. Interactions between the amino acid side chains guide the folding of the polypeptide chain to form secondary, tertiary, and (sometimes) quaternary structures, which cooperate in stabilizing the native conformation of the protein. In addition, a specialized group of proteins named chaperones is required for the proper folding of many species of proteins. Protein denaturation results in the unfolding and disorganization of the protein’s structure, which are not accompanied by hydrolysis of peptide bonds. Denaturation may be reversible or, more commonly, irreversible. Disease can occur when an apparently normal protein assumes a conformation | Biochemistry_Lippinco. Central to understanding protein structure is the concept of the native conformation (Fig. 2.15), which is the functional, fully folded protein structure (for example, an active enzyme or structural protein). The unique three-dimensional structure of the native conformation is determined by its primary structure, that is, its amino acid sequence. Interactions between the amino acid side chains guide the folding of the polypeptide chain to form secondary, tertiary, and (sometimes) quaternary structures, which cooperate in stabilizing the native conformation of the protein. In addition, a specialized group of proteins named chaperones is required for the proper folding of many species of proteins. Protein denaturation results in the unfolding and disorganization of the protein’s structure, which are not accompanied by hydrolysis of peptide bonds. Denaturation may be reversible or, more commonly, irreversible. Disease can occur when an apparently normal protein assumes a conformation |
Biochemistry_Lippincott_76 | Biochemistry_Lippinco | which are not accompanied by hydrolysis of peptide bonds. Denaturation may be reversible or, more commonly, irreversible. Disease can occur when an apparently normal protein assumes a conformation that is cytotoxic, as in the case of Alzheimer disease (AD) and the transmissible spongiform encephalopathies (TSE), including Creutzfeldt-Jakob disease. In AD, normal proteins, after abnormal chemical processing, take on a unique conformational state that leads to the formation of neurotoxic amyloid β peptide (Aβ) assemblies consisting of β-pleated sheets. In TSE, the infective agent is an altered version of a normal prion protein that acts as a template for converting normal protein to the pathogenic conformation. | Biochemistry_Lippinco. which are not accompanied by hydrolysis of peptide bonds. Denaturation may be reversible or, more commonly, irreversible. Disease can occur when an apparently normal protein assumes a conformation that is cytotoxic, as in the case of Alzheimer disease (AD) and the transmissible spongiform encephalopathies (TSE), including Creutzfeldt-Jakob disease. In AD, normal proteins, after abnormal chemical processing, take on a unique conformational state that leads to the formation of neurotoxic amyloid β peptide (Aβ) assemblies consisting of β-pleated sheets. In TSE, the infective agent is an altered version of a normal prion protein that acts as a template for converting normal protein to the pathogenic conformation. |
Biochemistry_Lippincott_77 | Biochemistry_Lippinco | Choose the ONE best answer. .1. Which one of the following statements concerning protein structure is correct? A. Proteins consisting of one polypeptide have quaternary structure that is stabilized by covalent bonds. B. The peptide bonds that link amino acids in a protein most commonly occur in the cis configuration. C. The formation of a disulfide bond in a protein requires the participating cysteine residues to be adjacent in the primary structure. D. The denaturation of proteins leads to irreversible loss of secondary structural elements such as the α-helix. E. The primary driving force for protein folding is the hydrophobic effect. | Biochemistry_Lippinco. Choose the ONE best answer. .1. Which one of the following statements concerning protein structure is correct? A. Proteins consisting of one polypeptide have quaternary structure that is stabilized by covalent bonds. B. The peptide bonds that link amino acids in a protein most commonly occur in the cis configuration. C. The formation of a disulfide bond in a protein requires the participating cysteine residues to be adjacent in the primary structure. D. The denaturation of proteins leads to irreversible loss of secondary structural elements such as the α-helix. E. The primary driving force for protein folding is the hydrophobic effect. |
Biochemistry_Lippincott_78 | Biochemistry_Lippinco | D. The denaturation of proteins leads to irreversible loss of secondary structural elements such as the α-helix. E. The primary driving force for protein folding is the hydrophobic effect. Correct answer = E. The hydrophobic effect, or the tendency of nonpolar entities to associate in a polar environment, is the primary driving force of protein folding. Quaternary structure requires more than one polypeptide, and, when present, it is stabilized primarily by noncovalent bonds. The peptide bond is almost always trans. The two cysteine residues participating in disulfide bond formation may be a great distance apart in the amino acid sequence of a polypeptide (or on two separate polypeptides) but are brought into close proximity by the three-dimensional folding of the polypeptide. Denaturation may be reversible or irreversible. | Biochemistry_Lippinco. D. The denaturation of proteins leads to irreversible loss of secondary structural elements such as the α-helix. E. The primary driving force for protein folding is the hydrophobic effect. Correct answer = E. The hydrophobic effect, or the tendency of nonpolar entities to associate in a polar environment, is the primary driving force of protein folding. Quaternary structure requires more than one polypeptide, and, when present, it is stabilized primarily by noncovalent bonds. The peptide bond is almost always trans. The two cysteine residues participating in disulfide bond formation may be a great distance apart in the amino acid sequence of a polypeptide (or on two separate polypeptides) but are brought into close proximity by the three-dimensional folding of the polypeptide. Denaturation may be reversible or irreversible. |
Biochemistry_Lippincott_79 | Biochemistry_Lippinco | .2. A particular point mutation results in disruption of the α-helical structure in a segment of the mutant protein. The most likely change in the primary structure of the mutant protein is: A. glutamate to aspartate. B. lysine to arginine. C. methionine to proline. D. valine to alanine. Correct answer = C. Proline, because of its secondary amino group, is incompatible with an α-helix. Glutamate, aspartate, lysine, and arginine are charged amino acids, and valine is a branched amino acid. Charged and branched (bulky) amino acids may disrupt an α-helix. [Note: The flexibility of glycine’s R group can also disrupt an α-helix.] .3. In comparing the α-helix to the β-sheet, which statement is correct only for the β-sheet? A. Extensive hydrogen bonds between the carbonyl oxygen (C=O) and the amide hydrogen (N−H) of the peptide bond are formed. B. It may be found in typical globular proteins. C. It is stabilized by interchain hydrogen bonds. | Biochemistry_Lippinco. .2. A particular point mutation results in disruption of the α-helical structure in a segment of the mutant protein. The most likely change in the primary structure of the mutant protein is: A. glutamate to aspartate. B. lysine to arginine. C. methionine to proline. D. valine to alanine. Correct answer = C. Proline, because of its secondary amino group, is incompatible with an α-helix. Glutamate, aspartate, lysine, and arginine are charged amino acids, and valine is a branched amino acid. Charged and branched (bulky) amino acids may disrupt an α-helix. [Note: The flexibility of glycine’s R group can also disrupt an α-helix.] .3. In comparing the α-helix to the β-sheet, which statement is correct only for the β-sheet? A. Extensive hydrogen bonds between the carbonyl oxygen (C=O) and the amide hydrogen (N−H) of the peptide bond are formed. B. It may be found in typical globular proteins. C. It is stabilized by interchain hydrogen bonds. |
Biochemistry_Lippincott_80 | Biochemistry_Lippinco | B. It may be found in typical globular proteins. C. It is stabilized by interchain hydrogen bonds. D. It is an example of secondary structure. E. It may be found in supersecondary structures. Correct answer = C. The β-sheet is stabilized by interchain hydrogen bonds formed between separate polypeptide chains and by intrachain hydrogen bonds formed between regions of a single polypeptide. The α-helix, however, is stabilized only by intrachain hydrogen bonds. Statements A, B, D, and E are true for both of these secondary structural elements. .4. An 80-year-old man presented with impairment of intellectual function and alterations in behavior. His family reported progressive disorientation and memory loss over the last 6 months. There is no family history of dementia. The patient was tentatively diagnosed with Alzheimer disease (AD). Which one of the following best describes AD? A. It is associated with β-amyloid, an abnormal protein with an altered amino acid sequence. | Biochemistry_Lippinco. B. It may be found in typical globular proteins. C. It is stabilized by interchain hydrogen bonds. D. It is an example of secondary structure. E. It may be found in supersecondary structures. Correct answer = C. The β-sheet is stabilized by interchain hydrogen bonds formed between separate polypeptide chains and by intrachain hydrogen bonds formed between regions of a single polypeptide. The α-helix, however, is stabilized only by intrachain hydrogen bonds. Statements A, B, D, and E are true for both of these secondary structural elements. .4. An 80-year-old man presented with impairment of intellectual function and alterations in behavior. His family reported progressive disorientation and memory loss over the last 6 months. There is no family history of dementia. The patient was tentatively diagnosed with Alzheimer disease (AD). Which one of the following best describes AD? A. It is associated with β-amyloid, an abnormal protein with an altered amino acid sequence. |
Biochemistry_Lippincott_81 | Biochemistry_Lippinco | A. It is associated with β-amyloid, an abnormal protein with an altered amino acid sequence. B. It results from accumulation of denatured proteins that have random conformations. C. It is associated with the accumulation of amyloid precursor protein. D. It is associated with the deposition of neurotoxic amyloid β peptide aggregates. E. It is an environmentally produced disease not influenced by the genetics of the individual. F. It is caused by the infectious β-sheet form of a host-cell protein. | Biochemistry_Lippinco. A. It is associated with β-amyloid, an abnormal protein with an altered amino acid sequence. B. It results from accumulation of denatured proteins that have random conformations. C. It is associated with the accumulation of amyloid precursor protein. D. It is associated with the deposition of neurotoxic amyloid β peptide aggregates. E. It is an environmentally produced disease not influenced by the genetics of the individual. F. It is caused by the infectious β-sheet form of a host-cell protein. |
Biochemistry_Lippincott_82 | Biochemistry_Lippinco | E. It is an environmentally produced disease not influenced by the genetics of the individual. F. It is caused by the infectious β-sheet form of a host-cell protein. Correct answer = D. Alzheimer disease (AD) is associated with long, fibrillar protein assemblies consisting of β-pleated sheets found in the brain and elsewhere. The disease is associated with abnormal processing of a normal protein. The accumulated altered protein occurs in a β-pleated sheet conformation that is neurotoxic. The amyloid β that is deposited in the brain in AD is derived by proteolytic cleavages from the larger amyloid precursor protein, a single transmembrane protein expressed on the cell surface in the brain and other tissues. Most cases of AD are sporadic, although at least 5% of cases are familial. Prion diseases, such as Creutzfeldt-Jakob, are caused by the infectious β-sheet form (PrPSc) of a host-cell protein (PrPC). | Biochemistry_Lippinco. E. It is an environmentally produced disease not influenced by the genetics of the individual. F. It is caused by the infectious β-sheet form of a host-cell protein. Correct answer = D. Alzheimer disease (AD) is associated with long, fibrillar protein assemblies consisting of β-pleated sheets found in the brain and elsewhere. The disease is associated with abnormal processing of a normal protein. The accumulated altered protein occurs in a β-pleated sheet conformation that is neurotoxic. The amyloid β that is deposited in the brain in AD is derived by proteolytic cleavages from the larger amyloid precursor protein, a single transmembrane protein expressed on the cell surface in the brain and other tissues. Most cases of AD are sporadic, although at least 5% of cases are familial. Prion diseases, such as Creutzfeldt-Jakob, are caused by the infectious β-sheet form (PrPSc) of a host-cell protein (PrPC). |
Subsets and Splits