throbber
DONALD VOET
`University ofPennsylvania
`
`JUDITH G. VOET
`
`Swarthmore College
`
`BIOCHEMISTRY
`
`SECOND EDITION
`
`JOHN WILEY & SONS, INC.
`New York Chichester Brisbane
`Toronto Singapore
`
`1EX' 2029
`CFAD V NPS
`IPR2015—01093
`
`Page 1
`
`Page 1
`
`NPS EX. 2029
`Part 1
`CFAD v. NPS
`IPR2015-01093
`
`

`
`Cover Art: Two paintings of horse heart cytochrome c by
`Irving Geis in which the protein is illuminated by its sin-
`gle iron atom. On the front cover the hydrophilic side
`chains are drawn in green, and on the back cover the hy-
`drophobic side chains are drawn in orange. The paintings
`are based on an X-ray structure by Richard Dickerson.
`
`Acquisitions Editor: Nedah Rose
`Marketing Manager: Catherine Faduska
`Designer: Madelyn Lesure
`Manufacturing Manager: Susan Stetzer
`Photo Researcher: Hilary Newman
`Illustration Coordinator: Edward Starr
`Illustrators: J/B Woolsey Associates
`Production Management: Pamela Kennedy Oborski and
`Suzanne Ingrao
`Cover and Part Opening Illustrations © Irving Geis
`This book was set in 9.5/12 Times Roman by Progressive
`Information Technologies and printed and bound by
`Von Hoffman Press. The cover was printed by Lehigh
`Press Lithographers. Color separations by Lehigh Press
`Colortronics.
`
`Recognizing the importance of preserving what has been written, it is a policy of John Wiley &
`Sons, Inc. to have books of enduring value published in the United States printed on acid-free
`paper, and we exert our best efforts to that end.
`
`Copyright © 1995, by John Wiley & Sons, Inc.
`
`All rights reserved. Published simultaneously in Canada.
`
`Reproduction or translation of any part of
`this work beyond that permitted by Sections
`107 and 108 of the 1976 United States Copyright
`Act without the permission of the copyright
`owner is unlawful. Requests for permission
`or further information should be addressed to
`the Permissions Department, John Wiley & Sons, Inc.
`
`ISBN: 0-471-58651-X
`
`Printed in the United States of America
`
`10987654
`
`Page 2
`
`Page 2
`
`

`
`
`
`CHAPTER
`
`
`
`Three-Dimensional
`Structures OfProteins
`
`
`
`The properties of a protein are largely determined by its
`three-dimensional structure. One might naively suppose
`that since proteins are composed of the same 20 types of
`amino acid residues, they would be more or less alike in
`their properties. Indeed, denatured (unfolded) proteins
`have rather similar characteristics, a kind of homogeneous
`
`“average” of their randomly dangling side chains. How-
`ever, the three-dimensional structure of a native (physio-
`logically folded) protein is specified by its primary structure
`so that it has a unique set of characteristics.
`In this chapter, we shall discuss the structural features of
`proteins, the forces that hold them together, and their hier-
`archical organization to form complex structures. This will
`form the basis for understanding the structure—function
`relationships necessary to comprehend the biochemical
`roles of proteins. Detailed consideration of the dynamic
`behavior of proteins and how they fold to their native struc-
`tures is deferred until Chapter 8.
`
`1. SECONDARY STRUCTURE
`
`A po1ymer’s secondary structure (2° structure) is defined as
`the local conformation of its backbone. For proteins, this
`has come to mean the specification of regular polypeptide
`backbone folding patterns: helices, pleated sheets, and
`turns. However, before we begin our discussion of these
`basic structural motifs let us consider the geometrical prop-
`erties of the peptide group because its understanding is pre-
`requisite to that of any structure containing it.
`
`
`
`Page 3
`
`1. Secondary Structure
`A. The Peptide Group
`B. Helical Structures
`C. Beta Structures
`D. Nonrepetitive Structures
`2. Fibrous Proteins
`A.
`or Keratin—A Helix of Helices
`B. Silk Fibroin—A B Pleated Sheet
`C. Collagen — A Triple Helical Cable
`D. Elastin —— A Nonrepetitive Coil
`3. Globular Proteins
`A.
`Interpretation of Protein X—Ray and NMR Structures
`B. Tertiary Structure
`
`4. Protein Stability
`Electrostatic Forces
`
`.m.U.O.==.>
`
`Hydrogen Bonding Forces
`Hydrophobic Forces
`Disulfide Bonds
`Protein Denaturation
`
`5. Quaternary Structure
`A. Subunit Interactions
`
`B. Symmetry in Proteins
`C. Determination of Subunit Composition
`
`Appendix: Viewing Stereo Pictures
`
`141
`
`Page 3
`
`

`
`
`142 Chapter 7. Three-Dimensional Structures OfProteins
`
`
`
`
`
`A. The Peptide Group
`
`In the 1930s and 1940s, Linus Pauling and Robert Corey
`determined the X-ray structures of several amino acids and
`dipeptides in an effort to elucidate the structural constraints
`on the conformations of a polypeptide chain. These studies
`indicated that the peptide group has a rigid, planar structure
`(Fig. 7-1) which, Pauling pointed out, is a consequence of
`resonance interactions that give the peptide bond an ~40%
`double-bond character:
`
`FIGURE 7-1. The standard dimensions (in angstroms, A, and
`degrees,‘’) of the planar trans-peptide group derived by
`averaging the results of X-ray crystal structure determinations of
`amino acids and peptides. [After Marsh, R.E. and Donohue,
`J ., Adv. Protein Chem. 22, 249 (1967).]
`
`This explanation is supported by the observations that a
`peptide’s C—N bond is 0.13 A shorter than its N—C,,
`single bond and that its C=O bond is 0.02 A longer than
`that of aldehydes and ketones. The peptide bond’s reso-
`nance energy has its maximum value, ~85 kJ - mol”‘, when
`the peptide group is planar because its 7:-bonding overlap is
`maximized in this conformation. This overlap, and thus the
`resonance energy, falls to zero as the peptide bond is twisted
`to 90° out of planarity, thereby accounting for the planar
`peptide group’s rigidity.
`Peptide groups, with few exceptions, assume the trans
`conformation: that in which successive C“ atoms are on
`opposite sides of the peptide bond joining them (Fig. 7-1).
`This is partly a result of steric interference which causes the
`cis conformation (Fig. 7-2) to be ~8 kJ -mol“ less stable
`than the trans conformation (this energy difference is
`somewhat less in peptide bonds followed by a Pro residue;
`indeed, ~10% of the Pro residues in proteins follow a cis
`peptide bond, whereas cis peptides are otherwise extremely
`rare).
`
`Polypeptide Backbone Conformations May Be Described
`by Their Torsion Angles
`The above considerations are important because they in-
`dicate that the backbone ofa protein is a linked sequence of
`rigid planar peptide groups (Fig. 7-3). We can therefore
`specify a polypeptide’s backbone conformation by the tor-
`sion angles (rotation angles or dihedral angles) about the
`
`Main chain ,
`
`
`
`
`
`cis-Peptide group
`
`FIGURE 7-2. The cis-peptide group.
`
`J-.-~.-_
`
`
`
`|1l_I!:Il.I_II!fi-;.,m-.._.._.
`
`Ca—N bond (<75) and the C,,—C bond (w) of each of its
`amino acid residues. These angles, (1) and I//, are both de-
`fined as 180° when the polypeptide chain is in its planar,
`fully extended (all-trans) conformation and increase for a
`
`FIGURE 7-3. A polypeptide chain in its fully extended conformation showing the planarity of
`each of its peptide groups. [Figure copyrighted © by Irving Geis.]
`
`Page 4
`
`

`
`
`
`
`Section 7-]. Secondary Structure
`
`143
`
`i
`
`Amide plane
`
`
`H
`
`H
`
`H
`
`H
`
`H
`
`H
`
`H H
`
`H
`H
`
`.
`
`H
`H
`
`(a) Staggered
`
`(b)
`
`Eclipsed
`
`
`
`FIGURE 7-5. Newman projections indicating the (a)
`staggered conformation and (b) eclipsed conformation of ethane.
`
`
`
`
`
`rangement because opposing hydrogen atoms are maxi-
`mally separated in this position. In contrast, the eclipsed
`conformation (Fig. 7-5b), which is characterized by a tor-
`sion angle of 0°, is least stable because in this arrangement
`
`the hydrogen atoms are closest to each other. The energy
`
`difference between the staggered and eclipsed conforma-
`
`tions in ethane is ~12 kJ - mol‘1, a quantity that represents
`
`an energy barrier to free rotation about the C—C single
`
`bond. Substituents other than hydrogen exhibit greater
`steric interference; that is, they increase the size of this en-
`ergy barrier due to their greater bulk. Indeed, with large
`Substituents, some conformations may be sterically for-
`bidden.
`
`
`
`
`
`Amide plane
`
`
`
`
`-u—¢--L».-——IJ4:
`
`
`
`FIGURE 7-4. A portion of a polypeptide chain indicating the
`torsional degrees of freedom of each peptide unit. The only
`reasonably free movements are rotations about the Ca—N bond
`((1)) and the C,,—C bond (1//). The torsion angles are both 180'‘
`in the conformation shown and increase, as is indicated, in a
`clockwise manner when viewed from C0,. [Figure copyrighted ©
`by Irving Geis.]
`
`Allowed Conformations of Polypeptides Are Indicated by
`the Ramachandran Diagram
`The sterically allowed values of d) and a// can be deter-
`mined by calculating the distances between the atoms of a
`tripeptide at all values of <1) and l// for the central peptide
`unit. Sterically forbidden conformations, such as the one
`shown in Fig. 7-6, are those in which any nonbonding inter-
`atomic distance is less than its corresponding van der Waals
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`clockwise rotation when viewed from C“ (Fig. 7-4). (In ear-
`lier literature, this fully extended conformation was defined
`as having gb = y/ = 0°. These values can be made to corre-
`spond to the present convention by subtracting 180° from
`both angles.)
`There are several steric constraints on the torsion angles,
`(I) and 1//, of a polypeptide backbone that limit its conforrna-
`tional range. The electronic structure of a single (0) bond,
`such as a C—C bond, is cylindrically symmetrical about its
`bond axis so that we might expect such a bond to exhibit
`free rotation. Ifthis were the case, then in ethane, for exam-
`ple, all torsion angles about the C-C bond would be
`equally likely. Yet, certain conformations in ethane are fa-
`vored because of repulsions between electrons in the C—H
`bonds. The staggered conformation (Fig. 7-5a), which
`occurs at a 180° torsion angle, is ethane’s most stable ar-
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`
`FIGURE 7-6. Steric interference between the carbonyl oxygen
`and the amide hydrogen on adjacent residues prevents the
`occurrence of the conformation (15 = —60°, l// = 30°. [Figure
`copyrighted © by Irving Geis.]
`
`
`
`Page 5
`
`

`
`
`
`144 Chapter 7. Three-Dimensional Structures OfProteins
`
`
`
`180
`
`90
`
`3
`
`‘I
`
`n
`
`1
`-1
`
`-I
`_l
`
`7
`
`-1
`
` 1'
`
`_I
`
`-90 -
`
`‘I
`
`-1
`
`I
`
`T
`
`'
`
`'
`
`1
`-1
`
`-1
`-1
`
`"
`
`-
`
`_
`
`480
`-180
`
`-90
`
`l_|__L
`o
`¢(deE)
`
`.n.
`90
`
`180
`
`FIGURE 7-7. A Ramachandran diagram (named in honor of
`its inventor, G.N. Ramachandran) shows the sterically allowed
`4» and VI angles for poly-L-alanine. The diagram was calculated
`using the van der Waals distances in Table 7-1. Regions of
`“normally allowed” (1) and VI angles are shaded in blue, whereas
`green-shaded regions correspond to conformations having
`“outer limit” van der Waals distances. The conformation angles,
`(I) and I//, of several secondary structures are indicated below:
`
`Secondary Structure
`
`4) (deg) w (deg)
`
`Right-handed oz helix (at)
`Parallel B pleated sheet (TT)
`Antiparallel B pleated sheet (T l)
`Right-handed 3,0 helix (3)
`Right-handed 1t helix (7t)
`2.27 ribbon (2)
`Left-handed polyglycine II and poly-L-proline
`II helices (II)
`Collagen (C)
`Left-handed oz helix (a,_)
`
`— 57
`- 1 19
`- 139
`— 49
`— 57
`—78
`-79
`
`— 5 1
`57
`
`-47
`1 13
`135
`— 26
`- 70
`59
`150
`
`153
`47
`
`[After Flory, P.J., Statistical Mechanics ofChain Molecules, p. 253,
`Interscience (1969); and IUPAC-IUB Commission on Biochemical
`Nomenclature, Biochemistry 9, 3475 (1970).]
`
`distance. Such information is summarized in a conforma-
`tion map or Ramachandran diagram (Fig. 7-7).
`Figure 7-7 indicates that most areas of the Ramachan-
`dran diagram (most combinations of (I) and VI) are confor-
`mationally inaccessible to a polypeptide chain. The partic-
`ular regions of the Ramachandran diagram that represent
`allowed conformations depend on the van der Waals radii
`chosen to calculate it. But with any realistic set of values,
`such as that in Table 7-1, only three small regions of the
`conformational map are physically accessible to a polypep-
`tide chain. Nevertheless, as we shall see, all ofthe common
`types of regular secondary structures found in proteins fall
`within allowed regions of the Ramachandran diagram. In-
`
`180hlL_Il__Ll_l
`-180
`-90
`o
`90
`lllldegl
`FIGURE 7-8. The conformation angle distribution of all
`residues but Gly and Pro in 12 precisely determined high-
`resolution X-ray structures with a superimposed Ramachandran
`diagram. [After Richardson, J.S. and Richardson, D.C., in
`Fasman, G.D. (Ed.), Prediction ofProtein Structure and the
`Principles ofProtein Conformation, p. 6, Plenum Press (1989).]
`
`180
`
`TABLE 7-1. VAN DER WAALS DISTANCES FOR '
`INTERATOMIC CONTACTS
`9
`
`Contact Type
`
`Normally Allowed (A)
`
`Outer Limit (A)
`
`-
`
`1.9
`2.0
`- H
`H -
`2.2
`2.4
`- O
`H -
`2.2
`2.4
`- N
`H -
`2.2
`2.4
`- C
`H -
`2.6
`2.7
`- O
`-
`O -
`2.6
`2.7
`- N
`-
`0 -
`2.7
`2.8
`- C
`-
`O -
`2.6
`2.7
`- N
`-
`N -
`2.8
`2.9
`N -- - C
`2.9
`3.0
`C -
`-
`- C
`3.0
`3.2
`C -
`-
`- CH2
`
`
`3.2CH, -- - CH2 3.0
`Source: Ramachandra, G.N. and Sasisekharan, V., Adv. Protein Chem.23,
`326 (1968).
`
`deed, the observed conformational angles of most non-Gly
`residues in proteins whose X-ray structures have been de-
`termined lie in these allowed regions (Fig. 7-8).
`Most points that fall in forbidden regions of Fig. 7-8 lie
`between its two fully allowed areas near VI = 0. However,
`these “forbidden” conformations, which arise from the col-
`lision of successive amide groups, are allowed if twists of
`only a few degrees about the peptide bond are permitted.
`This is not unreasonable since the peptide bond offers little
`resistance to small deformations from planarity.
`
`Page 6
`
`Page 6
`
`

`
`Gly, the only residue without a C, atom, is much less
`sterically hindered than the other amino acid residues. This
`is clearly apparent in comparing the Ramachandran dia-
`
`Section 7-1. Secondary Structure
`
`145
`
`gram for Gly in a polypeptide chain (Fig. 7-9) with that of
`other residues (Fig. 7-7). In fact, Gly often occupies posi-
`tions where a polypeptide backbone makes a sharp turn
`which, with any other residue, would be subject to steric
`interference.
`
`Figure 7-7 was calculated for three consecutive Ala resi-
`dues. Similar plots for larger residues that are unbranched
`
`at C5, such as Phe, are nearly identical. In Ramachandran
`diagrams of residues that are branched at C5, such as Thr,
`the allowed regions are somewhat smaller than for Ala. The
`cyclic side chain of Pro limits its (1) to the range -60“ i
`25 °, making it, not surprisingly, the most conformationally
`restricted amino acid residue. The conformations of resi-
`
`dues in chains longer than tripeptides are even more re-
`stricted than the Ramachandran diagram indicates be-
`cause, for instance, a polypeptide chain cannot assume a
`conformation in which it passes through itself. We shall see,
`however, that despite the great restrictions that peptide
`bond planarity and side chain bulk place on the conforma-
`tions of a polypeptide chain, every unique primary struc-
`ture has a correspondingly unique three-dimensional struc-
`ture.
`
`-180
`-180
`
`-90
`
`O
`
`90
`
`180
`
`B. Helical Structures
`
`¢.'(de8)
`FIGURE 7-9. The Ramachandran diagram of Gly residues in
`a polypeptide chain. “Normally allowed” regions are shaded
`in blue, whereas green-shaded regions correspond to “outer
`limit” atomic distances. Gly residues have far greater
`conformational freedom than do other (bulkier) amino acid
`residues as the comparison of this figure with Fig. 7-7 indicates.
`[After Ramachandran, G.N. and Sasisekharan, V., Adv. Protein
`Chem. 23, 332 (I968).]
`
`Helices are the most striking elements of protein 2° struc-
`ture. If a polypeptide chain is twisted by the same amount
`about each of its C, atoms, it assumes a helical conforma-
`tion. As an alternative to specifying its ()5 and z// angles, a
`helix may be characterized by the number, n, of peptide
`units per helical turn, and its pitch, p, the distance the helix
`rises along its axis per turn. Several examples of helices are
`diagrammed in Fig. 7-10. Note that a helix has chirality;
`
`n=5
`
`n=4
`
`n.=3
`
`n=2
`
`n=-3
`
`\
`
`\ ©
`
`llzvwa
`65i5
`
`FIGURE 7-10. Examples of helices indicating the definitions
`of the helical pitch, p, the number of repeating units per turn,
`n, and the helical rise per repeating unit, d = p/n. Right- and
`left-handed helices are defined, respectively, as having positive
`
`and negative values of n. For n = 2, the helix degenerates to a
`nonchiral ribbon. For 1) = 0, the helix degenerates to a closed
`ring. [Figure copyrighted © by Irving Geis.]
`
`Page 7
`
`

`
`l l2 44 t
`
`i
`£-
`:5.
`
`3.
`
`A polypeptide helix must, of course, have conformation
`angles that fall within the allowed regions of the Rama-
`chandran diagram. As we have seen, this greatly limits the
`possibilities. Furthermore, ifa particular conformation is to
`have more than a transient existence, it must be more than
`
`just al1owed—it must be stabilized. The “glue” that holds
`polypeptide helices and other 2° structures together is, in
`part, hydrogen bonds.
`
`The at Helix
`
`Only one helical polypeptide conformation has simulta-
`neously allowed conformation angles and a favorable hy-
`drogen bonding pattern: the C! helix (Fig. 7-1 I), a particu-
`larly rigid arrangement of the polypeptide chain.
`Its
`discovery through model building, by Pauling in 1951,
`ranks as one of the landmarks of structural biochemistry.
`For a polypeptide made from L-a-amino acid residues,
`the oz helix is right handed with torsion angles (1) = - 57°
`and 1/1 = -47 °, n = 3.6 residues per turn, and a pitch of 5.4
`A. (An oz helix of D-oz-amino acid residues is the mirror
`image of that made from L-amino acid residues: It is left
`handed with conformation angles (12 = + 57° and 1/1 =
`+47° but with the same values of n and p.)
`Figure 7-11 indicates that the hydrogen bonds of an a
`helix are arranged such that the peptide C=O bond of the
`nth residue points along the helix towards the peptide
`N—H group of the (n + 4)th residue. This results in a
`strong hydrogen bond that has the nearly optimum
`N -
`-
`- 0 distance of 2.8 A. In addition, the core of the a
`helix is tightly packed; that is, its atoms are in van der Waals
`contact across the helix, thereby maximizing their associa-
`tion energies (Section 7-4A). The R groups, whose posi-
`tions, as we saw, are not fully dealt with by the Ramachan-
`dran diagram, all project backward (downward in Fig. 7-1 1)
`and outward from the helix so as to avoid steric interference
`
`with the polypeptide backbone and with each other. Such
`an arrangement can also be seen in Fig. 7-12. Indeed, a
`major reason why the left-handed ct helix has never been
`observed (its helical parameters are but mildly forbidden;
`Fig. 7-7) is that its side chains contact its polypeptide back-
`bone too closely. Note, however, that 1 to 2% ofthe individ-
`ual non-Gly residues in proteins assume this conformation
`(Fig. 7-8).
`The at helix is a common secondary structural element of
`both fibrous and globular proteins. In globular proteins, at
`helices have an average span of ~ 12 residues, which corre-
`sponds to over three helical turns and a length of 18 A.
`However, oz helices with as many as 53 residues have been
`found.
`
`Other Polypeptide Helices
`Figure 7-13 indicates how hydrogen bonded polypeptide
`helices may be constructed. The first two, the 2.27 ribbon
`and the 31., helix, are described by the notation n,,, where n,
`as before, is the number of residues per helical turn, and m
`is the number of atoms, including H, in the ring that is
`
`Page 8
`
`146 Chapter 7. Three-Dimensional Structures OfProteins
`
`that is, it may be either right handed or left handed (a right-
`handed helix turns in the direction that the fingers of a right
`hand curl when its thumb points along the helix axis in the
`direction that the helix rises). In proteins, moreover, n need
`not be an integer and, in fact, rarely is.
`

`IRVING
`6215
`
`FIGURE 7-11. The right-handed oz helix. Hydrogen bonds
`between the N—H groups and the C=O groups that are four
`residues back along the polypeptide chain are indicated by
`dashed lines. [Figure copyrighted © by Irving Geis.]
`
`Page 8
`
`

`
`Section 7-]. Secondary Structure
`
`147
`
`FIGURE 7-12. A stereo, space-filling representation of an oz
`helical segment of sperm whale myoglobin (its E helix) as
`determined by X-ray crystal structure analysis. In the main
`chain, carbon atoms are green, nitrogen atoms are blue, oxygen
`
`atoms are red, and hydrogen atoms are white. The side chains
`are yellow. Instructions for viewing stereo diagrams are given in
`the appendix to this chapter.
`
` v
`
`6535
`
`l
`t
`_ _ _ _ _ _ __ /_ ——____.____/ ______._......_../
`/310 helix
`I on helix
`2.27 ribbon
`
`1
`//1: helix
`
`FIGURE 7-13. The hydrogen bonding pattern of several polypeptide helices. In the cases shown,
`the polypeptide chain curls around such that the C=O group on residue n forms a hydrogen
`bond with the N—H groups on residues (n + 2),
`(n + 3), (n + 4), or (n + 5).
`[Figure copyrighted © by Irving Geis.]
`
`
`
`Page 9
`
`Page 9
`
`

`
`
`
`148 Chapter 7. Three-Dimensional Structures OfProteins
`
`\_.
`
`on helix
`310 helix
`FIGURE 7-14. Two polypeptide helices that occasionally
`occur in proteins compared with the commonly occurring oz
`helix. (a) The 3,0 helix is characterized by 3.0 peptide units per
`turn and a pitch of 6.0 A, which makes it thinner and more
`elongated than the at helix. (b) The oz helix has 3.6 peptide units
`
`1: helix ¢
`
`per turn and a pitch of 5.4 A (also see Fig. 7-11). (c) The n helix,
`with 4.4 peptide units per turn and a pitch of 5.2 A, is wider
`and shorter than the oz helix. The peptide planes are indicated.
`[Figure copyrighted © by Irving Geis.]
`
`closed by the hydrogen bond. With this notation, an an helix
`is a 3.613 helix.
`The right-handed 3 ,0 helix (Fig. 7- 14a), which has a pitch
`of 6.0 A, is thinner and rises more steeply than does the an
`helix (Fig. 7-14b). Its torsion angles place it in a mildly
`forbidden zone ofthe Ramachandran diagram that is rather
`near the position of the oz helix (Fig. 7-7) and its R groups
`experience some steric interference. This explains why the
`3 ,0 helix is only occasionally observed in proteins, and then
`mostly in short segments that are frequently distorted from
`the ideal 3,0 conformation. The 3,0 helix most often occurs
`as a single-tum transition between one end ofan oz helix and
`the adjoining portion of a polypeptide chain.
`The 1: helix (4.416 helix), which also has a mildly forbid-
`den conformation (Fig. 7-7), has only rarely been observed
`and then only as segments of longer helices. This is proba-
`bly because its comparatively wide and flat conformation
`(Fig. 7- 14c) results in an axial hole that is too small to admit
`water molecules but yet too wide to allow van der Waals
`associations across the helix axis; this greatly reduces its
`
`stability relative to more closely packed conformations.
`The 2.27 ribbon, which as Fig. 7-7 indicates, has strongly
`forbidden conformation angles, has never been observed.
`Certain synthetic homopolypeptides assume conforma-
`tions that are models for helices in particular proteins. Po-
`lyproline is unable to assume any common secondary
`structure due to the conformational constraints imposed by
`its cyclic pyrrolidine side chains. Furthermore, the lack ofa
`hydrogen substituent on its backbone nitrogen precludes
`any polyproline conformation from being knit together by
`hydrogen bonding. Nevertheless, under the proper condi-
`tions, polyproline precipitates from solution as a left-
`handed helix of all-trans peptides that has 3.0 residues per
`helical turn and a pitch of 9.4 A (Fig. 7-15). This rather
`extended conformation permits the Pro side chains to avoid
`each other. Curiously, polyglycine, the least conformation-
`ally constrained polypeptide, precipitates from solution as a
`helix whose parameters are essentially identical to those of
`polyproline, the most conformationally constrained poly-
`peptide (although the polyglycine helix may be either right
`
`Page 10
`
`Page 10
`
`

`
`(a) Antiparallel
`
`Section 7- 1. “Secondary Structure
`
`149
`
`
`
`Parallel
`
`(13)
`
`FIGURE 7-15. The polyproline II
`helix. Polyglycine forms a nearly
`identical helix (polyglycine II). [Figure
`copyrighted © by Irving Geis.]
`
`FIGURE 7-16. The hydrogen bonding associations in B pleated sheets. Side chains are
`omitted for clarity. (a) The antiparallel /3 pleated sheet. (b) The parallel fl pleated sheet.
`[Figure copyrighted © by Irving Geis.]
`
`, C4——N\‘.
`
`I
`
`or left handed because Gly is nonchiral). The structures of
`the polyglycine and polyproline helices are of biological
`significance because they form the basic structural motif of
`collagen, a structural protein that contains a remarkably
`high proportion of both Gly and Pro (Section 7-2C).
`
`bonding capacity of the polypeptide backbone. In /Ipleated
`sheets, however, hydrogen bonding occurs between neigh-
`boring polypeptide chains rather than within one as in oz
`helices.
`
`/3 Pleated sheets come in two varieties:
`
`C. Beta Structures
`
`In 1951, the year that they proposed the at helix, Pauling
`and Corey also postulated the existence of a different poly-
`peptide secondary structure, the ,8 pleated sheet. As with the
`a helix, the B pleated sheet’s conformation has repeating ti)
`and y/ angles that fall in the allowed region of the Rama-
`chandran diagram (Fig. 7-7) and utilizes the full hydrogen
`
`
`
`1. The antiparallel ,6 pleated sheet, in which neighboring
`hydrogen bonded polypeptide chains run in opposite
`directions (Fig. 7- 1 6a).
`
`2. The parallel fl pleated sheet, in which the hydrogen
`bonded chains extend in the same direction (Fig. 7- 16b).
`
`The conformations in which these [3 structures are opti-
`mally hydrogen bonded vary somewhat from that of a fully
`extended polypeptide (cl) = my = i 180°) as indicated in
`Fig. 7-7. They therefore have a rippled or pleated edge-on
`
`Page 11
`
`Page 11
`
`

`
`150 Chapter 7. Three-Dimensional Structures 0fProteins
`
`appearance (Fig. 7-17), which accounts for the appellation
`“pleated sheet.” In this conformation, successive side
`chains of a polypeptide chain extend to opposite sides ofthe
`pleated sheet with a two-residue repeat distance of 7.0 A.
`/3 Sheets are common structural motifs in proteins. In
`globular proteins, they consist of from 2 to as many as 15
`polypeptide strands, the average being 6 strands, which
`have an aggregate width of ~25 A. The polypeptide chains
`in a /3 sheet are known to be up to 15 residues long, with the
`average being 6 residues that have a length of ~21 A. A
`6-stranded antiparallel /3 sheet, for example, occurs in the
`jack bean protein concanavalin A (Fig. 7-18).
`Parallel ,8 sheets of less than five strands are rare. This
`observation suggests that parallel fl sheets are less stable
`than antiparallel /3 sheets, possibly because the hydrogen
`bonds ofparallel sheets are distorted in comparison to those
`of the antiparallel sheets (Fig. 7-16). Mixed parallel-
`antiparallel ,8 sheets are common but occur with only ~40%
`of the frequency that would be expected for the random
`mixing of strand directions.
`The B pleated sheets in globular proteins invariably ex-
`hibit a pronounced right-handed twist when viewed along
`
`FIGURE 7-17. A two-stranded B antiparallel pleated sheet
`drawn so as to emphasize its pleated appearance. Dashed lines
`indicate hydrogen bonds. Note that the R groups (purple balls)
`on each polypeptide chain alternately extend to opposite sides
`of the sheet and that they are in register on adjacent chains.
`
`[Figure copyrighted © by Irving Geis.]
`to this chapter.
`
`FIGURE 7-18. A stereo, space-filling
`representation of the six-stranded
`antiparallel fl pleated sheet in jack
`bean concanavalin A as determined
`by X-ray crystal structure analysis.
`In the main chain, carbon atoms are
`green, nitrogen atoms are blue,
`oxygen atoms are red, and hydrogen
`atoms are white. R groups are
`represented by large purple balls.
`Instructions for viewing stereo
`drawings are given in the appendix
`
`Page 12
`
`Page 12
`
`

`
`
`, S
`
`ection 7-]. Secondary Structure
`
`151
`
`FIGURE 7-19. Polypeptide chain folding in proteins
`illustrating the right-handed twist of B sheets. Here the
`polypeptide backbones are represented by ribbons with or helices
`shown as coils and the strands of B sheets indicated by arrows
`pointing towards the C-terminus. Side chains are not shown.
`((1) Bovine carboxypeptidase A, a 307-residue protein, contains
`an eight-stranded mixed B sheet that forms a saddle-shaped
`curved surface with a right-handed twist. (b) Triose phosphate
`isomerase, a 247-residue chicken muscle enzyme, contains an
`eight-stranded parallel B sheet that forms a cylindrical structure
`known as a B barrel [here viewed from the top (left) and from
`the side (right)]. Note that the crossover connections between
`successive strands of the B barrel, which each consist
`predominantly of an a helix, are outside the B barrel and have a
`right-handed helical sense. [After drawings by Jane Richardson,
`Duke University.]
`
`(a)
`
`
`
`
`
`
`
`their polypeptide strands (e.g., Fig. 7-19). Such twisted B
`sheets are important architectural features of globular pro-
`teins since B sheets often form their central cores (Fig. 7- 19).
`Conformational energy calculations indicate that a B
`sheet’s right-handed twist is a consequence of nonbonded
`interactions between the chiral L—amino acid residues in the
`
`sheet’s extended polypeptide chains. These interactions
`tend to give the polypeptide chains a slight right-handed
`helical twist (Fig. 7-19) which distorts and hence weakens
`the B sheet’s interchain hydrogen bonds. A particular B
`sheet’s geometry is thus the result of a compromise between
`optimizing the conformational energies of its polypeptide
`chains and preserving its hydrogen bonds.
`The topology (connectivity) of the polypeptide strands in
`a B sheet can be quite complex; the connecting links ofthese
`assemblies often consist of long runs of polypeptide chain
`which frequently contain helices (e.g., Fig. 7-19). The link
`connecting two consecutive antiparallel strands is topologi-
`cally equivalent to a simple hairpin turn (Fig. 7-20a). How-
`ever, tandem parallel strands must be linked by a cross-over
`connection that is out of the plane of the B sheet. Such
`crossover connections almost always have a right-handed
`
`
`FIGURE 7-20. The connections between adjacent
`polypeptide strands in B pleated sheets: (a) The hairpin
`connection between antiparallel strands is topologically in the
`plane of the sheet. (b) A right-handed crossover connection
`between successive strands of a parallel B sheet. Nearly all such
`crossover connections in proteins have this chirality (see, e.g.,
`Fig. 7-1911). (C) A left-handed cross-over connection between
`parallel B sheet strands. Connections with this chirality are rare.
`[After Richardson, J .S., Adv. Protein Chem. 34, 290, 295 (1981).]
`
`
`
`Page 13
`
`
`Page 13
`
`

`
`152 Chapter 7. Three-Dimensional Structures OfProteins
`
` FIGURE 7-21. A possible folding scheme illustrating how
`
`right-handed polypeptide chain twisting favors the formation of
`right-handed cross-over connections between successive strands
`of a parallel B sheet.
`
`helical sense (Fig. 7-20b), which is thought to better fit the B
`sheets’ inherent right-handed twist (Fig. 7-21).
`
`D. Nonrepetitive Structures
`
`Regular secondary structures—/telices and B sheets-
`comprise around half of the average globular protein. The
`protein’s remaining polypeptide segments are said to have a
`coil or loop conformation. That is not to say, however, that
`these nonrepetitive secondary structures are any less or-
`dered than are helices or B sheets; they are simply irregular
`and hence more difficult to describe. You should therefore
`not confuse the term coil conformation with the term ran-
`
`dom coil, which refers to the totally disordered and rapidly
`fluctuating set of conformations assumed by denatured
`proteins and other polymers in solution.
`Globular proteins consist
`largely of approximately
`straight runs of secondary structure joined by stretches of
`
`polypeptide that abruptly change direction. Such reverse
`turns or B bends (so named because they often connect
`successive strands of antiparallel B sheets) almost always
`occur at protein surfaces; indeed, they partially define these
`surfaces. Most reverse turns involve four successive amino
`
`acid residues more or less arranged in one oftwo ways, Type
`I and Type II, that differ by a 180° flip of the peptide unit
`linking residues 2 and 3 (Fig. 7-22). Both types of B bends
`are stabilized by a hydrogen bond although deviations from
`these ideal conformations often disrupt this hydrogen
`bond. Type I B bends may be considered to be distorted
`sections of 3,0 helix. In Type II B bends, the oxygen atom of
`residue 2 crowds the C,, atom ofresidue 3, which is therefore
`usually Gly. Residue 2 of either type ofB bend is often Pro
`since it can facilely assume the required conformation.
`Almost all proteins of >60 residues contain one or more
`loops of 6 to 16 residues that are not components of hel

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket