Abstract
The structure of mitochondrial pyruvate dehydrogenase kinase isozyme 2 is of interest because it represents a family of serine-specific protein kinases that lack sequence similarity with all other eukaryotic protein kinases. Similarity exists instead with key motifs of prokaryotic histidine protein kinases and a family of eukaryotic ATPases. The 2.5-Å crystal structure reported here reveals that pyruvate dehydrogenase kinase isozyme 2 has two domains of about the same size. The N-terminal half is dominated by a bundle of four amphipathic α-helices, whereas the C-terminal half is folded into an α/β sandwich that contains the nucleotide-binding site. Analysis of the structure reveals this C-terminal domain to be very similar to the nucleotide-binding domain of bacterial histidine kinases, but the catalytic mechanism appears similar to that of the eukaryotic serine kinases and ATPases.
Mitochondrial pyruvate dehydrogenase complex (PDC)1 is a large, multienzyme complex composed of multiple copies of its four components: E1, E2, E3, and E3-binding protein. PDC catalyzes the irreversible conversion of pyruvate to acetyl-CoA (1), which then proceeds to oxidation through the tricarboxylic acid cycle or conversion to fatty acids for storage (2). The flux through this complex is controlled by the phosphorylation state of three serine residues on the E1α subunit of pyruvate dehydrogenase (E1). When glucose and insulin levels are high, the serine residues are dephosphorylated by an active pyruvate dehydrogenase phosphatase, which results in an active complex and the flow of energy from glycolysis to oxidation or storage. In fasting or diabetes, the kinase associated with the complex (3) is activated and also induced, resulting in phosphorylation and inactivation of the E1 enzyme. This serves to preserve the three-carbon skeleton of the pyruvate for gluconeogenesis, which maintains serum glucose levels during fasting but also contributes to pathologic hyperglycemia in diabetes (1).
The activity of the kinase is, in part, regulated by the substrates and products of the pyruvate dehydrogenase reaction. The substrates (pyruvate, NAD+, and CoA) inhibit kinase activity, whereas the products (NADH and acetyl-CoA) stimulate kinase activity. It is generally believed that only pyruvate directly regulates the kinase activity through a dedicated allosteric site. The effects of NAD+/NADH and CoA/acetyl-CoA are mediated by the oxidation, reduction, and acetylation state of the lipoyl group (4). Structurally, this domain is an 80-amino acid, free-folding domain in the N-terminal region of E2. It is connected to the globular C-terminal E2 domain, which forms the icosahedral core of the PDC, by an extended, unstructured linker sequence (5). Pyruvate dehydrogenase kinase (PDK) binds to the lipoyl domain, where it monitors changes in the reduced and acetylated states of the lipoic acid ligand and adjusts its activity accordingly.
Based on its substrate specificity, PDK2 is a strictly serine-specific protein kinase. Structurally, all serine/threonine kinases identified to date can be classified as one large superfamily (6). However, sequence determination of PDK showed that it lacked the characteristic structural elements of serine/threonine kinases (7) explicitly described by Hanks and Hunter (6). Subsequent work has resulted in the identification of ~25 genes in the PDK family. Four of these PDKs are expressed in mammalian cells. PDK2, the subject of this study, appears to be the most abundantly expressed isozyme (8). Comparison of these sequences shows that these proteins typically consist of two almost equal parts: a poorly conserved N-terminal half and a highly conserved C-terminal half. Within the C-terminal domain, there are four uniformly conserved subdomains, the N box (Glu-X-X-Lys-Asn-X-X-X-Ala), the G1 box (Asp-X-Gly-X-Gly), and the G2 box (Gly-X-Gly-X-Gly), which are analogous to the motifs found in bacterial histidine kinases (their occurrence in the topology of PDK2 is shown in Fig. 1d). The results of site-directed mutagenesis experiments showed that the nucleotide-binding domain is in the C-terminal half of PDK2, and that the G subdomains are important in nucleotide binding (9). This x-ray crystallographic study was undertaken to determine the protein fold and the catalytic mechanism of unusual serine-specific kinase, as well as to gain insight into the molecular mechanisms responsible for regulation of the kinase activity.
FIG. 1. Dimer/monomer structure.
a, ribbon representation of the PDK2 dimer viewed down the non-crystallographic 2-fold axis. α-Helices are colored cyan, and β-strands are colored yellow. b shows one protomer rotated by ~60° to more clearly define the domain structure. The α-helices and β-strands are labeled. Note that this figure is in the same orientation as the topology diagram in d. c shows a third view of the dimer, with the non-crystallographic 2-fold axis in the plane of the paper. d, diagram showing the secondary structural elements with their numbers as referred to in the text and also the topology of the PDK2 monomer. The diagram is oriented as in b. The ribbon diagrams were generated in Swiss-PdbViewer (38) and rendered with POVRay for Windows (downloaded from www.povray.org). Several short flexible loops not observed in the electron density maps were modeled using Swiss-Pdb-Viewer (38) and are indicated by hash marks.
MATERIALS AND METHODS
PDK2 from Rattus novegicus was cloned into a Novagen Pet28a vector and co-expressed in BL21[DE3] cells with the GroESL plasmid. Induction and purification using the N-terminal His6 tag were performed as described by Popov et al. (10), with an approximate yield of 40 mg of PDK2 from 1 liter of cells. After passage through a MonoQ (Amersham Pharmacia Biotech) anion exchange column, the protein was concentrated (Millipore, Centricon 10) to 10 mg/ml and crystallized via vapor diffusion from a solution of either 100 mm imidazole or 100 mm MES, pH 6.5−7.0, 500 mm KCl, 3.5 mm 3-iodopropionic acid, 3.5 mm ADP, 7 mm MgCl2, 6.5% polyethylene glycol 6000, and 1.0% ethylene glycol. The crystals grew in 1 week to ~0.3 × 0.3 × 0.1 mm. Cryoprotection was achieved with the well solution + 25% glycerol in stepped (5, 12.5, and 25%) increments. Selenium protein was produced (11) and purified in the same fashion, substituting a host bacterium auxotrophic for methionine (Novagen B834) and providing selenomethionine in a defined growth medium. Crystallizations in the case of the selenomethionine protein were carried out in a nitrogen atmosphere but were otherwise identical to the native conditions.
A three-wavelength multiwavelength anomalous diffraction (12) experiment was conducted on the selenomethionine protein crystals at 110 °K at the X12C beamline of the National Synchrotron Light Source, Brookhaven National Laboratory. These crystals diffracted to 2.8 Å and were collected in a highly redundant fashion. The data were processed utilizing DENZO (13) with the statistics listed in Table I, and care was taken to keep the Freidel pairs independent. A Patterson self-rotation function performed on the selenomethionine data in X-PLOR (14) identified a strong non-crystallographic symmetry axis in the PDK2 data, with a peak height 94% of that of the crystallographic 2-fold axis. The program SOLVE (version 1.0) (15) was used to locate the positions of the selenium diffractors, utilizing the MAD experiment as a special case of multiple isomorphous replacement (16). 21 of the 36 anticipated selenium sites were located and used in the initial estimation of phases. From the selenium positions, the exact positioning of the non-crystallographic symmetry axis with respect to the unit cell was identified. An initial electron density map showed the outline of the two monomers in the asymmetric unit cell predicted by analysis of the cell dimensions using the Mathews equation. The map was improved with density modification and 2-fold non-crystallographic symmetry averaging (17). In this map, α-helical and β-strand secondary structural elements were clearly visible. Approximately 40% of the structure could be built from this map with O (20) using the selenium positions to identify methionine residues and so help set the register of the model. Phase improvement was accomplished by incorporating the experimental and model phases in a Sigma A weighted protocol (17). More of the model was built as it became evident with the improved phases. Phase refinement was initially done in X-PLOR (14) to an Rfree of ~32%, after which no improvement was achieved. Reexamination of the data indicated that the crystals were pseudomerohedrally twinned along the b axis, with a statistically estimated twin fraction of 45% (18). Reexploration of crystallization space yielded native crystals diffracting to >2.5 Å, but the twinning remained. With the added resolution, phase refinement of the model could be shifted to SHELXL, which was capable of refining the pseudomerohedrally twinned x-ray diffraction data (19). Final rounds of model building and phase improvement were accomplished with building in O (20) and refinement in SHELXL (21).
TABLE I.
Data and refinement statistics
Data set | Sel-inflection | Sel-peak | Sel-remote | Native |
---|---|---|---|---|
Location | X12C | X12C | X12C | Home (Raxis) |
Space group | P21 | P21 | P21 | P21 |
Wavelength (Å) | 0.9803 | 0.9802 | 0.9300 | 1.54 |
Cell (Å) | ||||
a = | 71.72 | 71.62 | 71.77 | 71.42 |
b = | 109.38 | 109.38 | 109.38 | 109.88 |
c = | 71.62 | 71.62 | 71.80 | 71.41 |
β (°) = | 99.5 | 99.5 | 99.5 | 104.5 |
Resolution (Å)a | 2.8 (2.90–2.80) | 2.8 (2.90–2.80) | 2.8 (2.90–2.80) | 2.5 (2.60–2.50) |
I/Sigla | 15.4 (2.7) | 29.6 (5.8) | 15.6 (3.0) | 23.1 (4.2) |
Redundancy | 3.3 | 8.3 | 3.1 | 3.3 |
Unique reflections | 26326 | 27236 | 25957 | 35685 |
Completeness (%)a | 95.8 (80.6) | 99.6 (97.3) | 95.0 (94.0) | 88.3 (76.1) |
Rmergea,b | 8.7 (34.9) | 9.1 (33.2) | 8.4 (33.9) | 4.2 (28.7) |
Se Sites | 21 | 21 | 21 | n/a |
Refinement | ||||
Atoms | 5590 | |||
Twinning fraction | 0.467 | |||
Rwork/Rfreec | 19.9/25.7 | |||
RMS deviation | ||||
Bonds (Å) | 0.008 | |||
Angles (°) | 1.4 | |||
Average B factor (Å2) | 35.9 | |||
Main chain (Å2) | 28.4 | |||
Side chain (Å2) | 42.7 |
Data from the last resolution bin is shown in parentheses.
Rmerge = ΣhklΣi∣I -〈I 〉∣/Σ〈I〉, whereI is the intensity.
R = Σ∣Fobs∣−∣Fcalc∣/Σ∣Fobs∣.Rfree is a control monitoring 1600 reflections left out of the refinement.
The final model contains 681 amino acid residues (85% of the molecule), 238 water molecules, and two ADP molecules (see Table I for refinement statistics). The missing sections of the molecule are two loops between N-terminal helices (residues 34–39 and 130–135), a loop between domains (residues 172–176), the “lid” over the nucleotide-binding site (residues 290–307), and the C terminus of the molecule after residue 365. The Ramachandran plot shows good geometry in the model, with 86.0% of the amino acid residues in the most favored region and 14.0% in the additionally allowed regions (PROCHECK (17)).
RESULTS AND DISCUSSION
Structure of the Monomer
The monomer of PDK2 folds into two well defined domains (9). The N-terminal domain extends from residue 1 to residue 179 and consists of 7 α-helices connected by flexible loops. Four of these long helices are amphipathic and form a helical bundle with a hydrophobic core (Fig. 1, a and b). The α5/α6 helices (Fig. 1b) almost form one continuous long helix except for a short loop region where Pro-120 disturbs the helical secondary structure. This “kink” in the helix orients the conserved His-115 toward the interior of the helical bundle (see Fig. 4d). The positioning of this residue is of particular interest as it is conserved throughout the PDK family of kinases and was a candidate as a possible phosphohistidine intermediate (22). Its orientation within the helical bundle, where it is not solvent-accessible, makes it unlikely that His-115 fills this role.
FIG. 4. Detailed interactions in the nucleotide-binding pocket and with His 115.
a, adenine binding. This figure shows the matrix of hydrogen bonding between the adenine group of ADP, the conserved hydration sphere, and those amino acid side chains that are important in stabilization of the binding pocket. Note the adenine specific hydrogen bonds between Asp-282 (G1 loop) and the adenine ring. The matrix of “low temperature factor” water molecules (red crosses) mediates a number of interactions between the protein and the nucleotide. Asn-247 (N-box) is shown interacting with the magnesium ion near the position of the ADP phosphate groups. b, phosphate interaction with the G2 loop. This figure shows the presumed position of the γ-phosphate of ATP modeled into the PDK2 structure from the structure of MutL (see under “Results and Discussion”). The proposed hydrogen bonding from the residues of the G2 loop to the γ-phosphate is shown. The prominent red cross is a water molecule found in the MutL structure that occupies the proposed position for the substrate serine hydroxyl group. Distances to the important catalytic residues Glu-243 and Lys-246 (both in the N-box motif) are shown. Glu-243 is proposed to be a general base to activate the serine hydroxyl group, and Lys-246 may stabilize the developing negative charge. See text for a complete discussion. c, Mg2+ binding. This figure shows the coordination of the Mg2+ ion to the α- and β-phosphate oxygen atoms, residue Asn-247 from the N-box, and two low temperature factor water molecules. The coordinating atom to complete the classic coordination sphere of the magnesium ion was not found in the electron density map. This position may be occupied by a γ-phosphate oxygen atom, as it is when the MutL ATP moiety is modeled into the position of the nucleotide in PDK2. See text for discussion. d, N-terminal His-115 binding within the helical bundle. This figure shows the hydrogen bonding pattern of residue His-115 within the N-terminal helical bundle. This histidine is conserved throughout the PDK family of kinases. It is also present in the histidine kinases, in which it forms a covalent linkage with the γ-phosphate before transfer to the activating aspartate residue. However, His-115 in PDK2 is not solvent-accessible, instead making a number of important structural hydrogen bonds within the N-terminal helical bundle. Thus, it appears to have evolved an important structural role in the PDK family, rather than the original role as intermediate host to the γ-phosphate. These figures were generated using the programs O (20) and Oplot.
Within the helical bundle, His-115 is stabilized by a network of hydrogen bonds between the imidazole ring and Ser-83 on the adjacent α4 helix and via a well ordered water molecule to Tyr-80 and Arg-154 (see Fig. 4d). The result of this irregularity in the long α5/α6 helix is to expose to solvent a number of arginine residue side chains. These positive charges are concentrated in a patch on the floor of the crevice that is formed between the N- and C-terminal domains (see Fig. 5).
FIG. 5. Electrostatic surface of PDK2 compared with the lipoyl domain.
This figure shows a surface map of the PDK2 monomer and a modeled lipoyl domain. The colors are based on charge density: red, negative charge; blue, positive charge; white, neutral or hydrophobic. Both surface models were generated in Swiss-PdbViewer (38) by mapping the Poisson-Boltzmann electrostatic potential of the protein to colors on the accessible molecular surface. Rendering was done with POV-Ray for Windows (Cason, C., www.povray.org). A model of a lipoyl domain was produced from the NMR structure of the L2 domain of human PDC (1FYC) with the rat L2 sequence (90% identical to human) threaded on and energy minimized (38) The colors are defined as for the PDK2 surface. The PDK2 monomer is rotated away from the dimer interface to show the cleft between the N- and C-terminal domains that is compatible in shape and size to the lipoyl domain.
The C-terminal nucleotide-binding domain is folded as a mixed α/β structure (Fig. 1, a, b, and d). The core consists of five β-strands forming a single β-sheet. Four α-helices are arranged on one face of the sheet toward the N-terminal domain. A hydrophobic core formed from side chains as the helices and strands pack together stabilizes the domain. The opposite face of the β-sheet is in contact with the equivalent β-sheet of the other monomer in the asymmetric unit and forms the dimer interface (Fig. 1, a and c). There is an unusual “left-handed” arrangement of β2-α9- β3 (Fig. 1b), which has also been observed in bacterial histidine kinase nucleotide-binding domains (23). The electron density for the ADP co-crystallized with PDK was quite strong and was located between the α-helices and the β-sheet as it widens toward one end (see Fig. 3a).
FIG. 3. Electron density of ADP and the catalytic residues.
a, ADP binding pocket. This figure shows a section from the final electron density map in the region of the nucleotide-binding pocket. The position assigned to the magnesium ion is depicted as an orange cross on the right. The G1 loop is on the left, where it is in position to interact with the adenine moiety. The G2 loop is on the right, interacting with the phosphate groups. Glu-243 from the catalytic N-box is noted in the upper right. Water molecules are depicted as red crosses. The ones noted here have low temperature factors, and so are thought to be quite stable. b, N-box and catalytic residues. This figure shows a diagram of the N-box of PDK2 and those residues thought to take part in the catalytic mechanism. The positioning of the Mg2+ ion is well shown here, depicted as an orange cross in the middle. The important residues of the N-box are on sequential turns of helix α9, which goes from right to left in the figure. The first is His-239, which is thought to orient and polarize Glu-243 (see under “Results and Discussion”). The next turn has Glu-243, which is thought to act as the general base for the reaction. This side chain is in proximity to the magnesium ion and the closely associated coordinating water molecules (red crosses). Lys-246 is on the third turn (see Fig. 4b) and may interact with the developing negative charge on the serine as the general base, Glu-243, activates it. Residue Asn-247 in the N-box motif coordinates with the magnesium ion (see Fig. 4, a and b). Both panels were generated using the software packages O (20) and Oplot, and the 2Fo - Fc electron density map was contoured at a level of 1.5 σ.
The overall fold of the C-terminal nucleotide-binding domain of PDK2 is remarkably similar to prokaryotic histidine kinases and ATPases. Aligning the published structures of histidine kinases (CheA (24) and EnvZ (25)) and ATPases (DNA Gyrase B (26), MutL (27), and Hsp90 (28)) with TOP3D from the CCP4 (17) suite gives an RMSD for the Cα atoms of less than 2 Å over the bulk of the C-terminal domain (Fig. 2). The topology of the ATPase nucleotide-binding domain is different from that of PDK2 or the histidine kinases, because it has a rearrangement of the secondary structural elements, but the overall fold is preserved.
FIG. 2. Alignment of the PDK2 nucleotide-binding domain with prokaryotic histidine kinases and ATPases.
This figure shows a composite of the Cα tracings of the ATP-binding domains of the structurally homologous ATPases Hsp90 (Protein Data Bank (36) code 1AM1) residues 26–209, MutL (1B62) residues 22–201, and DNA Gyrase B (1EI1) residues 452–619, as well as the histidine kinase CheA (1B3Q) residues 354–538, all of which were aligned with the C-terminal domain of PDK2, residues 183–353, using the program TOP3D (17). The backbone for all structures is shown in gray, and the nucleotide is shown in color. Visualization was done with Swiss-PdbViewer (38), and rendering was done with POV-Ray for Windows.
The C-terminal and N-terminal domains of PDK2 are connected by a flexible, poorly ordered loop from Ser-171 to Pro-177 (α7/α8). They are stabilized relative to each other by the antiparallel interaction between β1 and β7, as well as hydrophobic interactions between the side chains of helices α4/α5/α6 in the N terminus and α11 in the C terminus. As a result of this association, there is a rather large solvent-exposed cavity between the two domains (see Fig. 5). Helices α9 and α11 form one side of the cavity and the N-terminal helices α4/α5/α6 the other side. Due to the kink in helix α5/α6 noted above, one side of the cavity carries a positive charge because of the exposed arginine side chains.
Dimer Interactions
Biochemical evidence suggests that PDK2 is a dimer. The asymmetric unit in the crystal structure of PDK2 contains two molecules related by a strong non-crystallographic symmetry axis, and these two molecules are presumed to form the functional dimer. Protomers in the dimer are in a head-to-head orientation, with the primary interaction between the β sheets of the C-terminal domains (Fig. 1a). The interface buries 2200 Å2 (18), or 12.4% of the total surface area (17,763 Å2 for monomer; crystallography and NMR system (CNS) (18), 1.4 Å probe) and is stabilized by a number of hydrophobic interactions and hydrogen bonds. The remainder of the molecules extends out in a horseshoe configuration, with the two N-terminal domains in the arms and the C-terminal domains forming the dimer interface at the base. In this configuration, the two nucleotide binding pockets are located on the outer, solvent-exposed surface of the dimer and are readily accessible.
ATP-binding Site
As mentioned previously, PDK2 shows sequence homology to the histidine kinases, specifically in the G1, G2, and N-box motifs. Observing the presence of these characteristic motifs in PDK2, Bowker-Kinley and Popov (9) have suggested that PDK belongs to the ATPase/kinase superfamily, which includes histidine protein kinases (EnvZ and CheA) and ATPases (Hsp90, DNA Gyrase B, and MutL). The present study confirms this observation, because fitting of the structure of PDK2 to other members of the family, the structures of which are known, shows that the overall fold of the C-terminal nucleotide-binding domain of PDK2 is structurally homologous to other members of this family. This domain similarity is shown in Fig. 2, in which the Cα tracing of the nucleotide-binding domains of the histidine kinase CheA and the ATPases Gyrase B, MutL, and Hsp90 are superimposed on PDK2. This structural homology is remarkable given that only the ADP/ATP moiety was used to fit these molecules. In addition, x-ray structural studies have shown that there is extremely high structural homology within the ATP-binding pocket in the histidine kinases and ATPases (29).
In the case of PDK2, the electron density corresponding to the bound ADP was quite clear, as is demonstrated in Fig. 3a, and was located at the center of a cone-shaped cavity formed between the β-sheet and the α-helices of the C-terminal domain (Fig. 1b). The ADP molecule itself adopts a skewed orientation (Figs. 2 and 3a), in contrast to the more open arrangement seen in cytoplasmic serine/threonine kinases but similar to that found in DNA Gyrase B, Hsp90, and MutL. The conserved G1 motif (residues Asp-282-X-Gly 284-X-Gly 286) is involved in binding the adenine moiety of the nucleotide. In PDK2, it forms one side of the binding pocket (Fig. 3a, left) and is a portion of the loop connecting helices α10 and α11 (residues 296–320). The G1 box is the basis of the adenine specific interactions between the nucleotide and the protein (compare Figs. 3a and 4a). These include a direct hydrogen bond from the C6 amino group (N-6) of the adenine ring to the carboxylate side chain of Asp-282 in the G1 box. This same aspartate residue also provides one of the three protein ligands for the water molecule that hydrogen bonds to the N1 imino nitrogen of the adenine ring (Fig. 4a). Two other protein ligands for this water molecule come from the side chain of the invariant Thr-346 and the peptide nitrogen of the invariant Gly-286 (G1 box). These interactions provide the specificity of PDK2 for adenyl nucleotides. The N-7 of the adenine ring is bound to another highly ordered water molecule that is part of an extensive hydrogen bonding network of water molecules and ionic side chains that extends across the nucleotide-binding site (Figs. 3a and 4a).
The central portion of this loop containing the G1 box extends over the nucleotide-binding site. In the family of ATPases, this loop forms a lid that is thought to limit access of ATP to the binding site (23) and may serve a similar role in PDK2. Part of this region is often disordered in the structures of the ATPases/kinases and also in PDK2, suggesting a high degree of mobility.
The other side of the binding pocket distal to the lid corresponds to the G2 box (Gly 317-X-Gly319-X-Gly321), again a conserved motif found in bacterial histidine kinases and structurally homologous ATPases (Fig. 3a, right) (9). The G2 box curves around the α- and β-phosphate groups of the ADP. These phosphate groups are stabilized by a number of direct hydrogen bonds between their oxygen atoms and the amide nitrogen atoms of residues Gly-317, Phe-318, Tyr-320, and Gly-321 as shown in Fig. 4b. Interestingly, Gly-319 (Fig. 4b) adopts a glycine-specific ϕ angle that is essential to the curve of this loop and the formation of these stabilizing hydrogen bonds, thus explaining the “dead” kinase when this residue is mutated (9). The positioning of the phosphate groups takes advantage of the helix dipole from α11 to stabilize their polyanionic character.
Additional interactions between the ADP and the protein include a direct hydrogen bond between the 2′ hydroxyl group of the ribose and the side chain of the invariant residue Thr-302 (ATP lid). The carbonyl oxygen of Thr-302 interacts with the 3′ hydroxyl group of the ribose, which, in turn, forms a hydrogen bond to the side chain of the invariant Arg-250 protruding from helix α9 across the nucleotide-binding pocket. The position of Arg-250 is stabilized through interactions with Glu-254 (helix α9) and the solvent hydrogen bonding network.
The electron density associated with the magnesium ion coordinated to each ADP molecule is easily seen (Fig. 3b). Each phosphate group provides an oxygen atom to coordinate with the Mg2+ ion (Fig. 4c). The side chain oxygen of Asn-247 together with two water molecules completes the observed coordination sphere (Fig. 4c). Asn-247 is part of the conserved N-box motif. These residues (Glu-243-X-X-Lys 246-Asn-247-X-X-Ala-250) are located on helix α9 (see Fig. 3b and Fig. 4, b and c) with the conserved residues occurring on successive turns of the helix. Glu-243 is thought to act as a general base in the reaction mechanism (see discussion below), whereas Asn-247 coordinates with the magnesium ion. The presumed sixth position to finish the classic octahedral coordination of the Mg2+ does not contain sufficient electron density to warrant the addition of a water molecule to the model. It is likely that the γ-phosphate of ATP will occupy this position when ATP is bound (see below).
Possible Mechanisms of Catalysis and Regulation of Enzymatic Activity
Because this structure was solved with bound nucleotide, a number of conclusions may be drawn about a possible mechanism of catalysis for the phosphate-transfer reaction. The position of the nucleotide within the binding pocket predetermines the approximate location of the γ-phosphate. A least-squares fitting program from the CCP4 suite (TOP3D) (17) was used to map the coordinates of the ATP-analogue found in the structure of MutL (30), the ATP-binding site of which is structurally homologous to that of PDK2 (see previous discussion), as a rigid body onto the ADP in the PDK2 structure. The adenine, ribose, α-phosphate, and β-phosphate groups of the MutL structure line up exactly with the ADP found in PDK2, and the resulting position of the γ-phosphate makes biochemical sense. The modeled γ-phosphate oxygen atoms could make four hydrogen bonds to the peptide nitrogen atoms of the G1 loop residues (Fig. 4b). In addition, the Mg2+ ion is in a position to coordinate to one of the γ-phosphate oxygen atoms, thus filling the sixth position that is missing in our model. This model would still conserve the stabilization of the polyanionic ATP by the helix dipole from the N-terminal end of helix α11.
Given this active site conformation, it is reasonable to ask which residues are in a position to carry out the chemistry of phosphate transfer and if formation of a phospho-histidine intermediate is reasonable. Examination of the structure of PDK2 in the vicinity of the putative γ-phosphate binding site provides no obvious candidate for a residue to carry out a nucleophilic attack. The only histidine conserved throughout the PDK family, and therefore a candidate for a histidine-phosphate intermediate, is His-115 (22). However, it is involved in making a number of important structural hydrogen bonds in the N-terminal domain and is not solvent-accessible (Fig. 4d). Given the absence of a nucleophile, it is quite likely that PDK uses a general base catalysis to activate the hydroxyl group of the substrate seryl residue for direct attack on the γ-phosphate group. It is interesting to examine the PDK2 catalytic pocket in light of what has been concluded about the very closely related catalytic pockets of the ATPases and histidine kinases.
In the ATPases/bacterial histidine kinases, the same basic structure is used for catalysis of histidine phosphorylation or ATP hydrolysis (i.e. transfer of the phosphate to water). In the histidine kinases, the phosphoacceptor (a histidine residue in the N-terminal domain) is itself a quite reactive nucleophile, and it has been suggested that it directly attacks the γ-phosphate of ATP. The nucleophilicity of the histidine residue is further increased through interaction with a neighboring glutamate residue (His-47 and Glu-70 in CheA) (31). It is noteworthy that both of these residues lie in the N-terminal phosphate acceptor domain of their respective kinases. Thus the N-terminal domain complements the catalytic center by providing a nucleophile for phosphotransfer and an activating glutamate residue. It is evident from our structure that this N-terminal complement to the catalytic pocket is not present in PDK2.
In the ATPases (DNA Gyrase B and Hsp90), the nucleophile (water) is much less reactive and needs to be activated in order to attack the γ-phosphate. In DNA Gyrase B, the activating glutamate (Glu-42) acts as a general base and is located within the ATPase nucleotide-binding pocket (32). Also, it is polarized and aligned by the neighboring His-38. Thus, the ATPase nucleotide-binding domain contains all of the circuitry needed for catalysis. This appears to be similar to the arrangement of catalytic residues in the nucleotide-binding domain of PDK2. The glutamate residue corresponding to Glu-42 of the DNA Gyrase B is Glu-243 (N-box, Fig. 3b) in PDK2 and is uniformly conserved in all mitochondrial protein kinases identified to date. It is positioned 3 Å from a water molecule in PDK2 that occupies the likely position for the serine substrate (Fig. 4b). This suggests that the mitochondrial protein kinases might use a mechanism similar to the general base catalysis of the ATPases to activate the serine residue of the protein substrate. In addition, His-239 of PDK2 might correspond to His-38 of DNA Gyrase B and may be involved in polarization and positioning of Glu-243 (Fig. 3b). It is in line with the modeled γ-phosphate group and located one turn along the α9 helix from residue Glu-243. The imidazole side chain of His-239 is thus positioned to orient the glutamate residue to interact with the putative serine substrate position. At the same time, the presence of His-239 could modify the effective pKa of Glu-234 to an appropriate range for the role of a general base in the reaction. It thus appears that the active site of PDK2 is capable of activating the serine substrate and transferring the γ-phosphate directly, without the necessity of a covalent intermediate.
The positive charge to stabilize the transition state, represented by Lys-246 (N-box) in PDK2 (Figs. 3b and 4b), is also present in the ATPases. Examination of sequence alignments of ATPases shows a variety of residues at the PDK2 Lys-246 position. However, the available x-ray crystal structures demonstrate that there is a positively charged residue in the catalytic pocket in all of the ATPases that is able to occupy the position in three-dimensional space of the lysine ∈-nitrogen. Examples of this would be Lys-737 in Gyrase B and Lys-307 in MutL. In other words, when these structures are aligned to the PDK2 model with TOP3D (17), they position a positive charge less than 2 Å from the ∈-nitrogen of Lys-246 in our structure. Given this catalytic similarity of the ATP-binding domain of PDK2 with the family of ATPases, it is not surprising that PDK2 has been found to also possess an ATPase activity, representing 1–3% of total kinase activity (33).
In contrast to the ATPases, an alignment of multiple histidine kinases to the PDK2 model shows that the position of the general base Glu-243 of PDK2 is occupied by a variety of residues and the positive charge contributed by Lys-246 in PDK2 is entirely absent. Thus, although the overall fold of the PDK2 is that of a histidine kinase, the chemistry of the active site is similar to that of an ATPase and is probably capable of a direct transfer of the γ-phosphate to the serine substrate.
PDK2 and the Pyruvate Dehydrogenase Complex
It is generally believed that the enzymatic activity of PDK is intimately dependent upon interactions with protein substrate and supporting structures of the PDC. Several lines of evidence suggest that the kinase is tightly bound to the lipoyl-bearing domain 2 of the E2 component of the complex (34). Association with the lipoyl-bearing domain appears to be responsible for the almost 10-fold increase in kinase activity as well as conferring the sensitivity of the kinase activity to metabolite concentrations (35). As noted above, the lipoyl domains are self-folding structures carrying a lipoic acid prosthetic group that is acetylated by the acetyl-TPP-E1 complex formed after pyruvate is decarboxylated.
The three-dimensional structure of these domains is conserved across species as a flattened β-barrel with a hydrophobic core and many exposed acidic residues (5). As noted earlier, examination of the three-dimensional structure of the PDK2 monomer revealed the existence of a rather large cavity between the C- and N-terminal domains (Fig. 5). The size and shape of the cavity appears to be complementary to the size and shape of the lipoyl-bearing domain. The cluster of positively charged residues exposed in the floor of the cavity by the discontinuity in helix α5/α6 is ideally located to interact with the negatively charged residues of the lipoyl domain. These observations bring about the intriguing possibility that the pocket between the domains of the monomer might comprise the lipoyl domain binding site. Orienting the lipoyl domain to fit into this putative binding cleft positions the lipoic acid ligand close to the nucleotide-binding pocket. This suggests that the lipoic acid could interact directly with the residues near the nucleotide-binding pocket. In this way, the state of the lipoic acid (acetylated, reduced, or oxidized) would be in a position to influence kinase activity.
The two putative lipoyl binding pockets are located on the outside of the PDK2 dimer, oriented in opposite directions. This suggests that PDK2 may bind to more than one lipoyl domain at a time. It has been shown that only a few kinase molecules are associated with any one PDC, and they must service many E1 molecules (34). This activity is presumed to involve the kinase moving over the surface of the complex without dissociation. The mechanism proposed for this movement is a “hand over hand” motion in which the kinase alternately binds and releases E2 domains while remaining bound to the complex (37). The orientation of the proposed binding sites in this model of PDK2 makes this alternating, multiple binding a very reasonable possibility.
Conclusions
The three-dimensional x-ray structure of PDK2 presented here clearly shows that this molecule shares a lineage with modern prokaryotic histidine kinases and ATPases. The fold of the nucleotide-binding C-terminal domain has been exactly conserved throughout this family to preserve the positioning of the nucleotide in the binding pocket. Variations of particular amino acids over time have allowed this domain to be adapted to three different biochemical roles: ATPase, histidine kinase, and serine kinase. In addition to this insight into molecular adaptation, the structure presented also illustrates the dimer and domain structure of PDK2. Taken together, our results suggest a mechanism whereby PDK2 can bind to the lipoyl domain of the pyruvate dehydrogenase complex and adjust its activity based on the state of the lipoic acid ligand. This has important implications for the modulation of PDC activity in normal intermediary metabolism. In addition, the inappropriate activity of PDK2 in diabetes provides a therapeutic opening for controlling hyperglycemia independent of insulin sensitivity. Clearly, future studies will involve determining the structure of PDK2 in complex with the lipoyl domain in order to better understand this important interaction.
Acknowledgments
We are grateful to Drs. Thomas D. Hurley and David E. Timm for helpful discussions. We also thank Dr. Robert Sweet for help in data collection at the X12C beamline at Brookhaven National Laboratories.
Footnotes
This work was supported by the American Diabetes Association (to J. A. H. and C. N. S.); United States Public Health Service Grants GM51262 (to K. M. P.), DK56898 (to K. M. P.), and DK47844 (to R. A. H.); and the Grace M. Showalter Trust (to R. A. H.).
The atomic coordinates and structure factors (code 1JM6) have been deposited in the Protein Data Bank, Research Collaboratory for Structural Bioinformatics, Rutgers University, New Brunswick, NJ (http://www.rcsb.org/).
The abbreviations used are: PDC, pyruvate dehydrogenase complex; PDK, pyruvate dehydrogenase kinase; PDK2, pyruvate dehydrogenase kinase isozyme 2; E1, pyruvate dehydrogenase component of PDC; E2, dihydrolipoyl acetyltransferase component of PDC; E3, dihydrolipoamide dehydrogenase component of PDC; MES, 2-[N-morpholino]ethane sulfonic acid.
References
- 1.Patel MS, Roche TE. FASEB J. 1990;4:3224–3233. doi: 10.1096/fasebj.4.14.2227213. [DOI] [PubMed] [Google Scholar]
- 2.Randle PJ. Biochem Soc Trans. 1986;14:799–806. doi: 10.1042/bst0140799. [DOI] [PubMed] [Google Scholar]
- 3.Stepp LR, Pettit FH, Yeaman SJ, Reed LJ. J Biol Chem. 1983;258:9454–9458. [PubMed] [Google Scholar]
- 4.Radke GA, Ono K, Ravindran S, Roche TE. Biochem Biophys Res Commun. 1993;190:982–991. doi: 10.1006/bbrc.1993.1146. [DOI] [PubMed] [Google Scholar]
- 5.Green JD, Laue ED, Perham RN, Ali ST, Guest JR. J Mol Biol. 1995;248:328–343. doi: 10.1016/s0022-2836(95)80054-9. [DOI] [PubMed] [Google Scholar]
- 6.Hanks SK, Hunter T. FASEB J. 1995;9:576–596. [PubMed] [Google Scholar]
- 7.Popov KM, Kedishvili NY, Zhao Y, Shimomura Y, Crabb DW, Harris RA. J Biol Chem. 1993;268:26602–26606. [PubMed] [Google Scholar]
- 8.Bowker-Kinley MM, Davis WI, Wu P, Harris RA, Popov KM. Biochem J. 1998;329:191–196. doi: 10.1042/bj3290191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Bowker-Kinley M, Popov KM. Biochem J. 1999;344:47–53. [PMC free article] [PubMed] [Google Scholar]
- 10.Popov KM, Kedishvili NY, Zhao Y, Gudi R, Harris RA. J Biol Chem. 1994;269:29720–29724. [PubMed] [Google Scholar]
- 11.Doublie S. In: Methods in Enzymology: Macromolecular Crystallography, Part A. Carter CW, Sweet RM, editors. Vol. 276. Academic Press; New York: 1997. pp. 523–530. [Google Scholar]
- 12.Hendrickson WA. Science. 1991;254:51–58. doi: 10.1126/science.1925561. [DOI] [PubMed] [Google Scholar]
- 13.Otwinowski Z, Minor W. In: Methods in Enzymology: Macromolecular Crystallography, Part B. Carter CW, Sweet RM, editors. Vol. 277. Academic Press; New York: 1997. pp. 307–326. [DOI] [PubMed] [Google Scholar]
- 14.Brünger AT. X-PLOR, Version 3.1, A System for X-ray Crystallography and NMR. Yale University Press; New Haven, CT: 1992. [Google Scholar]
- 15.Terwilliger TC. In: Methods in Enzymology: Macromolecular Crystallography, Part A. Carter CW, Sweet RM, editors. Vol. 276. Academic Press; New York: 1997. pp. 530–537. [Google Scholar]
- 16.Ramakrishnan V, Biou V. In: Methods in Enzymology: Macromolecular Crystallography, Part A. Carter CW, Sweet RM, editors. Vol. 276. Academic Press; New York: 1997. pp. 538–557. [Google Scholar]
- 17.Collaborative Computational Project Number 4. Acta Crystallogr Sect D Biol Crystallogr. 1999;50:760–763. [Google Scholar]
- 18.Brünger AT, Adams PD, Clore GM, Delano WL, Gros P, Grosse-Kunstlever RW, Jiang JS, Kuszewski J, Nilges N, Pannu NS, Read RJ, Rice LM, Simonson T. Acta Crystallogr Sect D Biol Crystallogr. 1998;54:905–921. doi: 10.1107/s0907444998003254. [DOI] [PubMed] [Google Scholar]
- 19.Yang F, Dauter Z, Wlodawer A. Acta Crystallogr Sect D Biol Crystallogr. 2000;56:959–964. doi: 10.1107/s0907444900007162. [DOI] [PubMed] [Google Scholar]
- 20.Jones TA, Zou JY, Kjeldgaard M. Acta Crystallogr Sect A. 1991;47:110–119. doi: 10.1107/s0108767390010224. [DOI] [PubMed] [Google Scholar]
- 21.Sheldrick GM, Schneider TR. In: Methods in Enzymology: Macromolecular Crystallography, Part B. Carter CW, Sweet RM, editors. Vol. 277. Academic Press; New York: 1997. pp. 319–343. [PubMed] [Google Scholar]
- 22.Harris RA, Popov KM, Zhao Y, Kedishvili NY, Shimomura Y, Crabb DW. Adv Enzyme Regul. 1995;35:147–162. doi: 10.1016/0065-2571(94)00020-4. [DOI] [PubMed] [Google Scholar]
- 23.Dutta R, Inouye M. Trends Biochem Sci. 2000;25:24–28. doi: 10.1016/s0968-0004(99)01503-0. [DOI] [PubMed] [Google Scholar]
- 24.Bilwes AM, Alex LA, Crane BR, Simon MI. Cell. 1999;96:131–141. doi: 10.1016/s0092-8674(00)80966-6. [DOI] [PubMed] [Google Scholar]
- 25.Tanaka T, Saha SK, Tomomori C, Ishima R, Liu D, Tong KI, Park H, Dutta R, Qin L, Swindells MB, Yamazaki T, Ono AM, Kainosho M, Inouye, Ikura M. Nature. 1998;396:88–92. doi: 10.1038/23968. [DOI] [PubMed] [Google Scholar]
- 26.Wigley DB, Davies GJ, Dodson EJ, Maxwell A, Dodson G. Nature. 1991;351:624–629. doi: 10.1038/351624a0. [DOI] [PubMed] [Google Scholar]
- 27.Ban C, Junop M, Yang W. Cell. 1999;97:85–97. doi: 10.1016/s0092-8674(00)80717-5. [DOI] [PubMed] [Google Scholar]
- 28.Prodromou C, Roe SM, O’Brien R, Ladbury JE, Piper PW, Pearl LH. Cell. 1997;90:65–75. doi: 10.1016/s0092-8674(00)80314-1. [DOI] [PubMed] [Google Scholar]
- 29.Bilwes AM, Quezada CM, Croal LR, Crane BR, Simon MI. Nat Struct Biol. 2001;8:353–360. doi: 10.1038/86243. [DOI] [PubMed] [Google Scholar]
- 30.Ban C, Yang W. Cell. 1998;95:541–552. doi: 10.1016/s0092-8674(00)81621-9. [DOI] [PubMed] [Google Scholar]
- 31.Zhou H, Dahlquist FW. Biochemistry. 1997;36:699–710. doi: 10.1021/bi961663p. [DOI] [PubMed] [Google Scholar]
- 32.Jackson AP, Maxwell A. Proc Natl Acad Sci U S A. 1993;90:11232–11236. doi: 10.1073/pnas.90.23.11232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Tuganova A, Yoder MD, Popov KM. J Biol Chem. 2001;276:17994–17999. doi: 10.1074/jbc.M009327200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ravindran S, Radke GA, Guest JR, Roche TE. J Biol Chem. 1996;271:653–662. doi: 10.1074/jbc.271.2.653. [DOI] [PubMed] [Google Scholar]
- 35.Baker JC, Yan X, Peng T, Kasten S, Roche TE. J Biol Chem. 2000;275:15773–15781. doi: 10.1074/jbc.M909488199. [DOI] [PubMed] [Google Scholar]
- 36.Breman H, Westbrook J, Feng Z, Gilliland G, Bhat T, Weissig H, Shindyalov I, Bourne P. Nucleic Acids Res. 2000;28:235–242. doi: 10.1093/nar/28.1.235. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Liu S, Baker JC, Roche TE. J Biol Chem. 1995;270:793–800. doi: 10.1074/jbc.270.2.793. [DOI] [PubMed] [Google Scholar]
- 38.Guex N, Peitsch M. Electrophoresis. 1997;18:2714–2723. doi: 10.1002/elps.1150181505. [DOI] [PubMed] [Google Scholar]