23
University of Groningen Structure and dynamics of peptides van der Spoel, David IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 1996 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): van der Spoel, D. (1996). Structure and dynamics of peptides: theoretical aspects of protein folding. s.n. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). The publication may also be distributed here under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license. More information can be found on the University of Groningen website: https://www.rug.nl/library/open-access/self-archiving-pure/taverne- amendment. Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 26-07-2022

University of Groningen Structure and dynamics of peptides

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: University of Groningen Structure and dynamics of peptides

University of Groningen

Structure and dynamics of peptidesvan der Spoel, David

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

Document VersionPublisher's PDF, also known as Version of record

Publication date:1996

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):van der Spoel, D. (1996). Structure and dynamics of peptides: theoretical aspects of protein folding. s.n.

CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

The publication may also be distributed here under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license.More information can be found on the University of Groningen website: https://www.rug.nl/library/open-access/self-archiving-pure/taverne-amendment.

Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

Download date: 26-07-2022

Page 2: University of Groningen Structure and dynamics of peptides

Chapter 6

MD Simulations of N-terminal Peptides from LactateDehydrogrenase

David van der Spoel, Hans J. Vogel and Herman J.C. Berendsen

PROTEINS Struct. Funct. Gen. 24:450-466 (1996)

Molecular dynamics (MD) simulations of N-terminal peptides from Lactate Dehydrogenase

(LDH) with increasing length and individual secondary structure elements were used to study

their stability in relation to folding. Ten simulations of 1-2 ns of di�erent peptides in water

starting from the coordinates of the crystal structure were performed. The stability of the

peptides was compared qualitatively by analyzing the root mean square deviation (RMSD)

from the crystal structure, radius of gyration, secondary and tertiary structure and solvent

accessible surface area. In agreement with earlier MD studies, relatively short (< 15 amino

acids) peptides containing individual secondary structure elements were generally found to

be unstable; the hydrophobic �1-helix of the nucleotide binding fold displayed a signi�cantly

higher stability, however. Our simulations further showed that the �rst ��� supersecondary

unit of the characteristic dinucleotide binding fold (Rossmann fold) of LDH is somewhat

more stable than other units of similar length and that the �2-helix, which unfolds by itself,

is stabilized by binding to this unit. This suggests that the �rst ��� unit could function

as an N-terminal folding nucleus, upon which the remainder of the polypeptide chain can

be assembled. Indeed, simulations with longer units (���� and ������) showed that all

structural elements of these units are rather stable. The outcome of our studies is in line with

suggestions that folding of the N-terminal portion of LDH in vivo can be a cotranslational

process which takes place during the ribosomal peptide synthesis.

Page 3: University of Groningen Structure and dynamics of peptides

84 Structure and Dynamics of Peptides

6.1 Introduction

The folding of a protein into its correct three dimensional structure is a topic of continuinginterest [33]. Following the original demonstration by An�nsen [15], that a denaturedprotein can regain its original structure when the denaturing agent is removed, it isclear that the information required for the folding of the protein is retained in its aminoacid sequence. Since this discovery it has been customary to study protein folding byfollowing the renaturation of the denatured polypeptide, which comprises the entire aminoacid sequence. These studies have led to the widely held view that so-called \moltenglobules" exist as obligatory intermediates along the in-vitro folding pathway of manyproteins [22{24, 239]. Such intermediates have been characterized in detail primarily byNMR hydrogen exchange measurements and have been found for various segments inthe amino acid sequence [240{242]. This view is no longer generally accepted however,as recent experiments have suggested that the partially folded forms of some proteinscould be trapped unproductive forms rather than obligatory folding intermediates [25, 26].Other studies aimed at solving the protein folding problem have relied on the assumptionthat the overall folding of a protein is dictated by the intrinsic folding properties of shortstretches of the amino acid sequence. Consequently, short linear peptides resemblingparts of the amino acid sequence of various proteins have been produced synthetically,and have been studied by high resolution NMR and CD spectroscopy for their foldingproperties [66, 212, 243]. While this work has clearly established the transient formationof elements of secondary structure in such peptides, the role of these secondary structureelements in directing the course of folding remains to be determined [212, 243].

The question remains whether protein folding in vivo really occurs from a completelysynthesized and unstructured random coil peptide. Several authors have argued thatit would be possible in principle to initiate folding in the N-terminal part of a proteinbecause the polypeptide chain emerges in an ordered manner from the ribosome duringprotein translation [1, 32, 244, 245]. This process could give rise to an N-terminal foldingnucleus, which can subsequently serve as a site of assembly for the remaining portionsof the polypeptide chain. This phenomenon is referred to as cotranslational rather thanposttranslational folding [244], and a number of molecular biology studies with expressedtruncated proteins [246{249] have indicated that the folding of certain classes of pro-teins may well proceed in this fashion. If such an N-terminal folding nucleus were tobe present within a group of proteins with identical folds, it should have a recognizablefeature near the N-terminal end of the protein. One group of proteins that has a well-characterized overall fold, which can be readily recognized on the basis of their amino acidhomology are the nucleotide binding proteins [250]. In particular the well characterizedRossmann fold, also known as the \classical dinucleotide binding chain" fold, is foundin all NAD+, NADP+ and FAD binding proteins [250, 251]. In addition, a related, butstructurally dissimilar \mononucleotide binding chain" fold is found in many proteins thatbind nucleotides such as ATP and GTP [250, 252, 253]. That this class of proteins may besuitable for studying the formation of an N-terminal folding nucleus was suggested in apaper by Wierenga et. al. [254]. They commented that the nucleotide binding ��� unit

Page 4: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 85

could be a possible nucleation site for protein folding: \... In the known structures these���-structural entities do always occur near the N terminus ... This suggests that these���-folds might function as a nucleation center for the folding of the complete domain ..."The same region was also indicated as a possible folding unit by an analysis of the riboso-mal translation rate [255]. Finally, it is known that N-terminal proteolytic fragments ofLactate Dehydrogenase (LDH) retain the capacity to bind dinucleotides, indicating thatthey are properly folded [256]. The ��� unit, which is an integral part of the Rossmannfold, is a common supersecondary structural element in proteins [257]. In all, at least 40proteins are presently known or predicted to have this fold, and considering the \age" ofthe fold, which was estimated to be over 3 billion years [257] there will probably be morethan these.

In this work we have used Molecular Dynamics (MD) calculations to obtain furtherinformation about the stability of peptides derived from the N-terminal portion of LDH, aprotein with the dinucleotide binding fold. Over the last few years MD calculations, whenused judiciously, have become a reliable means for studying the stability and exibilityof small proteins in water [43, 258]. With continuing improvements in force �elds andcomputing capacity, it has become possible to study processes such as protein denatura-tion [37, 38, 41, 42, 258]. In addition, simulations of peptides in aqueous solution can nowbe performed routinely and such studies have been done to assess the stability of turnsand helices in short linear peptides [36, 40, 202, 206, 213, 238, 259]. It is not possible toasses the thermodynamic stability of a peptide or protein in solution directly from MDsimulations, due to limited sampling of unfolded as well as partially folded states. How-ever, it is possible to investigate the kinetic stability i.e. the resistance against unfoldingand compare this for various peptides or proteins in di�erent environments or at di�erenttemperatures. This has recently been done in a number of cases, four regarding �-helicalpeptides [44, 67, 225, 260] and in another case a �-hairpin was studied [147]. Finally, ithas also been noted that MD simulations can assist in an analysis of protein foldingpathways [261].

Tsou [244] has suggested that an N-terminal folding nucleus should have a structurethat may not be identical, but should resemble the �nal structure in the completed protein.Therefore we decided to use the folded parts as determined in the crystal structure of theprotein as starting structures for our MD calculations. In order to simulate the e�ectof the ribosomal protein synthesis, we studied the N-terminal region, and systematicallyincreased the length of the simulated polypeptide chain. For comparison a number ofother units derived from the same amino acid sequence were also simulated. While thisapproach does not assess folding directly, the kinetic stability of the various peptidescan be compared in this manner, as was done previously in MD studies of �-helices and�-turns [36, 40, 44, 67, 147, 202, 213, 225, 238].

Page 5: University of Groningen Structure and dynamics of peptides

86 Structure and Dynamics of Peptides

6.2 Methods

Ten MD simulations of

20 30 40 50 60 70 80 90Residue

Sim. β1

Sim. α1

Sim. βα

Sim. β2

Sim. βαβ1

Sim. βαβ−

Sim. α2

Sim. βαβ2

Sim. βαβα

Sim. βαβαββ

β α β α β β

1.0 ns

2.0 ns

2.0 ns

1.0 ns

2.0 ns

2.0 ns

2.0 ns

2.0 ns

1.0 ns

1.0 ns

Figure 6.1: Overview of the peptides names and sequences used.

peptides from LDH in wa-ter, varying in length from6 to 76 amino acid residues,were performed. Oneof these was a deproto-nated peptide, in whichthe sidechain of a Ly-sine residue (Lys22) wasmade neutral, in order toassess the in uence of asalt bridge on the stability.In our simulations we ig-nored the �rst 20 residuesof the protein, because inthe complete protein theyform a motif which is stick-ing out into the solvent.This loop is necessary for

oligomerization, where the loop stabilizes the tetramer [256]; therefore the loop presum-ably is less important for the structure of the monomeric folded protein.

Number 2 3 4 5 6

1234567890123456789012345678901234567890

Amino Acid NKITVVGVGAVGMACAISILMKDLADEVALVDVMEDKLKG

Sec. Struct. S EEEEE SHHHHHHHHHHHHHT SEEEEE S HHHHH

Hydrophobic oooo o oo ooooo ooo oo oooo oo o

Charge + +- -- - --+ +

Number 6 7 8 9

123456789012345678901234567890123456

Amino Acid EMMDLQHGSLFLHTAKIVSGKDYSVSAGSKLVVITA

Sec. Struct. HHHHHHHHH GGG SEEEEESSGGGT S SEEEE

Hydrophobic oo o ooo oo oo o o oooooo

Charge - - + + + +- +

Table 6.1: Sequence characteristics (H = �-helix, G = 310-helix, T = Hydrogen Bonded Turn,S = Bend, E = Extended strand participating in Beta Sheet, Space = No de�ned secondary

structure).

All but four of the peptides start at residue Asn21. To distinguish the simulations we

Page 6: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 87

named them �1, �1, ��, �2, ���1, �2, ���2, ���� and ������ respectively, wherethe names refer to the structural elements simulated. The simulation of the deprotonatedpeptide was named ����. An overview of the simulations is given in Fig. 6.1. Simulations�1, ��, �2, ���1, ���

� and ���2 were 2.0 ns, the other simulations were 1.0 ns long.The sequence for the longest ������ peptide along with some characteristics are givenin Table 6.1. Plots of the ���1 and ������ peptides are given in Fig. 6.2 and Fig. 6.3respectively.

The coordinates for the peptides were taken

Figure 6.2: Plot of the ��� unit withLys22 and Glu47 in all atom model. Plotwas made using Molscript [170]

from the crystal structure of dog�sh apo-lactate dehydrogenase which was determinedby Abad-Zapatero [262]. The structure wasretrieved from the Brookhaven pdb database,entry 6LDH. They were solvated in a rectangu-lar box of SPC water [164] using a shell of atleast 0.8 nm on every side of the peptide. Al-though the use of counterions would make thetotal system electrostatically neutral, we feelthat this would not improve the results fromthe simulation without using a proper long-range electrostatic method such as Ewald sum-mation [53, 197] or particle-particle, particle-mesh (PPPM) [236, 237]. Although these meth-ods are well-established there are some seriousdrawbacks to using them: 1) a crystalline envi-ronment is imposed on a system that is meantto be a peptide in in�nite dilution, 2) a cor-rect lattice sum method requires the presenceof counterions which, when not sampled prop-erly will induce artifacts, 3) the CPU time is

much longer due to di�erent scaling, O(N2 log2 N) for Ewald vs. O(N) for simple cut-o�s, although the required CPU time for the PPPM method is comparable to a simplecut-o� [237]. Reaction �eld methods assume a dielectric continuum outside a speci�edcut-o� range around a given particle [263], which implies that the whole peptide shouldbe within cut-o� distance; this is impractical for peptides with a typical length of 2 nm,twice the cut-o� distance. A further complication towards using e.g. the PPPM methodis, to our knowledge, that there is no implementation of this method on message passingparallel computers which we have used for our simulations.

Some promising methods to deal with long-range electrostatic interactions are currentlyunder development [233, 264]; these will be implemented on our parallel computers in thenear future.

All simulations were performed using periodic boundary conditions. The solvated pep-tides were subsequently energy minimized with the steepest descent method for 100 steps.

Page 7: University of Groningen Structure and dynamics of peptides

88 Structure and Dynamics of Peptides

Then the system was run with position restraining on the peptide for 50 ps to allow forfurther relaxation of the solvent molecules. During these restrained MD runs the temper-ature was controlled using weak coupling to a bath of constant temperature (T0 = 300K, coupling time �T = 0.1 ps) and the pressure was controlled using weak coupling to abath of constant pressure (P0 = 1 bar, coupling time �P = 0.5 ps) [168]. The produc-tion runs were done with the same pressure- and temperature-coupling constants as therestrained runs. Protein and solvent were coupled to the temperature bath separately inrestrained as well as free MD. The center of mass motion of the entire simulation systemwas removed every step. The Gromos-87 force�eld [11] was used with modi�cations assuggested in [159]. The timestep was 2.0 fs, SHAKE [169] was used for all covalent bonds,and the cut-o� for nonbonded interactions was set to 1.0 nm. Neighbourlists were used,and updated every 20 fs.

To each peptide we added

Figure 6.3: Plot of the ������ unit, was made using

Molscript [170]

a CH3C=O group at the N-terminus and an NH2 group at theC-terminus; these neutral groupswere used in order to avoid dis-turbing the peptides by introduc-ing charges on their N- and C-terminal ends. For comparisonwith future NMR and CD experi-ments, residue Cys35 was mutatedto Ser to eliminate the possibil-ity of disulphide bridge formationin the test tube; we think thissubstitution is unlikely to havea great impact on the structureof the peptides. The number ofatoms and the number of watermolecules for each peptide simu-lation are given in Table 6.2. Thesimulations were carried out us-ing the GROmacs software pack-age [265] on custom-built parallelcomputers [51, 52] with Intel i 860CPU's and on a Silicon GraphicsPower Challenge. The total com-puter time for all simulations wasapproximately 2000 h.

Page 8: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 89

Residues # Residues # Waters #Atoms Charge

�1 Asn21 - Gly29 9 1346 4120 +1�1 Gly27 - Asp43 17 1612 4974 0�� Asn21 - Asp43 23 1886 5854 +1�2 Asp46 - Val51 6 1394 4238 -2���1 Asn21 - Asp52 32 1929 6057 -2���

� Asn21 - Asp52 32 1947 6110 -3�2 Asp56 - Ser69 14 1627 5020 0���2 Asp46 - Asp82 37 1919 6105 -2���� Asn21 - Leu72 52 2340 7480 -3������ Asn21 - Ala96 76 2510 8202 0

Table 6.2: Overview of the peptides used in the simulations, the number of water molecules andthe net charge on each peptide.

6.3 Results

When analysing long MD simulations of peptides or proteins it is useful to de�ne prop-erties that give a relative estimate of the stability. We have analysed the RMS deviationfrom the crystal structure, the radius of gyration and the solvent accessible hydrophobicsurface area. These values can be plotted as a function of time; in a stable protein thesevalues uctuate around an equilibrium value, in a changing conformation there will be adrift in one or more of these observables. Since we do not expect all our peptides to bestable in the simulation it is also interesting to look at the dynamical processes takingplace during the simulation. To illustrate this we have also looked at secondary struc-ture analysis, distance matrices and salt bridges involving Lys22. These items will beaddressed in subsequent sections.

6.3.1 RMS deviation from the crystal structure

The root mean square deviation (RMSD) of the backbone atoms (N,C�,C) with respect tothe crystal structure was calculated by least-squares �tting the backbone atom positionsof each peptide to the crystal structure and subsequently calculating the RMSD (eqn. 6.1).

RMSD(t) =

"1

NS

Xi2S

(ri(t)� ri(0))2

# 1

2

(6.1)

where ri(t) is the position of atom i at time t and S is a given subset of all atoms withsize NS . In Fig. 6.4 the RMSD averaged over all residues is plotted as a function of timefor all the simulations.

The RMSD of the �-strands in Sim. �1 and �2 are very high; these peptides deviate verymuch from the crystal structure. �2 deviates gradually, whereas �1 is more uctuating.

Page 9: University of Groningen Structure and dynamics of peptides

90 Structure and Dynamics of Peptides

It should be mentioned that an RMS deviation of 0.7 nm for a 9 residue peptide (�1)means that it is not very meaningful to �t the structures on top of each other. After aninitial rise of the RMSD to 0.3 nm, the isolated �1-helix is stable for 800 ps at an RMSDof 0.2 nm (from 350 until 1150 ps).

After this point in the

0 500 1000Time (ps)

0.0

0.2

0.4

0.6

RM

S (

nm)

β1

β2

βαβαβαβαββ

0 500 1000 1500 2000Time (ps)

0.0

0.2

0.4

RM

S (

nm)

βαβ1

βαβ2

βαβ−

0.0

0.2

0.4

RM

S (

nm)

α1

α2

βα

Figure 6.4: Root Mean Square Deviation of all backbone atomsfrom the X-ray structure. A running average over 25 ps is given

to improve clarity.

simulation the RMSD grad-ually rises to more than 0.4nm due to unfolding of theC-terminal part of the �-helix. The �2 peptide de-viates up to 0.4 nm fromthe crystal structure within800 ps and remains at thatlevel for the rest of the 2.0ns simulation. In the ��

simulation the RMSD goesup after 600 ps due to un-folding of the helix startingfrom the N-terminal side,i.e. from the exible loop.Around 1.0 ns the �� pep-tide reaches an RMSD of0.5 nm, after which pointthe RMSD decreases to 0.3nm at the end of the simu-lation.

In the runs of the ���1,the ���

� and the ���2

peptides the RMS devia-tion slowly increases to 0.4nm after 1000 ps, but after1250 ps it decreases againuntil it stabilises at 0.3 nm.In all three ��� units the

RMSD is then virtually constant for over 700 ps until the end of the 2.0 ns simulation.In the two longest peptides the RMSD slowly rises to 0.2 (������) resp 0.3 (����)nm. In both simulations there is still a small drift at the end of the simulation (1000 ps)indicating that the peptides are not fully equilibrated.

Page 10: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 91

The RMSD was also averaged over the backbone atoms per residue over the last 1000ps of each simulation (ie. over the entire trajectory for Sims. �1, �2, ���� and ������)(see Fig. 6.5). Here we can compare how parts of secondary structure behave in di�erentpeptides.

In most simulations

20 30 40 50 60 70 80 90Residue

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4

0.0

0.4 β1

α1

βα

β2

βαβ1

βαβ−

α2

βαβ2

βαβα

βαβαββ

RM

S (

nm)

Figure 6.5: Root Mean Square Deviation from the X-ray struc-ture per residue, averaged over the last 1000 ps of each simula-tion.

that contain the �1-helix(��, ���1, ���

� and, toa lesser extent in ����)the C-terminal end of the�1-helix is exible. The�1-helix is exible at theends only whereas the�2-helix is exible over itstotal length, especially nearGly60. In all simulationscontaining residues 70-72(Leu-Phe-Leu) (���2,���� and ������) thissection of the polypeptidechain is very exible be-cause three hydrophobicside chains are solventexposed, and are trying tomove away from the sol-vent. There is no suitableplace to go for these sidechains however, so theyremain solvated. Anothernoticeable feature is theGly-rich loop (residues 27-32) which shows a peak in�1, �� and ���

� thoughvarying in height.

6.3.2 Radius of gyration

For each peptide the radius of gyration Rg was calculated according to eqn. 6.2

Rg =

�Pi r

2imiP

imi

� 1

2

(6.2)

where mi is the mass of atom i and ri the position of atom i with respect to the centerof mass of the peptide. The radius of gyration is a rough measure for the compactness

Page 11: University of Groningen Structure and dynamics of peptides

92 Structure and Dynamics of Peptides

of a structure. In all but the longest peptides (���2, ���� and ������) we �nd somechange in Rg (Fig. 6.6).

In Sim. �1 the decrease

0 500 1000Time (ps)

0.3

0.5

0.7

0.9

1.1

Rg

(nm

)

β1

β2

βαβαβαβαββ

0.0 500.0 1000.0 1500.0 2000.0Time (ps)

0.7

0.8

0.9

1.0

Rg

(nm

)

βαβ1

βαβ2

βαβ−

0.6

0.7

0.8

Rg

(nm

)

α1

α2

βα

Figure 6.6: Radius of Gyration during the MD simulations. Arunning average over 25 ps is given to improve clarity.

at about 200 ps is due toa collapse of the peptidewhere a �-turn-like struc-ture is formed, which laterunfolds again. �2 collapsessomewhat and remains sta-ble after 700 ps. In �1

Rg decreases gradually be-cause the loops at bothends of the �1-helix some-times curl inward. In con-trast the large uctuationsin Rg in Sim. �2 are dueto unfolding and refoldingof parts of the �2-helix. Inthe �rst 300 ps of the ��simulation we see Rg de-creasing with 0.1 nm be-cause the C-terminal loopat the end of the helixcollapses inward, making ashort lived hydrogen bondbetween the sidechain NH2of Asn21 and the carbonyloxygen of residue Asp43.Movements in the Gly-richloop (res 27-32) also con-tribute to the change in Rg .

After 1.0 ns Rg decreases slowly to 0.75 nm. In Sim. ���� Rg decreases somewhat after800 ps due to movements in the second loop region; in ���1 in contrast Rg rises slightlyafter 800 ps although the �nal value (0.9 nm) is close to the initial one (0.87 nm).

6.3.3 Secondary structure analysis

A secondary structure analysis was performed with the DSSP program which determinesthe existence of hydrogen bonds as criteria for the presence of secondary structure [65].In Fig. 6.7 we have plotted the secondary structure elements of each residue in the dif-ferent peptides as a function of time. Colors were used to distinguish between secondarystructure types, the most important colors being red for �-sheet and blue for �-helix.

Page 12: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 93

0 200 400 600 800 1000 1200 1400 1600 1800 2000

5660α2

4650

60

70

80

βαβ2

21

30

40

50

βαβ−

21

30

40

50

βαβ1

21

30

40

βα

2730

40

α1

0 100 200 300 400 500 600 700 800 900 1000

21

30

40

50

60

70

80

90

βαβα

ββ

21

30

40

50

60

70

βαβα

4650β2

21β1

Time (ps)Coil B-Sheet B-Bridge Bend Turn A-Helix 5-Helix 3-Helix

Figure 6.7: Secondary structure as a function of time for all simulations.

Page 13: University of Groningen Structure and dynamics of peptides

94 Structure and Dynamics of Peptides

From Fig. 6.7 it can be seen that the isolated �1 (Res. 21-27) and loop (Res. 28-29)form a hydrogen bonded turn centered around Gly27 sometimes, and a bend most of thetime. In �2, a hydrogen bonded turn appears to be the most stable form in the simulation,during formation of the turn structure a transient 310 Helix is found, at 320 ps and around375 ps. Neither of the two �-strands remains even close to their starting structure. The�1-helix is rather stable, compared to the crystal structure only two residues on each sidecome o� initially; three residues in the middle of the �1-helix also loose their �-helicitytemporarily, but they rapidly regain their original structure. After 1500 ps however, 4residues on the C-terminal end lose their helicity although they remain hydrogen bondedin a turn. As can be seen in the �1 simulation (around 125 ps) the helix secondarystructure type (blue) can be rapidly changed into the turn structure type (yellow) andback, indicating that these secondary structure types can interconvert rapidly within thede�nitions of the DSSP program. In the �� unit the hydrophobic �1-strand and theGly-rich loop apparently are \pulling" on the �1-helix thereby unfolding it from the N-terminal side; at 1000 ps �1 is completely unfolded in this peptide. The �1-helix rapidlyreforms however, starting from the N-terminal end at Met33 which directly follows a Glyresidue: an 9 residue �-helix is formed in 200 ps. Simulations ���1 and ���

� can becompared in detail using this analysis too; both units are altogether stable, the number ofhydrogen bonds in the �-sheet varies but is never less than two. In the �-helical region ofthese peptides there is a di�erence between the two units, the �1-helix loses three residueson the C-terminal end in the ���1 simulation whereas it loses seven �-helical residues inthe C-terminal end in the ���� simulation. The loss of helicity is due to rearrangementsinvolving the hydrophobic residues in the second loop region and at the end of the �1-helix.After 2.0 ns 8 helical residues remain in simulation ���1 and 4 in ���

�, and a stable�-sheet in both. Considerable structural changes occur in the �2 peptide simulation,with �-helical residues transforming into �-helix or a hydrogen bonded turn, initially atthe C-terminal end only, but after 600 ps the �2-helix unfolds from the N-terminal sideand for several periods of up to 100 ps (1150 to 1250 ps), all �-helix is gone. A centralGly residue (Gly60) probably is the reason for breaking. During the whole simulation wesee �-helical parts in either the N-terminus or the C-terminus of the peptide but rarelysimultaneously (except around 1700 ps). In the ���2 peptide we see that the �2-helixunfolds from the C-terminus. All �-helical residues after Gly60 are gone after 850 ps. Incontrast the �-sheet is very stable, it restricts the conformational space of the residues inbetween. It is also interesting to note that considerable non-native secondary structureis formed after the C-terminal part of the �2-helix has unrolled. When the �2-helix isconnected to the ���1 peptide to form the ���� peptide the �2-helix is more stable thanin isolated form in solution (Fig. 6.7). In the ���� peptide the most exible part (apartfrom the three C-terminal residues) is the Gly-rich loop (Res. 27-32), which unfolds the�1-helix from the N-terminal side, similar to what was observed in the �� simulation. Asin the �2 simulation the �2-helix breaks at Gly60, but at either side of Gly60 the �2-helixis rather stable. The longest ������ peptide is stable during the whole simulation, onlythe �2-helix (which breaks at Gly60 again) and the three solvent-exposed hydrophobic

Page 14: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 95

residues (70-72) change their secondary structure, although almost always 10 residuesremain in �-helical conformation in the �2-helix. In both the ���� and ������ the�-sheet is stable; there are almost always two or more hydrogen bonds connecting theindividual �-strands; nevertheless in the ������ peptide the �rst and second �-strandare even more stable than in the ����, due to the addition of the two extra �-strandson either side of the original �-sheet (�-strands are in order �4�1�2�3). Although thisincreased �xation of �-strands also gives rise to a more stable �1-helix the secondarystructure pattern of the �2-helix in the ������ peptide is similar to that in the ����peptide.

6.3.4 Distance matrices

Although the secondary structure analysis is very important for understanding the sim-ulation, it does not tell all about the tertiary structure of the protein. We have thereforecalculated distance matrices, in which a matrix of residue-residue distances is computedas a function of time. We de�ned the distance between two residues Ai and Aj as thesmallest distance between any pair of atoms (i 2 Ai, j 2 Aj). The distance matrix is thussymmetric by de�nition. We have visualised this information using greyscales to representdistances. In Fig. 6.8 the distance matrices for �ve di�erent time frames in simulations���1, ���

� and ���2 are plotted. The distances have been discretized in thirteen levelsfrom 0 (black) to 1.2 nm (white) as is depicted in the �gure legend. All distances greaterthan 1.2 nm are plotted in white. Although the �gures appear to be very similar, acloser look reveals very detailed information in these �gures. The �1-helix can be seenin the center of the �gures, where a subdiagonal (and a superdiagonal) line close to thediagonal, shows that residues at position (n+4) in the sequence are close to residues atposition (n), where (n) runs from residue 29 to 38. At 800 ps in both simulations partof the �1-helix has unfolded at the C-terminal end, but more so in the ���� simulationthan in the ���1 simulation. Parallel �-strands show up as sub (super) diagonal linesfurther removed than four residues from the diagonal. It can also be seen that residues21 through 27 are close to residues 47 through 52, which is exactly the �-sheet region inthe peptide. The �-sheet remains intact during the entire simulation of these peptides.Finally, it can also be seen from these distance matrices that unfolding of the C-terminalpart of the �1-helix coincides with closer contact of almost all residues in the loop regionthat connects the �-helix to the second �-strand. ���2 behaves rather similar to theother ��� unit, the �-sheet is stable and the end of the �2-helix is somewhat disorderedafter 1000 ps. Also at t=1000 ps we see events take place around residue Gly60: somelocal close contacts between residues 59, 60 and 61 have formed. In Fig. 6.9 we have plot-ted the distance matrices of the larger ���� and ������ peptides at 0, 500 and 1000ps. As in the smaller peptide the secondary structure elements such as the two �-helicesare easily recognized in the distance matrix.

Page 15: University of Groningen Structure and dynamics of peptides

96 Structure and Dynamics of Peptides

21 30 40 50

21

30

40

50

t=0

ps

21

30

40

50

t=50

0 ps

21

30

40

50

t=10

00 p

s

21

30

40

50

t=15

00 p

s21

30

40

50

t=20

00 p

s

Residue21 30 40 50

21

30

40

50

21

30

40

50

21

30

40

50

21

30

40

50

21

30

40

50

Residue

0 Distance (nm) 1.2

4650 60 70 80

4650

60

70

80

t=0

ps4650

60

70

80

t=50

0 ps

4650

60

70

80

t=10

00 p

s

4650

60

70

80

t=15

00 p

s

4650

60

70

80

t=20

00 p

s

Residue

Figure 6.8: Residue-Residue distance matrices from �ve time frames of simulations ���1 (left)

and ���� (center) and ���2 (right). The distance between residues Ai and Aj is the smallestdistance between any pair (i,j) of atoms (i 2 Ai, j 2 Aj).

In Sim. ������ around residue 85 the 310-helix is visible as a dark region and it isalso possible to deduce the order in which the four �-strands comprise the �-sheet fromthis plot (�4�1�2�3). From the distance matrices at 500 ps and 1000 ps it can be seenthat the �-helices have moved somewhat closer, giving rise to more �-� contacts (roughlyresidues 30-40 and 55-65) in both ���� and ������ and more close (n,n+4) contactsat 1000 ps in ������ only. The most important conclusion from these plots is howeverthat the tertiary structure from the crystal structure is still present at the end of each of

Page 16: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 97

the simulations that have at least a ��� unit.

21 30 40 50 60 70

21

30

40

50

60

70

t=0

ps

21

30

40

50

60

70

t=50

0 ps

21

30

40

50

60

70

t=10

00 p

s

Residue

0 Distance (nm) 1.2

21 30 40 50 60 70 80 90

21

30

40

50

60

70

80

90

t=0

ps

21

30

40

50

60

70

80

90

t=50

0 ps

21

30

40

50

60

70

80

90

t=10

00 p

s

Residue

Figure 6.9: Residue-Residue distance matrices of three time frames from simulation ���� (left)and ������ (right). The distance between residues Ai and Aj is the smallest distance between

any pair (i,j) of atoms (i 2 Ai, j 2 Aj).

Page 17: University of Groningen Structure and dynamics of peptides

98 Structure and Dynamics of Peptides

6.3.5 Solvent accessible surface area

To examine the e�ect

0 500 1000Time (ps)

0

10

20

30

Are

a (n

m2 )

β1

β2

βαβαβαβαββ

0 500 1000 1500 2000Time (ps)

12

14

16

18

Are

a (n

m2 )

βαβ1

βαβ2

βαβ−

5

10

15

Are

a (n

m2 )

α1

α2

βα

Figure 6.10: Solvent accessible hydrophobic surface area during the

simulations. A running average over 25 ps is given to improve clar-ity.

of the solvent on hy-drophobic residues, thesolvent accessible sur-face area was calcu-lated with the DSSPprogram [65]. Theprogram calculates thesurface area and plotsit per residue. Wehave summed up the to-tal area for hydropho-bic side chains and plot-ted that as a function oftime in Fig. 6.10.

In some of the pep-tides we see that thetotal hydrophobic sur-face area is reduced overthe course of the sim-ulation (�2, ��, ���

�,���2 and ����). In���1 we note a cleardecrease in the solventaccessible hydrophobicarea around 900 ps, fol-lowed by a sharp in-crease. Eventually thearea is similar to thevalue at t=0. Thus, insome peptides there is a

clear hydrophobic e�ect which seems to drive the conformational changes that occur inour simulations. When we average the solvent accessible hydrophobic surface area perresidue over the simulation (Fig. 6.11), we see that some residues are protected fromsolvent almost completely in the larger peptides whereas they are exposed in the smallerones.

This applies especially to the sequence Ile23 through Gly28 comprising the �rst �-strand and Val48 through Val51 which largely makes up the second �-strand that areexposed to solvent in the ���1 unit, while they are protected in the larger ������peptide.

Page 18: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 99

20 30 40 50 60 70 80 90Residue

012012012012012012012012012012

Are

a (n

m2 )

β1

α1

βα

β2

βαβ1

βαβ−

α2

βαβ2

βαβα

βαβαββ

Figure 6.11: Average solvent accessible surface area per residue. Filled bars represent hydrophilicresidues, Open bars represent hydrophobic residues.

6.3.6 Salt bridges

It is interesting that oppositely charged residues can be found at the start of the �rst(Lys22) and the second strand (Asp46, Glu47) of the parallel �-sheet in ���1. In thecrystal structure of LDH Glu47 forms a salt bridge with Lys22, while the side chain ofAsp46 (O�1) forms a hydrogen bond with the backbone NH of Glu47.

From the beginning of simulation ���1 we see salt bridges occurring intermittentlybetween Lys22 and Asp46 or Glu47 (Fig. 6.12). Although the Lys22 to Glu47 salt bridgeis present most of the time before 1.0 ns, Glu47 is replaced by Asp46 for some time around300 ps; after 1.0 ns Asp46 and Glu47 both form saltbridges with Lys22 but neither ofthem is very persistent.

In simulation ����, where we deprotonated the Lys22 side chain from NH+

3to an

NH2 group, hydrogen bonds between Lys22 and Glu47 also occur, but less frequently. A

Page 19: University of Groningen Structure and dynamics of peptides

100 Structure and Dynamics of Peptides

hydrogen bond between Asp46 and Lys22 is never seen (Fig. 6.12). The large peak inthe distance between Lys22 and Glu47 of more than 1 nm coincides with a temporaryreduction in the number of hydrogen bonds in the �-sheet. From this �gure it can also beseen that the motion of the negative side chains is coupled in this simulation, they moveas a couple all the time, whereas in simulation ���1 there is no such coupling. This maybe because Glu47 is very much immobilized by the salt bridge whereas the Asp46 sidechain is still free to move. In the ���� and ������ simulations the salt bridge betweenLys22 and Glu47 is present almost continuously (Fig. 6.12), implying that the residuesare increasingly �xed in their position due to surrounding residues. In sim. ���� theAsp46 to Lys22 is sometimes present simultaneously with the Glu47 to Lys22 saltbridge.

6.4 Discussion

It has often been suggested that protein folding is a spontaneous process that starts duringprotein synthesis on the ribosome [1]. There has been some controversy on the topic how-ever, because there is currently not a lot of direct experimental evidence to support thisnotion. Nevertheless Tsou [244] has strongly argued that protein folding during biosyn-thesis is at least partially cotranslational, while the extent of folding may vary for di�erentproteins. It has also been found that the fully unfolded form of a complete polypeptidedoes not exist within the living cell [32]. In recent years some evidence for cotranslationalfolding has been provided; for example Fedorov [245] demonstrated the correct folding ofthe N-terminal end of the � subunit of tryptophane synthetase during its synthesis onthe ribosome. In addition, it has been demonstrated that the ribosome itself can inducefolding [266, 267]. Proteins with repeating �� structures with parts of their C-terminalends truncated, such as ras-protein [246], phosphoglycerate kinase [247], glyceraldehyde-3-phosphate dehydrogenase [249] or phosphoribosyl anthranilate isomerase [248], all foldinto a stable native like structure. Interestingly, in the case of the latter enzyme, thecarboxy-terminal fragment was unstructured by itself, but rapidly acquired its nativestructure upon binding to the N terminal domain. Experimental work has shown thatN-terminal fragments of porcine muscle lactate dehydrogenase can bind dinucleotides,indicating that the Rossmann fold is present in these fragments [256, 268]. Althoughalso C-terminal fragments of LDH were found to grossly possess the native structure,the stability of these fragments is drastically reduced compared to the native protein[33]. Taken together these observations suggest that the N terminal region of these pro-teins might function as a folding nucleus upon which the remainder of the structure isassembled. Recent MD simulation work on amphiphilic �-helical peptides in aqueoussolution [36, 40, 44, 202, 213, 259] have shown that most isolated �-helices have a tendencyto unfold in less than 200 ps under these conditions. In another case, of myoglobin �-helices, this was less clear; only one of these (helix F) unfolds within one nanosecondat 298 K, of the other �-helices helix A, G and H seem to be most stable [67]. In yetanother simulation study of the Helix I/Loop I fragment of Barnase, which is assumed tobe an initation site for folding of the enzyme, the authors found the �-helical part to be

Page 20: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 101

rather stable, mainly due to hydrophobic interactions [260]; from their simulation datathe authors build a detailed folding pathway for the �-helix. Nevertheless, most pep-tide �-helices are unstable in aqueous solution; water penetration and insertion into thepeptide backbone appears to be the major mechanism that gives rise to helix instability[238, 259].

We have found that the

0 500 1000Time (ps)

0.0

0.5

1.0

Lys22−Asp46Lys22−Glu47

0.0

0.5

1.0

0 500 1000 1500 2000Time (ps)

0.0

0.5

1.0

1.5

0.0

0.5

1.0βαβ1

βαβ−

βαβα

βαβαββ

Dis

tanc

e (n

m)

Dis

tanc

e (n

m)

Figure 6.12: Distance between charged groups of side chains ofLys22 and Asp46 and between Lys22 and Glu47 during the MDsimulations. A running average over 25 ps is given to improve

clarity.

stability of the isolated �1-helix in water is signif-icantly higher than most�-helices in MD simula-tions; the hydrophobic sur-face of the �-helical pep-tide apparently shields thebackbone from water inser-tion. In contrast, the am-phiphilic �2-helix of LDHsimulated in water is rela-tively unstable and behavesvery similar to other �-helical peptides in MD sim-ulations. In Fig. 6.7 itcan be seen that both endsof the peptide are some-times �-helical, most no-tably the C-terminal end.In the center of this �-helixis a Gly residue which fa-cilitates water insertion inthe �-helix backbone. Thestability in the �1-helix islikely to be real, because ofthe hydrophobic characterof the peptide. Althoughsimulation �1 is not equi-librated after 2.0 ns, thesame �-helix in simulation�� fully unfolds and subse-quently refolds again whichstrongly suggests that the �-helical structure is most favorable for this sequence. Dillet. al. [34] have suggested that such structures could form and be stabilized through hy-drophobic contacts between residues in close proximity in a polypeptide chain. Indeed itseems that upon refolding the hydrophobic surface area of the �� peptide is reduced (see

Page 21: University of Groningen Structure and dynamics of peptides

102 Structure and Dynamics of Peptides

Fig. 6.10). Recent MD simulations of a very hydrophobic membrane spanning peptide inwater have revealed that such �-helical peptides are of comparable stability to the LDH�1-helix [225]; the A helix of myoglobin which is very stable in a simulation also has avery hydrophobic sequence in the middle of the peptide spanning residues W7-A15 [67].

MD simulations of short peptides in water have demonstrated that turn-like structurescan rapidly form, disappear and reform [213, 214]. This notion has been con�rmed byNMR studies of short turn-forming peptides [212, 269, 270]. Indeed our simulations ofthe isolated �1, �2 and also �2-helix show similar behaviour. In the simulations of longerpeptides, the most exible regions are generally the turn forming regions and the ends ofthe peptides. All proteins with a nucleotide binding fold contain a Gly-rich region betweenthe �1 and �1-helix that forms a characteristic turn in the �nal structure [251, 271]. Inthe larger structures, particularly in the ������ peptide the exibility of this region issubstantially reduced. We believe that the glycines in this region may form a exiblejoint that facilitates the overall folding of the ���1 unit. Folding studies of turn formingpeptides are in agreement with this, as they show that a central Gly residue allows thetwo ends of the peptide to approach each other faster [272]. It is interesting to notethat the �rst 45 residues of adenylate kinase, which contain its nucleotide binding foldand its Gly-rich loop, can fold into a structure of two helices and three �-strands, whichresembles the folding in the intact protein [273]. Like the intact protein this peptidecan bind MgATP which stabilizes the fold [274]. In our simulation of the �� unit theN-terminal strand is connected to the helix by the Gly-rich loop. Due to this loop we seemuch more exibility in the �1-helix than in the isolated �1-helix simulation.

Most MD simulations in water performed to date, have focused on well characterizedsmall globular proteins or isolated peptides with elements of secondary structure. Herewe have simulated isolated elements of supersecondary structure derived from LDH. Wenote that in our simulations the secondary and tertiary structure of the ���1 and ���2units is quite well preserved. In general, the structure in these characteristic units ofthe Rossmann fold seems less well preserved than in compactly folded proteins, but sig-ni�cantly better than in linear peptides. In none of our simulations does a break occurin the �-sheet; once formed, this unit seems extremely stable. Perhaps this stability isrelated to the need for cooperative and simultaneous disruption of three or more hydrogenbonds, as was also found from a model study [275]. In addition, hydrophobic contactsbetween the �-sheet and the �1-helix contribute signi�cantly to the stability of this su-persecondary structure (for contribution of electrostatics, see below). In the absence ofMD simulations of isolated �-sheets it is di�cult at this stage to interpret our data ingreat detail. However, it has been reported that the �-sheet of lysozyme for example isvery resistant to denaturation as well [276]. In order to compare our data for the ���1unit we also simulated the ���2 unit of LDH. This unit is also comparatively stable,although somewhat less so, because the �2-helix is not fully preserved unlike the �1-helixin the ���1 unit. However, also in the ���2 simulation the small �-sheet is stable anddoes not come apart. Again hydrophobic contacts between �-helix and �-sheet stabilizethis unit.

Page 22: University of Groningen Structure and dynamics of peptides

N-terminal Peptides from LDH 103

A minimal requirement for a folding nucleus is that it can stabilize structure in polypep-tide segments that are subsequently added to it. Indeed the ���1 unit of LDH ful�lls thisrequirement (see Results and below). Since the ���1 unit is the �rst N-terminal supersec-ondary structure element that could fold into a stable structure, it can in principle ful�llthe role of an N-terminal folding nucleus in LDH. Even when later parts of the sequencewould be more stable thermodynamically in an isolated peptide, during synthesis on theribosome they are formed later; when our ���1 unit does indeed form during transla-tion, this will be the nucleus upon which the remainder of the protein folds. Clearly,studies with synthetic peptides are required to verify this behaviour experimentally. Pro-vided that solubility of these peptides is su�cient, such studies should demonstrate that a\nascent" ���1 unit can form in aqeous solution, analogous to the well-known \nascent"�-helix formation [66, 212].

A simulation of the ���� unit was set up with the purpose to investigate the role ofthe salt bridge between Lys22 and Glu47. Removal of this electrostatic interaction givesrise to a destabilization of the �1-helix in the ���1 unit (see Fig. 6.7), underscoring theimportance of this interaction for the overall structural stability of the ���1 unit. It ispossible that the long-range electrostatic interaction between Lys22 and the combinationAsp46 and Glu47 can help to drive the formation of the �-sheet of ���1. Studies withmutated proteins and synthetic peptides have indeed shown that attractive electrostaticinteractions can place secondary structure elements in their correct orientation; moreoverthey can even accelerate their folding [277{279].

We have also simulated regions that extend beyond the �rst ��� unit. We note thatthe �2-helix is notoriously unstable on its own, however as part of the ���� structureit gains considerably in structural stability. When further elements of the structure areadded to form the ������ structure also the extra �-strands are stable in the simulation.The reduction in solvent accessible surface area of the �-sheet and �-helix regions in the������ unit also contributes to a remarkable stabilization of these units. This is relatedto the more e�cient internal packing of hydrophobic residues that can occur in largerunits.

The inaccurate treatment of long-range electrostatic interactions requires further con-sideration. It has been well established that for uncorrelated dipoles (such as in water)the use of a cut-o� of 1.0 nm is acceptable, but if full charges are present artefacts mayoccur [197]. For distances below the cut-o� the charge interactions may be somewhatexaggerated due to the neglect of screening e�ects and the omission of electronic polaris-ability. For distances above the cut-o� the interactions are ignored. In aqueous solutionscreening is such that e�ective charge interactions are very small: this is clearly exempli-�ed by the di�erence in pK values for the two acid groups in adipic acid which are at adistance of approx. 0.6 nm and di�er by 0.02 pK unit, corresponding to a negligable 0.12kJ mol�1.

While our MD simulations do not prove in themselves that folding of this part of theprotein occurs in a cotranslational manner, our results are consistent with this suggestion,because we establish here that from the ��� unit onwards, the folded portion is relatively

Page 23: University of Groningen Structure and dynamics of peptides

104 Structure and Dynamics of Peptides

stable. This is a necessary minimum requirement for cotranslational folding. The outcomeof our studies hints at the possibility that the folding of most proteins containing aRossmann fold, which all share amino-acid homology, can in principle proceed in a similarcotranslational fashion. This agrees with other work which indicates that the �rst 50amino acids of many dehydrogenases contain a folding unit [255]. Whether this processcan play a role in the folding of other classes of proteins remains to be determined.However, a statistical analysis of the compactness, number of neighbouring amino acids,and stability of secondary structure in a large number of proteins suggests that it may bewidespread [280].

6.5 Conclusions

Our MD simulations show that the �1-helix of LDH is rather stable, when compared tosimilar simulations of a range of �-helices. The hydrophobic character of this helix makesit an ideal candidate for spontaneous formation through hydrophobic collapse [34]; itshydrophobic surface can subsequently interact with other hydrophobic residues of the �1and �2 strands as noted throughout our simulations. The ���1 unit is the �rst supersec-ondary structure element in the sequence of LDH; this unit is stable in our simulations, inparticular its �-sheet never falls apart and the �1-helix is not destabilized. Its formationmay be facilitated by the presence of the Gly-rich loop; indeed proper refolding of the�1-helix in the �� unit could be observed in our simulations. The ���1 unit is further sta-bilized by the formation of the �-sheet and the cross-strand ion pair. If it is stable, it canfunction as a nucleus upon which the remainder of the structure can be assembled [254]:this is demonstrated by the markedly increased structural stability of the �2-helix in the���� unit compared to the simulation of the isolated �2-helix. Also the strands �3 and�4 are extremely stable when simulated as part of the ������ unit; as isolated units,they would probably behave like the structurally exible �1 and �2 strands. These MDresults suggest a potential pathway for cotranslational folding of the nucleotide bindingfold. Future high resolution NMR and CD experiments with N-terminal peptides of LDHwith increasing length should test the validity of this proposal.

Acknowledgements

The authors would like to thank Dr. A.R. van Buuren and Dr. R.M. Scheek for usefulcomments. DvdS acknowledges support from the Netherlands Foundation for ChemicalResearch (SON) with �nancial aid from the Netherlands Organization for Scienti�c Re-search (NWO). HJV acknowledges support from the Medical Research Council of Canadaand from the Netherlands Organization for Scienti�c Research (NWO) for a visiting pro-fessorship.