Background: It is unknown how enzymes actually achieve the catalytic pathways proposed for ATP hydrolysis.
Results: A quantum mechanical analysis of myosin ATPase quantifies the relevant interactions that stabilize a metaphosphate intermediate.
Conclusion: The protein is designed to stabilize this metaphosphate, key to the enzymatic mechanism.
Significance: This yields a chemically consistent model for the catalytic strategy of nucleotide-hydrolyzing enzymes in general.
Keywords: ATP, ATPases, Bioenergetics, Biophysics, Computational Biology, Computer Modeling, Enzyme Catalysis, Enzyme Mechanisms, Myosin, Quantum Chemistry
Abstract
It has been proposed recently that ATP hydrolysis in ATPase enzymes proceeds via an initial intermediate in which the dissociated γ-phosphate of ATP is bound in the protein as a metaphosphate (PγO3−). A combined quantum/classical analysis of this dissociated nucleotide state inside myosin provides a quantitative understanding of how the enzyme stabilizes this unusual metaphosphate. Indeed, in vacuum, the energy of the ADP3−·PγO3−·Mg2+ complex is much higher than that of the undissociated ATP4−. The protein brings it to a surprisingly low value. Energy decomposition reveals how much each interaction in the protein stabilizes the metaphosphate state; backbone peptides of the P-loop contribute 50% of the stabilization energy, and the side chain of Lys-185+ contributes 25%. This can be explained by the fact that these groups make strong favorable interactions with the α- and β-phosphates, thus favoring the charge distribution of the metaphosphate state over that of the ATP state. Further stabilization (16%) is achieved by a hydrogen bond between the backbone C=O of Ser-237 (on loop Switch-1) and a water molecule perfectly positioned to attack the PγO3− in the subsequent hydrolysis step. The planar and singly negative PγO3− is a much better target for the subsequent nucleophilic attack by a negatively charged OH− than the tetrahedral and doubly negative PγO42− group of ATP. Therefore, we argue that the present mechanism of metaphosphate stabilization is common to the large family of nucleotide-hydrolyzing enzymes. Methodologically, this work presents a computational approach that allows us to obtain a truly quantitative conception of enzymatic strategy.
Introduction
Adenosine triphosphate (ATP) is the universal energy currency of biology (1, 2) in the form of its hydrolytic decomposition into adenosine diphosphate (ADP) and inorganic phosphate (see Fig. 1). ATP is very stable in solution, as indicated by its low rate of spontaneous hydrolysis (with rate constants of 3.2 × 10−5 s−1 and 17.5 × 10−5 s−1 at pH 8.4 and pH 4–5, respectively) (3, 4). These rates correspond to an activation free energy barrier in the 23–24 kcal mol−1 range (5). The energy barrier for the associative and dissociative methyl triphosphate hydrolysis in vacuum was calculated to be 40.0 and 42.0 kcal mol−1, respectively.2 Nevertheless, ATP hydrolysis is the most frequent chemical reaction occurring in the human body, (7) because it is catalyzed by ATPases (8, 9) to provide the energy required for most cellular processes (10, 11, 12). ATPases are enzymes that lower the energy barrier of ATP hydrolysis to as low as 10–15 kcal mol−1, thus allowing for a fast turnover rate (on the millisecond time scale). Much effort has gone into studying the catalytic mechanism of ATPases, (3, 13–18), but the catalytic strategy of these enzymes remains largely controversial. A Car-Parrinello molecular dynamics simulation (19) has shown that, in water, the dissociative mechanism (P–O bond cleavage precedes the attack of the hydrolyzed water; Fig. 1B) is slightly preferred over the associative mechanism (nucleophilic attack and breaking of the Pγ–Oβγ bond are concerted; Fig. 1A). Recent quantum mechanical investigations of F1-ATPase (15) and myosin-ATPase (16, 17) have proposed a dissociative mechanism (Fig. 1B), in which the initial event in the enzyme is the cleavage of the bond between the Pγ and the oxygen Oβγ of the triphosphate moiety prior to attack of the Pγ phosphorus by a water molecule. This results in the formation of a planar PO3− metaphosphate, which is bound in the protein between the ADP3−, the Mg2+ ion, and the attacking water. In both studies, the energy of the protein in this metaphosphate state (henceforth called Pm)3 is found to be quite low, only 8.1 kcal mol−1 (myosin) (16, 17) to 14.5 kcal mol−1 (F1-ATPase) (15) above the ATP reactant state. This is surprising, given that the hydrated ADP3−·PO3−·Mg2+ complex is very unstable in the absence of the proteins. Neither study explains how the enzymes manage to bring the energy of the Pm state to such low levels. Moreover, the two studies disagree about whether the Pm state is a local energy minimum (as found for F1-ATPase) (15) or rather an unstable transition state structure (as claimed for myosin) (16, 17). Notwithstanding, we propose that the formation of the Pm state is central to the whole catalytic strategy of ATPases and P–O bond-hydrolyzing enzymes in general. Indeed, the PO3− is planar and has only a single negative charge, which makes it a much better target for attack by an OH− hydroxyl group than the tetrahedral and doubly charged PγO32− of ATP (Fig. 1, compare A and B). Therefore, the stabilization of the Pm state emerges as a key element of the enzymatic mechanism. The questions addressed here are the following. 1) What is the energy of the Pm state, and is it a stable or rather a metastable intermediate? 2) How can the protein achieve this stabilization? 3) What are the protein residues responsible for Pm stabilization, and how much do they respectively contribute? We answer these questions by analyzing the Pm state in the myosin ATPase, using combined quantum/classical (QM/MM) calculations with density functional theory at a highly accurate level.
FIGURE 1.

Associative versus dissociative mechanisms. The formal charges on the phosphate groups are shown. A, in the “associative” mechanism, the nucleophilic attack of an OH− ion (hydrolyzed water) on the tetrahedral γ-phosphate of ATP4− is concerted with Pγ–Oβγ bond cleavage. B, in the dissociative mechanism, an intermediate with a planar PγO3− results from the Pγ–Oβγ bond cleavage occurring before the attack by the hydrolyzed water (arrow).
Myosin II is a molecular motor responsible for muscle contraction (20, 21). It uses the energy derived from ATP hydrolysis to drive the cyclic interactions between myosin and the actin filament that produce motion (Lymn-Taylor cycle) (20). During this cycle, myosin undergoes several conformational changes, which involve the successive closing of two loops (called Switch-1 and Switch-2; see Fig. 2A) around the ATP binding site, leading to a catalytically active ATPase (22–24). In this conformation, the rebinding of myosin to the actin filament and the concurrent power stroke can only occur after ATP has been hydrolyzed. We refer hereafter to this pre-power stroke conformation with ATP bound as the “reactant” state of the hydrolysis reaction. In this state, and looking at myosin-2 of Dictyostelium discoideum, the triphosphate moiety of ATP is tightly held by 15 hydrogen bonds with the surrounding protein (Fig. 2A). The attacking (Wa)3 and a helping (Wh) water are positioned by a dense hydrogen bonding network between the ATP and loops Switch-1 and Switch-2. Starting from this reactant state, the experimental rate of catalyzed hydrolysis is kcat = 102 s−1, which is 106 times faster than in solution and corresponds to a free energy barrier of 14.4 kcal mol−1 (25).
FIGURE 2.

Catalytic site of myosin. A, protein interactions with ATP in the reactant state. Only the triphosphate moiety of ATP is shown. The attacking (Wa) and helping (Wh) water molecules are labeled. Hydrogen bonds are indicated by thin dotted lines. Coordination bonds of the Mg2+ are shown as thick dotted lines (two more coordinating water molecules are not shown for clarity; see B). Switch-1, Switch-2, and P-loop are colored in magenta, green, and orange, respectively. Residue numbering is for myosin-2 of Dictyostelium discoideum. B, formation of the Pm state. Shown is overlap of the triphosphate moiety of ATP in the reactant state (ATP in gray) with the Pm state (ADP·PγO3− in color). The protein and the adenosine are not shown. All atoms of what is called here the “substrate” (except for the Mg2+-coordinating alcohol groups of Ser-237 and Thr-186; see A) are shown: the Mg2+ ion, the Wa and Wh water molecules, and two Mg2+-coordinating waters, W1 and W2.
We find that the structure of the Pm state is very similar to the crystallographic reactant state, the only major difference involving a 1.1-Å displacement of the γ-phosphorus atom (see Fig. 2B) as the substrate dissociates into ADP3− and PγO3−. The PγO3− is close to the Wa attacking water (Pγ·OWa distance is 1.95 Å), as had been observed in the Pm state of F1-ATPase (15), thus forming a hydrate complex. Our calculations show that in the absence of the protein, the energy of the substrate is 44.3 kcal mol−1 higher in the Pm than in the reactant state. In the presence of myosin, this energy difference is reduced to 2.3 kcal mol−1. To understand this strong stabilizing effect from the protein, the electrostatic interactions of the protein with the substrate in the reactant state were compared with those in the Pm state. Moreover, those interactions were decomposed into contributions from each protein moiety surrounding the substrate. With this approach, it was possible to pinpoint which groups are responsible for the stabilization and to precisely quantify their contribution. We identified eight peptide groups and three side chains that contribute most of the stabilizing interactions (Table 1). For example, the single most prominent peptide interaction is the H-bond between the backbone C=O of Ser-237 (on loop Switch-1) and the attacking water, which lowers the energy of the Pm state (relative to the reactant state) by 12 kcal mol−1, thus accounting by itself for 16% of the overall stabilization effect. Interactions with the backbone of the P-loop (residues 182–187) contribute ∼50% of the stabilization by forming hydrogen bonds with the oxygens of the α- and β-phosphates. Among the side chains, Lys-185 induces a major stabilizing effect (25% of the overall effect) by improving its interactions with the β-phosphate in the Pm state. More stabilizing interactions are listed in Table 1. Together, these form a pattern from which the catalytic strategy of the protein clearly emerges. Most of these interactions exploit the shift of one negative charge from the γ-phosphate in the reactant substrate (See Fig. 1A) to the α- and β-phosphates in the Pm substrate (Fig. 1B). Meanwhile, the Ser-237 carbonyl polarizes the Wa, thus promoting a stabilization of the PγO3−·H2OWa hydrate complex. The resulting metaphosphate species is then a much improved target for the attack by the nucleophilic OH− group resulting from the subsequent hydrolysis of water Wa. Exactly equivalent interactions are made in the structures of other nucleotide-hydrolyzing enzyme, such as Ras-RasGAP or the F1-ATPase (see “Discussion”), indicating that the metaphosphate-stabilizing strategy demonstrated here for myosin is used by other phosphate-hydrolyzing enzymes.
TABLE 1.
Main H-bonds between protein and substrate that stabilize the Pm state
| Protein residue | Protein moiety | Substrate moietya | ΔΔEb | Percentage of totalc |
|---|---|---|---|---|
| Gly-182 | Backbone NH | Oβγ | −8.2 | 11 (17)d |
| Ala-183, Gly-184, Lys-185, Thr-186, Glu-187 | Backbone NH | O1α, O1β, Oβγ, Oαβ | −26.6 | 35 |
| Ser-237 | C=O | Hydrogen of Wa | −12.0 | 16 |
| Gly-457 | C=O | Hydrogen of Wh | −3.9 | 5 |
| Lys-185 | -NH3+ | O1β and O1γ | −18.9 | 25 |
| Asn-233 | Side chain NH2 | O2α, O2β, O2γ | −2.9 | 4 |
| Ser-236 | -OH | Wa·PγO3− hydrate | −3.4 | 4 |
| Arg-238 | Guanidinium+ | NAe | +24.1 | NAe |
| Glu-459 | -COO− | Hydrogen of Wh | −13.9 | NAe |
a See Fig. 2B for the atomic nomenclature of the substrate.
b Electrostatic stabilization energy of interaction (in kcal mol−1) contributed by the protein moiety, same as plotted in Figs. 4B and 6B.
c ΔΔE of a given residue as a percentage of the sum of all favorable ΔΔE values in the table (excluding the 238/459 salt bridge).
d In parenthesis is the percentage contribution from the Gly-182 backbone NH when its quantum effects on stabilization (−5.08 kcal mol−1; see Table 2) are added to the ΔΔE.
e NA, not applicable.
MATERIALS AND METHODS
Computational Setup
The protein structure and classical energy function used here are mostly identical to those described in our previous QM/MM study of ATPase in myosin (5, 25). The structure of the protein was taken from a crystal of myosin complexed to the ATP analog ADP·Be·F3. The positions of the Wa and Wh molecules were taken from the crystal structure for the complex with ADP·VO4 (Protein Data Bank entry 1VOM) (26) as well as waters 1581 and 1178 of 1VOM. The side chain orientations of Glu-187, Asn-219, and Lys-241 were also taken from the 1VOM structure. The protein was divided into four concentric regions: 1) a central region composed mostly of the substrate, in which the atoms are QM-treated (all other atoms are treated classically); 2) a surrounding region consisting of all atoms within 14 Å of the γ-phosphate (these atoms are free to move); 3) a region of 8 Å around the previous region, in which a harmonic force (constant of 0.05 kcal mol−1 Å−2) is applied to each atom to keep it near its position in the crystal; and 4) a region consisting of all remaining atoms, which remain fixed in their positions in the crystal. The solvent screening of electrostatic interactions with the classical atoms was accounted for by using non-uniform charge scaling, an implicit solvent method described previously (27). In non-uniform charge scaling, the charges of the classically treated atoms are rescaled in such a way that their mutual electrostatic interactions reproduce the corresponding interactions obtained from a solution of the Poisson-Boltzmann equation.
The main methodological difference relative to our previous study was the choice of the quantum method. Here we use the hybrid density functional B3LYP (28, 29) with the 6–311+G** basis set. The classical atoms were subjected to the CHARMM force field, using parameters described previously (30, 31). QM/MM boundaries were treated using the link atom approach, as described previously (32, 33). Another difference is the choice of the QM region, which was extended. The QM region consisted of 84 atoms. These include the triphosphate moiety of ATP and the functional side chain groups of Ser-181, Thr-186, Ser-236, Ser-237, Arg-238, and Glu-459 residues. For residues Thr-186, Ser-236, and Ser-237, the link atom was placed between the Cα and Cβ atoms. In the case of Glu-459, the link atom was placed between Cβ and Cγ. For Arg-238, the link atom was placed between Cγ and Cδ. The peptide group [COCα]Ser-181[NH]Gly-182 was also included in the QM region. When a peptide is taken into the QM region, the point charges on the two neighboring peptide groups were modified as shown in Fig. 3. This avoids having a non-zero partial atomic charge on the two classical atoms making a bond to the QM peptide (i.e. Cα(i − 1) and C(i + 1) in Fig. 3) while preserving the dipole moments of these two peptide groups.
FIGURE 3.

Partial atomic charges on the backbone peptide groups. When a given peptide group i is treated quantum mechanically, the preceding (i − 1) and the following (i + 1) classical peptides are given non-standard partial atomic charges (shown in red). This is done to achieve a zero charge on the two classical atoms (i.e. Cα(i − 1) and C(i + 1)) that are directly bonded to the quantum moiety (peptide i). For reference, the standard classical atomic charges are shown in blue on peptide i (6).
Polarization Factors
Peptide groups are known to get easily polarized (i.e. their dipole moment significantly increases when they are in contact with a moiety carrying a net charge) (34–37). Thus, some of the peptide groups making hydrogen bonds with the triphosphate moiety (which carries a net charge of −4 ē) are expected to get highly polarized. In some cases, this can affect the energy difference ΔE between the reactant and the Pm states. For example, treating peptide 182/183 quantum mechanically instead of classically further stabilizes the Pm state by ΔΔE(182/183) = −1.24 kcal mol−1. The second column in Table 2 shows this energy change, ΔΔE(i), for each peptide group contacting the triphosphate. Ideally, all of the peptide groups around the triphosphate would be treated quantum mechanically, because this allows for the necessary polarization of their electron density. However, this would add so many atoms into the QM region that it is not computationally feasible. Therefore, the polarization was accounted for here by individually scaling up the classical atomic charges of each peptide. The charges of peptide group i are multiplied by a polarization factor PFi, (PFi > 1). The value of PFi is chosen so that the resulting change in the stabilization energy, ΔΔE(i), matches the effect of moving this peptide into the QM region. When the atoms of a peptide group i are added to the QM region, the atomic charges of the adjacent peptide groups (i − 1 and i + 1) were modified as shown in Fig. 3. Note that the polarization of peptide 181/182 is so strong that its stabilizing effect (ΔΔE(181/182) = −5.08 kcal mol−1; Table 2) could not be achieved by a mere upscaling of its atomic partial charges. This is the reason why this peptide was included in the QM region.
TABLE 2.
Polarization factors of peptide groups in direct contact with the triphosphate
| Peptide i/i + 1a | ΔΔE(i)b | ΔΔEsc(i)c | Self-consistent polarization factor, PFsc(i) |
|---|---|---|---|
| kcal mol−1 | kcal mol−1 | ||
| 181/182 | −5.08 | —d | — |
| 182/183 | −1.24 | −0.83 | 1.44 |
| 183/184 | −1.43 | −0.07 | 1.00 |
| 184/185 | −1.54 | −1.18 | 1.36 |
| 185/186 | −0.69 | −0.18 | 1.00 |
| 186/187 | +0.05 | — | 1.00 |
| 236/237 | +2.39 | +3.41 | 1.7 |
| 237/238 | −3.39 | −4.18 | 1.7 |
| 455/456 | −0.20 | −0.3 | 1.0 |
| 456/457 | +0.58 | +0.1 | 1.0 |
| 457/458 | +0.01 | — | 1.0 |
| Asn-233 side chain | −1.35 | −1.49 | 1.42 |
a Backbone peptide i/i + 1 consists of atoms C and O of residue i and atoms N, H, and Cα of residue i + 1.
b Effect on the stabilization energy of the Pm state from allowing single peptide polarization. ΔΔE(i) = ΔEQM(i) − ΔEMM(i), where ΔEQM(i) is the energy difference E(Pm) − E(reactant) when peptide i/i + 1 is treated quantum mechanically, and ΔEMM(i) is the corresponding energy difference when peptide i/i + 1 is treated classically, using standard (i.e. non-upscaled charges) for all of the other peptides. ΔΔE(i) < 0 indicates a stabilization effect, and ΔΔE(i) > 0 indicates a destabilization of Pm.
c Same as ΔΔE(i) in column 2, but accounting for the simultaneous polarization of the other peptide groups (see “Materials and Methods”). ΔΔEsc(i) is obtained as described in Footnote b, but using charges for all other peptides that are upscaled by their respective self-consistent polarization factor PFsc(i) (given in the rightmost column).
d Unless otherwise mentioned, peptide 181/182 was included in the QM region for all other calculations.
Because the polarization of the different peptide groups affects each other (either directly or via pulling electrons across the triphosphate moiety), all of the PF(i) values were modified iteratively until a self-consistent set of PF(i) values was obtained. The resulting set of the self-consistent polarization factors, PFsc(i), is listed in column 4 of Table 2. When |ΔΔEsc(i)| is less than 0.5 kcal mol−1, then the standard charges of peptide i were left unchanged (i.e. PFsc(i) = 1 (e.g. for peptide 183/184). These PFsc(i) values were used to multiply the atomic point charges of their respective peptides during the geometry optimization of the reactant and the Pm structures. The energy was minimized to a gradient of <0.01 kcal mol−1 Å−1, using the Turbomole version 5.9 program package (38) interfaced to CHARmm version c28b (39).
Electrostatic Interactions with the Substrate
In order to quantify the amount of stabilization of the Pm state due to electrostatic interaction of individual protein moieties (Figs. 4 and 6), the QM region was reduced so as to contain only the substrate (defined under “Results” and shown in Fig. 2B), totaling 45 QM atoms. In this case, peptide group 181/182 was treated classically, and its atomic charges were upscaled by a factor of 1.7 to partially account for its polarization. The geometries of the reactant and the Pm structures were then energy-optimized again at the B3LYP/6–31G* level. The protein residues around the QM region were subdivided into backbone peptide and side chains moieties. The stabilization effect ΔΔE(i) of a given moiety i is then obtained as the change in ΔE (equal to E(Pm) − E(reactant)) when the atomic charges of this moiety are all set to zero (as described in more detail under “Results”).
FIGURE 4.
Electrostatic stabilization of the Pm state by backbone peptide groups. The amount of stabilization by interactions with a given peptide group i/i + 1 (linking residues i and i + 1) is computed as described under “Results.” A negative value of ΔΔE(i) (in kcal mol−1) means that the peptide group i/i + 1 stabilizes the Pm state more than the reactant state. Only backbone peptides with significant contributions are shown. The substrate (as described in the legend to Fig. 2B) is treated quantum mechanically, and the rest of the system is treated classically. A, stabilization from the interaction of peptide i/i + 1 with the whole system (i.e. substrate and protein), ΔΔEall(pepi). B, stabilization from the interaction of peptide i/i + 1 with only the substrate, ΔΔEsubstr(pepi). Labels on the bars indicate when the peptide group is making an H-bond to one of the phosphate oxygens (α, β, γ, αβ, or βγ) or waters of the substrate.
FIGURE 6.
Electrostatic stabilization of the Pm state by side chains (in kcal mol−1). This figure is the same as Fig. 4 but for the interactions with a given side chains of residue i (rather than peptide). The contributions from the Arg-238 and Glu-459 side chains, which form a salt bridge (see Fig. 2A), have been combined into a single (rightmost) bar. A, stabilization from interactions of side chain i with the whole system, ΔΔEall(sidei). B, stabilization from interactions of side chain i with only the substrate, ΔΔEsubstr(sidei).
RESULTS
The Reactant and Pm Structures
In the energy-optimized reactant state (see “Materials and Methods”), the Pγ–Oβγ bond length is 1.78 Å (Fig. 2B), which is significantly longer than the distance typical for phosphoanhydride P–O–P bonds (∼1.6 Å). This indicates that a certain degree of ground state destabilization is happening in the reactant state, preparing for the dissociation of the γ-phosphate group. Using the 6–311+G** basis set, we find that the Pm structure is only 2.27 kcal mol−1 higher than the reactant structure. Minimum energy path calculations of the catalytic pathway with the conjugate peak refinement method (40) show that this Pm structure is in a local energy minimum and is separated by clear energy barriers from the states that precede and follow along the reaction pathway. These barriers are 6.0 and 7.7 kcal mol−1 for the back-step toward the reactant ATP state and for the forward water attack step, respectively.4 The Pγ–Oβγ bond cleavage results in an increase of the distance between Pγ and Oβγ to 2.94 Å. The attacking water oxygen (OWa) is then 1.95 Å from the Pγ (Fig. 2B), much closer than the 3 Å in the reactant (Fig. 2A) but not yet at a bonded distance. Thus, the PγO3− forms a “hydrated complex” with water Wa. The PγO3− metaphosphate species has a single negative net charge (Fig. 1B), which fundamentally distinguishes this group from the doubly charged -PγO42− phosphate moiety in the ATP state (Fig. 1A). This difference in charge distribution between reactant and Pm states (i.e. the γ-phosphate group going from a formal charge of −2 in ATP to −1 and the β-phosphate group going from a formal charge of −1 to −2) is used by the protein to stabilize the Pm state (as described under “Interactions Stabilizing the Pm State”). Note that although the three oxygens bonded to Pγ are nearly coplanar with Pγ and the “apical” oxygens Oβγ and OWa are co-linear with the Pγ (angle = 178.3°), this is not a bipyramidal structure as would be formed in an associative transition state, because Oβγ and OWa have different distances from the Pγ (2.94 and 1.95 Å, as mentioned above) (25). The three oxygens bonded to Pγ remain nearly stationary during the dissociation, and the difference between the ATP and Pm substrates is mostly due to a small motion (1.1 Å) of the γ-phosphorus atom when it becomes coplanar with its three oxygens (see Fig. 2B). The Mg2+ ion is hardly moving and remains hexacoordinated. There are no significant motions in the surrounding protein.
Interactions Stabilizing the Pm State
The “substrate” is defined hereafter as the atoms of the methyl-triphosphate moiety ((PO3)3OCH24−), the Wa and Wh water, and the Mg2+ and its four coordinating groups (waters W1 and W2, alcohol groups of Thr-186 and Ser-237), as shown in Fig. 2B. For all calculations of interaction energies, the substrate was treated quantum mechanically at the RB3LYP/631G* level. To estimate how much the protein stabilizes the Pm state, the energy difference between the Pm and the reactant structures was computed for the substrate alone (i.e. by removing all other atoms of the system). This energy difference is 44.3 kcal mol−1. In other words, the presence of the protein reduces the energy difference between the reactant and the Pm structures by 44.3 − 2.9 = 41.4 kcal mol−1. This is an astonishingly large stabilization effect, considering that the coordinates of the reactant and the Pm states are so similar (as mentioned above).
In order to understand how this stabilization is achieved, the stabilizing contribution from each protein moiety (breaking the protein into backbone peptides and side chains) was analyzed. The electrostatic interaction energy of a moiety in the Pm state is compared with its interaction energy in the reactant state. Because the substrate is treated quantum mechanically, the electrostatic effect due to the atomic partial charges of a given protein moiety (numbered i) is obtained here by setting these partial charges to zero and comparing the resulting total electrostatic energy to the corresponding energy obtained with standard (i.e. non-zeroed) partial charges. This yields Equation 1,
![]() |
where the “all” subscript indicates that all electrostatic interactions are included (i.e. interactions of moiety i with both the substrate and the rest of the system). Doing this for both the reactant and the Pm state structures gives ΔEall(i)Reactant and ΔEall(i)Pm, respectively. The stabilization effect due to group i is then obtained as follows.
This was done for each backbone peptide group within 14 Å of the triphosphate. A given peptide moiety i is composed of the atoms [C,O] of residue i and [N,H,Cα] of residue i + 1. Such a peptide group has a total net charge of zero (the partial charges of a peptide group are shown in Fig. 3. The resulting stabilization values, ΔΔEall(pepi), are plotted in Fig. 4A (only peptide groups with a significant contribution are shown). This clearly indicates which backbone peptide groups dominantly contribute to stabilizing the Pm state: peptide 237/238 on the Switch-1 loop (−13.2 kcal mol−1, a negative sign meaning that the interaction lowers the energy of the Pm state), peptide 457/458 on the Switch-2 loop (−3.9 kcal mol−1), and peptides 181–187 on the P-loop (−5.9 to −2.3 kcal mol−1). Interestingly, these are all peptides that form H-bonds with either the Wa, the Wh, or the α- and β-phosphate groups (see Fig. 2B). Because these peptides all directly interact with the substrate, we examined how much of the stabilization in Fig. 4A is due to interactions between each peptide group and the substrate itself (i.e. excluding the electrostatic interactions of the peptide with the rest of the system). This is described below.
Backbone Peptide Groups That Favor the Pm Substrate
The stabilization of the Pm state from interactions between a peptide group i and the substrate, ΔΔEsubst(pepi), is calculated as described above for ΔΔEall(pepi), except that Esubstr is used instead of Eall. Esubstr is the quantum mechanical energy of the substrate feeling the electrostatic field of the atomic partial charges of the classically treated protein. The resulting ΔΔEsubstr(pepi) values are plotted in Fig. 4B. Overall, Fig. 4B much resembles Fig. 4A, which confirms that the stabilization effect from the peptide groups is mainly due to their direct interaction with the substrate.
Prominent among the strongly stabilizing backbone peptides in Fig. 4B are six NH groups contiguously located on the P-loop, starting from the NH of Gly-182 to the NH of Gly-187. All six act as donors in H-bonds with perfect geometry to the oxygens of the α- and β-phosphates, both in the reactant (Fig. 2A) and in the Pm (Fig. 8) states. Their stabilizing effect on the Pm state is explained by the transfer of one negative charge from the γ-phosphate in the reactant state (Fig. 1A) to the α-β-diphosphate in the Pm state (Fig. 1B). This charge increase on the α-β-diphosphate (from approximately −2 to −3) allows for stronger H-bonds with the above mentioned backbone peptides in the Pm state of the substrate. Note that Fig. 1B provides only a formal view of the charge distribution. In reality, the transferred charge does not sit only on the oxygen Oβγ but is delocalized over the whole α-β-diphosphate group (e.g. see Fig. 5). Peptide 181/182 contributes the most stabilization (−8.2 kcal mol−1) (see Fig. 4B and Table 1) because the H-bond distance of Oβγ to the NH of Gly-182 shortens by 0.2 Å in the Pm state. The other five H-bonds between the P-loop and the α-β-diphosphate contribute stabilizations ranging from −2.4 kcal mol−1 to −6.5 kcal mol−1 (Fig. 4B), for a total of −26.6 kcal mol−1 (Table 1).
FIGURE 8.
Hydrogen bonding network in the Pm state. Hydrogen bonds are shown with dashed lines. The Switch-1, Switch-2, and P-loop residues are colored in magenta, green, and orange, respectively. The attacking (Wa) and helping (Wh) water molecules are labeled.
FIGURE 5.

Partial atomic charges of phosphate groups. A, ATP4−. B, PO3− + ADP3−. The partial charges of ATP and ADP are those of the CHARMM parameter set 27 (31). Note that this charge distribution is somewhat simplified. For example, the α-phosphate atoms have the same partial charges in ATP and ADP. A more accurate charge distribution for ADP would have a little more negative charge delocalized from the β- to the α-phosphate atoms.
Besides the P-loop, two other peptide groups make large contributions to stabilize the Pm substrate. Most prominent in Fig. 4B is peptide group 237/238 (on loop Switch-1), which contributes −12 kcal mol−1 stabilization (Table 1). Its Ser-237 C=O carbonyl accepts an H-bond from water Wa (Fig. 2A), thereby polarizing this water molecule (i.e. increasing the electron density on its oxygen), thus helping to make the oxygen of water Wa a stronger nucleophile. This reinforces the hydrate complex between Wa and the PγO3−, leading to the unusually close distance between the oxygen and phosphorus atoms (1.95 Å; see Fig. 2B). Finally, peptide 457/458 (on loop Switch-2) has its carbonyl C=O of Gly-457 accepting an H-bond from water Wh, which is itself H-bonded to Wa (Fig. 2A). The H-bond distance between the oxygen of Wh and the hydrogen of Wa goes from normal (1.75 Å) in the reactant state to very close (1.5 Å) in the Pm state (Fig. 2B). This indicates a cooperative polarization of waters Wa and Wh as water Wh relays the polarization effect from peptide group 457/458 to water Wa. In this way, peptide 457/458 contributes to further stabilizing (by −3.9 kcal mol−1; Table 1) the Wa·PγO3− hydrate complex.
The only backbone peptide group that has a clearly destabilizing effect on the Pm substrate is peptide 236/237 (+4 kcal mol−1; Fig. 4B). This is due to the fact that its NH of Ser-237 donates an H-bond to the γ-phosphate (Fig. 2A). This H-bond becomes weakened when the negative charge shifts away from the γ-phosphate (and toward the α-β-diphosphate) in the Pm state. Thus, a self-consistent picture of stabilization is emerging, in which the question of whether a protein group is either favoring or disfavoring the Pm state can be explained in terms of its interaction with either the α-β-diphosphate or the γ-phosphate groups (and the charge shift between the two).
Side Chains That Stabilize the Pm Substrate
The total electrostatic interaction of a given side chain i with the rest of the system, ΔΔEall(sidei), was computed in the same way as described above for the peptide moieties (ΔΔEall(pepi)). The result is shown in Fig. 6A for the residues making significant contributions. The corresponding side chain contributions from interactions with only the substrate, ΔΔEsubstr(sidei), were computed as described for the peptide (ΔΔEsubstr(pepi)) and are plotted in Fig. 6B. Comparing Figs. 6A and 4B shows that the dominant stabilizations are again mostly due to direct interactions with the substrate, as observed above for the peptide groups.
By far the strongest stabilization comes from the side chain of Lys-185+ (Fig. 6B), ΔΔEsubstr(Lys-185) = −18.9 kcal mol−1 (Table 1). This is somewhat surprising at first, because its positively charged -NH3+ group makes H-bonds with both the γ- and β-phosphates (oxygens Oγ1 and Oβ1; see Fig. 1 for the atomic nomenclature) with the same perfect geometry (2.6 and 2.7 Å, respectively) in both the reactant (Fig. 2A) and the Pm states. Therefore, a charge transfer from the γ- to β-phosphates would seem to have little effect on the interaction with this side chain. However, in the reactant, the distance of its ammonium nitrogen atom to the other atoms of the β-phosphate group is much closer (3.3–3.6 Å; see Table 3) than to the other atoms of the γ-phosphate (3.6–4.9 Å). In the Pm state, this difference in distance is even more pronounced (3.2–3.3 Å to the β-phosphate atoms, 3.8–4.8 Å to the γ-phosphate atoms). Therefore, and because the negative charge density on the β-phosphate is distributed over all of its atoms (e.g. see Fig. 5), the increase of this negative charge allows the positive -NH3+ group to interact much more favorably with the phosphates in the Pm than in the reactant substrate. This is shown in Fig. 7, which plots the contributions from the electrostatic interaction between the Lys-185+ side chain and the moieties of the substrate. For this purpose, classical partial charges were assigned to the atoms of the substrate. In this charge distribution (Fig. 5), a 0.6 ¯e charge is shifted from the γ- to the β-phosphate when going from the ATP (Fig. 5A) to the ADP·Pi (Fig. 5B) states. Fig. 7 clearly shows that the electrostatic interaction with the γ-phosphate is much less favorable in the Pm than in the reactant (ATP) substrate (+53.4 kcal mol−1), whereas the interaction with the β-phosphate stabilizes the Pm substrate by −67.1 kcal mol−1. In sum, this results in the net stabilizing effect of Lys-185+. Obviously, the interaction energies plotted in Fig. 7 are only indicative, because the classical charge distribution used in Fig. 5 is approximate and does not account for electronic polarization effects. Nevertheless, their relative amplitudes are meaningful. Indeed, the sum of the classical interactions in Fig. 7 is −15.6 kcal mol−1, which is comparable with the quantum interaction with Lys-185+ of −18.9 kcal mol−1 (mentioned above), showing that the charges in Fig. 5 capture the essence of the charge shift from the γ- to the β-phosphates.
TABLE 3.
Distance (in Å) of substrate atoms to the nitrogen atom of the Lys-185+ side chain
| Substrate moiety | Atoms | Reactant state | Pm state |
|---|---|---|---|
| γ-Phosphate | Oγ1 | 2.59 | 2.66 |
| Oγ2 | 4.23 | 4.01 | |
| Oγ3 | 4.93 | 4.81 | |
| Pγ | 3.63 | 3.75 | |
| β-phosphate | Oβ1 | 2.68 | 2.69 |
| Oβγ | 3.55 | 3.30 | |
| Oβ2 | 3.61 | 3.35 | |
| Pβ | 3.34 | 3.20 |
FIGURE 7.

Stabilization from interactions of the Lys-185+ side chain with substrate moieties. The ΔΔEsubstr(side = Lys-185) contribution from Fig. 6B is broken into subcontributions from electrostatic interactions between the Lys-185+ side chain and fragments of the substrate (using classical partial atomic charges for the substrate; see Fig. 5). Pa, PαO4CH2; Pb, PβO3; Pg, PγO3; Mg-Water, waters coordinating to Mg2+.
Another side chain that exploits the γ-to-β charge shift is Asn-233. It makes a rather poor H-bond (3.1 Å) with oxygen Oγ2 in the reactant (Fig. 2A), which breaks in the Pm state to make a perfect H-bond (2.8 Å) with oxygen Oβγ (Fig. 8). This contributes −2.9 kcal mol−1 of stabilization (Fig. 6B and Table 1) of the substrate in the Pm state. Ser-236 is the last side chain that makes significant stabilizing interactions with the substrate (Fig. 6B). In the reactant state, it donates an H-bond to water Wa (the OSer-236–OWa distance is 2.6 Å), which breaks in the Pm state to form an H-bond to oxygen Oγ1 of the γ-phosphate (Fig. 8). This switch from a neutral (water) to a charged (PO3−) H-bond acceptor contributes −3.4 kcal mol−1 stabilizing interaction with the substrate (Table 1).
The salt bridge between Arg-238+ (on Switch-1) and Glu-459− (on Switch-2) is known to be essential for the catalytic activity of myosin (41, 42). However, Fig. 6A shows that its role in stabilizing the Pm state is negligible. This is due in part to the fact that the electrostatic interactions of these two side chains with the substrate compete with each other; the γ- and β-phosphate moieties are proximal and distal, respectively, to the positively charged Arg-238+ side chain. The favorable interactions with the proximal γ-phosphate in the reactant state become less favorable (ΔΔEsubstr(Arg-238) = +24.1 kcal mol−1) as the negative charge shifts toward the distal β-phosphate. The opposite effect occurs for the negatively charged carboxyl group of the Glu-459− side chain; the unfavorable interaction of the -COO− group with the proximal and doubly negative -PγO42− becomes less unfavorable as the negative charge shifts further away onto the distal β-phosphate group, thus favoring the Pm state of the substrate (ΔΔEsubstr(Arg-238) = −13.5 kcal mol−1). In terms of the total electrostatic energy, the stabilizing contribution from Glu-459− (ΔΔEall(Glu-459) = −9.5 kcal mol−1) and the destabilizing contribution from Arg-238+ (ΔΔEall(Arg-238) = 10.6 kcal mol−1) practically cancel out. Nevertheless, the carboxyl -COO− of Glu-459 makes a strong H-bond to water Wh (Fig. 2A), thereby reinforcing the cooperative polarization of Wh and Wa described above and thus somewhat contributing to the stabilization of the Pm state.
DISCUSSION
We have presented here a comparison of the energetics of the reactant (ATP) and the Pm states of myosin using combined quantum/classical methods. In the Pm state, the Oβγ–Pγ bond of ATP is broken, and the resulting quasiplanar PO3− group forms a hydrate complex with the water that will attack the PγO3− in the subsequent step of water hydrolysis. The structure of the protein is very similar in the Pm and the reactant states. The energy difference between the two structures is between 2.3 and 2.7 kcal mol−1 (depending upon the basis set; Table 4). This is even more stable than the 8.1 kcal mol−1 that had been found before for the Pm state (16, 17). The lower value is probably due to the fact that the strong polarization of all peptide groups in contact with the triphosphate was included here explicitly (Table 2), which results in stronger interactions and therefore in more stabilization. For the same reason, the Pm state is separated here from other states along the hydrolysis reaction by higher barriers than had been found before, where the Pm state had been identified as transition state-like (16, 17).
TABLE 4.
Energy of the Pm state (relative to the reactant), in kcal mol−1
| Method | ΔEtotala | ΔΔEQMb |
|---|---|---|
| kcal mol−1 | kcal mol−1 | |
| RB3LYP/6-31G* | 2.7 | 47.4 (44.3)c |
| RB3LYP/6-311+G** | 2.3 | 51.6 |
a Difference in the total energy of the protein between the Pm and reactant states, ΔEtotal = Etotal(Pm) − Etotal(reactant), where the conformations of both Pm and reactant have been optimized; without zero point energy correction.
b Stabilization of the QM region caused by electrostatic interactions with the partial charges of the surrounding protein. ΔΔEQM = ΔEQM(classical charges = 0) − ΔEQM(normal protein charges), with ΔEQM = EQM(Pm) − EQM(reactant), where EQM is the quantum mechanical energy of the QM region (84 atoms; see “Materials and Methods”) in the same conformation as that used for ΔEtotal (column 2). Note that these values somewhat overestimate the stabilization of the substrate, because this QM region contains also four side chains (Ser-181, Ser-236, Arg-238, Glu-459) in addition to the substrate atoms.
c In parenthesis is the value of ΔΔEQM when only the substrate (shown in Fig. 2B) is in the QM region, and the conformations of Pm and reactant are energy-optimized for this reduced QM region.
When the protein is removed, the energy of the Pm substrate is 44.3 kcal mol−1 above that of ATP. This shows that an extraordinary stabilization of the substrate in the Pm state is achieved by interactions with the surrounding protein. The electrostatic interactions between the individual protein groups and the substrate were computed to find which protein groups contribute toward the Pm stabilization. This has revealed that the protein combines two complementary strategies. 1) The protein exploits the shift of negative charge from the γ-phosphate group to the α-β-diphosphate moiety, as illustrated in Figs. 1 and 5. It does so by placing positively charged groups (Lys-185+) and H-bond donors (peptide groups on the P-loop backbone and the Asn-233 side chain) that preferentially interact with the α- and β-phosphate groups, thereby favoring the charge shift. Approximately 75% of the overall stabilization is achieved in this way (Table 1). 2) The other strategy, accounting for the remaining 25% of stabilization, consists in reinforcing the hydrate complex of the PγO3− with water Wa. To do this, the protein highly polarizes water Wa, so that water oxygen OWa bears a higher electron density. This allows OWa to approach very closely to the electrophilic phosphorus atom of the PγO3− (1.95 Å length; Fig. 2B). Two backbone peptide groups achieve this polarization, either via a direct H-bond to water Wa (C=O of Ser-237) or indirectly via water Wh (itself polarized by an H-bond to C=O of Gly-457). Note that this polarization of water Wa has the added advantage of preparing the breakup of water Wa during the subsequent hydrolysis step (i.e. the generation of the attacking OH− hydroxyl).
Protein Groups Participating in Stabilizing Strategy 1
The group that provides the single largest stabilizing contribution (25%) is the Lys-185+ side chain (Table 1). As explained above, the -NH3+ group of Lys-185 is globally closer to the β- than the γ-phosphate group (Table 3), so that it interacts more favorably upon the shift of negative charge from γ to β. It is known that the K185Q point mutation inhibits ATP hydrolysis in myosin (43, 44). In the crystal structure of the human Ras-RasGAP GTPase (45), a lysine (Lys-16) has its -NH3+ group at exactly the same position relative to the phosphate groups, so that it can promote the dissociation of the metaphosphate in the same way as described here for myosin. In the structure of F1-ATPase, the -NH3+ of Lys-162 is located in an exactly equivalent position (46).
The next largest contribution comes from the backbone peptide 181/182 (located on the P-loop). This is not surprising, because the NH group of Gly-182 is making an H-bond with the oxygen Oβγ of ATP (i.e. the atom right next to the Pβ–Oβγ bond that breaks upon formation of the metaphosphate). This H-bond is maintained in the Pm state, but its length shortens from 2.9 to 2.7 Å (Oβγ to N182 distance), making this interaction stronger in the Pm state. Note that the stabilization effect of peptide 181/182 is stronger than indicated by the purely coulombic interaction energy of −8.2 kcal mol−1 (Table 1). This becomes apparent when this peptide is treated quantum mechanically instead of classically, which further lowers the energy of the Pm state by 5.08 kcal mol−1 (Table 2). This effect could not be reproduced by upscaling the classical charges of the peptide group (as was done for some other peptide groups to account for their polarization; see “Materials and Methods”). This means that this quantum effect is due not only to a polarization of the peptide group itself but also to some degree of charge redistribution on the substrate. Therefore, Gly-182 emerges as one of the residues crucial in stabilizing the Pm intermediate. Altogether, its contribution can be estimated to be around 17%. Exactly the same H-bond between a backbone peptide and Oβγ is found in the structure of other NTPases (e.g. in Ras-RasGAP with Gly-13 (45) and in the F1-ATPase with Gly-159 of the βDP chain (46)).
Five more peptide groups of the P-loop (from residue 183 to 187) together contribute 35% (Table 1) of stabilization. They all make H-bonds to the α- and β-phosphate groups, thus directly favoring the charge shift. The P-loop forms a cradle around the α-β-diphosphate moiety (Fig. 2A). In total, eight H-bonds exist between NH groups of the P-loop backbone and the oxygen atoms of the α- and β-phosphates (Fig. 8). The corresponding residues in the Ras-RasGAP structure are at positions 14–18 (45), and the corresponding residues in F1-ATPase are 160–164 (46). Their backbone peptides all make exactly the same H-bonds with the α-β-diphosphate as described here for residues 183–187 in myosin. Thus, the P-loop is clearly designed to pull electronic charge from the γ- to the α-β-groups, thereby promoting the dissociation of the metaphosphate. This explains why the P-loop is a characteristic structural feature in other NTPases, where it serves not only as a binding motif (47) but, more importantly, as a device to stabilize the Pm state.
Finally, the peptide side chain of Asn-233 is positioned so that it can make an optimal H-bond with Oβγ when this oxygen becomes accessible after the Pβ–Oβγ bond has broken (Fig. 8). This contributes 4% of the total stabilization (Table 1). Experiments have shown that point mutations of Asn-233 affect the ATPase activity (43). In the Ras-RasGAP GTPase, Arg-789 is making exactly the same interaction with the oxygen Oβγ of GTP (45). The corresponding interaction is found in F1-ATPase with Arg-373 (46). Note that several of the peptide groups mentioned above are highly polarized (see Table 2) by their interaction with the strongly charged triphosphate, thereby augmenting their stabilizing contribution. This effect was included here to account for the full amount of stabilization (see “Materials and Methods”).
Protein Groups Participating in Stabilization Strategy 2
Of the groups that reinforce the polarization of water Wa, peptide 237/238 contributes 16% of the overall stabilization. The C=O group of Ser-237 on Switch-2 forms an H-bond with the hydrogen H1 of water Wa (Fig. 2A). This H-bond polarizes one of the two O–H bonds of the water (OWa-H1; Fig. 2B). The corresponding backbone interaction with the attacking water in Ras-RasGAP is with the C=O of Thr-35 (45), and with the C=O of Ser-344 in the F1-ATPase (46).
The protein indirectly achieves the polarization of the other bond in water Wa (OWa–H2) by an H-bond with water Wh, itself polarized by an H-bond with the C=O of Gly-457 (Fig. 2A). This contributes 5% of the overall stabilization (Table 1). The -COO− group of Glu-459 reinforces this effect by also making an H-bond to water Wh (Fig. 2A). As a result of having both of its hydrogens engaged in H-bonds, water Wa is highly polarized and can form a stronger hydrate complex with the PγO3−. Finally, this complex is favored by the side chain of Ser-236, which is forming an H-bond with the PγO3− that is not formed in the reactant state, contributing a further 4% stabilization.
The Metaphosphate Is Central to the Catalytic Strategy
Pm state stabilization is essential to allow a dissociative mechanism, which in turn is the key to lowering the barrier of the ATP hydrolysis reaction. The uncatalyzed reaction is so slow in physiological conditions because the tetrahedral and doubly negative -PγO42− group is a very poor target for attack by an OH− hydroxyl. Moreover, the generation of this OH− by hydrolysis of a water molecule is also a slow process. An NTPase like myosin accelerates the reaction by using a dual strategy: 1) promoting the P–O bond dissociation that leads to formation of a PO3− and 2) deprotonation of the attacking water molecule by a nearby base to form the OH− group. The PγO3− is a much better target for attack by the OH− hydroxyl group as compared with the PγO42− of ATP, because its planar geometry facilitates access to the phosphorus atoms (compared with the tetrahedral PγO42−), and it has one less negative charge. To stabilize the Pm state, the protein exploits the difference in charge distribution between the ATP and Pm substrates, with approximately one negative charge shifting from the γ- to the β-phosphate. The protein groups involved in these stabilizing interactions are mainly the Lys-185+ side chain and the backbone NH groups of the P-loop (in particular Gly-182), the Ser-237 C=O moiety, and the Asn-233 side chain (both on the Switch-1 loop). Their relative contributions to the stabilization are listed in Table 1. Equivalent interactions are found in most other NTPases, indicating that the stabilizing mechanism and the catalytic strategy described here are used by NTPases in general.
Reversibility of ATP Hydrolysis in Myosin
Isotope exchange experiments have shown that the oxygen atoms of the terminal γ-PO3 of ATP and the attacking water molecule can exchange during myosin-catalyzed ATP hydrolysis (48–50). This implies that, in the protein, hydrolytic cleavage of ATP is reversible (51, 52). A rapid equilibrium is established between the unhydrolyzed myosin·ATP state and the hydrolyzed myosin·ADP·Pi state, resulting in a reversible cycle that consists of three steps (53, 54): 1) the forward hydrolytic step during which the oxygen of the attacking water is incorporated into the dissociated Pγ; 2) Pγ rotation, during which the phosphate oxygen atoms interchange their positions, and 3) reverse hydrolysis during which the reaction proceeds backwards, followed by water release. In each cycle, one solvent oxygen atom is incorporated into the γ-phosphate of ATP and one oxygen atom of Pγ is released as part of a water molecule (48–50). After several cycles, ATP undergoes almost complete oxygen exchange with the solvent. This oxygen exchange is also observed in the presence of actin (49) but at a slower rate (55, 56), because instead of favoring phosphate rotation, the acto-myosin complex induces product release. The rates and equilibrium constants of the forward and reverse ATP hydrolysis reaction in Dictyostelium myosin II at varying temperatures have been reported (57). The reaction gives a forward rate of 100 ± 10/s, and an equilibrium constant of 13–79 at 293 K, which correspond to a free energy barrier of 14.4 kcal/mol and a reaction free energy of −1.5 to −2.6 kcal/mol. The metaphosphate state presented here is fully compatible with this reversibility of ATP hydrolysis in myosin, both structurally and energetically. In terms of the structural pathway, the dissociative as well as the associative mechanisms allow for the formation of an ADP/Pi product in which the PγO4 product molecule can rotate in the binding pocket and interchange its oxygen positions before reassociating to form ATP and water. In other words, the isotope exchange experiments do not discriminate between dissociative (with a metaphosphate intermediate) and associative mechanisms in the protein. In terms of energetics, the energy of the Pm state (2.3 kcal/mol) is much lower than the 14.4 kcal/mol free energy barrier of the hydrolysis step, so that it does not introduce a rate-limiting step and consequently does not interfere with the reversibility of the reaction. In steady-state conditions, the Pm state is expected to be populated ∼100 times less than the ATP reactant state and 4000 times less than the APD/Pi product state. This should make it difficult to detect in kinetic experiments.
This work was supported by the German Science Foundation (DFG) through the module “SFB 623, Molecular Catalysis: Structure and Function Design.”
Using B3LYP density functional with a 6-31+G** basis set, we computed the hydrolysis barrier of methyl triphosphate4− complexed with Mg2+ in the presence of four Mg2+-coordinated water molecules and one attacking water molecule.
F. A. Kiani and S. Fischer, manuscript in preparation.
- Pm
- metaphosphate state of protein with PO3− and ADP bound
- Wa and Wh
- attacking and helping water molecule, respectively
- QM
- quantum mechanical
- MM
- molecular mechanical
- H-bond
- hydrogen bond.
REFERENCES
- 1. Rodgers A. J., Wilce M. C. (2000) Structure of the γ-ϵ complex of ATP synthase. Nat. Struct. Biol. 7, 1051–1054 [DOI] [PubMed] [Google Scholar]
- 2. Pebay-Peyroula E., Dahout-Gonzalez C., Kahn R., Trézéguet V., Lauquin G. J., Brandolin G. (2003) Structure of mitochondrial ADP/ATP carrier in complex with carboxyatractyloside. Nature 426, 39–44 [DOI] [PubMed] [Google Scholar]
- 3. Trentham D. R., Eccleston J. F., Bagshaw C. R. (1976) Kinetic analysis of ATPase mechanisms. Q. Rev. Biophys. 9, 217–281 [DOI] [PubMed] [Google Scholar]
- 4. Miller D. L., Westheimer F. H. (1966) The hydrolysis of γ-phenylpropyl di- and triphosphates. J. Am. Chem. Soc. 88, 1507–1511 [DOI] [PubMed] [Google Scholar]
- 5. Schwarzl S. M. (2006) Understanding the ATP Hydrolysis Mechanism in Myosin Using Computer Simulation Techniques, Ph.D. thesis, Ruprecht-Karls University of Heidelberg [Google Scholar]
- 6. Brooks B. R., Bruccoleri R. E., Olafson B. D., States D. J., Swaminathan S., Karplus M. (1983) CHARMM: A program for macromolecular energy, minimization, and dynamics calculations. J. Comput. Chem. 4, 187–217 [Google Scholar]
- 7. Boyer P. D. (2003) Nobel Lectures, Chemistry, 1996–2000, p. 120, World Scientific Publishing, Singapore [Google Scholar]
- 8. Kull F. J., Sablin E. P., Lau R., Fletterick R. J., Vale R. D. (1996) Crystal structure of the kinesin motor domain reveals a structural similarity to myosin. Nature 380, 550–555 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Vale R. D. (1996) Switches, latches, and amplifiers. Common themes of G proteins and molecular motors. J. Cell Biol. 135, 291–302 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Howard J., Hudspeth A. J., Vale R. D. (1989) Movement of microtubules by single kinesin molecules. Nature 342, 154–158 [DOI] [PubMed] [Google Scholar]
- 11. Block S. M., Goldstein L. S., Schnapp B. J. (1990) Bead movement by single kinesin molecules studied with optical tweezers. Nature 348, 348–352 [DOI] [PubMed] [Google Scholar]
- 12. Finer J. T., Simmons R. M., Spudich J. A. (1994) Single myosin molecule mechanics. Piconewton forces and nanometre steps. Nature 368, 113–119 [DOI] [PubMed] [Google Scholar]
- 13. Li G., Cui Q. (2004) Mechanochemical coupling in myosin. A theoretical analysis with molecular dynamics and combined QM/MM reaction path calculations. J. Phys. Chem. B 108, 3342–3357 [Google Scholar]
- 14. Yu H., Ma L., Yang Y., Cui Q. (2007) Mechanochemical coupling in the myosin motor domain. I. Insights from equilibrium active-site simulations. PLoS Comput. Biol. 3, e21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Hayashi S., Ueno H., Shaikh A. R., Umemura M., Kamiya M., Ito Y., Ikeguchi M., Komoriya Y., Iino R., Noji H. (2012) Molecular mechanism of ATP hydrolysis in F1-ATPase revealed by molecular simulations and single-molecule observations. J. Am. Chem. Soc. 134, 8447–8454 [DOI] [PubMed] [Google Scholar]
- 16. Grigorenko B. L., Rogov A. V., Topol I. A., Burt S. K., Martinez H. M., Nemukhin A. V. (2007) Mechanism of the myosin catalyzed hydrolysis of ATP as rationalized by molecular modeling. Proc. Natl. Acad. Sci. U.S.A. 104, 7057–7061 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Grigorenko B. L., Kaliman I. A., Nemukhin A. V. (2011) Minimum energy reaction profiles for ATP hydrolysis in myosin. J. Mol. Graph. Model. 31, 1–4 [DOI] [PubMed] [Google Scholar]
- 18. McGrath M. J., Kuo I.-F., Hayashi S., Takada S. (2013) Adenosine triphosphate hydrolysis mechanism in kinesin studied by combined quantum-mechanical/molecular-mechanical metadynamics simulations. J. Am. Chem. Soc. 135, 8908–8919 [DOI] [PubMed] [Google Scholar]
- 19. Akola J., Jones R. O. (2003) ATP hydrolysis in water: A density functional study. J. Phys. Chem. B 107, 11774–11783 [Google Scholar]
- 20. Lymn R. W., Taylor E. W. (1971) Mechanism of adenosine triphosphate hydrolysis by actomyosin. Biochemistry 10, 4617–4624 [DOI] [PubMed] [Google Scholar]
- 21. Rayment I., Holden H. M., Whittaker M., Yohn C. B., Lorenz M., Holmes K. C., Milligan R. A. (1993) Structure of the actin-myosin complex and its implications for muscle contraction. Science 261, 58–65 [DOI] [PubMed] [Google Scholar]
- 22. Kühner S., Fischer S. (2011) Structural mechanism of the ATP-induced dissociation of rigor myosin from actin. Proc. Natl. Acad. Sci. U.S.A. 108, 7793–7798 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Koppole S., Smith J. C., Fischer S. (2007) The structural coupling between ATPase activation and recovery stroke in the myosin II motor. Structure 15, 825–837 [DOI] [PubMed] [Google Scholar]
- 24. Yang Y., Yu H., Cui Q. (2008) Extensive conformational transitions are required to turn on ATP hydrolysis in myosin. J. Mol. Biol. 381, 1407–1420 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Schwarzl S. M., Smith J. C., Fischer S. (2006) Insights into the chemomechanical coupling of the myosin motor from simulation of its ATPase mechanism. Biochemistry 45, 5830–5847 [DOI] [PubMed] [Google Scholar]
- 26. Smith C. A., Rayment I. (1996) X-ray structure of the magnesium(II)·ADP·vanadate complex of the Dictyostelium discoideum myosin motor domain to 1.9 Å resolution. Biochemistry 35, 5404–5417 [DOI] [PubMed] [Google Scholar]
- 27. Schwarzl S. M., Huang D., Smith J. C., Fischer S. (2005) Nonuniform charge scaling (NUCS). A practical approximation of solvent electrostatic screening in proteins. J. Comput. Chem. 26, 1359–1371 [DOI] [PubMed] [Google Scholar]
- 28. Becke A. D. (1993) Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 98, 5648–5652 [Google Scholar]
- 29. Lee C., Yang W., Parr R. G. (1988) Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 37, 785–789 [DOI] [PubMed] [Google Scholar]
- 30. Neria E., Fischer S., Karplus M. (1996) Simulation of activation free energies in molecular systems. J. Chem. Phys. 105, 1902–1921 [Google Scholar]
- 31. MacKerell A. D., Jr., Bashford D., Bellott M., Dunbrack R. L., Jr., Evanseck J. D., Field M. J., Fischer S., Gao J., Guo H., Ha S., Joseph-McCarthy D., Kuchnir L., Kuczera K., Lau F. T. K., Mattos C., Michnick S., Ngo T., Nguyen D. T., Prodhom B., Reiher W. E., III, Roux B., Schlenkrich M., Smith J. C., Stote R., Straub J., Watanabe M., Wiorkiewicz-Kuczera J., Yin D., Karplus M. (1998) All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 102, 3586–3616 [DOI] [PubMed] [Google Scholar]
- 32. Field M. J., Bash P. A., Karplus M. (1990) A combined quantum-mechanical and molecular mechanical potential for molecular-dynamics simulations. J. Comput. Chem. 11, 700–733 [Google Scholar]
- 33. Gao J. (1996) Methods and applications of combined quantum mechanical and molecular mechanical potentials. in Reviews in Computational Chemistry, Vol. 7 (Lipkowitz K. B., Boyd D. B., eds) pp. 119–185, John Wiley & Sons, Inc., Hoboken, NJ [Google Scholar]
- 34. Roux B., Karplus M. (1991) Ion transport in a model gramicidin channel. Structure and thermodynamics. Biophys. J. 59, 961–981 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Van Belle D., Couplet I., Prevost M., Wodak S. J. (1987) Calculations of electrostatic properties in proteins. Analysis of contributions from induced protein dipoles. J. Mol. Biol. 198, 721–735 [DOI] [PubMed] [Google Scholar]
- 36. Lybrand T. P., Kollman P. A. (1985) Water-water and water-ion potential functions including terms for many body effects. J. Phys. Chem. 83, 2923–2933 [Google Scholar]
- 37. Guo H., Gresh N., Roques B. P., Salahub D. R. (2000) Many-body effects in systems of peptide hydrogen-bonded networks and their contributions to ligand binding. A comparison of the performances of DFT and polarizable molecular mechanics. J. Phys. Chem. B 104, 9746–9754 [Google Scholar]
- 38. Ahlrichs R., Bar M., Haser M., Horn H., Kolmel C. (1989) Electronic structure calculations on workstation computers: The program system turbomole. Chem. Phys. Lett. 162, 165–169 [Google Scholar]
- 39. Brooks B. R., Brooks C. L., 3rd, Mackerell A. D., Jr., Nilsson L., Petrella R. J., Roux B., Won Y., Archontis G., Bartels C., Boresch S., Caflisch A., Caves L., Cui Q., Dinner A. R., Feig M., Fischer S., Gao J., Hodoscek M., Im W., Kuczera K., Lazaridis T., Ma J., Ovchinnikov V., Paci E., Pastor R. W., Post C. B., Pu J. Z., Schaefer M., Tidor B., Venable R. M., Woodcock H. L., Wu X., Yang W., York D. M., Karplus M. (2009) CHARMM. The biomolecular simulation program. J. Comput. Chem. 30, 1545–1614 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Fischer S., Karplus M. (1992) Conjugate peak refinement: An algorithm for finding reaction paths and accurate transition states in systems with many degrees of freedom. Chem. Phys. Lett. 194, 252–261 [Google Scholar]
- 41. Furch M., Fujita-Becker S., Geeves M. A., Holmes K. C., Manstein D. J. (1999) Role of the salt-bridge between switch-1 and switch-2 of Dictyostelium myosin. J. Mol. Biol. 290, 797–809 [DOI] [PubMed] [Google Scholar]
- 42. Onishi H., Kojima S., Katoh K., Fujiwara K., Martinez H. M., Morales M. F. (1998) Functional transitions in myosin. Formation of a critical salt-bridge and transmission of effect to the sensitive tryptophan. Proc. Natl. Acad. Sci. U.S.A. 95, 6653–6658 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Li X.-D., Rhodes T. E., Ikebe R., Kambara T., White H. D., Ikebe M. (1998) Effects of mutations in the γ-phosphate binding site of myosin on its motor function. J. Biol. Chem. 273, 27404–27411 [DOI] [PubMed] [Google Scholar]
- 44. Shimada T., Sasaki N., Ohkura R., Sutoh K. (1997) Alanine scanning mutagenesis of the switch I region in the ATPase site of Dictyostelium discoideum myosin II. Biochemistry 36, 14037–14043 [DOI] [PubMed] [Google Scholar]
- 45. Scheffzek K., Ahmadian M. R., Kabsch W., Wiesmüller L., Lautwein A., Schmitz F., Wittinghofer A. (1997) The Ras-RasGAP complex. Structural basis for GTPase activation and its loss in oncogenic Ras mutants. Science 277, 333–338 [DOI] [PubMed] [Google Scholar]
- 46. Bowler M. W., Montgomery M. G., Leslie A. G., Walker J. E. (2007) Ground state structure of F1-ATPase from bovine heart mitochondria at 1.9 Å resolution. J. Biol. Chem. 282, 14238–14242 [DOI] [PubMed] [Google Scholar]
- 47. Rayment I., Rypniewski W. R., Schmidt-Bäse K., Smith R., Tomchick D. R., Benning M. M., Winkelmann D. A., Wesenberg G., Holden H. M. (1993) Three-dimensional structure of myosin subfragment-1. A molecular motor. Science 261, 50–58 [DOI] [PubMed] [Google Scholar]
- 48. Levy H. M., Koshland D. E., Jr. (1958) Evidence for an intermediate in the hydrolysis of ATP by muscle proteins. J. Am. Chem. Soc. 80, 3164–3165 [Google Scholar]
- 49. Levy H. M., Koshland D. E., Jr. (1959) Mechanism of hydrolysis of adenosine triphosphate by muscle proteins and its relation to muscular contraction. J. Biol. Chem. 234, 1102–1107 [PubMed] [Google Scholar]
- 50. Levy H. M., Sharon N., Lindemann E., Koshland D. E., Jr. (1960) Properties of the active site in myosin hydrolysis of adenosine triphosphate as indicated by the O18-exchange reaction. J. Biol. Chem. 235, 2628–2632 [PubMed] [Google Scholar]
- 51. Dempsey M. E., Boyer P. D., Benson E. S. (1963) Characteristics of an orthophosphate oxygen exchange catalyzed by myosin, actomyosin, and muscle fibers. J. Biol. Chem. 238, 2708–2715 [PubMed] [Google Scholar]
- 52. Swanson J. R., Yount R. (1966) The properties of heavy meromyosin and myosin catalyzed “medium” and “intermediate” 18O-phosphate exchange. Biochem. Z. 345, 395–409 [Google Scholar]
- 53. Bagshaw C. R., Trentham D. R., Wolcott R. G., Boyer P. D. (1975) Oxygen exchange in the γ-phosphoryl group of protein-bound ATP during Mg2+-dependent adenosine triphosphatase activity of myosin. Proc. Natl. Acad. Sci. U.S.A. 72, 2592–2596 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Webb M. R., Trentham D. R. (1981) The mechanism of ATP hydrolysis catalyzed by myosin and actomyosin, using rapid reaction techniques to study oxygen exchange. J. Biol. Chem. 256, 10910–10916 [PubMed] [Google Scholar]
- 55. Shukla K. K., Levy H. M. (1977) Mechanism of oxygen exchange in actin-activated hydrolysis of adenosine triphosphate by myosin subfragment 1. Biochemistry 16, 132–136 [DOI] [PubMed] [Google Scholar]
- 56. Sleep J. A., Boyer P. D. (1978) Effect of actin concentration on the intermediate oxygen exchange of myosin. Relation to the refractory state and the mechanism of exchange. Biochemistry 17, 5417–5422 [DOI] [PubMed] [Google Scholar]
- 57. Málnási-Csizmadia A., Pearson D. S., Kovács M., Woolley R. J., Geeves M. A., Bagshaw C. R. (2001) Kinetic resolution of a conformational transition and the ATP hydrolysis step using relaxation methods with a Dictyostelium myosin II mutant containing a single tryptophan residue. Biochemistry 40, 12727–12737 [DOI] [PubMed] [Google Scholar]




