+ All Categories
Home > Documents > Folding-based molecular simulations reveal mechanisms of ... · data with simulation results, we...

Folding-based molecular simulations reveal mechanisms of ... · data with simulation results, we...

Date post: 20-Oct-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
6
Folding-based molecular simulations reveal mechanisms of the rotary motor F 1 –ATPase Nobuyasu Koga* and Shoji Takada* †‡ *Graduate School of Science and Technology, Kobe University, Rokkodai, Nada, Kobe 657-8501, Japan; and Core Research for Evolutional Science and Technology, Japan Science and Technology Corporation, Rokkodai, Nada, Kobe 657-8501, Japan Edited by Peter G. Wolynes, University of California at San Diego, La Jolla, CA, and approved February 13, 2006 (received for review November 5, 2005) Biomolecular machines fulfill their function through large confor- mational changes that typically occur on the millisecond time scale or longer. Conventional atomistic simulations can only reach mi- croseconds at the moment. Here, extending the minimalist model developed for protein folding, we propose the ‘‘switching Go model’’ and use it to simulate the rotary motion of ATP-driven molecular motor F 1 –ATPase. The simulation recovers the unidirec- tional 120° rotation of the -subunit, the rotor. The rotation was induced solely by steric repulsion from the 3 3 subunits, the stator, which undergoes conformation changes during ATP hydro- lysis. In silico alanine mutagenesis further elucidated which resi- dues play specific roles in the rotation. Finally, regarding the mechanochemical coupling scheme, we found that the tri-site model does not lead to successful rotation but that the always- bi-site model produces 30° and 90° substeps, perfectly in accord with experiments. In the always-bi-site model, the number of sites occupied by nucleotides is always two during the hydrolysis cycle. This study opens up an avenue of simulating functional dynamics of huge biomolecules that occur on the millisecond time scales involving large-amplitude conformational change. always-bi-site model energy landscape funnel switching Go model B iomolecular machines, such as the ribosome, transporter, and molecular motors, fulfill their function through large- amplitude conformational change. Structural information be- fore and after the conformational change has been provided by x-ray crystallography and other methods for many cases (1). However, these methods do not directly observe the molecular dynamics that connects two-end structures. These dynamical aspects can be observed directly by fluorescence and other time-resolved spectroscopy; however, the latter methods moni- tor local structure but do not give global structural information. In this sense, molecular dynamics simulation is potentially powerful because it can provide full time-dependent structural information about biomolecular machines (2–5). Yet, functional cycles of these systems typically take milliseconds or longer, which is far beyond the current reach of molecular simulations with all-atom standard force-fields (6). Thus, a complementary approach may be to coarse-grain the molecular representation, thereby enabling the simulation orders of magnitude-longer time scales. Coarse-graining drops some details from the model, while hopefully keeping the essence. To achieve them, we need some perspective on the physical process we are trying to simply. Upon conformational change, some portions of molecules do not significantly change their structures; here, the atomic motion can be well approximated by harmonic fluctuations. It has been known that very simple C models, such as the elastic network model (7), can reproduce low-frequency harmonic fluctuation fairly well. Conversely, large-scale conformational change in- volves rearrangement of nonlocal contacts between amino acids, which is beyond the harmonic approximation (8, 9). For approx- imating these rearrangements, we need to take into account for partial disorder or partial unfolding, which induces conforma- tional diversity (8). To this end, we rely on the energy landscape perspective of proteins, which has been established in protein- folding studies. In this perspective, proteins evolved to have funnel-like energy landscapes, where the native structure is at the bottom of the funnel with the lowest effective energy and disordered states have higher energies and large conformational entropy (10, 11). Technically, a funnel-like energy landscape can easily be realized by Go models, which have become one of the standard models for folding simulation studies (12–18). For representing large-amplitude conformational dynamics, Go models are suitable because they account for both small fluc- tuations around the native basin and large fluctuations that involves local unfolding. Here, we propose a previously unde- scribed computational framework, which we call a ‘‘switching Go model,’’ for simulating large-amplitude motion of biomolecular machines. We then apply it to the rotary motion of F 1 –ATPase. F 1 –ATPase, the catalytic part of the energy conversion en- zyme F 0 F 1 –ATP synthase, is known to act as a rotary motor upon ATP hydrolysis (19–24). Single-molecule experiments have vi- sualized the rotation and elucidated some of the details; the -subunit rotates unidirectionally in discrete 120° steps with each ATP hydrolysis (25), and the 120° step can be divided into 90° and 30° substeps at lower ATP concentration (26). The dwell time before the 90° substep depends on ATP concentration, and the one before the 30° substep includes two rates of an 1-ms time scale. These findings suggest that the 90° substep involves ATP binding, and the 30° substep may correspond to ATP hydrolysis and release of products (ADP, phosphate, or both). These experiments have given much insight into the rotational mechanisms as described above; however, the resolution of those studies is limited to 40 nm, which cannot provide a complete structural view of the F 1 dynamics. The minimal catalytic complex of F 1 contains seven subunits 3 3 , whose x-ray structure (27) is depicted in Fig. 1a. The central stalk is a rotor, whereas the alternately arranged 3 and 3 subunits form a ring acting as the stator of the rotary motor. There are three catalytic nucleotide-binding sites, which are at the interfaces of -subunits with the neighboring -subunits. In the crystal structure determined in 1994 (27), one catalytic site is filled with ATP analog and another with ADP, and the last is empty. The structures of the ATP-bound subunits are denoted as TP and TP , where TP takes on the ‘‘closed form’’ in its C terminus (Fig. 1b Right Lower) (28) and the catalytic interface is loosely packed. The ADP-bound -subunit, DP , also takes the closed form, and the interface to DP is tightly packed. The structures of nucleotide-free subunits are denoted as E and E , where E has ‘‘open form’’ in its C terminus (Fig. 1b Left Lower) (28). From the static structure, we can deduce that the catalysis- induced bending motion of between closed and open forms produce torque on leading to rotation (27, 29). Given the static x-ray structures at high resolution and the dynamic single Conflict of interest statement: No conflicts declared. This paper was submitted directly (Track II) to the PNAS office. To whom correspondence should be sent at the * address. E-mail: [email protected]. © 2006 by The National Academy of Sciences of the USA www.pnas.orgcgidoi10.1073pnas.0509642103 PNAS April 4, 2006 vol. 103 no. 14 5367–5372 BIOPHYSICS Downloaded by guest on March 30, 2021
Transcript
  • Folding-based molecular simulations revealmechanisms of the rotary motor F1–ATPaseNobuyasu Koga* and Shoji Takada*†‡

    *Graduate School of Science and Technology, Kobe University, Rokkodai, Nada, Kobe 657-8501, Japan; and †Core Research for Evolutional Scienceand Technology, Japan Science and Technology Corporation, Rokkodai, Nada, Kobe 657-8501, Japan

    Edited by Peter G. Wolynes, University of California at San Diego, La Jolla, CA, and approved February 13, 2006 (received for review November 5, 2005)

    Biomolecular machines fulfill their function through large confor-mational changes that typically occur on the millisecond time scaleor longer. Conventional atomistic simulations can only reach mi-croseconds at the moment. Here, extending the minimalist modeldeveloped for protein folding, we propose the ‘‘switching Go�model’’ and use it to simulate the rotary motion of ATP-drivenmolecular motor F1–ATPase. The simulation recovers the unidirec-tional 120° rotation of the �-subunit, the rotor. The rotation wasinduced solely by steric repulsion from the �3�3 subunits, thestator, which undergoes conformation changes during ATP hydro-lysis. In silico alanine mutagenesis further elucidated which resi-dues play specific roles in the rotation. Finally, regarding themechanochemical coupling scheme, we found that the tri-sitemodel does not lead to successful rotation but that the always-bi-site model produces �30° and �90° substeps, perfectly in accordwith experiments. In the always-bi-site model, the number of sitesoccupied by nucleotides is always two during the hydrolysis cycle.This study opens up an avenue of simulating functional dynamicsof huge biomolecules that occur on the millisecond time scalesinvolving large-amplitude conformational change.

    always-bi-site model � energy landscape � funnel � switching Go� model

    B iomolecular machines, such as the ribosome, transporter,and molecular motors, fulfill their function through large-amplitude conformational change. Structural information be-fore and after the conformational change has been provided byx-ray crystallography and other methods for many cases (1).However, these methods do not directly observe the moleculardynamics that connects two-end structures. These dynamicalaspects can be observed directly by fluorescence and othertime-resolved spectroscopy; however, the latter methods moni-tor local structure but do not give global structural information.In this sense, molecular dynamics simulation is potentiallypowerful because it can provide full time-dependent structuralinformation about biomolecular machines (2–5). Yet, functionalcycles of these systems typically take milliseconds or longer,which is far beyond the current reach of molecular simulationswith all-atom standard force-fields (6). Thus, a complementaryapproach may be to coarse-grain the molecular representation,thereby enabling the simulation orders of magnitude-longer timescales.

    Coarse-graining drops some details from the model, whilehopefully keeping the essence. To achieve them, we need someperspective on the physical process we are trying to simply. Uponconformational change, some portions of molecules do notsignificantly change their structures; here, the atomic motion canbe well approximated by harmonic fluctuations. It has beenknown that very simple C� models, such as the elastic networkmodel (7), can reproduce low-frequency harmonic fluctuationfairly well. Conversely, large-scale conformational change in-volves rearrangement of nonlocal contacts between amino acids,which is beyond the harmonic approximation (8, 9). For approx-imating these rearrangements, we need to take into account forpartial disorder or partial unfolding, which induces conforma-tional diversity (8). To this end, we rely on the energy landscape

    perspective of proteins, which has been established in protein-folding studies. In this perspective, proteins evolved to havefunnel-like energy landscapes, where the native structure is atthe bottom of the funnel with the lowest effective energy anddisordered states have higher energies and large conformationalentropy (10, 11). Technically, a funnel-like energy landscape caneasily be realized by Go� models, which have become one of thestandard models for folding simulation studies (12–18). Forrepresenting large-amplitude conformational dynamics, Go�models are suitable because they account for both small f luc-tuations around the native basin and large fluctuations thatinvolves local unfolding. Here, we propose a previously unde-scribed computational framework, which we call a ‘‘switching Go�model,’’ for simulating large-amplitude motion of biomolecularmachines. We then apply it to the rotary motion of F1–ATPase.

    F1–ATPase, the catalytic part of the energy conversion en-zyme F0F1–ATP synthase, is known to act as a rotary motor uponATP hydrolysis (19–24). Single-molecule experiments have vi-sualized the rotation and elucidated some of the details; the�-subunit rotates unidirectionally in discrete 120° steps with eachATP hydrolysis (25), and the 120° step can be divided into 90°and 30° substeps at lower ATP concentration (26). The dwelltime before the 90° substep depends on ATP concentration, andthe one before the 30° substep includes two rates of an �1-mstime scale. These findings suggest that the 90° substep involvesATP binding, and the 30° substep may correspond to ATPhydrolysis and release of products (ADP, phosphate, or both).These experiments have given much insight into the rotationalmechanisms as described above; however, the resolution of thosestudies is limited to �40 nm, which cannot provide a completestructural view of the F1 dynamics.

    The minimal catalytic complex of F1 contains seven subunits�3�3�, whose x-ray structure (27) is depicted in Fig. 1a. Thecentral stalk � is a rotor, whereas the alternately arranged �3 and�3 subunits form a ring acting as the stator of the rotary motor.There are three catalytic nucleotide-binding sites, which are atthe interfaces of �-subunits with the neighboring �-subunits. Inthe crystal structure determined in 1994 (27), one catalytic siteis filled with ATP analog and another with ADP, and the last isempty. The structures of the ATP-bound subunits are denotedas �TP and �TP, where �TP takes on the ‘‘closed form’’ in its Cterminus (Fig. 1b Right Lower) (28) and the catalytic interface isloosely packed. The ADP-bound �-subunit, �DP, also takes theclosed form, and the interface to �DP is tightly packed. Thestructures of nucleotide-free subunits are denoted as �E and �E,where �E has ‘‘open form’’ in its C terminus (Fig. 1b Left Lower)(28). From the static structure, we can deduce that the catalysis-induced bending motion of � between closed and open formsproduce torque on � leading to rotation (27, 29). Given the staticx-ray structures at high resolution and the dynamic single

    Conflict of interest statement: No conflicts declared.

    This paper was submitted directly (Track II) to the PNAS office.

    ‡To whom correspondence should be sent at the * address. E-mail: [email protected].

    © 2006 by The National Academy of Sciences of the USA

    www.pnas.org�cgi�doi�10.1073�pnas.0509642103 PNAS � April 4, 2006 � vol. 103 � no. 14 � 5367–5372

    BIO

    PHYS

    ICS

    Dow

    nloa

    ded

    by g

    uest

    on

    Mar

    ch 3

    0, 2

    021

  • molecule observations at low resolution, currently missing is thetime-dependent structural insight that connects the two.

    In this work, by using our proposed switching Go� model, werealize structure-based molecular simulations of F1–ATPaserotation upon hydrolysis, providing us with time-dependentstructural insights on F1 rotation. We address which parts ofthe molecule play what roles in rotation in a site-specificmanner. Furthermore, combining all available experimentaldata with simulation results, we identify the mechanochemicalcoupling mechanism, which has long been under debate in thefield. This work opens an avenue of simulating large-scalemotion involved in dynamical function of large biomolecularcomplexes by folding-based model.

    ResultsSwitching Go� Model. There are many versions of Go� models, whichconcisely realize perfect funnel energy landscape (13) of proteins(13–18). Among them, we use the version of Clementi et al. (18),where amino acids are represented by single spheres centered atC� atoms, which are sequentially connected by virtual bonds, andinteractions are specified so that structures closer to the nativestructure are more stable (30). See Materials and Methods fordetails.

    The switching Go� model is based on the idea that the shapeof the funnel depends on chemical composition of the system andthus is altered upon ligand binding (31). We illustrate this notionfor the conformational transition in �-subunit of F1–ATPasefrom the nucleotide-free state ‘‘E’’ to the ATP-bound state ‘‘TP’’upon binding to ATP (Fig. 1b). In this case, we consider twofunnels; one for E state, at the bottom of which is the �Estructure (Fig. 1b, cyan funnel), and another one for the TP state,at the bottom of which is the �TP structure (Fig. 1b, greenfunnel). Go� potentials of the �-subunit for the states E and TPare denoted as V�(R��E�) and V�(R��TP�), respectively, where

    R� collectively represents the coordinates of amino acids in the�-subunit, and E and TP indicate the nucleotide state.

    We start simulations from the �E structure using the Go�potential V�(R��E�) appropriate for the nucleotide-free state E�(blue funnel in Fig. 1b). Upon binding ATP, we then switch theenergy to another Go� potential V�(R��TP�), appropriate for theATP-bound state TP� (green funnel in Fig. 1b). Fig. 1c plotsdegree of conformational change as a function of time. Theamount of conformational change is defined by a set of aminoacid pairs that are in the native contact at the final structure, �TP,and that are not in the native contact at the initial structure, �E(for the definition of the native contact, see Materials andMethods). Within the set, the fraction that is formed is used asthe measure of conformational change. We performed molec-ular dynamics simulations (32, 33) at a temperature T1 andswitched the Go� potential at the 10,000th time step, observingthe structural relaxation of the � from �E to �TP in 200 steps (Fig.1c). Simulation of relatively large-scale conformational changeswas achieved within a very short time by switching Go� model.

    Modeling the �3�3� Complex of F1–ATPase. Next, we move on to themodel of the F1 complex. The model of the �3�3� complex ofF1 has several important features. (i) A Go� potential V� funnelsthe simulated �-subunit into its structure of the 1994 complexso that the � only f luctuates around its native structure. (ii) Forthe �3�3 ring, we used the switching Go� potentialV�3�3(R��1,��2,��3�X��1,��2,��3), where the ring structure is fun-neled to the conformation characterized by nucleotide statesof the three ��-subunits X��1,��2,��3. Each of the three ��-subunits may have nucleotide states of either TP, DP � Pi, DP,or E states (here, we omit the possibility that Pi alone is bound,but the ADP is released based on structural observation).Thus, X��1��2��3 varies among 4

    3 � 64 types of nucleotidestates. Here, there is some room of uncertainty in mapping thenucleotide states to the reference structures, which will bedescribed thoroughly later. For example, when X��1,��2,��3 isthe nucleotide state {DP, E, TP}��1,��2,��3, the stable confor-mations of three ��-subunits are ��DP

    1994, ��E1994, and ��TP

    1994.[The superscript 1994 means that structures are taken from thex-ray structure determined in 1994 (27).] To induce confor-mational changes of the �3�3 ring, this Go� potential isswitched, e.g., for the simulation of conformational changeupon ATP binding at the catalytic site of the ��2, we switch theGo� potential from V�3�3(R��1,��2,��3�{DP, E, TP}��1,��2,��3) to

    Fig. 2. F1–ATPase rotary simulation by all-at-once catalysis approximation.The vertical axis is the � rotation angle. Each line represents the individualtrajectory, where the energy is switched from V�3�3(R��1,��2,��3�{DP, E,TP}��1,��2,��3) to V�3�3(R��1,��2,��3�{E, TP, DP}��1,��2,��3) at the 30,000th step. (a)Time courses of � rotation upon single ATP hydrolysis at the temperature T1.The � angles averaged over the time steps 20,000–30,000 and 50,000–60,000are �15.5° and 105.8°, respectively. (b) The � angle vs. degree of conforma-tional change of the �3�3 ring (horizontal axis). For the latter, 0 means thestructure of {��DP

    1994, ��E1994, ��TP

    1994}, and 1 corresponds to that of {��E1994, ��TP

    1994,��DP

    1994}.

    Fig. 1. Structure of F1–ATPase and illustration of switching Go� model. (a)Structure of F1–ATPase by Abrahams et al. (27), viewed from the outermembrane side. The central stalk � is the rotor. The ��-subunits of theATP-bound (yellow), the ADP-bound (gray), and the nucleotide-free parts aredenoted as (�TP, �TP), (�DP, �DP), and (�E, �E). (b) Concept of switching Go�model. In the simulation of ATP binding, we start the simulation withV�(R��E�), in which the simulated structure of the �-subunit is funneled to the�E structure, and switch it at a certain time to V�(R��TP�), where the structureof simulated � is funneled to the �TP structure. (c) Time courses of simulationsof a single �-subunit upon ATP binding at temperature T1. The switch fromV�(R��E�) to V�(R��TP�) is conducted at the 10,000th step. The value of thevertical axis close to 0.0 (1.0) means the structure is close to �E (�TP).

    5368 � www.pnas.org�cgi�doi�10.1073�pnas.0509642103 Koga and Takada

    Dow

    nloa

    ded

    by g

    uest

    on

    Mar

    ch 3

    0, 2

    021

  • V�3�3(R��1,��2,��3�{DP, TP, TP}��1,��2,��3). (iii) For the interac-tion between the � and the �3�3 ring V�3�3��, we used a simple,short-ranged steric repulsion between amino acids. Very im-portantly, the switching Go� potential was imposed only for theintra-�3�3 ring interactions, but not the � rotation, and thus therotation is not biased by design. (iv) The N termini of the three�-subunits and the C termini of � are constrained to theirinitial positions by springs Vspring, which mimics the His-tagused to fix F1 in the single molecule experiments (19). The totalpotential is the sum of four terms: V � V� � V�3�3 � V�3�3�� �Vspring. See Materials and Methods and Table 1, which ispublished as supporting information on the PNAS web site, fordetails.

    F1 Simulation of All-at-Once Catalysis. The first set of F1 simulationsstarted with the 1994 structure and the potentialV�3�3(R��1,��2,��3�{DP, E, TP}��1,��2,��3) for which the lowestenergy state corresponds to the 1994 structure, {��DP

    1994, ��E1994,

    ��TP1994} (27). After the 30,000th time steps, the potential is

    switched to V�3�3(R��1,��2,��3�{E, TP, DP}��1,��2,��3), in which thesimulated structures are funneled to the structure, {��E

    1994,��TP

    1994, ��DP1994}. The switching corresponds to a single cycle of the

    catalysis consisting of ADP release from the catalytic site of the��1 (termed site 1 hereafter), ATP binding at site 2, and ATPhydrolysis at site 3, which are conducted all at once. (Here, the

    phosphate, Pi, release may occur before or after reaching thenucleotide state of the 1994 structure.) Repeating the simula-tions 20 times, we monitored the rotary angle of � as a functionof time. In all trajectories, the � angle weakly fluctuated around�15.5° before the switching, and after the switching, the �rotated �120° in a counterclockwise direction, consistent withthe observations of the single-molecule experiment (19, 25) (Fig.2a; see also Movie 1, which is published as supporting informa-tion on the PNAS web site). Fig. 2b plots the � angle and thedegree of structural relaxation in the �3�3 ring, indicating thatstructural relaxation in the �3�3 ring is followed by the � rotation.The � rotation was induced solely by steric repulsion upon theconformational change in the �3�3 ring. Although factors notincluded in this simulation, such as hydrophobic and side-chaininteractions, may perturb the results quantitatively, structuralcomplementarities are of primary importance for the rotation.

    Critical Residues for Producing Torque. To decipher which parts ofthe � and �3�3 are essential for transferring the torque, weperformed a series of simulations analogous to Ala-scanningmutagenesis (34). In silico mutagenesis removed the stericinteractions of five consecutive residues in �, from �(i � 2) to�(i � 2), and for each i we repeated 40 simulations of all-at-oncecatalysis just as described in the previous subsection. Thenumber of trajectories that successfully completed the 120°rotation is shown at two different temperatures (T1 and T2 �5T1) in Fig. 3 as a function of central residue number i. At thelower temperature T1, we found three critical regions requiredfor the � rotation: �Ala-1–Asp-5, �Ile-13–Met-25, and ��la-231–Ala-235. We highlighted these three portions in white andthe portions of the �-subunits that contact them in yellow in Fig.4. Both �Ile-13–Met-25, and ��la-231–Ala-235 form contactswith the portion of � that shows the major structural change (theyellow portion in the C-terminal of the �-subunit colored by cyanin Fig. 4 a and c). This portion of � is not present in the highlyconserved DELSEED region but is localized nearby. This resultis consistent with the experimental inference that the DEL-SEED region does not play a critical role in torque transfer (35).The portions around the �Ile-13–Met-25 and �Ala-231–Ala-235regions are likely the sites of action of the � rotation. The otherimportant site, �Ala-1–Asp-5, keeps contacts with �- and �-sub-units at their middle portion, the switch II region that does notshow significant structural change (Fig. 4 b and c). Thus,�Ala-1–Asp-5 probably works as a fulcrum. Interestingly, anytrajectories that did not achieve a 120° rotation of � in thesemutagenesis simulations still showed a � rotation up to 30° from

    Fig. 3. Simulations analogous to Ala-scanning mutagenesis (34). The num-ber of completely rotated trajectories, out of 40 trials, is plotted against thecentral residue i of �, around which five consecutive residues are ‘‘mutated.’’All-at-once catalysis simulations were conducted at two temperatures, T1 (red)and T2 � 5 � T1 (blue).

    Fig. 4. Residues that transfer the torque. According to the results represented in Fig. 3, the three critical parts in � are plotted in white: �Ala-1–Asp-5,�Ile-13–Met-25, and �Ala-231–Ala-235. a–c are different views of the identical structure taken from a simulated snapshot after 30° rotation. The � subunits ingreen, cyan, and magenta are in the nucleotide states Tp, E, and DP, respectively, before all-at-once catalysis switch. Shown are views from �TP (a), �E (b), and�E (c). The subunit in pink (only in b) is the �-subunit in DP state. In b, only �1–44 is represented. The residues in � that are within 11 Å from the three criticalparts of � are in yellow.

    Koga and Takada PNAS � April 4, 2006 � vol. 103 � no. 14 � 5369

    BIO

    PHYS

    ICS

    Dow

    nloa

    ded

    by g

    uest

    on

    Mar

    ch 3

    0, 2

    021

  • the initial 1994 structure (27). (We show an example of thesimulations in Fig. 7, which is published as supporting informa-tion on the PNAS web site.) This result implies that the initial 30°rotation is less sensitive to steric interactions. At higher tem-peratures, the results were less clear, although a similar tendencywas found.

    Mechanochemical Coupling: Tri-Site Model vs. Always-Bi-Site Model.Next, we address how the catalytic reactions are related toconformational changes in �3�3 and � rotation substeps. Exper-iments showed that the 90° substeps are driven by ATP binding(26), and two, and only two, �-subunits are occupied by nucle-otides before the ATP binding (36). We first assumed that the1994 structure and its corresponding nucleotide state {DP, E,TP}��1,��2,��3 are on the pathway. Starting from this state, thenext step suggested by the above experiments is ATP binding atsite 2 of the 1994 structure (27), which is expected to make a 90°substep (Fig. 5a). This scenario is consistent with the modelgenerally called the tri-site model (37). To test this hypothesis,we performed simulations starting from V�3�3(R��1,��2,��3�{DP,E, TP}��1,��2,��3), of which the lowest energy state correspondsto the 1994 structure, {��DP

    1994, ��E1994, ��TP

    1994}, and at the30,000th step, we switched it to V�3�3(R��1,��2,��3�{DP, TP,TP}��1,��2,��3), which funneled the simulated structures to{��DP

    1994, ��TP1994, ��TP

    1994}. However, the simulations exhibited afatal collision between the �3�3 and � and did not show thesubstep or any unidirectional rotation (Fig. 5b). This resultsuggests that, for avoiding the steric clash, ��1 must open itsconformation before or concomitant with the closure of theC-terminal domain of ��2 upon ATP binding. All conceivablecatalytic schemes starting from the 1994 structure with {DP, E,TP}��1,��2,��3 were simulated. Those failed to reproduce resultsconsistent with experiments. As Abrahams et al. (27) suggested,the nucleotide state of {DP, E, TP}��1,��2,��3 with the 1994

    structure may correspond to the MgADP-inhibition state and isnot on the pathway.

    Here, three additional implications from experiments furtherinform the mechanistic hypothesis. (i) A FRET experimentsuggests that the 1994 structure corresponds to the one formedafter a 90° substep (38). (ii) A structural change in � uponphosphate release has been reported (39). These two observa-tions and the above simulation results led us to redefine thenucleotide state of the 1994 structure as {DP � Pi, E,TP}��1,��2,��3, which is on-pathway, and to assume that the�-subunit changes its structure upon phosphate release fromclosed to ‘‘half-closed’’ form (28), the latter taken from thestructure reported in 2001 (40). Furthermore, (iii) one scenariothat is consistent with the experiments of Nishizaka et al. (36) isthat the site occupancy of nucleotides is always 2 (41). Finally,we suggest simultaneous reactions of ATP binding and ADPrelease. The catalysis scheme thus starts from {DP � Pi, E,TP}��1,��2,��3 with the 1994 structure. Then, the phosphate isreleased at site 1, which is followed by the simultaneous reactionsof ATP binding at site 2 and the ADP release at site 1. Finally,ATP hydrolysis occurs at site 3 (Fig. 6a).

    Simulations consistent with this mechanism were conducted

    Fig. 5. Simulation of a step in the tri-site model. (a) Cartoon diagrams. Thediagrams correspond to the part of the tri-site model. (Upper) The transitionof the nucleotide states in the catalytic scheme. (Lower) The reference struc-tures used in the simulation. Each circle represents an ��-subunit, and char-acters in it indicate the nucleotide state of the ��-subunit at the top and thereference structure at the bottom. Red arrow indicates the angle of � obtainedby simulations. (b) Time courses of � angle based on the catalytic model. Eachline represents individual trajectory. ATP binding was conducted at the30,000th step.

    Fig. 6. Simulation with the always-bi-site model. (a) Cartoon diagrams. Seethe legend of Fig. 5 for description of Upper and Lower. (b) Time courses of �angle based on the catalytic model. Each line represents individual trajectory.Simulations contain four stages shown in a, and each stage corresponds to30,000 time steps. (c) Histogram of the � angle from simulations shown in b.

    5370 � www.pnas.org�cgi�doi�10.1073�pnas.0509642103 Koga and Takada

    Dow

    nloa

    ded

    by g

    uest

    on

    Mar

    ch 3

    0, 2

    021

  • (Fig. 6 a and b). We first simulated with V�3�3(R��1,��2,��3�{DP �Pi, E, TP}��1,��2,��3), where site 1 is funneled to ��DP of the1994 structure; the reference structure is {��DP

    1994, ��E1994,

    ��P1994}. At the 30,000th step, this potential is switched to

    V�3�3(R��1,��2,��3�{DP, E, TP}��1,��2,��3), where site 1 has thefunnel centered at the ��DP�

    2001, the half-closed � structure in the2001 complex (40); the reference structure is {��DP�

    2001, ��E1994,

    ��TP1994}. Next, we switched the potential to V�3�3(R��1,��2,��3�{E,

    TP, TP}��1,��2,��3) at the 60,000th step, and V�3�3(R��1,��2,��3�{E,TP, DP}��1,��2,��3) at the 90,000th step. The former potentialfunnels the simulated structures to {��E

    1994, ��TP1994, ��TP

    1994} andthe latter to {��E

    1994, ��TP1994, ��DP

    1994}. The resulting trajectoriesand histogram are shown in Fig. 6 b and c (see Movie 2, whichis published as supporting information on the PNAS web site)where we found the �30° substep upon the phosphate releaseand the �80° substep upon the coupled reactions of the ATPbinding and ADP release, which are perfectly consistent withexperiments. Interestingly, the �-subunit rotated �10° uponATP hydrolysis. This finding might be related to the discrepancybetween 90° � 30° substeps with ATP as a substrate and 80° �40° substeps with slowly hydrolysable substrate ATP[�-S] (42).

    Discussion and ConclusionThe results from the experiment and simulation led us to thefollowing rotational mechanisms. Starting from the 1994 struc-ture and the nucleotide state {DP � Pi, E, TP}��1,��2,��3,remodeling of the site 1 �-subunits from closed to half-closedform upon phosphate release makes space between the �3�3 andthe �, leading to 30° rotation of �. Next, ATP-binding at site 2induces the �-subunits closing motion, which pushes the �-sub-unit primarily through contacts near �Ile-13–Met-25, and ��la-231–Ala-235 leading to � rotation. This motion further induces� opening at site 1, thus accelerating the ADP release. The latter� opening motion at site 1 is crucial for making space for � tocomplete �80° rotation (see Movie 2). The ATP-binding, the 80°rotation, and ADP release are so tightly coupled that they are inthe single kinetic process. Hydrolysis might induce � rotation of�10°. We note that current simulations do not rule out theso-called bi-site model (43) at very low ATP concentration.

    It is surprising that steric repulsion between the � and the�3�3 ring alone reproduced not only 120° rotation, but also 30°and 90° substeps. Here, the shape complementarities betweenthe � and �3�3 play crucial roles in functional dynamics of F1.It may be viewed as the ‘‘dynamic version of lock-and-key’’mechanism; the shape of the keyhole, the �3�3 ring, changesin the catalytic cycle, which rotates the key, the �-subunit, bystructural complementarities. It may be one reason for F1 toexhibit more deterministic motion (25) than that of linearmolecular motors such as myosin (44).

    As in the F1–ATPase rotary motion, biomolecular functionalmotion takes milliseconds, a time scale that is orders of magni-tude longer that the current reach of atomistic molecular dy-namics simulations (3, 4). A folding-based simulation modelprovides a powerful tool of deducing functional dynamics oflarge biomolecular complexes from static structural information,once structures in multiple states are experimentally supplied.Here, we reported an initial application of this approach toF1–ATPase.

    There are many ways to further improve the current model ofF1. First of all, a remarkable feature of F1–ATPase is itsreversibility. By enforcing the rotation of �, F1–ATPase cansynthesize ATP from ADP and Pi (45). For simulating it, we needto take into consideration the feedback mechanism from the�3�3 ring structure to the chemical reaction of nucleotide.Second, the current switching Go� model introduced the ‘‘verticalexcitation’’ for jumping from the initial funnel to the final funnel(see Fig. 1b). Physically, this way of simulation corresponds to‘‘photoexcitation,’’ whereas the real process studied is of course

    thermally activated. The present approach may be viewed as arough approximation of the real process for addressing struc-ture–function relationship. For pursuing more physical aspects,however, the present approach needs to be modified. Recently,methods in this direction were investigated (8, 9, 46). Third, wedid not explicitly include nucleotide molecules, such as ATP,here. Explicit inclusion of ligand molecules is desired. Improve-ment toward these directions is needed.

    Materials and MethodsMolecular Simulation with Go� Models. Clementi’s version of the C�Go� model has the effective energy function

    V�R�X � �bonds

    Kr�r i � rX,i2 � �angles

    K��� i � �X,i2

    � �dihedral

    K��1�1 � cos�� i � �X,i�

    � K��3�1 � cos 3�� i � �X,i�

    � �i�j�3

    nativecontact

    � 5� rX,ijr ij �12

    � 6� rX,ijr ij �10�

    � �i�j�3

    nonnativecontact

    �D1r ij �12

    . [1]

    Here, R collectively represents the Cartesian coordinates of C�atoms. X signifies the native structure of the simulated protein.In this equation, the first term keeps the chain connectivity; ri isthe length of the ith virtual bond between the ith and i�1-thamino acids. The second and the third terms represent the localtorsional interactions; �i, and i stand for the virtual bond anglebetween the i�1-th and ith bonds and the virtual dihedral anglearound the i-h bond, respectively. The fourth and fifth terms arenonlocal interactions. The former includes interactions between‘‘native contact’’ pairs of amino acids that are spatially close inthe native structure (see below for details), and the last term isa nonspecific repulsion between amino acid pairs. The rij is thedistance between the C� atoms of the ith and jth amino acids.Structural information on X is used for setting the parameters;all of the constant parameters with the subscript X indicate thevalues of the corresponding variables at the structure X. Forexample, because the variable rij is the distance between the ithand jth residues, the parameter rXij has the value of rij in the Xstructure. Native contact in the fifth term is defined as below. Ifone of the nonhydrogen atoms in the ith amino acid is within adistance of 6.5 Å from any nonhydrogen atom in the jth aminoacid, we define the pair of the ith and jth amino acids as beingnative contact (amino acid pairs that are not in the native contactare called nonnative pairs). We note that all of the terms exceptthe last one are set up so that each term has the lowest energywhen the conformation R coincides with the structure X: theseterms realize the funnel-like energy landscape. Parameters Kr �100.0, K� � 20.0, K�

    (1) � 1.0, K�(3) � 0.5, � 0.36, and D1 � 4.0

    are used throughout the present work. In the switching Go� modelproposed here, the reference structure X is switched uponchange in the nucleotide state.

    The energy function for the F1 complex consists of four terms:V � V� � V�3�3 � V�3�3�� � Vspring. V� is a Go� model for �subunit, in which the bottom structure X corresponds to the �structure in the 1994 complex. V�3�3(R��1,��2,��3�X��1,��2,��3) is theswitching Go� model for the �3�3 ring, where the referencestructure X depends on the nucleotide states. For example, whenX � {DP, E, TP}, the stable conformations of three ��-subunits

    Koga and Takada PNAS � April 4, 2006 � vol. 103 � no. 14 � 5371

    BIO

    PHYS

    ICS

    Dow

    nloa

    ded

    by g

    uest

    on

    Mar

    ch 3

    0, 2

    021

  • are ��DP1994, ��E

    1994, and ��TP1994 (see Crystal Structures Used in

    Materials and Methods). V�3�3�� is the short-range steric repul-sion defined as

    V�3�3�� � �i��3�3, j��

    �D2�rij12, [2]

    where D2 � 6.0 Å. Vspring is introduced to keep the F1 complexon a plane and is defined as

    Vspring � �i��19,�29,�39,�272

    � k�Li � Li,02 �Li � Li,0 ,0 �Li � Li,0 ,[3]

    where k � 0.18, Li is the distance of the ith residue from its initialposition, and Li,0 � 1.5 Å (i � �19, �29, �39), and Li,0 � 2.0 Åfor i � �272. See Supporting Text, which is published as sup-porting information on the PNAS web site, for some moredetails.

    Molecular dynamics simulations were performed by integrat-ing Newton equation with ‘‘velocity Verlet’’ algorithm (32), andthe velocity rescaling proposed by Berendsen et al. (33) was usedto realize constant temperature simulation. Temperatures usedare T1 � 4 � 10�3 and 5T1. The mass for all amino acids wasidentical.

    Crystal Structures Used. We used the reference structures of the�3�3� complex that correspond to the bottoms of the funnels inGo� potentials. All of the reference structures were taken fromthe two crystal structures determined in 1994 and 2001 by

    Walker and coworkers, which we refer as the 1994 (27) and the2001 (40) structures, respectively. As for the � stalk, we used theconformation in the 1994 structure as the reference structurethroughout the work. Conversely, the reference structure of the�3�3 ring varied depending on the states of three catalyticbinding sites. Because the catalytic nucleotide binding sites arelocated at the interfaces between the �- and �-subunits, wetreated the �- and �-subunits that sandwich the catalytic sitesgrouped together, which are colored by analogous hue in Fig. 1a.From the 1994 and 2001 structures, we took four differentreference structures of grouped ��-subunits. We denoted eachof them as ��A

    y : ��DP1994, ��E

    1994, ��TP1994, or ��DP�

    2001. The superscripty represents the parent structure, from which the structure of ��is taken; y is either 1994 or 2001. In the 1994 structure, A is TP,DP, or E. ��TP

    1994 has ATP-analog in the catalytic site; thestructure of the �-subunit is in the closed form in C-terminalregion, and the interface is loosely packed. ��DP

    1994 includes ADPin the catalytic site; the �-subunit structure is similar to that inTP1994, whereas the interface is tightly packed. ��E

    1994 does nothave nucleotides at the catalytic site, and the �-subunit takes theopen form in C-terminal region. In the 2001 structure, we pickedup one ��-structure as the reference structure, which the�-subunit takes as a half-closed form (28): its C-terminal isbetween the closed form in ��TP

    1994 and the open form in ��E1994.

    We call it ��DP�2001.

    We thank Jeffery Saven and Peter Wolynes for many helpful suggestions.N.K. is supported by Japan Society for the Promotion of ScienceResearch Fellowship for Young Scientists. This work was supported inpart by the ‘‘Water and Biomolecules’’ program from the Ministry ofEducation, Culture, Sports, Science, and Technology (Japan).

    1. Gerstein, M. & Echols, N. (2004) Curr. Opin. Chem. Biol. 8, 14–19.2. Ma, J., Flynn T. C., Cui, Q., Leslie, A. G., Walker, J. E. & Karplus, M. (2002)

    Structure (London) 10, 921–931.3. Kuhlman, B., O’Neill, J. W., Kim, D. E., Zhang, K. Y. & Baker, D. (2002) J.

    Mol. Biol. 315, 471–477.4. Bockmann, R. A. & Grubmuller, H. (2002) Nat. Struct. Biol. 9, 198–202.5. Sun, S., Chandler, D., Dinner, A. R. & Oster, G. (2003) Eur. Biophys. J. 32,

    676–683.6. Daggett, V. (2003) Protein Simulations (Elsevier Academic, San Diego).7. Tirion, M. M. (1996) Phys. Rev. Lett. 77, 1905–1908.8. Miyashita, O., Onuchic, J. N. & Wolynes, P. G. (2003) Proc. Natl. Acad. Sci.

    USA 100, 12570–12575.9. Maragakis, P. & Karplus, M. (2005) J. Mol. Biol. 352, 807–822.

    10. Bryngelson, J. D. & Wolynes, P. G. (1987) Proc. Natl. Acad. Sci. USA 84,7524–7528.

    11. Leopold, P. E., Montal, M. & Onuchic, J. N. (1992) Proc. Natl. Acad. Sci. USA89, 8721–8725.

    12. Go� , N. (1983) Annu. Rev. Biophys. Bioeng. 12, 183–210.13. Nymeyer, H., Garcia, A. E. & Onuchic, J. N. (1998) Proc. Natl. Acad. Sci. USA

    95, 5921–5928.14. Shoemaker, B. A., Wang, J. & Wolynes, P. G. (1997) Proc. Natl. Acad. Sci. USA

    94, 777–782.15. Alm, E. & Baker, D. (1999) Proc. Natl. Acad. Sci. USA 96, 11305–11310.16. Galzitskaya, O. V. & Finkelstein, A. V. (1999) Proc. Natl. Acad. Sci. USA 96,

    11299–11304.17. Munoz, V. & Eaton, W. A. (1999) Proc. Natl. Acad. Sci. USA 96, 11311–11316.18. Clementi, C., Nymeyer, H. & Onuchic, J. N. (2000) J. Mol. Biol. 298, 937–953.19. Noji, H., Yasuda, R., Yoshida, M. & Kinosita, K., Jr. (1997) Nature 386,

    299–302.20. Boyer, P. D. & Kohlbrenner, W. E. (1981) in Energy Coupling in Photosynthesis,

    eds. Selman, B. R. & Selman-Reimer, S. (Elsevier, Amsterdam), pp. 231–40.21. Oosawa, F. & Hayashi, S. (1983) J. Phys. Soc. Jpn. 52, 4019–4028.22. Cox, G. B., Jans, D. A., Fimmel, A. L., Gibson, F. & Hatch, L. (1984) Biochim.

    Biophys. Acta 768, 201–208.23. Duncan, T. M., Bulygin, V. V., Zhou, Y., Hutcheon, M. L. & Cross, R. L. (1995)

    Proc. Natl. Acad. Sci. USA 92, 10964–10968.24. Sabbert, D., Engelbrecht, S. & Junge, W. (1996) Nature 381, 623–625.25. Yasuda, R., Noji, H., Kinosita, K., Jr., & Yoshida, M. (1998) Cell 93, 1117–1124.

    26. Yasuda, R., Noji, H., Yoshida, M., Kinosita, K., Jr., & Itoh, H. (2001) Nature410, 898–904.

    27. Abrahams, J. P., Leslie, A. G., Lutter, R. & Walker, J. E. (1994) Nature 370,621–628.

    28. Tanaka, K., Waki, H., Ido, Y., Akita, S., Yoshida, Y., Yoshida, T. (1988) RapidCommun. Mass Spectrom. 2, 151–153.

    29. Wang, H. & Oster, G. (1998) Nature 396, 279–282.30. Koga, N. & Takada, S. (2001) J. Mol. Biol. 313, 171–180.31. Kumar, S., Ma, B., Tsai, C. J., Sinha, N. & Nussinov, R. (2000) Protein Sci. 9,

    10–19.32. Swope, W. C., Andersen, H. C., Berens, P. H. & Wilson, K. R. (1982) J. Chem.

    Phys. 76, 637–649.33. Berendsen, H. J., Postma, J. P. M., van Gunsteren, W. F., DiNola, A. & Haak,

    J. R. (1984) J. Chem. Phys. 81, 3684–3690.34. Cunningham, B. C. & Wells, J. A. (1989) Science 244, 1081–1085.35. Hara, K. Y., Noji, H., Bald, D., Yasuda, R., Kinosita, K., Jr., & Yoshida, M.

    (2000) J. Biol. Chem. 275, 14260–14263.36. Nishizaka, T., Oiwa, K., Noji, H., Kimura, S., Muneyuki, E., Yoshida, M. &

    Kinosita, K., Jr. (2004) Nat. Struct. Mol. Biol. 11, 142–148.37. Weber, J., Wilke-Mounts, S., Lee, R. S., Grell, E. & Senior, A. E. (1993) J. Biol.

    Chem. 268, 20126–20133.38. Yasuda, R., Masaike, T., Adachi, K., Noji, H., Itoh, H. & Kinosita, K., Jr. (2003)

    Proc. Natl. Acad. Sci. USA 100, 9314–9318.39. Masaike, T., Muneyuki, E., Noji, H., Kinosita, K., Jr., & Yoshida, M. (2002)

    J. Biol. Chem. 277, 21643–21649.40. Menz, R. I., Walker, J. E. & Leslie, A. G. (2001) Cell 106, 331–341.41. Kinosita, K., Adachi, K. & Itoh, H. (2004) Annu. Rev. Biophys. Biomol. Struct.

    33, 245–268.42. Shimabukuro, K., Yasuda, R., Muneyuki, E., Hara, K. Y., Kinosita, K., Jr., &

    Yoshida, M. (2003) Proc. Natl. Acad. Sci. USA 100, 14731–14736.43. Boyer, P. D. (1997) Annu. Rev. Biochem. 66, 717–749.44. Kitamura, K., Tokunaga, M., Iwane, A. H. & Yanagida, T. (1999) Nature 397,

    129–134.45. Itoh, H., Takahashi, A., Adachi, K., Noji, H., Yasuda, R., Yoshida, M. &

    Kinosita, K. (2004) Nature 427, 465–468.46. Best, R. B., Chen, Y. G. & Hummer, G. (2005) Structure (London) 13,

    1755–1763.

    5372 � www.pnas.org�cgi�doi�10.1073�pnas.0509642103 Koga and Takada

    Dow

    nloa

    ded

    by g

    uest

    on

    Mar

    ch 3

    0, 2

    021


Recommended