co-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 m. soloviov and...

12
CO-dynamics in the active site of cytochrome c oxidase Maksym Soloviov and Markus Meuwly Citation: The Journal of Chemical Physics 140, 145101 (2014); doi: 10.1063/1.4870264 View online: http://dx.doi.org/10.1063/1.4870264 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/140/14?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Reaction of the C3(X1Σg +) carbon cluster with H2S(X1A1), hydrogen sulfide: Photon-induced formation of C3S, tricarbon sulfur J. Chem. Phys. 141, 204310 (2014); 10.1063/1.4901891 Monitoring equilibrium reaction dynamics of a nearly barrierless molecular rotor using ultrafast vibrational echoes J. Chem. Phys. 141, 134313 (2014); 10.1063/1.4896536 A combined experimental/theoretical investigation of the near-infrared photodissociation of IBr − ( CO 2 ) n J. Chem. Phys. 129, 224304 (2008); 10.1063/1.3033746 Femtosecond study on the isomerization dynamics of NK88. I. Ground-state dynamics after photoexcitation J. Chem. Phys. 125, 044512 (2006); 10.1063/1.2210482 Collisional and photoinitiated reaction dynamics in the ground electronic state of Ca–HCl J. Chem. Phys. 123, 064301 (2005); 10.1063/1.1995700 This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP: 192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Upload: others

Post on 10-Mar-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

CO-dynamics in the active site of cytochrome c oxidaseMaksym Soloviov and Markus Meuwly Citation: The Journal of Chemical Physics 140, 145101 (2014); doi: 10.1063/1.4870264 View online: http://dx.doi.org/10.1063/1.4870264 View Table of Contents: http://scitation.aip.org/content/aip/journal/jcp/140/14?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Reaction of the C3(X1Σg +) carbon cluster with H2S(X1A1), hydrogen sulfide: Photon-induced formation of C3S,tricarbon sulfur J. Chem. Phys. 141, 204310 (2014); 10.1063/1.4901891 Monitoring equilibrium reaction dynamics of a nearly barrierless molecular rotor using ultrafast vibrational echoes J. Chem. Phys. 141, 134313 (2014); 10.1063/1.4896536 A combined experimental/theoretical investigation of the near-infrared photodissociation of IBr − ( CO 2 ) n J. Chem. Phys. 129, 224304 (2008); 10.1063/1.3033746 Femtosecond study on the isomerization dynamics of NK88. I. Ground-state dynamics after photoexcitation J. Chem. Phys. 125, 044512 (2006); 10.1063/1.2210482 Collisional and photoinitiated reaction dynamics in the ground electronic state of Ca–HCl J. Chem. Phys. 123, 064301 (2005); 10.1063/1.1995700

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 2: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

THE JOURNAL OF CHEMICAL PHYSICS 140, 145101 (2014)

CO-dynamics in the active site of cytochrome c oxidaseMaksym Soloviov and Markus Meuwlya)

Department of Chemistry, University of Basel, Klingelbergstrasse 80, 4056 Basel, Switzerland

(Received 20 January 2014; accepted 21 March 2014; published online 8 April 2014)

The transfer of CO from heme a3 to the CuB site in Cytochrome c oxidase (CcO) after photolysisis studied using molecular dynamics simulations using an explicitly reactive, parametrized poten-tial energy surface based on density functional theory calculations. After photodissociation from theheme-Fe, the CO ligand rebinds to the CuB site on the sub-picosecond time scale. Depending onthe simulation protocol the characteristic time ranges from 260 fs to 380 fs which compares withan estimated 450 fs from experiment based on the analysis of the spectral changes as a functionof time delay after the photodissociating pulse. Following photoexcitation ≈90% of the ligands arefound to rebind to either the CuB (major component, 85%) or the heme-Fe (minor component, 2%)whereas about 10% remain in an unbound state. The infrared spectra of unbound CO in the ac-tive site is broad and featureless and no appreciable shift relative to gas-phase CO is found, whichis in contrast to the situation in myoglobin. These observations explain why experimentally, un-bound CO in the binuclear site of CcO has not been found as yet. © 2014 AIP Publishing LLC.[http://dx.doi.org/10.1063/1.4870264]

I. INTRODUCTION

The dynamics of small ligands in proteins strongly de-pends on their environment. One of the routinely employedprobe molecules is carbon monoxide (CO) which is oftenused in ligand binding-rebinding experiments.1–3 In such ex-periments, the ligand is dissociated from its binding partner,which is typically the iron atom of a heme group, by usinga UV/visible laser pulse that promotes the system to an elec-tronically excited and repulsive state. Subsequently, the time-resolved data of the probe molecule allows to follow the time-varying environmental changes by spectroscopic means.4, 5

The structural interpretation of the data, including the natureof the electrostatic environment,6 the ligand migration path-ways and reactivities,7, 8 or the ligand-protein coupling9 is,however, far from straightforward. Usually, the conclusionsthat can be drawn from experiment alone about the protein en-vironment are indirect except for circumstances under whichspectroscopic and structural signatures can be recorded simul-taneously which is only possible in rare cases.10 Alternatively,validated molecular dynamics (MD) simulations provide therequired temporal and spatial resolution to complement ex-perimental investigations and allow to interpret the dynamicsat atomic resolution.11, 12

The ultrafast dynamics of cytochrome c oxidase(CcO) has previously been investigated using a numberof experimental techniques13, 14 which provide informationabout the protein environment and ligand binding-rebindingreactions.15, 16 The physiological function of CcO is to reducemolecular oxygen to water. The protein contains two a-typeheme groups: heme a (low spin) is hexacoordinated and me-diates electron transfer from exogenous cytochrome c towardsthe active site whereas heme a3 (high-spin), near the copper

a)Author to whom correspondence should be addressed. Electronic mail:[email protected]

atom CuB ≈5 Å away from the iron atom of heme a3, acts asthe binding site for molecular oxygen and its reaction inter-mediates during four-electron reduction. The electron equiv-alents are provided by cytochrome c in the outside of themitochondrial inner membrane (the intermembrane space) viaa copper site (CuA) and a low-spin heme (heme a) to the O2

reduction site. The protons used for water formation fromO2 are transferred from the inside of the mitochondrial innermembrane (the matrix space) through two hydrogen-bondednetworks known as the K and D pathways. In addition to O2,heme a3 also binds other diatomic ligands, including nitricoxide (NO) or carbon monoxide (CO) which are physiolog-ically generated and act as biological messengers.17 Becauseheme binds CO tightly, its accessibility to the heme a3 in CcOmust be well controlled.

The active site of interest for the current study is theone where the four electron reduction takes place. It containsheme a3, bound to the protein via histidine His411, and a three-histidine-coordinated Cu center (CuB) located 4.5–5.0 Åaway from the Fe of heme a3. Together they form a bimetal-lic active site. Before binding to heme a3, ligands such as O2

and CO bind to CuB intermediately. The reverse reaction, in-volving ligand transfer from heme a3 out of the protein viaCuB, can be studied using flash photolysis. CO is known topermanently inhibit CcO by forming a thermodynamicallystable complex with the heme a3-Fe. As in other protein en-vironments, the frequency of the CO stretch is sensitive to thechanges in the chemical environment.6, 18, 19 This property hasbeen extensively used in numerous spectroscopic studies ofenzymatic active sites as a probe. In CcO, the CO moleculeis known to bind transiently to the CuB site before form-ing a thermodynamically more stable complex with the hemea3-Fe.3 The reverse reaction can be induced by photolysis ofthe Fe–CO bond. Subsequent reactions involve the transfer ofCO to the CuB site which occurs on the 1–2 ps time scale.20

0021-9606/2014/140(14)/145101/11/$30.00 © 2014 AIP Publishing LLC140, 145101-1

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 3: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

In order to better understand the atomistic details under-lying CO transfer dynamics in the bimetallic center of CcO,MD simulations together with a parametrized force field areemployed. Such simulations are capable of characterizing theligand transfer trajectories and relate structural and temporalaspects of the process. One of the open questions is whetherfree CO is formed at all and on what time scales it is expectedto exist under such circumstances. One suitable observableto compare with is the infrared spectroscopy of CO whichchanges characteristically between metal-bound and free CO.Furthermore, it is of interest to explicitly follow the ligandunbinding and rebinding dynamics and to determine the typ-ical time scales for this process which is expected to be onthe sub-picosecond to picosecond time scale. This makes thepresent system particularly relevant for direct comparison be-tween experiment and simulation because extensive averagingis possible.

II. METHODS

A. Molecular dynamics simulations

All simulations of Cytochrome C oxidase with boundand unbound CO were carried out using CHARMM21 withthe CHARMM22 force field.22 For CcO, the X-ray struc-ture with the Protein Data Bank reference 1AR123 was used.The computational model of the protein contains subunits A(529 amino acids) and B (252 amino acids), 3 copper atoms,2 heme groups (heme a and heme a3), calcium and mag-nesium ions, 9 molecules of lauryldimethylamine oxide, thecarbon monoxide molecule, and 42 water molecules, whichhave been found in the X-ray structure. For simulations withFe-bound CO of heme a3, the initial Fe–C(CO) bond lengthwas set to 1.90 Å. The crosslink between His276 and Tyr280

(Nε–Cε) known for bovine CcO23, 24 has been introduced viaa CHARMM patch. The two copper atoms, which are not partof the active site of interest are treated by applying harmonicconstraints to the bond lengths and the angles between thesecopper atoms and the protein. The 1AR1 structure was usedas this was also the reference structure in previous computa-tional work.20 The active site structures of unligated, reducedCcO of P. denitrificans (1AR1), Th. thermophilus (3EH5), andBos taurus (2EIJ) are structurally very similar (see Figure 1in the supplementary material89).

Hydrogen atoms were added using CHARMM and theprotein was solvated in a pre-equilibrated water box (80× 117 × 74 Å3) which leads to a total system size of 58 092atoms. The protonation states of the residues were deter-mined using PROPKA25–28 and Karlsberg+.29, 30 For the mostrelevant active site residues His325, His326, His411, the stan-dard δ-protonation state was found. However, it is worth-while to mention that “standard” protonation states can bemisleading as for the His64 residue in Myoglobin whichhas been shown to be ε-protonated based on computationsand experiments.31, 32 All simulations used periodic boundaryconditions and the TIP3 model for water.33 The total chargeof the system is −4.0e and was neutralized by replacing ran-dom TIP3 waters by potassium ions. Some water moleculeswere replaced by potassium and chloride ions so that the

concentration of KCl in the water box was ≈ 0.15 M. TheSHAKE algorithm34, 35 with a tolerance of 10−6 was appliedto all bonds which included hydrogen atoms. A time stepof 1 fs was used, and the non-bonded interactions were cutoff at 14 Å. First, the system was minimized with the steep-est descent algorithm, then heated to 300 K during 300 ps.The early equilibration was a 300 ps NpT simulation (Nose-Hoover thermostat with coupling constant 5.0 ps−1) whichleads to a pressure of 1 atm at the end of the first equilibrationphase. Following this, the box size was adjusted and a 700 psNV T equilibration was run. Then, production runs were car-ried out for 5 ns, during which the initial configurations forthe excitation were collected every 10 ps.

B. Intermolecular interactions

The global PES V ( �X) is a function of the coordinates �Xof all atoms involved. For the particular case of CO transferbetween the bimetallic site in CcO, we write the total energyas

Vtot( �X) = VFF( �Q) + V (r, R, θ ). (1)

The force field VFF( �Q) is the standard CHARMM22 forcefield36 and �Q contains all configurational coordinates of thesystem except R, ρ, and θ (see below) which describe the co-ordinate dependence of the PES involving the binuclear siteand the ligand in the electronic ground state, see Figure 1.The coordinates of the subsystem are R = | �R| the distance be-tween the Fe atom of heme a3 (labeled Fea3 in the following)and the CuB site, the vector r = |�r| between Fea3 and the cen-ter of mass (COM) of CO, and θ which is the angle between�r and the vector along the CO-bond. Thus, θ corresponds tothe CO-rotation. In the ligand-bound state, the CO stretchingcoordinate was parametrized as a harmonic potential with aforce constant of 1333.1 kcal/mol and an equilibrium separa-tion of 1.128 Å, the charges are +0.021 and −0.021 for C andO, respectively.36, 37

For the three-dimensional PES V (r, R, θ ), more than3300 reference energies were calculated using densityfunctional theory (DFT) calculations with the B3LYP38–41

functional and the 6-31G(d,p) basis set.42, 43 All electronic

FIG. 1. (a) The model of the system (red – active site of interest, cyan –protein backbone. (b) Diagram showing variables of the potential.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 4: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-3 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

structure calculations were carried out with Gaussian09.44 Forthe electronic structure calculations, a grid was defined whichincludes Fe–Cu distances between 4.3 and 5.5 Å with a spac-ing of 0.1 Å, Fe–CO(CoM) distances from 1.5 to 3.5 Å andangles θ corresponding to an 11-point Gauss-Legendre grid.Such a definition for the angular grid was found to enhancethe stability of the fit and allow for a convenient represen-tation of the angular degree of freedom in fitting the three-dimensional PES.45–47 We found it convenient to introducethe dimensionless coordinate ρ defined as ρ = r−σ

R−2σwhere

σ = 1.5 Å is the minimal Fe–C and Cu–C distance en-countered in DFT scans, high up on the repulsive wall. Thistransformation maps ρ ∈ [0, 1] for all separations R whichsimplifies and stabilizes the fitting.47, 48 For an analyticalrepresentation of the interaction energies, the followingparametrization is used:

V (R, ρ, θ ) =10∑

λ=0

Vλ(R, ρ)Pλ(cos(θ )), (2)

where Pλ are Legendre polynomials and the Vλs are doubleMorse-like potentials

Vλ(R, ρ) = D1,λ(R) · (1 − e−β1,λ(R)(ρ−ρe1 ,λ(R)))2

+D2,λ(R)(1 − e−β2,λ(R)·(ρe2 ,λ(R)−ρ))2

−α(R). (3)

Each parameter Xλ, including Di(R), β i(R), and ρei(R), is

fitted to an expression Xλ(R) = 12p0 tanh(p1(R − p2)) + p3.

This parametrization scheme was already found useful forinvestigating proton transfer between a donor and an acceptorand allows to work with analytical derivatives required forMD simulations.47, 48

The parameters pi were determined from least-squaresfitting by using a Nelder-Mead simplex49, 50 followed bya nonlinear least squares refinement with the I-Nollssoftware.51, 52

For the unligated state, the dissociated CO molecule wasdescribed by a three-point fluctuating charge model53

qi(rCO) = Ai + Bi · rCO + Ci · r2CO + Di · r3

CO, (4)

where the qi are the charges on the C and O atoms, respec-tively (see Table I), and an additional charge was placed atthe COM of CO with

qCOM = qC + qO. (5)

Such a model was preferred in the present case over a morerigorous multipolar model for computational efficiency.54

The C–O bond potential is an anharmonic and spectro-scopically accurate rotational Rydberg-Klein-Rees (RRKR)potential.55–57 This model has performed well in previousstudies.53, 58

TABLE I. Parameters used for 3-point charge model.

Atoms A (e) B (e/Å) C (e/Å2) D (e/Å3)

C –11.011 20.205 –13.149 3.0461O –11.254 22.309 –15.981 3.9834

C. Photodissociation and ligand transfer

Photodissociation was simulated based on the suddenapproximation.59 Starting from the bound state (FeCO), theground state PES was instantaneously switched to the param-eters of the unbound state and a repulsive term between theFe and the CCO atoms was added

Vrepulsive(r) = C ·( r

σ

)−12, (6)

where σ is the equilibrium distance of the Fe–C bond(σ = 1.9 Å) and C = 10 kcal/mol is an energy parameter.This potential mimics the repulsive part of the excited statePES to which the Fe–C bond is pumped by the photodis-sociating laser pulse.4, 60 We note that such a preparation ofthe system does not necessarily redistribute the excess energycorrectly into the available degrees of freedom of the leav-ing CO-fragment. However, as we are primarily interested inpreparing a representative ensemble of photodissociated ini-tial structures and because the dissociative potential is typi-cally only active for a few vibrational periods (a few tenthsof femtoseconds) in the present case, a more rigorous treat-ment was not deemed necessary. More appropriate treatmentsof such processes have been recently discussed in the contextof photodissociation of small molecules together with finalstate analysis.61 After applying the photodissociating pulse,the system is switched back – after time delays ranging from20 to 200 fs – to the ground state surface in order to followthe reactive dynamics. For this the bond between Fe and COwas removed and replaced by Eq. (6) which was active for100 fs after which the Fe–CO nonbonded interactions werereintroduced. This included in particular the van der Waalsand electrostatic interactions between the CO ligand andits environment. For the electrostatic part, the fluctuat-ing three-point charge model (see Sec. II) was employed.The fitted, three-dimensional PES is reactive and allows todescribe both, the Fe–CO and Cu–CO states. However, alter-native procedures to follow chemical reactions in condensed-phase simulations are available, including reactive molecu-lar dynamics,59, 62–65 reformulations of it,66 or the empiricalvalence bond (EVB) formalism.67

III. RESULTS AND DISCUSSION

A. PES fitting

The PES was parametrized by fitting to DFT-referenceenergies. Figure 2 shows a model of the active site that wasused for generating the three-dimensional PES correspond-ing to the singlet ground state. The model includes the trun-cated Heme a3 in which the isoprenoid chain was replaced bya methyl group, the CO ligand, the Cu+ cation, and the His276,His325, His326, and His411 side chains which were truncated tothe methyl-imidazole core. In order to estimate how stronglythe structure deforms when CO is bound to either CuB or theHeme a3, partial optimizations were carried out with the po-sitions of the His-Cβ carbons and the carbon atoms of thepropionate-carboxylic groups being fixed (see Figure 2). Asit is expected the CO–CuB bound structure (θ = 180◦) is≈16 kcal/mol higher in energy than the CO-heme a3 one

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 5: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-4 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

FIG. 2. Optimized structure of the active site at the B3LYP/6-31G(d,p) levelfor CO bound to heme-Fe. The histidine Cβ carbons and the carbon atomsof the propionate-carboxylic groups, marked blue, are fixed at their X-raypositions in the partial optimizations.

(θ = 0). In the partially optimized structure, the most sig-nificant structural differences are found for the CuB atom (itsdisplacement relative to the reference structure is 0.38 Å) andthe heme-a3-Fe (displacement 0.15 Å). However, the overallstructure is well preserved which establishes that using theX-ray structure is a meaningful reference for scanning thePES.

The minimum barrier of the transfer reaction on theground state PES is 38.1 kcal/mol and the root mean squareerror between the parametrized PES and the DFT referenceenergies is 1.2 kcal/mol. A direct comparison between the ref-erence data and the fit is reported in Figure 3(c) for the cutwith an Fe–Cu distance of 5.2 Å and demonstrates the highquality of the PES.

B. Dynamics of CO bound to heme a3 Fe

As a first validation of the parametrized PES, NV T

simulations were carried out for Fe-bound CO. The average

Fe–C bond length is 1.93 ± 0.067 Å, the average Fe–C–Ovalence angle is 176 ± 2.04◦ – close to linearity – and theaverage CuB–O distance is 2.39 ± 0.538 Å. The structuraldata reported for the bovine heart CcO5 and the one of Th.thermophilus68 are summarized in Table II. Significant dif-ferences are, however, observed for the CO-tilt (Fe–C–O an-gle). The canonical Fe–CO conformation is linear18 which isalso found in the present MD simulations but is very differentin Th. thermophilus. Part of the difference could be relatedto the rather low resolution of the X-ray diffraction experi-ments of 2.8–2.9 Å and 2.8–3.2 Å and the low temperatureof 100 K at which these experiments were carried out.5, 68

The tilt observed in the X-ray structures has been proposed toresult from “repulsion between the O-atom of the Fe-boundCO and CuB.”68 However, given the experimental data onrelated systems and the fact that the partial optimizations atthe B3LYP/6-31G(d,p) level (see above) based on the X-raystructure yield an almost linear Fe–CO arrangement with anangle of 178◦, this appears to be unlikely. Furthermore, a morerecent structure for CcO from Bos taurus at 2.2 Å resolu-tion shows an Fe–C–O angle of 164◦, considerably closer tothe computed value (see Table II).69 The current results arealso supported by previous computational studies of heme-CO systems.11, 53, 59, 70, 71 Even at the mixed QM/MM level forMbCO in which the His-Heme-CO system was treated quan-tum mechanically and the rest of the protein was representedas point charges, a linear Fe–CO arrangement was found.31

The Fe–Cu distance during the equilibrium dynamics is4.98 ± 0.08 Å (see Figure 8). This compares with an averagedistance of 4.4 ± 0.34 Å from simulations for the unligatedprotein (see below) and 4.5 Å from the X-ray structure.23

Hence, the presence of the ligand in the active site pushesthe two metal centers apart and photodissociation of theligand allows the Fe- and Cu-atoms to move closer. Also,the comparison between simulations and experiment for the

FIG. 3. (a) PES based on DFT points (red) and fitted potential (black) for R = 5.2 Å. (b) The DFT energies (black circles) and the corresponding values of thefitted PES (red) (upper frame), and the absolute error of the potential for each of the points (lower frame) for R = 5.2 Å.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 6: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-5 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

TABLE II. Characteristic bond lengths and angles in the active site from X-ray structures of CO-bound CcOsand the present simulations.

Reference Fe–C [Å] Fe–C–O [deg] Cu–O [Å]

CcO (1AR1) (MD, current work) 1.93 ± 0.067 176 ± 2.04 2.39 ± 0.538CcO of Th. thermophilus (3QJQ, 2.9 Å)68 1.95 ± 0.04 126 ± 8 2.42 ± 0.11CcO of Bos taurus (1OCO, 2.8 Å)5 1.90 152 2.4CcO of Bos taurus (3AG1, 2.2 Å)69 1.85 164 2.72

unligated system are an additional validation for the forcefield employed in the present work.

C. Free CO dynamics

In order to better characterize the active site, simulationswere also carried out for photodissociated CO in the bimetal-lic active site. For this, a quadrupolar electrostatic model forthe free CO molecule was employed. This model describes theinteraction of CO with the protein better than a conventionalpoint charge model.11, 53 In particular, the molecular dipoleand quadrupole moments are correctly captured and thus fluc-tuate with CO bond length. This also allows to determine theCO-infrared spectrum from the time-varying dipole-momentautocorrelation function C(t) = 〈 μ(t)μ(0)〉.

The total simulation time for free CO in the bimetallicsite was 11 ns. This allowed to map out the shape and theeffective volume of the cavity around the active site to whichfree CO has access, as reported in Figure 4. In the free dynam-ics simulations, we find three possible locations to which COescapes after photodissociation: the distal pocket, the water-protein interface, and the “K-path” shown in Figure 4(b).Most of the time CO remains in the distal pocket. In one

simulation, CO moved towards the protein-water interfaceand in one it diffused into the “K-path” pocket. In cases whereCO left the distal pocket, the escape occurred on a timescaleof ≈ 0.5 ns.

During these simulations the heme a3 Fe changed froma hexacoordinated to a pentacoordinated state. Hence, the Femoves below the heme a3-plane. This coordination changecan be described by the distance between the plane definedby the four heme-nitrogen atoms and the Fe atom. In the orig-inal crystal structure, the heme-Fe separation is ≈ 0.42 Å be-low the plane (in the direction away from the distal pocket)which compares with 0.50 Å from the present simulations(Fig. 5) and 0.35 Å from previous work on myoglobin.59 Af-ter excitation, CO moves in the direction normal to the hemeplane. Hence, the CO dynamics right after photodissociationis rather translational than rotational. After 100–250 fs the COhits the protein and its dynamics is affected by the structureof the pocket of the active site.

As the major function of CcOs is proton transfer, theproton channels such as the K-path often contain cavitiessufficiently large to accept water molecules. The K-path isa proton uptake pathway, in which protons are transferredtowards heme a3 and heme a sites, and therefore essential

FIG. 4. Probability distributions for (a) Fe-bound CO and (b) photodissociated CO in the active site of CcO. Important residues in the active site are labelled.The small green and orange spheres are the heme-Fe and CuB atoms, respectively. In (b) several docking sites, including the “K-path” and distal site pocketscan be distinguished.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 7: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-6 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

FIG. 5. Out of plane motion of Fe in heme a3 during free CO dynamics.μ – expectation, σ – standard deviation.

for the function of the protein. In crystal structures of CcOsof other bacteria, e.g., Th. thermophilus,72 up to 3 watermolecules can be found in the area close to the “K-path”pocket.73 Hence, this pocket may be similar in character tothe well-known Xe-binding sites in CcO74 or myoglobin75, 76

which are also accessible to external ligands, including COand NO. It was therefore of particular interest to investigatethe dynamics of the free CO molecule inside the active siteof CcO, as experiments20, 68 do not give definitive answersabout the existence of free CO after photodissociation.

Infrared spectroscopy is suitable to address this point butmay lack the sensitivity and resolution for unambiguous de-tection. For a computational characterization, ten individual500 ps simulations of photodissociated CO in the active sitewere run. This time scale is appropriate to obtain sufficientlyresolved and converged spectra as recent work on NO in Mbhas shown.77 The simulations were started from Fe-boundCO, the repulsive potential to induce photodissociation wasactive for 100 fs after which the system was switched to the3A state. The average IR spectrum of the photodissociated,free CO molecule was calculated with a resolution better than1 cm−1 and is reported in Figure 6. For the IR spectra, the real-time dipole-dipole autocorrelation function C(t) = 〈μ(t)μ(0)〉is calculated and its Fourier transform C(ω) yields the IRspectrum according to

A(ω) ∝ ω(1 − exp(−¯ω/(kBT )))C(ω), (7)

where kB is the Boltzmann constant, ¯ is the Planck con-stant, and T is the temperature. The individual spectra differas they reflect different environments sampled by the ligandbut all of them are broad and centered around 2180 cm−1

which is the value for a gas-phase simulation of isolatedCO.53, 54 The difference to the 2143 cm−1 band known ex-perimentally originates from the finite step size used in theMD simulations and the classical treatment of the anhar-monic oscillator.54, 78, 79 Consequently, an insignificant shiftfor photodissociated CO in the active site of CcO is predictedrelative to gas-phase CO which contrasts with the situationencountered for free CO in the active sites of Myoglobinor Neuroglobin.8, 10, 11, 19, 53, 54, 57, 70, 77, 80 Figure 6 reports three

individual spectra (colored traces) together with the averageover all 10 trajectories (black). Of the 10 trajectories, 9 showa largely unshifted spectrum (similar to the red and greentraces in Figure 6) relative to gas-phase CO whereas one tra-jectory (blue trace) has a spectrum shifted to the red. How-ever, even detailed analysis of the atomic motions did notreveal any substantial differences between the configurationsthat lead to the blue and red/green spectral features. This isshown in Figure 3 in the supplementary material.89 Hence,the current simulations do not support the notion of pocket-specific spectra such as those found for diatomic ligands inMb or Ngb.10, 11, 19, 53, 54, 57, 77, 80

As anticipated from previous experiments,20 the expectedspectrum of photodissociated CO in the active site is broadwhich is consistent for the dynamics of a free diatomicmolecule in a highly inhomogeneous electric field and alargely open chemical environment: the pocket of the activesite is formed by both polar (the tri-histidine (His276, His325,His326) coordinated CuB, a Tyr280 with a hydrogen bond to thehydroxyl of the heme a3) and apolar groups (Val279, Ile315,Ile347). The pocket volume was analyzed using SURFNET.81

This program finds cavities by covering space lined by agiven collection of residues using spheres of different radii(Rmin = 1.0Å, Rmax = 3.0 Å), which define the volume ofthe pocket. For estimating the pocket volume relevant in thepresent work, 400 frames along a trajectory with photodis-sociated CO were analyzed. For this, the CO molecule wasremoved from the structure. The average size of the pocketis 105 ± 32 Å3, which is approximately three times largerthan the cavity of the distal pocket in myoglobin.82 This, to-gether with the low signal-to-noise ratio in the experimentscould partially explain the absence of the free CO signal inthe experiments.20

D. CO transfer following excitation

In a next step, the ligand transfer reaction was investi-gated. To this end, 1000 independent excitation trajectoriesfor each excitation time τ e were run and analyzed. Initially,the excitation was simulated by a repulsive term (Eq. (6))which was already used in the free dynamics. At defined timesafter the excitation – including τ e = 20, 30, 50, 100, 150, and200 fs (for each initial configuration) – the interactions wereswitched to the interaction potential Eq. (2) describing theground state PES supporting both, Fe–CO and Cu–CO-boundstates. For the heme a3, separate sets of simulations were car-ried out with the penta- and the hexacoordinated force fields.In the following, only the results with the pentacoordinatedforce field are described and comparisons with results fromsimulations with the hexacoordinated force field are madewhere it is appropriate. The supplementary material89 reportsthe results from simulations with the hexacoordinated forcefield.

The analysis of excitation trajectories shows that theprobability for ligand transfer to yield Cu–CO is only in-significantly affected by increasing the quench delay time (seeFigure 7). For all times τ e, the final state population isprominently peaked at the Cu–CO state which suggests that

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 8: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-7 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

FIG. 6. (a)–(c) Location of the free CO in the active site from 5 ns (10 × 500 ps) simulations ((a) xy-plane, (b) yz-plane, (c) xz-plane). The black bold solid linein panels (b) and (c) shows the approximate linear dimensions of the heme a3. The green and orange circles in panels (a)–(c) correspond to the positions of theheme a3 Fe and the CuB atom, respectively. (d) Averaged infrared spectra of the free CO in the active site of the CcO, calculated from ten 500-ps trajectories at300 K using the three-point fluctuating charge model (black bold line). The colored lines are from three representative trajectories, independently normalized.The dashed line corresponds to the position of the IR absorbance peak of the free CO in vacuum, calculated using the three-point fluctuating charge model.

following photodissociation from the heme-Fe the CO ligandis efficiently transferred to the copper atom. After photodis-sociation, the Fe atom moves out of the heme plane definedby the four heme nitrogens to a distance of ≈ 0.5 Å below theplane within the first ≈ 100 fs of the simulation. This has alsobeen found in previous work on Myoglobin59, 83 and the timefor breaking the Fe–C bond has been estimated to be on theorder of 100 fs.84 The Fe-out-of-plane displacement is closeto that in the experimental 1AR1 X-ray structure for pentaco-ordinated heme a3, where it is 0.42 Å.

In equilibrium simulations for Fe-bound CO, the Fe–Cu distance is typically around 5.0 Å (see Figure 8) whichincreases to 5.5 Å during the transfer within a picosecond.

This displacement is caused partially by the out-of-planemotion of the heme Fe. Therefore, the transfer leads to mod-erate but functionally important changes to the structure ofthe active site. In particular, the pentacoordinated heme-Feis “hidden” below the plane which makes it less availablefor reactions with external ligands. This is in contrast to thetetrahedrally coordinated Cu-ion which always presents anavailable ligand-binding site to free CO.

The analysis of the trajectories suggests that θ is a mean-ingful reaction coordinate to describe the transition betweenthe Fe–CO (θ ≈ 0) and Cu–CO (θ ≈ 150◦) states (as θ isthe angle between heme a3, the center of the CO and the Cof the CO, the minimum corresponding to the Cu–CO has a

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 9: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-8 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

FIG. 7. Probability of final states (after 5 ps) as a function of the quenchdelay time (based on 1000 trajectories). θ = 0 corresponds to Fe–CO,θ > 150◦ to Cu–CO.

value of 150◦ in the coordinate system used). The time seriesθ (t) obtained from the trajectories (Figure 9) show that COarrives at the CuB site typically within 1 ps. The transfer it-self consists of two phases – a ballistic and a diffusive one(Figure 10).

The kinetic curves show that there is a lag time for thetransfer reaction, which corresponds to the minimal time re-quired for the CO molecule to move towards the CuB site,followed by the rotation (Figure 10). In the simulations, thislag time is the sum of two components: the quench delay timeafter which the dynamics is driven by the repulsive potential(Eq. (6)) and the true ballistic time, when the potential de-scribing the ground PES is active (Eq. (2)). The estimated lagtime is not greater than 300 fs and the true ballistic time is≈ 100 fs (Figure S6 in the supplementary material89).

FIG. 8. Probability of Fe–Cu distances for simulations (5 ns) of Fe-boundCO (Fe–CO) (black; μ = 4.98 Å, σ = 0.08 Å) and for 1000 excitation sim-ulations showing the structural changes of the active site (red). The quenchdelay times τ e are reported in the panels.

To give a quantitative interpretation of the ligand trans-fer reaction, we further analyze the reactive trajectories.Similar to proton transfer reactions, it is useful to considergeometrical criteria with which to distinguish educt fromproduct states and to subsequently determine the sensitivity ofthe conclusions on the specific criteria chosen.85, 86 The geo-metrical criterion for ligand transfer is ρ ≥ 0.64 (CO is closerto CuB site) and θ ≥ 150◦, which are the average values of thecorresponding variables obtained from the final state analysis(ρCuB−CO ≥ 0.64 ± 0.05, θCuB−CO = 150 ± 5◦). Hence, if atrajectory matches this criterion, the transfer is consideredto be complete. In the following, the process of CO transferis treated as a first-order chemical reaction. For determiningthe rate of conversion from Fe–CO to Cu–CO, we use theconversion fraction (�) as a concentration term where � isthe fraction of the trajectories for which CO is transferred tothe CuB site at a particular time. Fitting the data (see insets inFigure 9) to a first order kinetics yields an estimated rateconstant of 3.83 ± 0.7 ps−1 and a characteristic time for theconversion to Cu–CO of ≈260 fs, which is somewhat morerapid than the value (450 fs) obtained from experiment whichis based on an exponential fit of the Cu–CO (2065 cm−1)differential transmission spectrum.20 To further quantifyconvergence of these results, bootstrapping was used.62, 87, 88

For this, only half of the data was randomly selected tocompute the observable and an associated error. This wasrepeated for 250 samples and yields a rate constant of3.73 ± 0.8 ps−1, in quite good agreement with the estimatedrate constant from using all the data.

For the hexacoordinated state in which the heme-Feremains closer to the heme plane, the Fe is still available forrebinding to the CO molecule. Hence, at very short quenchdelay times (τ e ≤ 100 fs) the probability of rebinding to theheme a3-Fe is high (see Figures S2– S5 in the supplemen-tary material89). The probability for ligand transfer to formCu–CO increases with quench delay time and reaches a max-imum of close to 100% (see Figure S3 in the supplementarymaterial89) for τ e > 100 fs and does not significantly changeeven up to 1 ps. After photodissociation, the Fe atom slightlymoves out of the heme plane in the similar fashion as in thepentacoordinated state but only to a distance of ≈ 0.10 Å be-low the heme plane within the first 20–100 fs of the simu-lation. Although the equilibrium structures of the penta- andhexacoordinated heme-group are quite different, the rates ofconversion from Fe–CO to Cu–CO are similar. The character-istic time increases from 260 fs to 300 fs when simulationsare carried out with the hexacoordinated heme force field.Hence, the conclusions about the time scale and nature ofthe rebinding reactions are robust with respect to force fieldparametrizations which characterize two extremes (penta-and hexacoordinated heme-Fe) of the physically relevantsituation.

E. Validation of the CO transfer protocol

The approach used so far assumes a certain timescale τ e

for the excited state dynamics. Beyond 100–200 fs – depend-ing on the coordination of the heme-Fe – the final state was

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 10: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-9 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

FIG. 9. CO transfer kinetics for different quench time delays – see labels in panels from analysis of 1000 trajectories. Main panels: black points are averages〈θ〉; grey overlapping lines are the individual trajectories, red dashed line shows the point where the quench has been performed. Insets: �Cu-CO is the conversionfraction to the Cu–CO state. At t = 0, �Cu-CO(t = 0) = 0. The slope of ln (1 − �) is the rate of forming Cu–CO, see text.

FIG. 10. 40 different reactive trajectories which lead to transfer for differ-ent quench time delays. The PES for a specific Fe–Cu separation of 5.2 Åis also reported for reference. Red lines – dynamics on the excited state sur-face (CO unbound), grey – dynamics on the ground state surface (CO boundto Cu). Green and orange areas show the approximate contour of the termi-nation criteria used to validate the approach for Fe–CO and Cu–CO statescorrespondingly, so that whenever the CO molecule is found within the areathe state is considered to be defined.

found to be insensitive to the particular time spent in the un-bound state. In order to independently validate this, additionalsimulations in the unbound state were carried out. Specifi-cally, 600 trajectories, each with a maximum simulation timeof 100 ps in length, were started from the bound state. TheFe–CO bond was broken and the repulsive potential was ac-tive for τ e = 50, 100, and 200 fs. Subsequently, the dynamicswas continued with the CHARMM standard force field for apentacoordinated heme a3 and the CO was treated with the3-point fluctuating charge model, analogous to the free COsimulations, i.e., without the explicit potential V (R, ρ, θ ) (seeEq. (2)).

The time of arrival at the CuB site or at the heme a3

was again determined when certain geometrical criteria weremet. They included a metal–CCO distance of less than 2.45 Åand a metal–C–O angle greater than 100◦. For comparison,the equilibrium Fe–C separation from the X-ray structure is1.9 Å and the Cu–C distance from DFT calculations is 1.87 Å.With these geometrical criteria, within 1.5 ps after photodis-sociation 85% (vs. 90% in the hexacoordinated state) of thesimulations have a defined state which is either CuB–CO orFea3 –CO. In the remaining cases, CO is still unbound andsamples the active site regions characterized earlier, see

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 11: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-10 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

Figure 4. Specifically, within 1.5 ps of breaking the Fe–Cbond in 512 (85.3%) simulations (out of a total of 600) theCO is bound to CuB, in 10 cases (1.7%) it rebinds to the Fea3

and for 78 (13%) simulations CO remains in a photodissoci-ated state. The kinetics analysis for the validation simulationsbased on the first order kinetics along the same lines as forthe reaction using the explicit reactive 3D-PES (see above)yield rate constants of 2.54 ps−1 (characteristic time 393 fs)and 3.11 ps−1 (characteristic time 321 fs) for penta- and hexa-coordinated heme-Fe, respectively. Within the limitations ofthe validation simulations this compares favorably with thesimulations on the reactive surface.

Similar to the case when simulations were run on thethree-dimensional PES suitable to explicitly describe bothbound states, Fe–CO and Cu–CO, additional simulations werecarried out in which the heme was treated with a hexacoordi-nate force field. Within the first 1.5 ps after the excitation 90%of the trajectories terminate in a metal-bound state (Figure S5in the supplementary material89). The probability of rebind-ing to the Fea3 is higher in this case as the iron atom is readilyaccessible to the CO ligand due to the smaller out-of planedistortion (0.1 Å vs. 0.5 Å). Hence, after 1.5 ps 484 (81%) ofthe simulations find the Cu–CO state, 59 (10%) terminate inFe–CO state, and in 57 (9%) case CO remains in a photodis-sociated state or rebind the metal sites on the longer timescale.

IV. CONCLUSIONS

In the present work, the dynamics of CO between theheme a3 and the CuB binding sites in CcO were studied basedon molecular dynamics simulations using an accurate three-dimensional reactive PES. Independent on the details of thesimulation protocols, the time scale for ligand transfer afterphotodissociation from the heme-Fe is on the sub-picosecondtime scale. Following photodissociation from the hemea3-Fe, approximately 90% of the ligands rebind on such timescales to either CuB (majority) or the heme a3 (minority).The data obtained from MD simulations and experimentaldata suggest a ballistic contribution to the transfer process.A time delay of ≈100 fs before rebinding starts is foundin the simulations. This qualitatively agrees with experimentwhere the time scales were determined by analyzing spectro-scopic signatures.20 The characteristic time scale for rebind-ing is ≈300 fs compared with 450 fs from experiment. Theremaining population (10%) consists of unbound CO sam-pling the active site which, however, are difficult to directlyobserve experimentally. Computational infrared spectroscopysuggests that the absorption features of unbound CO are broadand centered around the gas-phase value. All these observa-tions agree with experiments which were interpreted based onspectroscopic data.20

ACKNOWLEDGMENTS

This work was supported by the Swiss National ScienceFoundation (NSF(CH)) through Grant Nos. 200021-117810,and the NCCR MUST.

1S. Park and L. P. Pan, Biophys. J. 71, 1036 (1996).2J. A. Sigman, B. C. Kwok, and Y. Lu, J. Am. Chem. Soc. 122, 8192(2000).

3D. Lemon, M. Calhoun, R. Gennis, and W. Woodruff, Biochemistry 32,11953 (1993).

4P. A. Anfinrud, C. Han, and R. M. Hochstrasser, Proc. Natl. Acad. Sci.U.S.A. 86, 8387 (1989).

5S. Yoshikawa, K. Shinzawa-Itoh, R. Nakashima et al., Science 280, 1723(1998).

6E. S. Park, M. R. Thomas, and S. G. Boxer, J. Am. Chem. Soc. 122, 12297(2000).

7A. Ostermann, R. Waschipky, F. G. Parak, and G. U. Nienhaus, Nature(London) 404, 205 (2000).

8K. Nienhaus, S. Lutz, M. Meuwly, and G. U. Nienhaus, Chem. Eur. J. 19,3558 (2013).

9N. Plattner and M. Meuwly, Biophys. J. 102, 333 (2012).10F. Schotte, M. Lim, T. A. Jackson, A. V. Smirnov, J. Soman, J. S. Olson,

G. N. Phillips, M. Wulff, and P. A. Anfinrud, Science (New York, N.Y.)300, 1944 (2003).

11D. R. Nutt and M. Meuwly, Proc. Natl. Acad. Sci. U.S.A. 101, 5998 (2004).12M. Porrini, V. Daskalakis, S. C. Farantos, and C. Varotsis, J. Phys. Chem.

B 113, 12129 (2009).13U. Liebl, G. Lipowski, M. Négrerie, J. C. Lambry, J. L. Martin, and M. H.

Vos, Nature (London) 401, 181 (1999).14E. Olkhova, M. C. Hutter, M. A. Lill, V. Helms, and H. Michel, Biophys.

J. 86, 1873 (2004).15E. Pinakoulaki, T. Ohta, T. Soulimane, T. Kitagawa, and C. Varotsis, J. Biol.

Chem. 279, 22791 (2004).16D. Heitbrink, H. Sigurdson, C. Bolwien, P. Brzezinski, and J. Heberle, Bio-

phys. J. 82, 1 (2002).17I. Hofacker and K. Schulten, Proteins 30, 100 (1998).18K. M. Vogel, P. M. Kozlowski, M. Z. Zgierski, and T. G. Spiro, Inorg. Chim.

Acta 297, 11 (2000).19K. Nienhaus, S. Lutz, M. Meuwly, and G. U. Nienhaus, Chem. Phys. Chem.

11, 119 (2010).20J. Treuffet, K. J. Kubarych, J.-C. Lambry, E. Pilet, J.-B. Masson, J.-L. Mar-

tin, M. H. Vos, M. Joffre, and A. Alexandrou, Proc. Natl. Acad. Sci. U.S.A.104, 15705 (2007).

21B. R. Brooks, C. L. Brooks III, A. D. J. Mackerell, L. Nilsson et al., J.Comput. Chem. 30, 1545 (2009).

22A. J. MacKerell, M. Feig, and C. I. Brooks, J. Comput. Chem. 25, 1400(2004).

23C. Ostermeier, A. Harrenga, U. Ermler, and H. Michel, Proc. Natl. Acad.Sci. U.S.A. 94, 10547 (1997).

24V. R. Kaila, M. P. Johansson, D. Sundholm, L. Laakkonen, and M. Wik-ström, Biochem. Biophys. Acta - Bioenerg. 1787, 221 (2009).

25H. Li, A. D. Robertson, and J. H. Jensen, Proteins 61, 704 (2005).26D. C. Bas, D. M. Rogers, and J. H. Jensen, Proteins 73, 765 (2008).27M. H. Olsson, C. R. Søndergard, M. Rostkowski, and J. H. Jensen, J. Chem.

Theory Comput. 7, 525 (2011).28C. R. Søndergard, M. H. Olsson, M. Rostkowski, and J. H. Jensen, J. Chem.

Theory Comput. 7, 2284 (2011).29G. Kieseritzky and E. Knapp, Proteins 71, 1335 (2008).30B. Rabenstein and E. Knapp, Biophys. J. 80, 1141 (2001).31M. Meuwly, Chem. Phys. Chem. 7, 2061–2063 (2006).32K. Nienhaus, J. S. Olson, S. Franzen, and G. U. Nienhaus, J. Am. Chem.

Soc. 127, 40–41 (2005).33W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L.

Klein, J. Chem. Phys. 79, 926 (1983).34J. P. Ryckaert, G. Ciccotti, and H. J. C. Berendsen, J. Comput. Phys. 23,

327 (1977).35M. Yoneya, H. J. C. Berendsen, and K. Hirasawa, Mol. Simul. 13, 395

(1994).36A. D. MacKerell, Jr., D. Bashford, M. Bellott, J. R. L. Dunbrack, J. D.

Evanseck, M. J. Field, S. Fischer, J. Gao, H. Guo, S. Ha, D. Joseph-McCarthy, L. Kuchnir, K. Kuczera, F. T. K. Lau, C. Mattos, S. Michnick, T.Ngo, D. T. Nguyen, B. Prodhom, I. W. E. Reiher, B. Roux, M. Schlenkrich,J. C. Smith, R. Stote, J. Straub, M. Watanabe, J. Wiorkiewicz- Kuczera, D.Yin, and M. Karplus, J. Phys. Chem. B 102, 3586 (1998).

37J. W. Petrich, J. C. Lambry, K. Kuczera, M. Karplus, C. Poyart, and J. L.Martin, Biochemistry 30, 3975 (1991).

38A. D. Becke, J. Chem. Phys. 98, 5648 (1993).39C. Lee, W. Yang, and R. Parr, Phys. Rev. B 37, 785 (1988).40S. Vosko, L. Wilk, and M. Nusair, Can. J. Phys. 58, 1200 (1980).

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54

Page 12: CO-dynamics in the active site of cytochrome c oxidase · 2015. 2. 17. · 145101-2 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014) In order to better understand the atomistic

145101-11 M. Soloviov and M. Meuwly J. Chem. Phys. 140, 145101 (2014)

41P. Stephens, F. Devlin, C. Chabalowski, and M. Frisch, J. Chem. Phys. 98,11623 (1994).

42G. Petersson and M. A. Al-Laham, J. Chem. Phys. 94, 6081 (1991).43G. Petersson, A. Bennett, T. Tensfeldt, M. A. Al-Laham, W. A. Shirley, and

J. Mantzaris, J. Chem. Phys. 89, 2193 (1988).44M. J. Frisch, G. W. Trucks, H. B. Schlegel et al., Gaussian 09, Revision

A.1., Gaussian, Inc., Wallingford, CT, 2009.45M. Meuwly and J. M. Hutson, J. Chem. Phys. 110, 8338 (1999).46D. R. Nutt, M. Karplus, and M. Meuwly, J. Phys. Chem. B 109, 21118

(2005).47S. Lammers, S. Lutz, and M. Meuwly, J. Comput. Chem. 29, 1048 (2007).48J. Huang and M. Meuwly, J. Chem. Theory Comput. 6, 467 (2010).49S. G. Johnson, “The NLopt nonlinear-optimization package,” see

http://ab-initio.mit.edu/nlopt.50J. A. Nelder and R. Mead, Comput. J. 7, 308 (1965).51M. M. Law and J. M. Hutson, Comput. Phys. Commun. 102, 252 (1997).52M. Devereux and M. Meuwly, J. Chem. Inf. Model. 50, 349 (2010).53D. R. Nutt and M. Meuwly, Biophys. J. 85, 3612 (2003).54N. Plattner and M. Meuwly, Biophys. J. 94, 2505 (2008).55J. N. Huffaker, J. Chem. Phys. 64, 3175 (1976).56J. N. Huffaker, J. Chem. Phys. 64, 4564 (1976).57S. Lutz, K. Nienhaus, G. U. Nienhaus, and M. Meuwly, J. Phys. Chem. B

113, 15334 (2009).58M. Devereux and M. Meuwly, Biophys. J. 96, 4363 (2009).59M. Meuwly, O. M. Becker, R. Stote, and M. Karplus, Biophys. Chem. 98,

183 (2002).60E. R. Henry, M. Levitt, and W. A. Eaton, Proc. Natl. Acad. Sci. U.S.A. 82,

2034 (1985).61S. Lutz and M. Meuwly, Chem. Phys. Chem. 13, 305 (2012).62D. R. Nutt and M. Meuwly, Biophys. J. 90, 1191 (2006).63J. Danielsson and M. Meuwly, J. Chem. Theory Comput. 4, 1083 (2008).64P. Zhang, S. W. Ahn, and J. E. Straub, J. Phys. Chem. B 117, 7190 (2013).65T. Nagy, J. Y. Reyes, and M. Meuwly, “Multisurface Adiabatic Reactive

Molecular Dynamics,” J. Chem. Theory Comput. (published online).66P. Dayal, S. A. Weyand, J. McNeish, and N. Mosey, J. Chem. Phys. Lett.

516, 263 (2011).67A. Warshel and R. M. Weiss, J. Am. Chem. Soc. 102, 6218 (1980).

68B. Liu, Y. Zhang, J. T. Sage, S. M. Soltis, T. Doukov, Y. Chen, C. D. Stout,and J. A. Fee, Biochem. Biophys. Acta - Bioenerg. 1817, 658 (2012).

69K. Muramoto, K. Ohta, K. Shinzawa-Itoh, K. Kanda, M. Taniguchi, H.Nabekura, E. Yamashita, T. Tsukihara, and S. Yoshikawa, Proc. Natl. Acad.Sci. U.S.A. 107, 7740 (2010).

70D. R. Nutt and M. Meuwly, Chem. Phys. Chem. 5, 1710 (2004).71N. Strickland and J. N. Harvey, J. Phys. Chem. B 111, 841 (2007).72T. Tiefenbrunn, W. Liu, Y. Chen, V. Katritch, C. D. Stout, J. A. Fee, and V.

Cherezov, PLoS ONE 6, e22348 (2011).73L. Buhrow, S. Ferguson-Miller, and L. Kuhn, Biophys. J. 102, 2158

(2012).74V. M. Luna, J. A. Fee, A. A. Deniz, and C. D. Stout, Biochemistry 51, 4669

(2012).75H. Frauenfelder, B. H. McMahon, R. H. Austin, K. Chu, and J. T. Groves,

Proc. Natl. Acad. Sci. U.S.A. 98, 2370 (2001).76R. F. Tilton, I. D. Kuntz, and G. A. Petsko, Biochemistry 23, 2849 (1984).77M. W. Lee and M. Meuwly, J. Phys. Chem. B 116, 4154 (2012).78P. Berens and K. Wilson, J. Chem. Phys. 74, 4872 (1981).79J. Danielsson and M. Meuwly, J. Phys. Chem. B 111, 218 (2007).80M. Lim, T. A. Jackson, and P. A. Anfinrud, J. Chem. Phys. 102, 4355

(1995).81R. A. Laskowski, J. Mol. Graph. 13, 323 (1995).82C. Bossa, M. Anselmi, D. Roccatano, A. Amadei, B. Vallone, M. Brunori,

and A. Di Nola, Biophys. J. 86, 3855 (2004).83J. Vojtechovsky, K. Chu, J. Berendzen, R. B. Sweet, and I. Schlichting,

Biophys. J. 77, 2153 (1999).84P. Stoutland, J. Lambry, J. Martin, and W. Woodruff, J. Phys. Chem. 95,

6406 (1991).85E. Helfand, J. Chem. Phys. 69, 1010 (1978).86M. Meuwly and M. Karplus, J. Chem. Phys. 116, 2572 (2002).87B. Efron, Ann. Stat. 7, 1–26 (1979).88S. Nangia, A. W. Jasper, T. F. Miller III, and D. G. Truhlar, J. Chem. Phys.

120, 3586–3597 (2004).89See supplementary material at http://dx.doi.org/10.1063/1.4870264 for

comparisons of the unligated forms of CcO, analysis of the trajectories ofunbound CO and the rebinding kinetics for the heme in its hexacoordinatedgeometry.

This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to IP:

192.33.103.194 On: Tue, 17 Feb 2015 14:19:54