13
Biophysical Chemistry 106 (2003) 111–123 0301-4622/03/$ - see front matter 2003 Elsevier B.V. All rights reserved. doi:10.1016/S0301-4622(03)00156-X MD simulation of a plastocyanin mutant adsorbed onto a gold surface Anna Rita Bizzarri*, Giulio Costantini, Salvatore Cannistraro 1 UnitaINFM, Dipartimento di Scienze Ambientali, Universitadella Tuscia, Viterbo I-01100, Italy Received 14 February 2003; received in revised form 23 May 2003; accepted 28 May 2003 Abstract MD simulation of plastocyanin, an electron transfer protein, adsorbed onto a gold surface, has been performed for 10 ns. Starting from the crystallographic structure of a poplar plastocyanin mutant engineered with the insertion of a disulfide bridge, the protein has been anchored to a gold substrate modeled by a cluster of three layers in the AuN111M configuration. A number of significant structural and dynamical properties of the protein molecule, covalently bound through either one or two sulfur atoms to the gold surface, has been extracted and compared with those of the free protein. Attention has been paid to investigate the dynamical aspects putatively related to the electron transfer process. In particular, the cross-correlation function between specific active site vibrations and all the other protein atom motions and the principal component analysis have been calculated in order to put into evidence dynamical correlation of some functional relevance. The results are discussed also in connection with related experiments. 2003 Elsevier B.V. All rights reserved. Keywords: MD simulation; Hybrid electron transfer systems; Bioelectronic devices 1. Introduction Recently, electron transfer (ET) proteins, chemi- sorbed onto gold substrates, have attracted large attention even for their possible application in biomolecule-based electronic components w1–3x. The strengths associated with utilizing biomole- cules in such devices reside largely in their capac- ity for self-assembly, specific surface recognition and a structure designed for efficient electron *Corresponding author. Tel.: q39-0761-357027; fax: q39- 0761-357119. E-mail address: [email protected] (A.R. Bizzarri). http:yywww.unitus.it ybiophysicsy 1 transfer. Native, or suitably engineered, thiol groups and disulfide bonds are mainly used to achieve a stable immobilization as well as exploit- able ET pathways w4x. Among different experi- mental approaches, scanning tunneling microscopy (STM) provides new progress in studying the ET process of these hybrid systems w5–7x. Azurin (AZ), a protein agent in redox systems, has been one of the most studied copper ET proteins, also in the perspective of building hybrid systems, in which the ET properties could be exploited to transduce chemical signals w8–13x. With the same aim, more recently, a mutant of poplar plastocyanin (PC), an ET protein similar to AZ, has been engineered by introducing a surface disulfide

MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

Biophysical Chemistry 106(2003) 111–123

0301-4622/03/$ - see front matter� 2003 Elsevier B.V. All rights reserved.doi:10.1016/S0301-4622(03)00156-X

MD simulation of a plastocyanin mutant adsorbed onto a goldsurface

Anna Rita Bizzarri*, Giulio Costantini, Salvatore Cannistraro1

Unita’ INFM, Dipartimento di Scienze Ambientali, Universita’ della Tuscia, Viterbo I-01100, Italy

Received 14 February 2003; received in revised form 23 May 2003; accepted 28 May 2003

Abstract

MD simulation of plastocyanin, an electron transfer protein, adsorbed onto a gold surface, has been performed for10 ns. Starting from the crystallographic structure of a poplar plastocyanin mutant engineered with the insertion of adisulfide bridge, the protein has been anchored to a gold substrate modeled by a cluster of three layers in the AuN111Mconfiguration. A number of significant structural and dynamical properties of the protein molecule, covalently boundthrough either one or two sulfur atoms to the gold surface, has been extracted and compared with those of the freeprotein. Attention has been paid to investigate the dynamical aspects putatively related to the electron transfer process.In particular, the cross-correlation function between specific active site vibrations and all the other protein atommotions and the principal component analysis have been calculated in order to put into evidence dynamical correlationof some functional relevance. The results are discussed also in connection with related experiments.� 2003 Elsevier B.V. All rights reserved.

Keywords: MD simulation; Hybrid electron transfer systems; Bioelectronic devices

1. Introduction

Recently, electron transfer(ET) proteins, chemi-sorbed onto gold substrates, have attracted largeattention even for their possible application inbiomolecule-based electronic componentsw1–3x.The strengths associated with utilizing biomole-cules in such devices reside largely in their capac-ity for self-assembly, specific surface recognitionand a structure designed for efficient electron

*Corresponding author. Tel.:q39-0761-357027; fax:q39-0761-357119.

E-mail address: [email protected](A.R. Bizzarri).http:yywww.unitus.itybiophysicsy1

transfer. Native, or suitably engineered, thiolgroups and disulfide bonds are mainly used toachieve a stable immobilization as well as exploit-able ET pathwaysw4x. Among different experi-mental approaches, scanning tunneling microscopy(STM) provides new progress in studying the ETprocess of these hybrid systemsw5–7x. Azurin(AZ), a protein agent in redox systems, has beenone of the most studied copper ET proteins, alsoin the perspective of building hybrid systems, inwhich the ET properties could be exploited totransduce chemical signalsw8–13x. With the sameaim, more recently, a mutant of poplar plastocyanin(PC), an ET protein similar to AZ, has beenengineered by introducing a surface disulfide

Page 2: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

112 A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 1. Schematic representation of PCSS anchored onto aAuN111M. Each of the two protein sulfurs is engaged in cova-lent bonds with three gold atoms(see text).

bridge opposite to the active sitew14x. It has beenshown that the resulting mutated structure(PCSS)retains the spectroscopic features of the wild typestructure and likely, the ET propertiesw14,15x.PCSS covalently binds to a gold surface throughthe disulfide bridge(see Fig. 1) and STM imagesof immobilized PCSS seem to indicate that thecopper active site is involved in the ET process.Furthermore, scanning tunneling spectroscopy atthe level of single molecule has revealed anintriguing rectifying-like behavior; such a peculiarmolecular conduction having been detected also inair and deserving possible applications in hybriddevicesw16x.We aimed, therefore, at modeling the adsorption

and the interaction of the mutated PCSS at a goldsurface to get more insights about the structureand dynamical features of this immobilized systemand their possible implications with the ET pro-cess. In this connection, it should be stressed thatwhile self assembled monolayers of thiol contain-ing small organic molecules on gold surfacesw4,17–21x have been widely studied, by ab initioand MD simulation approaches, no similar infor-mation is available, to the best of our knowledge,about proteins. Only a Brownian dynamics studyhas been reported on a physically adsorbed proteinw22x. However, MD simulations have emerged asuseful means, complementary to various experi-mental techniques, to provide structural anddynamical properties of protein systems. In partic-ular, MD simulations have been widely employed

by some of the authors to investigate the structural,dynamical and ET properties of wild PCw23–25x;the effect of introducing a disulfide linkage-bondin poplar PC has also been studiedw26x.Here, we report MD simulations of PCSS

adsorbed onto a gold substrate, which have beenperformed after having covalently anchored theprotein either with one or two sulfurs to the goldsurface. As the first stage of analysis, we havefocussed our attention to dry systems. In thisrespect, it should be mentioned that many STMand atomic force microscopy(AFM) studies onprotein systems adsorbed onto metal substrates areperformed, in air or even in vacuo, on dry samplesw27,28x.The potential energy, the root mean square

displacements(RMSD) and the gyration ratio fromthe initial structure of both the adsorbed systemshave been monitored during 10 ns of MD simula-tion and compared with those of free PCSS. Theorientation and precession of the protein macro-molecule about the normal to the gold surfacehave been investigated during the MD simulatedtrajectories for both of the differently anchoredsystems. Additionally, the height of the macromol-ecule with respect to the gold substrate has beencalculated for a set of trajectories and the corre-sponding histograms have been extracted. Finally,the cross-correlation function between the Cu–S(Cys84) bond distance and each protein atommotions and the principal component analysis(PCA) have been performed along the MD trajec-tory. This analysis shows a stability of both theanchored systems up to 10 ns. In addition, itreveals the existence of a dynamical correlationbetween the copper active site and protein regionsfar from it, including the anchoring region. Theseresults are also discussed in connection with relat-ed experiments.

2. Computational methods

MD simulations have been performed by usingthe CHARMM packagew29x including a standardmolecular mechanics potential form. The localinteractions include harmonic bonds, angles,improper torsions and dihedral angles. The long-range interactions include Lennard–Jones potential

Page 3: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

113A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

and electrostatic potential between atom-basedchargesw29x:

2 2Vs k (ryr ) q k (uyu )r 0 0 08 8bonds angles

2q k (vyv ) q kv 0 f8 8impropers dihedrals

w x= 1qcos(nfyf )0B EA Bij ijC Fq y8 12 6D Gr rij ijnon-bonded pairs

q qi jq (1)8 rijnon-bonded pairs

Parameters for protein atoms are taken fromCHARMM27 force field w30x. Initial coordinatesof PCSS have been taken from the X-ray structure,at 0.16 nm resolution, of the mutant poplar PCbearing a disulfide bridge introduced by replacingIle21 and Glu25 by two cysteines(1JXG entry ofBrookhaven Protein Data Bank) w15x.PC is an eight-stranded anti-parallelb-barrel

macromolecule with a copper atom liganded bythe side-chains of two histidine(His37 and His87),a cysteine(Cys84) and a methionine(Met92) ina peculiar, tetrahedral geometryw31x (see Fig. 1).With the exception of the copper ligands His37,Cys84, His87, all the ionizable residues areassumed to be in their fully charged state, accord-ing to the pH value. As in previous works, acovalent bond approach for the copper site hasbeen adoptedw23,32,33x. In particular, the copperhas been considered to be bound to three ligands(two nitrogens from His37 and His87, respectively,and one sulfur from Cys84), while the muchweaker interaction of copper with the sulfur fromMet92 has been treated by a non-bonded approachw34x. The partial charges for the copper and itsligand residues have been modified by a linearinterpolation of the standard CHARMM chargesbetween the protonated and the un-protonated res-idues to match the redistribution of the charges inthe residues as due to the reduction of the coppercharge fromq2 toq0.98, by following Ref.w33x;such kind of approach having allowed us to relia-bly reproduce the experimental main vibrationalfeatures of PCw32x. Additionally, we remark thatthese charges are rather similar to those recentlyobtained by quantum calculationsw35,36x.

Starting from the crystallographic structure, toappropriately release the conformational stressesand the steric constraints as due to the presence ofthe gold substrate, the minimization of the PCSShas been carried out with a combination of threedifferent algorithms: 500 steps with the steepestdescent, 500 steps with the conjugate gradient andfinally, 1000 steps with the adopted basis Newton–Raphson minimization. An equilibration procedurehas been then performed by rising the temperatureof the system from 0 K to 300 K with incrementsof 5 K every 0.2 ps, by assigning the atomicvelocities according to a Maxwellian distribution.Subsequently, a MD simulation at 300 K has beencarried out for 200 ps with a Gaussian velocityreassignment when the temperature deviated ofmore than 10 K from the assigned temperature.After that, the dynamics of the system have

started by sampling at intervals of 0.1 ps for 10ns; a careful scaling down of the velocity duringthe simulations having been performedw37,38x.The Shake constraint algorithmw39x has been usedfor all hydrogens. A cut-off ratio of 11 and 13 A˚has been used for the Lennard–Jones and theelectrostatic energy functions, respectively. TheMD simulations of free PCSS have been conductedin the micro canonical ensemble(NVE). In orderto assess a good exploration of the configurationin the phase space, a scaling down of the velocityhas been performed during the simulation. Tomodel the Au(111) surface, the gold atoms havebeen arranged hexagonally into a cluster of threelayers each one with 22=25 atoms; the charge ofthe gold atoms has been put to zero. The nearestneighbor distance has been assumed to be 2.88 A˚and the positions of the gold atoms have beenfixed during all the simulations.According to experiments, PCSS is expected to

be covalently anchored to the gold surface byinvolving the sulfurs of the engineered S–S bridge(see Fig. 1) w14x. Since one or both sulfurs mightbe involved in the anchoring of the protein to thesubstrate, two different systems have been inves-tigated. First, to increase the external accessibilityof the sulfurs, monitored by the Lee and Richardsalgorithm w40x, the disulfide bridge has been bro-ken in agreement with the experimental evidencepointing out that chemisorption occurs after the

Page 4: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

114 A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Table 1Force field parameters describing the interactions betweenPCSS and the gold surfacew43x. These interaction parametersare added to the CHARMM27 all-atoms interactions

Harmonic bond interactionr (A)0

˚ k (kcalymol A )2r˚

Au–S 2.531 198S–C 1.836 205

Harmonic angle interactionu (deg)0 k (kcalymol)u

Au–S–C 109 46.347

Dihedral angle interactionf (deg)0 k (kcalymol)f n*

Au–S–C–C 180 0.31 2S–C–C–C y19 0.22 2

Lennard–Jones interaction**

r (A)min˚ E (kcalymol)min

Au 2.0736 0.078

multiplicity of dihedral angles.*

see ref.w43x.**

S–S bond dissociationw4,41x. After a minimiza-tion, an accessible surface of approximately 2 A2˚for Cys25 sulfur has been obtained. Subsequently,a minimization of the protein approaching the goldsurface has been carried out. Protein–gold inter-actions have been described by Lennard–Jonespotentials; with the interaction parameters beingreported in Table 1. At the end of this furtherminimization, the Cys25 sulfur exhibits an acces-sibility of approximately 17 A. A covalent bond2˚between this sulfur and the gold surface has beenthen established. In particular, the sulfur atom hasbeen bound with three-fold fcc hollow site goldatoms, placed at the center of the area surface inagreement with Ref.w42x; such a geometricalarrangement having been shown to be energeticallymore favorablew19x. The second sulfur of thebridge has been saturated with a hydrogen atomas a standard cysteine; such a configuration havinglabeled as PCSS-I. The parameters used to describethe gold protein interactions are reported in Table1 w43x.In the configuration labeled as PCSS-II, both

Cys25 and Cys21 sulfurs have been covalentlyanchored to the gold substrate. Starting from thePCSS-I configuration, to attach the second sulfur

(Cys21) to the gold surface, the disulfide bridgehas been re-established and the system has beenminimized. An accessible surface of 14 and 21A for Cys21 and Cys25 sulfurs, respectively, has2˚been obtained. After the breaking of the disulfidebridge, the Cys21 sulfur has been anchored at afcc three-fold hollow site of the gold. A schematicpicture of the PCSS bound to the gold surface bytwo covalent bonds is shown in Fig. 1. Thisconfiguration will be referred to as PCSS-II in thefollowing.For both the systems, a minimization and an

equilibration procedure has been started, followedby 10 ns of data collection(simulation PCSS-Iand simulation PCSS-II). To identify the correlatedatomic motions, the MD trajectories of PCSS,PCSS-I and PCSS-II have been analyzed accordingto the PCA or essential dynamicsw44,45x. Thisapproach aims at identifying a new reference framesuch that only a subset of coordinates, usually onthe order of a few percent of the total, is sufficientto describe the overall dynamics of the system.The remaining degrees of freedom, correspondingto constrained harmonic oscillations, can beneglected. Such a method is based on the diagon-alisation of the covariance matrix, built from theatomic fluctuations in a MD trajectory from whichoverall translational and rotational motions havebeen removed:

C sN(XyX )(X yX )M (2)ij i i,0 j j,0

where X are thex-, y- and z-coordinates of theatoms fluctuating around their average positionsX and whereN...M denotes an average over time.0

To construct the protein covariance matrices, wehave used theC atom trajectories, including thea

copper atoms. Upon diagonalisation of the covar-iance matrix, a set of eigenvalues and eigenvectorsis obtained. The eigenvectors of the covariancematrix correspond to directions in a 3N dimension-al space(where N is the number ofC atomsa

including the copper atom) and motions alongsingle eigenvectors correspond to concerted fluc-tuations of atoms. The eigenvalues represent thetotal mean square fluctuation of the system alongthe corresponding eigenvectors. If the eigenvectorsare ordered according to decreasing eigenvalues,

Page 5: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

115A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 2. Temporal evolution of the total potential energy, alonga 10 ns MD trajectory for the PCSS-I(circles) and PCSS-II(triangles) systems. Inset: temporal evolution of the totalpotential energy for free PCSS.

Table 2Average value, standard deviation and drift, expressed in kcalmol ns , of total potential energy and of the Lennard–Jonesy1 y1

components for PCSS, PCSS-I and PCSS-II, as extracted froma 10 ns trajectory. The drift has been calculated from a linearregression over the total trajectory

Mean Standard deviation Drift

Total potential energy(Kcalymol)PCSS y4704 53 0.013PCSS-I y4810 54 0.015PCSS-II y4814 44 0.012

Lennard–Jones energyPCSS y71 27 0.004PCSS-I y16 18 y0.0012PCSS-II y59 26 0.004

Au-PCSS Lennard–Jones energyPCSS-I y8 0.9 y6=10y5

PCSS-II y12 1.2 0.0002

the first one describes the largest scale correlatedmotions, whereas the last one will correspond tosmall-amplitude vibrationsw44x.

3. Results and discussion

The potential energy trend observed during 10ns of MD simulation of both the adsorbed PCSS-I and PCSS-II systems are shown in Fig. 2. Forcomparison, that of free PCSS is also plotted(insetof Fig. 2).After a short initial time, the potential energies

of free PCSS, PCSS-I and PCSS-II appear to bepractically constant in time with a small drift(seeTable 2); the energy of free PCSS appearing to beslightly higher than that of the bound proteinsystems. Such a trend seems to indicate that theadsorption of the protein to the gold substrate doesnot substantially affect the stability of the system.From Table 2, it is inferred that high values for

the standard deviation are also detected for thetotal potential energy of PCSS-II and PCSS-I. Aslong as the spread of the potential energy reflectsthe sampling of the conformational substates andlocal minima of the protein energy landscapew46–48x, such a behavior is indicative of the fact thatan exploration of the energy landscape also occursfor the immobilized systems. In addition, theslightly lower value for the standard deviation ofPCSS-II with respect to PCSS-I, suggests that the

binding of the protein to the gold surface by twosulfurs might yield some restriction in the energylandscape region accessible to the protein duringits dynamical evolution.An analysis of the Lennard–Jones energy, sep-

arately reported in Table 2, shows an increase ofthis energy component, with respect to free PCSS,for both the immobilized systems; such an effectbeing particularly evident for PCSS-I. Moreover,since the Lennard–Jones energy is essentiallymade up by protein atom–atom and protein–atomgold–atom interactions, the contribution as due tothe protein–gold interaction has been extracted.We note from Table 2 that the Au–PCSS Lennard–Jones energy constitutes a significant part of thetotal Lennard–Jones energy for both the immobi-lized systems; such a contribution to the totalLennard–Jones energy is more significant inPCSS-I than in PCSS-II.To get further insight into the effect of gold-

adsorption on the structural and dynamical featureof the protein, we have calculated the RMSD fromthe equilibrated structure. The RSMDs of all theprotein atoms as a function of time are plotted inFig. 3 for both PCSS-I and PCSS-II, and for freePCSS in inset of Fig. 3. During 10 ns of MDsimulation run, the RMSDs of both the PCSSsystems anchored to the gold substrate undergo afew jumps. In particular, the most significant jumps

Page 6: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

116 A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 3. RMSD from the starting structure of PCSS-I(line andcircles) and PCSS-II(line and triangles) along a 10 ns MDtrajectory. Inset: RMSD of the free PCSS. The RMSD havebeen obtained by averaging over all the protein atoms.

Fig. 4. Main-chainC traces of the PCSS-II as obtained froma

the time-averaged structure over 2.0–2.5 ns(light grey) and4.0–4.5 ns time interval(dark grey). The T1 and T8 turns, theb-strand S4 and thea-helix are the region where the mostsignificant differences between the two structures are detected.The copper atom and the Cys21 and Cys25 sulfur are shown.

are observed at approximately 3.1 ns, for PCSS-I,and at 3.0 ns and 7.2 ns for PCSS-II, respectively,and at approximately 2.1 ns for free PCSS. Inorder to investigate if these jumps arise fromconformational changes, we have analyzed theprotein conformation before and after these tran-sitions. The most relevant changes appear at 2.5–4.0 ns jump of PCSS-II. Indeed, Fig. 4, showinga superposition between the structure averagedover a 0.5 ns time interval before and after thejump detected for the PCSS-II system, revealssmall changes in correspondence of the turns T1and T8, of the S1 strand and of thea-helix.However, the main protein structure is preserved;a similar behavior occurring for other jumps.A comparison between the main-chainC tracesa

of PCSS-I, PCSS-II and of free PCSS is plotted inFig. 5. The structures have been obtained byaveraging all the atom positions over the entiresimulation. Generally, we note that the overallprotein structure is substantially preserved uponadsorbing the macromolecule onto the gold sub-strate, though some small structural changes canbe put into evidence. At a visual inspection, themost significant differences are detected at theturns T1, T3, T5 and T8 and at theb-strand S4for PCSS-I. The changes at T3 and S4, close tothe anchored protein sulfur, are reasonably due tothe direct involvement of these regions into thebinding to the gold atoms. Concerning PCSS-II,

we again observe changes at T3, T5, T8, S4 and,in addition, at theb-strand S3. The region closerto the gold anchorage site exhibits larger confor-mational changes in comparison with PCSS-I. Thispoints out that the presence of two anchoringpoints can result in more relevant structural anddynamical constraints on the protein structure.Furthermore, small structural changes at the copperactive site can be also detected for both PCSS-Iand PCSS-II.Additional information about the protein confor-

mation evolution during the dynamics can beobtained by analyzing the gyration ratioR : valuesg

for R of (12.395"0.044)A for the free PCSS,g˚

and of (12.373"0.012)A and (12.547"0.031)A˚ ˚for PCSS-I and PCSS-II, respectively, by averagingover 10 ns. According to our previous work, thegyration ratio of free PCSS is not significantlydifferent from that of wild PCw26x. Moreover, itturns out that PCSS-I shows aR value practicallyg

Page 7: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

117A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 5. Main-chainC traces, time-averaged over 10 ns, of free PCSS(dark grey), of PCSS-Iwlight grey (a)x and PCSS-IIwlighta

gray(b)x structures, respectively. The turns(T1, T3, T5 and T8) andb-strands(S3 and S4) are the regions where the most significantchanges are observed. The copper atom and the Cys21 and Cys25 sulfurs are shown.

equal to that of free PCSS; a slightly higher valueappearing for PCSS-II.An interesting aspect, which might be relevant

in connection with some experimental results, isthe overall orientation of the macromolecule withrespect to the gold substrate. Actually, topologicalinformation as obtained by AFM, on immobilizedprotein systems are related to the orientation andstructural arrangements of the macromolecular atthe gold surfacew13,49x. At the same time, sincethe copper active site plays a crucial role in themechanisms regulating the ET processw34x, theposition of the copper and the related electronlevels are expected to be relevant for the protein–gold ET paths in these hybrid systems; theseaspects having been investigated also by STMw11,12x. On such a ground, we have described theprotein orientation with respect to the gold surfaceby u andf angles as depicted in Fig. 6:u is the

angle that the axis , joining the sulfur of Cys25™pand the copper atom, forms with the normal to the

gold surface;f is the precession of about the™pgold surface normal. The trend with time ofu and

f, calculated during the MD trajectory, are plotted,for both the PCSS-I and PCSS-II systems, in Fig.6a and Fig. 6b, respectively. We note that in both

systems a significant deviation of from the™pnormal axis is registered. In addition, the meanvalue ofu is smaller for PCSS I than for PCSS-II

(see legend of Fig. 6); this means that the axis™pis closer to the normal for the protein systemanchored by a single bond. Such a behavior couldfind a correspondence with the lower Lennard–Jones energy calculated in this case and might beindicative of a less extensive contact between theprotein atoms and the gold surface. However,anchoring the protein with both sulfurs drives the

axis closer to the gold surface plane(higheru).™pConcerningf, PCSS-I and PCSS-II show rather

similar average values. A much larger variabilityis observed for the PCSS-I in comparison withPCSS-II (see legend of Fig. 6). In addition, it isinteresting to note the presence, in both the sys-tems, of fast oscillations to which slower oscilla-tions are overimposed; the origin of such an effectbeing not clear.

Page 8: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

118 A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 6. Temporal evolution, along a 10 ns MD trajectory, of theu (a) andf (b) angles describing the orientation and the pre-

cession, respectively, of the protein axis with respect to gold™psurface normal, for the PCSS-I(dark gray) and PCSS-II(lightgray). The average value and the standard deviation ofu are(27"4)8 and (36"3)8 for PCSS-I and PCSS-II, respectively.The average values and the standard deviations off are(112"12)8 and (105"6)8 for PCSS-I and PCSS-II, respec-tively. The inset depicts theu andf angles.

Fig. 7. Statistical analysis of the molecular height above theAu(111) substrate, as extracted from the MD simulated trajec-tories for PCSS-I and PCSS-II. The vertical dimension of pro-teins has been estimated from 1000 molecules sampled every10 ps of the MD simulated trajectory.

These results can provide some hints in theinterpretation of the experimental topological data,as obtained by AFM, of immobilized proteins. Inthis connection, the protein is represented by anellipsoid centered at the center of mass of theprotein and oriented as its main inertia axes(asan example see inset of Fig. 6). The ellipsoid,evaluated from the MD simulated trajectories,gives an overall representation of the macromole-cule thus providing an appropriate description ofthe protein height in order to be compared with

the related data from topological AFMw13,50x.The ellipsoid and its corresponding height overthe gold substrate have been extracted from theMD simulated trajectories by sampling structuresevery 10 ps. The histograms of the heights, asobtained by 1000 structures, are shown in Fig. 7.From this figure, it shows that a single modedistribution is registered for both systems. Themean values are, in both cases, close to thatexpected for the crystallographic data by assumingthat protein is anchored to gold via the disulfidebridge(h ; 3 nm). However, we note that PCSS-I is characterized by a mean value slightly higherthan that of PCSS-II(see Fig. 7). Such a findingis consistent with the previous results, showingthat the protein anchored by a single bond canassume, on average, a configuration closer to thenormal to the surface. For both systems, a spreadof heights is registered(see Fig. 7). Generally, thismeans that the macromolecule dynamically sam-ples its accessible conformations; assuming a vari-ety of arrangements, and then of heights, withrespect to the gold substrate. However, we notethat PCSS-I is characterized by a wider distributionin comparison with PCSS-II. As a consequence,molecules covalently bound to gold through onesulfur atom show slightly higher flexibility thanmolecules anchored via two covalent bonds. By

Page 9: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

119A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 8. Cross correlationC as calculated from Eq.(2), forx,i

free PCSS(continuous line), PCSS-I(a) (dashed line and cir-cles) and PCSS-II(b) (dashed line and triangles), averagedover 10 ns MD simulation trajectory, plotted vs. backboneatoms (N, C , C and O). Some of the structural elementsa

involved in the ET process are shown.

comparing these results, we note that the heightdistribution, as extracted by MD simulation, resultsmarkedly narrower than that derived by AFMw50x.This finding can be explained by taking intoaccount the stress exerted by the tip on the biom-olecule during the AFM measurements. Indeed,even in tapping mode(in which the tip is expectednot to be in a direct contact with the sample),some kind of interaction between the tip and thebiomolecule can occur. However, AFM is sensitiveto protein movements on a much longer time scale(at least milliseconds if we refer to tapping modewith a frequency in the range of kilohertz). Addi-tionally, we remark that AFM data refer to acollection of molecules, likely in different startingarrangements and anchored to gold either by oneor both sulfur atoms, whereas, MD simulationsconsider only an isolated molecule in two differentanchoring situations. Nevertheless, the present MDsimulations suggest that the protein, even if sam-pled in a restricted temporal window, can assumea variety of orientations with respect to the goldsubstrate.It is well-ascertained that the ET process occur-

ring in a protein system, results to be coupled tosome nuclear motions involving atoms far fromthe active sitew51,52x. Moreover, vibrations of ETproteins are of main concern for the functionalitysince they operate a fine-tuning of the redoxpotential and hence of the electron transfer processw51x. In this connection, the study of the dynamicalbehavior of PCSS, anchored to a gold substrate, isof utmost relevance to investigate if the proteinretains its native dynamical and ET functionalproperties. In wild PC, it has been shown thatsome macromolecular motions, even far from theactive site, along putative ET paths, are stronglycorrelated to the fluctuations of Cu–S(Cys84)bond distancew33x. Notably, the power spectra asextracted from the temporal fluctuations of such abond distance during a dynamical evolution areclosely similar to those detected in the resonantRaman spectraw32,33x.On such a ground, and in order to put into

evidence the coupling between protein motionsand the vibrations of the active site, relevant tothe ET process, we have performed an analysis ofthe cross-correlation function of the Cartesian

motions of each protein atom with the Cu–S(Cys84) bond distancex w33x:

N M N Mr y r xy xN MŽ .Ž .i,a i,a

C s (3)1y2xi,a 1y2w z w z2 2x | x |

N M N Mr y r xy xN M N MŽ . Ž .i,a i,ay ~ y ~

wherer designates one of the threea coordinatesi,a

of the position of the atomi. A single atomicvalue can be obtained by means of the geometricalaverage ofC . Fig. 8 shows theC as a functionx,i x,i

of the backbone atoms, extracted by averagingover a 10 ns MD trajectory, for free PCSS, PCSS-I and PCSS-II.Free PCSS reveals high values forC through-x,i

out the protein residues. In particular, the mostevident peaks are detected at Phe14, Cys25, Ile27,

Page 10: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

120 A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Met57, Leu62, Tyr83. Generally, most of thesepeaks find a correspondence with those observedin wild PC by Ungar et al.w33x; such an agreementconfirming that the insertion of a disulfide bridgedoes not alter the ET properties of the system.Indeed, some of these peaks involve protein resi-dues, which are known to be implied into the ETprocess. In particular, the peak at Phe14, spatiallyclose to Leu12, involves the PC hydrophobicpatch, while the peak at Leu62 is located aroundthe negative patch; both these regions being rele-vant for binding reaction partners as well forassisting the ET processw34x. Moreover, the strongvibrational coupling of Tyr83 is consistent withthe hypothesis that this residue could be a mainelement of the ET path from the PC partnercytochromef w34x. In addition, it is worth notingthat a strong correlation between the functionalactive site and the disulfide bridge is observed.Dynamical correlation extending over large dis-tances should be considered quite reasonably inthe light of the experimental finding, showing thatthe spectroscopic features of copper proteins aresensitive to the protein amino acid composition up12 A from the copper metalw52x.˚Regarding theC cross-correlation for PCSS I,x,i

most of the peaks observed in free PCSS aresuppressed. However, the peaks around Leu14,Glu61 and Tyr83 are still detected even if with alower intensity. An additional peak at Phe41 isobserved. Notably, the protein region containingthe engineered residues Cys21 and Cys25 exhibitslower values forC with respect to free PCSS. Inx,i

particular, a small peak is detected at the Cys25gold anchoring site, while practically no correla-tion is detected at the Cys21 residue. Such abehavior indicates that the removal of the covalentdisulfide bridge, in connection with the insertionof a single protein gold anchoring bond, cansignificantly modulate the dynamical correlationproperties of this region with the active site.On the contrary, theC trend for PCSS-IIx,i

appears qualitatively rather similar to that of freePCSS; three additional major peaks, located atresidues Ser22, Phe41 and Gly78, can be detected.At the same time, lower values in comparison tofree PCSS are observed around the residues Ser17,Cys25, Met57 and Glu60. It is worth noting that

the anchoring region, now involving two proteinsulfur gold bonds, exhibits high values forC .x,i

Such a finding could indicate that the dynamicalcross-correlation of the gold–protein anchoringregion with the active site, is strongly dependenton the structure of the binding site. However, thefact that the double bond anchorage is dynamicallycoupled to the copper active site suggests that itmight play a role in the ET process toward thegold electrode. This could be of some relevancefor bioelectronic application in hybrid systems, asa number of STM experiments seem to indicatew11–13,16x. Indeed, in these hybrid systems, thecovalent anchorage of the protein to the goldsurface is expected to take part in the ET pathwayjoining the copper active site to the metallicsubstrate; even if electric field fluctuations shouldalso be taken into considerations.To provide further information about the effect

of anchoring the macromolecule to a gold surfaceon its overall motions also in connection with itsET functionality, we have applied the PCA to theC atoms of all the analyzed systems. Similar toa

that observed in our previous simulations for wildtype and mutated plastocyaninsw24,26x, it hasbeen found that approximately 80% of the totalprotein motions are described by the first 30eigenvectors(data not shown). The absolute valueof the components of the first four eigenvectors ofPCSS, PCSS-I and PCSS-II are shown in Fig. 9.For free PCSS, we observe correlated motions,mainly involving turns(see Fig. 9). In particular,in the first eigenvector, concerted motions amongT3, T4 and T5 are detected. Similarly, in thesuccessive eigenvectors, concerted motions involv-ing the turns T3, T7, T1, T3, T4 and T8 areevident. Generally, such a behavior is qualitativelysimilar to that observed in fully hydrated PCSSand wild type PCw24,26x.At a visual inspection, the absolute values of

the first eigenvectors of both PCSS-I and PCSS-IIreveal concerted motions involving a larger portionof the macromolecule with respect to the freePCSS. Indeed, for both the systems, we note largepeaks in correspondence of regions containing bothturns andb-strands. Furthermore, it is interestingto remark that by comparing the first three eigen-vectors of PCSS-I and PCSS-II, the main peaks

Page 11: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

121A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

Fig. 9. Absolute values of the component of the first foureigenvectors for PCSS, PCSS-I and PCSS-II as a function ofresidue number. The positions of the turns T1, T3, T4, T5, T6,T7 and T8 in the secondary structure are also shown.

are approximately located at the same position.This suggests that anchoring the protein to thegold substrate might result in the selection of largerespiratory motions; irrespective of the fact thatone or both the protein sulfurs are covalentlybound to gold.Concerning the first two eigenvalues of PCSS-I

and PCSS-II, we note the appearance of correlatedmotions in the 1–19, 27–41, 54–72 and 82–97residues. These motions involve all the hydropho-bic patch and a large part of the negative patch;with both of these patches playing a role in thebinding with ET partners and being rather close tothe active sitew34x. At the same time, the copperligand residues are also found to be involved inthe correlated motions of these eigenvalues. In thethird eigenvalue of PCSS-I and PCSS-II, a markedpeak involving the 38–62 residues, at which thenegative and the hydrophilic patch is located, canbe detected. Finally, it is interesting to note thatthe fourth eigenvalue exhibits many sharp peakswith a behavior approaching that observed for freePCSS; such a trend being more evident for thesuccessive eigenvectors(not shown). These resultssuggest that a protein anchored to a gold substrateselectively enhances peculiar concerted motions inthe first eigenvectors, while higher eigenvectorsshow correlated motions rather similar to those

detected in the free system; the latter motionsbeing likely involved in the ET functionality.Accordingly, it could be hypothesized that themotions peculiar to the ET process could be stillactive in the immobilized systems.

4. Conclusions

The present study constitutes the first MD sim-ulation of a protein, covalently bound to a goldsubstrate. Our results point out that, anchoring adisulfide engineered plastocyanin, through eitherone or two covalent bonds to a gold surface doesnot substantially alter the structural properties ofthe macromolecule; the overall structure appearingstable for the 10 ns MD simulation run.The analysis of the orientation of the axis

joining the binding site to the copper ion, showsa different dynamical arrangement with respect tothe gold surface when anchoring by one or twocovalent bonds are considered. A statistical anal-ysis of the height of the macromolecule withrespect to the gold substrate has allowed us tocompare MD simulation data and topologicalexperimental data as obtained by AFM.Finally, the study of the Cu–S(Cys84) bond

distance cross-correlation and the PCA investiga-tions have provided us some information about thedynamical couplings relevant to the ET processand susceptible to be investigated by STM.MD modeling of these immobilized systems

could be of some help in elucidating the ETprocess also towards metallic electrode, whosecontrol should constitute the first step to buildhybrid systems by using ET biomolecules.

Acknowledgments

This work has been partially supported by the‘Molecular Nanodevices’ FIRB project, ‘SingleMolecule Spectroscopy’ INFM-PAIS project andthe EC Project SAMBA(V Frame FET).

References

w1x J.M. Shifman, B.R. Gibney, R.E. Sharp, P.L. Dutton,Functionalized de novo designed proteins: mechanismof proton coupling to oxidationyreduction in hemo

Page 12: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

122 A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

protein maquettes, Biochemistry 39 (2000)14 813–14 827.

w2x I. Willner, V. Geleg Shbtai, E. Katz, H.K. Rau, W.Haehnel, Integration of a reconstituted de novo synthe-sized hemoprotein and native metalloproteins with elec-trode supports for bioelectronic and bioelectrocatalyticapplications, J. Am. Chem. Soc. 121(1999) 6455–6468.

w3x G. Gilardi, A. Fantuzzi, S.J. Sadeghi, Engineering anddesign in the bioelectrochemistry of metalloproteins,Curr. Opin. Struct. Biol. 11(2001) 491–499.

w4x A. Ulman, Formation and structure of self-assembledmonolayers, Chem. Rev. 96(1996) 1533–1554.

w5x J. Chen, M.L. Reed, M.A. Rawlett, J.M. Tour, Largeon–off ratios and negative differential resistance in amolecular electronic device, Science 256(1999)1550–1552.

w6x C. Joachim, J.K. Gimzewski, Analysis of low-voltageI(V) characteristics of a single C60 molecule, Europhys.Lett. 30 (1995) (1995) 409–414.

w7x E.P. Friis, J.E.T. Anderson, Y.I. Kharkats, A.M. Kuznet-sov, R.J. Nichols, J.D Zhang, et al., An approach tolong range electron transfer mechanisms in metallopro-teins: in situ scanning tunneling microscopy with sub-molecular resolution, Proc. Natl. Acad. Sci. USA 96(1999) 1379–1384.

w8x P.B. Lukins, T. Oates, Single-molecule high-resolutionstructure and electron conduction of photosystem IIfrom scanning tunneling microscopy and spectroscopy,Biochim. Biophys. Acta 1409(1998) 1–11.

w9x G.B. Khomutov, L.V. Belovolova, S.P. Gubin, V.V.Khanin, A.Yu. Obydenov, A.N. Sergeev Cherenkov, etal., STM study of morphology and electron trans-port.features in cytochromec and nanocluster moleculemonolayers, Bioelectrochemistry 55(2002) 177–181.

w10x J.J. Davis, H.A.O. Hill, A.M. Bond, The application ofelectrochemical scanning probe microscopy to the inter-pretation of metalloprotein voltammetry, Coord. Chem.Rev. 200(2000) 411–442.

w11x J. Zhang, Q. Chi, Q.A.M. Kuznetsov, A.G. Hansen, H.Wackerbarth, H.E.M. Chris-tensen, et al., Electronicproperties of functional biomolecules at metalyaqueoussolution interfaces, J. Phys. Chem. B 106(2002)1131–1152.

w12x P. Facci, D. Alliata, S. Cannistraro, Potential-inducedresonant tunneling through a redox metalloprotein inves-tigated by electrochemical scanning probe microscopy,Ultramicroscopy 89(2001) 291–298.

w13x J.J. Davis, H.A.O. Hill, The scanning probe microscopyof metalloproteins and metalloenzymes, Chem. Comm.(2002) 393–401.

w14x L. Andolfi, S. Cannistraro, G.W. Canters, P. Facci, A.G.Ficca, I.M.C. Van Amsterdam, et al., A poplar plasto-cyanin mutant suitable for adsorption onto gold surfacevia disulphide bridge, Arch. Biochem. Biophys. 399(2002) 81–88.

w15x M. Milani, L. Andolfi, S. Cannistraro, M.Ph. Verbeet,M. Bolognesi, The 1.6 resolution crystal structure of a

mutant plastocyanin bearing a 21–25 engineered disul-fide bridge, Acta Cryst. D 57(2001) 1735–1738.

w16x L. Andolfi, S. Cannistraro, G.W. Canters, J.J. Davis,H.A.O. Hill, M. Ph. Verbeet, to be submitted.

w17x J. Hautman, M.L. Klein, Simulation of a monolayer ofalkyl thiol chains, J. Chem. Phys. 91(1989) 4994–5001.

w18x H. Groenbcck, A. Curioni, W. Andreoni, Thiols anddisulfides on the Au(111) surface: the headgroup-goldinteractions, J. Am. Chem. Soc. 122(2000) 3839–3842.

w19x M.C. Vargas, P. Giannozzi, A. Selloni, A.J. Scoles,Coverage-dependent adsorption of CH S and(CH S) )3 3 2

on Au(111): a density functional theory study, J. Phys.Chem. B 105(2001) 9509–9513.

w20x Z. Zhang, T.L Beck, J.T. Young, F.J. Boerio, Molecularstructure of manolayors from thiol-terminated polyimidemodel compounds on gold. 2. molecular dynamicssimulations, Langmuir 12(1996) 1227–1234.

w21x M. Tarek, K. Tu, M.L. Klein, D.J. Tobias, Moleculardynamics simulations of supported phospholipidyalkan-thiol bilayers on a gold(111) surface, Biophys. J. 77(1999) 964–972.

w22x S. Ravichandran, J. Talbot, Mobility of adsorbed pro-teins: a Brownian dynamics study, Biophys. J. 78(2000) 110–120.

w23x A. Ciocchetti, A.R. Bizzarri, S. Cannistraro, Long termmolecular dynamics simulation of copper plastocyaninin water, Biophys. Chem. 69(1997) 185–198.

w24x C. Arcangeli, A.R. Bizzarri, S. Cannistraro, Concertedmotion in copper plastocyanin and azurin: an essentialdynamics study, Biophys. Chem. 90(2001) 45–56.

w25x A.R. Bizzarri, S. Cannistraro, Molecular dynamics sim-ulation evidence of anomalous diffusion of proteinhydration water, Phys. Rev. E 53(1996) 3040–3043.

w26x C. Arcangeli, A.R. Bizzarri, S. Cannistraro, MD simu-lation and essential dynamics study of mutated plasto-cyanin: structural, dynamical and functional effects of adisulphide bridge insertion at the protein surface, Bio-phys. Chem. 92(2001) 183–199.

w27x J.J. Davis, D. Djuric, K.K.W. Lo, W.L.L. Wong, H.A.O.Hill, A scanning tunneling study of immobilized cyto-chrome P450, Faraday Discuss 116(2000) 15–22.

w28x S.T. Tang, A.J. McGhie, Imaging individual chaperoninand immunoglobulin G molecules with scanning tun-neling microscope, Langmnir 12(1996) 1088–1093.

w29x C.L. Brooks, R.E. Bruccoleri, B.D. Olafson, D.J. Statos,S. Swaminathan, M. Karplus, CHARMM: a programfor macromolccular energy, minimization, and dynamicscalculations, J. Comp. Chem. 4(1983) 187–217.

w30x A.D. MacKerrell Jr, D. Bashford, M. Bellott, R.L.Dunbrack Jr, J.D. Evanseck, M.J. Field, et al., All-atomempirical potential for molecular modeling and dynam-ics studies of proteins, J. Phys. Chem. B 102(1998)3586–3616.

w31x J.M. Guss, H.D. Bartunik, H.C. Freeman, Accuracy andprecision in protein structure analysis; restrained least-squares refinement of the structure of poplar plastocy-

Page 13: MD simulation of a plastocyanin mutant adsorbed onto a ...193.205.144.19/dipartimenti/disa/progetti/Biophysics/RicercaEn_file/... · MD simulation of plastocyanin, an electron transfer

123A.R. Bizzarri et al. / Biophysical Chemistry 106 (2003) 111–123

anin at 1.33 A resolution, Acta Cryst. B48(1992)˚790–811.

w32x A.R. Bizzarri, S. Cannistraro, Intensity fluctuations ofthe copper site resonant vibrational modes as observedby MD simulation in single plastocyanin molecule,Chem. Phys. Lett. 349(2001) 497–503.

w33x L.W. Ungar, N.F. Seherer, G.A. Voth, Classical molec-ular dynamics simulation of the photoinduced electrontransfer dynamics of plastocyanin, Biophys. J. 72(1997) 5–17.

w34x M.R. Redinbo, T.O. Yeates, S. Merchant, Plastocyanin:structural and functional analysis, J. Bioener. Biomembr.26 (1994) 49–66.

w35x J.O.A. De Kerpel, U. Ryde, Protein strain in blue copperproteins studied by free energy perturbations, Proteins:Struct. Fund. Genet. 36(1999) 157–174.

w36x P. Coniba, R. Remenyi, A new molecular mechanicsforce field for the oxidized form of blue copper proteins,J. Comp. Chem. 23(2002) 697–705.

w37x P.J. Steinbach, B.R. Brooks, Protein hydration elucidatedby molecular dinamics simulation, Proc. Natl. Acad.Sci. USA 90(1993) 9135–9139.

w38x P.J. Steinbach, B.R. Brooks, Hydrated myoglobin’sanharmonic fluctuations are not primarily due to dihe-dral transitions, Proc. Natl. Acad. Sci. USA 93(1996)55–59.

w39x J.P. Ryckaert, G. Ciccotti, H.J.C. Berendsen, Numericalintegration of the Cartesian equations of motion of asystem with constraints: molecular dynamics ofn-alkanes, J. Comput. Phys. 23(1977) 327–341.

w40x B. Lee, F.M. Richards, The interpretation of proteinstructures: estimation of static accessibility, J. Mol. Biol.55 (1971) 379–400.

w41x D.J. Lavrich, S.M. Wetterer, S.T. Bernasek, G. Scoles,Physisorption and chemisorption of alkanethiols andalkyl sulfides on Au(111), J. Phys. Chem. B 102(1998)3456–3465.

w42x H.H. Jung, Y.D. Won, S. Shin, K. Kim, Moleculardynamics simulation of benzenethiolate and benzyl mer-captide on Au(111), Langmuir 15(1999) 1147–1154.

w43x J. Qian, R. Hentschke, W. Knoll, Superstructures ofcyclodextrin derivatives on Au(111): a combined ran-dom planting-molecular dynamics approach, Langmuir13 (1997) 7092–7098.

w44x A. Amadei, A.B.M. Linssen, H.J.C. Bercndsen, Essen-tial dynamics of proteins, Proteins: Struct. Funct. Genet.17 (1993) 412–425.

w45x A.E. Garcia, Large amplitude nonlinear motions inproteins, Phys. Rev. Lett. 68(1992) 2696–2699.

w46x A.R. Bizzarri, S. Cannistraro, Flickering noise in thepotential energy fluctuations of proteins as investigatedby MD simulation, Phys. Lett. A 236(1997) 596–601.

w47x H. Frauenfelder, F. Parak, R.D. Young, Conformationalsubstates in proteins, Ann. Rev. Biophys. Biophys.Chem. 17(1988) 451–479.

w48x P.H. Hueneberger, A.E. Mark, W.F. van Gunsteren,Fluctuations and cross-correlation analysis of proteinmotions observed in nanosecond molecular dynamicssimulations, J. Mot. Biol. 252(1995) 492–503.

w49x H. Assender, V. Bliznyuk, K. Porfyrakis, How surfacetopography relates to materials’ properties, Science 297(2002) 97ex3–976.

w50x L. Andolfi, B. Bonanni, G.W. Canters, M.Ph. Verbeet,S. Cannistraro, SPM characterization of gold-chemi-sorbed poplar plastocyanin mutants, Surf. Sci. 530(2003) 181–194.

w51x R.A. Marcus, N. Sutin, Electron transfer in chemistryand biology, Biochim. Biophys. Acta 811(1985)268–322.

w52x G.R. Loppnow, E. Fraga, Proteins as solvents: the roleof amino acid composition in the excited-state chargetransfer dynamics of plastocyanins, J. Am. Chem. Soc.119 (1997) 896–905.