an atomic-level mechanism for molybdenum nitrogenase. part 2. proton reduction, inhibition of...

10
An Atomic-Level Mechanism for Molybdenum Nitrogenase. Part 2. Proton Reduction, Inhibition of Dinitrogen Reduction by Dihydrogen, and the HD Formation Reaction ² Marcus C. Durrant* Computational Biology Group, John Innes Centre, Norwich Research Park, Colney Lane, Norwich NR4 7UH, U.K. ReceiVed February 4, 2002; ReVised Manuscript ReceiVed August 6, 2002 ABSTRACT: Quantum calculations have been used to examine the energetics of the reactions of diazene and isodiazene with H 2 and the properties of the Fe and Mo sites of the nitrogenase iron-molybdenum cofactor with respect to the binding of H and H 2 . The results have been used to extend the model for N 2 reduction by nitrogenase given in the preceding paper to describe the formation of HD from D 2 . The proposed mechanism for HD formation invokes a combination of two well-established chemical reactions, namely, competitive protonation of metal N 2 species at either the metal or at N 2 , followed by scrambling of D 2 at a metal hydride. The model is evaluated against the available biochemical data for the nitrogenase HD formation reaction and extended to account for H 2 inhibition of N 2 reduction and the reduction of H + in the absence of other substrates. In the preceding paper, a detailed mechanism for the reduction of dinitrogen by Mo nitrogenase was proposed on the basis of theoretical calculations, published biochemical data, and established chemical precedents. In addition to their essential biological function in the reduction of dinitrogen to ammonia, all known nitrogenases catalyze the production of dihydrogen from protons. Under turnover conditions but in the absence of any other reducible substrate, nitrogenase catalyses the reduction of H + to H 2 , a process which has been referred to as general hydrogen evolution (1). When N 2 is introduced, the rate of H 2 evolution decreases until a limiting ratio of N 2 reduced to H 2 produced is reached; this nonsuppressible H 2 evolution is known as obligatory hydro- gen evolution (OHE) 1 (1). The exact N 2 :H 2 ratio varies with different nitrogenases; for the Mo enzyme, the accepted limiting H 2 /N 2 ratio is 1(2), but for V nitrogenase, it is 3.5 (3). Several other facets of the chemistry of Mo nitrogenase with protons, H 2 , and N 2 have been revealed by experiment. Thus, H 2 is a competitive inhibitor for the reduction of N 2 , but of no other substrate (4, 5). One of the most intriguing properties of the enzyme is its ability to produce HD from D 2 , the so-called HD formation reaction (2, 4, 5). Although hydrogenases also produce HD when operating under an atmosphere of D 2 , the nitrogenase- catalyzed reaction has a number of striking differences. Most notably, HD formation by nitrogenase requires the presence of N 2 as well as D 2 (6). Since the steady-state kinetics of HD production and of H 2 inhibition of N 2 reduction by nitrogenase were found to be the same, it has generally been assumed that these two phenomena are different manifesta- tions of the same underlying molecular process (2). Recent experimental studies on the R-Asn 195 mutant nitrogenase from A. Vinelandii have qualified this view (5). This nitrogenase binds N 2 , apparently in the same way as the wild-type enzyme, but does not reduce it. H 2 was found to inhibit N 2 binding, but the mutant gave no HD when turning over under aD 2 /N 2 atmosphere. Hence, H 2 inhibition of N 2 binding (distinguished from N 2 reduction) appears not to be directly related to HD formation. The authors suggested that initial binding of N 2 at the FeMoco suffers from direct competition by H 2 , whereas a reduced form of N 2 may be required for HD formation. The idea that a reduced N 2 intermediate is somehow involved in HD formation is consistent with the observations that one electron is consumed for each molecule of HD produced and that these electrons are diverted exclusively from N 2 reduction, i.e., OHE is unaffected by HD formation (7). Another key difference in HD formation by nitrogenase compared to hydrogenase concerns the exchange of H + /D + with the solvent. Hydrogenases generally react with D 2 in water to give HD, H 2 , and HOD. This has been interpreted in terms of the reversible heterolytic cleavage of D 2 to give a metal hydride and a protonated amino acid residue (8- 12). Both of the D atoms can then exchange with H 2 O, but the hydride exchanges more slowly than the protonated amino acid (8-10). In contrast, tritium labeling experiments on nitrogenase showed that the rate of incorporation of T + into the aqueous phase was much slower than the rate of HD formation under the same conditions (13). This suggests that unlike the hydrogenase-catalyzed reaction, cleavage of D 2 by nitrogenase does not produce readily exchangeable protons. ² Support was provided by the Biotechnology and Biological Sciences Research Council, U.K. * Telephone: +44 (0)1603 450704. Fax: +44 (0)1603 450021. E-mail: [email protected] 1 Abbreviations: Av1, Kp1, molybdenum nitrogenases from A. Vinelandii and K. pneumoniae, respectively; BE, binding energy; DFT, density functional theory; FeMoco, iron-molybdenum cofactor; MP2, MP4, second- and fourth-order Møller-Plesset perturbation theory, respectively; OHE, obligatory hydrogen evolution; zpe, zero-point energy. 13946 Biochemistry 2002, 41, 13946-13955 10.1021/bi025624r CCC: $22.00 © 2002 American Chemical Society Published on Web 10/30/2002

Upload: marcus-c

Post on 13-Feb-2017

214 views

Category:

Documents


0 download

TRANSCRIPT

An Atomic-Level Mechanism for Molybdenum Nitrogenase. Part 2. ProtonReduction, Inhibition of Dinitrogen Reduction by Dihydrogen, and the HD

Formation Reaction†

Marcus C. Durrant*

Computational Biology Group, John Innes Centre, Norwich Research Park, Colney Lane, Norwich NR4 7UH, U.K.

ReceiVed February 4, 2002; ReVised Manuscript ReceiVed August 6, 2002

ABSTRACT: Quantum calculations have been used to examine the energetics of the reactions of diazeneand isodiazene with H2 and the properties of the Fe and Mo sites of the nitrogenase iron-molybdenumcofactor with respect to the binding of H and H2. The results have been used to extend the model for N2

reduction by nitrogenase given in the preceding paper to describe the formation of HD from D2. Theproposed mechanism for HD formation invokes a combination of two well-established chemical reactions,namely, competitive protonation of metal N2 species at either the metal or at N2, followed by scramblingof D2 at a metal hydride. The model is evaluated against the available biochemical data for the nitrogenaseHD formation reaction and extended to account for H2 inhibition of N2 reduction and the reduction of H+

in the absence of other substrates.

In the preceding paper, a detailed mechanism for thereduction of dinitrogen by Mo nitrogenase was proposed onthe basis of theoretical calculations, published biochemicaldata, and established chemical precedents. In addition to theiressential biological function in the reduction of dinitrogento ammonia, all known nitrogenases catalyze the productionof dihydrogen from protons. Under turnover conditions butin the absence of any other reducible substrate, nitrogenasecatalyses the reduction of H+ to H2, a process which hasbeen referred to as general hydrogen evolution (1). WhenN2 is introduced, the rate of H2 evolution decreases until alimiting ratio of N2 reduced to H2 produced is reached; thisnonsuppressible H2 evolution is known as obligatory hydro-gen evolution (OHE)1 (1). The exact N2:H2 ratio varies withdifferent nitrogenases; for the Mo enzyme, the acceptedlimiting H2/N2 ratio is ∼1 (2), but for V nitrogenase, it is3.5 (3). Several other facets of the chemistry of Monitrogenase with protons, H2, and N2 have been revealed byexperiment. Thus, H2 is a competitive inhibitor for thereduction of N2, but of no other substrate (4, 5). One of themost intriguing properties of the enzyme is its ability toproduce HD from D2, the so-called HD formation reaction(2, 4, 5). Although hydrogenases also produce HD whenoperating under an atmosphere of D2, the nitrogenase-catalyzed reaction has a number of striking differences. Mostnotably, HD formation by nitrogenase requires the presenceof N2 as well as D2 (6). Since the steady-state kinetics of

HD production and of H2 inhibition of N2 reduction bynitrogenase were found to be the same, it has generally beenassumed that these two phenomena are different manifesta-tions of the same underlying molecular process (2). Recentexperimental studies on theR-Asn195mutant nitrogenase fromA. Vinelandii have qualified this view (5). This nitrogenasebinds N2, apparently in the same way as the wild-typeenzyme, but does not reduce it. H2 was found to inhibit N2binding, but the mutant gave no HD when turning over undera D2/N2 atmosphere. Hence, H2 inhibition of N2 binding(distinguished from N2 reduction) appears not to be directlyrelated to HD formation. The authors suggested that initialbinding of N2 at the FeMoco suffers from direct competitionby H2, whereas a reduced form of N2 may be required forHD formation. The idea that a reduced N2 intermediate issomehow involved in HD formation is consistent with theobservations that one electron is consumed for each moleculeof HD produced and that these electrons are divertedexclusively from N2 reduction, i.e., OHE is unaffected byHD formation (7).

Another key difference in HD formation by nitrogenasecompared to hydrogenase concerns the exchange of H+/D+

with the solvent. Hydrogenases generally react with D2 inwater to give HD, H2, and HOD. This has been interpretedin terms of the reversible heterolytic cleavage of D2 to givea metal hydride and a protonated amino acid residue (8-12). Both of the D atoms can then exchange with H2O, butthe hydride exchanges more slowly than the protonatedamino acid (8-10). In contrast, tritium labeling experimentson nitrogenase showed that the rate of incorporation of T+

into the aqueous phase was much slower than the rate ofHD formation under the same conditions (13). This suggeststhat unlike the hydrogenase-catalyzed reaction, cleavage ofD2 by nitrogenase does not produce readily exchangeableprotons.

† Support was provided by the Biotechnology and Biological SciencesResearch Council, U.K.

* Telephone: +44 (0)1603 450704. Fax:+44 (0)1603 450021.E-mail: [email protected]

1 Abbreviations: Av1, Kp1, molybdenum nitrogenases fromA.VinelandiiandK. pneumoniae, respectively; BE, binding energy; DFT,density functional theory; FeMoco, iron-molybdenum cofactor; MP2,MP4, second- and fourth-order Møller-Plesset perturbation theory,respectively; OHE, obligatory hydrogen evolution; zpe, zero-pointenergy.

13946 Biochemistry2002,41, 13946-13955

10.1021/bi025624r CCC: $22.00 © 2002 American Chemical SocietyPublished on Web 10/30/2002

A central point of controversy in models for the nitroge-nase HD formation reaction centers on the chemical natureof the species involved in this process. As noted above, somefeatures of the reaction are best explained in terms of HDformation mediated by a partially reduced N2 intermediate(5), such as diazene (13). However, the D-D bond is strong(106 kcal mol-1), and chemical systems for H2 activationmost commonly involve metal hydrides. This was interpretedby Lowe and Thorneley (2, 14) in terms of homolyticcleavage of D2 at the nitrogenase active site, followed byattack of H+ on the resulting dideuteride to give HD. Thisis consistent with known chemistry (1) and fits in with theLowe-Thorneley scheme for reduction of N2 (2, 14-17).In the Lowe-Thorneley scheme, N2 binding can only occurby the reversible displacement of H2; HD formation and H2inhibition of N2 reduction then arise from displacement ofN2 by D2 and H2, respectively. Although this model providesa chemically reasonable pathway for D-D bond cleavagevia metal deuterides, it does not invoke any reduced N2

intermediate; hence, it cannot easily be reconciled with thenew observation from theR-Asn195 mutant studies (5) thatbinding of N2 to the FeMoco is of itself not sufficient togive HD.

An important requirement for any atomic model fornitrogenase action is that it should describe not just N2

reduction, but also the H+, H2 and D2 chemistry of theenzyme. This paper deals with the extension of the modeldescribed in the preceding paper to accommodate thischemistry. A new model for the HD formation reaction isproposed on the basis of competition for protons betweenN2 and the FeMoco core. As noted above, some aspects ofthe HD formation reaction are best explained in terms of areduced N2 intermediate; others suggest that metal hydridesare required. Much of the discussion in the literature contraststhese two aspects of the reaction, which are generally viewedas mutually exclusive possibilities. A key feature of themodel presented in this paper is that it reconciles these twofacets of the reaction by invoking both a reduced N2 speciesand a metal hydride as intermediates.

MATERIALS AND METHODS

All quantum calculations were performed using GAUSS-IAN 98W (18). DFT calculations were carried out exactlyas described in the preceding paper. Calculations on thereactions of free N2H2 isomers with H2 were done followingthe methodology of McKee et al. (19). Full geometryoptimizations (calculated without any symmetry constraints)and frequency calculations were carried out at the MP2/6-31G* level. Final single-point calculations at the calculatedgeometries were done at the MP4/6-31+G(2d, p) level oftheory. Protein coordinates were retrieved from the Brook-haven Protein Data Bank (20), accession codes 1QGU (Kp1)and 2MIN (Av1). Figures were prepared using RASMOLmolecular graphics software (21).

RESULTS

Direct Reaction of N2H2 Species with H2. As mentionedin the Introduction, the nitrogenase HD formation reactionhas been widely interpreted in terms of the direct reactionof an N2H2 intermediate with D2. In principle, this couldinvolve either diazene or isodazene (Scheme 1), either bound

to the cofactor or having dissociated into solution. Freediazene is very reactive and, in the absence of other reagents,undergoes rapid disproportionation in solution to give N2

plus hydrazine. Note thattrans-diazene is the more stableisomer; however, the cis form required for disproportionationor reaction with H2 is only ∼6 kcal mol-1 higher in energy.The disproportionation reaction (Scheme 1a) was investigatedquantum mechanically by McKee et al. (19) using Møller-Plesset perturbation theory to account for the effects ofelectron correlation. The results were in very good agreementwith experiment and showed that there is essentially nopotential energy barrier for this reaction. The same meth-odology has been applied in the present study to thehypothetical reactions of free diazene and isodiazene withH2 (panels b and c, respectively, of Scheme 1). As expected,the diazene reaction b was calculated to be stronglyexothermic (MP4 energy-50.5 kcal mol-1, or -58.3including zpe’s). However, the transition state had a calcu-lated MP4 energy of+26.6 kcal mol-1 (+25.6 kcal mol-1

including zpe’s). The transition state structure was identifiedas a genuine saddle point by virtue of its single imaginaryfrequency (-2021 cm-1). For the sake of comparison, thereaction was also examined by DFT (B3LYP/LanL2DZ),giving a barrier of+24 kcal mol-1. The isodiazene reactionc is even more exothermic (MP4 energy-71.4 kcal mol-1,or -79.1 kcal mol-1 including zpe’s), but the energy barrierwas found to be higher; in this case the calculated MP4transition state energy was+33.0 kcal mol-1 (+29.6 kcalmol-1 including zpe’s; single imaginary frequency of-2208cm-1), compared to a DFT value of+35 kcal mol-1.

In order for the free N2H2 species to react with D2 asdescribed above, it would first need to dissociate from theFeMoco. The calculated DFT energies for the dissociationof cis-diazene and isodiazene from the FeMoco werecalculated by comparison of structures1 and2 with structure5 in Figure 1, as+44 and+56 kcal mol-1, respectively;combining these figures with the DFT energies for thereactions discussed above gives overall energy barriers forthe dissociation of diazene and isodiazene from the FeMocofollowed by reaction with H2 of +69 and+90 kcal mol-1,respectively.

Scheme 1: (a) Disproportionation ofcis-Diazene;(b) Hypothetical Reaction ofcis-Diazene with H2;(c) Hypothetical Reaction of Isodiazene with H2

a

a Dashed bonds are used to indicate the transition state structures.

Mechanism of Molybdenum Nitrogenase Part 2 Biochemistry, Vol. 41, No. 47, 200213947

An alternative possibility is that H2 reacts directly withthe N2H2 intermediate without dissociation of the latter fromthe FeMoco. The calculated overall energies for reaction ofthe bound diazene and isodiazene species,1 and2, respec-tively, in Figure 1, with H2 to give N2 (bound at Mo) plus2H2 were -4 and+1 kcal mol-1, respectively. Therefore,both reactions are thermodynamically allowed. However, thetransition state structures3 and4 had energies of+69 and+76 kcal mol-1 compared to1 and 2, respectively. Thegeometries of these bound-N2H2 transition state structuresare very similar to those for the free species discussed above.

Binding of H2 at FeMoco Sites.The model complexes usedto evaluate the H2 binding characteristics at different FeMocosites, namely, the trigonal FeII atoms and the five-coordinateMoIII atom generated by opening of the homocitrate ring,are shown in Figure 1. Comparison of complexes5, 6, and7 in Figure 1 gave H2 binding energies (BE’s) of+1 and-2 kcal mol-1 for the FeII and MoIII sites, respectively.Hence, binding of H2 at both sites is extremely weak andthe energies involved are of the same order as van der Waalsinteractions. The result for the Fe site is in good agreement

with the calculations of Rod and Nørskov (22), which gaveH2 BE’s of +2 to+10 kcal mol-1, depending on the numberof H atoms added to the cluster core (as with N2 binding,the simpler model employed in the current work does nottake into account the cost of distortion of the FeMoco onbinding ligands at Fe). The calculated H-H bond lengthsof 0.751 and 0.752 Å in6 and7, respectively, compared to0.744 Å calculated for free H2 (experimental value 0.741Å), suggest essentially no weakening of the H-H bond uponcoordination to either metal. Hence, the FeMoco seems tobe incapable of strong binding and activation of H2 withrespect to oxidative addition to give a dihydride at either itsFe or Mo sites. Structure9 in Figure 1 indicates that thepresence of bound N2 on Mo has no detectable effect on H2

binding at Fe. Indeed, initial geometry optimization on thiscomplex resulted in complete dissociation of the H2, andstructure9 could only be generated by constraining oneFe-H distance to 2.30 Å; the H2 BE for 9 is +1 kcal mol-1.

Perturbation of N2 Protonation by H2. In the precedingpaper, it was found that addition of H to N2 terminallycoordinated at Mo (as in Figure 1,8) resulted in spontaneousrearrangement to a bridging diazenido(1-) ligand, structure10. This suggests a possible role for H2 in inhibiting thereduction of N2; although H2 binds very weakly at Fe, thepresence of H2 on Fe as in9 could destabilize the diazenido-(1-) state by preventing access to this bridging mode. Thestructure of the resulting terminally bound, H2-blockedcomplex is shown in Figure 1,11. Assuming that the H2 BEin this complex is the same as that for complex9, thediazenido(1-) ligand in 11 is destabilized by 5 kcal mol-1

compared to10. It is important to note immediately that thisdestabilization is a kinetic effect, since dissociation of H2 toallow bridging of the NNH ligand is thermodynamicallyfavorable. Nevertheless, kinetic stabilization of thermody-namically disfavored intermediates is the central principleof enzyme catalysis.

Destabilization of the initial protonated form of N2 couldhave important consequences, since this state is well-knownto be the most energetically unfavorable in the course ofreduction of N2 to NH3. Specifically, it has been known formany years from chemical studies on dinitrogen complexesthat protonation of N2 competes with protonation at the metal(23-28). This possibility was investigated in the presentwork. First of all, the effects of adding a single H atom toeither the Fe or Mo site of the model to give a metal hydridewere evaluated. The Mo hydride proved to be more stablethan the Fe hydride; indeed, for some spin states, the H atomspontaneously migrated from Fe to Mo during the geometryoptimization. The final minimum-energy structure is shownin Figure 2,12. The Mo-H bond energy in this structure is45 kcal mol-1. Although the H atom appears to be bridgingbetween the metal atoms, the Fe-H and Mo-H bond lengths(2.21 and 1.79 Å, respectively) show that it is boundprimarily to Mo. Next, an H atom was added to the MoN2

complex at the Mo atom rather than at the terminal N. Thisresulted in complete dissociation of the N2 ligand (Figure 2,13).

H2 Scrambling and H2 Production Properties of theHydride Ligand.Formation of a metal hydride by the processoutlined above raises the possibility of scrambling of thehydride with H2 or D2. This type of reaction is wellestablished in complexes containing both hydride and H2

FIGURE 1: Reactions of bound diazene and isodiazene intermediateswith H2, binding characteristics of H2 at the trigonal Fe andhomocitrate ring-opened Mo sites of the FeMoco, and destabiliza-tion of the NNH intermediate by H2. Color code: C, gray; H, paleblue; N, dark blue; O, red; Fe, green; Mo, purple; S, yellow. Theincoming H2 molecules in structures3 and 4 are highlighted ingreen.

13948 Biochemistry, Vol. 41, No. 47, 2002 Durrant

ligands (29). Binding of H2 at the Fe atom of the Mo hydride12 to give14 was found to be thermoneutral (as with9, oneFe-H distance had to be fixed at 2.30 Å to preventdissociation of the H2). In the case of D2, transfer of the D2to Mo, giving a MoHD2 species, would then allow scram-bling to a molybdenum deuteride plus HD. Such a scramblingreaction would necessarily be approximately thermoneutral;the transition state for this process is approximated bystructure15, in which the H-H distances between the centraland peripheral H atoms of the H3 ligand were constrainedto be equal in the geometry optimization. Comparison ofstructures14 and 15 gave a barrier of approximately+23kcal mol-1 for the scrambling reaction.2

A final consideration is the stability of the Mo hydride/deuteride species itself. As shown in Figure 2, the hydride16, which is formally MoIV and has a protonated carboxylategroup, is thermodynamically unstable with respect to elimi-nation of H2 and reclosure of the homocitrate ring to give17. Thus, there would probably be a limited time scale forscrambling to occur before elimination of H2 returned theenzyme to an earlier state in the Lowe-Thorneley scheme.

DISCUSSION

Mechanism of the HD Formation Reaction.Production ofHD from D2 by nitrogenase requires capture of D2 fromsolution followed by cleavage of the D-D bond. Asmentioned in the Introduction, there are two longstandingopinions as to how this reaction might occur. The first ofthese dates from 1960 and a proposal by Hoch (30) that thereaction involves a diazene level intermediate. Subsequentdetailed experimental work on the reaction (13) lent strongbiochemical support to this view. However, it was also shownin the 1960s that N2 can displace two hydride ligands, asH2, from a metal to give a dinitrogen complex (31). Thisobservation was incorporated into the Lowe-Thorneleyscheme such that N2 can only bind to the FeMoco viadisplacement of two hydride ligands as H2. In the Lowe-Thorneley scheme, HD is produced via reversible displace-ment of N2 by D2.

As expected, the calculations on free diazene and isodia-zene presented in this work indicate that their reaction withD2 to give 2HD plus N2 is energetically highly favorable.Even when bound to the FeMoco, these reactions are stillthermodynamically allowed. However, the very high calcu-lated energy barriers for the reactions of bound N2H2 species,of +69 kcal mol-1 or more, indicate that the direct reactionof a diazene level intermediate with D2 is not a kineticallyviable route to HD in the nitrogenase catalytic cycle.Furthermore, there are no close chemical precedents for thistype of reaction. Perhaps the most relevant chemistrydescribed to date concerns atrans-diazene complex, [µ-N2H2-{Ru(PCy3)(“S4”)}2] [Cy )cyclohexyl, “S4” ) (CH2SC6H4S-)2],described by Sellmann (32). This complex reacts with D2 togive HD plus the equivalent N2D2 complex. However, thisreaction clearly does not proceed via direct interaction ofD2 with the diazene ligand, since this is in an enforced transgeometry. Rather, the reaction was found to involve hetero-lytic cleavage of D2 to give a metal deuteride.

Given the problems attending the production of HD bynitrogenase via direct reaction of a diazene level intermediate,the alternative possibility of HD production via metalhydrides requires further consideration. The factors control-ling this type of chemistry are now quite well understood.Since the first isolation of ‘nonclassical’ dihydrogen com-plexes in 1984 (33), a great deal of experimental andtheoretical work has been directed toward elucidating theinterconversion of dihydrogen and hydride ligands, and thetheoretical aspects of this topic have recently been reviewed(29). Two pathways leading to cleavage of the H-H bondhave been described, namely, oxidative addition andσ-bondmetathesis (Scheme 2).

Oxidative addition involves the formation of two M-Hbonds, with concomitant H-H bond cleavage and oxidationof the metal by two units. The Lowe-Thorneley schemefor HD formation implies the formation of such a dihydride(2, 14). By combining the production of hydrides from H2

with reversible mutual displacement of H2 and N2, Loweand Thorneley produced a mechanism which accounts forthe experimental data relating to H2 inhibition, HD formation,and the limiting H2/N2 stoichiometry. However, this modeldoes not readily explain theR-Asn195 mutant data, discussedin the Introduction, which implicate a reduced N2 intermedi-ate in HD formation (5). A further problem relates to thecomplete lack of activation of H2 toward oxidative additionon both the Fe and Mo sites of the FeMoco, as indicated bythe DFT results presented above and also, in the case of theFe sites, by the DFT studies of Rod and Nørskov (22). Thisis in line with expectations from other studies, which showthat second- and third-row metals, in relatively low oxidationstates and with strongly electron-releasing co-ligands such

2 Of all the structures considered in this and the preceding paper,the relative energies of14 and15 are most sensitive to the constraintsplaced on the Mo coordination sphere. This appears to be a reflectionof the very small size of the hydride ligand, together with the substantialtrans interaction between the hydride and SH ligand on Mo.

FIGURE 2: Binding characteristics of H, effect on N2 binding, H2/Hscrambling, and H2 formation at the Mo site. Color code as inFigure 1.

Scheme 2: Schematic Mechanisms for (a) OxidativeAddition and (b)σ-bond Metathesis

Mechanism of Molybdenum Nitrogenase Part 2 Biochemistry, Vol. 41, No. 47, 200213949

as phosphines, are generally required for oxidative additionof H2 (29). For example, DFT calculations on a series ofcomplexes prepared by oxidative addition of H2 to [MoX2-(PMe3)4] (X ) F, Cl, Br, or I) gave Mo-H bond energiesof at most 64 kcal mol-1; experimentally, these complexesare quite unstable, and the iodide complex, for which thecalculated Mo-H bond energy was 55 kcal mol-1, couldnot be isolated at all (34). The Mo-H bond energy of 45kcal mol-1 determined for12 in the present work is less thanthis value. Hence, it is difficult to imagine that any site onthe FeMoco could sustain cleavage of D2 via oxidativeaddition during turnover.

The second established pathway for H-H bond cleavage,σ-bond metathesis, has also been considered as the basis forvarious HD formation pathways in nitrogenase. PublishedDFT calculations suggest that metathesis of H2 on theFeMoco to give an iron hydride and a protonated sulfide isenergetically feasible (22), while elegant synthetic chemistryhas provided experimental precedents for a model in whichH2 is initially cleaved to an iron hydride and a protonatedsulfur at a diazene level intermediate (32). Furthermore, itis highly likely that H-H bond cleavage occurs via hetero-lytic σ-bond metathesis to give a metal hydride and aprotonated cysteine in hydrogenases (8-12). However, thecomparison with hydrogenases also highlights a criticaldifference with nitrogenase. In hydrogenase, HD formationoccurs via D+/H+ exchange; hence, HD formation is ac-companied by stoichiometric release of D+ into solution. Incontrast, tritium labeling experiments showed that the rateof HD formation by nitrogenase is∼50 times greater thanthe rate of incorporation of T+ into solution (13). If σ-bondmetathesis of the type shown in Scheme 2 was to occur innitrogenase, then the most likely acceptor sites for the D+

would be the three central S atoms of the FeMoco (22).However, two of these are connected to hydrogen bondedchains which would readily allow exchange of D+ with bulkwater (35). Hence, it is difficult to reconcile aσ-bondmetathesis process involving the metal and sulfide sites ofthe FeMoco with the available experimental data on nitro-genase.

There is, however, a special type ofσ-bond metathesiswhich could also occur on the FeMoco. This is the case ofhydrogen atom exchange, which is now well established byboth experiment and theory (29). In nitrogenase, exchangeof a hydride with D2 could provide a mechanism for HDformation with minimal leakage of deuterium into solution,since metal hydrides generally exchange more slowly withwater than do protonated sulfides (8, 10, 12). Furthermore,hydride/dihydrogen exchange is by definition thermoneutraland frequently has a low activation energy. To operate innitrogenase, such a mechanism would require the formationof a hydride contingent upon the presence of both N2 andD2 during catalysis. This criterion can be met by consider-ation of a well-established feature of dinitrogen reductionin model chemical complexes, namely, that initial protonationof N2 competes with protonation at the metal. This was firstrecognized in the equilibrium between the complexes [MX-(N2H)(dppe)2] and [MXH(N2)(dppe)2] (M ) Mo or W; X )F, Cl or Br; dppe) Ph2PCH2CH2PPh2) (26), and a recentinfrared and theoretical study of these complexes confirmedand refined this picture (27). The equilibrium between[WX(N2H)(dppe)2]0/+ and [WXH(N2)(dppe)2]0/+ (X ) F or

Br) has also been observed electrochemically in solution (36).Kinetic studies on the protonation of [Mo(N2)2(dppe)2]highlighted competition for protonation at the metal and theN2 ligand (23, 24); it was found that protonation at the metaltends to labilize the N2 ligand, as well as reducing its basicity.

In the present context, it can be envisaged that protonationof N2 bound to the Mo atom of the FeMoco competes withprotonation of the FeMoco core itself. The DFT calculationspresented above indicate that addition of H to the Mo atomis preferred over addition to the Fe atoms and, furthermore,that protonation of the MoN2 species at Mo results indissociation of N2 from the FeMoco. In the absence of H2

or D2, the diazenido(1-) complex10 (Figure 1) is energeti-cally preferred over the Mo hydride13 (Figure 2) by 5 kcalmol-1. This, however, depends on the diazenido(1-) ligand’sability to bridge to the neighboring Fe atom. If D2 (or H2) isallowed to occupy the Fe site before protonation of N2, asin complex9, the resulting linear diazenido(1-) species11is then formed, and this is essentially isoenergetic with thecorresponding Mo hydride14 plus free N2.

These considerations provide a model for HD formationas shown in Scheme 3. In the absence of D2, N2 binds toMo at state E3 to give structureA and is initially reduced toa bridging diazenido(1-) species,B, as described in thepreceding paper. In the presence of D2, the neighboring Fesite in the initial N2 complexA can become blocked by D2,giving C. This results in kinetic destabilization of thediazenido(1-) intermediateD, allowing competitive forma-tion of the Mo hydride and loss of N2 as inE. The overallinterconversion betweenA andE depends on the concentra-tion of both D2 and N2. The D2 ligand inE is then very wellplaced to undergo scrambling with the Mo hydride viahydrogen atom exchange. Loss of the first molecule of HDleaves a molybdenum deuteride; addition of another H+/e-

couple to give a state equivalent to E4 results in the formationof a second molecule of HD via rapid hydrolysis, and theenzyme is returned to state E2. This gives an overallstoichiometry for HD formation from each molecule of D2

which undergoes scrambling as shown in eq 1

The assignment of the critical step in the HD formationreaction to state E3 is consistent with the kinetic analysis ofLowe and Thorneley, who showed that H2 inhibition mustoccur before state E4 (16). The available experimental datafor the nitrogenase HD formation reaction constitute anexacting test of any atomic model for this process. Thepresent model accommodates these data as described below.

1. Requirement for N2. Rigorous experimental analysisshowed that the nitrogenase HD formation reaction requiresboth D2 and N2 (6). In the present model, the N2 requirementallows two inferences to be drawn about the mechanism, asfollows. First, state E3 in the catalytic cycle must undergohomocitrate ring opening in the presence of N2, but not whenH+ is the only available substrate. This is supported by theanalysis of homocitrate ring opening in the preceding paper;in the absence of N2, ring opening is endothermic, butconcomitant binding of N2 at Mo gives a process which isexothermic overall. This also allows a rationalization of thefact that the Mo atom is coordinatively saturated in the resting

D2 + 2H+ + 2e- f 2HD (1)

13950 Biochemistry, Vol. 41, No. 47, 2002 Durrant

state of the enzyme; in the absence of N2, the Mo site isprotected by homocitrate at the E3 level, and so cannot beprotonated to give a catalytically counterproductive hydride.The second inference is that a hydride formed at the E3 levelis sufficiently long-lived to undergo scrambling with D2, buta hydride at the more reduced E4 level is hydrolyzed toorapidly to give HD. If this were not so, then the Mo hydrideformed at the E4 level in the absence of N2 (see below andScheme 4) would be capable of N2-independent HD forma-tion. The hypothesis of different hydride lifetimes at statesE3 and E4 can be justified by the Lowe-Thorneley rateconstants for evolution of H2 from these two states, 8 and400 s-1, respectively (2).

2. The Electrons for HD Formation Are DiVerted fromN2 Reduction.It has been established that the electrons usedin HD formation are diverted exclusively from N2 reduction

rather than from OHE (7). In the present model, this followsnaturally from the competitive protonation between N2 andthe FeMoco core, as discussed above. OHE is considered tobe an entirely separate process, unaffected by the presenceof either N2 or D2 (see below).

3. TurnoVer under HD GiVes No D2. When nitrogenasewas turned over under a mixture of HD and N2, no D2 wasdetected in the products (37). This observation suggests thatonly a single scrambling event can occur on the catalytictime scale. This is reasonable given that there is alwaysnecessarily one molecule of D2 or, in this case, HD, at theactive site when the Mo hydride is formed, but the localFeMoco environment is relatively isolated from bulk solution(see preceding paper); hence, diffusion of D2 or HD fromthe protein exterior to the FeMoco is relatively slow. Notethat a high pHD might allow for multiple scrambling events,

Scheme 3: Proposed Mechanism for the HD Formation Reactiona

a Each of the bold arrows represents the transfer of a single e- from the Fe protein to the MoFe protein. Circled H+ represent nonlocated protonsassociated with the protein or homocitrate.

Mechanism of Molybdenum Nitrogenase Part 2 Biochemistry, Vol. 41, No. 47, 200213951

resulting in some D2; however, the pressure of HD used inthese experiments was only 0.5 Atm.

4. Minimal Exchange with SolVent. It was found that therate of incorporation of T+ into solution is∼50 times slowerthan the rate of HD formation (13). This is consistent withthe present model, in that work on hydrogenases has shownthat exchange of hydride with solvent is relatively slowcompared to exchange involving protonated nonmetal atoms(8, 10, 12).

5. CompetitiVe Inhibition of N2 Binding and Reduction.The observation that H2 inhibits reduction of N2, but not ofH+, N2O, N3

-, C2H2, CN-, MeNC, or N2H4 (4), is readilyunderstood in that only N2 achieves critical stabilization ofa reduced intermediate by bridging between the Mo and Featoms of the FeMoco. Of all known nitrogenase substrates,N2 is the hardest to reduce; moreover, the first H+/e- additionis the most difficult step of N2 reduction. Hence, althoughother substrates might bind at Mo via homocitrate ringopening, their initial reduction will be significantly easierthan that of N2, and consequently they will not suffer fromcompetitive formation of hydride on the FeMoco core cluster.Alternative substrates may or may not be bound at Mo; forexample, C2H2 binds to states equivalent to E1 and E2 in theLowe-Thorneley cycle (38), for which Mo is unavailableaccording to the present model. In this context, it isinteresting to note that N2 inhibits the reduction of C2H2,both in whole nitrogenase (39) and in solutions of extractedFeMoco (40). Following the analysis presented in the

preceding paper, this can be interpreted in terms of competi-tive weak binding of N2 and C2H2 at the central Fe sites ofthe FeMoco in states E1 and E2.

It has been shown that HD production is suppressed asthe pN2 is increased (41). This is expected in the presentmodel, given that species13 in Figure 2 shows dissociationof N2; thus, a high pN2 will shift the equilibrium betweenDandE in Scheme 3 in favor ofD. At first sight, it might beargued that a high pN2 would itself destabilize intermediateD by blocking the Fe site with a second N2 ligand. However,this destabilization is not a problem since high pN2 also shiftsthe equilibrium between the MoNNH and MoH+ N2 statesback toward the former.

Recent experiments have indicated that in addition tosuppressing N2 reduction, H2 also competes directly with N2for binding sites on the FeMoco (5). In the present model,competition between H2 and N2 for binding sites can beviewed as a result of the cage effect discussed in thepreceding paper. Thus, the FeMoco-containing pocket canhold a strictly limited number of small molecules. In the earlystages of catalysis up to state E3, H2, and N2 bind weaklyand nonspecifically at the FeMoco Fe sites, and so competefor occupancy of the pocket. TheR-Asn195 mutant behavesnormally in this regard; however, the lack of a hydrogenbond to S2B of the FeMoco in this case (35) means that theFeMoco cannot sustain reduction to the level required forthe first protonation of N2, hence no HD can be produced.

Scheme 4: Proposed Mechanism for H2 Evolution in the Absence of Other Substratesa

a Curved arrows represent loss of H2 via the k+7, k+8, andk+9 pathways in the Lowe-Thorneley scheme (14). Other symbols have the samemeanings as those in Scheme 2.

13952 Biochemistry, Vol. 41, No. 47, 2002 Durrant

6. Limiting HD/H2 Stoichiometries.Experiments designedto measure the limiting HD/H2 stoichiometry using pD2’sof up to 400 kPa (4 Atm) gave a limiting value for this ratioof 2 at infinite pD2 (41). For the model shown in Scheme 3,the limiting HD/H2 ratio can be calculated as follows. Asnoted above, scrambling should be limited to the single D2

molecule found at the active site upon formation of themolybdenum hydride, giving the following stoichiometry:

Addition of a further H+/e- couple results in hydrolysis ofthe hydride/deuteride; hence

Overall, the products of eq 3 require two H+/e- couples,since one is used to make the initial Mo hydride. Underlimiting conditions for N2 reduction, six out of every eightelectrons are utilized for N2 reduction, while two are allocatedto OHE. Since all electrons used in HD production arediverted from N2 reduction, the limiting equations for HDproduction therefore consist of the HD component, eq 4, plusthe OHE component, eq 5:

Combining eqs 4 and 5 gives an overall limiting HD/H2 ratioof 2, in agreement with experiment. Interestingly, however,use of still higher pD2, up to 50 Atm, gave a different limitingHD/H2 ratio, of 6.0-6.8 (42). There are two possibleexplanations for this. First, very high pressures should resultin an increase in the number of molecules of D2 which canaccess the FeMoco pocket, enabling multiple scramblingevents to occur. As the mean number of scramblingencounters increases, the HD/H2 ratio should approach alimiting value of 6. However, since all of the electrons usedfor HD formation are diverted from N2 reduction, themaximum possible HD/H2 ratio should be 6, regardless ofthe mechanism of HD formation. Hence, the reported HD/H2 ratios of up to 6.8 are especially intriguing. Since allchemical reactions are equilibria, a sufficiently high pD2 willsuppress obligatory H2 evolution. If OHE is not strictlyrequired for binding and initial reduction of N2, which arethe processes associated with HD formation, then HD/H2

ratios greater than 6 become possible. Taken at face value,then, these experimental data provide evidence that OHE isnot an essential feature of nitrogenase mechanism.

Production of H2 by Nitrogenase.In the absence of anyother reducible substrate, nitrogenase catalyses the productionof H2 from protons. A scheme for proton reduction can bededuced by extension of the analysis of the HD formationreaction given above, Scheme 4. The first part this schemeis identical to that for N2 reduction given in the precedingpaper. In the absence of N2, however, the homocitrate ringremains closed at state E3. Addition of a further H+/e- coupleto the FeMoco core to give a state corresponding to E4 allowshomocitrate ring opening and concomitant protonation at Mo,with subsequent rapid loss of H2 returning the system to state

E2, similar to the scheme for HD formation given above.Experimentally, it has been shown that reduction of a Mocarboxylate complex results in dechelation of the carboxylateligand and protonation of the metal (43). In addition, eachof the states E2-E4 can lose H2, as indicated in the Lowe-Thorneley scheme (2, 15), by migration of an H atom fromS to Fe on the FeMoco core and subsequent hydrolysis (35).Conceptually, Scheme 4 exactly parallels the mechanism ofreduction of H+ to H2 by [Fe4S4(SPh)4]3-, proposed byGronberg et al. (44). In both cases, protonation of the clustercore at sulfur is followed by dissociation of a protonatedleaving group [homocitrate in Scheme 4, thiophenol inreference (44)]. Further protonation then occurs at theexposed metal site, and the resulting hydride is lost as H2,allowing religation of the leaving group. The assignment ofthe reduced iron hydride cluster [Fe4S4H3(SPh)3], rather thanthe initially produced [Fe4S4H3(SPh)3]+, as the primarysource of H2 also lends further support to the hypothesis offaster release of HD from state E4 compared to E3 in Scheme3, as discussed above.

An interesting test of this model is provided by work onNifV - mutant nitrogenase, whose FeMoco contains citraterather than homocitrate (45, 46). In both of these studies,the rate of H2 evolution by NifV- nitrogenase under argonwas found to be∼92% of that of the wild type. Similarly,the rate of H2 evolution of the wild-type enzyme in thepresence of a low concentration of CO was virtually identicalto the rate in the absence of CO. However, the NifV- mutantevolved H2 at significantly lower rates under low COconcentrations, resulting from CO-induced uncoupling ofATP hydrolysis from electron transfer. It has been shownthat CO perturbs the profile of H2 evolution versus pH forwild-type nitrogenase (47), an effect which was explainedin terms of a shift in pKa of a group associated with theFeMoco from∼9 to ∼8.5. We have previously suggestedthat differences in the reactivities of homocitrate- and citrate-containing cofactors are best understood in terms of openingof the (homo)citrate chelate ring, which allows a specifichydrogen bond to Av1R-His442 (Kp1 R-His440) in the caseof homocitrate but not of citrate, thereby enhancing thebasicity of the Mo atom (48). In state E4 of Scheme 4, thereis an equilibrium between protonation at Mo and anunidentified amino acid residue close to the FeMoco. Theposition of the equilibrium will be determined by the pKa ofthe amino acid residue, which should be relatively invariant,and the pKa of the Mo hydride, which will be influenced bythe disposition of the FeMoco. Binding of CO at a trigonalFe site will reduce the basicity of the FeMoco, as will lossof the hydrogen bond between homocitrate andR-His442.Combining both effects then reduces the pKa of the Mohydride below that of the protonatable amino acid residue,switching the position of the equilibrium in state E4 backfrom G to F. SpeciesF cannot easily be reduced, since ithas no sites available to receive further protons, resulting inthe uncoupling of ATP hydrolysis from electron-transfernoted above. Linear regression analysis gave a maximuminhibition of H2 production by CO in these experiments of73% (45); the residual H2 production can be accounted forif OHE continues at its normal rate.

Another set of experiments which can be interpreted usingthe present model concerns the effects of site-directedmutagenesis on Av1 nitrogenase residueR-His195 (49). The

MoH + D2 f 2/3MoD + 1/3MoH + 2/3HD + 1/3D2

(2)

2/3MoD + 1/3MoH + 2/3HD + H+ + e- f4/3HD + 1/3H2 (3)

2D2 + 6H+ + 6e- f 4HD + H2 (4)

2H+ + 2e- f H2 (5)

Mechanism of Molybdenum Nitrogenase Part 2 Biochemistry, Vol. 41, No. 47, 200213953

wild-type enzyme exhibits a significantly lower rate of H2

evolution under an atmosphere of N2 compared to that underAr, due to nitrogen fixation. TheR-Gln195 mutant showed avery similar inhibition of H2 evolution under N2, even thoughin this case the rate of N2 fixation is minimal (50), whilethe R-Asn195 and other mutants showed a low rate of H2

evolution under both Ar and N2 atmospheres (49). It waspreviously suggested (35) that these data can be interpretedin terms of a successive attenuation of the FeMoco’s basicityin going from the wild type through the Gln mutant to theAsn mutant, meaning that the Gln mutant can access acondition in which N2 is bound but only inefficiently reduced,whereas the Asn mutant cannot reduce N2 at all. Thisinterpretation can now be made more specific, as follows.The Gln mutant would operate normally in the absence ofN2, giving rapid H2 evolution via a state similar toG inScheme 4, except that S2B of the FeMoco would receiveonly a hydrogen bond fromR-Gln195, rather than the formaltransfer of a proton as occurs in the wild-type enzyme. Inthe presence of N2, this mutant would be diverted into anonproductive state with the Mo site occupied by N2 ratherthan H. The other mutants, lacking even a hydrogen bondto S2B, would not be able to reach the level of reductionrequired for homocitrate ring opening, resulting in a lowerrate of H2 production even in the absence of N2.

Obligatory Hydrogen EVolution. The model presentedabove accommodates competitive inhibition of N2 bindingand reduction by H2, the HD formation reaction, and thecharacteristics of H+ reduction in the absence of othersubstrates. The one remaining feature of nitrogenase hydro-gen chemistry which requires discussion is obligatoryhydrogen evolution. Is this a necessary feature of reductionof N2 by nitrogenase, or merely a consequence of unavoid-able leakage of H2 from reduced states of the enzyme (2)?Regardless of the answer to this question, it can be reasonablyassumed on chemical grounds that OHE involves metalhydrides. This raises a general problem in seeking aconstructive role for OHE in N2 fixation; addition of H atomsto the nonmetallic elements of the FeMoco (sulfide orhomocitrate) reduces the total oxidation state of the metals,enhancing their ability to bind and protonate N2, whereasaddition of H to the metals has the opposite effect. This isconsistent with experiments on model complexes (23, 24)and is illustrated by the model for HD formation presentedabove, where protonation at the metal competes withprotonation at N2.

It could be argued that formation of metal hydrides allowssome crucial change in the FeMoco structure, which isrequired for N2 binding. Subsequent reductive eliminationof H2 would then allow protonation of N2 to proceed. Thequestion is, what could the structural change be? Onepossibility is that formation of a Mo hydride allows openingof the homocitrate chelate ring (43). This fits in very wellwith the model for H+ reduction in the absence of N2 shownin Scheme 4 (cf. structureG), but the results shown in Figure2 structure13 suggest that the resulting hydride blocksbinding of N2 at Mo almost as effectively as does homoci-trate.

The limiting H2/N2 stoichiometry for Mo nitrogenase is1, but vanadium nitrogenase has a limiting H2/N2 stoichi-ometry of 3.5 (3). There are three possibilities which couldexplain this difference. First, both limiting stoichiometries

might genuinely reflect the enzymes’ mechanisms. In thiscase, V nitrogenase would operate by a mechanism requiringevolution of 3.5 molecules of H2 per N2 reduced, whichseems singularly unlikely. Second, both nitrogenases mightrequire release of one H2 per N2, and the balance of 2.5 H2for V nitrogenase would result from unavoidable leakage ofreducing potential. Third, all the H2 produced by bothenzymes might result from unavoidable leakage. This lastpossibility is in line with Occam’s razor and also fits in bestwith the results from both the present and other (22)theoretical studies, which failed to uncover any potentialconstructive role for OHE. Finally, it was argued in aprevious study (35) that under the experimental conditionsused to measure the limiting H2/N2 ratio (51), wasteful releaseof H2 from states E2 - E4 in the Lowe-Thorneley schemewould not be fully suppressed. The rate of leakage wouldbe essentially constant, dependent only upon the rate ofassociation/dissociation of the nitrogenase component pro-teins. If correct, this would explain both the phenomenon ofOHE and the different limiting stoichiometries for thedifferent enzymes, as well as the HD/H2 ratios greater than6 which were found at very high pD2, as discussed above.

ACKNOWLEDGMENT

I thank C. J. Pickett, D. J. Lowe and B. E. Smith (JohnInnes Centre), R. L. Richards (University of Sussex), andR. A. Henderson (University of Newcastle) for helpfuldiscussions.

REFERENCES

1. Pickett, C. J. (1996) The Chatt cycle and the mechanism ofenzymic reduction of molecular nitrogen,J. Biol. Inorg. Chem.1, 601-606.

2. Burgess, B. K., and Lowe, D. J. (1996) Mechanism of molybdenumnitrogenase,Chem. ReV. 96, 2983-3011.

3. Eady, R. R. (1996) Structure-function relationships of alternativenitrogenases,Chem. ReV. 96, 3013-3030.

4. Guth, J. H., and Burris, R. H. (1983) Inhibition of nitrogenase-catalyzed NH3 formation by H2, Biochemistry 22, 5111-5122.

5. Fisher, K., Dilworth, M. J., and Newton, W. E. (2000) Differentialeffects on N2 binding and reduction, HD formation, and azidereduction withR-195His- andR-191Gln-substituted MoFe proteinsof AzotobacterVinelandii nitrogenase,Biochemistry 39, 15570-15577, and references therein.

6. Li, J., and Burris, R. H. (1983) Influence of pN2 and pD2 on HDformation by various nitrogenases,Biochemistry 22, 4472-4485.

7. Newton, W. E., Bulen, W. A., Hadfield, K. L., Stiefel, E. I., andWatt, G. D. (1977) HD formation as a probe for intermediates inN2 reduction, in Recent DeVelopments in Nitrogen Fixation(Newton, W., Postgate, J. R., and Rodriguez-Barreuco C., Eds.)pp 119-130, Academic Press, London.

8. Tamiya, N., and Miller, S. L. (1963) Kinetic studies on hydro-genase,J. Biol. Chem. 238, 2194-2198.

9. Sellmann, D., Geipel, F., and Moll, M. (2000) [Ni(NHPnPr3)(‘S3’)],the first nickel thiolate complex modeling the nickel cysteinatesite and reactivity of [NiFe] hydrogenase,Angew. Chem., Int. Ed.39, 561-563.

10. Yagi, T., Tsuda, M., and Inokuchi, H. (1973) Parahydrogen-orthohydrogen conversion and hydrogen-deuterium exchangereactions,J. Biochem. 73, 1061-1081.

11. Siegbahn, P. E. M., Blomberg, M. R. A., Wirstam, M., andCrabtree, R. H. (2001) The mechanism of the Ni-Fe hydroge-nases: a quantum chemical perspective,J. Biol. Inorg. Chem. 6,460-466.

12. Zorin, N. A., Dimon, B., Gagnon, J., Gaillard, J., Carrier, P., andVignais, P. M. (1996) Inhibition by iodoacetamide and acetyleneof the H-D-exchange reaction catalyzed byThiocapsa roseo-persicinahydrogenase,Eur. J. Biochem. 241, 675-681.

13. Burgess, B. K., Wherland, S., Newton, W. E., and Stiefel, E. I.(1981) Nitrogenase reactivity: insight into the nitrogen-fixing

13954 Biochemistry, Vol. 41, No. 47, 2002 Durrant

process through hydrogen-inhibition and HD-forming reactions,Biochemistry 20, 5140-5146.

14. Lowe, D. J., and Thorneley, R. N. F. (1984) The mechanism ofKlebsiella pneumoniaenitrogenase action. The determination ofrate constants required for the simulation of the kinetics of N2

reduction and H2 evolution,Biochem. J. 224, 895-901.15. Lowe, D. J., and Thorneley, R. N. F. (1984) The mechanism of

Klebsiella pneumoniaenitrogenase action. Pre-steady-state kineticsof H2 formation,Biochem. J. 224, 877-886.

16. Thorneley, R. N. F., and Lowe, D. J. (1984) The mechanism ofKlebsiella pneumoniaenitrogenase action. Pre-steady-state kineticsof an enzyme-bound intermediate in N2 reduction and of NH3formation,Biochem. J. 224, 887-894.

17. Thorneley, R. N. F., and Lowe, D. J. (1984) The mechanism ofKlebsiella pneumoniaenitrogenase action. Simulation of thedependences of H2-evolution rate on component-protein concen-tration and ratio and sodium dithionite concentration,Biochem.J. 224, 903-909.

18. Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E.,Robb, M. A., Cheeseman, J. R., Zakrzewski, V. G., Montgomery,J. A., Jr., Stratmann, R. E., Burant, J. C., Dapprich, S., Millam,J. M., Daniels, A. D., Kudin, K. N., Strain, M. C., Farkas, O.,Tomasi, J., Barone, V., Cossi, M., Cammi, R., Mennucci, B.,Pomelli, C., Adamo, C., Clifford, S., Ochterski, J., Petersson, G.A., Ayala, P. Y., Cui, Q., Morokuma, K., Malick, D. K., Rabuck,A. D., Raghavachari, K., Foresman, J. B., Cioslowski, J., Ortiz,J. V., Stefanov, B. B., Liu, G., Liashenko, A., Piskorz, P.,Komaromi, I., Gomperts, R., Martin, R. L., Fox, D. J., Keith, T.,Al-Laham, M. A., Peng, C. Y., Nanayakkara, A., Gonzalez, C.,Challacombe, M., Gill, P. M. W., Johnson, B. G., Chen, W.,Wong, M. W., Andres, J. L., Head-Gordon, M., Replogle, E. S.,and Pople, J. A. (1998)Gaussian 98, revision A.7, Gaussian, Inc.,Pittsburgh, PA.

19. McKee, M. L., Squillacote, M. E., and Stanbury, D. M. (1992)Ab initio investigation of dihydrogen transfer fromcis-1,2-diazene,J. Phys. Chem. 96, 3266-3272.

20. Berman, H. M., Westbrook, J., Feng, Z., Gilliland, G., Bhat, T.N., Weissig, H., Shindyalov, I. N., and Bourne, P. E. (2000) TheProtein Data Bank,Nucleic Acids Res. 28, 235-242.

21. Sayle, R. (1993-1995)RasMol, version2.4, Glaxo Research andDevelopment, Greenford, U.K.

22. Rod, T. H., and Nørskov, J. K. (2000) Modeling the nitrogenaseFeMo cofactor,J. Am. Chem. Soc. 122,12751-12763.

23. Henderson, R. A. (1982) The mechanism of formation of hy-drazido(2-)- and hydrido- complexes by the reaction of dinitrogencomplexes with acids in tetrahydrofuran,J. Chem. Soc., DaltonTrans., 917-925.

24. Henderson, R. A. (1984) Reactions betweentrans-bis[1, 2- bis-(diethylphosphino)ethane]bis-(dinitrogen)molybdenum and hy-drogen chloride in tetrahydrofuran: a mechanistic study,J. Chem.Soc., Dalton Trans, 2259-2263.

25. Cusanelli, A., and Sutton, D. (1996) Cis and trans rhenium hydridodinitrogen complexes from protonation reactions of Cp*Re-(PMe3)2(N2), Organometallics 15, 3717-3722.

26. Chatt, J., Kan, C. T., Leigh, G. J., and Pickett, C. J. (1980)Transition-metal binding sites and ligand parameters,J. Chem.Soc., Dalton Trans., 2032-2038.

27. Lehnert, N., and Tuczek, F. (1999) The reduction pathway of end-on coordinated dinitrogen. I. Vibrational spectra of Mo/W-N2,-NNH, and-NNH2 complexes and quantum chemistry assistednormal coordinate analysis,Inorg. Chem. 38, 1659-1670.

28. Komiya, S., Akita, M., Yoza, A., Kasuga, N., Fukuoka, A., andKai, Y. (1993) Isolation of a zerovalent iron dinitrogen complexwith 1,2-bis(diethylphosphino)ethane ligands,J. Chem. Soc.,Chem. Commun., 787-788.

29. Maseras, F., Lledo´s, A., Clot, E., and Eisenstein, O. (2000)Transition metal polyhydrides: from qualitative ideas to reliablecomputational studies,Chem. ReV. 100, 601-636.

30. Hoch, G. E., Schneider, K. C., and Burris, R. H. (1960) Hydrogenevolution and exchange, and conversion of N2O to N2 by soybeanroot nodules,Biochim. Biophys. Acta 37, 273-279.

31. Sacco, A., and Rossi, M. (1967) Hydride and nitrogen complexesof cobalt,Chem. Commun.316.

32. Sellmann, D., Fu¨rsattel, A., and Sutter, J. (2000) The nitrogenasecatalyzed N2 dependent HD formation: a model reaction and itssignificance for the FeMoco function,Coord. Chem. ReV. 200-202, 545-561.

33. Kubas, G. J., Ryan, R. R., Swanson, B. I., Vergamini, P. J., andWasserman, H. J. (1984) Characterization of the first examples

of isolable molecular hydrogen complexes, M(CO)3(PR3)2(H2)(M ) Mo, W; R ) Cy, i-Pr). Evidence for a side-on bonded H2

ligand,J. Am. Chem. Soc. 106, 451-452.34. Hascall, T., Rabinovich, D., Murphy, V. J., Beachy, M. D.,

Friesner, R. A., and Parkin, G. (1999) Mechanistic and theoreticalanalysis of the oxidative addition of H2 to six-coordinatemolybdenum and tungsten complexes M(PMe3)4X2 (M ) Mo, W;X ) F, Cl, Br, I): an inverse equilibrium isotope effect and anunprecedented halide dependence,J. Am. Chem. Soc. 121, 11402-11417.

35. Durrant, M. C. (2001) Controlled protonation of iron-molybde-num cofactor by nitrogenase: a structural and theoretical analysis,Biochem. J. 355, 569-576.

36. Le Grand, N. (2001) Contribution a` l’etude des processus detransfert de couples{H+/e} et de transformation de ligands enrelation avec la re´duction de l′azote mole´culaire, Ph.D. Thesis,L’Universite de Bretagne Occidentale, Brest, France.

37. Burgess, B. K., Corbin, J. L., Rubinson, J. F., Li, J. G., Dilworth,M. J., and Newton, W. E. (1984) Nitrogenase reactivity, inAdVances in Nitrogen Fixation Research(Veeger, C., and Newton,W. E. Eds.), p146, Nijhof/Junk, Pudoc, The Netherlands.

38. Lowe, D. J., Fisher, K., and Thorneley, R. N. F. (1990)Klebsiellapneumoniaenitrogenase. Mechanism of acetylene reduction andits inhibition by carbon monoxide,Biochem. J. 272, 621-625.

39. Davis, L. C., and Wang, Y.-L. (1980) In vivo and in vitro kineticsof nitrogenase,J. Bacteriol. 3, 1230-1238.

40. Bazhenova, T. A., Bazhenova, M. A., Petrova, G. N., Mironova,S. A., and Strelets, V. V. (2000) Catalytic behavior of thenitrogenase iron-molybdenum cofactor extracted from the enzymein the reduction of C2H2 under nonenzymatic conditions,Kinet.Catal. 41, 499-510.

41. Jensen, B. B., and Burris, R. H. (1985) Effect of high pN2 andhigh pD2 on NH3 production, H2 evolution, and HD formation bynitrogenases,Biochemistry 24,1141-1147.

42. Simpson, F. B., Thiele, J., Liang, J., and Burris, R. H. (1985) Theeffect of high pressures of H2 and D2 on HD formation bynitrogenase, inNitrogen Fixation Research Progress(Evans, H.J., Bottomley, P. J., and Newton, W. E., Eds.), p 614, MartinusNijhoff Publishers, Dordrecht, The Netherlands.

43. Hughes, D. L., Ibrahim, S. K., Pickett, C. J., Querne, G., Laouenan,A., Talarmin, J., Queiros, A., and Fonseca, A. (1994) Oncarboxylate as a leaving group at the active site of Mo nitroge-nase: electrochemical reactions of some Mo and W carboxylates,formation of mono-, di- and tri-hydrides and the detection of anMoH2(N2) intermediate,Polyhedron 13, 3341-3348.

44. Gronberg, K. L. C., Henderson, R. A., and Oglieve, K. E. (1998)A unified mechanism for the stoichiometric reduction of H+ andC2H2 by [Fe4S4(SPh)4]3- in MeCN,J. Chem. Soc., Dalton Trans.3093-3104.

45. McLean, P. A., Smith, B. E., and Dixon, R. A. (1983) Nitrogenaseof Klebsiella pneumoniae nif Vmutants. Investigation of the novelcarbon monoxide-sensitivity of hydrogen evolution by the mutantenzyme,Biochem. J. 211, 589-597.

46. Hoover, T. R., Imperial, J., Liang, J., Ludden, P. W., and Shah,V. K. (1988) Dinitrogenase with altered substrate specificity resultsfrom the use of homocitrate analogues for in vitro synthesis ofthe iron-molybdenum cofactor,Biochemistry 27, 3647-3652.

47. Pham, D. N., and Burgess, B. K. (1993) Nitrogenase reactivity:effects of pH on substrate reduction and CO inhibition,Biochem-istry 32, 13725-13731.

48. Gronberg, K. L. C., Gormal, C. A., Durrant, M. C., Smith, B. E.,and Henderson, R. A. (1998) Why R-homocitrate is essential tothe reactivity of FeMo-cofactor of nitrogenase: studies on NifV--extracted FeMo-cofactor,J. Am. Chem. Soc. 120, 10613-10621.

49. Kim, C.-H., Newton, W. E., and Dean, D. R. (1995) Role of theMoFe proteinR-subunit histidine-195 residue in FeMo-cofactorbinding and nitrogenase catalysis,Biochemistry 34, 2798-2808.

50. Dilworth, M. J., Fisher, K., Kim, C.-H., and Newton, W. E. (1998)Effects on substrate reduction of substitution of Histidine-195 byglutamine in theR-subunit of the MoFe protein ofAzotobacterVinelandii nitrogenase,Biochemistry 37, 17495-17505.

51. Simpson, F. B., and Burris, R. H. (1984) A nitrogen pressure of50 atm does not prevent evolution of hydrogen by nitrogenase,Science 224, 1095-1097.

BI025624R

Mechanism of Molybdenum Nitrogenase Part 2 Biochemistry, Vol. 41, No. 47, 200213955