throbber
MYLAN EXHIBIT - 1012 (Part 1)
`Mylan Pharmaceuticals, Inc. v. Bausch Health Ireland, Ltd. - IPR2022-00722
`
`

`

`Lehninger Principles of Biochemistry Third Edition
`
`David L. Nelson and Michael M. Cox
`
`Copyright © 2000, 1993, 1982 by Worth Publishers
`
`All rights reserved
`Printed in the United States of America
`
`Library of Congress Cataloging-in-Publication Data
`
`Nelson, David L.
`Lehninger principles of biochemistry / David L. Nelson, Michael M. Cox.— 3rd ed.
`p.cm.
`Includes index.
`ISBN 1-57259-153-6
`
`1, Biochemistry. II. Cox, Michael M._III. Title.I. Nelson, David L. (David Lee), 1942-
`
`
`QD415 .L44 2000
`572—dc21
`
`99-049137
`
`Printing:
`
`5
`
`4
`
`3
`
`2
`
`1
`
`Year:
`
`04
`
`03
`
`02
`
`O01
`
`00
`
`DevelopmentEditor: Morgan Ryan, with Linda Strange and Valerie Neal
`Project Editor: Elizabeth Geller
`Art Director: Barbara Rusin
`
`Design: Paul Lacy
`Production Supervisor: Bernadine Richey
`Layout: York Graphic Services and Paul Lacy
`Photo Editor: Deborah Goodsite
`
`Mlustrations: SusanTilberry (with Alan Landau and Joan Waites), J.B. Woolsey & Associates,
`Laura Pardi Duprey, and York Graphic Services
`Molecular Graphics: Jean-Yves Sgro
`Composition: York Graphic Services
`Printing and Binding: R.R. Donnelley and Sons
`
`Cover (from top to bottom): Cut-away view of GroEL, a protein complex in-
`volved in protein folding; cut-away viewof tobacco mosaic virus, an RNAvirus;
`ribbon modelof a 8-barrel structural domain from UDP N-acetylglucosamine
`acyltransferase; cut-awayviewof the F, subunit of ATP synthase, with bound
`ATP shown as a stick structure; mesh surface imageof the electron-transfer
`protein cytochromec, with its heme group shown asa stick structure.
`
`Cover images created by Jean-Yves Sgro.
`
`Illustration credits begin on p. IC-1 and constitute a continuation of the
`copyright page.
`
`Steenbock Memorial Library
`University of/isconsin-Madisan
`550 BabcockDrive
`Madison, Wi 53706-1293
`
`Worth Publishers
`41 Madison Avenue
`
`NewYork, NY 10010
`
`

`

`
`
`This view of Earth from space shows that most of the
`planet's surface is covered with water. The seas, where
`life probably first arose, are today the habitat of countless
`organisms.
`
`82
`
`Water
`
`Water is the most abundant substanceinliving systems, making up 70% or
`more of the weight of most organisms. The first living organisms doubtless
`arose in an aqueous environment, and the course of evolution has been
`shaped bythe properties of the aqueous mediumin whichlife began.
`This chapter begins with descriptions of the physical and chemical
`properties of water, to which all aspects of cell structure and function are
`adapted. The attractive forces between water molecules and the slight ten-
`dencyof waterto ionize are of crucial importanceto the structure and func-
`tion of biomolecules. We will review the topic of ionization in terms of equi-
`librium constants, pH, and titration curves, and consider how aqueous
`solutions of weak acids or bases and their salts act as buffers against pH
`changes in biological systems. The water molecule andits ionization prod-
`ucts, H* and OH, profoundly influence the structure, self-assembly, and
`propertiesofall cellular components, including proteins, nucleic acids, and
`lipids. The noncovalent interactions responsible for the strength and speci-
`ficity of “recognition” among biomolecules are decisively influenced by the
`solvent properties of water.
`
`Weak Interactions in Aqueous Systems
`Hydrogen bonds between water molecules provide the cohesive forces that
`make water a liquid at room temperature and that favor the extreme or-
`dering of moleculesthatis typical of crystalline water (ice). Polar biomole-
`cules dissolve readily in water because they can replace water-water inter-
`actions with more energetically favorable water-solute interactions,
`In
`contrast, nonpolar biomolecules interfere with water-water interactions but
`are unable to form water-solute interactions—consequently, nonpolar mol-
`ecules are poorly soluble in water. In aqueous solutions, nonpolar molecules
`tend to cluster together.
`Hydrogen bonds andionic, hydrophobic (Greek, “water-fearing”), and
`van der Waals interactions are individually weak, but collectively they have
`a very significant influence on the three-dimensional structures of proteins,
`nucleic acids, polysaccharides, and membranelipids.
`
`Hydrogen Bonding Gives Water Its Unusual Properties
`Water has a higher melting point, boiling point, and heat of vaporization
`than most other commonsolvents (Table 4-1). These unusual properties
`
`

`

`Chapter 4 Water
`
`83
`
`table 4—1]
`of Some CommonSolvents
`porization
`Melting Point, Boiling Point, and Heat of Vaporizati
`
`Water
`Methanol (CH,0H)
`Ethanol (CH;CH,OH)
`Propanol (CH;CH,CH,OH)
`Butanol (CH,(CH,),CH,OH)
`Acetone (CH,COCH;)
`Hexane (CH.(CH,),CH,)
`Benzene (CgH,)
`Butane (CH,(CH,).CH;)
`Chloroform (CHCl,;)
`
`:
`Melting
`point (°C)
`0
`—98
`=117
`-127
`—90
`—95
`-98
`6
`-—135
`-63
`
`Boiling
`point (°C)
`100
`65
`78
`97
`117
`56
`69
`80
`-0.5
`61
`
`Heat of
`vaporization
`(J/g)*
`2,260
`1,100
`854
`687
`590
`523
`423
`394
`381
`247
`
`"The heat energy required to convert 1.0g of a liquid at its boiling point, at atmospheric
`PVOSSUTE, jnto ftsgaseous state at the samme temperature. ft is a direct measure ofthe energy
`required te overcome attractive forces between molecules in the liquid phase.
`
`are a consequence of attractions between adjacent water molecules that
`give liquid water great internal cohesion. A look at the electron structure of
`the H,O molecule reveals the cause of these intermolecular attractions.
`Each hydrogen atom of a water molecule shares an electron pair with
`the oxygen atom. The geometry of the molecule is dictated by the shapes of
`the outer electron orbitals of the oxygen atom, which are similar to the
`bonding orbitals of carbon (see Fig. 3-4a). These orbitals describe a rough
`tetrahedron, with a hydrogen atom at each of two corners and unshared
`electronpairs at the other two corners (Fig. 4-1a). The H—O—Hbondan-
`gle is 104.5°, slightly less than the 109.5° of a perfect tetrahedron because
`of crowding by the nonbonding orbitals of the oxygen atom.
`The oxygen nucleusattracts electrons more strongly than does the hy-
`drogen nucleus (a proton); oxygenis more electronegative (see Table 3-2).
`The sharing of electrons between H and O is therefore unequal; the elec-
`trons are more often in the vicinity of the oxygen atom than of the hydro-
`gen. The result of this unequal electron sharing is two electric dipoles in the
`water molecule, one along each of the H—O bonds; the oxygen atom bears
`a partial negative charge (26), and each hydrogena partial positive charge
`(6°). As a result, there is an electrostatic attraction between the oxygen
`atom of one water molecule and the hydrogen of another (Fig. 4-1c), called
`a hydrogen bond. Throughout this book, we will represent hydrogen
`bonds with three parallel blue lines, as in Figure 4—1c.
`
`
`
`Hydrogen bond
`0.177 nm
`
`Covalent bond
`0.0965 nm
`
`
`
`
`figure 4-1
`Structure of the water molecule. The dipolar nature of
`the H,O molecule is shown by (a) ball-and-stick and
`(b) space-filling models. The dashed lines in (a) represent
`the nonbonding orbitals. There is a nearly tetrahedral
`arrangementof the outer-shell electron pairs around the
`oxygen atom; the two hydrogen atoms have localized
`
`(c)
`partial positive charges (8*) and the oxygen atom has a
`partial negative charge (26 ). (c) Two HO molecules
`joined by a hydrogen bond (designated here, and
`throughout this book, by three blue lines) between the
`oxygen atom of the upper molecule and a hydrogen atom
`of the lower one. Hydrogen bondsare longer and weaker
`than covalent O—H bonds.
`
`

`

`Hydrogen bonding in ice. Each water molecule forms the
`maximum of four hydrogen bonds,creating a regular
`crystal lattice. In liquid water at room temperature and
`atmospheric pressure, by contrast, each water molecule
`hydrogen bonds with an average of 3.4 other water mole-
`cules. The crystal lattice of ice occupies more space than
`that occupied by the same number of HO molecules in
`liquid water; ice is less dense than—and thus floats on—
`liquid water.
`
`Foundations of Biochemistry figure 4-2
`
`Hydrogen bonds are weaker than covalent bonds. The hydrogen bonds
`in liquid water have a bond dissociation energy (the energy required to
`break a bond) of about 20 kJ/mol, compared with 348 kJ/mol for the cova-
`lent C—C bond. At room temperature, the thermal energy of an aqueous so-
`lution (the kinetic energy of motionof the individual atoms and molecules)
`is of the same order of magnitude as that required to break hydrogen bonds.
`When water is heated, the increase in temperature reflects the faster mo-
`tion of individual water molecules. Although at any given time most of the
`moleculesin liquid water are engaged in hydrogen bonding, the lifetime of
`each hydrogen bondis less than 1 * 107® s. The apt phrase “flickering clus-
`ters” has been applied to the short-lived groups of hydrogen-bonded mole-
`cules in liquid water. The sum ofall the hydrogen bonds between molecules
`nevertheless confers great internal cohesion on liquid water.
`The nearly tetrahedral arrangement of the orbitals about the oxygen
`atom (Fig. 4—1a) allows each water molecule to form hydrogen bonds with
`as many as four neighboring water molecules. In liquid water at room tem-
`perature and atmospheric pressure, however, water molecules are disorga-
`nized and in continuous motion, so that each molecule forms hydrogen
`bonds with an averageof only 3.4 other molecules. In ice, on the other hand,
`each water moleculeis fixed in space and forms hydrogen bonds with four
`other water moleculesto yield a regular lattice structure (Fig. 4-2). Break-
`age of a sufficient number of hydrogen bondsto destabilize the crystal lat-
`tice of ice requires much thermal energy, which accounts for the relatively
`high melting point of water (Table 4-1). When ice melts or water evapo-
`rates, heat is taken up bythe system:
`
`H,0(s) —> H,0(1)
`
`AH = +5.9 kJ/mol
`
`H,0O(1) —> H,0(g)
`
`AH = +44.0kJ/mol
`
`During melting or evaporation, the entropy of the aqueous system in-
`creases as more highly ordered arrays of water molecules relax into the less
`orderly hydrogen-bonded arrays in liquid water or the wholly disordered
`gaseous state. At room temperature, both the melting of ice and the evapo-
`ration of water occur spontaneously; the tendency of the water molecules
`to associate through hydrogen bondsis outweighed by the energetic push
`toward randomness. Recall that the free-energy change (AG) must have a
`negative value for a process to occur spontaneously: AG = AH — T AS,
`where AG represents the driving force, AH the enthalpy change from mak-
`ing and breaking bonds, and AS the change in randomness. Because AH is
`positive for melting and evaporation, it is clearly the increase in entropy
`(AS) that makes AG negative and drives these transformations.
`
`

`

`Chapter 4 Water
`
`85
`
`Hydrogen
`acceptor
`Hydrogen
`donor
`
`No”
`“ae
`I wyxny ova V7
`oO
`O
`N
`oO
`0”
`N
`H
`H
`H
`H
`H
`H
`|
`|
`|
`|
`|
`|
`oO
`Oo
`O
`N
`N
`N
`|
`|
`|
`|
`|
`|
`
`figure 4-3
`Common hydrogen bondsin biological systems. The
`hydrogen acceptor is usually oxygen or nitrogen.
`
`Water Forms Hydrogen Bonds with Polar Solutes
`Hydrogen bonds are not unique to water. They readily form between an
`electronegative atom (the hydrogen acceptor, usually oxygen or nitrogen
`with a lone pair of electrons) and a hydrogen atom covalently bonded to an-
`other electronegative atom (the hydrogen donor) in the same or another
`molecule (Fig. 4-3). Hydrogen atoms covalently bonded to carbon atoms
`(which are not electronegative) do not participate in hydrogen bonding.
`The distinction explains why butanol (CH3(CH,).CH,OH) hasa relatively
`high boiling point of 117 °C, whereas butane (CH3(CHz)2CHs3) has a boiling
`point of only —0.5 °C. Butanol has a polar hydroxyl group and thus can form
`intermolecular hydrogen bonds.
`Uncharged but polar biomolecules such as sugars dissolve readily in
`water because of the stabilizing effect of hydrogen bonds between the hy-
`droxyl groups or carbonyl oxygen of the sugar and the polar water mole-
`cules. Alcohols, aldehydes, ketones, and compounds containing N—H
`bondsall form hydrogen bonds with water molecules (Fig. 4—4) and tend to
`be soluble in water.
`
`Between the
`hydroxyl group
`of an alcohol
`and water
`
`Between the
`carbonyl group
`of a ketone
`and water
`
`Between peptide
`groups in
`polypeptides
`
`Between
`complementary
`bases of DNA
`
`figure 4-4
`Somebiologically important hydrogen bonds.
`
`Thymine
`
`Adenine
`
`H
`
`R
`
`H
`
`cH3
`
`RC.
`x6
`S27 ne i Sow
`v
`Ze~ yh
`i
`OO
`acted
`i
`sigh
`pots ae
`=
`|
`Cc
`C
`N
`NH
`|
`I
`“ot” Se”
`R
`Cc
`Cc
`
`i~
`
`o* “Nn
`a
`
`R®
`
`H
`
`Bi,
`
`0
`H
`=
`O
`H -H
`
`Bho
`C
`Oo
`H
`id
`“H
`
`Hydrogen bondsare strongest when the bonded moleculesare oriented
`to maximize electrostatic interaction, which occurs when the hydrogen
`atom and the two atoms that share it are in a straight line—that is, when
`the acceptor atom is in line with the covalent bond between the donor atom
`and H (Fig. 4-5). Hydrogen bondsare thus highly directional and capable
`of holding two hydrogen-bonded molecules or groups in a specific geomet-
`ric arrangement. As weshall seelater, this property of hydrogen bonds con-
`fers very precise three-dimensional structures on protein and nucleic acid
`molecules, which have manyintramolecular hydrogen bonds.
`
`0
`HStrong
`- hydrogen bond oo
`
`|
`¥ Weaker
`
`hydrogen bond
`
`—P
`xe
`
`figure 4-5
`Directionality of the hydrogen bond. The attraction
`betweenthe partial electric charges (see Fig. 4-1) is
`greatest when the three atomsinvolved(in this case O, H,
`and 0) lie in a straight line. When the hydrogen-bonded
`moieties are structurally constrained (as when they are
`parts of a single protein molecule, for example), this ideal
`geometry may not be possible and the resulting hydrogen
`bond is weaker.
`
`

`

`86
`
`Part | Foundations of Biochemistry
`
`WaterInteracts Electrostatically with Charged Solutes
`Wateris a polar solvent. It readily dissolves most biomolecules, which are
`generally charged or polar compounds (Table 4-2); compounds that dis-
`solve easily in water are hydrophilic (Greek, “water-loving”). In contrast,
`nonpolar solvents such as chloroform and benzene are poor solvents for po-
`lar biomolecules but easily dissolve those thatare hydrophobic—nonpolar
`molecules such as lipids and waxes.
`Water dissolves salts such as NaCl by hydrating andstabilizing the Na™
`and Cl- ions, weakening the electrostatic interactions between them and
`thus counteracting their tendencyto associate in a crystalline lattice (Fig.
`
`tanie —2
`tahia 4
`Some Examples of Polar, Nonpolar, and Amphipathic Biomolecules
`(Shown as lonic Formsat pH 7)
`Polar
`Glucose
`
`CH,OH
`
`
`
`—|
`
`__| Polar groups
`en Nonpolar groups
`
`Glycine
`Aspartate
`
`Lactate
`
`Glycerol
`
`Nonpolar
`Typical wax
`
`+NH,—CH,—COO-
`ae
`-00C—CH,—CH—COO-
`iialome
`OH
`
`oH
`HOCH,.—CH—CH,OH
`
`|
`CH,(CH2);—CH=CH—(CH2)g—CH2—C.0
`|
`CH;(CH,);—CH=CH—(CH2);—CHe
`
`Amphipathic
`Phenylalanine
`
`Phosphatidylcholine
`
`*NH,
`
`()-ctt-Gx —CO0O0-
`
`9
`CH,(CH»);;CH,—C—O—CH,
`|
`CHs(CHa)isCHs—¢—O—CH
`
`0
`
`eae
`
`

`

`Chapter 4 Water
`
`87
`
`
`
`
`
`
`figure 4-6
`Water dissolves manycrystalline salts by hydrating their
`componentions. The NaCl crystal lattice is disrupted as
`water molecules cluster about the Cl” and Na* ions. The
`ionic charges are partially neutralized, and the electrostatic
`attractions necessary forlattice formation are weakened.
`
`4-6). The same factors apply to charged biomolecules, compounds with
`functional groups suchas ionized carboxylic acids (—COO), protonated
`amines (—NH}), and phosphate esters or anhydrides. Water readily dis-
`solves such compounds by replacing solute-solute hydrogen bonds with
`solute-water hydrogen bonds, thus screening the electrostatic interactions
`between solute molecules.
`Water is especially effective in screening the electrostatic interactions
`between dissolved ions because of its high dielectric constant, a physical
`property reflecting the number of dipoles in a solvent. The strength, or
`force (F), of ionic interactions in a solution depends upon the magnitude of
`the charges (Q), the distance between the charged groups (7), and the di-
`electric constant (€) of the solvent in which the interactions occur:
`
`Fe= QQ.
`er?
`
`For water at 25 °C, e (which is dimensionless) is 78.5, and for the very non-
`polar solvent benzene, ¢ is 4.6. Thus, ionic interactions are much strongerin
`less polar environments. The dependence on r? is such that ionic attractions
`or repulsions operate only over short distances—in the range of 10 to
`40 nm (depending on the electrolyte concentration) when the solventis
`water.
`
`Entropy Increases as Crystalline Substances Dissolve
`Asa salt such as NaCl dissolves, the Na* and Cl ions leaving the crystallat-
`tice acquire far greater freedom of motion (Fig. 4-6). The resulting in-
`crease in the entropy (randomness) of the system is largely responsible for
`the ease of dissolving salts such as NaCl in water. In thermodynamic terms,
`formation of the solution occurs with a favorable change in free energy:
`AG = AH— T AS, where AH has a small positive value and T ASa large
`positive value; thus AG is negative.
`k= °A aesT&S
`-~% = x
`
`AY ZO
`igs =O
`
`

`

`88
`
`Part | Foundations of Biochemistry
`
`Nonpolar Gases Are Poorly Soluble in Water
`The moleculesof the biologically important gases COs, O., and Np are non-
`polar, In O, and Ng, electrons are shared equally by both atoms. In CO., each
`C=Obondis polar, but the two dipoles are oppositely directed and cancel
`each other (Table 4-3). The movement of molecules from the disordered
`gas phase into aqueous solution constrains their motion and the motion of
`water molecules and therefore represents a decrease in entropy. The non-
`polar nature of these gases and the decrease in entropy when they enterso-
`lution combine to make them very poorly soluble in water (Table 4-3).
`Some organisms have water-soluble carrier proteins (hemoglobin and myo-
`globin, for example) that facilitate the transport of O. Carbon dioxide
`forms carbonic acid (HyCO3) in aqueous solution andis transported as the
`HCO; (bicarbonate) ion, either free—bicarbonate is very soluble in water
`(~100 g/L at 25 °C)—or bound to hemoglobin.
`Two other gases, NH3 and H,S, also have biological roles in some or-
`ganisms; these gases are polar and dissolve readily in water.
`
`table 4-3 0
`| Solubilities of Some Gases in Water
`
`
`Solubility
`Gas in water (g/L)'arnaeeeStructure* Polarity
`
`
`
`Nitrogen
`N=N
`Nonpolar
`0.018 (40 °C)
`
`| Oxygen
`|
`_ Carbon
`|
`dioxide
`
`| Ammonia
`i
`i
`| Hydrogen
`|
`‘sulfide
`iL
`
`o=0
`8
`5
`Sarr iTt
`0=C=0
`
`8
`
`HHav
`og |
`N
`H
`H |
`tae
`
`hy
`
`Nonpolar
`
`Nonpolar
`
`Polar
`
`Polar
`
`0.035 (50 °C)
`
`0.97 (45 °C)
`
`|
`4
`
`900 (10°C)
`
`1,860 (40 °C)
`
`*The arrows representelectric dipoles; there is a partial negative charge (6~) at the head of the
`arrow, a partial positive charge (6°; not shown here) at thetail.
`‘Note that polar molecules dissolve far better even at low temperatures than do nonpolar
`molecules at relatively high temperatures.
`
`Nonpolar CompoundsForce Energetically Unfavorable Changes
`in the Structure of Water
`When water is mixed with benzene or hexane, two phases form; neither liq-
`uid is soluble in the other. Nonpolar compounds such as benzene and
`hexane are hydrophobic—they are unable to undergo energetically favor-
`able interactions with water molecules, and they actually interfere with the
`hydrogen bonding among water molecules. All molecules or ions in aqueous
`solution interfere with the hydrogen bonding of some water molecules in
`their immediate vicinity, but polar or charged solutes (such as NaCl) com-
`pensate for lost water-water hydrogen bonds by forming new solute-water
`interactions. The net change in enthalpy (AH) for dissolving these solutes
`is generally small. Hydrophobic solutes, however, offer no such compensa-
`tion, and their addition to water may therefore result in a small gain of en-
`thalpy; the breaking of hydrogen bonds between water molecules takes up
`
`

`

`Chapter 4 Water
`
`89
`
`Dispersion of
`lipids in H,O
`Eachlipid
`molecule forces
`surrounding H)O
`molecules to become
`highly ordered.
`
`Clusters oflipid
`molecules
`
`
`
`bo3ReaaeaybPSpiotaie amFgeob
`
`
`& &
`ay)
`
`Only lipid portions
`at the edge of
`the cluster force the
`ordering of water.
`Fewer H0 molecules
`#%, are ordered, and
`entropy is increased.
`
`Hydrophilic
`"head group"
`
`molecules in bulk phase
`
`"Flickering clusters" of Hg0
`
`Highly ordered HzO molecules form
`“cages” around the hydrophobic alkyl chains
`
`4
`
`(a)
`
`
`
`energy from the system. Furthermore, dissolving hydrophobic compounds
`in water produces a measurable decreasein entropy. Water moleculesin the
`immediate vicinity of a nonpolar solute are constrained in their possible ori-
`entations as they form a highly ordered cagelike shell around each solute
`molecule. These water molecules are not as highly ordered as those in the
`crystalline compoundof a nonpolar solute and water (a clathrate), but the
`effect is the same in both cases: the ordering of water molecules reduces
`All hydrophobic
`entropy. The numberof ordered water molecules, and therefore the magni-
`groups are
`tude of the entropy decrease,is proportional to the surface area of the hy-
`sequestered from
`water; ordered
`drophobic solute enclosed within the cage of water molecules. The free-
`shell of HgO
`energy changefor dissolving a nonpolar solute in wateris thus unfavorable:
`molecules is
`minimized, and
`AG = AH — T AS, where AH hasapositive value, AS has a negative value,
`entropyis further
`increased.
`and AG is positive.
`Amphipathie compounds contain regions that are polar (or charged)
`and regions that are nonpolar (Table 4-2). When an amphipathic com-
`pound is mixed with water, the polar, hydrophilic region interacts favorably
`with the solvent andtendsto dissolve, but the nonpolar, hydrophobic region
`tends to avoid contact with the water (Fig. 4-7a). The nonpolar regions of
`the molecules cluster together to present the smallest hydrophobic area to
`the aqueous solvent, and the polar regions are arranged to maximize their
`interaction with the solvent (Fig. 4—7b). These stable structures of amphi-
`pathic compounds in water, called micelles, may contain hundreds or
`thousands of molecules. The forces that hold the nonpolar regions of the
`molecules together are called hydrophobic interactions. The strength of
`hydrophobic interactions is not due to any intrinsic attraction between non-
`polar moieties. Rather, it results from the system's achieving greatest ther-
`modynamic stability by minimizing the numberof ordered water molecules
`required to surround hydrophobic portions of the solute molecules.
`
`Micelles
`
`(b)
`
`figure 4-7
`(a) Long-
`Amphipathic compounds in aqueous solution.
`chain fatty acids have very hydrophobic alkyl chains,
`each of which is surrounded by a layer of highly ordered
`water molecules. (b) By clustering together in micelles,
`the fatty acid molecules expose the smallest possible
`hydrophobic surface area to the water, and fewer water
`molecules are required in the shell of ordered water. The
`energy gained by freeing immobilized water molecules
`stabilizes the micelle.
`
`

`

`90
`
`Part |
`
`Foundations of Biochemistry
`
`Ordered water
`
`substrate and enzyme
`
`interacting with
`“eea.
`
`p o&
`
`Sr
`Robo

`Seec
`
`Enzyme
`
`Disordered water
`displaced by
`enzyme-substrate
`
`interaction
`
`Enzyme-substrate
`interaction stabilized
`by hydrogen-bonding,
`ionic, and hydrophobic
`interactions
`
`figure 4-8
`Release of ordered water favors formation of an
`enzyme-substrate complex. While separate, both enzyme
`and substrate force neighboring water molecules into an
`ordered shell. Binding of substrate to enzyme releases
`some of the ordered water, and the resulting increase in
`entropy provides a thermodynamic push toward formation
`of the enzyme-substrate complex.
`
`Many biomolecules are amphipathic; proteins, pigments, certain vita-
`mins, and the sterols and phospholipids of membranes all have polar and
`nonpolar surface regions. Structures composed of these molecules are sta-
`bilized by hydrophobic interactions among the nonpolar regions. Hy-
`drophobic interactions amonglipids, and betweenlipids and proteins, are
`the most important determinants of structure in biological membranes. Hy-
`drophobic interactions between nonpolar amino acids also stabilize the
`three-dimensional folding patterns of proteins.
`Hydrogen bonding between water and polar solutes also causes some
`ordering of water molecules, but the effectis less significant than with non-
`polar solutes. Part of the driving force for binding of a polar substrate (re-
`actant) to the complementary polar surface of an enzymeis the entropy in-
`crease as the enzymedisplaces ordered water fromthe substrate (Fig. 4-8).
`
`Van der Waals Interactions Are Weak Interatomic Attractions
`When two uncharged atoms are brought very close together, their sur-
`rounding electron clouds influence each other. Random variations in the po-
`sitions of the electrons around one nucleus may create a transient electric
`dipole, which induces a transient, opposite electric dipole in the nearby
`atom. The two dipoles weakly attract each other, bringing the two nuclei
`closer. These weakattractions are called van der Waals interactions. As
`the two nuclei draw closer together, their electron clouds begin to repel
`eachother. At the point when the vander Waalsattraction exactly balances
`this repulsive force, the nuclei are said to be in van der Waals contact. Each
`atom has a characteristic van der Waals radius, a measure of how close
`that atom will allow another to approach (see Table 3-1). In the “space-
`filling” molecular models shown throughout this book (e.g., Fig. 83-7c) the
`atoms are depicted in sizes proportional to their van der Waals radii.
`
`Weak Interactions Are Crucial to Macromolecular
`Structure and Function
`The noncovalent interactions we have described (hydrogen bonds and
`ionic, hydrophobic, and van der Waals interactions) (Table 4-4) are much
`weaker than covalent bonds. An input of about 350 kJ of energy is required
`to break a mole of (6 x 10°°) C—Csingle bonds, and about 410 kJ to break
`a mole of C—H bonds,butaslittle as 4 kJ is sufficient to disrupt a mole of
`typical van der Waals interactions. Hydrophobic interactions are also much
`weaker than covalent bonds, although they are substantially strengthened
`by a highly polar solvent (a concentratedsalt solution, for example). Ionic
`interactions and hydrogen bondsare variable in strength, depending on the
`polarity of the solvent, but they are alwayssignificantly weaker than cova-
`lent bonds.
`In aqueous solvent at 25 °C, the available thermal energy can
`be of the same order of magnitude as the strength of these weak interac-
`tions, and the interaction between solute and solvent (water) molecules is
`nearly as favorable as solute-solute interactions. Consequently, hydrogen
`bondsandionic, hydrophobic, and van der Waals interactions are continu-
`ally formed and broken.
`Although these four types of interactions are individually weak relative
`to covalent bonds, the cumulative effect of many such interactions with a
`protein or nucleic acid can be very significant. For example, the noncova-
`lent binding of an enzymeto its substrate may involve several hydrogen
`bonds and one or moreionic interactions, as well as hydrophobic and van
`der Waals interactions. The formation of each of these weak bonds con-
`tributes to a net decrease in the free energy of the system. The stability of
`a noncovalentinteraction such as that of a small molecule hydrogen-bonded
`to its macromolecular partneris calculable from the binding energy. Stabil-
`
`

`

`Chapter 4 Water
`
`table 4-4
`
`Four Types of Noncovalent (“Weak”) Interactions
`among Biomolecules in Aqueous Solvent
`
`Hydrogen bonds
`Between neutral groups
`
`.
`Between peptide bonds
`
`\
`ae 1 H-—O—
`
`NN
`a
`
`C=O'''H—N
`
`rs
`
`lonic interactions
`
`Attraction
`

`
`~O—C—
`
`—+*NH;
`
`~<
`
`Repulsion
`
`Hydrophobic interactions
`
`Van der Waals interactions
`
`Any two atoms in
`close proximity
`
`ity, as measured by the equilibrium constant (see below) of the binding re-
`action, varies exponentially with binding energy. The dissociation of two
`biomolecules associated noncovalently by multiple weak interactions (such
`as an enzyme andits bound substrate) requires all these interactions to be
`disrupted at the same time. Because the interactions fluctuate randomly,
`such simultaneous disruptions are very unlikely. The molecular stability be-
`stowed by two or five or 20 weak interactions is therefore much greater
`than would be expected intuitively from a simple summation of small bind-
`ing energies.
`Macromolecules such as proteins, DNA, and RNA contain so manysites
`of potential hydrogen bonding or ionic, van der Waals, or hydrophobic in-
`teractions that the cumulative effect of the many small binding forces is
`enormous. For macromolecules, the most stable (native) structure is usu-
`ally that in which weak-bonding possibilities are maximized. The folding of
`a single polypeptide or polynucleotide chain into its three-dimensional
`shape is determined bythis principle. The binding of an antigen to a spe-
`cific antibody depends on the cumulative effects of many weak interactions.
`As noted earlier, the energy released when an enzyme binds noncovalently
`to its substrate is the main source of the enzyme’s catalytic power. The
`binding of a hormone or a neurotransmitter to its cellular receptor protein
`is the result of weakinteractions. One consequence of the large size of en-
`gvmes and receptors Is that their extensive surfaces provide many oppor-
`tunities for weak interactions. At the molecular level, the complementarity
`between interacting biomolecules reflects the complementarity and weak
`interactions between polar, charged, and hydrophobic groups on the sur-
`faces of the molecules.
`
`
`
`

`

`Qo = H,0
`Q= Solute
`
`Forming
`
`ice crystal
`
`(a)
`
`(b)
`
`In pure water,
`every molecule at
`the surface is H,O, and
`all contribute to the
`vapor pressure. Every
`molecule in the bulk
`solution is H,O, and
`can contribute to
`formation of ice crystals.
`
`In this solution, the
`effective concentration
`of HO is reduced; only
`3 of every 4 molecules
`at the surface and in the
`bulk phase are H,O.
`The vapor pressure of
`water andthe tendency
`of liquid water to enter
`a crystal are reduced
`proportionately.
`
`figure 4-9
`Solutes alter the colligative properties of aqueous
`solutions. (a) At 101 kPa (1 atm) pressure, pure water
`boils at 100 °C and freezes at 0 °C. (b) The presence of
`solute molecules reduces the probability of a water mole-
`cule leaving the solution and entering the gas phase,
`thereby reducing the vapor pressure of the solution and
`increasing the boiling point. Similarly, the probability of a
`water molecule colliding with and joining a forming ice
`crystal is reduced when someof the molecules colliding
`with the crystal are solute, not water, molecules. The
`effect is depression of the freezing point.
`
`figure 4-10
`Osmosis and the measurement of osmotic pressure.
`(a) Theinitial state. The tube contains an aqueous solu-
`tion, the beaker contains pure water, and the semiperme-
`able membrane allows the passage of water but not
`solute. Water flows from the beaker into the tube to
`equalize its concentration across the membrane. (b) The
`final state. Water has moved into the solution of the non-
`permeant compound, diluting it and raising the column of
`water within the tube. At equiliorium, the force of gravity
`operating on the solution in the tube exactly balances the
`tendency of water to move into the tube, where its con-
`centration is lower. (c) Osmotic pressure (II) is measured
`as the force that must be applied to return the solution in
`the tube to the level of that in the beaker. This force is
`proportional to the height, h, of the column in (b).
`
`Solutes Affect the Colligative Properties of Aqueous Solutions
`Dissolved solutesofall kinds alter certain physical properties of the solvent,
`water: its vapor pressure, boiling point, melting point (freezing point), and
`osmotic pressure. These are called colligative (“tied together”) proper-
`ties because the effect of solutes on all four properties has the samebasis:
`the concentration of water is lower in solutions than in pure water. Theef-
`fect of solute concentration on the colligative properties of water is inde-
`pendent of the chemical properties of the solute; it depends only on the
`numberof solute particles (molecules, ions) in a given amount of water. A
`compound such as NaCl, which dissociates in solution, has twice the effect
`on osmotic pressure, for example, as an equal number of moles of a nondis-
`sociating solute such as glucose.
`Dissolved solutes alter the colligative properties of aqueous solutions
`by lowering the effective concentration of water. For example, when a sig-
`nificant fraction of the molecules at the surface of an aqueous solution are
`not water but solute, the tendency of water molecules to escape into the va-
`por phase—the vapor pressure—is lowered (Fig. 4—9). Si

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