1. Trang chủ
  2. » Kỹ Thuật - Công Nghệ

Biochemistry, 4th Edition P22 pps

10 651 0

Đang tải... (xem toàn văn)

THÔNG TIN TÀI LIỆU

Nội dung

6.5 How Do Protein Subunits Interact at the Quaternary Level of Protein Structure? 173 6.5 How Do Protein Subunits Interact at the Quaternary Level of Protein Structure? Many proteins exist in nature as oligomers, complexes composed of (often sym- metric) noncovalent assemblies of two or more monomer subunits. In fact, subunit association is a common feature of macromolecular organization in biology. Most intracellular enzymes are oligomeric and may be composed either of a single type of monomer subunit (homomultimers) or of several different kinds of subunits (het- eromultimers). The simplest case is a protein composed of identical subunits. Liver al- cohol dehydrogenase, shown in Figure 6.41, is such a protein. Alcohol consumed in a beer or mixed drink is oxidized in the liver by alcohol dehydrogenase. Hormonal signals modulate blood sugar levels by controlling the activity of glycogen phos- phorylase, an elegantly regulated homodimeric muscle enzyme. Oxygen is carried in the blood by hemoglobin, which contains two each of two different subunits (het- erotetramer). A counterpoint to these small clusters is made by the proteins that form large polymeric aggregates. Proteins are synthesized on large complexes of many pro- tein units and several RNA molecules called ribosomes. Muscle contraction depends on large polymer clusters of the protein myosin sliding along filamentous polymers of another protein, actin. The way in which separate folded monomeric protein subunits associate to form the oligomeric protein constitutes the quaternary struc- ture of that protein. Table 6.3 lists several proteins and their subunit compositions (see also Table 4.2). Proteins with two to four subunits predominate in nature, but many cases of higher numbers exist. The subunits of an oligomeric protein typically fold independently and then in- teract with other subunits. The surfaces at which subunits interact are similar in na- ture to the interiors of the individual subunits—closely packed with both polar and hydrophobic interactions. Interacting surfaces must therefore possess complemen- tary arrangements of polar and hydrophobic groups. Oligomeric associations of protein subunits can be divided into those between identical subunits and those between nonidentical subunits. Interactions among identical subunits can be further distinguished as either isologous or heterologous. In isologous interactions, the interacting surfaces are identical and the resulting structure is necessarily dimeric and closed, with a twofold axis of symmetry (Figure 6.42). If any additional interactions occur to form a trimer or tetramer, these must use different interfaces on the protein’s surface. Many proteins, such as transthy- retin, form tetramers by means of two sets of isologous interactions (Figure 6.43). Such structures possess three different twofold axes of symmetry. In contrast, het- erologous associations among subunits involve nonidentical interfaces. These sur- faces must be complementary, but they are generally not symmetric. FIGURE 6.41 The quaternary structure of liver alcohol dehydrogenase.Within each subunit is a six-stranded parallel sheet. Between the two subunits is a two-stranded antiparallel sheet (pdb id ϭ 1CDO).(Jane Richardson.) Number of Protein Subunits Alcohol dehydrogenase 2 Malate dehydrogenase 2 Superoxide dismutase 2 Triose phosphate isomerase 2 Glycogen phosphorylase 2 Aldolase 3 Bacteriochlorophyll protein 3 Concanavalin A 4 Glyceraldehyde-3-phosphate 4 dehydrogenase Immunoglobulin 4 Lactate dehydrogenase 4 Prealbumin 4 Pyruvate kinase 4 Phosphoglycerate mutase 4 Hemoglobin 2 ϩ 2 Insulin 6 Aspartate transcarbamoylase 6 ϩ 6 Glutamine synthetase 12 TMV protein disc 17 Apoferritin 24 Coat of tomato bushy stunt 180 virus TABLE 6.3 Aggregation Symmetries of Globular Proteins 174 Chapter 6 Proteins: Secondary,Tertiary, and Quaternary Structure There Is Symmetry in Quaternary Structures Many multimeric proteins are symmetric arrangements of asymmetric objects (the monomer subunits). All of the polypeptide’s ␣-carbons are asymmetric, and the polypeptide nearly always folds to form a low-symmetry structure. (The long helical arrays formed by some synthetic polypeptides are an exception.) Thus, protein sub- units do not have mirror reflection planes, points, or axes of inversion. The only sym- metry operation possible for protein subunits is a rotation. The most common sym- metries observed for multisubunit proteins are cyclic symmetry and dihedral symmetry. In cyclic symmetry, the subunits are arranged around a single rotation axis, as shown in Figure 6.44. If there are two subunits, the axis is referred to as a twofold rotation axis. Rotating the quaternary structure 180° about this axis gives a structure identical to the original one. With three subunits arranged about a three- fold rotation axis, a rotation of 120° about that axis gives an identical structure. Dihedral symmetry occurs when a structure possesses at least one twofold rotation axis perpendicular to another n-fold rotation axis. This type of subunit arrangement (Figure 6.44) occurs in annexin XII (where n ϭ 3). Quaternary Association Is Driven by Weak Forces Weak forces stabilize quaternary structures. Typical dissociation constants for simple two-subunit associations range from 10 Ϫ8 to 10 Ϫ16 M. These values correspond to free energies of association of about 50 to 100 kJ/mol at 37°C. Dimerization of subunits is accompanied by both favorable and unfavorable energy changes. The favorable in- teractions include van der Waals interactions, hydrogen bonds, ionic bonds, and hy- drophobic interactions. However, considerable entropy loss occurs when subunits in- teract. When two subunits move as one, three translational degrees of freedom are lost for one subunit because it is constrained to move with the other one. In addition, (a) Isologous association (b) Heterologous association (c) Heterologous tetramer (d) Isologous tetramer Symmetry axis FIGURE 6.42 Isologous and heterologous associations between protein subunits.(a) An isologous interaction be- tween two subunits with a twofold axis of symmetry perpendicular to the plane of the page.(b) A heterologous interaction that could lead to the formation of a long polymer. (c) A heterologous interaction leading to a closed structure—a tetramer. (d) A tetramer formed by two sets of isologous interactions. Monomer A Monomer A Monomer B Monomer B Monomer AЈ Monomer BЈ (a) (b) FIGURE 6.43 Many proteins form tetramers by means of two sets of isologous interactions.The dimeric (a) and tetrameric (b) forms of transthyretin (also known as prealbumin) are shown here (pdb id ϭ 1GKE).The monomers of this protein form a dimer in a manner that extends the large monomer ␤-sheet.The tetramer is formed by isologous interactions between the large ␤-sheets of two dimers. 6.5 How Do Protein Subunits Interact at the Quaternary Level of Protein Structure? 175 Tetramer B. anthracis dihydrodipicolinate synthase (pdb id = 1XL9) Pentamer Shiga-like toxin I B (pdb id = 1CZG) Trimer E. blattae acid phosphatase (pdb id = 2EOI) Dimer ARNT PAS-B (pdb id = 2HV1) Bundled hexamer Uridylate kinase (pdb id = 2A1F) Trimer of dimers Annexin XII (pdb id = 1DM5) Cyclic hexamer Circadian clock protein KaiC (pdb id = 1TF7) Octamer Limulus polyphemus SAP-like pentraxin ( p db id = 1QTJ) Heptamer M. tuberculosis chaperonin-10 ( p db id = 1HX5) Dodecamer Lactococcus lactis MG1363 DpsB protein ( p db id = 1ZS3) FIGURE 6.44 Multimeric proteins are symmetric arrangements of asymmetric objects. A variety of symmetries is displayed in these multimeric structures. 176 Chapter 6 Proteins: Secondary,Tertiary, and Quaternary Structure many peptide residues at the subunit interface, which were previously free to move on the protein surface, now have their movements restricted by the subunit associa- tion. This unfavorable energy of association is in the range of 80 to 120 kJ/mol for temperatures of 25° to 37°C. Thus, to achieve stability, the dimerization of two sub- units must involve approximately 130 to 220 kJ/mol of favorable interactions. 1 Van der Waals interactions at protein interfaces are numerous, often running to several hundred for a typical monomer–monomer association. This would account for about 150 to 200 kJ/mol of favorable free energy of association. However, when solvent is removed from the protein surface to form the subunit–subunit contacts, nearly as many van der Waals associations are lost as are made. One subunit is simply trading water molecules for peptide residues in the other subunit. As a result, the energy of subunit association due to van der Waals interactions actually contributes little to the stability of the dimer. Hydrophobic interactions at the subunit–subunit interface, however, are generally very favorable. For many proteins, the subunit association process effectively buries as much as 20 nm 2 of surface area previously exposed to sol- vent, resulting in as much as 100 to 200 kJ/mol of favorable hydrophobic interactions. Together with whatever polar interactions occur at the protein–protein interface, this is sufficient to account for the observed stabilization that occurs when two protein subunits associate. An additional and important factor contributing to the stability of subunit asso- ciations for some proteins is the formation of disulfide bonds between different sub- units. All antibodies are ␣ 2 ␤ 2 -tetramers composed of two heavy chains (53 to 75 kD) and two light chains (23 kD). In addition to intrasubunit disulfide bonds (four per heavy chain, two per light chain), two intersubunit disulfide bridges hold the two heavy chains together and a disulfide bridge links each of the two light chains to a heavy chain (Figure 6.45). 1 For example, 130 kJ/mol of favorable interaction minus 80 kJ/mol of unfavorable interaction equals a net free energy of association of 50 kJ/mol. SS S S SS S S SS SS SS SS SS SS S S S S N N S S S S S S S S C (CH 2 O) n addition site C C H 2C H 3 446 C H 1 V H V L C L Hinge region N N Heavy Light 214 Antigen binding Antigen binding 4.5 nm FIGURE 6.45 Schematic drawing of an immunoglobulin molecule, showing the intermolecular and intra- molecular disulfide bonds.Two identical L chains are joined with two identical H chains. Each L chain is held to an H chain via an interchain disulfide bond.The variable regions of the four polypeptides lie at the ends of the arms of the Y-shaped molecule.These regions are responsible for the antigen recognition function of antibody molecules. For purposes of illustration, some features are shown on only one or the other L chain or H chain, but all features are common to both chains. 6.5 How Do Protein Subunits Interact at the Quaternary Level of Protein Structure? 177 Open Quaternary Structures Can Polymerize All of the quaternary structures we have considered to this point have been closed structures, with a limited capacity to associate. Many proteins in nature associate to form open heterologous structures, which can polymerize more or less indefinitely, creating structures that are both esthetically attractive and functionally important to the cells or tissue in which they exist. One such protein is tubulin, the ␣␤-dimeric protein that polymerizes into long, tubular structures that are the structural basis of cilia, flagella, and the cytoskeletal matrix. The microtubule thus formed (Figure 6.46) may be viewed as consisting of 13 parallel filaments arising from end-to-end aggregation of the tubulin dimers. Human immunodeficiency virus, HIV, the causative agent of AIDS (also discussed in Chapter 14), is enveloped by a spherical shell composed of hundreds of coat protein subunits, a large-scale, but closed, qua- ternary association. There Are Structural and Functional Advantages to Quaternary Association There are several important consequences when protein subunits associate in oligomeric structures. Stability One general benefit of subunit association is a favorable reduction of the protein’s surface-to-volume ratio. The surface-to-volume ratio becomes smaller as the radius of any particle or object becomes larger. (This is because surface area is a function of the radius squared and volume is a function of the radius cubed.) Because interactions within the protein usually tend to stabilize the protein ener- getically and because the interaction of the protein surface with solvent water is of- ten energetically unfavorable, decreased surface-to-volume ratios usually result in more stable proteins. Subunit association may also serve to shield hydrophobic residues from solvent water. Subunits that recognize either themselves or other sub- units avoid any errors arising in genetic translation by binding mutant forms of the subunits less tightly. Genetic Economy and Efficiency Oligomeric association of protein monomers is genetically economical for an organism. Less DNA is required to code for a monomer that assembles into a homomultimer than for a large polypeptide of the same molecular mass. Another way to look at this is to realize that virtually all of the information that determines oligomer assembly and subunit–subunit interac- tion is contained in the genetic material needed to code for the monomer. For ex- ample, HIV protease, an enzyme that is a dimer of identical subunits, performs a catalytic function similar to homologous cellular enzymes that are single polypep- tide chains of twice the molecular mass (see Chapter 14). A DEEPER LOOK Immunoglobulins—All the Features of Protein Structure Brought Together The immunoglobulin structure in Figure 6.45 represents the con- fluence of all the details of protein structure that have been thus far discussed. As for all proteins, the primary structure determines other aspects of structure. There are numerous elements of sec- ondary structure, including ␤-sheets and tight turns. The tertiary structure consists of 12 distinct domains, and the protein adopts a heterotetrameric quaternary structure. To make matters more in- teresting, both intrasubunit and intersubunit disulfide linkages act to stabilize the discrete domains and to stabilize the tetramer itself. One more level of sophistication awaits. As discussed in Chap- ter 28, the amino acid sequences of both light and heavy immuno- globulin chains are not constant! Instead, the primary structure of these chains is highly variable in the N-terminal regions (first 108 residues). Heterogeneity of the amino acid sequence leads to vari- ations in the conformation of these variable regions. This variation accounts for antibody diversity and the ability of antibodies to rec- ognize and bind a virtually limitless range of antigens. This full po- tential of antibodyϺantigen recognition enables organisms to mount immunological responses to almost any antigen that might challenge the organism. 8.0 nm 3.5- to 4.0-nm subunit α β FIGURE 6.46 The structure of a typical microtubule, showing the arrangement of the ␣- and ␤-monomers of the tubulin dimer. 178 Chapter 6 Proteins: Secondary,Tertiary, and Quaternary Structure Bringing Catalytic Sites Together Many enzymes (see Chapters 13 to 15) derive at least some of their catalytic power from oligomeric associations of monomer sub- units. This can happen in several ways. The monomer may not constitute a com- plete enzyme active site. Formation of the oligomer may bring all the necessary cat- alytic groups together to form an active enzyme. For example, the active sites of bacterial glutamine synthetase are formed from pairs of adjacent subunits. The dis- sociated monomers are inactive. Oligomeric enzymes may also carry out different but related reactions on differ- ent subunits. Thus, tryptophan synthase is a tetramer consisting of pairs of different subunits, ␣ 2 ␤ 2 . Purified ␣-subunits catalyze the following reaction: Indoleglycerol phosphate34indole ϩ glyceraldehyde-3-phosphate and the ␤-subunits catalyze this reaction: Indole ϩ L-serine34L-tryptophan Indole, the product of the ␣-reaction and the reactant for the ␤-reaction, is passed directly from the ␣-subunit to the ␤-subunit and cannot be detected as a free inter- mediate. Cooperativity There is another, more important consequence when monomer subunits associate into oligomeric complexes. Most oligomeric enzymes regulate catalytic activity by means of subunit interactions, which may give rise to coopera- tive phenomena. Multisubunit proteins typically possess multiple binding sites for a given ligand. If the binding of ligand at one site changes the affinity of the protein for ligand at the other binding sites, the binding is said to be cooperative. Infor- mation transfer in this manner across long distances in proteins is termed allostery, literally “action at another site.” Increases in affinity at subsequent sites represent positive cooperativity, whereas decreases in affinity correspond to negative cooper- ativity. The points of contact between protein subunits provide a mechanism for this signal transduction through the protein structure and for communication between the subunits. This in turn provides a way in which the binding of ligand to one sub- unit can influence the binding behavior at the other subunits. Such cooperative be- havior, discussed in greater depth in Chapter 15, is the underlying mechanism for regulation of many biological processes. HUMAN BIOCHEMISTRY Faster-Acting Insulin: Genetic Engineering Solves a Quaternary Structure Problem Insulin is a peptide hormone secreted by the pancreas that regu- lates glucose metabolism in the body. Insufficient production of insulin or failure of insulin to stimulate target sites in liver, muscle, and adipose tissue leads to the serious metabolic disorder known as diabetes mellitus. Diabetes afflicts millions of people worldwide. Diabetic individuals typically exhibit high levels of glucose in the blood, but insulin injection therapy allows these individuals to maintain normal levels of blood glucose. Insulin is composed of two peptide chains covalently linked by disulfide bonds (see Figure 5.8). This “monomer” of insulin is the active form that binds to receptors in target cells. However, in solution, insulin spontaneously forms dimers, which them- selves aggregate to form hexamers. The surface of the insulin molecule that self-associates to form hexamers is also the surface that binds to insulin receptors in target cells. Thus, hexamers of insulin are inactive. Insulin released from the pancreas is monomeric and acts rapidly at target tissues. However, when insulin is administered (by injection) to a diabetic patient, the insulin hexamers dissociate slowly and the patient’s blood glucose levels typically drop slowly (over several hours). In 1988, G. Dodson showed that insulin could be genetically engineered to prefer the monomeric (active) state. Dodson and his colleagues used recombinant DNA technology (discussed in Chapter 12) to produce insulin with an aspartate residue replac- ing a proline at the contact interface between adjacent subunits. The negative charge on the Asp side chain creates electrostatic repulsion between subunits and increases the dissociation con- stant for the hexamer 34 monomer equilibrium. Injection of this mutant insulin into test animals produced more rapid de- creases in blood glucose than did ordinary insulin. This mutant in- sulin, marketed by the Danish pharmaceutical company Novo as NovoLog in the United States and as NovoRapid in Europe, may eventually replace ordinary insulin in the treatment of diabetes. NovoLog has a faster rate of absorption, a faster onset of action, and a shorter duration of action than regular human insulin. It is particularly suited for mealtime dosing to control postprandial glycemia, the rise in blood sugar following consumption of food. Regular human insulin acts more slowly, so patients must usually administer it 30 minutes before eating. SUMMARY 6.1 What Noncovalent Interactions Stabilize Protein Structure? Sev- eral different kinds of noncovalent interactions are of vital importance in protein structure. Hydrogen bonds, hydrophobic interactions, electrosta- tic bonds, and van der Waals forces are all noncovalent in nature yet are extremely important influences on protein conformations. The stabiliza- tion free energies afforded by each of these interactions are highly de- pendent on the local environment within the protein. Hydrogen bonds are generally made wherever possible within a given protein structure. Hydrophobic interactions form because non- polar side chains of amino acids and other nonpolar solutes prefer to cluster in a nonpolar environment rather than to intercalate in a polar solvent such as water. Electrostatic interactions include the attraction between opposite charges and the repulsion of like charges in the pro- tein. Van der Waals interactions involve instantaneous dipoles and in- duced dipoles that arise because of fluctuations in the electron charge distributions of adjacent nonbonded atoms. 6.2 What Role Does the Amino Acid Sequence Play in Protein Struc- ture? All of the information necessary for folding the peptide chain into its “native” structure is contained in the amino acid sequence of the peptide. Just how proteins recognize and interpret the information that is stored in the polypeptide sequence is not yet well understood. It may be assumed that certain loci along the peptide chain act as nucleation points, which initiate folding processes that eventually lead to the cor- rect structures. Regardless of how this process operates, it must take the protein correctly to the final native structure, without getting trapped in a local energy-minimum state, which, although stable, may be different from the native state itself. 6.3 What Are the Elements of Secondary Structure in Proteins, and How Are They Formed? Secondary structure in proteins forms so as to maximize hydrogen bonding and maintain the planar nature of the peptide bond. Secondary structures include ␣-helices, ␤-sheets, and tight turns. 6.4 How Do Polypeptides Fold into Three-Dimensional Protein Struc- tures? First, secondary structures—helices and sheets—form when- ever possible as a consequence of the formation of large numbers of hy- drogen bonds. Second, ␣-helices and ␤-sheets often associate and pack close together in the protein. There are a few common methods for such packin g to occur. Third, because the peptide segments between secondary structures in the protein tend to be short and direct, the pep- tide does not execute complicated twists and knots as it moves from one region of a secondary structure to another. A consequence of these three principles is that protein chains are usually folded so that the sec- ondary structures are arranged in one of a few common patterns. For this reason, there are families of proteins that have similar tertiary struc- ture, with little apparent evolutionary or functional relationship among them. Finally, proteins generally fold so as to form the most stable struc- tures possible. The stability of most proteins arises from (1) the forma- tion of large numbers of intramolecular hydrogen bonds and (2) the re- duction in the surface area accessible to solvent that occurs upon folding. 6.5 How Do Protein Subunits Interact at the Quaternary Level of Protein Structure? The subunits of an oligomeric protein typically fold into ap- parently independent globular conformations and then interact with other subunits. The particular surfaces at which protein subunits interact are similar in nature to the interiors of the individual subunits. These in- terfaces are closely packed and involve both polar and hydrophobic in- teractions. Interacting surfaces must therefore possess complementary arrangements of polar and hydrophobic groups. Problems 179 PROBLEMS Preparing for an exam? Create your own study path for this chapter at www.cengage.com/login. 1. The central rod domain of a keratin protein is approximately 312 residues in length. What is the length (in Å) of the keratin rod domain? If this same peptide segment were a true ␣-helix, how long would it be? If the same seg ment were a ␤-sheet, what would its length be? 2. A teenager can grow 4 inches in a year during a “growth spurt.” As- suming that the increase in height is due to vertical growth of col- lagen fibers (in bone), calculate the number of collagen helix turns synthesized per minute. 3. Discuss the potential contributions to hydrophobic and van der Waals interactions and ionic and hydrogen bonds for the side chains of Asp, Leu, Tyr, and His in a protein. 4. Pro is the amino acid least commonly found in ␣-helices but most commonly found in ␤-turns. Discuss the reasons for this behavior. 5. For flavodoxin (pdb id ϭ 5NLL), identify the right-handed cross- overs and the left-handed cross-overs in the parallel ␤-sheet. 6. Choose any three regions in the Ramachandran plot and discuss the likelihood of observing that combination of ␾ and ␺ in a pep- tide or protein. Defend your answer using suitable molecular mod- els of a peptide. 7. A new protein of unknown structure has been purified. Gel filtra- tion chromatography reveals that the native protein has a molecu- lar weight of 240,000. Chromatography in the presence of 6 M guanidine hydrochloride yields only a peak for a protein of M r 60,000. Chromatography in the presence of 6 M guanidine hydro- chloride and 10 mM ␤-mercaptoethanol yields peaks for proteins of M r 34,000 and 26,000. Explain what can be determined about the structure of this protein from these data. 8. Two polypeptides, A and B, have similar tertiary structures, but A normally exists as a monomer, whereas B exists as a tetramer, B 4 . What differences might be expected in the amino acid composition of A versus B? 9. The hemagglutinin protein in influenza virus contains a remarkably long ␣-helix, with 53 residues. a. How long is this ␣-helix (in nm)? b. How many turns does this helix have? c. Each residue in an ␣-helix is involved in two H bonds. How many H bonds are present in this helix? 10. It is often observed that Gly residues are conserved in proteins to a greater degree than other amino acids. From what you have learned in this chapter, suggest a reason for this observation. 11. Which amino acids would be capable of forming H bonds with a lysine residue in a protein? 12. Poly- L-glutamate adopts an ␣-helical structure at low pH but becomes a random coil above pH 5. Explain this behavior. 13. Imagine that the dimensions of the alpha helix were such that there were exactly 3.5 amino acids per turn, instead of 3.6. What would be the consequences for coiled-coil structures? Preparing for the MCAT Exam 14. Consider the following peptide sequences: EANQIDEMLYNVQCSLTTLEDTVPW LGVHLDITVPLSWTWTLYVKL QQNWGGLVVILTLVWFLM CNMKHGDSQCDERTYP YTREQSDGHIPKMNCDS AGPFGPDGPTIGPK 180 Chapter 6 Proteins: Secondary,Tertiary, and Quaternary Structure Which of the preceding sequences would be likely to be found in each of the following: a. A parallel ␤-sheet b. An antiparallel ␤-sheet c. A tropocollagen molecule d. The helical portions of a protein found in your hair 15. To fully appreciate the elements of secondary structure in proteins, it is useful to have a practical sense of their structures. On a piece of paper, draw a simple but larg e zigzag pattern to represent a ␤-strand. Then fill in the structure, drawing the locations of the atoms of the chain on this zigzag pattern. Then draw a simple, large coil on a piece of paper to represent an ␣-helix. Then fill in the structure, drawing the backbone atoms in the correction loca- tions along the coil and indicating the locations of the R groups in your drawing. 16. The dissociation constant for a particular protein dimer is 1 micro- molar. Calculate the free energ y difference for the monomer to dimer transition. FURTHER READING General Branden, C., and Tooze, J., 1991. Introduction to Protein Structure. New York: Garland Publishing. Chothia, C., 1984. Principles that determine the structure of proteins. Annual Review of Biochemistry 53:537–572. Fink, A., 2005. Natively unfolded proteins. Current Opinion in Structural Biology 15:35-41. Greene, L., Lewis, T., Addou, S., et al., 2006. The CATH domain struc- ture database: New protocols and classification levels give a more comprehensive resource for exploring evolution. Nucleic Acids Re- search 35:D291–D297. Hardie, D G., and Coggins, J. R., eds., 1986. Multidomain Proteins: Struc- ture and Evolution. New York: Elsevier. Harper, E., and Rose, G. D., 1993. Helix stop signals in proteins and peptides: The capping box. Biochemistry 32:7605–7609. Judson, H. F., 1979. The Eighth Day of Creation. New York: Simon and Schuster. Lupas, A., 1996. Coiled coils: New structures and new functions. Trends in Biochemical Sciences 21:375–382. Petsko, G., and Ringe, D., 2004. Protein Structure and Function. London: New Science Press. Richardson, J. S., 1981. The anatomy and taxonomy of protein struc- ture. Advances in Protein Chemistry 34:167–339. Schulze, A. J., Huber, R., Bode, W., and Engh, R. A., 1994. Structural as- pects of serpin inhibition. FEBS Letters 344:117–124. Smith, T., 2000. Structural Genomics—special supplement. Nature Struc- tural Biology Volume 7, Issue 11S. This entire supplemental issue is devoted to structural genomics and contains a trove of information about this burgeoning field. Tompa, P., 2002. Intrinsically unstructured proteins. Trends in Biochemi- cal Sciences 27:527–533. Tompa, P., Szasz, C., and Buday, L., 2005. Structural disorder throws new light on moonlighting. Trends in Biochemical Sciences 30:484–489. Uversky, V. N., 2002. Natively unfolded proteins: A point where biology waits for physics. Protein Science 11:739–756. Webster, D. M., 2000. Protein Structure Prediction—Methods and Protocols. New Jersey: Humana Press. Protein Folding Aurora, R., Creamer, T., Srinivasan, R., and Rose, G. D., 1997. Local in- teractions in protein folding: Lessons from the ␣-helix. The Journal of Biological Chemistry 272:1413–1416. Baker, D., 2000. A surprising simplicity to protein folding. Nature 405: 39–42. Creighton, T. E., 1997. How important is the molten globule for correct protein folding? Trends in Biochemical Sciences 22:6–11. Deber, C. M., and Therien, A. G., 2002. Putting the ␤-breaks on mem- brane protein misfolding. Nature Structural Biology 9:318–319. Dill, K. A., and Chan, H. S., 1997. From Levinthal to pathways to fun- nels. Natur e Structural Biology 4:10–19. Dinner, A. R., Sali, A., Smith, L. J., Dobson, C. M., and Karplus, M., 2001. Understanding protein folding via free-energy surfaces from theory and experiment. Trends in Biochemical Sciences 25:331–339. Han, J H., Batey, S., Nickson, A., et al., 2007. The folding and evolution of multidomain proteins. Nature Reviews Molecular Cell Biology 8: 319–330. Kelly, J., 2005. Structural biology: Form and function instructions. Nature 437:486–487. Mirny, L., and Shakhnovich, E., 2001. Protein folding theory: From lat- tice to all-atom models. Annual Review of Biophysics and Biomolecular Structure 30:361–396. Mok, K., Kuhn, L., Goez, M., et al., 2007. A pre-existing hydrophobic collapse in the unfolded state of an ultrafast folding protein. Nature 447:106–109. Murphy, K. P., 2001. Protein Structure, Stability, and Folding. New Jersey: Humana Press. Myers, J. K., and Oas, T. G., 2002. Mechanisms of fast protein foldin g. Annual Review of Biochemistry 71:783–815. Orengo, C., and Thornton, J., 2005. Protein families and their evolution— a structural perspective. Annual Review of Biochemistry 74:867–900. Radford, S. E., 2000. Protein folding: Progress made and promises ahead. Trends in Biochemical Sciences 25:611–618. Raschke, T. M., and Marqusee, S., 1997. The kinetic folding intermedi- ate of ribonuclease H resembles the acid molten globule and par- tially unfolded molecules detected under native conditions. Nature Structural Biology 4:298–304. Srinivasan, R., and Rose, G. D., 1995. LINUS: A hierarchic procedure to predict the fold of a protein. Proteins: Structure, Function and Genetics 22:81–99. Secondary Structure Xiong, H., Buckwalter, B., Shieh, H., and Hecht, M. H., 1995. Periodic- ity of polar and nonpolar amino acids is the major determinant of secondary structure in self-assembling oligomeric peptides. Proceed- ings of the National Academy of Sciences 92:6349–6353. Structural Studies Bradley, P., Misura, K., and Baker, D., 2005. Toward high-resolution de novo structure prediction for small proteins. Science 309:1868–1871. Hadley, C., and Jones, D., 1999. A systematic comparison of protein struc- ture classifications: SCOP, CATH, and FSSP. Structure 7:1099–1112. Lomas, D., Belorgey, D., Mallya, M., et al., 2005. Molecular mousetraps and the serpinopathies. Biochemical Society Transactions 33 (part 2): 321–330. Wagner, G., Hyberts, S., and Havel, T., 1992. NMR structure determina- tion in solution: A critique and comparison with X-ray crystallogra- phy. Annual Review of Biophysics and Biomolecular Structure 21:167–242. Wand, A. J., 2001. Dynamic activation of protein function: A view emerg- ing from NMR spectroscopy. Nature Structural Biology 8:926–931. Diseases of Protein Folding Bucchiantini, M., et al., 2002. Inherent toxicity of aggregates implies a common mechanism for protein misfolding diseases. Nature 416: 507–511. Sifers, R. M., 1995. Defective protein folding as a cause of disease. Na- ture Structural Biology 2:355–367. Stein, P . E., and Carrell, R. W., 1995. What do dysfunctional serpins tell us about molecular mobility and disease? Nature Structural Biology 2:96–113. Thomas, P. J., Qu, B-H. and Pedersen, P. L., 1995. Defective protein fold- ing as a basis of human disease. Trends in Biochemical Sciences 20: 456–459. © Burstein Collection/CORBIS 7 Carbohydrates and the Glycoconjugates of Cell Surfaces Carbohydrates are the single most abundant class of organic molecules found in na- ture. Energy from the sun captured by green plants, algae, and some bacteria dur- ing photosynthesis (see Chapter 21) converts more than 250 billion kilograms of carbon dioxide into carbohydrates every day on earth. In turn, carbohydrates are the metabolic precursors of virtually all other biomolecules. Breakdown of carbo- hydrates provides the energy that sustains animal life. In addition, carbohydrates are covalently linked with a variety of other molecules. These glycoconjugates are important components of cell walls and extracellular structures in plants, animals, and bacteria. In addition to the structural roles such molecules play, they also serve in a variety of processes involving recognition between cell types or recognition of cel- lular structures by other molecules. Recognition events are important in normal cell growth, fertilization, transformation of cells, and other processes. All of these functions are made possible by the characteristic chemical features of carbohydrates: • the existence of at least one and often two or more asymmetric centers • the ability to exist either in linear or ring structures • the capacity to form polymeric structures via glycosidic bonds • the potential to form multiple hydrogen bonds with water or other molecules in their environment. 7.1 How Are Carbohydrates Named? The name carbohydrate arises from the basic molecular formula (CH 2 O) n , where n ϭ 3 or more. (CH 2 O) n can be rewritten (CиH 2 O) n to show that these substances are hydrates of carbon. Carbohydrates are generally classified into three groups: monosaccharides (and their derivatives), oligosaccharides, and polysaccharides. The monosaccharides are also called simple sugars and have the formula (CH 2 O) n . Monosaccharides cannot be broken down into smaller sugars under mild conditions. Oligosaccha- rides derive their name from the Greek word oligo, meaning “few,” and consist of from two to ten simple sugar residues. Disaccharides are common in nature, and trisaccharides also occur frequently. Four- to six-sugar-unit oligosaccharides are usually bound covalently to other molecules, including glycoproteins. As their name suggests, polysaccharides are polymers of the simple sugars and their deriv- atives. They may be either linear or branched polymers and may contain hundreds or even thousands of monosaccharide units. Their molecular weights range up to 1 million or more. “The Discovery of Honey”—Piero di Cosimo (1492). Sugar in the gourd and honey in the horn, I never was so happy since the hour I was born. Turkey in the Straw, stanza 6 (classic American folk tune) KEY QUESTIONS 7.1 How Are Carbohydrates Named? 7.2 What Is the Structure and Chemistry of Monosaccharides? 7.3 What Is the Structure and Chemistry of Oligosaccharides? 7.4 What Is the Structure and Chemistry of Polysaccharides? 7.5 What Are Glycoproteins, and How Do They Function in Cells? 7.6 How Do Proteoglycans Modulate Processes in Cells and Organisms? 7.7 Do Carbohydrates Provide a Structural Code? ESSENTIAL QUESTION Carbohydrates are a versatile class of molecules of the formula (CH 2 O) n .They are a major form of stored energy in organisms, and they are the metabolic precursors of virtually all other biomolecules. Conjugates of carbohydrates with proteins and lipids perform a variety of functions, including recognition events that are important in cell growth, transformation, and other processes. What is the structure, chemistry, and biological function of carbohydrates? Create your own study path for this chapter with tutorials, simulations, animations, and Active Figures at www.cengage.com/ login 182 Chapter 7 Carbohydrates and the Glycoconjugates of Cell Surfaces 7.2 What Is the Structure and Chemistry of Monosaccharides? Monosaccharides Are Classified as Aldoses and Ketoses Monosaccharides consist typically of three to seven carbon atoms and are described either as aldoses or ketoses, depending on whether the molecule contains an alde- hyde function or a ketone group. The simplest aldose is glyceraldehyde, and the simplest ketose is dihydroxyacetone (Figure 7.1). These two simple sugars are termed trioses because they each contain three carbon atoms. The structures and names of a family of aldoses and ketoses with three, four, five, and six carbons are shown in Figures 7.2 and 7.3. Hexoses are the most abundant sugars in nature. Nev- ertheless, sugars from all these classes are important in metabolism. HCOH CHO CH 2 OH 2 1 3 D-Glyceraldehyde Carbon number HCOH CH 2 OH 3 4 D-Erythrose HCOH CHO 2 1 HCOH CH 2 OH D-Threose HOCH CHO HCOH CH 2 OH 4 5 D-Ribose (Rib) HCOH 3 HCOH CHO 2 1 HCOH CH 2 OH D-Arabinose (Ara) HCOH HOCH CHO HCOH CH 2 OH D-Xylose (Xyl) HOCH HCOH CHO HCOH CH 2 OH D-Lyxose (Lyx) HOCH HOCH CHO HCOH CH 2 OH 5 D-Allose HCOH 4 HCOH3 HCOH CHO 2 1 Carbon number Carbon number Carbon number 6 HCOH CH 2 OH D-Altrose HCOH HCOH HOCH CHO HCOH CH 2 OH D-Glucose (Glc) HCOH HOCH HCOH CHO HCOH CH 2 OH D-Mannose (Man) HCOH HOCH CHO HOCH HCOH CH 2 OH D-Gulose HOCH HCOH CHO HCOH CH 2 OH D-Idose HCOH HOCH CHO HCOH HOCH HCOH CH 2 OH D-Galactose (Gal) HOCH HCOH CHO HCOH CH 2 OH D-Talose HOCH CHO HOCH HOCH HOCH ALDOTRIOSE ALDOTETROSES ALDOPENTOSES ALDOHEXOSES FIGURE 7.2 The structure and stereochemical relationships of D-aldoses with three to six carbons. The configu- ration in each case is determined by the highest numbered asymmetric carbon (shown in pink). In each row, the “new” asymmetric carbon is shown in yellow. Blue highlights indicate the most common aldoses. HO C H CH 2 OH H or OHC CH 2 OH H L-Isomer D-Isomer Glyceraldehyde CH 2 OH CH 2 OH Dihydroxy- acetone O C O C HO C HHO CHO CH 2 OH L-Glyceraldehyde OHH CHO CH 2 OH D-Glyceraldehyde CC FIGURE 7.1 Structure of a simple aldose (glyceralde- hyde) and a simple ketose (dihydroxyacetone). Glyco: A generic term relating to sugars.

Ngày đăng: 06/07/2014, 14:20

TỪ KHÓA LIÊN QUAN

w