ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

Embed Size (px)

Citation preview

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    1/13

    Review

    Synthase (H ATPase): coupling between catalysis, mechanical work,

    and proton translocation

    Masamitsu Futai *, Hiroshi Omote, Yoshihiro Sambongi, Yoh Wada

    Division of Biological Sciences, Institute of Scientic and Industrial Research, Osaka University, CREST,

    Japan Science and Technology Corporation, Ibaraki, Osaka 567-0047, Japan

    Abstract

    Coupling with electrochemical proton gradient, ATP synthase (F0F1) synthesizes ATP from ADP and phosphate.

    Mutational studies on high-resolution structure have been useful in understanding this complicated membrane enzyme. We

    discuss mainly the mechanism of catalysis in the L subunit of F1 sector and roles of the Q subunit in energy coupling. The

    Q-subunit rotation during catalysis is also discussed. 2000 Elsevier Science B.V. All rights reserved.

    Keywords: F0F1 ; Catalytic site; ATP synthesis; Proton transport; H ATPase; Rotational catalysis

    1. Introduction

    The ATP synthases (H-ATPases) from the mem-branes of mitochondria, chloroplasts and bacteria

    are similar in structure and mechanism [1^7]. They

    are formed from a membrane extrinsic F1 and a

    transmembrane F0 sectors. The F1 is a catalytic sec-

    tor consisting of ve subunits, K, L, Q, N, and A with a

    stoichiometry of 3:3:1:1:1. The F0 is a membrane

    sector formed from three subunits, a, b, and c with

    a stoichiometry of 1:2:12 (Fig. 1a). The X-ray dif-

    fraction of the bovine crystal solved the K3L3Q com-

    plex structure at 2.8 A resolution [8], while the struc-

    ture of the entire F0 sector has only been observed byelectron or atomic force microscopy [9^11].

    The complicated structure enables the enzyme to

    carry out ATP synthesis, a simple chemical reaction

    `ATPHADP+Pi', with high eciency in the dened

    cytoplasmic compartments. The unique feature ofthis enzyme mechanism is its strong catalytic coop-

    erativity and coupling to an electrochemical proton

    gradient, vWH, formed across membranes. The

    vWH is not required for the chemical reaction at

    the catalytic site synthesizing ATP from ADP and

    Pi (inorganic phosphoric acid), but serves primarily

    for releasing product ATP. The entire ATP synthesis

    is carried out by the binding-change mechanism pro-

    posed by Boyer [5].

    Escherichia coli ATP synthase has been studied

    extensively. The obvious advantage of studying thebacterial enzyme is that both forward and reverse

    genetic approach can be easily applied. Mutant en-

    zymes revealed amino acid residues responsible for

    catalysis, proton transport and energy coupling [1^3].

    The functional interaction of dierent domains can

    be studied genetically by identifying the second mu-

    tation(s) which suppresses the initial mutation. Cata-

    lytic residues identied by the combined biochemical

    and genetic approaches are actually found nearby the

    0005-2728 / 00 / $ ^ see front matter 2000 Elsevier Science B.V. All rights reserved.

    PII: S 0 0 0 5 - 2 7 2 8 ( 0 0 ) 0 0 0 8 0 - 3

    * Corresponding author. Fax: +81-6-6875-5724;

    E-mail: [email protected]

    Biochimica et Biophysica Acta 1458 (2000) 276^288

    www.elsevier.com/locate/bba

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    2/13

    phosphate moiety of ATP or ADP in the X-ray

    structure of the bovine F1 [8]. Mutants defective in

    energy coupling between catalysis and proton trans-

    port have been isolated, providing important infor-

    mation about the mechanism.

    It has been speculated that ATP synthesis or hy-

    drolysis by the binding-change mechanism is coupled

    to proton transport by the mechanical rotation of theQ^A^c assembly extending through F1F0 ; vWH

    drives the Q^A^c assembly rotation, which causes se-

    quential conformational changes (binding-change) of

    the three L subunits, resulting in product ATP release

    [5,12]. The high-resolution structure of K3L3Q com-

    plex [8] is consistent with the rotation of the Q sub-

    unit during catalysis and three dierent conforma-

    tions of the L subunits correlate well with the

    binding-change mechanism. The rotation during

    ATP synthesis or hydrolysis has been suggested by

    biochemical and biophysical studies [13^18], includ-

    ing post-photobleaching recovery of a probe linked

    to the Q subunit [17,18]. A more convincing result is

    the direct observation of the rotation. The ATP-hy-

    drolysis-dependent rotation of actin lament at-

    tached to the thermophilic bacterial Q subunit in

    the K3L3Q complex was observed and video-recorded

    [19]. The Q-subunit rotation of the E. colienzyme was

    also directly observed [20]. These results indicate that

    the two bacterial enzymes produce essentially the

    same frictional torque. Thus, ATP synthesis or hy-

    drolysis by F0F1 is a coupling of three distinct steps:

    catalysis, mechanical work and proton transport

    (Fig. 1b). From the previous work, it became possi-

    ble to study the detailed mechanism utilizing awealth of information obtained by combined bio-

    chemical and genetic approaches [1^4].

    As ATP hydrolysis-dependent rotation of the

    Q subunit in F1 has been shown clearly, the ATP-de-

    pendent H translocation (reverse reaction of ATP

    synthesis) should include rotation. The energy of

    ATP hydrolysis drives Q-subunit rotation, followed

    by proton transport to establish vWH. The two cou-

    pling processes are obviously present: coupling be-

    tween ATP hydrolysis and the Q-subunit rotation

    (mechanical work) and that between the rotation

    and proton translocation. In ATP synthesis, proton

    transport driven by vWH rotates the Q subunit and

    this rotation changes the conformation of the cata-

    lytic site and nally releases product ATP. More

    work will certainly be required to understand the

    coupling mechanism of the dierent steps. We still

    have many questions to solve. For example, it is un-

    known as to whether vWH actually drives the Q-sub-

    unit rotation.

    The catalysis by the F1 sector of the ATP synthase

    is discussed in this article, mainly focusing on the

    mechanism of the basic chemical reaction and cata-lytic cooperativity in the steady-state reactions lead-

    ing to the Q-subunit rotation. Also, the coupling of

    the Q-subunit rotation with proton transport is dis-

    cussed. Emphasis is given to the E. coli enzyme, be-

    cause it has been studied extensively using genetic

    approaches. We discuss E. coli mutant results refer-

    ring to the high-resolution structure of the bovine

    enzyme, supported by the homology of the amino

    acid sequences between the two species. This ap-

    proach could contribute greatly to an understanding

    of the enzyme, although it is desirable to have an

    E. coli crystal structure. In this regard, the thermo-

    philic bacterial K3L3 complex structure is closely sim-

    ilar to bovine K3L3Q [21]. This article is not intended

    to be a comprehensive review, but rather summarizes

    important aspects of the enzyme and discusses inter-

    esting questions. Naturally, amino acid residues dis-

    cussed below are numbered following the E. coli se-

    quence. Excellent studies not cited in this article have

    been reviewed previously from dierent aspects [1^7].

    Fig. 1. (a) A model of ATP synthase F0F1. Schematic subunit

    organization of ATP synthase. Catalysis and proton transloca-

    tion are shown schematically. (b) Catalysis, mechanical work

    and proton translocation during ATP synthesis and hydrolysisby ATP synthase.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288 277

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    3/13

    2. Mechanism of catalysis in F1 sector of

    ATP synthase

    2.1. Kinetics

    The catalytic site of the enzyme is located mostly

    in the L subunit of the F1 sector, and the binding-

    change mechanism predicts that all three sites, each

    in a single L subunit, participate in catalysis sequen-

    tially [5]. The participation of only two sites in cat-

    alysis is not consistent with kinetic data and the

    high-resolution structure, as discussed previously.

    Boyer and coworkers showed that F1 catalyzes ex-change between the oxygen of water and oxygen of

    phosphate during steady-state ATP hydrolysis at low

    ATP concentration [5,22]. However, this exchange

    was not observed with high ATP concentration.

    These results indicate that ATP hydrolysis and syn-

    thesis at the catalytic site are reversible at low ATP

    concentration, but the reaction proceeds only in the

    direction of hydrolysis at high ATP concentration,

    due to the catalytic cooperativity.

    The two types of F1-ATPase catalysis have been

    clearly shown: unisite (single site) and multisite

    (steady-state) catalysis can be measured in the pres-

    ence of a sub-stoichiometric and excess amount of

    ATP, respectively [6,23,24] (Fig. 2). Bisite catalysis

    was also shown. Values of Km (or Kd) for uni, bi-

    and multi (or tri-) site catalysis of bovine F1 were

    shown to be 10312 M, 3U1035 M, and 1.5U1034 M,

    respectively [24]. Kinetic analysis of unisite catalysis

    indicated that F16ATP (ATP-bound F1) can be

    formed from F16ADPWPi (ADP and Pi bound F1)

    with little free energy change [22]. These observations

    predict that vWH is required for dissociation of

    ATP from the catalytic site, but not for the synthesis

    of F16ATP from F16ADPWPi, supporting the

    binding-change mechanism [5]. The rate of multisitecatalysis is 105^106-fold higher than unisite catalysis,

    due to the cooperativity between three catalytic sites.

    The basic kinetic feature of the enzyme, briey de-

    scribed here, was discussed in detail previously

    [1,5,6]. Presence of unisite catalysis and catalytic co-

    operativity was questioned recently by measuring the

    disappearance of ATP during ATP hydrolysis by mi-

    tochondrial F1 [25]. However, the reported results

    seem not to disprove the previous kinetic studies.

    Biochemical and genetic results, including selective

    loss of multisite catalysis by mutations or inhibitors,

    obviously support positive catalytic cooperativity

    and unisite catalysis [1^5].

    2.2. Nucleotide binding to F1-ATPase

    F1 has three non-catalytic nucleotide binding sites,

    located mostly in the K subunit, together with the

    three catalytic sites [1^5]. Mutational studies indi-

    cated that non-catalytic sites do not participate in

    catalysis [26,27]. Thus, binding of ATP or ADP to

    catalytic sites had been dicult to analyze speci-

    cally. The X-ray structure clearly indicates that thebovine residue corresponding to the LTyr331 is lo-

    cated close to the adenine ring of ADP or ATP

    (AMPWPNP) bound to the catalytic site [8]. This res-

    idue is also shown to be the 2-azido ATP-binding site

    [28]. From the structural consideration, Senior and

    coworkers promptly replaced this residue by trypto-

    phan [29,30]. The LTyr331Trp mutant is essentially

    similar to the wild type, in catalysis and energy cou-

    pling. The intrinsic uorescence of the tryptophan

    residue introduced in the LTyr331Trp mutant could

    be a signal from the active site.

    As expected from its location, LTrp331 uores-

    cence was quenched upon ATP binding to the cata-

    lytic site [29,30]. Titration of the uorescence with

    MgATP clearly indicated the presence of three cata-

    lytic nucleotide binding sites with negative coopera-

    tivity: Kd1, 9 5U1038 M; Kd2, 10

    36 M; Kd3, 1035 M

    [29]. These results suggest that all three catalytic sites

    are occupied under physiological conditions

    (ATPs 1 mM). On the other hand, the bovine X-

    Fig. 2. Kinetics of ATP hydrolysis by F1 sector. The kinetic

    mechanism of ATP hydrolysis by puried F1 is shown sche-

    matically. In multisite catalysis, three catalytic sites participate

    in the reaction, although two sites are shown in the gure.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288278

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    4/13

    ray structure shows three dierent forms of the

    L subunit: LTP and LDP (AMPcPNP and ADP

    bound forms, respectively) and LE (empty form)

    [8]. Consistent with the structure, two catalytic sites

    of LTyr331Trp mutant were lled and one site re-mained empty under crystallization conditions [31].

    Thus, the crystal F1 corresponds to an enzyme spe-

    cies that releases product during steady-state cataly-

    sis, as suggested previously.

    2.3. ATP synthesis/hydrolysis at the catalytic site

    2.3.1. Residues in the P-loop (phosphate-binding

    loop)

    In the bovine LTP and LDP subunits, one lysine,

    aspartic acid, and threonine together with two argi-

    nine and glutamate residues are found near the L/Q

    phosphate moiety of the bound nucleotide. The cor-

    responding E. coli residues are LLys155, LThr156,

    LGlu181, LArg182, LGlu185, and LAsp242 of the

    L subunit and KArg376 of the K subunit (Fig. 3).

    The functional importance of these residues was in-

    dicated by extensive mutagenesis before the crystal

    structure became available. The LLys155, LThr156,

    LGlu181 and LArg182 were suggested to be catalytic

    residues mainly by biochemical studies of mutant

    enzymes.

    The LLys155 and LThr156 are in the glycine-richsequence or P-loop (phosphate binding loop ; consen-

    sus, Gly-X-X-X-X-Gly-Lys-Thr). The importance of

    the loop has been suggested from its high degree of

    conservation among nucleotide binding proteins in-

    cluding F1L subunits. Mutations [32^36] and anity

    labeling [37] of the residues in the corresponding re-

    gion (LGly149^LThr156) indicated importance of the

    loop. Replacements of even non-conserved residues

    such as LAla151 drastically altered enzyme catalysis

    [32,33]. LLys155Ala or LLys155Ser mutants showed

    low multisite and unisite catalysis. The rate of ATP

    binding (k1, Fig. 2) was 10-fold lower than the wild-

    type enzyme [34]. Kinetic studies of the puried mu-

    tant F1 suggested that the A-amino moiety of the

    LLys155 side chain interacts with the Q phosphate

    of ATP at the catalytic site, contributing necessary

    binding energy to drive catalysis [35,36]. Analysis of

    the tryptophan signal of LLys155Gln/LTyr331Trp

    enzyme concluded that LLys155 interacts mainly

    with the Q phosphate of MgATP and is primarily

    important for site 1 (the highest anity catalytic

    site) and site 2 [38]. The X-ray structure of bovine

    F1 supports these results. The distance between the

    A-amino moiety of LLys155 to the nearest Q-phos-

    phate oxygen atom of MgAMPcPNP is 2.7 A , and

    to the L phosphate is 3.3 A [8].

    LThr156 could be replaced by Ser, while maintain-

    ing enzyme activity, but not by other residues [34].

    Furthermore, the threonine residue could not be

    moved to position 157 [33]. Measurement of Mg2-

    binding to the L subunit^aurovertine complex sug-

    Fig. 3. Catalytic site in ATP synthase. Catalytic site of ATP

    synthase is shown with residues required for catalysis and coop-

    erativity in LTP (ATP bound), LDP (ADP bound) and LE

    (without nucleotide). The positions of amino acid residues were

    taken from the bovine structure [9] but shown by E. coli num-

    bering. This gure was prepared using Molscript [72].

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288 279

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    5/13

    gested that LThr156, possibly its hydroxyl moiety,

    contributes to Mg2 binding to the catalytic site

    [34]. Consistent with this interpretation, the distance

    between Mg2 and oxygen of LThr156 side chain is

    the shortest (2.2 A ) in the LTP of the bovine crystalstructure [8].

    2.3.2. Residues in the GERXXE sequence and a model

    of catalysis

    The GERXXE (LGly180^LGlu185) is a sequence

    conserved among the L subunit of the F1 sector and

    subunit A of V-ATPase from various species [39,40].

    Replacements ofLGlu181 with Gln or Ala gave a F1with no multisite and very slow unisite catalysis [39].

    The mutant defects can be attributed to the 102^103-

    fold decreases of k2 and k32 (rates of ATP hydro-

    lysis and synthesis [F16ATPHF16ADPcPi] at the

    catalytic site, respectively) of unisite catalysis, while

    mutations showed k1 and k31 (rates of ATP binding

    and release, respectively) 10-fold less than the wild

    type. Analysis of LGlu181Gln/LTyr331Trp enzyme

    suggests that LGlu181 contributes little to MgATP

    binding [38]. The position of the carboxyl side chain

    of LGlu181 should be critical, because LGlu181Asp

    mutant enzyme showed a 10-fold lower k2, k32, and

    overall rate of unisite catalysis than wild type. This

    side chain is hydrogen-bonded to the water molecule

    located close to the Q-phosphate moiety of ATP inthe high-resolution structure. The distance between

    the water molecule and side-chain oxygen is 2.8 A

    [8]. Therefore, this residue could activate the water

    molecule when the enzyme hydrolyzes ATP.

    Although the K and L subunits have signicant

    sequence homology, the K-subunit residue corre-

    sponding to LGlu181 is glutamine, consistent with

    the absence of catalytic activity in the K subunit. In

    this regard, the P-loop lysine residue in the K subunit

    P-loop contributed to nucleotide binding, but its re-

    placement had only slight eects on catalysis [26,27].

    Mutant studies also suggested that the positive

    charge of LArg182 side chain is required for ATP

    binding [41]. The LArg182Gln enzyme showed a

    102-fold lower k1 (rate of ATP binding to F1) and

    unisite catalysis, while the LArg182Lys enzyme was

    similar to the wild type.

    From the studies briey summarized above, an

    overall scheme of ATP synthesis and hydrolysis

    (F16ATPHF16ADPcPi) at the catalytic site can

    be proposed, as shown in Fig. 4. For ATP synthesis,the L-phosphate of ADP and inorganic phosphate

    bind to A amino group of the LLys155 side chain

    and Mg through LThr156. The L-phosphate bound

    to Mg forms a phosphoester bond with Pi, releasing

    water a molecule together with the hydrogen of the

    carboxyl moiety ofLGly181. In the reverse reaction,

    the L- and Q-phosphate of ATP bind to LLys155 and

    Mg through LThr156, and the carboxyl moiety of

    LGlu181 activates the water molecule, which attacks

    PQ to break the P^O bond [42]. There is a consensus

    on the role ofL

    Glu181 [5,8]. TheL

    Glu185 and

    KArg376 are also at the catalytic site, but have roles

    in cooperativity rather than the elemental catalytic

    steps as discussed below.

    2.4. Cooperativity initiated from the catalytic site and

    K/L subunit interaction

    2.4.1. L subunit residues for cooperativity

    The F0F1 shows cooperativity, positive in catalysis

    Fig. 4. Basic chemical reaction at the catalytic site in F 1. Mod-

    els of ATP synthesis (a) and hydrolysis (b) at the catalytic site

    of ATP synthase are shown together with amino acid residues.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288280

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    6/13

    and negative in nucleotide binding, as described

    above. The cooperativity is a very complex phenom-

    ena to describe at the molecular level; it includes

    binding of nucleotides to the three catalytic sites,

    signal transmission between these sites, and theirconformation change, resulting in nucleotide release.

    The cooperativity also includes conformation trans-

    mission between the K and L subunits, rotation of Q,

    and interaction between the L and Q. It seems ob-

    vious that the site to site conformational transmis-

    sion is initiated from the changes of the amino acid

    side chain orientation in a single catalytic site. As

    expected, certain K and L subunit residues located

    at or near the catalytic site are found to be essential

    for cooperativity, rather than catalysis itself. Re-

    placements of these residues gave enzymes having

    unisite, but no multisite catalysis; they retain the

    basic chemical reaction at the catalytic site but are

    defective in site-to-site cooperativity. From this as-

    pect, LGlu185 is of special interest.

    The LGlu185 is the last residue of the conserved

    GERXXE sequence, which includes catalytic

    LGlu181 and LArg182 residues, as discussed above.

    Mutants with replacement of LGlu185, except

    LGlu185Asp, had no multisite catalysis, although

    they maintained substantial unisite catalysis [43].

    LGlu185Asp mutant could grow by oxidative phos-

    phorylation, and its F1 sector had 30^50% multisitecatalysis of the wild type, depending on Mg2 con-

    centration. The rate of multisite catalysis of the

    LCys185 enzyme was at least 106-fold lower than

    wild type, whereas the same enzyme treated with

    iodoacetate for S-carboxylmethylation had 1^30%

    of the wild-type activity, depending on Mg2 concen-

    tration. These results indicate that the carboxyl side

    chain at position 185 (glutamate, aspartate, or S-car-

    boxylmethyl cysteine) is required for catalytic coop-

    erativity, but its side chain length can be exible (3^6

    A ). The two mutant enzymes (aspartate and S-car-

    boxylmethyl cysteine) lost sensitivity to inhibition

    with high Mg2 concentrations, suggesting that the

    carboxyl side chain ofLGlu185 interact with Mg2 at

    the catalytic site. Consistent with this suggestion, the

    side chain oxygen of LGlu185 is about 4.0 A from

    Mg in the LDP or LTP subunit [8]. No further kinetic

    analysis of the LGlu185 mutant enzymes could be

    carried out, because nucleotide-bound mutant en-

    zymes could not be obtained. Other L subunit resi-

    dues, including LAsp242 and LArg246 are located at

    or near the catalytic site and may have some roles in

    catalytic cooperativity.

    2.4.2. K-Subunit residues for cooperativityOne requirement for catalytic cooperativity is a

    functional interaction between the K and L subunits.

    This has been conrmed by early mutagenesis

    around position K370 such as KSer373Phe and

    KArg376Cys [44^46]. In these cases, the mutant mul-

    tisite catalysis was three orders of magnitude lower

    than that of the wild type, but exhibited substantial

    unisite catalysis. These residues are suggested to be

    located at the subunit interface and participate in the

    L^K inter-subunit conformational transmission [45^

    47]. After the crystal structure became available,

    KArg376 was found to be close to the Q-phosphate

    moiety of ATP. The distance between the nearest

    oxygen atom of the ATP Q-phosphate and the nitro-

    gen of the guanidino moiety ofKArg376 is estimated

    to be 3.1 A [8] (Fig. 3). KArg376 was the impetus for

    further detailed kinetic studies by the authors. The

    rates of multisite catalysis of the K376 mutants were

    103^104-fold lower than wild type [48]. The three

    catalytic sites of these mutants retained negative co-

    operativity in nucleotide binding, with three binding

    constants similar to the wild type. The binding an-

    ity of ATP to the unisite was 10^20-fold lower, butthe basic catalytic rates (k2 and k32) for E6AT-

    PHE6ADP were not changed. These results indi-

    cate that the major defect of the mutant enzyme was

    the lack of stimulation of product release from the

    rst site by ATP binding to the second and third

    catalytic sites. These studies suggest that KArg376

    is required for the initial step of the transmission

    of conformation changes between catalytic sites, so

    that products are released rapidly.

    For catalytic cooperativity, the conformational

    changes initiated from a single catalytic site should

    be transmitted to other sites through interactions be-

    tween subunits. As each L subunit is located next to

    the K subunit in the hexagonal K3L3 arrangement,

    conformational change [5] of one L should be trans-

    mitted to another L through the adjacent K subunit.

    The requirement of the conformational transmission

    between the two subunits is supported by the mu-

    tants defective in multisite catalysis, mapped to the

    residues located at the interface between the K and L :

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288 281

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    7/13

    they are KPro281, KAla285, KArg296, KGlu299,

    KArg303, KAla306 and KArg376 (Fig. 5) [45^47].

    The conformational transmission between the K

    and L subunits, as a part of the catalytic steps,

    should be ultimately coupled to proton translocation.

    The functional interaction(s) of the two subunits forenergy coupling was indicated clearly by results in-

    dicating suppression of the defective L mutant by the

    second-site mutation in the K subunit. The LSer174-

    Phe mutant could not couple ATP hydrolysis to pro-

    ton transport, and synthesize ATP only slowly

    [49,50]. This mutation was suppressed by the second

    mutation in the K subunit, KArg296Cys. The

    KCys296/LPhe174 double mutant had essentially the

    same ATPase activity as LSer174Phe single mutant,

    but restored proton transport and ATP synthesis

    [51]. The conformational transmission between cata-

    lytic sites should be also coupled to the rotation of

    the Q subunit, as discussed below.

    3. The QQ subunit is required for catalysis and energy

    coupling

    3.1. Roles of Qsubunit in multisite catalysis and

    energy coupling

    3.1.1. Q-Subunit residues for catalysis and coupling

    The amino and carboxyl terminal K-helices of the Q

    subunit are positioned in the central space of the

    K3L3 hexamer, according to the X-ray structure [8],

    and the amino acid sequences of the helical regions

    are highly conserved among dierent species [52].

    Consistent with the location at the center of the

    K3L3 assembly, early studies indicated that the Q sub-

    unit is required for the in vitro reconstitution of the

    minimal catalytic K3L3Q assembly [53] or in vivo as-

    sembly of the entire ATP synthase [54]. Retention of

    ATPase activity in K3L3 assembly is possible only

    with the thermophilic bacterial enzyme [55]. Further-

    more, the entire enzyme assembly was lost by Qmu-

    tations such as QGln261End (QGln261CEnd) [56,57]

    and a deletion between QLys21 and QAla28 [58].

    We were interested in introducing mutations in the

    region between QGln261 and the carboxyl terminus

    before the high-resolution structure became available

    (Fig. 6) [57]. Multisite catalysis was inhibited by 10^

    Fig. 5. K-Subunit residues important for catalytic cooperativity. Amino acid residues important for catalytic cooperativity are shown

    at the interface between K and L subunits. Replacement of these residues reduce catalytic cooperativity.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288282

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    8/13

    80%, depending on the site of mutations, indicating

    that the Q subunit has an important role, mainly for

    catalytic cooperativity. The mutants attracted our

    special interests were those having the same reduced

    ATPase activities, but which formed ATP hydrolysis-

    dependent electrochemical proton gradients of dier-

    ent magnitudes. The QGlu269Leu mutant had AT-

    Pase activity similar to QGlu275Lys and QThr277End,

    but formed a much weaker vWH [57]. The low

    vWH formation was possibly due to the defective

    energy coupling between catalysis and proton trans-

    port, giving the rst indication that the Q subunit

    participates in energy coupling. Inspired by these re-

    sults, we replaced conserved residues in the amino

    terminal region systematically and studied catalysis

    and energy coupling of the mutant enzymes [59].

    3.1.2. An energy-coupling mutation QMet23Lys

    Most of the mutations in the amino terminal re-

    gion of the Q subunit had little eect on growth by

    oxidative phosphorylation, ATPase activity, or pro-

    ton pumping [59]. The notable exception was ob-

    tained from substitutions of the conserved QMet23

    residue. The QMet23Arg and QMet23Lys mutants

    could not grow by oxidative phosphorylation. Their

    membrane vesicles had essentially the same ATPase

    activities as the wild type, but formed much lower

    ATP-dependent electrochemical proton gradients [59]

    and had reduced ATP synthesis [60], relative to the

    wild type. Other substitutions such as QMet23Asp,

    QMet23Glu and QMet23Leu were similar to the wild

    type, indicating that QMet23 is not an essential resi-

    due, but introduction of the positively-charged side

    chain reduced energy coupling between catalysis and

    proton transport. Al-Shawi and coworkers studied

    the QMet23Lys mutation extensively [60^62] and sug-

    gested that the side chain of the mutant lysine resi-

    due forms an ionized hydrogen bond with the car-

    Fig. 6. Mutational studies of Q subunit. The Q-subunit mutations for catalysis and energy coupling are summarized in the high-resolu-

    tion structure. See text and [63,64] for details.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288 283

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    9/13

    boxyl side chain of LGlu381 of the DELSEED seg-

    ment (LAsp380^LAsp386) of the LDP subunit. From

    the studies of ATP hydrolysis by F1 and F0F1, they

    suggested that the mutation perturbed the functional

    interaction of F0 and F1, which is mediated by the Qand A subunits [60].

    The defective energy coupling of QMet23Lys was

    suppressed by a series of second-site mutations in the

    carboxyl terminal region [63]. The results of muta-

    tion/suppression studies predicted that QMet23,

    QArg242 and the region between QGlu269 and

    QVal280 are close to each other and interact to medi-

    ate energy coupling. Similarly, QGlu269Glu or

    QThr273Val mutation was suppressed by replacement

    of amino terminal residues at position 18, 34, and 35

    [64]. The suppression of QMet23Lys by QArg242Cys

    could be explained by the altered interaction between

    DELSEED and the replaced residues, because

    QArg242 and QMet23, in the carboxyl and amino ter-

    minal K-helices, respectively, are located in close vi-

    cinity in the crystal structure [8]. A more interesting

    observation was that the QMet23Lys mutation was

    suppressed by the mutations between QGlu269 and

    QVal280, which does not interact directly with the

    QMet23 residue or those located at the domain near-

    by. Furthermore, the second mutations included a

    variety of substitutions, and could not be grouped

    into a single type. These results suggest that thetwo K-helices of the Q subunit interact closely and

    form a domain favorable for long range conforma-

    tion transmission, leading to the mechanical rotation

    of the Q subunit as discussed below.

    3.2. Interactions between L and Qsubunits

    Successive interactions of the Qsubunit helices with

    the three L subunits seem to be essential for catalytic

    cooperativity and energy coupling during ATP syn-

    thesis or hydrolysis. Such interactions have been

    shown both by biochemical and genetic approaches.

    The enzyme with a Q-subunit frameshift mutation

    having 16 unrelated carboxyl terminal residues

    showed no ATP synthesis and hydrolysis [57]. Due

    to a nucleotide deletion from QGlu278 codon, the

    Q-subunit frameshift mutant has seven additional res-

    idues at its carboxyl terminus, together with nine

    altered residues downstream of QThr277. The longer

    carboxyl terminus of the frameshift may interact

    with the upper L-barrel or a part of the nucleotide

    binding domain of the L subunit [8], and cause de-

    fective conformation transmission between the L and

    Q subunits.The second-site mutation of the L subunit,

    LArg52Cys or LGly150Asp, could suppress the dele-

    terious eect(s) of the frameshift mutation [65]. The

    LArg52Cys mutation suppressed the frameshift, pos-

    sibly by restoring the deleterious interactions of the

    16 unrelated carboxyl terminal regions with the

    L-barrel domain, where LArg52 is located. The defec-

    tive energy coupling of the frameshift was also sup-

    pressed by the LGly150Asp mutation in the P-loop.

    This loop may show a large conformation change

    during catalysis, because its structure in LE and

    LTP or LDP are strikingly dierent [8]. The

    LGly150Asp mutation may change the structure of

    this loop and aect orientation of another loop

    (LAsp301^LPro306), located above the P-loop. The

    P-loop residue, LGly150, is close to LAsp301; dis-

    tance between the CK ofLGly150 and the side chain

    oxygen of LAsp301 is 3.3 A in the crystal structure

    [8]. When LGly150 is replaced by Asp, the mutant

    side chain collides with that of LAsp301 and aects

    the entire conformation of the loop (LAsp301^

    LPro306). As described below, this loop interacts

    with the Qsubunit around QGln269. These mutationalalterations may change the mode of conformational

    transmission between the catalytic site in the L sub-

    unit and the frameshift Q subunit. The importance of

    the interaction between the P-loop and the Q subunit

    helices are also indicated by the cross-linking studies.

    The region between LVal145-LLys155 (P-loop, posi-

    tions 149^155) of the LDP subunit is shown to inter-

    act with QSer8 [66].

    The L subunit DELSEED loop domain (LAsp380^

    LAsp386) has been shown to interact with the Q sub-

    unit aroundQ

    Cys87, one of the two cysteines of the

    Qsubunit, in the short K helix [8]. This QCys87 can be

    cross-linked to a Cys residue, replacing LGlu381, in-

    dicating that the two residues are located nearby

    (approximately 4 A apart) [67]. QCys87 is shielded

    to various maleimides, but became reactive when

    the LGlu381 was replaced by a smaller side chain

    [68]. Furthermore, the reactivity of QCys87 was nu-

    cleotide-dependent, showing no reactivity in the pres-

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288284

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    10/13

    ence of ADP+Mg. These results indicate L/Q interac-

    tion in this region is important for catalysis or energy

    coupling.

    The region around QGln269 is close to the L sub-

    unit loop between LAsp301 and LPro306(DDLTDP), located above the P-loop (LGly150 is

    located close to LAsp301). The bovine structure sug-

    gests that LAsp302 and LThr304 form hydrogen

    bonds with QGln269 [8]. Mutations in the loop be-

    tween LAsp301 and LPro306 yielded variable eects

    on F1ATPase activity, stability and binding to F0[68]. Replacement of QGln269 had a similar eect.

    LGln269Glu and LGln269Asp F1 became dissociated

    into subunits once they were solubilized. Another

    loop (LAla256^LThr270) can be found near the top

    of the L subunit between L-barrel and nucleotide

    binding domain, and close to the Q subunit residues

    between QIle272 and QVal286. However, the interac-

    tion of this loop with the Q subunit is not so impor-

    tant because the Q subunit lacking this region

    (QThr277End) is still active in ATP synthesis and hy-

    drolysis [57].

    3.3. Q-Subunit rotation

    3.3.1. Q-Subunit rotation and torque generation during

    ATP hydrolysis

    The binding-change mechanism predicted confor-mation transmission among three catalytic sites, via

    Q-subunit rotation. This rotation was initially hy-

    pothesized, and later conrmed experimentally [5].

    The rotation is required for catalytic cooperativity

    and energy coupling during ATP synthesis or hydro-

    lysis. The high-resolution structure indicates the pres-

    ence ofLTP, LDP, and LE, the three L subunits with

    dierent catalytic site conformations. Furthermore,

    they demonstrate diering interactions with the cen-

    tral Q-subunit helices [8]. This structure clearly sug-

    gests the rotation of theQ

    subunit during catalysis, so

    that the Q subunit can interact alternately with the

    three L subunits.

    The Q-subunit rotation has been shown by a vari-

    ety of approaches. The observations suggesting the

    rotation come from following studies: cryoelectron

    microscopy of F1 [13,14], L/Q subunit cross-linking

    [15,16], and analysis of polarized absorption recovery

    after photo bleaching of the probe (eosin dye) linked

    to the Q subunit of chloroplasts [17,18]. It may be

    appropriate to briey summarize the cross-linking

    approach here because the L/Q interaction was uti-

    lized. Duncan and coworkers observed a specic di-

    sulde bond formation ofQCys87 with a cysteine res-

    idue in the LAsp380Cys (LAsp380 in DELSEED)mutation upon oxidation. Following reduction of

    the disulde bond, ATP hydrolysis or synthesis was

    carried out, and the disulde bond was formed by

    oxidation. They indicated a reorientation of QCys87

    relative to the three L subunits during catalysis, sup-

    porting the Q-subunit rotation [15,16].

    Finally, the rotation of the Q subunit was directly

    observed and video-recorded by Noji and coworkers

    [19]. The thermophilic bacterial K3L3Q complex was

    attached to a glass surface and the uorescently la-

    beled actin lament was attached to the Q subunit.

    The rotation of the actin lament was dependent on

    ATP hydrolysis, anti-clockwise viewed from the

    membrane side and having 120 steps. The rotation

    became slower with the increase of the lament

    length. The titration curve between the lament

    lengths and rotational rates indicates that the rota-

    tion generated a constant frictional torque ofV40

    pNcnm, although it was dicult to determine the

    precise value because of the scatter of experimental

    points. Free energy of ATP hydrolysis (vGATP) is

    V80 pNcnm under physiological conditions and

    comparable with a value of the torque times 2Z/3(50 pNcnm 2Z/3; work done in one third of revolu-

    tion) [69]. These results indicate that the thermody-

    namic eciency is close to 100%. Similar approaches

    showed the A-subunit rotation with the Qsubunit [70].

    We have recently shown that the Q subunit in the

    E. coli F1 sector could also rotate utilizing the energy

    of ATP hydrolysis [20]. Upon addition of ATP, con-

    tinuous anti-clockwise rotation of the lament was

    observed. Furthermore, the titration of the lament

    lengths against rotational rates gave essentially the

    same results as those of the thermophilic bacterial

    enzyme [69]. These results established that ATP syn-

    thesis or hydrolysis by F0F1 is a combination of cat-

    alysis, mechanical work (Q-subunit rotation and tor-

    que generation) and proton transport.

    3.3.2. Rotation and energy coupling

    It became possible to analyze the rotational catal-

    ysis in detail, utilizing the wealth of information ob-

    tained from E. coli mutants. The calculated titration

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288 285

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    11/13

    curve between rotational rates and actin lament

    lengths ts with the experimental points [20] when

    the lament length is suciently long; the rate of

    2 Wm lament calculated assuming the constant torque

    of 40 pNcnm is 0.92 s

    31

    , while experimental valuesare 0.97^1.7 s31. However, the same calculation gives

    2800 s31 with the 0.1 Wm lament. This value is

    impossible to obtain experimentally because the rota-

    tional rates should not exceed ATPase turnover

    (V60 s31). Obviously, the rotational rates become

    independent of torque generation with the lighter

    load. Thus, it may be possible that a certain un-

    coupled mutant subunit rotates similar to the wild

    type but cannot generate enough torque to drive

    proton transport, or the Q-subunit rotation of other

    uncoupled mutants may not be strictly obligatory for

    ATPase activity. Some mutations may even cause

    slipping of the Q subunit. Such a mutant enzyme

    has ATPase activity but cannot rotate Q because of

    the slip. In this regard, the thermophilic K3L3 assem-

    bly lacking the Q subunit had about 25% ATPase

    activity of K3L3Q complex [55].

    Thus, it is interesting to study the Q rotation in F1sectors with mutations-induced defective energy cou-

    pling between catalysis and proton transport. Ana-

    lyzing these mutants may indicate the role(s) of the

    rotation in the mechanism of ATP hydrolysis or syn-

    thesis. We were interested in the uncoupled mutationQMet23Lys [59]. The mutation caused serious defects

    on energy coupling of the F1 engineered for rotation

    studies, similar to the original mutant. The mutant

    enzyme had wild-type ATPase activity, but showed

    no ATP-dependent proton transport. However, the

    QMet23Lys mutant Q subunit rotated similar to the

    wild type. This mutation had no eect on the rota-

    tional rate and the generation of frictional torque,

    indicating that defective energy coupling of the

    QMet23Lys enzyme is not directly related to the

    Q-subunit rotation in F

    1[20]. Thus, the defect of

    Q-subunit Met23 mutant is after Q rotation, possibly

    at the step where the rotation couples to proton

    transport. For ATP synthesis, conformation change

    of the c subunit oligomer, possibly its rotation,

    should be transmitted (or transformed) to the Q-sub-

    unit rotation. The QMet23Lys enzyme may be defec-

    tive in such interactions between F1 and F0 sectors.

    Rotation of the Qsubunit in ATP synthesis by proton

    transport is suggestive from the present data, but it

    should be noted that the mechanism is still specula-

    tive.

    For the rotation of the Q subunit, the K3L3 assem-

    bly should be xed with F0 in membranes through

    the non-rotating subunits, and the Q-subunit rota-tions should be transmitted to the F0 c subunit olig-

    omer [5,12]. The stator structure may be formed

    from the two b subunits interacting with the K and

    the a subunit of F0. The N subunit is interacting with

    K, while A is rotating with the Q subunit. The F1F0with two stalks, possibly stator and rotor structures,

    was observed recently by electron microscopy [71].

    4. Summary and perspectives

    ATP synthase is a unique enzyme which couples

    chemical reaction, mechanical work, and transport of

    protons (Fig. 1). Most of the catalytic residues were

    identied [1^5] before the high-resolution structure

    was obtained by X-ray crystallography [8]. However,

    it was dicult to interpret the results unambiguously

    without the high-resolution structure. The bovine

    catalytic site structure and kinetics of E. coli mutant

    enzymes obviously deepened our understanding of

    the mechanism of ATP synthesis and hydrolysis.

    The catalytic residues identied are conserved in

    the A subunit of V-type ATPase [39,40], suggestingthat the two ATPases have the same mechanism.

    The rotation of the Q subunit in the K3L3Qcomplex

    has been established by biochemical studies [12^18]

    and physical observation of actin lament rotation

    attached to the Q subunit [19,20]. The ATPase ac-

    tually generated frictional torque with high thermo-

    dynamic eciency. In ATP hydrolysis, the c^Q^A as-

    sembly, extending through F0F1, rotates to complete

    ATP-dependent proton transport. It is of interest to

    know whether the Q-subunit rotation observed in im-

    mobilized F1

    (K3L3Q

    ) is a part of the rotation of the

    c^Q^A assembly, or if the torque generated by the

    Q subunit is transformed to the rotation of the c sub-

    unit oligomer. Such a basic question on coupling

    between catalysis, mechanical work, and proton

    transport can be answered at least partly from stud-

    ies of uncoupled mutants. In this regard, we observed

    that an actin lament connected to the c subunit of

    F0F1 could rotate using the energy of ATP hydroly-

    sis [73].

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288286

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    12/13

    We predicted that the uncoupled mutations gener-

    ate less torque than wild type or show no Q-subunit

    rotation. However, we observed that the rotation of

    the Q subunit of uncoupled QMet23Lys mutant could

    rotate and generate essentially the same degree offrictional torque as that of the wild type. Thus the

    defect of energy coupling in QMet23Lys mutation

    takes place after the Qrotation and torque generation

    step, possibly at the step where rotation drives pro-

    ton transport. It may be reasonable to assume that

    the rotation of the Qsubunit in the F1 sector is trans-

    mitted to the rotation of the c subunit oligomer in

    the entire F0F1. We can still expect other mutations

    with low torque generation and/or no Q rotation.

    These mutations may indicate the mechanism of

    the rotation. We believe that E. coli is still extremely

    useful for studying the energy coupling during ATP

    synthesis.

    Acknowledgements

    Studies of our laboratory cited in this article are

    supported by Japanese Ministry of Education, Sci-

    ence and Culture and Japan Science and Technology

    Corporation. We are grateful to Dr. Ashley Spies for

    the critical reading of the manuscript.

    References

    [1] M. Futai, T. Noumi, M. Maeda, Annu. Rev. Biochem. 58

    (1989) 111^136.

    [2] M. Futai, H. Omote, in : W.N. Konings, H.R. Kaback, J.S.

    Lolkema (Eds.), Handbook of Biological Physics vol. 2,

    Elsevier, Amsterdam, 1996, pp. 47^74.

    [3] J. Weber, A.E. Senior, Biochim. Biophys. Acta 1319 (1997)

    19^58.

    [4] R.H. Fillingame, Curr. Opin. Struct. Biol. 6 (1996) 491^498.

    [5] P.D. Boyer, Annu. Rev. Biochem. 66 (1997) 717^749.

    [6] H.S. Penefsky, R.L. Cross, Adv. Enzymol. Relat. Areas

    Mol. Biol. 64 (1991) 173^214.

    [7] R.K. Nakamoto, J. Membr. Biol. 151 (1996) 101^111.

    [8] J.P. Abrahams, A.G.W. Leslie, R. Lutter, J.E. Walker, Na-

    ture 370 (1994) 621^628.

    [9] R. Birkenhager, M. Hoppert, G. Deckers-Hebestreit, F.

    Mayer, A. Altendorf, Eur. J. Biochem. 230 (1995) 58^67.

    [10] S. Singh, P. Turiana, C.J. Bustamante, D.J. Keller, R.A.

    Capaldi, FEBS Lett. 397 (1996) 30^34.

    [11] K. Takeyasu, H. Omote, S. Nettikadan, F. Tokumasu, A.

    Iwamoto, M. Futai, FEBS Lett. 392 (1996) 110^113.

    [12] W. Junge, H. Lill, S. Engelbrecht, Trends Biochem. Sci. 22

    (1997) 420^423.

    [13] E.P. Gogol, E. Johnston, R. Aggeler, R.A. Capaldi, Proc.

    Natl. Acad. Sci. USA 87 (1990) 9585^9589.

    [14] R. Aggeler, M.A. Haughton, R.A. Capaldi, J. Biol. Chem.

    270 (1995) 9185^9191.[15] T.M. Duncan, V.V. Bulygin, Y. Zhou, M.L. Hutcheon, R.L.

    Cross, Proc. Natl. Acad. Sci. USA 92 (1995) 10964^

    10968.

    [16] Y. Zhou, T.M. Duncan, R.L. Cross, Proc. Natl. Acad. Sci.

    USA 94 (1997) 10583^10587.

    [17] D. Sabbert, S. Engelbrecht, W. Junge, Nature 381 (1996)

    623^625.

    [18] D. Sabbert, S. Engelbrecht, W. Junge, Proc. Natl. Acad. Sci.

    USA 94 (1997) 4401^4405.

    [19] H. Noji, R. Yasuda, M. Yoshida, K. Kinoshita Jr., Nature

    386 (1997) 299^302.

    [20] H. Omote, N. Sanbonmatsu, K. Saito, Y. Sambongi, A.

    Iwamoto-Kihara, T. Yanagida, Y. Wada, M. Futai, Proc.

    Natl. Acad. Sci. USA 96 (1999) 7780^7784.

    [21] Y. Shirakihara, A.G.W. Leslie, J.P. Abrahams, J.E. Walker,

    T. Ueda, Y. Sekimoto, M. Kambara, K. Saika, Y. Kagawa,

    M. Yoshida, Structure 5 (1997) 825^836.

    [22] P.D. Boyer, in : C.P. Lee, G. Shatz, L. Ernster (Eds.), Mem-

    brane Bioenergetics, Addison-Wesley, Reading, MA, 1979,

    pp. 461^479.

    [23] C. Grubmeyer, R.L. Cross, H.S. Penefsky, J. Biol. Chem.

    257 (1982) 12092^12100.

    [24] R.L. Cross, C. Grubmeyer, H.S. Penefsky, J. Biol. Chem.

    257 (1982) 12101^12105.

    [25] B.D. Reynafarje, P.L. Pedersen, J. Biol. Chem. 271 (1996)

    32546^32550.

    [26] M. Jounouchi, M. Maeda, M. Futai, J. Biochem. 114 (1993)171^176.

    [27] R. Rao, J. Pagan, A.E. Senior, J. Biol. Chem. 263 (1988)

    15957^15963.

    [28] R.L. Cross, D. Cunningham, C.G. Miller, Z.X. Xue, J.-M.

    Zhou, P.D. Boyer, Proc. Natl. Acad. Sci. USA 84 (1987)

    5715^5719.

    [29] J. Weber, S. Wilke-Mounts, R.S. Lee, A.E. Senior, J. Biol.

    Chem. 268 (1993) 20126^20133.

    [30] J. Weber, S. Wilke-Mounts, A.E. Senior, J. Biol. Chem. 269

    (1994) 20462^20467.

    [31] S. Lo bau, J. Weber, A.E. Senior, FEBS Lett. 404 (1997) 15^

    18.

    [32] S.-Y. Hsu, T. Noumi, M. Takeyama, M. Maeda, S. Ishiba-

    shi, M. Futai, FEBS Lett. 218 (1987) 222^226.

    [33] M. Takeyama, K. Ihara, Y. Moriyama, T. Noumi, K. Ida,

    A. Tomioka, A. Itai, M. Maeda, M. Futai, J. Biol. Chem.

    265 (1990) 21279^21284.

    [34] H. Omote, M. Maeda, M. Futai, J. Biol. Chem. 267 (1992)

    20571^20576.

    [35] A.E. Senior, M.K. Al-Showi, J. Biol. Chem. 267 (1992)

    21471^21478.

    [36] A.E. Senior, S. Wilke-Mounts, M.K. Al-Shawi, J. Biol.

    Chem. 268 (1993) 6989^6994.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288 287

  • 7/31/2019 ATP Synthase Coupling Between Catalysis Mechanical Work and Proton Translocation

    13/13

    [37] K. Ida, T. Noumi, M. Maeda, M. Futai, J. Biol. Chem. 266

    (1991) 5424^5429.

    [38] S. Lo bau, J. Weber, S. Wilke-Mounts, A.E. Senior, J. Biol.

    Chem. 272 (1997) 3648^3656.

    [39] K. Inatomi, S. Eya, M. Maeda, M. Futai, J. Biol Chem. 264

    (1989) 10954^10959.[40] M. Futai, H. Omote, in: S. Papa, F. Gueieri, J.M. Tager

    (Eds.), Frontiers of Cellular Bioenergetics: Molecular Biol-

    ogy, Biochemistry and Pathology, Plenum, London, 1999,

    pp. 399^421.

    [41] M.-Y. Park, H. Omote, M. Maeda, M. Futai, J. Biochem.

    116 (1994) 1139^1145.

    [42] M. Futai, H. Omote, M. Maeda, Biochem. Soc. Trans. 23

    (1995) 785^789.

    [43] H. Omote, N.P. Le, M.-Y. Park, M. Maeda, M. Futai,

    J. Biol. Chem. 270 (1995) 25656^25660.

    [44] T. Noumi, M. Futai, H. Kanazawa, J. Biol. Chem. 259

    (1984) 10076^10079.

    [45] S. Soga, T. Noumi, M. Takeyama, M. Maeda, M. Futai,

    Arch. Biochem. Biophys. 268 (1989) 643^648.

    [46] M.B. Maggio, J. Pagan, D. Personage, L. Hatch, A.E. Se-

    nior, J. Biol. Chem. 262 (1987) 8981^8984.

    [47] J. Pagan, A.E. Senior, Arch. Biochem. Biophys. 277 (1990)

    283^289.

    [48] N.P. Le, H. Omote, M. Al-Shawi, R.K. Nakamoto, Y.

    Wada, M. Futai, Biochemistry (2000) in press.

    [49] A. Iwamoto, H. Omote, H. Hanada, N. Tomioka, A. Itai,

    M. Maeda, M. Fuati, J. Biol. Chem. 266 (1991) 16350^

    16355.

    [50] A. Iwamoto, M.-Y. Park, M. Maeda, M. Futai, J. Biol.

    Chem. 269 (1994) 10265^10269.

    [51] H. Omote, M.-Y. Park, M. Maeda, M. Futai, J. Biol. Chem.

    268 (1993) 3156^3160.[52] R.K. Nakamoto, K. Shin, A. Iwamoto, H. Omote, M. Mae-

    da, M. Futai, Ann. NY Acad. Sci. 671 (1992) 335^344.

    [53] M. Futai, Biochem. Biophys. Res. Commun. 79 (1977) 1231^

    1237.

    [54] J. Miki, M. Maeda, M. Futai, J. Bacteriol. 170 (1988) 179^

    183.

    [55] Y. Kagawa, S. Ohta, Y. Otawara-Hamamoto, FEBS Lett.

    249 (1989) 67^69.

    [56] J. Miki, M. Takayama, T. Noumi, H. Kanazawa, M. Mae-

    da, M. Futai, Arch. Biochem. Biophys. 251 (1986) 458^464.

    [57] A. Iwamoto, J. Miki, M. Maeda, M. Futai, J. Biol. Chem.

    265 (1990) 5043^5048.[58] H. Kanazawa, H. Hama, B.P. Rosen, M. Futai, Arch. Bio-

    chem. Biophys. 241 (1985) 364^370.

    [59] K. Shin, R.K. Nakamoto, M. Maeda, M. Futai, J. Biol.

    Chem. 267 (1992) 20835^20839.

    [60] M.K. Al-Shawi, C.J. Ketchum, R.K. Nakamoto, J. Biol.

    Chem. 272 (1997) 2300^2306.

    [61] M.K. Al-Shawi, R.K. Nakamoto, Biochemistry 36 (1997)

    12954^12960.

    [62] M.K. Al-Shawi, C.J. Ketchum, R.K. Nakamoto, Biochem-

    istry 36 (1997) 12961^12969.

    [63] R.K. Nakamoto, M. Maeda, M. Futai, J. Biol. Chem. 268

    (1993) 867^872.

    [64] R.K. Nakamoto, M. Al-Shawi, M. Futai, J. Biol. Chem. 270

    (1995) 14042^14046.

    [65] C. Jeanteur-De Beukelar, H. Omote, A. Iwamoto-Kihara,

    M. Maeda, M. Futai, J. Biol. Chem. 270 (1995) 22850^

    22854.

    [66] R. Aggeler, S.X. Cai, J.F.W. Kaena, T. Koike, R.K. Capa-

    ldi, J. Biol. Chem. 268 (1993) 20831^20837.

    [67] Z. Feng, R. Aggeler, M.A. Haughton, R.A. Capaldi, J. Biol.

    Chem. 271 (1996) 17986^17989.

    [68] H. Omote, K. Tainaka, K. Fujie, A. Iwamoto-Kihara, Y.

    Wada, M. Futai, Arch. Biochem. Biophys. 358 (1998) 277^

    282.

    [69] R. Yasuda, H. Noji, K. Kinoshita Jr., M. Yoshida, Cell 93

    (1998) 1117^1124.

    [70] Y. Kato-Yamada, H. Noji, R. Yasuda, K. Kinoshita Jr., M.Yoshida, J. Biol. Chem. 273 (1998) 19375^19377.

    [71] S. Wilkens, R.A. Capaldi, Nature 393 (1998) 29.

    [72] P.J. Kraulis, J. Appl. Cryst. 24 (1991) 946 950.

    [73] Y. Sambongi, Y. Iko, M. Tanabe, H. Omote, A. Iwamoto-

    Kihara, I. Ueda, T. Yanagda, Y. Wada, M. Futa, Science

    286 (1999) 1722^1724.

    M. Futai et al. / Biochimica et Biophysica Acta 1458 (2000) 276^288288