throbber
Acc. CTwm. Res. 1996, £9, 331-339
`
`331
`
`Carbonic Anhydrase: Evolution of the Zinc Binding Site by
`Nature and by Design
`
`Department o£ Chemistzy, University of Pennsylvania, Philadelphia, Pennsylvania 19104 6323
`
`DAVID W. CHRISTIANSON*
`
`Department of Biochemistry, Box 3711, Duke University Medical Centez; Durham, North Carolina 27710
`
`Received December 18, 1995
`
`CAROL A. FIERKE*
`
`Ever since its discovery at the University of Penn-
`sylvaniaJ and the University of Cambridge2 more than
`60 years ago, the zinc enzyme carbonic anhydrase has
`occupied a prominent position at the frontiers of
`biochemistry, medicinal chemistry, and protein engi-
`neering. Seven genetically-distinct forms of this en-
`zyme (known as isozymes I-VII) have evolved in
`numerous tissues and cellular locations, and each
`contains a catalytically-obligatory zinc ion.3,4 This
`diversity reflects the ubiquitous biological requirement
`for rapid hydration of carbon dioxide to yield bicarbon-
`ate ion plus a proton.5-9 Although a deceptively
`simple reaction, the chemical and structural aspects
`of this mechanism have only recently been delineated
`for an isozyme found in the red blood cell, carbonic
`anhydrase II (CAII) (Figure 1). The three-dimensional
`structure of CAIIm has stimulated research probing
`the determinants of the substrate association site and
`the pathway the product proton traverses through the
`enzyme active site. On the basis of phylogenetic
`comparisons,3,4 the substrate and zinc binding sites
`are mainly conserved among all catalytic carbonic
`anhydrase isozymes found in n~ammals. However, the
`trajectory of catalytic proton transfer to bulk solvent
`has diverged during the evolution of the seven
`isozymes.
`We begin this Account with a brief review of the
`CAII mechanism, emphasizing recent developments
`(see previous Accounts for additional information6,r).
`Notably, more X-ray crystallographic and enzymologi-
`cal studies have been performed on CAII and its site-
`specific variants than any other metalloenzyme, and
`these studies uniquely illuminate the molecular de-
`tails of catalysis. Importantly, this work provides an
`elegant example of how the complementary methods
`
`After receiving A.B., A.M., and Ph.D. degrees in Chemistry from Harvard University,
`David W. Christianson moved to the University of Pennsylvania in 1988 where he is
`currently Associate Professor of Chemistry. While at Penn he received a Searle
`Scholar Award, the Young Investigator Award from the Office of Naval Research, an
`Alfred P. Sloan Research Fellowship, and a Camille and Henry Dreyfus Teacher-
`Scholar Award. As a protein crystallographer, Christianson has spent more than a
`decade studying structure-function relationships in metalloenzymes such as carbonic
`anhydrase, and a 1989 Account outlines his work on the related zinc enzyme
`carboxypeptidase A.
`Carol A. Fierke received her B.A. and Ph.D. degrees from Carleton College and
`Brandeis University, respectively. After a postdoctoral fellowship in chemistry at
`Pennsylvania State University she joined the faculty of the Duke University Medical
`School where she is currently Associate Professor of Biochemistry and Chemistry
`as well as a member of the Duke University Cancer Center. While at Duke she
`received an American Heart Association Established Investigator Award, a David
`and Lucile Packard Foundation Fellowship, and an American Cancer Society Junior
`Faculty Research Award. During her career she has investigated molecular
`determinants of biological catalysis in metalloenzymes such as carbonic anhydrase
`and ribonuclease P.
`
`of molecular biology and structural biology can be used
`to probe the structure, function, and stability of a zinc
`binding site in a metalloenzyme. Additionally, these
`studies set a useful foundation for understanding the
`evolution of carbonic anhydrase into noncatalytic
`biological roles, such as signal transduction. For
`example, in the nervous system an extracellular, CAII-
`like domain of receptor protein tyrosine phosphatase
`f! binds an axonal cell recognition molecule (contactin)
`important for neuronal development and differ-
`entiation. J J- J 5 Intriguingly, the putative zinc binding
`site of this domain has partially evolved away from
`that found in CAII.
`We then review recent progress in the "directed
`evolution"--i.e., the structure-based redesign--of the
`CAII zinc binding site, following Nature’s example. We
`describe novel structural determinants of protein-
`metal affinity and the chemical reactivity of zinc-
`bound solvent. Not only does this work represent the
`first molecular dissection of structure-function rela-
`tionships in a protein-zinc binding site (and therefore
`serves as a paradigm for the design of de novo metal
`sites in other proteins), it also sets the foundation for
`the development and optimization of CAII as a metal
`ion biosensor. Remarkably, this ubiquitous metallo-
`protein can be engineered and exploited as a sensitive
`tool for analytical chemistry and biotechnology.
`
`(1) Stadie, W. C.; O’Brien, H. J. Bio]. Chem. 1933, 103, 521-529.
`(2) Meldrum, N. U.; Roughton, F. J. W. ~. Physiol. 1933, 80, 113-
`142.
`(3) Tashian, R. E. Adv. Genet. 1982, 30, 321-356.
`(4) Hewett Emmett, D.; Tashian, R. E. Moi. Phyiogenet. ~voi. 1996,
`5, 50-77.
`(5) Coleman, J. E. J. Biol. Chem. 1967, 242, 5212-5219.
`(6) Bertini, I.; Luchinat, C. Ace Chem. Res. 1983, 16, 272-279.
`(7) Silverman, D. N.; Lindskog, S. Acc. Chem. Res. 1988, 21, 30-36.
`(8) Lindskog, S.; Liljas, A. Cur±: Op#~. Ntruct. Biol. 1993, 3, 915-
`920.
`(9) Silverman, D. N. Methods ~nzymol. 1995, 249, 479-503.
`(10) Liljas, A.; Kannan, K. K.; Bergsten, P. C.; Waara, I.; Fridborg,
`K.; Strandberg, B.; Carlbom, U.; Jarup, L.; Lovgren, S.; Peter, M. Nature
`NewBiol. 197Z, 235, 131-137.
`(11) Krueger, N. X.; Saito, H. Proc. Natl. Acad NcL {ZN.A. 199Z, 89,
`7417-7421.
`(12) Barnea, G.; Silvennoinen, O.; Shaanan, B.; Honegger, A. M.;
`Canoll, P. D.; D’Eustachio, P.; Morse, B.; Levy, J. B.; LaForgia, S.;
`Huebner, K.; Musacchio, J. M.; Sap, J.; Schlessinger, J. Mol. Cell. Biol.
`1993, 13, 1497-1506.
`(13) Levy, J. B.; Cannol, P. D.; Silvennoine, O.; Barnea, G.; Morse,
`B.; Honegger, A. M.; Haung, J. T.; Cannizzaro, L. A.; Park, S. H.; Druck,
`T.; Huebner, K.; Sap, J.; Ehrlich, M.; Musacchio, J. M.; Schlessinger, J.
`~L Biol. Chem. 1993, 26& 10573-10581.
`
`(14) Barford, D.; Jia, Z.; Tonks, N. K. Nature St±~ct. Biol. 1995, 2,
`1043-1053.
`(15) Peles, E.; Nativ, M.; Campbell, P. L.; Sakurai, T.; Martinez, R.;
`Lev, S.; ClaW, D. O.; Schilling, J.; Barnea, G.; Plowman, G. D.; Grumet,
`M.; Schlessinger, J. Cell 1995, 82, 251-260.
`
`S0001-4842(95)00123-3 CCC: $12.00
`
`© 1996 American Chemical Society
`
`Akermin, Inc.
`Exhibit 1016
`Page 1
`
`

`

`332 Acco CT~em. Res., Vol. 29, No. 7, 1996
`
`C7~ristianson and FierCe
`
`o
`
`Zn2÷
`
`H~s 94
`
`His 119
`H is-96
`
`~Z /
`
`H2°
`
`HC03"
`
`Zrl2+
`
`. ./.\.I
`
`Hm 119
`His-16
`
`H~s 94
`
`.. .
`
`His I~9
`His-96
`
`His 94
`
`(a)
`
`(b)
`
`H
`
`HIS 94
`
`HIS 119
`His-96
`
`H* to buffer via His-64 shuttle
`
`Figure 1. Summary of the CAII mechanism. Zinc bound hydroxide attacks the carbonyl carbon of CO2 to form zinc bound bicarbonate.
`The initial mode of bicarbonate binding (a) may reflect the structure of either a discrete intermediate or the transition state.
`Bicarbonate may then isomerize (b), representing either a productive or a nonproductive complex. Following the exchange of a water
`molecule for zinc bound bicarbonate, a proton is transferred flom zinc bound water to solvent via His 64 to regenerate the zinc
`hydroxide species.
`
`~:.~,....~.." ~_,, ~.~(cid:128) ~.~
`
`,.
`
`"i; li’t’4i " " ’.’" xi~" ~li~°:1;/
`
`,..,,.; .....
`! .... ....
`
`Figure 2. CAII active site, complexed with the competitive inhibitor phenol bound in the hydrophobic locket.31 Important active
`site residues are labeled. Note that inhibitor (and therefore substrate) binding does not require the displacement of zinc bound
`solvent (unlabeled gray sphere). Figure generated with MOLSCRIPT.89
`
`Mechanism of Catalysis: Substrate Binding
`and Proton Transfer
`
`The substrate association site of CAII is a hydro-
`phobic pocket adjacent to zinc-bound hydroxide,1°,16
`formed in large part by residues Val-143 at its base
`and Val- 121, Trp-209, and Leu- 198 at its neck (Figure
`2). This pocket is highly conserved among all active
`isozymes on the basis of phylogenetic comparisons,3,4
`although in isozyme III the pocket is somewhat
`constricted by the bulky side chain of Phe-198.lz The
`structure-based dissection of this pocket in CAII
`delineates the minimum size and shape of the pocket
`required for enzymatic activity; for instance, the
`occlusion of the pocket by the radical Val-143 ~ Phe
`substitution obliterates catalytic activity due to the
`
`(16) Lindskog, S. In Z#2c ~nzymes; Bertini, I., Luchinat, C., Maret,
`W., Zeppezauer, M., Eds.; Birkhauser: Boston, 1986; pp 307-316.
`(17) Eriksson, A. E.; Liljas, A. Proteins: St±~ct., Funct. Genet. 199a,
`16, 29-42.
`
`loss of the substrate association site.18,19 More moder-
`ate amino acid substitutions in the pocket (e.g., Val-
`143 ~ Gly or Val-121 ~ Leu) modestly affect the
`stability of the transition state for CO2 hydration
`(rather than substrate affinity) and alter the structure
`of active site solvent, leading to slight changes in the
`rate constant for proton transfer from zinc-bound
`water to His-64 (see Figure 1). The hydrophobic
`pocket therefore has a minimum width2°-23 and
`depth18,19 for efficient catalysis, and linear free energy
`relationships indicate that the volume of the amino
`
`(18) Alexander, R. S.; Nair, S. K.; Christianson, D. W. BiochemistW
`1991, 30 11064-11072.
`(19) Fierke, C. A.; Calderone, T. L.; Krebs, J. F. ~iochemist±’y 1991,
`30, 11054-11063.
`(20) Nair, S. K.; Calderone, T. L.; Christianson, D. W.; Fierke, C. A.
`~L £~1ol. Chem. 1991, 266, 17320-17325.
`
`(21) Nair, S. K.; Christianson, D. W. £~iochemist±’y 1993, 32, 4506-
`4514.
`(22) Krebs, J. F.; Rana, F.; Dluhy, R. A.; Fierke, C. A. £~iocheraist±y
`1993, 32, 4496-4505.
`(23) Nair, S. K.; Krebs, J. F.; Christianson, D. W.; Fierke, C. A.
`£~iochemist±y 1995, 34, 3981-3989.
`
`Akermin, Inc.
`Exhibit 1016
`Page 2
`
`

`

`Carbonlc Anhydrase
`
`Acc. CTwm. Res., Vol. 29, No. Z 1996 333
`
`acid residue at the base of the pocket is critical for
`activity,~9 whereas the hydrophobicity of residues at
`the neck of the pocket is likewise important.2°,22,23
`These results are consistent with molecular dynam-
`ics simulations that point to the hydrophobic pocket
`as a substrate association site.24-26 Experimental
`evidence supporting the catalytic role of the hydro-
`phobic pocket comes from a variety of spectroscopic
`studies,at,aS including Fourier transform infrared spec-
`troscopy.32,39 The shift of the asymmetric stretching
`vibration of COa m a lower wavenumber upon associa-
`tion with CAII is consistent with the transfer of COa
`from aqueous solution to a hydrophobic environment.
`These experiments also indicate that the affinity of
`CAII for COa is low (~0.1 M), as required by the high
`turnover number of the enzyme which necessitates a
`rapid product dissociation rate constant. Finally,
`although the precatalytic enzyme-substrate complex
`is too short-lived to be observed by traditional X-ray
`crystallographic methods, the structure of the only
`known competitive inhibitor of CAII-catalyzed CO2
`hydration, phenol,3° has been solved in complex with
`the enzyme:a~ phenol binds in the hydrophobic pocket
`and makes van der Waals contacts with Val- 121, Val-
`143, Leu-198, and Trp-209 while its hydroxyl group
`hydrogen bonds with zinc-bound hydroxide (Figure 2).
`Since a competitive inhibitor must bind in the same
`location as the substrate, COa must therefore bind in
`the hydrophobic pocket prior to catalysis. Impor-
`tantly, COa binding does not displace zinc-bound
`hydroxide, although long-range (i.e., >3 A) weakly-
`polar interactions with zinc may contribute to sub-
`strate orientation.
`Product bicarbonate ion is formed by the nucleo-
`philic attack of zinc-bound hydroxide at COa im-
`mobilized in the hydrophobic pocket, and the binding
`mode of bicarbonate ion to zinc has been the subject
`of some controversy. Central to this controversy is the
`possible role of Thr-199 as a "doorkeeper".aa-a4 That
`is to say, the side chain hydroxyl group of this residue
`is proposed to allow only anions capable of donating
`a hydrogen bond to Thr-199 access to zinc binding.
`X-ray crystallographic structures of Co2+-substituted
`and Thr-199 ~ Ala CAIIs in complex with bicarbonate
`are consistent with this interpretation.35,36 However,
`the binding of azide anion, a competitive inhibitor of
`bicarbonate dehydration, demonstrates that Thr- 199
`is not a doorkeeper all the time:37,38 azide anion binds
`to zinc but does not hydrogen bond with Thr-199.
`
`(24) Merz, K. M. ~L Mol. ;3iol. 1990, 214, 799-802.
`(25) Liang, J. Y.; Lipscomb, W. N. Proc Natl. Acad. ScL [ZS.A. 1990,
`87, 3675-3679.
`(26) Merz, K. M. ~L Am. Chem. Soc. 1991, 113, 406-411.
`(27) Williams, T. J.; Henkens, R. W. ;3iochemist±y 1985, 24, 2459-
`2462.
`(28) Bertini, T.; Luchinat, C.; Monnanni, R.; Roelens, S.; Moratal, J.
`M. J. Am. Chem. Soc. 1987, 109, 7855-7856.
`(29) Riepe, M. E.; Wang, J. H. J. ;3~ol. Chem. 1968, 243, 2779-2787.
`(30) Simonsson, I.; Jonsson B. H.; Lindskog, S. ;3ioc!~em. ;3iop!~ys. Res.
`Commun. 1982, 108, 1406-1412.
`(31) Nair, S. K.; Ludwig, P. A.; Christianson, D. W. ft. Am. Chem.
`1994, 116, 3659 3660.
`(32) HAkansson, K.; Carlsson, M.; Svensson, L. A.; Liljas, A. ft.
`;3ioi. 1992, 227, 1192 1204.
`(33) Liljas, A.; HAkansson, K.; Jonsson, B. H.; Xue, Y. Eu±: J. ;3iochem.
`1994, 219, 1-10.
`(34) Lindahl, M.; Svensson, L. A.; Liljas, A. P[’oteJ±~s: Struct.,
`Genet. 1993, 15, 177 182.
`(35) HAkansson, K.; Wehnert, A. ft. Moi. ;3~oi. 1992, 228, 1212-1218.
`(36) Xue, Y.; Liljas, A.; Jonsson, B. H. P±’ote#~s: StFuct., Funct. Genet.
`1993, 17, 93-106.
`(37) Nair, S. K.; Christianson, D. W. ~u±’. J. ;3iochem. 1993, 213, 507-
`515.
`
`Instead, the zinc-bound azide nitrogen makes a non-
`hydrogen-bonded, van der Waals interaction with the
`Thr-199 hydroxyl group. Such a binding mode is
`likewise possible for product bicarbonate, as illus-
`trated in Figure 1. Although such an interaction does
`not contribute to enzyme-product affinity, it is never-
`theless likely m contribute to efficient catalysis by
`facilitating rapid product dissociation. Consistent
`with this interpretation are studies of Thr-199 vari-
`ants showing that deletion of the Thr-199 hydroxyl
`group stabilizes bicarbonate binding39,4° and alters the
`structure of the bound bicarbonate ion.36 Therefore,
`the Thr-199 side chain promotes maximum catalytic
`efficiency by destabilizing the product complex to
`provide for rapid dissociation and by selecting for a
`catalytically competent bicarbonate-zinc complex (Fig-
`ure 1). Thus, it is clear that the zinc binding site and
`its environment have evolved for optimal catalytic
`activity.
`Following bicarbonate dissociation is the rate-
`determining step of proton transfer to regenerate the
`active catalyst, zinc-bound hydroxide (Figure 1).5-9
`The product proton is not transferred from zinc-bound
`water directly to bulk solvent; instead, it is first
`transferred to an intermediate "shuttle" residue, and
`then transferred to bulk solvent. Because of its
`greater exposure to solvent, the shuttle residue can
`more efficiently transfer a proton to a variety of
`acceptors with pK~ values higher than that of water]
`Interestingly, the proton transfer pathway has diver-
`gently evolved among the carbonic anhydrase isozymes.
`In isozyme II, His-64 is the catalytic proton shuttle,41,42
`and it exhibits significant conformational mobility
`which accompanies its function.43-45 This residue is
`too far from zinc-bound solvent m allow for direct
`proton transfer; instead, proton transfer is achieved
`across two intervening, hydrogen-bonded solvent mol-
`ecules in the native enzyme (Figure 3). Isozymes IV,
`VI, and VII also contain His-64, which may similarly
`function as a proton shuttle (isozyme I contains His-
`64, but His-200 is the major proton shuttle group in
`this isozyme46 ). However, isozyme III contains Lys-
`64, and the main proton transfer pathway is direct
`transfer to bulk solvent.4r Intriguingly, isozyme V
`contains Tyr-64, which plays no major role in proton
`transfer48 due in part to steric effects arising from the
`bulky adjacent side chain of Phe-65.49 The three-
`
`(38) Jonsson, B. M.; Hfikansson, K.; Liljas, A. FEBSLelr. 1993, 322,
`186-190.
`(39) Krebs, J. F.; Ippolito, J. A.; Christianson, D. W.; Fierke, C. A. ~L
`;3~oi. Chem. 1993, 268, 27458-27466.
`
`(40) Liang, Z.; Xue, Y.; Behravan, G.; Jonnson, B. H.; Lindskog, S.
`~u±’. d. Biochem. 1993, 211, 821-827.
`(41) Steiner, H.; Jonsson, B. H.; Lindskog, S. Eu±: J. t3iochem. 1975,
`59, 253-259.
`(42) Tu, C. K.; Silvennan, D. N.; Forsman, C.; Jonsson, B.H.;
`Lindskog, S. Biochemist±y 1989, 28, 7913-7918.
`(43) Krebs, J. F.; Fierke, C. A.; Alexander, R. S.; Christianson, D. W.
`Biochemist±y 1991, 30, 9153-9160.
`(44) Nair, S. K.; Christianson, D. W. ~ Am. Chem. $oc. 1991, 113,
`9455-9458.
`(45) Taoka, S.; Tu, C.; Kistler, K. A.; Silverman; D. N. J. Biol. Chem.
`1994, 269, 17988 17992.
`(46) Engstrand, C.; Johnson, B. H.; Lindskog, S. Eur. J. t3iochem.
`1995, 229, 696-702.
`(47) Jewell, D. A.; Tu, C.; Paranawithana, S. R.; Tanhauser, S. M.;
`LoGrasso, P. V.; Laipis, P. J.; Silvennan, D. N. Biochemist±y 1991, 30,
`1484-1490.
`(48) Heck, R. W.; Tanhauser, S. M.; Manda, R.; Tu, C. K.; Laipis, P.
`J.; Silverman, D. N. J. Biol. Chem. 1994, 269, 24742-24746.
`(49) Boriack Sjodin, P. A.; Heck, R. W.; Laipis, P. J.; Silvennan, D.
`N; Christianson, D. W. Proc. N~tl. Acid. Sci. ~S.A. 1995, 92, 10955-
`10959.
`
`Akermin, Inc.
`Exhibit 1016
`Page 3
`
`

`

`334 Acc. CT~em. t~es., Vol. 29, No. 7, 1996
`
`CT~ristianson and Fierke
`
`Lessons learned in the structure-based redesign of
`the CAII zinc binding site are also relevant to the
`design and construction of de novo transition metal
`binding sites in other proteins,5J-57 since only a few
`examples are structurally characterized by atomic
`resolution X-ray crystallographic methods to provide
`ultimate proof-of-design.58,59 Indeed, examples char-
`acterized by other biophysical techniques sometimes
`illustrate the limitations of current structure-based
`design approaches despite the best attempts of com-
`putational structure analysis--for example, where
`"designed" metal ligands do not coordinate to the
`target metal ion,6°’61 or where zinc affinity is at least
`4 orders of magnitude weaker5~ than that measured
`for CAII.62,63 Furthermore, none of the designed sites
`function as efficient catalysts. Hence, there is much
`yet to be learned regarding structure-activity and
`structure-stability relationships in protein-metal
`binding sites, and the zinc binding site of CAII is a
`universally-recognized and easily-characterized para-
`digm. We note that there is a novel application of
`redesigned CAII zinc binding site variants in analyti-
`cal chemistry and biotechnology: a CAII-based metal
`ion biosensor.64 Following Nature’s example in the
`evolution of carbonic anhydrase for different biological
`functions, the "directed evolution", or structure-based
`redesign, of the CAII zinc binding site allows for the
`optimization of its properties65 in the development of
`a zinc biosensor.
`Altering the Direct Zinc Ligands. A thorough
`understanding of three factors must precede the
`molecular dissection and redesign of direct metal
`ligands in any protein-metal binding site. First, the
`chemical nature of the target metal must be consid-
`ered. It could be "hard", like the small, highly-charged
`magnesium ion, it could be "soft", like the large,
`highly-polarizable mercury ion, or it could be inter-
`mediate between these two extremes (i.e., "border-
`line").66 An optimized metal binding site in a protein
`molecule will exhibit hardness complementary to that
`of its target metal ion. Since zinc is a metal of
`borderline hardness, it is satisfactorily coordinated by
`the harder ligands aspartate and glutamate, the
`borderline ligand histidine, and the softer ligand
`cysteine. Second, the conformations of the amino acid
`side chains that comprise the metal coordination site
`
`(51) Tainer, J. A.; Roberts, V. A.; Getzoff, E. D. Cu±’~: Opin. Biotechnol.
`199Z, 3, 378-387.
`(52) Regan, L. Trends Biochem. ScL 1995, 20, 280-285.
`(53) Christianson, D. W. Adv. ])rotein Chem. 1991, 42, 281-355.
`(54) Higaki, J. N.; Fletterick, R. J.; Craik, C. S. TrendsBiochem. ScL
`199Z, 17, 100-104.
`(55) Klemba, M.; Gardner, K. H.; Marino, S.; Clarke, N. D.; Regan,
`L. Nature Stz~ct. Biol. 1995, 2, 368-373.
`(56) Elling, C. E.; Nielsen, S. M.; Schwartz, T. W. Nature 1995, 374,
`74-77.
`(57) Mfiller, H. N.; Skerra, A. Biochemistry 1994, 33, 14126-14135.
`(58) McGrath, M. E.; Haymore, B. L.; Summers, N. L.; Craik, C. S.;
`Fletterick, R. J. BJochemJst±y 1993, 32, 1914-1919.
`(59) Browner, M. F.; Hackos, D.; Fletterick, R. J. Nature Struct. Biol.
`1994, 1, 327-333.
`(60) Hellinga, H. W.; Richards, F. M. J. Mol. Biol. 1991, 222, 763-
`785.
`(61) Hellinga, H. W.; Caradonna, J. P.; Richards, F. M. ~/. Mol. Biol.
`1991, 222, 787-803.
`(62) Lindskog, S.; Nyman, P. O. Biochim. Biophys. Acta 1964, 85, 462-
`474.
`(63) Kiefer, L. L.; Krebs, J. F.; Paterno, S. A.; Fierke, C. A. Biochem
`istzy 1993, 32, 9896 9900.
`(64) Thompson, R. B.; Jones, E. R. Anal. Chem. 1993, 65, 730-734.
`(65) Kiefer, L. L.; Paterno, S. A.; Fierke, C. A. ~ Am. Chem. Soc. 1995,
`117, 6831-6837.
`(66) Pearson, R. G. J. Am. Chem. Soc. 1963 85, 3533-3539.
`
`Figure 3. Proton shuttle His 64 adopts the "in" conforrnation
`in native CAII and engages in a hydrogen bonded solvent
`network (red) with zinc bound solvent at pH 8.5.32 Proton
`transfer across this network regenerates the reactive zinc bound
`hydroxide species in catalysis (Figure 1). Upon protonation at
`lower pH values, His 64 swings away from the active site to
`the "out" conformation.44 Thus, His 64 is conformationally
`mobile, as indicated. Figure generated with MOLSCRIPT.89
`
`dimensional structure of isozyme V suggests that a
`residue in the vicinity of Tyr-131 (e.g., the phenolic
`side chain of this residue or one of three flanking
`lysines) may serve as a proton shuttle group,49 but
`further experiments are necessary to confirm this
`speculation. Rate constants for proton transfer in the
`carbonic anhydrases are dependent on the difference
`in pKa between zinc-bound water and the proton
`shuttle group, and are limited by organization of the
`intervening hydrogen-bonded solvent molecules.9,45,5°
`Individual components of the zinc binding site may
`therefore affect catalytic proton transfer by modulat-
`ing the pKa of zinc-bound water.
`
`Structure-Based Protein Engineering:
`Redesigning the Zinc Binding Site
`
`As summarized in the preceding section and in
`Figure 1, the zinc ion in the CAII active site plays a
`central role in the mechanism of CO2 hydration. The
`principal role of zinc is that of an electrostatic catalyst,
`since it stabilizes the negatively-charged transition
`state leading to bicarbonate formation. It also de-
`presses the pK~ of its bound water molecule to provide
`an ample supply of nucleophilic hydroxide ion for
`catalysis at neutral pH values. Logically, the protein
`environment of zinc is critical for optimizing the
`electrostatic properties of the metal ion for catalysis,
`and this environment includes not only the direct
`metal ligands but also the residues hydrogen bonding
`with these ligands (i.e., "indirect" or "second shell"
`ligands) (Figure 4). By combining the techniques of
`molecular and structural biology, we have dissected
`the determinants of affinity and catalysis in the zinc
`binding site of CAII.
`
`(50) Silverman, D. N.; Tu, C.; Chen, X.; Tanhauser, S. M.; Kresge, A.
`J.; Laipis, P. J. BJochemJstzy 1993, 32, 10757-10762.
`
`Akermin, Inc.
`Exhibit 1016
`Page 4
`
`

`

`CaFbonlc Anhydz~se
`
`Acc. ~7;em. Res., Vol. 29, No. 7, 1996 335
`
`Direct Metal Ligands
`Indirect Metal Ligands
`
`Figure 4. Scheme of the CAII zinc binding site. Variants with characterized properties and three dimensional structures are indicated
`by italics at the site of substitution.
`
`Table 1. Properties of Selected CA Variants with
`Altered Zinc Binding Sitesa
`
`variant
`
`kcatlKM/uM 1 s 1)
`
`wild type
`H94D
`H94C
`H96C
`Hll9C
`H119D
`
`110
`0.11
`0.11
`0.073
`0.11
`3.8
`
`pKa
`
`6.8
`_>9.6
`_> 9.5
`8.5
`nd
`8.6
`
`zinc Kd (nM)
`
`0.004
`15
`33
`60
`50
`25
`
`Data from ref 75. nd = no data.
`
`must be favorable.54,67 If an engineered ligand must
`incur too high an energetic cost to coordinate to a
`metal ion, it will not do so. Even so, certain regions
`of a protein structure--e.g., loops, ~-helices, or f!-
`strands--can be sufficiently pliable to allow for modest
`conformational changes or segmental shifts which
`optimize metal coordination by an engineered ligand.
`Finally, and perhaps most importantly, the separation
`and stereochemistry of ligand-metal coordination
`must be reasonable. The geometric preferences of
`carboxylate,68,69 imidazole,7° and thiolate71 ligands for
`protein-metal ion coordination have been outlined,
`and the results of these studies provide structural
`criteria by which metal binding site designs can be
`evaluated and optimized.
`The alteration of direct zinc ligands in CAII yields
`important structural and functional insight as the
`properties of selected variants (Table 1) are inter-
`preted in light of their three-dimensional struc-
`
`(67) Ponder, J. W.; Richards, F. M. ~. Mol. J3~o1. 1987, 193, 775-791.
`(68) Christianson, D. W.; Alexander, R. S. ~. Am. Chem. Soc. 1989,
`111, 6412-6419.
`(69) Chakrabarti, P. P±’ote#~ Eng. 1990, 4, 49-56.
`(70) Chakrabarti, P. P±’ote#~ ~ng. 1990, 4, 57-63.
`(71) Chakrabarti, P. J3iochemist±y 1989, 28, 6081-6085.
`
`Lures.72-76 As a point of reference, recall that the
`protein ligands to zinc in wild-type CAII are His-94,
`His-96, and His-ll9 (Figures 2 and 4). A notable
`feature in the CAII zinc binding site is that metal
`binding to the histidine ligands is very cooperative:
`the deletion of any one protein ligand by a histidine
`~ alanine substitution decreases the zinc affinity75 by
`a factor of ~105. This is much larger than the 10-
`100-fold decrease in metal affinity observed for remov-
`ing one ligand from Cys2His2 zinc binding sites,
`including a zinc finger peptide77 and a de novo
`designed zinc coordination polyhedron in the 4~-
`helical bundle protein designated Z~4.7~
`Similarly, substitution of a neutral histidine ligand
`by the negatively-charged side chains of aspartate,
`glutamate, or cysteine decreases the zinc affinity ~ 104-
`fold.75 The structure of His-94 ~ Asp CAII reveals
`that this loss of protein-metal affinity arises prima-
`rily from the 0.9 ik movement of zinc that accom-
`modates the shorter side chain of the ligand. Intrigu-
`ingly, the structure of His-94 ~ Cys CAII demon-
`strates that the zinc ion occupies the same site in the
`His-94 ~ Asp and His-94 ~ Cys variants;75,76 ad-
`ditionally, a 1 ik deformation of the f!-strand contain-
`ing Cys-94 ensures that this shortest ligand side chain
`coordinates to zinc (Figure 5). Notably, this f!-strand
`
`(72) Alexander, R. S.; Kiefer, L. L.; Fierke, C. A.; Christianson, D. W.
`;3iochemist±’y 1993, 32, 1510-1518.
`(73) Ippolito, J. A.; Nair, S. K.; Alexander, R. S.; Kiefer, L. L.; Fierke,
`C. A.; Christianson, D. W. ]3±’ote#~ ~ng. 1995, 8, 975-980.
`(74) Kiefer, L. L.; Ippolito, J. A.; Fierke, C. A.; Christianson, D. W. ~
`Am. Chem. $oc. 1993, 115, 12581-12582.
`(75) Kiefer, L. L.; Fierke, C. A. BJochemJsr±’y1994, 33, 15233-15240.
`(76) Ippolito, J. A.; Christianson, D. W. BJochemJsr±y1994, 33, 15241-
`15249.
`(77) Merkle, D. L., Schmidt, M. H.; Berg, J. M. ~ Am. Chem. Soc.
`1991, 113, 5450-5451.
`(78) Klemba, M.; Regan, L. ;3iochemisr±y 1995, 34, 10094-10100.
`
`Akermin, Inc.
`Exhibit 1016
`Page 5
`
`

`

`336 Acc. CTwm. Reso, Vol. £9, No. 7, 1996
`
`C7~rlstlanson and Fler~e
`
`Figure 6. log of the pH independent second order rate constant
`for CO2 hydration versus the zinc-water pKa for wild type and
`variants of CAII,43,65,r~ including indirect ligand substitutions
`at position 92 (Gln ~ Ala, Leu, AsH, Glu) and position 117 (Glu
`~ Asp, Ala) (@), direct ligand substitutions (His 94 ~ Cys, His
`94 ~ Asp, His 119 ~ Asp) (~), and position 199 variants (Thr
`~ Set, Ala, Val, Pro) (~). The data are fit to a line yielding R =
`0.92 and slope -0.65.
`
`small molecule nucleophile attacking an amide, ester,
`or carbonyl carbon where nucleophilicity increases
`with basicity.8° The negative slope observed for CAII
`variants indicates that electrostatic stabilization of the
`transition state by the positively-charged zinc ion,
`rather than the nucleophilicity of the zinc hydroxide,
`is a dominant factor in catalysis. One possible inter-
`pretation of these data is that the protein ligand-zinc
`separation changes to facilitate electrostatic stabiliza-
`tion of the transition state by maintaining the bond
`valence sum of zinc.8~,82 Therefore, any diminution of
`the positive charge in the zinc binding site has dire
`consequences for catalysis; however, variants with a
`substituted neutral ligand, such as His-119 ~ Gin,
`retain reasonable catalytic activity (C.-c. Huang and
`C. A. Fierke, unpublished results) and maintain active
`site structure (C. A. Lesburg and D. W. Christianson,
`unpublished results). The neutral histidine zh~c ligands
`in CAII are essential for attaining high catalytic
`efficiency by enhancing the net positive charge of the
`metal ion site.75,79 Intriguingly, the CAII-like domains
`of receptor protein tyrosine phosphatases contain
`histidine ~ glutamine substitutions, but neither zinc
`binding properties nor catalytic activity has been
`documented for these systems.~-~5
`Amino acid substitutions elsewhere in the direct
`metal coordination polyhedron of CAII reveal that not
`all metal ligands reside on pliable segments of active
`site/9-strands: the His-119 ~ Cys substitution yields
`an anomalously long zinc-sulfur separation of 2.8 ~,
`and the His-96 ~ Cys substitution leads to an unex-
`pected conformational change for the Cys-96 side
`chain that places it 6.1 ~k away from zinc (a water
`molecule moves into the intended coordination position
`of Cys-96, thereby maintaining the tetrahedral geom-
`etry).76 Each of these amino acid substitutions also
`yields a 104-fold loss of protein-metal affinity.75
`Despite the unexpected consequences of the His-96 ~
`Cys and His-119 ~ Cys substitutions, it is clear that
`(1) the less buried an engineered metal ligand is in
`the protein scaffolding, the more capable it will be of
`achieving metal coordination through a plastic struc-
`tural change and (2) no matter what amino acid
`substitution is made in the protein-zinc coordination
`polyhedron, the coordination geometry of zinc remains
`
`(80) Fersht, A. R. E±2zyme Structure and Mechanism, 2rid ed.; W. H.
`Freeman & Co.: New York, 1985; pp 83 85.
`(81) Thorp, H. H. I±*org. Chem. 1992, 31, 1585-1588.
`(82) Xiang, S. B.; Short, S. A.; Wolfenden, R.; Carter, C. W. BJochem
`Jst±y 1996, 35, 1335-1341.
`
`Figure 5. Superposition of wild type (yellow), His 94 ~ Asp
`(cyan), and His 94 ~ Cys (red) CAIIs. Note the plasticity of the
`f! strand containing residue 94, as well as the movement ofZn2+
`which accommodates amino acid substitutions. Zn2+ moves to
`an identical location in the His 94 ~ Asp and His 94 ~ Cys
`variants. Reprinted with permission from ref 76. Copyright 1994
`American Chemical Society.
`
`deformation results in only an additional 2-fold loss
`of protein-metal affinity relative to that of His-94 ~
`Asp CAII. Thus, the f!-strand containing metal ligands
`His-94 and His-96 is sufficiently pliable to accom-
`modate various metal ligand residues at position 94.
`However, it is the movement of zinc from its naturally-
`evolved wild-type position--and not the compensatory
`plasticity of the f!-strand containing the metal
`ligands--that most seriously compromises protein-
`metal affinity. It is intriguing that f!-sheet structure
`and stability can be regulated by the construction of
`transition metal binding sites, and this feature can
`be exploited in protein engineering experiments with
`many other systems.
`The introduction of a negative charge into the zinc
`coordination polyhedron yields a significant increase
`in the pKa of zinc-bound water as assayed by the pH
`dependence of esterase activity (in addition to cata

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