throbber
Page 1
`
`NPS EX. 2028
`CFAD v. NPS
`IPR2015-01093
`
`

`
`Principles of Biochemistry
`
`Second Edition
`
`Albert L. Lehninger, David L. Nelson, and Michael M. Cox
`
`Copyright (TC) 1993, 1982 by Worth Publishers, Inc.
`
`All rights reserved
`
`Printed in the United States of America
`
`Library of Congress Catalog Card No. 91-67492
`ISBN: 0—87901-500-4
`
`Printing: 5 4 3 2
`
`Year: 97 96 95 94 93
`
`Development Editor: Valerie Neal
`
`Design: Malcolm Grear Designers
`Art Director: George Touloumes
`
`Project Editor: Elizabeth Geller
`Production Supervisor: Sarah Segal
`
`Layout: Patricia Lawson
`Picture Editor: Stuart Kenter
`
`Illustration Design: Susan Tilberry
`
`Illustrators: Susan Tilberry, Alan Landau, and Joan Waites
`Computer Art: Laura Pardi Duprey and York Graphic Services
`
`Composition: York Graphic Services
`Printing and binding: R.R. Donnelley and Sons
`Cover: The active site of the prnteolytic ensynie chynit-trypsin, show-
`ing the substrate (blue and pllrplel and the amino acid residues (red
`and orange) critical to catalysis. Determination of the rletailetl
`reaction inechsnism of this enzyme [dt-ascribed on pp. 223—'.’.2G}
`helped to establish the general principles of en '.?._YT!1e action.
`
`Froritispicce: A view of tobacco rihuiose-1,5~bisphuspl1ste carboxyluse
`(ruhisco 1. This en:r.yme is central to photosynthetic carbon dioxide
`fixation; it is the most abundant euzyine in the biosphere. Different
`subunits are shown in blues and gruys. Importimt active site residtles
`are shown in red. Sullhtes bound at the active site (an artifact of the
`
`crystallization procedure) are shown in yellow.
`
`Cover, frontispiece, and part opening images produced by Alisa Zapp (see
`Molecular Modeling credits, p. IC-4) and enhanced by Academy Arts.
`
`Illustration credits begin on p.
`copyright page.
`
`IC—1 and constitute a continuation of the
`
`Worth Publishers
`
`33 Irving Place
`
`New York, NY 10003
`
`Page 2
`
`Page 2
`
`

`
`CHAPTER
`
`-lino Acids and Peptides
`
`' are the most abundant macromolecules in living cells, occur-
`all cells and all parts of cells. Proteins also occur in great vari-
`'ugg'nd.=s of different kinds may be found in a single cell. More-
`glproteins exhibit great diversity in their biological function. Their
`_
`" role is made evident by the fact that proteins are the most
`is final products of the information pathways discussed in
`of this book. In a sense, they are the molecular instruments
`which genetic information is expressed. It is appropriate to
`he study of biological macromolecules with the proteins, whose
`erives from the Greek protos, meaning “first” or “foremost.”
`latively simple monomeric subunits provide the key to the struc-
`of the thousands of different proteins. All proteins, whether from
`
`5
`
`flare constructed from the same ubiquitous set of 20 amino acids,
`Hiéiitly linked in characteristic linear sequences. Because each of
`' mino acids has a distinctive side chain that determines its
`
`_
`
`"
`
`a1 properties, this group of 20 precursor molecules may be re-
`as the alphabet in which the language of protein structure is
`
`Proteins are chains of amino acids, each joined to its neighbor by a
`fie" type of covalent bond. What is most remarkable is that cells
`produce proteins that have strikingly different properties and ac-
`: .by joining the same 20 amino acids in many different combina-
`and sequences. From these building blocks different organisms
`kc such widely diverse products as enzymes, hormones, anti-
`5 the lens protein of the eye, feathers, spider webs, rhinoceros
`(_F1g- 5-1), milk proteins, antibiotics, mushroom poisons, and a
`, ‘‘..d Of Other substances having distinct biological activities.
`F1'°l3e1_I1 structure and function is the topic for the next four chap-
`I _th1s chapter we begin with a description of amino acids and the
`st bonds that link them together in peptides and proteins.‘
`
`Figure 5-1 The protein keratin is formed by all
`vertebrates. It is the chief structural component of
`hair, scales, horn, wool, nails, and feathers. The
`black rhinoceros is nearing extinction in the wild
`because of the myths prevalent in some parts of the
`world that a powder derived from its horn has aph-
`rodisiac properties. In reality, the chemical proper-
`ties are no different from those of powdered bovine
`hooves or human fingernails.
`
`Page 3
`
`

`
`
`
`112
`
`Part II Structure and Catalysis
`
`Amino Acids
`
`Proteins can be reduced to their constituent amino acids by a variet .;-
`methods, and the earliest studies of proteins naturally focused on .-'|
`free amino acids derived from them. The first amino acid to be disc"
`ered in proteins was asparagine, in 1806. The last of the 20 to be 1"on 3-]
`threonine, was not identified until 1938. All the amino acids have t 5|
`ial or common names, in some cases derived from the source F"
`which they were first isolated. Asparagine was first found in EI..'=;pa
`gus, as one might guess; glutamate was found in wheat gluten; ty
`sine was first isolated from cheese (thus its name is derived from .i
`Greek tyros, “cheese”); and glycine (Greek glykos, “sweet”) was I
`named because of its sweet taste.
`
`'
`
`Amino Acids Have Common Structural Features
`
`All of the 20 amino acids found in proteins have a carboxyl group .-'
`an amino group bonded to the same carbon atom (the oz carbon) {_
`it
`5-2). They differ from each other in their side chains, or R grou
`which vary in structure, size, and electric charge, and influence ;
`solubility of amino acids in water. When the R group contains a.
`tional carbons in a chain, they are designated ,8, 7, 5, 6, etc., proceed’
`out from the oz carbon. The 20 amino acids of proteins are often 1'oI"o
`to as the standard, primary, or normal amino acids, to distingu
`them from amino acids within proteins that are modified after ,
`proteins are synthesized, and from many other kinds of amino ucl
`present in living organisms but not in proteins. The standard am
`acids have been assigned three-letter abbreviations and one-let
`symbols (Table 5-1), which are used as shorthand to indicate the so I
`position and sequence of amino acids in proteins.
`1
`We note in Figure 5-2 that for all the standard amino acids exc
`one (glycine) the oz carbon is asymmetric, bonded to four different as
`-j
`stituent groups: a carboxyl group, an amino group, an R group, an;
`hydrogen atom. The oz-carbon atom is thus a3'chi1:a‘l:(;'ér'1'ter.(see *-
`:. ro-
`3—9). Because of the tetrahedral arrangement of the bonding or.
`around the oz-carbon atom of amino acids, the four different substi .
`ent groups can occupy two different arrangements in space, which if
`nonsuperimposable mirror images of each other (Fig. 5-3). These _j
`forms are called enantiomers or stereoisomers (see Fig. 3-9). "
`molecules with a chiral center are also optically active—i.e., Ll
`can rotate plane-polarized light, with the direction of the rotation I
`fering for different stereoisomers.
`
`Figure 5-3 (a) The two stereoisomers of alanine.
`L- and D-alanine are nonsuperimposable mirror
`images of each other. (b, c) Two different conven-
`tions for showing the configurations in space of
`stereoisomers. In perspective formulas (b) the
`wedge-shaped bonds project out of the plane of the
`paper, the dashed bonds behind it. In projection
`formulas (c) the horizontal bonds are assumed to
`project out of the plane of the paper, the Vertical
`bonds behind. However, projection formulas are
`often used casually without reference to stereo-
`chemical configuration.
`
`Page 4
`
`C00’
`
`*
`H3N— |3—H
`
`3 H
`
`Glycine
`
`C00’
`+
`3:
`H3N— —H
`
`Amino acid
`
`Figure 5-2 General structure of the amino acids
`found in proteins. With the exception of the nature
`of the R group, this structure is common to all the
`oz-amino acids. (Proline, because it is an imino acid,
`is an exceptional component of proteins.) The a car-
`bon is shown in blue. R (in red) represents the R
`group or side chain, which is different in each
`amino acid. In all amino acids except glycine
`(shown for comparison) the oz-carbon atom has four
`different substituent groups.
`
`
`
`L-Alanine
`
`
`
`Q00‘
`H313-—¢<H
`CH3
`L-Alanine
`
`(|300’
`H;.IiI—-U—I[
`:_L.H_...
`L-Alanine
`
`(a)
`
`(b)
`
`(c)
`
`Q00"
`H»¢«-KIH3
`CH3
`D-Alanine
`
`C1100"
`H—(l:—1?IH3
`CH3
`D-Alanine
`
`
`
`Page 4
`
`

`
`Chapter 5 Amino Acids and Peptides
`
`properties and conventions associated with the standard amino acids
`-
`‘I
`
`PK1
`(—COOH)
`
`PK2
`(—NH§ )
`
`PKR
`(R group)
`
`pl
`
`Hydropathy
`index*
`
`Occurrence
`in Proteins (%)T
`
`?H_l"J
`T-I—-( ‘ ---()H
`
`FC
`
`H._.OH
`D-G-1-yceraldehyde
`
`(‘.200
`:
`4
`
`[El-—? —-N'[-I3
`CH3
`D-Alanine
`
`‘('7-HU
`
`Ho-44¢ -H
`3&1-IZOH
`L-Glycersltlehyde
`
`(_:0(J
`r1_.,1§-—g".-H
`
`L-Alanine
`
`Figure 5-4 Steric relationship of the stereoisomers
`of alanine to the absolute configuration of L- and
`D-glyceraldehyde. In these perspective formulas, the
`carbons are lined up vertically, with the chiral atom
`in the center. The carbons in these molecules are
`numbered beginning with the aldehyde or carboxyl
`carbons on the end, or 1 to 3 from top to bottom
`as shown. When presented in this way, the R group
`of the amino acid (in this case the methyl group
`of alanine) is always below the a carbon. L-Amino
`acids are those with the a-amino group on the left,
`and D-amino acids have the a-amino group on the
`right.
`
`Page 5
`
`i_.‘Dm_bi'uing liydi-up!-1uIJici'ty sml hydmphilicity; can be used to predict. which amino
`b'e'f'ou'ml in an aqueous uJwi1'umnen1'. E— values] and which will be found in a
`'_iu environment H values). sea em; in-2. Ftjcrnl Kyte. .1. & Doolittle. er. uses}
`I61. 157, 1Dl'i—1.'J2.
`
`occurrence in over 200 proteins. From Klapper, MH. (1977) Biochem. Biophys. Res.
`78, 1018—1024.
`
`.
`
`’
`
`'_
`
`_
`«._=
`
`lB"'cl"a1ssificatiou and naming of stereoisomers is based on the ab-
`I-Ijilnfiguration of the four substituents of the asymmetric car-
`For this purpose a reference compound has been chosen, to
`-.013. er optically active compounds are compared. This refer-
`Pflund is the 3-carbon sugar glyceraldehy-de (Fig. 5-4}, the
`fiflgar to have an asymmetric carbon atom. The naming of
`4 al3.1.0l'l8 of both simple sugars and amino acids is based on the
`' Wfifigllration of glyc'e.ra1del'1yde, as established by x-ray dif-
`"3_B'«alysis. Th_e stereoisomers of all chiral compounds having at,
`atétion related to that of L-glyceraldehyde are designated 1. (fo
`ry, derlvecl from two, meaning ‘‘left”), and the stereoisomer
`__ ET-:;:"D‘g1yw1'aldehyde are designated a {for dextrorotatory, dei.
`-_ dextro, meaning “right”l. T-he symbols 1. and D thus refer to
`Solute configuration of the four substituents around the chiral
`
`Page 5
`
`

`
`114
`
`Part Il Structure and Catalysis
`
`
`
`Proteins Contain L-Amino Acids
`
`Nearly all biological compounds with a chiral center occur naturally in
`only one stereoisomeric form, either D or L. _’.l._‘he=-amino‘ a-cids*in1:irotei-1-.1.
`moleculesare the'L“ stereoisomers. D-Amino acids have been found unly
`in small peptides of bacterial cell walls and in some peptide autibiot.i¢'3
`(see Fig. 5-19).
`It is remarkable that the amino acids of proteins are all L ste1*eo130_
`mers. As we noted in Chapter 3, when chiral compounds are formed by.
`ordinary chemical reactions, a racemic mixture of D and 1. isomers 1-E.
`sults. Whereas the L and D forms of chiral molecules are difficult for a.
`chemist to distinguish and isolate, they are as different as nighl; and
`day to a living system. The ability of cells to specifically synthesize the.
`L isomer of amino acids reflects one of many extraordinary properties
`of enzymes (Chapter 8). The stereospecificity of the reactions catalyzed
`by some enzymes is made possible by the asymmetry of their active
`sites. The characteristic three-dimensional structures of proteins
`(Chapter 7), which dictate their diverse biological activities, require
`that all their constituent amino acids be of one stereochemical series_
`
`Amino Acids Are Ionized in Aqueous Solutions
`
`Amino acids in aqueous solution are ionized and can act as acids or
`bases. Knowledge of the acid—base properties of amino acids is ex-
`tremely important in understanding the physical and biological prop-
`erties of proteins. Moreover, the technology of separating, identifying,
`and quantifying the different amino acids, which are necessary steps
`in determining the amino acid composition and sequence of protein‘
`molecules, is based largely on their characteristic acid-base behavior.
`Those a-amino acids having a single amino group and a single
`carboxyl group crystallize from neutral aqueous solutionsias fully ion-
`
`ized species known as. .'fGerman for “hybrid ions”), each
`having both a positive and a negative charge (Fig. 5-5). These ions are
`electrically neutral and remain stationary in an electric field. The di-
`polar nature of amino acids was first suggested by the observation that
`crystalline amino acids have melting points much higher than those of
`other organic molecules of similar size. The crystal lattice of amino"
`acids is held together by strong electrostatic forces between. positively-
`and negatively charged functional groups of neighboring molecules,
`resembling the stable ionic crystal lattice of NaCl (see Fig. 4-6).
`
`Amino Acids Can Be Classified by R Group
`
`An understanding of the chemical properties of the standard amino
`acids is central to an understanding of much of biochemistry. The topic
`can be simplified by grouping the amino acids into classes based on the
`properties of their R groups (Table 5—1], in particular, their polarity
`or tendency to interact with water at biological pH {near pH 7.0J. The
`polarity of the R groups varies widely, from totally nonpolar or hydro-
`phobic (water-insouble) to highly polar or hydrophilic (water-soluble].
`The structures of the 20 standard amino acids are shown in FigI1T9
`5-6, and many of their properties are listed in Table 5-1. There 31'“
`five main classes of amino acids, those whose R groups are: nonpolar
`and aliphatic; aromatic {generally nonpolar); polar but uncharged?
`negatively charged; and positively charged. Within each class there
`are gradations of polarity, size, and shape of the R groups.
`
`Page 6
`
`“F00”
`H2N—C—H
`l:
`Nonionic
`form
`
`“[00
`.
`H3N—(|?—H
`R
`Zwitterionic
`form
`
`Figure 5-5 Nonionic and zwitterionic forms of
`amino acids. Note the separation of the + and ~
`charges in the zwitterion, which makes it an elec-
`tric dipole. The nonionic form does not occur in sig-
`nificant amounts in aqueous solutions. The zwitter-
`ion predominates at neutral pH.
`
`Page 6
`
`

`
`Chapter 5 Amino Acids and Peptides
`
`Nonpolar, aliphatic R groups
`
`C00’
`I
`
`+
`
`H3N—(|]—H
`H
`
`C00”
`I
`
`+
`
`H3N—Cl)‘—H
`CH3
`
`Alanine
`
`C00 '
`l
`
`+
`
`H3N—i—H
`cfia C11,,
`Valine
`
`H3KI—
`
`G00"
`I
`
`—H
`
`CH2
`
`Aromatic R groups
`COO’
`I
`éflg
`
`+
`
`COO _
`+
`I
`H3N—(1—H
`&H2
`L
`
`as
`
`' H
`
`Phenylalanine
`
`Tyrosine
`
`Tryptophan
`
`Positively charged R groups
`
`C00"
`I
`+
`H3N—C—H
`
`C00"
`I
`+
`H3N—Cl)—H
`CH2
`(IZHZ
`is
`$11 +
`(|3=NH2
`NH2
`
`C00-
`+ J;
`—-H
`H3N
`(Ia;
`J3-NH\
`(I43/“H
`H
`
`Arginine
`
`Histidine
`
`Negatively charged R groups
`C00‘
`C00-
`+
`+
`I
`I
`I-I3N—(iJ—H
`H3N——(l)—H
`(|3H2
`(EH2
`COO’
`
`' O0‘
`
`Aspartate
`
`Glutamate
`
`Figure 5-6 The 20 standard amino acids of pro-
`teins. They are shown with their amino and car-
`boxyl groups ionized, as they would occur at
`pH 7.0. The portions in black are those common to
`all the amino acids; the portions shaded in red are
`the R groups.
`
`COO"
`H3f{I—(E—H
`H_g|:_CH3
`CH2
`$3,
`Isoleucine
`
`H
`+ /(Ii _
`H,_l\|«I
`EH2
`I-I30-—— H2
`
`Leucine
`
`Polar, uncharged R groups
`
`coo‘
`H31iI—(|J—H
`CH2
`H
`
`IS
`
`Cysteine
`
`C|!OO'
`+
`H3N—‘C£3—H
`H— —0}{
`$113-
`Threonine
`
`(|300_
`+
`H3N——(|3—H
`
`CH20H
`
`Serine
`
`coo‘
`H3fiI—(l:—H
`(EH2
`(|3H2
`s
`5H3
`Methionine
`
`(|100'
`H35J—c|:—H
`(I-111,
`(EH2
`C
`/ %
`
`o
`Hm
`Glutamine
`
`-
`
`.
`
`'
`
`'
`
`Asparagine
`
`lfmri Aliphatic R Groups The hydrocarbon R groups in this
`- ammo acids are nonpolar and hydrophobic (Fig. 5-6). The
`e chains of alanine, valine, leucine, and isoleucine, with
`.-si31“.°I‘»1V§ shapes, are important in promoting hydrophobic in-
`Within protein structures. Glycine has the simplest amino
`iiure. Where it is present in a protein, the minimal steric
`9f the glycine side chain allows much more structural flexi-
`a etlfcte other amino acids. Proline represents the opposite
`- 0 X Pelme. The secondary amino (imino) group is held in a
`--
`..
`. Tmatlon that reduces the structural flexibility of the protein
`- Point.
`
`Page 7
`
`

`
`116
`
`Part 11 Structure and Catalysis
`
`6'3
`
`Tryptophan
`
`UV
`
`0::
`
`
`
`Tyrosine
`
`Absorbance
`
`
`Phenylalanine
`i
`{, I I
`23:: 240 250 260 270 280 290 360 310
`Wavelength (nm)
`
`_‘_
`([700 T
`H3N—(|3—H
`CH2—:_'"l
`Cysteine
`
`(FOOT
`H,,1¢1—(|:—H
`“H—CH2
`Cysteine
`
`H313
`
`coo’
`<|.'.oo
`(I: —H H3f\i—o—H
`ca.-—-a H CH2
`Cystine
`
`Figure 5-7 Comparison of the light absorbance
`spectra of the aromatic amino acids at pH 6.0. The
`amino acids are present in equimolar amounts
`(10’3 M) under identical conditions. The light ab-
`sorbance of tryptophan is as much as fourfold
`higher than that of tyrosine. Phenylalanine absorbs
`less light than either tryptophan or tyrosine. Note
`that the absorbance maximum for tryptophan and
`tyrosine occurs near a wavelength of 280 nm.
`
`Aromatic R Groups Phenylalanine, tyrosine, and tryptophan,
`with their aromatic side chains (Fig. 5-6), are relatively nonpolar (hy-
`drophobic). All can participate in hydrophobic interactions, which are
`particularly strong when the aromatic groups are stacked on one an-
`other. The hydroxyl group of tyrosine can form hydrogen bonds, and it
`acts as an important functional group in the activity of some enzymes.
`Tyrosine and tryptophan are significantly more polar than phenylala-
`nine because of the tyrosine hydroxyl group and the nitrogen of the
`tryptophan indole ring.
`Tryptophan and tyrosine, and to a lesser extent phenylalanine,
`absorb ultraviolet light (Fig. 5-7 and Box 5-1). This accounts for the
`characteristic strong absorbance of light by proteins at a wavelength of
`280 nm, and is a property exploited by researchers in the characteriza-
`tion of proteins.
`
`Polar, Uncharged R Groups The R groups of these amino acids (Fig.
`5-6) are more soluble in water, or hydrophilic, than those of the non-
`polar amino acids, because they contain functional groups that form
`hydrogen bonds with water. This class of amino acids includes serine,
`threonine, cysteine, methionine, asparagine, and glutamine.
`The polarity of serine and threonine is contributed by their hydroxyl
`groups; that of cysteine and methionine by their sulfur atom; and that
`of asparagine and glutamine by their amide groups.
`Asparagine and glutamine are the amides of two other amino acids
`also found in proteins, aspartate and glutamate, respectively, to which
`asparagine and glutamine are easily hydrolyzed by acid or base. Cys-
`teine has an R group (a thiol group) that is approximately as acidic as
`the hydroxyl group of tyrosine. Cysteine requires special mention for
`another reason. It is readily oxidized to form a covalently linked di-
`meric amino acid cal1ed~'e§7ns't'i1"1e,.in wliicH‘tvi{6__<_:§§11e_ine .molecules'are
`joined-by-a-disulfide-bridge. Disulfide bridges of this kind occur in
`many proteins, stabilizing their structures.
`
`Negatively Charged (Acidic) R Groups The two amino acids having
`R groups with a net negative charge at pH 7.0 are aspartate and
`glutamate, each with a second carboxyl group (Fig. 5-6). These amino
`acids are the parent compounds of asparagine and glutamine, respec-
`tively.
`
`Positively Charged (Basic) R Groups The amino acids in which the R
`groups have a net positive charge at pH 7.0 are lysine, which has a
`second amino group at the e position on its aliphatic chain; arginine,
`which has a positively charged guanidino group; and histidine, con-
`taining an imidazole group (Fig. 5-6). Histidine is the only standard
`amino acid having a side chain with a pKa near neutrality.
`
`Page 8
`
`
`
`Page 8
`
`

`
`Chapter 5 Amino Acids and Peptides
`
`117
`
`BOX 54
`
`Absorption of Light by Molecules: The Lambert—Beer Law
`
`Measurement of light absorption is an important
`tool for analysis of many biological molecules. The
`fraction of the incident light absorbed by a solution
`at a given wavelength is related to the thickness of
`the absorbing layer (path length) and the concen-
`tration of the absorbing species. These two rela-
`tionships are combined into the Lambert—Beer
`law, given in integrated form as
`I
`log 70 = ecl
`
`where 10 is the intensity of the incident light, I is
`the intensity of the transmitted light, 5 is the
`molar absorption coefficient (in units of liters per
`mole-centimeter), c the concentration of the ab-
`sorbing species (in moles per liter), and l the path
`length of the light-absorbing sample (in centime-
`ters). The Lambert—Beer law assumes that the in-
`cident light
`is parallel and monochromatic and
`that the solvent and solute molecules are randomly
`
`oriented. The expression log (I0/I) is called the ab-
`sorbance, designated A.
`It is important to note that each millimeter path
`length of absorbing solution in a 1.0 cm cell absorbs
`not a constant amount but a constant fraction of
`
`the incident light. However, with an absorbing
`layer of fixed path length, the absorbance A is di-
`rectly proportional to the concentration of the ab-
`sorbing solute.
`The molar absorption coefficient varies with the
`nature of the absorbing compound, the'solvent, the
`wavelength, and also with pH if the light-absorb-
`ing species is in equilibrium with another species
`having a different spectrum through gain or loss of
`protons.
`In practice, absorbance measurements are usu-
`ally made on a set of standard solutions of known
`concentration at a fixed wavelength. A sample of
`unknown concentration can then be compared with
`the resulting standard curve, as shown in Figure 1.
`
` I
`
`0
`
`0
`
`I
`
`I
`40
`
`i
`
`i
`80
`
`I_
`
`_L!_
`120
`
`I
`
`I
`160
`
`I
`
`1
`200
`
`Figure 1 Eight standard solutions containing
`known amounts of protein and one sample contain-
`ing an unknown amount of protein were reacted
`with the Bradford reagent. This reagent contains a
`dye that shifts its absorption maximum to 595 nm
`when it binds amino acid residues. The A595 (ab-
`sorbance at 595 nm) of the standard samples was
`plotted against the protein concentration to create
`the standard curve, shown here. The A595 of the
`unknown sample, 0.58, corresponds to a protein
`concentration of 122 pig/mL.
`
`Protein concentration (pig/mL)
`
`
`
`Cells Also Contain Nonstandard Amino Acids
`
`In addition to the 20 standard amino acids that are common in all
`Proteins, other amino acids have been found as components of only
`i‘-e1‘tain types of proteins (Fig. 5~8aJ. Each of these is derived from one
`Of: the 20 standard amino acids, in a modification reaction that occurs
`after the standard amino acid has been inserted into a protein. Among
`the nrmstandard amino acids are 4-hydrO'xypr0li-ne,- a derivative of
`Droline, and 5-hydroxylysine; the former is found in plant cel1—wall
`pmteins, and both are found in the fibrous protein collagen of con-
`nective tissues. N-Methyllysine is found in myosin, a contractile
`protein of muscle. Another important nonstandard amino acid is
`
`‘
`
`
`
`Page 9
`
`,
`
`
`
`Page 9
`
`

`
`
`
`118
`
`Part II Structure and Catalysis
`
`H3FI~—CH2—CH2—CH2—Cl3H—COO_ HZN C If CH2 CH2 CH2
`
`(IJH COO"
`
`Ornithine
`
`+NH3
`
`ll
`O H
`
`Citrulline
`
`(b)
`
`:y-cafl‘.i'OY§?‘"gl'at—amate, fourld in the blood-clotting protein prothrom-
`bin as well as in certain other proteins that bind Ca2+ in their biologi-
`cal function. More complicated is the nonstandard amino acid des-
`mosi_ne, a derivative of four separate lysine residues, found in the
`fibrous protein elastin. 'S_éleno(_:}7s‘t‘éi_‘ne contains selenium rather
`than the oxygen of serine, and is found in glutathione peroxidase and a
`few other proteins.
`Some 300 additional amino acids have been found in cells and have
`a variety of functions but are not substituents of proteins. Orni-thine
`and ‘citrulline (Fig. 5—8b) deserve special note because they are key
`intermediates in the biosynthesis of arginine and in the urea cycle.
`These pathways are described in Chapters 21 and 17, respectively.
`
`Amino Acids Can Act as Acids and as Bases
`
`When a crystalline amino acid, such as alanine, is dissolved in water, it
`exists in solution as the dipolar ion, or zwitterion, which can act either
`as an acid (proton donor):
`
`1?
`5‘
`R—C—COO‘ xi‘ R—-(|J—COO‘ + H*
`*NH3
`NH3
`
`or as a base (proton acceptor):
`
`‘F
`‘F
`R—(|3—COO‘ + H’“ ,1 R—C|J—COOH
`+NH3
`"NH3
`
`Substances having ‘this dual nature are'ampl1'o‘t’e’ri?:-and are often
`called ififimtesf from “amphoteric
`electrolytes.” A simple
`monoamino monocarboxylic a—amino acid, such as alanine, is actually
`a diprotic acid when it is fully protonated, that is, when both its car-
`boxyl group and amino group have accepted protons. In this form it has
`two groups that can ionize to yield protons, as indicated in the follow-
`ing equation:
`
`W 1%
`Ir
`R—(|‘.—C!OOI.I. ——”—» R—(|}—COO‘
`
`+NH3
`
`+NH3
`
`1%
`W
`=7
`» R—(|3—COO‘
`
`NH2
`
`Amino Acids Have Characteristic Titration Curves
`
`Titration involves the gradual addition or removal of protons. Figure
`5-9 shows the titration curve of the diprotic form of glycine. Each
`molecule of added base results in the net removal of one proton from
`
`Page 10
`
`in‘
`IN r—(|,‘. ($112
`11.0.‘ __ /_(1H—COO‘
`_N
`
`\H
`H’
`4—Hydroxyproline
`
`H31¢I—cH2—cH—cH2—CH2—cH—coo‘
`I H I
`+NH3
`5-Hydroxylysine
`
`CH3 NH CH2 CH2 CH9 CH«_, CH COO"
`I
`‘NI-1;,
`
`6-N-Methyllysine
`
`COO‘
`I
`-ooc—cH—cH._.—(|3H—-coo
`+NH_.,
`‘y-C arboxyglutamate
`
`..
`I-l3N
`
`C00‘
`
`NH3
`<c1«12;g—o§I
`coo"
`
`>CH—(Gm:l2-I5’
`I
`
`+N
`
`»/
`
`FUQC
`
`+
`
`C00 '
`H;,N
`Desmosine
`
`HSe—CH2—-C|}H—COO‘
`+NH3
`Selenocysteine
`
`(a)
`
`Figure 5-8 (a) Some nonstandard amino acids
`found in proteins; all are derived from standard
`amino acids. The extra functional groups are shown
`in red. Desmosine is formed from four residues of
`lysine, whose carbon backbones are shaded in gray.
`Selenocysteine is derived from serine. (b) Ornithine
`and citrulline are intermediates in the biosynthesis
`of arginine and in the urea cycle. Note that two
`systems are used to number carbons in the naming
`of these amino acids. The ca, [3, y system used for
`y-carboxyglutamate begins at the a carbon (see Fig.
`5-2) and extends into the R group. The oz-carboxyl
`group is not included. In contrast, the numbering
`system used to identify the modified carbon in
`4-hydroxyproline, 5-hydroxylysine, and 6-N-methyl-
`lysine includes the oz-carboxyl carbon, which is des-
`ignated carbon 1 (or C-1).
`
`Page 10
`
`

`
`
`
`mlec-Hie of arni.no acid. The plot has two distinct stages, each cor-
`O-ne finding to the removal oi" one proton firoin glycine. Each of the two
`tiesllzes ,.,._.~.(y.ini1l.e.~= in aliape the titration curve of a monoprotic acid,
`_"5.1"a%1ea_u,
`-.}L.-..:1-.ic acid {see Fig. 4-10), and can be analyzed iii the same
`:l__Jacy__ At very low pH,
`the predominant
`ionic species of glycine is
`.rH,.N———Cll-,;~-COOH, the fully protonated form. At the midpoint in
`‘ti-i,gJf‘i1'i5l. stage ofthe titration, in which the —COOH group of glycine
`110%35
`its
`proton,
`equimolar
`concentrations
`of
`proton-donor
`(_.,. H3N,C'[I5;—(1(JOIi} and proton-‘acceptor l_ " I-I,-.,N--_C.H2—CO_C) ' } Spg_
`cl-@531-epreHent.Ai'. the lnidpoint of a titration (see Fig. 4-11.), the pH is
`"equal to the pK,, of the protonated group being titrated. For glycine, the
`H at the midpoint is 2.34, thus its —COOH group has a 13K“ of 2.34.
`[Recall that pH and pK,, are simply convenient no1;a1.io'ns for proton
`concenti.'al.ion and the equilibrium Constant for ioliiization, respectively
`(Chapter 4). The pk], is a measure of the tendency of a group to give up
`a proton, with that tendency decreasing tenfold as the pK,, increases by
`one unit.J As the titration proceeds, another important point is reached
`at pH 5.97. Here there is a_1p6i3't{(.‘)'f!‘i'r;ff'1'é'c't'i'()'I1,‘iat which-removal-of'the
`first-proton is essentially complete, and removal of the second has just
`begun. At this pH the glycine is present largely as the dipolar ion
`+H3N-CH2—COO’. We shall return to the significance of this inflec—
`tion point in the titration curve shortly.
`The second stage of the titration corresponds to the removal of a
`proton from the —NH§“ group of glycine. The pH at the midpoint of this
`stage is 9.60, equal to the pK,, for the_—NH§“ group. The titration is
`complete at a pH of about 12, at which point the predominant form of
`glycine is H2N—CH2—COO’.
`From the titration curve of glycine we can derive several important
`pieces of information. First, it gives a quantitative measure of the pK,,
`of each of the two ionizing groups, 2.34 for the —COOH group and 9.60
`for the —NH§‘ group. Note that the carboxyl group of glycine is over
`100 times more acidic (more easily ionized) than the carboxyl group of
`acetic acid, which has a pK,, of 4.76. This effect is caused by the nearby
`positively charged amino group on the oz—carbon atom, as described in
`Figure 5-10.
`The second piece of information given by the titration curve of gly-
`cine (Fig. 5-9) is that this amino acid has two regions of buffering
`power (see Fig. 4-12). One of these is the relatively flat portion of the
`curve centered about the first pl?” of 2.34, indicating that glycine is a
`good buf'I"or near this pH. The other bui'l'e'ring zone extends for "-1.2 pH
`units centered around pH 9.60. Note also that glycine is not a good
`buffer at the pH of intracellular fluid or blood, about 7.4.-.
`The I--Iondcrson-Haaseibalch equation {Chapter 4) can be used to
`Ciilculatr-. the proportions of proton-donor and proton-acceptor species
`Oi Elyciiie required to make a buffer at a given pH within the buffering
`ranges ofglycine; it also makes it possible to solve other kinds of buffer-
`l31"0l1len'13 involving amino acids (see Box 4-2).
`
`Chapter 5 Amino Acids and Peptides
`
`Kin,
`CH2 %
`COOH
`
`NH3
`
`(|3oo‘
`
`13
`
`NH2
`
`(|300’
`
`:.-.-:
`
` .________....-_.._.__.________;.'
`
`
`0
`
`0:5
`
`1
`
`1.5
`
`2
`
`OH” (equivalents)
`
`Figure 5-9 The titration curve of 0.1 M glycine
`at 25 °C. The ionic species predominating at key
`points in the titration are shown above the graph.
`The shaded boxes, centered about pK1 = 2.34 and
`pK2 = 9.60, indicate the regions of greatest buffer-
`ing power.
`
`Figure 5-10 (a) Interactions between the a-amino
`and av-carboxyl groups in an ct-amino acid. The
`nearby positive charge of the —NH;;* group makes
`ionization of the carboxyl group more likely (i.e.,
`lowers the pK, for —COOH). This is due to a stabi-
`lizing interaction between opposite charges on the
`zwitterion and a repulsive interaction between the
`positive charges of the amino group and the depart-
`ing proton. (b) The normal pKn for a carboxyl
`group is approximately 4.76, as for acetic acid.
`
`Acetic acid
`
`Lu-Amino acid (glycine)
`+1TIH3
`-1TiI'I_-i
`H -L|.‘.—COOH LT‘ H—(|3—COO" + H F
`H
`
`H (
`
`pKn = 2.34
`
`CH,-COOH _‘-é CH,,—COO" + H+
`
`pKl : 476
`
`a)
`
`(b)
`
`Page 11
`
`
`
`Page 11
`
`

`
`
`
`120
`
`Part II Structure and Catalysis
`
`The Titration Curve Predicts the
`
`Electric Charge of Amino Acids
`
`Another important piece of information derived from the titration
`curve of an amino acid is the relationship between its net electric
`charge and the pH of the solution. At pH 5.97, the point of inflection
`between the two stages in its titration curve, glycine is present as its
`dipolar form, fully ionized but with no net electric charge (Fig. 5-9).
`This characteristic pH is called the isoelectric point.or isoelectric
`- pH; designated- pI- or pH1. For an amino acid such as glycine, which has
`no ionizable group in the side chain, the isoelectric point is the arith-
`metic-mean of the two pKa values:
`1
`P1 = §(PK1 + PK2)
`
`which in the case of glycine is
`
`pl = %(2.34 + 9.60) = 5.97
`
`As is evident in Figure 5-9, glycine has a net negative charge at any
`pH above its pl and will thus move toward the positive electrode (the
`anode) when placed in an electric field. At any pH below its pl, glycine
`has a net positive charge and will move toward the negative electrode,
`the cathode. The farther the pH of a glycine solution is from its isoelec-
`tric point, the greater the net electric charge of the population of gly-
`cine molecules. At pH 1.0, for example, glycine exists entirely as the
`form "'H;,N—CH3—COOH, with a net positive charge of 1.0. At pH
`2.34, where there is an equal mixture of ‘H3N—CH2—C0OH and
`“H3N—CH2—COO‘, the average or net positive charge is 0.5. The
`sign and the magnitude of the net charge of any amino acid at any pH
`can be predicted in the same way.
`This information has practical importance. For a solution contain-
`ing a mixture of amino acids, the different amino acids can be sepa-
`rated on the basis of the direction and relative rate of their migration
`when placed in an electric field at a known pH.
`1
`
`Amino Acids Differ in Their Acid—Base Properties
`
`The shared properties of many amino acids permit some simplifying
`generalizations about the acid-base behavior of different classes of
`amino acids.
`
`All amino acids with a single a-amino group, a single oz-carboxyl
`group, and an R group that does not ionize have titration curves resem-
`bling that of glycine (Fig. 5-9). This group of amino acids is character-
`ized by havi

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