throbber
TABLE 7-6. TORSION ANGLES (dz, w) FOR RESIDUES 24 T0 73 or HEN EGG WHITE
`LYSOZYME
`
`Residue
`Number
`
`Amino
`Acid
`
`dz (deg)
`
`Residue
`
`Number
`
`Amino
`
`Acid
`
`([2 (deg)
`
`I]! (deg)
`
`Problems
`
`189
`
`Source: lmoto, T., Johnson, L.N., North, A.C.T., Phillips, D.C., and Rupley, J.A., in Boyer, P.D. (Ed.), The
`Enzymes (3rd ed.), Vol. 7, pp. 693 -695, Academic Press (1972).
`
`probable identity of residue 54? (d) What is the secondary
`structure of residues 69 — 7 1? (e) What additional information
`besides the torsion angles, (I) and 1//, of each of its residues are
`required to define the three-dimensional structure of a pro-
`tein?
`
`. Hair splits most easily along its fiber axis, whereas fingernails
`tend to split across the finger rather than along it. What are the
`directions of the keratin fibrils in hair and in fingernails? Ex-
`plain your reasoning.
`
`. What structural features are responsible for the observations
`that 0: keratin fibers can stretch to over twice their normal
`
`length, whereas silk is nearly inextensible?
`
`. What is the growth rate, in turns per second, ofthe ct helices in
`a hair that is growing 15 cm -year"?
`
`. Can polyproline form a collagenlike triple helix? Explain.
`
`. As Mother Nature’s chief engineer, you have been asked to
`design a five-tum a helix that is destined to have half its cir-
`Cumference immersed in the interior ofa protein. Indicate the
`helical wheel projection of your prototype (X helix and its
`amino acid sequence (see Fig. 7-45).
`
`- B.-Aminopropionitrile is effective in reducing excessive scar
`tissue formation after an injury (although its use is contraindi-
`cated by side effects). What is the mechanism of action of this
`lathyrogen?
`
`11. Proteins have been classified as oz, B, a/B, or 0: + fl proteins
`depending on whether their tertiary structures, respectively,
`consist of mostly oz helices, mostly B sheets, alternating a
`helices and B sheets, or some 0: helices and /3 sheets that tend to
`aggregate together rather than alternate along the polypeptide
`chain. By inspection, classify the following proteins according
`to this nomenclature and, where possible, identify their super-
`secondary structures: carboxypeptidase A (Fig. 7-19a), triose
`phosphate isomerase (Fig. 7- 1 9b), myoglobin (Fig. 7-42), con-
`canavalin A (Fig. 7-43), carbonic anhydrase (Fig. 7-44), gly-
`ceraldehyde-3-phosphate dehydrogenase (Fig. 7-47), and
`prealbumin (Fig. 7-59).
`
`. The coat protein of tomato bushy stunt virus consists of 180
`chemically identical subunits, each of which is composed of
`~386 amino acid residues. The probability that a wrong
`amino acid residue will be biosynthetically incorporated in a
`polypeptide chain is 1 part in 3000 per residue. Calculate the
`average number ofcoat protein subunits that would have to be
`synthesized in order to produce a perfect viral coat. What
`would this number be if the viral coat were a single polypep-
`tide chain with the same number of residues that it actually
`has?
`
`. State the rotational symmetry of the following objects: (a) a
`starfish, (b) a square pyramid, (c) a rectangular box, and (d) a
`
`trigonal bipyramid. NPS EX. 2029
`Part 3
`CFAD v. NPS
`
`IPR2015-00990
`
`page 51
`
`Page 51
`
`NPS EX. 2029
`Part 3
`CFAD v. NPS
`IPR2015-00990
`
`

`
`
`
`190 Chapter 7. Three-Dimensional Structures OfProteins
`
`14.
`
`15.
`
`16.
`
`17.
`
`Myoglobin and the subunits of hemoglobin are polypeptides
`of similar size and structure. Compare the expected ratio of
`nonpolar to polar amino acid residues in myoglobin and in
`hemoglobin.
`Why are London dispersion forces always attractive?
`Membrane-bound proteins are generally closely associated
`with the nonpolar groups of lipid molecules (Section 1 1-3A).
`Explain how detergents affect the structural integrity of mem-
`brane bound proteins in comparison to their effects on normal
`globular proteins.
`
`Sickle-cell hemoglobin (HbS) differs from normal human
`adult hemoglobin (HbA) by a single mutational change, Glu
`B6 —> Val, which causes the HbS molecules to aggregate under
`proper conditions (Section 6-3A). Under certain conditions,
`the HbS filaments that form at body temperature disaggregate
`when the temperature is lowered to 0°C. Explain.
`
`18.
`
`19.
`
`Indicate experimental evidence that is inconsistent with the
`hypothesis that urea and guanidinium ion act to denature
`proteins by competing for their internal hydrogen bonds.
`Proteins in solution are often denatured if the solution is
`shaken violently enough to cause foaming. Indicate the mech-
`
`20.
`
`*21.
`
`*22.
`
`anism of this process. (Hint: The nonpolar groups of deter-
`gents extend into the air at air—water interfaces.)
`An oligomeric protein in a dilute buffer at pH 7 dissociates to
`its component subunits when exposed to the following agents.
`Which ofthese observations would not support the contention
`that the quaternary structure ofthe protein is stabilized exclu-
`sively by hydrophobic interactions? Explain. (a) 6M guanidi-
`nium chloride, (b) 20% ethanol, (c) 2M NaCl, (d) tempera-
`tures below 0°C, (e) 2-mercaptoethanol, (f) pH 3, and (g)
`0.01M SDS.
`
`What are the relative amounts of each isozyme formed when
`equimolar amounts ofpure heart and muscle lactate dehydro-
`genase (H4 and M,,) are hybridized?
`The SDS—polyacrylamide gel electrophoresis of a protein
`yields two bands corresponding to molecular masses of 10 and
`17 kD. After cross-linking this protein with dimethylsuberi-
`midate under sufficient dilution to eliminate intermolecular
`cross-linking, SDS—polyacrylamide gel electrophoresis of the
`product yields 12 bands with molecular masses 10, 17, 20, 27,
`30, 37, 40, 47, 54, 57, 64, and 74 kD. Assuming that dimethyl-
`suberimidate can cross-link only contacting subunits, dia-
`gram the quaternary structure of the protein.
`
`
`
`
`
`‘-:l.n.I|—...|.:uni-nun.-J-n,..>==.-_-.....
`
`'.;n-‘lg
`
`Page 52
`
`Page 52
`
`

`
`CHAPTER
`
`
`
`1
`
`Enzymatic Catalysis
`
`1. Catalytic Mechanisms
`A. Acid—Base Catalysis
`B. Covalent Catalysis
`C. Metal Ion Catalysis
`D. Electrostatic Catalysis
`E. Catalysis through Proximity and Orientation Effects
`F. Catalysis by Preferential Transition State Binding
`
`2. Lysozyme
`A. Enzyme Structure
`B. Catalytic Mechanism
`C. Testing the Phillips Mechanism
`3. Serine Proteases
`
`A. Kinetics and Catalytic Groups
`B. X-Ray Structures
`C. Catalytic Mechanism
`D. Testing the Catalytic Mechanism
`E. Zymogens
`4. Glutathione Reductase
`
`Enzymes, as we have seen, cause rate enhancements that
`are orders of magnitude greater than those of the best
`chemical catalysts. Yet, they operate under mild conditions
`and are highly specific as to the identities of both their
`substrates and their products. These catalytic properties are
`so remarkable that many nineteenth century scientists con-
`cluded that enzymes have characteristics that are not
`shared by substances of nonliving origin. To this day, there
`are few enzymes for which we understand in more than
`cursory detail how they achieve their enormous rate accel-
`erations. Nevertheless, it is now abundantly clear that the
`catalytic mechanisms employed by enzymes are identical
`
`371
`
`to those used by chemical catalysts. Enzymes are simply
`better designed.
`In this chapter we consider the nature of enzymatic catal-
`ysis. We begin by discussing the underlying principles of
`chemical catalysis as elucidated through the study of or-
`ganic reaction mechanisms. We then embark on a detailed
`examination of the catalytic mechanisms of several of the
`best characterized enzymes: lysozyme, the serine proteases,
`and glutathione reductase. Their study should lead to an
`appreciation of the intracacies of these remarkably elficient
`catalysts as well as of the experimental methods used to
`elucidate their properties.
`
`1. CATALYTIC MECHANISMS
`
`Catalysis is a process that increases the rate at which a
`reaction approaches equilibrium. Since, as we discussed in
`Section 13-1C, the rate of a reaction is a function of its free
`energy of activation (AG T ), a catalyst acts by lowering the
`height of this kinetic barrier; that is, a catalyst stabilizes the
`transition state with respect to the uncatalyzed reaction.
`There is, in most cases, nothing unique about enzymatic
`mechanisms of catalysis in comparison to nonenzymatic
`mechanisms. What apparently make enzymes such power-
`ful catalysts are two related properties: their specificity of
`substrate binding combined with their optimal arrangement
`of catalytic groups. An enzyme’s arrangement of binding
`and catalytic groups is, of course, the product of eons of
`evolution: Nature has had ample opportunity to “fine
`tune” the performances of most enzymes.
`
`Page 53
`
`Page 53
`
`

`
`372 Chapter 14. Enzymatic Catalysis
`
`The types of catalytic mechanisms that enzymes employ
`have been classified as:
`
`In aqueous solvents, the initial rate of mutarotation of(1-1)-
`glucose, as monitored by polarimetry (Section 4-2A), is ob-
`served to follow the relationship:
`
`v:
`
`_ d[a-D-glucose]
`dt
`
`= obs [oz-D-glucose]
`
`[l4. 1]
`
`where kob, is the reaction’s apparent first-order rate con-
`stant. The mutarotation rate increases with the concentra-
`
`tions of general acids and general bases; they are thought to
`catalyze mutarotation according to the mechanism:
`
`H—/ls
`H
`
`H)F\'.B—
`_/
`0L-D-Glucose
`
`H‘A
`—O{ ’{1‘H :B"
`
`J \H
`
`B-D-Glucose
`
`A /
`
`H :l-\'
`
`CH —/via‘ H - B
`_/
`
`Linear form
`
`This model is consistent with the observation that in aprotic
`solvents such as benzene, 2,3,4,6-0-tetramethyl-a-I)-glu-
`cose (a less polar benzene-soluble analog)
`
` H
`
`OCH3
`2,3,4,6-0-Tetramethy1- oi-D-glucose
`
`. Acid—base catalysis.
`
`. Covalent catalysis.
`
`Metal ion catalysis.
`
`. Electrostatic catalysis.
`
`Proximity and orientation effects.
`
`Preferential binding of the transition state complex.
`
`In this section, we examine these various phenomena. In
`doing so we shall frequently refer to the organic model
`compounds that have been used to characterize these cata-
`lytic mechanisms.
`
`A. Acid - Base Catalysis
`
`General acid catalysis is a process in which partial proton
`transfer from a Bronsted acid (a species that can donate
`protons; Section 2-2A) lowers thefree energy ofa reaction ‘s
`transition state. For example, an uncatalyzed keto—enol
`tautomerization reaction occurs quite slowly as a result of
`the high energy of its carbanionlike transition state (Fig.
`14-la). Proton donation to the oxygen atom (Fig. 14-lb),
`however, reduces the carbanion character of the transition
`state, thereby catalyzing the reaction. A reaction may also
`be stimulated by general base catalysis ifits rate is increased
`by partial proton abstraction by a Bransted base (a species
`that can combine with a proton; Fig. 14-1 c). Some reactions
`may be simultaneously subject to both processes: a con-
`certed general acid— base catalyzed reaction.
`
`Mutarotation Is Catalyzed by Acids and by Bases
`The mutarotation of glucose provides an instructive ex-
`ample of acid—base catalysis. Recall that a glucose mole-
`cule can assume either of two anomeric cyclic forms
`through the intermediacy ofits linear form (Section 10- 1 B):
`
`
`
`H
`
`OH
`
`H
`
`OH
`
`oc-D-Glucose
`[0c]]2)° = 112.2°
`
`\‘
`argon
`
`1/
`
`B-D-Glucose
`[0(,]]2)0 = 1s.7°
`
`does not undergo mutarotation. Yet, the reaction is cata-
`lyzed by the addition of phenol, a weak benzene-soluble
`acid, together with pyridine, a weak benzene-soluble base,
`according to the rate equation:
`
`v = k[phenol] [pyridine] [tetramethyl-oz-D-glucose]
`
`[l4.2]
`
`Moreover, in the presence of oz-pyridone, whose acid and
`base groups can rapidly interconvert between two tauto-
`
` H
`
`OH
`Linear form
`
`Page 54
`
`Page 54
`
`

`
`Keto
`
`R
`
`Transition state
`
`Enol
`
`R
`
`5_
`I
`1
`(a) foe» (ll:-=0
`cH2
`CH3"
`H
`H5+
`
`(b)
`
`c=o+ H"-A —> Q--—-O5-H---A_T> c—o—H + A"
`ll
`H.
`CH2
`1’.
`‘EH2
`R
`H
`1118+
`
`
`
`HQ
`
`2
`
`H—A + OH-
`
`R
`
`R
`
`(c)
`
`C=O+:B ——> cl-=0
`
`(IJH2
`H
`
`CH2
`Hm
`156+
`
`FIGURE 14-1. Mechanisms of keto—eno1 tautomerization: (a) uncatalyzed, (b) general acid
`catalyzed, and (c) general base catalyzed.
`
`meric forms and are situated so that they can simulta-
`neously catalyze mutarotation,
`
`on-Pyridone / \
`H/1:‘ Q.)
`jg
`
`H
`
`/C\
`Glucose
`
`/ \
`4/ (Cl)
`TO)
`
`the reaction follows the rate law
`
`v = k’[oz-pyridone][tetramethyl-oz-D-glucose]
`
`[l4.3]
`
`where k’ = 7000M X k. This increased rate constant indi-
`
`cates that at-pyridone does, in fact, catalyze mutarotation in
`a concerted fashion since 1M a-pyridone has the same cata-
`
`
`
`lytic effect as impossibly high concentrations of phenol and
`pyridine (e.g., 70M phenol and 100M pyridine).
`Many types of biochemically significant reactions are
`susceptible to acid and/or base catalysis. These include the
`hydrolysis ofpeptides and esters, the reactions ofphosphate
`groups, tautomerizations, and additions to carbonyl groups.
`The side chains of the amino acid residues Asp, Glu, His,
`Cys, Tyr, and Lys have pK’s in or near the physiologicalpH
`range (Table 4-1) which, we shall see, permits them to act in
`the enzymatic capacity ofgeneral acid and/or base catalysts
`in analogy with known organic mechanisms. Indeed, the
`ability ofenzymes to arrange several catalytic groups about
`their substrates makes concerted acid- base catalysis a
`common enzymatic mechanism.
`
`The RNase A Reaction Incorporates General Acid—Base
`Catalysis
`Bovine pancreatic ribonuclease A (RNase A) provides an
`illuminating example of enzymatically mediated general
`acid—base catalysis. This digestive enzyme functions to hy-
`drolyze RNA to its component nucleotides. The isolation
`of 2’,3’-cyclic nucleotides from RNase A digests of RNA
`
`Page 55
`
`Page 55
`
`

`
`
`
`FIGURE 14-2. The pH dependence of V(m /K,1, in the RNase
`A-catalyzed hydrolysis of cytidine-2’,3’-cyclic phosphate where
`Vfm /Kj, is given in units of M"s“. Analysis of this curve
`(Section 13-4) suggests the catalytic participation of groups with
`pK’s of 5.4 and 6.4. [After del Rosario, E.J. and Hammes, G.G.,
`Biochemistry 8, 1887 (1969).]
`
`in Fig. 8-2). Evidently, the RNase A reaction is a two-step
`process (Fig. 14-3):
`
`1. His 12, acting as a general base, abstracts a proton from
`an RNA 2’-OH group, thereby promoting its nucleo-
`philic attack on the adjacent phosphorus atom while His
`119, acting as a general acid, promotes bond scission by
`protonating the leaving group.
`
`2. The 2’,3’-cyclic intermediate is hydrolyzed through
`what is essentially the reverse of the first step in which
`water replaces the leaving group. Thus His 12 acts as a
`general acid and His 119 as a general base to yield the
`hydrolyzed RNA and the enzyme in its original state.
`
`B. Covalent Catalysis
`
`Covalent catalysis involves rate acceleration through the
`transient formation of a catalyst—substrate covalent bond.
`The decarboxylation of acetoacetate, as chemically cata-
`lyzed by primary amines, is an example of such a process
`(Fig. 14-4). In the first stage of this reaction, the amine
`nucleophilically attacks the carbonyl group of acetoacetate
`to form a Schifl‘ base (imine bond).
`
`374 Chapter 14. Enzymatic Catalysis
`
`indicates that the enzyme mediates the following reaction
`sequence:
`
`
`
`0i
`
`-0 — II’ = 0
`0
`!
`
`(I)
`-0 —l“°—O-—CH2
`0
`H
`
`H
`
`0
`
`Base
`
`H
`
`H
`
`1)‘?/ft)
`-0" wt)
`
`2',3'-Cyclic nucleotide
`
`H20
`H+
`
`
`
`The RNase A reaction exhibits a pH rate profile that peaks
`near pH 6 (Fig. 14-2). Analysis of this curve (Section 13-4),
`together with chemical derivatization and X-ray studies,
`indicates that RNase A has two essential His residues, His
`12 and His 1 19, which act in a concerted manner as general
`acid and base catalysts (the structure ofRNase A is sketched
`
`H
`
`H
`H
`+ \(‘=O —‘—lE1"—é-Gin —“*—IL'+— r‘;
`—N‘’
`"\r--19" V "
`I
`E H \ +
`E‘
`{V
`Schiff
`:13
`H—A
`base
`
`I‘:
`
`H
`
`Page 56
`
`Page 56
`
`

`
`Section 14- 1. Catalytic Mechanisms 375
`
`The protonated nitrogen atom ofthe covalent intermediate
`then acts as an electron sink (Fig. 14-4, bottom) so as to
`reduce the otherwise high-energy enolate character of the
`transition state. The formation and decomposition of the
`Schifi" base occurs quite rapidly so that it is not rate deter-
`mining in this reaction sequence.
`
`Covalent Catalysis Has Both Nucelophilic and
`Electrophilic Stages
`As the preceding example indicates, covalent catalysis
`may be conceptually decomposed into three stages:
`
`1. The nucleophilic reaction between the catalyst and the
`substrate to form a covalent bond.
`
`2. The withdrawal of electrons from the reaction center by
`the now electrophilic catalyst.
`
`3. The elimination of the catalyst, a reaction that is essen-
`tially the reverse of stage 1.
`
`Reaction mechanisms are somewhat arbitrarily classified as
`occurring with either nucleophilic catalysis, or electrophilic
`catalysis depending on which of these effects provides the
`greater driving force for the reaction, that is, which cata-
`lyzes its rate-detennining step. The primary amine-cata-
`lyzed decarboxylation of acetoacetate is clearly an electro-
`philically catalyzed reaction since its nucleophilic phase,
`Schiff base formation, is not its rate-determining step. In
`other covalently catalyzed reactions, however, the nucleo-
`philic phase may be rate determining.
`The nucleophilicity of a substance is closely related to its
`basicity. Indeed, the mechanism of nucleophilic catalysis
`resembles that of general base catalysis except that, instead
`of abstracting a proton from the substrate, the catalyst nu-
`cleophilically attacks it so as to fonn a covalent bond. Con-
`sequently, if covalent bond formation is the rate-determin-
`ing step of a covalently catalyzed reaction, the reaction rate
`tends to increase with the covalent catalyst’s basicity (pK).
`An important aspect of covalent catalysis is that the more
`stable the covalent bond formed, the less facilely it will
`decompose in the final steps of a reaction. A good covalent
`catalyst must therefore combine the seemingly contradic-
`tory properties of high nucleophilicity and the ability to
`form a good leaving group, that is, to easily reverse the bond
`formation step. Groups with high polarizabilities (highly
`mobile electrons), such as imidazole and thiol functions,
`have these properties and hence make good covalent cata-
`lysts.
`
`Certain Amino Acid Side Chains and Coenzymes Can
`Serve as Covalent Catalysts
`Enzymes commonly employ covalent catalytic mecha-
`nisms as is indicated by the large variety of covalently
`linked enzyme— substrate reaction intermediates that have
`been isolated. For example, the enzymatic decarboxylation
`ofacetoacetate proceeds, much as described above, through
`Schifl base formation with an enzyme Lys residue’s
`
`Page 57
`
`
`
`FIGURE 14-3. The bovine pancreatic RNase A-catalyzed
`hydrolysis of RNA is a two-step process with the intermediate
`formation of a 2’,3’-cyclic nucleotide.
`
`Page 57
`
`

`
`376 Chapter 14. Enzymatic Catalysis
`
`FIGURE 14-4. The uncatalyzed reaction
`mechanism for the decarboxylation of
`acetoacetate (top) and the reaction
`mechanism as catalyzed by primary amines
`(bottom).
`
`0 l
`
`l
`CH3— c— CH3
`
`Acetone
`
`1
`(.03
`
`C O-
`/‘v
`I
`CH3— c= CH2
`
`Enolate
`
`+
`
`H
`I
`
`RNH,
`
`R
`
`II
`
`+
`
`2
`
`I
`
`i
`
`C?‘
`CH3— c =k-EH2
`
`L.
`
`CH3—C—CH3
`
`/0
`“if
`/"'—"'\
`[
`GIL — {‘. — C1-l'.-.-- (I
`.5
`-
`\,10
`
`Acetoacetate
`
`R N I I .3
`
`on "
`
`|.-
`
`,1:
`N.
`ll
`/
`cH_.,—c—”cii._.\—c
`
`o
`
`’1o
`
`I \
`
`chapters in conjunction with discussions of specific enzyme
`mechanisms.
`
`Metal Ions Promote Catalysis through Charge
`Stabilization
`
`In many metal ion—catalyzed reactions, the metal ion
`acts in much the same way as a proton to neutralize nega-
`tive charge, that is, it acts as a Lewis acid. Yet, metal ions are
`often much more eflective catalysts than protons because
`metal ions can be present in high concentrations at neutral
`pH ‘s and can have charges > + I . Metal ions have therefore
`been dubbed “superacids.”
`The decarboxylation of dimethyloxaloacetate, as cata-
`lyzed by metal ions such as Cu“ and Ni“ , is a nonenzyma-
`tic example of catalysis by a metal ion:
`
`'Ml1+
`
`—o'
`0 CH 0
`\C <7 é 3//
`//
`KW v'\
`0
`CH3 0-
`
`Dimethyloxaloacetate
`
`Schiff base
`
`oz-amino group. The covalent intermediate, in this case, has
`been isolated through NaBH4 reduction ofits imine bond to
`an amine, thereby irreversibly inhibiting the enzyme. Other
`enzyme functional groups that participate in covalent ca-
`talysis include the imidazole moiety of His, the thiol group
`of Cys, the carboxyl function of Asp, and the hydroxyl
`group of Ser. In addition, several coenzymes, most notably
`thiamine pyrophosphate (Section 16-3B) and pyridoxal
`phosphate (Section 24-IA), function in association with
`their apoenzymes mainly as covalent catalysts.
`
`C. Metal Ion Catalysis
`
`Nearly one third ofall known enzymes require the presence
`of metal ions for catalytic activity. There are two classes of
`metal ion-requiring enzymes that are distinguished by the
`strengths of their ion — protein interactions:
`
`1. Metalloenzymes contain tightly bound metal ions, most
`commonly transition metal ions such as Fe“, Fe”,
`Cu“, Zn“, Mn“ or C0”.
`
`2. Metal-activated enzymes loosely bind metal ions from
`solution, usually the alkali and alkaline earth metal ions
`Na+, K+, Mg“ , or Ca“.
`
`Metal ions participate in the catalytic process in three
`major ways:
`
`1 . By binding to substrates so as to orient them properly for
`reaction.
`
`2. By mediating oxidation—reduction reactions through
`reversible changes in the metal ion’s oxidation state.
`
`3. By electrostatically stabilizing or shielding negative
`charges.
`
`In this section we shall be mainly concerned with the
`third aspect of metal ion catalysis. The other forms of en-
`zyme-mediated metal ion catalysis are considered in later
`
`co,
`
`CH
`
`3
`
`CH3
`
`_
`
`MIl+
`
`-o
`
`'o
`
`\c C/:50/I
`
`//
`0
`
`U \
`
`H+
`
`CH3
`
`0
`"0
`/
`ll
`\
`c— c — OH
`//
`\
`0
`
`CH3
`
`+ M~+
`
`Page 58
`
`Page 58
`
`

`
`
`
`Section 14- 1. Catalytic Mechanisms 377
`
`in a process facilitated through general base catalysis
`most probably by His 64 (Fig. 14-5). Although His 64 is
`too far away from the Zn“-bound water to directly re-
`move its proton, these entities are linked by two inter-
`vening water molecules to form a hydrogen bonded net-
`work that is thought to act as a proton shuttle.
`
`2. The resulting Zn“-bound OH‘ nucleophilically attacks
`the nearby enzymatically bound CO2, thereby convert-
`ing it to HCO§.
`
`Im
`|
`Im—z|n4+~-<|,~
`Im ll
`
`O
`||
`(If
`(0
`
`+
`
`P“
`Im—Zn2'+'
`|
`Im
`
`0//
`C
`\
`
`0‘
`
`0
`|
`I-I
`
`Wk H20
`
`Im
`0
`I
`,+
`//
`Im—Z|n—(‘)“ + H+ + H " O "C\
`Im H
`
`0-
`
`Im = imidazole
`
`3. The catalytic site is then regenerated by the binding and
`ionization of another H20 to the Zn“, possibly before
`the departure of the HCO; ion, so as to transiently form
`a 5-coordinated Zn“ complex.
`
`Metal Ions Promote Reactions through Charge Shielding
`Another important enzymatic function of metal ions is
`charge shielding. For example, the actual substrates of ki-
`nases (phosphoryl-transfer enzymes utilizing ATP) are
`Mg2+—ATP complexes such as
`
`\l:_'_{
`
`Adenine
`
`Ribose
`
`if
`0 fl’
`O
`
`O
`
`if
`fl’
`O
`
`O
`
`(I)-
`E O’
`0
`
`rather than just ATP. Here, the Mg“ ion’s role, in addition
`to its orienting effect, is to shield electrostatically the nega-
`tive charges of the phosphate groups. Otherwise,
`these
`charges would tend to repel the electron pairs of attacking
`nucleophiles, especially those with anionic character.
`
`D. Electrostatic Catalysis
`
`The binding of substrate generally excludes water from an
`enzyme’s active site. The local dielectric constant of the
`active site therefore resembles that in an organic solvent,
`where electrostatic interactions are much stronger than
`
`Page 59
`
`Here the metal ion (M"+), which is chelated by the dimethy-
`loxaloacetate, electrostatically stabilizes the developing en-
`olate ion of the transition state. This mechanism is sup-
`ported by the observation that acetoacetate, which cannot
`form such a chelate, is not subject to metal ion—catalyzed
`decarboxylation. Most enzymes that decarboxylate oxalo-
`acetate require a metal ion for activity.
`
`Metal Ions Promote Nucleophilic Catalysis via Water
`Ionization
`
`A metal ion ‘s charge makes its bound water molecules
`more acidic than free H20 and therefore a source of OH‘
`ions even below neutral pH ‘s. For example, the water mole-
`cule of (NH3)5Co3+(H2O) ionizes according to the reaction:
`
`(NH3)5Co3+(H2O) :‘ (NH3)5Co3+(OH‘) + H’“
`
`with a pK of 6.6, which is some 9 pH units below the pK of
`free H2O. The resulting metal ion — bound hydroxyl group is
`a potent nucleophile.
`An excellent example of this phenomenon occurs in the
`catalytic mechanism of carbonic anhydrase (Section 9- 1 C),
`a widely occurring enzyme that catalyzes the reaction:
`
`C0, + H20 : Hco; + H+
`
`Carbonic anhydrase contains an essential Zn“ ion that is
`implicated in the enzyme’s catalytic mechanism as follows:
`
`1. The crystal structure of human carbonic anhydrase (Fig.
`7-44) reveals that its Zn“ lies at the bottom of a 15-A-
`deep active site cleft, where it is tetrahedrally coordi-
`nated by three evolutionarily invariant His side chains
`and a H20 molecule. This Zn“-polarized H20 ionizes
`
`
`
`
`
`Glu 106
`
`GIU 117
`
`FIGURE 14-5. The active site of human carbonic anhydrase.
`The light grey ligand to the Zn“ indicates the probable fifth
`211“ coordination site. The arrow points towards the opening of
`the active site cavity. [After Sheridan, R. P. and Allen, L. C.,
`1- Am. Chem. Soc. 103, 1545 (1981).]
`
`Page 59
`
`

`
`Here k}, the pseudo-first-order rate constant, is 0.0018 s“‘
`when [imidazole] = 1M (d) = phenyl). However, for the
`intramolecular reaction
`
`3?
`C—t': ~ ¢N0._,
`
`N:
`
`l
`./_
`‘QR; _,.N‘l-l
`
`k2
`—?>
`
`‘H’
`C
`\
`N1;
`W
`:1’
`\\_/N1-I
`
`.
`+ IJr-
`
`,
`£L‘iVL|._.
`
`the first-order rate constant k, = 0.043 s”‘.; that is, k2 =
`24 kg. Thus, when the 1M imidazole catalyst is covalently
`attached to the reactant, it is 24-fold more efl‘ective than
`when it is free in solution; that is, the imidazole group in the
`intramolecular reaction behaves as if its concentration is
`24M. This rate enhancement has contributions from both
`
`proximity and orientation.
`
`Proximity Alone Contributes Relatively Little to Catalysis
`Let us make a rough calculation as to how the rate of a
`reaction is affected purely by the proximity of its reacting
`groups. Following Daniel Koshland’s treatment, we shall
`make several reasonable assumptions:
`
`1. Reactant species, that is, functional groups, are about
`the size of water molecules.
`
`2. Each reactant species in solution has 12 nearest-neigh-
`bor molecules, as do packed spheres of identical size.
`
`3. Chemical reactions occur only between reactants that
`are in contact.
`
`4. The reactant concentration in solution is low enough so
`that the probability of any reactant species being in si-
`multaneous contact with more than one other reactant
`
`molecule is negligible.
`
`Then the reaction:
`
`A + B “—'> A—B
`
`obeys the second-order rate equation
`
`_ d[A—B]
`_
`dt
`
`v
`
`=
`
`= k2[A9B]pairs
`
`where [A,B],,,,,-,3 is the concentration of contacting mole-
`cules of A and B. The value of this quantity is
`
`[A9B]pairs
`
`_ 12[A][B]
`‘ 55.5M
`
`[
`
`since there are 12 ways that A can be in contact with B, and
`[A]/55.5M is the fraction of sites occupied by A in water
`solution ([H2O] = $5.5M in dilute aqueous solutions) and
`hence the probability that a molecule of B will be next to
`one of A. Combining Eqs. [14.5] and [l4.6] yields
`
`v = k.(%§)[A,B1,,..»,. = 4.6k.[A,B1,..,.
`
`£14.71
`
`Page 60
`
`378 Chapter 14. Enzymatic Catalysis
`
`they are in aqueous solutions (Section 7-4A). The charge
`distribution in a medium of low dielectric constant can
`greatly influence chemical reactivity. Thus, as we have
`seen, the pK’s of amino acid side chains in proteins may
`vary by several units from their nominal values (Table 4-1)
`because of the proximity of charged groups.
`Although experimental evidence and theoretical analy-
`ses on the subject are still sparse, there are mounting indica-
`tions that the charge distributions about the active sites of
`enzymes are arranged so as to stabilize the transition states
`of the catalyzed reactions. Such a mode of rate enhance-
`ment, which resembles the form of metal ion catalysis dis-
`cussed above, is termed electrostatic catalysis. Moreover,
`in several enzymes, these charge distributions apparently
`serve to guide polar substrates towards their binding sites so
`that the rates ofthese enzymatic reactions are greater than
`their apparent diflusion-controlled limits (Section I3-2B).
`
`E. Catalysis through Proximity and
`Orientation Eflects
`
`Although enzymes employ catalytic mechanisms that re-
`semble those of organic model reactions, they are far more
`catalytically eflicient than these models. Such efliciency
`must arise from the specific physical conditions at enzyme
`catalytic sites that promote the corresponding chemical re-
`actions. The most obvious efl'ects are proximity and orien-
`tation: Reactants must come together with theproper spatial
`relationship for a reaction to occur. For example, in the
`bimolecular reaction of imidazole with p-nitropheny1ace-
`tate,
`
`.'-!|J',
`
`Un
`
`CH;}T(1
`
`/1
`ll‘
`(' p-Nitrophenylacetate
`
`-"~'..
`
`.
`1r
`\\
`‘— H
`[midazole
`
`CH3\C +
`N
`
`II
`
`5:
`
`N--.
`
`V+E
`
`-—- NH
`
`p-Nitrophenolate
`
`N-Acetylimidazolium
`
`the progress of the reaction is conveniently monitored by
`the appearance ofthe intensely yellowp-nitrophenolate ion:
`
`61 [I7-N02¢0']
`dt
`
`= k,[imidazole][p-NO2d>Ac]
`= k{[p-NO2d>Ac]
`
`[l4.4]
`
`Page 60
`
`

`
`‘T
`
`
`
`Section 14 — I . Catalytic Mechanisms 379
`
`Thus, in the absence of other effects, this model predicts
`that for the intramolecular reaction,
`
`TABLE 14-1. RELATIVE RATES OF ANIIYDRIDE
`FORMATION FOR ESTERS POsEssING DIFFERENT DEGREES
`OF MOTIONAL FREEDOM IN THE REACTION:
`
`I3 2
`A 13::-A B
`
`1;:
`
`k2 = 4.6k, , which is a rather small rate enhancement. Fac-
`tors that will increase this value other than proximity alone
`clearly must be considered.
`
`Arresting Reactants’ Relative Motions and Properly
`Orienting Them Can Result in Large Catalytic Rate
`Enhancements
`
`The foregoing theory is, of course, quite simple. For ex-
`ample, it does not take into account the motions of the
`reacting groups with respect to one another. Yet, in the
`transition state complex, the reacting groups have little rel-
`ative motion. In fact, as Thomas Bruice demonstrated, the
`rates of intramolecular reactions are greatly increased by
`arresting a molecule’s internal motions (Table 14-1). Thus,
`when an enzyme brings two molecules together in a bimo-
`lecular reaction, as William Jencks pointed out, not only
`does it increase their proximity, but it freezes out their rela-
`tive translational and rotational motions (decreases their
`entropy), thereby enhancing their reactivity. Table 14-1 in-
`dicates that such rate enhancements can be enormous.
`
`Another efl"ect that we have neglected in our treatment of
`proximity is that of orientation. Molecules are not equally
`reactive in all directions as Koshland’s simple theory as-
`sumes. Rather, they react most readily only ifthey have the
`proper relative orientation (Fig. 14-6). For example, in an
`SN2 (bimolecular nucleophilic substitution) reaction, the
`incoming nucleophile optimally attacks its target along the
`direction opposite to that of the bond to the leaving group
`(backside attack). The approaches of reacting atoms along a
`trajectory that deviates by as little as 10 ° from this optimum
`direction results in a significantly reduced reactivity. It has
`been estimated that properly orientating substrates can in-
`crease reaction rates by a factor of up to ~100. In a re-
`lated phenomenon, a molecule may be maximally reactive
`only when it assumes a conformation that aligns its various
`orbitals in a way that minimizes the electronic energy of
`
`
`
`Productive
`
`Unproductive
`
`FIGURE 14-6. Molecules are susceptible to chemical attack
`over only limited regions of their surfaces (represented by the
`colored areas). Without the proper relative orientation (left),
`reactions do not occur (right).
`
`0
`ll
`R,—C--O
`
`R —C—O“
`2
`ll
`O
`
`Br
`
`0
`
`/
`R,—C\
`
`R -—
`2
`
`\0
`
`Reactants"
`
`Relative Rate Constant
`
`CH3COO<t>Br
`+
`
`CH3COO“
`COO¢>Br
`
`:§
`
`s-E
`
`0
`
`CO0‘
`
`C00:;llBr
`
`C00"
`
`COO</>Br
`
`CO0’
`
`1.0
`
`~1 X 103
`
`~2.2 X 10-‘
`
`~5 X 107
`
`" Curved arrows indicate rotational degrees of freedom.
`Source: Bruice, T.C., Annu. Rev. Biochem. 45, 353 (1976).
`
`its transition state, an elfect
`assistance.
`
`termed stereoelectronic
`
`Enzymes, as we shall see in Sections 14-2 and 14-3, bind
`substrates in a manner that both immobilizes them and
`
`aligns them so as to optimize their reactivities. The free
`energy required to do so is derivedfrom the specific binding
`energy ofsubstrate to enzyme.
`
`F. Catalysis by Preferential Transition State
`Binding
`
`The rate enhancements effected by enzymes are often
`greater than can be reasonably accounted for by the cata-
`lytic mechanisms so far discussed. However, we have not
`yet considered one of the most important mechanisms of
`enzymatic catalysis: The binding ofthe transition state to an
`enzyme with greater aflinity than the corresponding sub-
`strates or products. When taken together with the previ-
`ously described catalytic mechanisms, preferential transi-
`tion state binding rationalizes the observed rates of
`enzymatic reactions.
`The original concept of transition stat

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