.--r "W _m< rd‘h Slip .-‘.- , .31"! . "Pf-y w. m. . H ~t ,.-‘ . .?:;‘v x ‘5 ‘ -1’ £4“ um'gi'r. ‘r . ‘3. -._€§; ‘2 I . . . :35 N J: - f. ‘1‘ ‘ 3"" t . #{éifiUL $5.167 _ l . b 15 .. 1 -., Glc-6-P + 111313.1492+ This reaction is the first reaction after glucose being transported. into cytoplasmic space, and is, in. essence, phosphorylation of glucose at the expense of high energy ATP (AG = -7.3 kcal/mole hydrolyzed at pH 7). The reaction forms a lower energy phosphoglucose, Glc-6-P (AG = -3.3 kcal/mole hydrolyzed at pH 7); therefore, the AG for this reaction is -4.0 kcal/mole at pH 7 (1). This reaction not only activates glucose for further metabolism but because of the large negative free energy change, the reaction is thermodynamically favored to the right and essentially irreversible (2). The product of this reaction, Glc-6-P, is a common substrate for phosphohexoisomerase, phosphoglucomutase, and Glc-6-P dehydrogenase, which introduce Glc-6-P into glycolysis, glycogen synthesis, and the pentose phosphate pathway, respectively. Thus, hexokinase plays an important role in admitting glucose into metabolism. As the name implies, hexokinases are able to phosphorylate a variety of hexoses (3), e.g. mannose, 2- deoxyglucose, glucosamine and, to a lesser extent, fructose and galactose. .All of these hexoses are phosphorylated at the 6-hydroxyl position, forming hexose 6-phosphate. 3 Evolution of mammalian hexokinases There are four distinct hexokinase isozymes found in mammalian tissues (3-5), designated as type I, II, III, and IV hexokinase. Type I, II, and III isozymes are monomeric with molecular weight of approximately 100 kDa, whereas type IV isozyme, better known as glucokinase, is only 50 kDa. Because the molecular weight of mammalian hexokinase I, II, and III is twice that of yeast hexokinase (50 kDa), it has been suggested first by Easterby (6) and Colowick (7) that mammalian loo-kDa hexokinases arose by duplication and fusion of a gene encoding an ancestral 50 kDa hexokinase similar to present-day yeast hexokinase. If this hypothesis is true, the original ancestral 100 kDa hexokinase should have two identical halves with one active site in each half (6,7). Easterby and Colowick further postulated that this original 100 kDa hexokinase in turn gave rise to modern hexokinase I, II, and III. By comparing amino acid sequences deduced from cDNAs of various sources, hexokinase I, II, and III are indeed very similar (5,8-10). In addition, they exhibit internal 50 kDa repetitions which are similar to the sequence encoding.50 kDa yeast hexokinase (5,8-10) . Furthermore, several residues involved in glucose binding or catalysis are well conserved in each repetition as well as in yeast hexokinase (5,8-10), supporting the theory that modern mammalian 100 kDa hexokinases are derived from a common ancestral hexokinase 4 which itself was a product of gene duplication and fusion (5,11,12). One important piece of evidence supporting this gene duplication and fusion theory is by Printz et al. (13) and Kogure et a1. (14) in which they found that the splicing sites and the exon sizes of hexokinase II gene showed a direct repetition between the N- and C- terminal halves. Interestingly, this intron/exon pattern is also observed in glucokinase:gene (13,14), and therefore they suggested that an ancestral gene encoding a 50 kDa hexokinase similar to glucokinase underwent gene duplication and fusion to form the hexokinase II gene. However, instead of glucokinase being the precursor of hexokinase II, Griffin et a1. (12) suggested an alternative view'in which.the glucokinase gene arose by resplitting of the hexokinase II gene and thus it retained the intron/exon pattern of the hexokinase II gene. In fact, the latter view is perhaps a better model, because it is better supported by the evolution scheme. In lampreys (Agnatha; jawless fish), only one hexokinase isozyme is present, which is about 90 kDa in size (15), while type I, II, III isozymes but not glucokinase are present in further evolved Osteichthys (bony fish). Glucokinase is not present until the emergence of Amphibia and higher animals (except for Aves; no glucokinase is found in Aves) (15,16). These findings would suggest that the gene duplication and fusion event of hexokinase occurred 5 by the time when primitive fish emerged, while glucokinase did not arise until the emergence of Amphibia. Furthermore, Griffin et a1. (12) found that glucokinase has a closer sequence similarity to the mammalian 100 kDa hexokinases than to yeast hexokinase. Therefore, the glucokinase gene seems more likely to be a product of the splitting of the hexokinase II gene, not the precursor. Properties of mammalian hexokinases Mammalian hexokinases I, II, and III are sometimes referred as low'Km hexokinases (4,5), because they have low Km, i.e. high affinity, for glucose (Km: 10-150 uM). Mammalian low lg, hexokinases are sensitive to inhibition by the product, G1c-6-P, with Ki being about 10-100 uM. Type IV isozyme, on the other hand, has lower affinity for glucose with K, being about 5-10 mM, and unlike low Km hexokinases, it is not sensitive to inhibition by physiological concentrations of Glc-6-P (3,5). While hexokinase I is the dominant isozyme found in most tissues, hexokinase II is mostly found in those insulin sensitive organs and tissues, e.g. muscles, liver, and adipose tissues, due to, partially at least, the up-regulation of hexokinase II by insulin (13,17). Because of this tissue distribution and hormonal regulation along with other kinetic properties, it has been speculated that the hexokinase I is responsible for introducing glucose into glycolysis, while 6 hexokinase II is responsible for converting glucose into glycogen in insulin sensitive organs and tissues (4,18) (also see Kinetics of hexokinases). Intracellular localization of hexokinases is not homogeneous (17), which is believed to bear functional significance (5). A large portion of hexokinase I and II was found to co-sediment with mitochondria (17). Later, hexokinase I and II were shown to be bound to the mitochondrial membrane (4,5,17) by interaction of the N- terminal hydrophobic sequences and, presumably, mitochondrial membrane lipids (19) and by interaction between the negative charge on hexokinase surface and presumably mitochondrial membrane phospholipids, bridged by divalent cations (20). In the case of brain hexokinase I, Glc-6-P is able to cause dissociation of hexokinase I from mitochondria to various extent depending on species; the degree of dissociation ranges from 90% in rats to 20% in human, with guinea pig and bovine in between (21). Whether hexokinase II has the same properties is not known. It is noteworthy that there persists a portion of bound hexokinase resistant to Glc- 6-P solubilization, which can be, however, solubilized by 0.5 M KSCN. Comparison of the Glc-6-P soluble and KSCN- solubilized.hexokinases showed.no»distinguishable difference, which led to the postulation by Kabir and Wilson (21) that there are two subtypes of hexokinase—binding-sites on the mitochondrial surface. That is, solubilities of the 7 'hexokinase bound on mitochondria are determined by the nature of hexokinase-binding-sites rather than by different subtypes of hexokinase or mitochondria. The functional significance of intracellular localization of hexokinase I on mitochondria is suggested by BeltrandelRio and. Wilson (22) as the feedback link between oxidative phosphorylation and glycolysis. Because of the close proximity, mitochondria-bound hexokinase I utilizes ATP primarily generated from mitochondria to phosphorylate glucose (22). In this scenario, when the energy charge state in the cell becomes low and mitochondrial oxidative phosphorylation is stimulated, the increased flux of ATP from mitochondria would elevate the flux of glucose phosphorylation (by hexokinase I) which in turn supports the need for carbon sources in glycolysis and Krebs cycle. Hexokinase III is found in many tissues but in much lower amount than other hexokinase isozymes (23). Intracellular localization of hexokinase III was thought to be in the cytoplasmic space, but.recently hexokinase III was found.to be localized at the nuclear periphery. The association of hexokinase III to the nuclear periphery is apparently rather weak, and.disrupted.when tissue is homogenized (23). The role of hexokinase III and its intracellular localization in cell physiology is not clear. Glucokinase is mainly found in liver (24, 25) and in pancreas (25). Because its Km for glucose is in the range of 8 circulating blood glucose levels, fluctuation of blood glucose can greatly influence the rate of glucose phosphorylation catalyzed by glucokinase. During a postmeal high blood glucose state, the high Km for glucose makes glucokinase ideal to convert excess blood glucose into glycogen in liver or to serve as a sensor governing the release of insulin in the pancreatic B-islets. A Glucokinase was thought to be a cytosolic enzyme; however, Miwa et a1. (26) have demonstrated that glucokinase is present in the nucleus as well as in the cytosol. The physiological significance of this intracellular localization of glucokinase is not known. Kinetics of hexokinases The hexokinase reaction proceeds by a ternary complex mechanism, i.e. hexokinase-glucose-ATP.Mg2+ must be formed before the catalysis can occur (4). Whether the substrate binding follows a sequential order or random order has been the subject of a long debate (4,5). Fromm and his coworkers reported that the binding order is a rapid equilibrium random mechanism (18,27,28); that is, either glucose or ATP.Mg“'can bind to hexokinase first, followed by the binding of the other substrate. However, most other investigators reported that substrate binding follows a sequential order in which glucose binds first (4,5,7). To prove the latter view and to avoid the inherent. ambiguity] of 'traditional product inhibition 9 kinetics, Gregoriou et al. (29) demonstrated, with the flux ratio method, that the substrate binding is a sequential ordered mechanism with glucose binding to the enzyme first. Later Ganson and Fromm (30) rebutted with results using the equilibrium isotope exchange method, and reported that the reaction operates by a random mechanism but with a preferred pathway, in which two thirds of the reaction proceeds with glucose binding first and.ATP.Mg“’later. Since the mid 1980, the argument about the kinetic mechanism of hexokinases seems to have subsided, and to this date, there is no clear and decisive report on whether hexokinases follow a sequential order or a random order in the substrate binding. Glc-6-P is a competitive inhibitor vs. ATP.Mg“’(3-5,7), indicating Glc-6-P and ATP.Mg“ are mutually exclusive when binding to hexokinase. Whether Glc-6-P binds at a distinct allosteric site or a site overlapping with the ATP.Mg2+ is another ongoing debate (see Allosteric inhibition of hexokinase and Functional organization of hexokinases in this chapter). Most investigators ihave reported. that. Glc-6-P is a noncompetitive inhibitor vs. glucose (3-5,31,32). However, their double reciprocal (Lineweaver-Burk) plots showed almost parallel lines (31,32), which. is characteristic of uncompetitive inhibition. In fact, Fromm and Zewe (33) reported that the inhibition is uncompetitive. This near parallel pattern, or near uncompetitive inhibition, is 10 probably due to the synergistic effect of glucose and Glc-6-P binding (34,35), in' which the converging point of the Lineweaver-Burk plot becomes distant from the origin (36). Thus, this unusual "mixed type" inhibition pattern closely resembles uncompetitive inhibition. One other.unique characteristic of hexokinase I is that inorganic phosphate can antagonize Glc-6-P inhibition at low mM concentrations (11,27,28,37,38). This unique characteristic is believed to play a significant role under certain physiological conditions in the brain (11). In brain, hexokinase I is considerably inhibited by its product, Glc-6- P, under normal conditions. When an energy demanding circumstance arises, the elevated phosphate concentration, resulted from hydrolysis of ATP, serves to relieve hexokinase I from Glc-6—P inhibition (11) “and thus hexokinase I introduces more glucose into glycolysis and-the Krebs cycle to meet the energy demand. At higher concentrations (>5 mM), however, phosphate starts to exhibit its inhibitory effect on hexokinase I (4,5,38). In the case of hexokinase II, antagonism of Glc-6-P inhibition by inorganic phosphate is not seen and.phosphate is solely an inhibitor with Ki being about 3 mM (18,38). This inhibitory effect of inorganic phosphate would suggest that during muscle exercise, elevated phosphate concentrations, again resulting from ATP hydrolysis, could inhibit hexokinase II and at the same time activate glycogen phosphorylase and 11 phosphofructokinase (18). As a result, myoglycogen is funneled into glycolysis to generate more ATP for muscle contractions. Such a mechanism indicates that the source of carbon fuel for muscle contraction is myoglycogen not free glucose; therefore, hexokinase II is responsible for converting glucose to myoglycogen rather than introducing glucose into glycolysis. Compared to hexokinase I and II, hexokinase III has lower Kfi for glucose (10-30 uM) and higher KifOr Glc-6-P (approx. loo‘uM) (3-5), and uniquely, hexokinase III exhibits substrate inhibition by glucose in the mid to high mM range (3,39). The physiological significance of these kinetic properties is not clear. Allosteric inhibition of hexokinases By using a series of analogues, Sols and Crane (40,41) discovered that brain hexokinase I has different specificities for the hexose moiety of substrates and inhibitors, and therefore suggested that the inhibitor occupies a distinct allosteric site' different from the active site where substrates bind. For example, both mannose and 2-deoxyglucose are good substrates for hexokinase I, but their phosphorylated counterparts, mannose-6-P and 2-deoxyglucose-6-P are poor inhibitors. This hypothesis was further tested by Sols (42) and Ureta et al. (43) with the reverse reaction catalyzed by hexokinase I, in which G1c-6-P serves both as a substrate and 12 as an inhibitor. The reverse reaction showed substrate inhibition when Glc-6-P was used as the substrate, whereas 2- deoxyglucose-6-P showed no substrate inhibition due to its poor ability to inhibit hexokinase I. Furthermore, yeast hexokinase, which is insensitive to Glc-6-P inhibition, showed no substrate inhibition when Glc-6-P was used as the substrate in the reverse reaction. These experiments (43) suggested that G1c-6—P exerts it inhibitory effect on hexokinase I by binding to a distinct allosteric site rather than the active site. The separate allosteric site in hexokinase I is also evident from the fact that Glc-6-P is not a competitive inhibitor vs. glucose (3). If Glc-6-P inhibits hexokinase I by binding to the active site, it should have been a competitive inhibitor vs. glucose, since both molecules cannot occupy the same active site at the same time. The fact that Glc-6-P is not a competitive inhibitor vs. glucose and that both G1c-6—P and glucose can bind to hexokinase I simultaneously (34,35) confirms the discrete allosteric and active sites on hexokinase I. In fact, Glc-6-P and glucose showed synergism when they bind to hexokinase I (34,35); that is, binding of either ligand will promote the binding of the other ligand. Thus, Glc-6-P must bind at a discrete allosteric site, which does not overlap with the active site where glucose binds. 13 Structure and ligand binding Although there is no mammalian 100 kDa hexokinase crystal structure available, a significantly related hexokinase from yeast (50 kDa) has been A '8‘ 50 _ <1 .3 _ 82 . ; . NIICI . 25 - H ‘ NICII 0 o I l l I O 5 10 15 20 Time (min) Figure 7. Heat stability of Type I, Type II, and chimeric hexokinases. Hexokinase activity in extracts of COS-1 cells transfected with cDNAs for Type I (Cl) , Type II (0 ) , NICII (A) , and NIICI (V) hexokinases was determined as a function of time at 40° C. Values shown are the average from two experiments (each with extracts from a different transfection) which gave similar results; average deviation between experiments was i 4% from the values shown. 54 Table II. Kinetic Properties of Type I and Type II Hexokinase and the Chimeric Hexokinases. Km for Ki for Ieeeyme Glucose(pM)‘ ATP (mM)‘ Angie-g-P (gu)° Type I 65 i 4 0.52 i 0.04 17 i 1 Type II 142 i 1 0.77 i 0.07 22 i 4 NICII 140 I 17 1.01 i 0.15 72 i 20 NIICI 82 i 13 0.64 i 0.06 6 i 1 ‘Mean 1 SD for 3 determinations, each with an extract from a different transfection. bMean 1 SD for 4 determinations, each with an extract from a different transfection. 55 and Type II isozymes, and indicate somewhat higher apparent affinity (lower Kh) of the Type I isozyme for both substrates. Kinetically, the chimeras resembled the parental isozyme from which the C-terminal half was derived, i.e., NIICI had a lower Km for both substrates than did NICII. As previously found for Glc-6-P itself (14,40) as well as its analog, AnGlc-6-P (44,45), the Type:I and Type II isozymes were quite similar in their sensitivity to inhibition by AnGlc-6-P (Table II). In contrast to the parental isozymes, the chimeric enzymes differed markedly in their sensitivity to this inhibitor, with NICII having aim approximately 10-fold greater than that seen with NIICI. For all forms, inhibition by AnGlc-6-P was competitive vs. ATP, as found in earlier studies with the parental isozymes (44,45). In contrast to the similarity in sensitivity of the Type I and Type II isozymes to inhibition by AnGlc-6-P (44, and Table II), Rose et al. (44) had found that the Type II isozyme was considerably more sensitive to inhibition by Glc-1,6- bisphosphate. This was confirmed in the present study (Fig. 8). As in the case of inhibition by AnGlc-6-P, the chimeric enzymes differed much more markedly than did the parental isozymes in their sensitivity to inhibition by Glc-1,6- bisphosphate, with the NIICI chimera again being much more susceptible than was the NICII chimeric hexokinase (Fig. 8). In agreement with Rose et a1. (44), inhibition was found to be competitive vs. ATP (data not shown) ; Ki values determined (one 56 100 '3: ‘5 y T T .3 .-. NICII 75 - i a I -.?.-. I 0 so - ' < ~ 11 . °\° .L . 25 - NIICI ° o I l J l 0 25 50 75 100 " Glc-1,6-bisP (uM) Figure 8. Inhibition of Type I, Type II, and chimeric hexokinases by Glc-1,6-bisphosphate. Activity was determined as 'a function of increasing Glc-1,6-bisphosphate concentration, with initial ATP concentration in each assay equal to the Km value (Table II) for the respective hexokinase. ([3), Type I; (0), Type II; (A), NICII; (V), NIICI. Values shown are mean f SD from three experiments, each with extracts from a different transfection; where no error bars are seen, they are obscured by the data point symbol. 57 experiment) for the Type I, Type II, NICII, and NIICI hexokinases were 40, 11, 143, and 4 pM, respectively. Although.P,is an inhibitor, competitive vs. ATP, for both Type:I and Type II hexokinase, the Type I isozyme was reported to be much less sensitive to this inhibitor, with a K,of 35 mM (27), well above the K, of 2.7 mM found for the Type II isozyme (30). We have confirmed this marked difference in sensitivity to P, with the Type I and Type II isozymes expressed in COS-1 cells (Fig. 9). The chimeric enzymes were indistinguishable from the parental isozyme from.which the N- terminal half was derived, i.e., NICII responded as the Type I isozyme, and NIICI as the Type II isozyme (Fig. 9). This same correlation with the origin of the N-terminal half was seen in the response of the chimeric enzymes to P,as an antagonist of inhibition by the Glc-6-P analog, AnGlc-6-P (Fig. 10). The Type I isozyme, expressed in COS-1 cells, responded as previously reported for the Type I isozyme from brain (26,27,29), i.e., low concentrations of P,reversed the inhibition while at higher concentrations, inhibition by P, itself became evident. The NICII chimera was virtually identical to the Type I isozyme in this response. In contrast, P, did not reverse inhibition of the Type II isozyme, either from rat skeletal muscle (30) or expressed in COS-1 cells (Fig. 10), and the NIICI chimeric hexokinase was indistinguishable from the Type II isozyme in this lack of response to P,as an antagonist of 58 100 »- I E‘- ': 'L ““\ u ‘- 75 x fifififififififi I *- u\ ..... I z: NICII eeq \ .2. n «I _L ..... 2 50 - ‘-~-_‘—__5 ~ 7 82 NIICI .L Phosphate (mM) Figure 9. Inhibition of Type I, Type II, and chimeric hexokinases by P,. Activity was determined as a function of increasing P, concentration, with initial ATP concentration in each assay equal to the Km value (Table II) for the respective hexokinase. (Cl), Type I; (0), Type II; (A), NICII; (V), NIICI; where no error bars are seen, they are obscured by the data point symbOl. 59 100 I 'ty 1V1 % Act Phosphate (mM) Figure 10. Effectiveness of P,as an antagonist of inhibition by' AnGlc-6-P. .Activity' was. determined as a function of increasing P,concentration, with initial ATP concentration in each assay equal to the KIn value (Table II) for the respective hexokinase, and AnGlc-6-P concentration sufficient to reduce the activity to approximately 50% of that seen in the absence of AnGlC-G-P- (ID). Type I; (0), Type II; (A). NICII; (V). NIICI. Values shown are mean f SD from three experiments, each with extracts from a different transfection; where no error bars are seen, they are obscured by the data point symbol. 60 inhibition by AnGlc-6-P. DISCUSSION There are at least two possible interpretations for the observation that ichimeric hexokinases can be expressed as highly active species with retention of kinetic characteristics.and.thermal stabilities that are.comparable to the parental isozymes. One would be that the N- and C-terminal halves of these molecules exist as quasi-independent domains. If this were the case, the properties of a particular domain would essentially be independent of the other, e.g., the catalytic activity of the C-terminal domain of Type I hexokinase (15) would be fully expressed in chimeric forms, and its catalytic and regulatory properties would not depend on the identity of the N-terminal domain with which it was fused. The present work demonstrates that the latter prediction is not fulfilled. The inhibition by P, and the effectiveness of P, as an antagonist of inhibition by AnGlc-6-P are clearly dependent on the origin of the N-terminal domain. Previous studies provide additional evidence against the view that the N- and C-terminal halves are essentially independent. Thus, if these regions were tethered together by a linking polypeptide sequence but with no significant noncovalent interactions between the halves, one might anticipate the linking polypeptide segment to be readily 61 susceptible to proteolysis, as is frequently the case with segments linking functionally and structurally independent domains (46). In fact, cleavage*within the segment linking the N- and. C-terminal halves of Type. I Ihexokinase requires perturbation of the structure with low concentrations of a denaturing agent, guanidine hydrochloride, implying that the linking segment is "buried" in the native structure as a result of close spatial interactions between the N- and C- terminal regions (15). Moreover, if the N- and C-terminal regions were quasi-independent, then one would anticipate that conformational changes induced by the binding of ligands to a particular domain would largely be restricted to that domain. However, there is abundant evidence that binding of various ligands, including hexoses, hexose 6-phosphates, ATP, and P“ evokes conformational effects throughout the molecule, affecting susceptibility to proteolytic attack or chemical modification of sulfhydryl or arginyl residues, thermal stability, and immunoreactivity with monoclonal antibodies recognizing conformational epitopes (25,47-51). Collectively, these observations provide a solid basis for rejecting the view that the N- and C-terminal halves represent functionally and structurally discrete,- quasi- independent entities. On the contrary, they lead to the conclusion that the N- and C-terminal halves of these isozymes are in intimate contact, providing a structural basis for functional interactions between these regions. Thus, an 62 alternative interpretation for the preservation of catalytic function in the chimeric enzymes would be that interactions between the N- and C-terminal domains have been maintained, without distortion of structural features critical for catalysis. In other words, it seems likely that interactions between the N- and C-terminal halves are similar in the Type I and Type II isozymes. In view of the extensive similarity between the amino acid sequences of these isozymes (and also that of the Type III isozyme) (1), which obviously must include segments involved in interactions between the N- and C-terminal domains, it is perhaps not surprising that this would be the case. As noted in the introduction to this paper, the location of the functional allosteric site t0‘which.the inhibitory Glc- 6-P is bound remains in dispute (20,22,23-25). The present work obviously does not resolve this controversy since there is no clear correlation between the relative susceptibility to inhibition by the Glc-6-P analogs, AnGlc-6-P and Glc-1,6- bisphosphate, and the parental origin of either the N- or C- terminal domain of the chimeric hexokinases (but see below). The present study demonstrates that the affinity for these inhibitors is markedly influenced by the identity of the N- and C-terminal domains comprising the 100 kDa enzymes. It thus seems likely that the apparent affinity for inhibitory hexose phosphates is determined by complex interactions involving both the N- and C-terminal domains, and depends on specific 63 aspects of the interdomain interactions that, despite an overall similarity, vary somewhat in both the parental isozymes as well as the chimeras. Such ambiguity is not seen in the response to P, as an . inhibitor and as an antagonist of inhibition by AnGlc-6-P. It is clear that this is correlated with the origin of the N- terminal domain. The results are fully consistent with the proposal (24,25) that the N-terminal half of Type I hexokinase is the location of an anion binding site for which P,and the 6-phosphate of Glc-6-P (or analogs) compete, accounting for both the mutually exclusive binding of these ligands (26,27) and the resulting antagonism of inhibition by the hexose 6- phosphates (with binding of P,at the N-terminal site.not being inhibitory), while the C-terminal half includes a site to which P, binds with lower affinity and with resulting inhibition. Finally, wenote that the present study has revealed that the chimeric hexokinases Idiffer significantly from the parental isozymes in their response to product inhibition, i.e., the ‘marked sensitivity’ of the ‘NIICI chimera, and relative insensitivity of the NICII chimera, to inhibition by Glc-6-P analogs (and presumably Glc-6-P itself). This offers an intriguing opportunity for examining the allosteric regulation of hexokinase activity in situ by transfection of cells with DNA. encoding" hexokinase forms differing substantially in their sensitivity to product inhibition. 10. 11. 12. 13. 14. 15. 16. REFERENCES Wilson,.J.E. (1994) Rev. Physiol. Biochem. Pharmacol., in press. Nishi, S. Seino, S., and Bell, G.I. (1988) Biochem. Biophys. Res. Commun. 157, 937-943. Schwab, D.A., and Wilson, J.E. (1989) Proc. Natl. Acad. Sci. USA 86, 2563-2567. Arora, K.K., Fanciulli, M., and Pedersen, P.L. (1990) J. Biol. Chem. 265, 6481-6488. Griffin, L.D., Gelb, B.D., Wheeler, D.A., Davison, D., Adams, V., and McCabe, E.R.B. (1991) Genomics 11, 1014- 1024. Thelen, A.P., and Wilson, J.E. (1991) Arch. Biochem. Biophys. 286, 645-651. Schwab, D.A., and Wilson, J.E. (1991) Arch. Biochem. Biophys. 285, 365-370. Deeb, S.S., Malkki, M., Laakso, M. (1993) Biochem. Biophys. Res. Commun. 197, 68-74. Bork, P., Sander, C., and Valencia, A. (1993) Prot. Sci. 2, 31-40. Kopetzki, E., Entian, K.-D., and Mecke, D. (1985) Gene 39, 95-102. Stachelek, C., Stachelek, J., Swan, J., Botstein, D., and Konigsberg, W. (1986) Nucleic Acids Res. 14, 945-963. Colowick, S.P. (1973) in The Enzymes (Boyer, P.D., Ed.), 3rd ed., Vol. 9, pp. 1-48, Academic Press, New York. Easterby, J.S., and O’Brien, M.J. (1973) Eur. J. Biochem. 38, 201-211. Ureta, T. (1982) Comp. Biochem. Physiol. 718, 549-555. White, T.K., and Wilson, J.E. (1989) Arch. Biochem. Biophys. 274, 375-393. Chothia, C., and Lesk, A.M. (1986) EMBO J. 5, 823-826. 64 17. 18. 19. 20. 21. 22. 230 240 25. 26. 270 28. 29. 30. 310 32. 65 Harrison, R. (1985) Crystallographic Refinement of Two Isozymes of Yeast Hexokinase and Relationship of Structure to Function. Ph.D. thesis, Yale University, New Haven, CT. Baijal, M., and. Wilson, J.E. (1992) Arch. Biochem. Biophys. 298, 271-278. Arora, K.K., Filburn, C.R., and Pedersen, P.L. (1991) J. Biol. Chem. 266, 5359-5362. Arora, K.K., Filburn, C.R., and Pedersen, P.L. (1993) J. Biol. Chem. 268, 18259-18266. Chen, L., Zeng, C., and Fromm, H.J. (1993) FASEB J. 7, A1176. Magnani, M., Bianchi, M., Casabianca, A., Stocchi, V., Daniele, A., Altruda, F., Ferrone, M., Silengo, L. (1992) Biochem. J. 285, 193-199. White, T.K., and Wilson, J.E. (1987) Arch. Biochem. Biophys. 259, 402-411. White, T.K., and Wilson, J.E. (1990) Arch. Biochem. Biophys. 277, 26-34. Hutny, J., and Wilson, J.E. (1990) Arch. Biochem. Biophys. 283, 173-183. Ellison, ‘W.R., Lueck, J.D., and Fromm, H.J. (1974) Biochem. Biophys. Res. Commun. 57, 1214-1220. Ellison, W.R., Lueck, J.D., and Fromm, H.J. (1975) J. Biol. Chem. 250, 1864-1871. Chou, A.C., and. Wilson, J.E. (1974) Arch. Biochem. Biophys. 165, 628-633. Purich, D.L., and Fromm, H.J. (1971) J. Biol. Chem. 246, 3456-3463. Lueck, J.D., and Fromm, H.J. (1974) J. Biol. Chem. 249, 1341-1347. Sambrook, J., Fritsch, E.F., and Maniatis, T. (1989) Molecular Cloning: A laboratory Manual, 2nd ed., Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Bird, P., Gething, M.-J., and Sambrook, J. (1987) J; Cell Biol. 105, 2905-2914. 33. 34. 350 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 66 DeWitt, D.L., El-Harith, E.A., Kraemer, S.A., Andrews, M.J., Yao, E.F., Armstrong, R.L., and Smith, W.L. (1990) J. Biol. Chem. 265, 5192-5198. Housley, P.R., and Forsthoefel, A.M. (1989) Biochem. Biophys. Res. Commun. 164, 480-487. Wilson, J.E. (1989) Prep. Biochem. 19, 13-21. Perrella, F.W. (1988) Anal. Biochem. 174, 437-447. Hill, H.D., and Straka, J.G. (1988) Anal. Biochem. 170, 203-208. Smith, A.D., and Wilson, J.E. (1991) Arch. Biochem. Biophys. 287, 359-366. Dixon, M., and Webb, E.C. (1964) Enzymes, Academic Press, New York. Grossbard, L., and Schimke, R.T. (1966) J. Biol. Chem. 241, 3546-3560. Katzen, H.M., and Schimke, R.'T. (1965) Biochem. Biophys. Res. Commun. 54, 1218-1225. Gonzalez, C., Ureta, T., Babul, J., Rabajille, E., and Niemeyer, H. (1967) Biochemistry 6, 460-468. Wilson, J.E., and Smith, A.D. (1985) J. Biol. Chem. 260, 12838-12843. Rose, I.A., Warms, J.V.B., and Kosow, D.P. (1974) Arch. Biochem. Biophys. 164, 729-735. Wilson, J.E., and Chung, V. (1989) Arch. Biochem. Biophys. 269, 517-525. Wilson, J.E. (1991) in Methods of Biochemical Analysis (Suelter, C.H., Ed.), Vol. 35, pp. 207-250, Wiley- Interscience, New York. Wilson, J.E. (1978) Arch. Biochem. Biophys. 185, 88-99. Wilson, J.E. (1979) Arch. Biochem. Biophys. 196, 79-87. Wilson, J.E. (1982) Arch. Biochem. Biophys. 218, 254-262. Smith, A.D., and Wilson, J.E. (1991) Arch. Biochem. Biophys. 291, 59—68. 67 51. Smith, A.D., and Wilson, J.E. (1992) Arch. Biochem. Biophys. 292, 165-178. FOOTNOTES ‘Abbreviations used: Glc-6-P, glucose 6-phosphate; AnGlc-6-P, 1,5-anhydroglucitol 6-phosphate; SDS, sodium.dodecyl sulfate; ANOVA, analysis of variance. 68 Chapter III Functional Organization of Mammalian Hexokinases: Both N- and C-Terminal Halves of the Rat Type II Isozyme Possess Catalytic Sites Archives of Biochemistry and Biophysics Vol.329, No. 1, May 1, pp. 17—23, 1996 69 ABSTRACT Previous ‘work. has shown that. catalytic function is associated exclusively with the C-terminal half of the Type I isozyme of mammalian hexokinase. In contrast, we now demonstrate that both halves of the Type II isozyme possess comparable catalytic activities. Mutation of a catalytically important Ser residue to Ala at analogous positions in either the N- or C-terminal halves (8155A or S603A, respectively) of the rat Type II isozyme resulted in approximately 60% reduction in specific activity of the enzyme, with more than 90% reduction in the doubly mutated enzyme (8155A/S603A). Catalytic activity was retained in a chimeric hexokinase comprised of the N-terminal half of Type II hexokinase and catalytically inactive (by site-directed.mutation) C-terminal half of the Type I isozyme. The N- and C-terminal catalytic sites of Type II hexokinase are similar in Vmu and Km (£130 pH) for Glc; however, the N-terminal site has a lower (0.45 mM vs. 1.1 mM) Km for ATP, is slightly 'more sensitive to inhibition by the product analog, 1, 5-anhydroglucitol-6-P, and is much more sensitive to inhibition by P,. It is suggested that the Type II isozyme most closely resembles the 100 kDa hexokinase which resulted from duplication and fusion of a gene encoding an ancestral 50 kDa hexokinase and.which.was the precursor for the contemporary Type I, Type II, and Type III mammalian isozymes. Subsequent evolutionary changes could then have led to functional differentiation of the N- and C- 70 71 terminal halves, as seen with the Type I (and possibly the Type III) isozyme. 72 Mammalian tissues contain three isozymes of hexokinase (ATP:D-hexose 6-phosphotransferase, EC 2.7.1.1), generally designated as Types I, II, and III. These exist as monomeric species with molecular weights of approximately 100,000 and show similarity in amino acid sequence (reviewed in Ref. 1) consistent with their membership in a closely related family which also includes other kinases (2). Moreover, these isozymes exhibit internal sequence repetition, i.e. there is extensive sequence similarity between the N-terminal and C- terminal halves of the enzymes (1,3,4), and these in turn are similar to sequences of 50 kDa hexokinases such as those found in yeast (5,6) or Schistosoma mansoni (7). Such observations have led to the current view that the 100 kDa mammalian isozymes have evolved by duplication and fusion of a gene encoding an ancestral 50 kDa hexokinase, which also gave rise to the 50 kDa hexokinases of contemporary organisms (1). Site-directed.mutagenesis studies (8-12) have identified several amino acid residues as being catalytically important, and all of these are conserved in both N- and C-terminal halves of all Type I, II, or III isozymes whose sequences have been determined (1). Despite this, it is clear that the two halves of the Type I isozyme are functionally distinct, with catalytic activity associated solely with the C-terminal half (9-11,13) while the N-terminal half is catalytically inactive and thought to serve a regulatory function (13-15). Whether this same functional organization exists in the Type II and 73 Type III isozymes has been uncertain. A recent study (15) of chimeric hexokinases, constructed by interchange of cDNA.segments encoding the:N- and C-terminal halves of the Type I and Type II isozymes followed by expression in C051 cells, confirmed that catalytic function was associated with.the C-terminal half of Type II hexokinase, as with the Type I isozyme. A chimera comprised of the catalytically inactive N-terminal half of Type I hexokinase and the C-terminal half of the Type II isozyme was catalytically active. However, we also noted that higher specific activities were associated with hexokinases possessing the N-terminal half of the Type II isozyme (either the Type II isozyme itself or the chimera produced by combination of the N-terminal half of Type II with the C- terminal half of Type I). As one possible explanation for this observation, we speculated that, unlike the N-terminal half of the Type I isozyme, the N-terminal half of the Type II isozyme might retain intrinsic catalytic activity. The present study has shown this to be the case. In addition to contributing to our ‘understanding' of ‘the functional organization. of the mammalian isozymes of hexokinase, these findings have implications with respect to the evolutionary relationship between these isozymes. MATERIALS AND METHODS Materials. Restriction enzymes, T4 DNA ligase, and Glc-6- ]? dehydrogenase 'were purchased from Boehringer' Mannheim (Indianapolis, IN) and other biochemicals from Sigma Chemical Co. (St. Louis, MO). Plasmid DNAs were purified using plasmid purification kits obtained from Qiagen, Inc. (Chatsworth, CA). DMEM (high glucose) for culture of €081 cells was a product of GIBCO BRL (Gaithersburg, MD), and supplemented with fetal bovine serum (2%) and defined bovine calf serum (8%) from HyClone Laboratories (Logan, UT). The BCA Protein Assay Reagent and BSA standard, as well as the Imject Activation Immunogen Conjugation Kit were purchased from Pierce Chemical Co. (Rockford, IL). Site-direCted mutagenesis and construction of vectors for expression in 0031 ceiie. A previously described (15) full- 1ength cDNA encoding rat Type II hexokinase was used. Sequence encoding the N-terminal half of the enzyme is encoded by an Eco RI-Nco I fragment, while a complementary Nco I-Pst I fragment encodes the C-terminal half. This division 'was convenient for selectively generating site-directed mutations within the N- or C-terminal halves, and then reconstructing full-length coding sequences with any desired mutations. Site-directed mutagenesis was done using the Altered Sites kit and pSelect vector from Promega (Madison, WI), as previously described in detail (10). An Nco I site had been 74 75 engineered (10) into the multiple cloning site of the pSelect vector to facilitate manipulations. The Eco RI-Nco»I or.Nco I- Pst I fragments were cloned into the modified pSelect to give vectors containing sequence encoding the N- or C-terminal halves, respectively. Mutations were made at a Ser residue that is conserved in both‘ N- and C-terminal halves of the mammalian hexokinases (1) and which was previously shown to be of critical importance for catalytic function in the Type I isozyme (8,10). Mutated forms of the rat Type II isozyme with Ser converted to Ala in the N-terminal half, the C-terminal half, or both halves are designated as $155A, S603A, or SISSA/S603A, respectively. The mutant 8155A was made using the oligonucleotide:GGTTTCACCTTCQCGTTCCCCTG, where the underlined base corresponds to conversion of the Ser codon (ECG) to the Ala codon (QCG) ; the analogous mutation in the C-terminal half, S603A, was made using the oligonucleotide GGTTTCACATTCQCCTTCCCTTG, with the underlined base resulting in conversion of the TCC coding for Ser to QCC coding for Ala. Complete coding sequences for the Type II isozyme, with any desired mutations, were reconstructed by ligation of complementary fragments and cloning in pUC18. The construct was then subcloned into pSVT7 for expresssion in COSl cells as previously described (15). All mutations were confirmed by sequencing, first in the pSelect vector in which the mutation had been made, and then again after transfer of the mutated construCt to pSVT7. 76 Expression of Type II hexokinase and mutants in COSl eelle. The transfection procedure was exactly as described previously. Sham transfected cells were treated identically except that no plasmid DNA was added. Cells were harvested 3 days after transfection and extracts prepared as previously described (15) except that the sonicate was centrifuged at 15,000 x g for 10 min prior to analysis. Protein and hexokinase activity were determined immediately after preparation of the extracts, which were then stored at -80° C for further use. Determination of hexokinase activity and protein. Hexokinase activity was determined using a spectrophotometric assay in which Glc-6-P formation is coupled to NADPH production, monitored at 340 nm, in the presence of excess Glc-6-P dehydrogenase (16). The Kms for substrates, Glc and ATP, were determined under the same conditions, except for appropriate variations in the concentration of the relevant substrate. Prior to determination of the Km for Glc, Glc present in the extraction buffer was removed by chromatography on spin columns of Sephadex G-25 (fine), equilibrated with 50 mM Hepes, 0.5 mM EDTA, 10 mM thioglycerol, pH 7.5; absence of Glc in the eluted enzyme was confirmed by enzymatic assay. All kinetic data conformed to Michaelis-Menten behavior and were analyzed using the EZ Fit program of Perrella (17) . Inhibition by the Glc-6-P analog, AnGlc-6-P, and by P,and its analogs, sulfate and arsenate, was determined as in previous studies 77 (10,15). Protein was assayed using the BCA Protein Assay reagent. Samples were pretreated with iodoacetamide (18) to avoid interference from ‘thioglycerol present. in ‘the extraction buffer. Eroduction of a pelyclonal antiserum specific for the Tyg II isozyme of hexokinase. While the isozymes of mammalian hexokinase exhibit extensive similarity throughout virtually all of their amino acid sequences, their C-terminal sequences are unique (1). An octapeptide having the sequence CIREAGQR was synthesized in the Macromolecular Structure, Sequencing & Synthesis Facility, Michigan State University, with purity and identity confirmed by MALDI TOF mass spectrometry. The last seven residues in this peptide correspond to the C-terminal sequence of the rat Type II isozyme of hexokinase, which is quite distinct from that of the Type I or Type III isozymes; the Cys residue at the N-terminus was included to facilitate conjugation to carrier protein for immunization. The latter was.done, with.keyhole limpet hemocyanin.as carrier, using the Imject Conjugation Kit from Pierce Chemical Co. The conjugate was provided to University Laboratory Animal Resources, Michigan State University, for production of antiserum. A rabbit was immunized with 100 pg of the conjugate, emulsified in Hunter's Titer Max, with subsequent booster injections approximately one and two months later. Titer was determined by ELISA, essentially as described previously (19), except 78 that wells of the microtiter plate were coated with BSA that had been conjugated with the above peptide, again using the conjugation kit from Pierce Chemical Co. Several bleedings showing good titer were combined and used for these experiments. SDS gel electrophoresis and impunoblottipg. Procedures for SDS gel electrophoresis were essentially as described earlier (11,20) except that blotting was done using a Trans- Blot SD Transfer Cell (BioRad Laboratories, Richmond, CA). Immunoblots were analyzed quantitatively (11) using the GDS- 2000 gel documentation system and associated software from UVP, Inc. (San Gabriel, CA). geguence comparisons. Sequences of the N- and C-terminal halves of the isozymes were compared using the BESTFIT routine of the GCG Sequence .Analysis Software Package (Genetics Computer Group, Inc., Madison, WI). Default weighting values (gap penalty = 3.00; gap extension penalty = 0.10) were used. Sequences used in comparisons were: residues 1-475, 1-475, and 1-488 for the N-terminal halves of Types I, II, and III, respectively; residues 476-918, 476-917, and 489-924 for the respective C-terminal halves. RESULTS Expreesien of Zyg II hexokinase and mutant forms in cos; cells. Specific activities of hexokinase in extracts from C081 cells transfected with pSVT7 constructs encoding the wild type 79 Type II isozyme or mutant forms are shown in Table I. As in the previous study (15), transfection with the cDNA for the wild type isozyme resulted in hexokinase activities that were well above those in sham tranfected cells, i.e., above endogenous levels of hexokinase in COS1 cells. Transfection with cDNAs encoding the singly mutated forms, SISSA or S603A, gave activities that were also well above those in sham transfected cells, but lower than that with wild type cDNA. Hexokinase activities in extracts from cells transfected with cDNA encoding the double mutant, 8155A/8603A, were markedly reduced.but still higher than those of sham transfected cells. Examination of extracts by SDS gel electrophoresis (Fig. 1) disclosed the expected presence of an additional component in extracts from transfected cells, migrating slightly ahead of the Type I hexokinase (16) used as marker; no other differences were detected between transfected cells and the sham transfected control. It has previously been noted (15) that, despite the close similarity in actual molecular weight of the Type I and Type II isozymes, the Type II isozyme migrates slightly faster on SDS gels. It was not possible to quantitate the relative amounts of these expresSed species by densitometric methods due to inadequate resolution from other closely adjacent components. However, visual estimation of the relative staining intensities indicated that the mutant forms (both single and double mutant) were present in amounts somewhat.greater than the*wild type enzyme. This was supported 80 Table I. Hexokinase Activity in Extracts of C081 Cells Transfected with cDNAs Encoding Wild Type and Mutant Forms of Type II Hexokinase cDNA used for Hexokinase activity‘ transfection (unitslmg protein) Sham control ‘ 0.05 i 0.02 Wild type 0.76 I 0.25 SlSSA . 0.52 i 0.16 S603A 0.51 i 0.21 $155A/8603A 0.09 i 0.03 'IMean 1 SD for 11 different transfections. ANOVA analysis and t-test indicated significant differences (p<0.01) between sham control and all other values, between wild type and all mutants, and between the single mutants and the double mutant; there was no significant difference between the S155A and 8603A mutants. 81 Figure 1. SDS-gel electrophoretic analysis of C081 cell extracts. Lane 1, Type I hexokinase purified from rat brain (16); lane 2, extract from sham-transfected C081 cells; lanes 3, extract from C081 cells transfected with cDNA encoding wild type Type II hexokinase; lanes 4-6, extracts from C081 cells transfected with cDNAs encoding mutants 8155A, 8603A, and the double mutant 8155A/S603A, respectively. The gel was stained with Coomassie Blue. The position of the Type II isozyme, and mutant forms, is indicated by the arrow at the right. 82 by immunoblotting results (Fig. 2) which indicated. more intense staining of bands corresponding to the mutant forms. Since the epitope recognized by the antiserum is the C- terminal sequence common to the wild type and.mutant forms, it is reasonable to assume that the immunoreactivity of the various forms is equivalent, and thus that more intense staining on the blot is indicative of higher levels of the particular protein. Relative specific activities of wild type and mutant forms of Type II hexokinase. Staining intensity of the Type II hexokinase band on immunoblots increased linearly with the amount of extract from cells transfected with the wild type enzyme (Fig. 3). Since it is reasonable to assume that mutant and wild.type forms exhibit equivalent.immunoreactivity, it is therefore possible to estimate the relative amounts of wild type and mutant proteins based on relative staining intensities. On this basis, the mutants 8155A, 8603A, and 8155A/S603A were expressed at levels 1.5- to 2-fold higher than the wild type enzyme. Knowing the relative amounts of these enzymes present in the extracts (from immunoblotting), as well as the specific activities of hexokinase present in the extracts (e.g., Table I), it ‘was possible to calculate the relative specific activities of the mutant and wild type forms. The specific activity of the endogenous C081 enzyme in cell extracts was assumed to be constant and equal to that of the sham 83 Figure 2. Immunoblot analysis of C081 cell extracts. Extracts were from: lane 1, sham-transfected cells; lanes 2-5, cells transfected with cDNAs encoding wild type Type II hexokinase, 8155A, 8603A, or 8155A/8603A, respectively. The blot was probed with antiserum raised against a peptide corresponding to sequence at the C-terminus of rat Type II hexokinase. 84 600 ° I 400 ° 200 Area (pixels) 0 20 40 60 80 100 Cell Extract (ill) Figure 3. Densitometric analysis of an immunoblot. Increasing amounts of an extract from C081 cells transfected with cDNA encoding the wild type Type II isozyme were used, and the intensity of the bands on the immunoblot was quantitated as described in Methods. The resulting standard curve was then used to estimate quantities of the mutant forms, 8155A, 8603A, or 8155A/8603A, present in extracts from cells transfected with the respective cDNAs; the latter extracts were included on the same blot used for generating the standard curve. 85 transfected control; this value was subtracted from. the 'specific activities seen in extracts from cells transfected with wild type or mutant cDNAs. The specific activity of a mutant form, relative to that of the wild type enzyme, could then be calculated as the ratio of the activity present in the "mutant" extract and that present in a "wild type" extract containing an equivalent amount of the expressed isozyme. On this basis, the specific activities of the 8155A, 8603A, and 8155A/S603A mutants were calculated to be 36%, 43%, and 4% of the wild type isozyme, respectively. In another determination, using extracts from a different transfection experiment, the corresponding values were quite similar, being 33%, 40%, and 7%, respectively. Kinetic parameters and refllatory properties of wild type and mutant forms of Tyg II hexokinase. The Kms of the expressed wild type enzyme as well as the 8155A and 8603A mutants are shown in Table II; the various forms were not distinguishable in. their apparent affinity for' Glc, but differed significantly in apparent affinity for ATP. The mutant and wild type isozymes did not differ greatly in their sensitivity to inhibition by AnGlc-6-P (Fig. 4), although 8155A was slightly less susceptible, and S603A slightly more susceptible, than the wild type Type II isozyme. In contrast, the 8155A mutant was markedly less sensitive, and the 8603A mutant considerably more sensitive than wild type Type II hexokinase to inhibition by P,(Fig. 5), previously shown to 86 Table II. Kinetic Properties of Wild Type and Mutant Forms of Type II Hexokinase Extract from cells transfected with cDNA for Wild type 146 8155A 126 S603A 143 i i i Glucose (uMl' 22 15 10 Km for ATP 0.64 1.09 0045 mM " i i i 0.04 0.20 0.05 ‘Mean i SD for three determinations, each with an extract from a different transfection. bMean f SD for six determinations, each with an extract from a different transfection. 87 100 80 29 .g 60 O 4 40 s 20 1,5-Anhydroglucitol (uM) Figure 4. Inhibition of wild type and mutant forms of Type II hexokinase by the Glc-6-P analog, 1,5-anhydroglucitol-6-P. Activity was determined as a function of increasing inhibitor concentrations, with initial ATP concentration in each assay equal to the Km value (Table II) for the respective enzyme. (. ) , Wild type Type II hexokinase; (CI) , 8155A; (0) , 8603A. ‘88 100 80 z.» .5 60- 0 <1 40 s 20 Phosphate (mM) Figure 5. Inhibition of wild type and mutant forms of Type II hexokinase by inorganic phosphate. Activity was determined as a function of increasing phosphate concentrations, with initial ATP concentration in each assay equal to the Km value (Table II) for the respective enzyme. (0) , Wild type Type II hexokinase; (D), 8155A; (0), 8603A. 89 inhibit competitively vs. ATP (21). Similar results (not shown) were seen for inhibition by the P,analogs, sulfate and arsenate. Retention of activity in a chimeric hexokinese. constrUcted from an inactivated C-terminal (catalytic) domaip of Type I.and the N-terminal half’of Type II.hexokinase. It is apparent that the most straightforward interpretation of the above results is that both N- and C-terminal halves of Type II hexokinase possess intrinsic, and comparable, catalytic activities. Thus, mutation of a catalytically important Ser residue in either half results in reduction of specific activity to somewhat less than half of wild type, while the double mutation results in almost complete loss of activity. This was further examined by an alternative strategy based on previous work from this laboratory (10,15). Baijal and Wilson (10) showed that mutation of Ser 603 to Ala resulted in virtually complete loss of catalytic activity in the Type I isozyme. Using the same strategy previously used to prepare chimeric enzymes from complementary halves of the Type I and Type II isozymes (15), a chimera was prepared in which the N-terminal half of Type II was combined with a C- terminal half of Type I, but with Ser 603 changed to Ala in the latter (10). It.has previously been shown (10,15) that the wild type Type I isozyme, the Type I mutant S603A, and the NIICI chimera (comprised of the N-terminal half of Type II hexokinase and the C-terminal half of the Type I isozyme) are 90 all expressed at similar levels under these conditions. This was confirmed in the present study and, based on staining intensity' 'with Coomassie Blue, the mutated chimera NIICI(8603A) was also expressed in amounts comparable to the other forms (results not shown). Activities seen in extracts of C081 cells transfected with these constructs are shown in Table III; similar results were seen in other experiments employing various subsets of these cDNAs. In confirmation of previous reports, these results demonstrate the devastating effect of the 8603A.mutation on activity of the Type I isozyme (8,10) and the enhanced activity seen with the NIICI chimera (15), the latter being the observation that led to the present study (see comments in Introduction). Most interesting in the present context is the last entry in Table III, demonstrating retention of substantial catalytic activity in the chimeric enzyme produced from the N-terminal half of Type II and an inactive C-terminal half of Type I hexokinase. This clearly supports assignation of catalytic activity directly to the N- terminal half of the Type II isozyme. DISCUSSION The catalytic sites in the N- and C-terminal halves of the Type II isozyme appear to function more-or-less independently' - for» example, there is no indication of cooperativity in binding of substrates. Moreover, they have 91 Table III. Hexokinase Activity in Extracts from C081 Cells Transfected with cDNAs Encoding Wild Type and Chimeric Forms of Hexokinase cDNA used for Hexokinase Activity transfection lunits/mg protein) Sham 0.05 Type I 1.14 Mutant Type I(S603A) 0.07 NIICI chimera 1.87 NIICI(8603A) chimera 0.66 92 comparable V5“ values, with mutation of analogous catalytically important residues (Ser 155 or Ser 603) in either half having similar effect on specific activityfl. The two sites do differ significantly in Km for ATP but not for Glc, with the catalytic site in the N-terminal half having the higher apparent affinity for this nucleotide substrate. In principle, differences in KIn should lead to deviation from monophasic Michaelis-Menten kinetics, but this would not be expected to be detectable with the relatively slight difference in the Kms of the N- and C-terminal sites for ATP (22). Indeed, simulation of kinetic results that would be expected from two sites functioning with the same V,m but with Khs for ATP of 1.09 and 0.45 mM (Table II) was consistent with monophasic Michaelis-Menten behavior and an apparent K, of 0.66 mM, virtually identical with the value found for the Type II isozyme (Table II). Deviation from simple Michaelis-Menten behavior might also be predicted for the single mutants, 8155A and 8603A, since they were comprised of N- and C-terminal halves with markedly different Vmax values (22). However, this Ser to Ala mutation results in reduction of Vpu to approximately 5% of Vm for the unmodified enzyme (8,10, and present work). Simulation demonstrates that the kinetics would be dominated by the unmodified site, consistent with the observation of monophasic Michaelis-Menten behavior for the mutant enzymes. 93 Okazaki et al. (23) have previously reported that the C- terminal half of Type II hexokinase was preferentially labeled by the ATP analog, 2’,3'-dialdehyde ATP. These authors took this to indicate that the catalytic site was associated solely with the C-terminal half, as with the Type I isozyme (9- 11,13). The present’ study indicates that an alternative explanation is necessary,‘ perhaps a difference in .the effectiveness with which this analog labels the ATP binding sites in the N- and C-terminal halves of Type II hexokinase. This is not unreasonable since the differences in Km for ATP do, in fact, indicate that the two sites are not identical. The two sites also