helix conformations

Upload: vaishali-shukla

Post on 13-Apr-2018

366 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/26/2019 Helix Conformations

    1/13

    The role of-, 310

    -, and -helix

    in helixcoil transitions

    ROGER ARMEN, DARWIN O.V. ALONSO, ANDVALERIE DAGGETT

    Department of Medicinal Chemistry, University of Washington, Seattle, Washington 98195, USA

    (RECEIVEDNovember 26, 2002; FINALREVISION March 18, 2003; ACCEPTEDMarch 21, 2003)

    Abstract

    The conformational equilibrium between 310

    - and -helical structure has been studied via high-resolu-

    tion NMR spectroscopy by Millhauser and coworkers using the MW peptide Ac-AMAAKAWAAKA

    AAARA-NH2. Their 750-MHz nuclear Overhauser effect spectroscopy (NOESY) spectra were interpreted

    to reflect appreciable populations of 310-helix throughout the peptide, with the greatest contribution at theN and C termini. The presence of simultaneous N(i,i+ 2) and N(i,i+ 4) NOE cross-peaks was proposed

    to represent conformational averaging between 310

    - and -helical structures. In this study, we describe

    25-nsec molecular dynamics simulations of the MW peptide at 298 K, using both an 8 and a 10 force-shifted nonbonded cutoff. The ensemble averages of both simulations are in reasonable agreementwith the experimental helical content from circular dichroism (CD), the 3JHNcoupling constants, and the57 observed NOEs. Analysis of the structures from both simulations revealed very little formation ofcontiguousi i +3 hydrogen bonds (310-helix); however, there was a large population of bifurcated i i+3 and i i+4 -helical hydrogen bonds. In addition, both simulations contained considerable popu-lations of-helix (i i+5 hydrogen bonds). Individual turns formed over residues 19, which we predictcontribute to the intensities of the experimentally observedN(i,i+2)NOEs. Here we show how samplingof both folded and unfolded structures can provide a structural framework for deconvolution of the con-formational contributions to experimental ensemble averages.

    Keywords: Molecular dynamics; -helix; 310-helix; force-shifted cutoff; conformational ensemble; helix-coil transition

    The 310-helix is the fourth most common type of secondarystructure in proteins after -helices, -sheets, and reverseturns (Barlow and Thornton 1988). Approximately 15%20% of all helices in protein structures are 310-helices,which are commonly found as N- or C-terminal extensionsto an -helix (Barlow and Thornton 1988). 310-Helices inproteins are typically only three to five residues long com-pared with a mean of 1012 residues for -helices (Rich-

    ardson and Richardson 1988). Their formation can be in-troduced by c

    -q-disubstituted amino acids, such as -ami-

    noisobutyric acid (AIB; Karle and Balaram 1990). Such

    peptides crystallize into pure 310-, pure-, or mixed 310/-helices depending on their length and relative AIB content,and these authors concluded that a six-residue helical pep-tide with only L amino acids is equally likely to form 310-and -helix (Karle and Balaram 1990).

    310-Helices have been proposed to be intermediates in thefolding/unfolding of -helices (Millhauser 1995). The ra-tionale for this suggestion is that there is a lower entropic

    penalty for the necessary loop closure for the formation ofii+3versusii+4hydrogen bonds. Helix-coil theoryhas been modified to include the 310-helix, indicating that310-helix should be populated in the helix-coil transition(Sheinerman and Brooks 1995; Rohl and Doig 1996). Thesemodified helix-coil theories also predict that 310-helical seg-ments will be short. There is also experimental evidencefrom solution electron spin resonance (ESR) spectroscopythat the population of 310-helix in an-helical peptide dem-

    Reprint requests to: Valerie Daggett, Department of Medicinal Chem-istry, University of Washington, Seattle, Washington 98195, USA; e-mail:[email protected]; fax: (206) 685-3252.

    Article and publication are at http://www.proteinscience.org/cgi/doi/10.1110/ps.0240103.

    Protein Science(2003), 12:11451157. Published by Cold Spring Harbor Laboratory Press. Copyright 2003 The Protein Society 1145

  • 7/26/2019 Helix Conformations

    2/13

    onstrates length dependence, with longer peptides favoringthe-helical conformation (Fiori et al. 1993), which is alsopredicted by the modified helix-coil theories (Sheinermanand Brooks 1995; Rohl and Doig 1996). For L amino acids,there is not a disallowed region between the -helical( 57, 47) and 310-helical ( 49, 27) conformations, allowing for free interconversionbetween the two. However, in AIB peptides, the steric in-teractions from the two methyl groups on the -carbon re-sult in the 310-helix geometry being favored energeticallyover the-helix (Zhang and Hermans 1994). Previous shortsimulations have addressed the /310-helix equilibrium inboth AIB-rich peptides (Zhang and Hermans 1994) and ala-nine-based peptides (Tirado-Rives et al. 1993). The resultsof these studies indicate that the 310-helical conformation isnot favored in L amino acids, although 310-helical interme-diates have been observed in transitions between -helicaland nonhelical states (Daggett and Levitt 1992).

    Millhauser and coworkers (1997) have investigated the

    /310-helix equilibrium in an alanine-based peptide withNMR spectroscopy. The MW peptide was designed to re-duce signal overlap for high-field NMR (Ac-AMAAKAWAAKAAAARA-NH2). This peptide is a variant of the 3Kpeptide (Marquesee et al. 1989), which has been the subjectof many ESR and fourier transform infrared (FTIR) studies(Todd and Millhauser 1991; Fiori et al. 1993, 1994; Miicket al. 1993, Smythe et al. 1995). The NOESY spectra of theMW peptide were acquired at 750 MHz (pH 5.0, 274 K) andrevealed 57 NOEs and 3JHNcoupling constants for 13 ofthe 16 residues (Millhauser et al. 1997). From the circulardichroism (CD) spectra, the peptide is expected to be48%helical at 274 K and 25% helical at 298 K (Millhauser et al.

    1997). The MW peptide displays both of the standardbenchmark NOEs used to distinguish 310-helix [N(i, i+2)]and -helix [N(i, i+4)]. Many helical peptides reflectboth connectivities simultaneously (Osterhout et al. 1989;Bradley et al. 1990; Merutka et al. 1993) and structuralinterpretation of these NOEs can be ambiguous. Isolatedturns are an expected component of random coil structureand may also contribute to the intensity of N(i, i+2)NOEs (Chandrasekhar et al. 1991).

    Molecular dynamics (MD) trajectories that sample bothhelical and nonhelical conformations can be useful in inter-preting the structural basis for ensemble averaged observ-ables like the N(i, i+2) NOESY cross-peaks. Conforma-

    tional ensembles of small peptides can be calculated withdistance-restrained MD simulations and/or simulated an-nealing with distance and torsional restraints taken fromNMR spectra (Gratias et al. 1998). Often the lack of ex-perimental distance restraints limits the usefulness of thesecalculated structures for interpreting structurally ambiguouscross-peaks. Another approach is to perform unrestrainedMD simulations that can sample a variety of conformationalstates and compare them with experiment. Here we take the

    latter approach and describe 25-nsec, room temperature mo-lecular dynamics simulations of the MW peptide starting asan-helix, using two different nonbonded cutoffs (8 and10 ). In addition, we describe the mechanisms of transi-tions between helical and nonhelical conformers and theimportance of 310- and -helix in this process.

    Results

    -Helical content and comparison with

    circular dichroism data

    The helical content during the simulation of the MW pep-tide was calculated based on two different properties: (1)repeating () helical structure and (2) the percentage ofintact i i+4 hydrogen bonds (O . . . HN distance 2.6; Fig. 1). In the dihedral definition, three or more con-

    secutive residues had to be in the helical region of confor-mational space (100 30 and 80 5;Daggett et al. 1991; Daggett and Levitt 1992). This defini-tion is very broad; the ideal values of-helix ( 57, 47), 310-helix ( 49, 27), and -helix( 57, 70) all fall into this region. For com-parison, the -region is defined as (170 50 and80 170 ).

    The peptide unfolded and refolded over the course of the8 cutoff trajectory and sampled nonhelical conformations.In contrast, the peptide remained very helical with the 10 cutoff using the dihedral metric. Both definitions of helicalstructure give similar curve shapes, but the absolute values

    and fluctuations differed over time (Fig. 1). Using the 10 cutoff, the peptide had a higher helix content using the ()definition, but the-helix content using the hydrogen bonddefinition was similar to that of the 8 trajectory.

    Our calculated helix contents cannot be compared quan-titatively with experiment because of the uncertainties instructural properties that contribute to the helical CD signaland the complications from absorbance of Trp (Chakra-bartty et al. 1993). Nevertheless, we can estimate the helixcontents from the CD spectra of the peptide, using themethod of Chakrabartty et al. (40,000[12.5/n] deg cm2/dmole and 0 deg cm2/dmole is used to represent 100% and0% helix, respectively, where n is the number of residues

    in the peptide; 1993). Using this approach, we estimate thepeptide to be48% helical at 274 K (the NMR spectra wereacquired at 274 K) and 24% at the simulation temperatureof 298 K. For the 8 cutoff, the helical content was41% 23% for the () definition, 38% 21% by the hy-drogen bond distance definition, and 27% 18% if thehydrogen bonds were also required to be within 45 oflinearity. Such alignment has been suggested to be impor-tant to CD signals (Gans et al. 1991). For the 10 cutoff,

    Armen et al.

    1146 Protein Science, vol. 12

  • 7/26/2019 Helix Conformations

    3/13

    the helical content was 85% 14% for the () definition,53% 17% by the hydrogen bond distance definition, and36% 19% with the angular constraint added. The helixcontent, as measured by i i+4 hydrogen bonds, is closeto the experimental value (24% at 298 K) for both the 8 (27%) and 10 (36%) cutoff simulations.

    Comparison with high-resolution NMR data

    Experimental3JHNscalar coupling constants (Millhauser etal. 1997) were used to calculate the average dihedralangles using the Karplus relationship (3JHN 6.4 cos

    2

    1.4 cos + 1.9, where | 60|; Karplus 1959). Aconservative estimate for the uncertainty of converting3JHNtousing this relationship is at least 0.7 Hz (Karplus1959; Wuthrich 1986). The experimental uncertainty in themagnitude of 3JHNconstants for the MW peptide is also inthe range of 0.5 Hz (Millhauser et al. 1996), and the fluc-tuations in calculated 3JHNconstants are of the same order

    of magnitude. The 1- to 25-nsec ensemble-averaged ()dihedral angles are presented with their fluctuations in Fig-ure 2, along with the experimental values for comparison.The () angles of the peptide in the 10 trajectory fluc-tuated much less compared with the 8 trajectory, as wouldbe expected for a predominantly helical ensemble. The angles explored in both simulations were in the range of theexperimentally derived values with the exception of Ala 6(Fig. 2). Averaging over this same time period, the r6

    weighted distances calculated for the 57 experimental NOEswere in good agreement (5 ) with experiment for boththe 8 and 10 trajectories with only a few minor viola-tions, all of which are within the cutoff if the standarddeviation in the distance is considered (Fig. 3). Because theNMR spectra were acquired at 274 K, the level of agree-

    ment between simulation and experiment is acceptablegiven that the peptide is known by CD to sample morenonhelical conformations at this temperature.

    Calculation of Zimm-Bragg helix-coil parameters

    In Zimm-Bragg helix-coil theory (Zimm and Bragg 1959),the propagation parameter s is the equilibrium constant be-tween an existing helical segment and the addition of a newii+4hydrogen bond. During an MD simulation, we canestimate this equilibrium constant by considering waitingtimes, , for the hcc hhc and hhc hhh (propagation)and the hhchcc and hhhhhc transitions as an average

    over time (Daggett et al. 1991; Daggett and Levitt 1992),where h indicates that a residue is in the helical region of() space and c means that it is not. The waiting timesreflect the populations of the different conformations, andthe ratio of the average waiting times is used to calculate theequilibrium constant between the two populations.

    sMD

    = Khcchhc

    =hhchcc

    hcchhc

    =khcchhc

    khhchcc

    Figure 1. Percentage of-helical structure for the 8 and 10 cutoff simulations. A and Cwere calculated using repeating helical() angles, andBandDwere calculated as a percentage of native (i i+4)hydrogen bonds. The plots in panels CandDhave beensmoothed.

    -, 310-, and -helixcoil transitions

    www.proteinscience.org 1147

  • 7/26/2019 Helix Conformations

    4/13

    The average waiting times to add a helical residue (e.g., hhchhh) were faster using the 10 than the 8 cutoff (0.4versus 3.4 psec), whereas thehhh hhcvalues were similar(6.6 versus 7.1 psec). These differences are due to theheightened stability of the helix using the longer cutoff(more details provided following). The peptide rarely left

    the helical basin of () space, so that excursions to coilwere very minor.

    Our average values ofsMDfor Khcc hhcand Khhc hhhfor the alanine residues were 1.5 and 2.0, respectively, usingthe 8 cutoff and 5.4 and 15.9 using 10 . Typical experi-mental values for s in alanine-based peptides like the oneinvestigated here are in the range of 1.5 to 2.0 (Marqusee et

    al. 1989; Lyu et al. 1990; Rohl et al. 1992), although thelysine residues used to solubilize these peptides appear tocontribute to the higher value compared with a value of1obtained using other background host peptides (Scheraga etal. 2002). The calculated sMDvalues for the 8 trajectoryare within the experimentally determined range of values,whereas they are too large for the 10 trajectory. With ourvalues ofsMDand the average helical contents (), we canestimate the macroscopic nucleation parameter, , fromequation 3b of Zimm and Bragg (1959):

    =n 3s 1 2 + n 3s 1 + 2ssn+2

    n 3s 11 + s 12 sn+1 n 3s 1 + ssn+2

    Using the hydrogen bond definition of-helical content, thecalculated macroscopic was 1.8 104 and 5.8 1013

    using the 8 and 10 cutoffs, respectively. For the 8 trajectory, the values for are within the range typical-

    Figure 2. Average dihedral () angles (averaged from 125 nsec) for both the 8 and 10 cutoff simulations. Experimental 3JHNcoupling constants were converted to using the Karplus relation and are shown in bold for comparison. The error bars on theexperimental values reflect a variation of 1 Hz.

    Figure 3. Summary of the experimentally observed NOEs and corre-sponding distances from the simulations. The distances between protonswere calculated as an ensemble average over the simulation by usingweighted averages, (rw) r61/6) where rw is the weighted distanceand ris the actual distance between the two hydrogens of interest. Thecalculated NOE weighted distances (averaged over 125 nsec) are shownfor the 8 cutoff simulation (top) and the 10 cutoff simulation (bottom).

    Armen et al.

    1148 Protein Science, vol. 12

  • 7/26/2019 Helix Conformations

    5/13

    ly assumed by experimentalists (1 103 to 1 105;Marqusee et al. 1989; Lyu et al. 1990; Rohl et al. 1992;Scheraga et al. 2002).

    Structural transitions and peptide hydrogen bonding

    The local region of conformational space occupied by eachresidue is shown for -helical (cyan) and -structure (red)as a function of time in Figure 4. There were many confor-mational transitions throughout the 8 cutoff simulationand fewer using the 10 cutoff. The most significant lossof -helical structure occurred from 9.8 to 16.5 nsec andagain after 21 nsec using an 8 cutoff. These transitionsinvolved numerous residues in the center of the sequence. Inaddition, large deviations from the helical state occurrednear the N terminus: Ala 1, Ala 3, and Ala 4 sampledextended and-regions of conformational space, which ledto the formation of transient -turns throughout the trajec-

    tory (Table 1).The total number of 310-,-, and-hydrogen bonds overtime are shown in Figure 4. For both simulations, the struc-tures from 0 to 10 nsec were primarily -helical, and thegreatest population of-helical conformers was from 10 to20 nsec. Both simulations exhibited major transitions from and , indicating that this transition is revers-ible on the 25-nsec timescale. The structures from 10 to 20nsec contained a mixture of 310-, -, and -hydrogen bonds,as well as turns (Fig. 4).

    Figure 5 displays the percentage of total simulation timethat hydrogen bonds were satisfied (O . . . H distance 2.6) for every possible 310, -, and -hydrogen bond in the

    sequence in both simulations. For the 8 cutoff, 310-hy-drogen bonding had the lowest population with 1.6 1.5hydrogen bonds, or 16% of the hydrogen bonds made by thepeptide. This was followed by -hydrogen bonding(3.2 1.5, or 30% of the hydrogen bonds) and -helicalhydrogen bonding (5.4 1.9 or 41%). For the 10 cutoff,310-hydrogen bonding had the lowest population with1.3 1.5 hydrogen bonds, or 16% of the total hydrogenbonds made. This was followed by - (5.5 1.5, or 40% ofthe hydrogen bonds) and -helical hydrogen bonding(7.4 1.9 or 53%). Thus, both the 8 and 10 simulationshad a lower population of 310- than -helix.

    Most of the occurrences of 310-helix were isolated, with

    only oneii+3hydrogen bond in the peptide. The largestnumber of 310-hydrogen bonds at any given time was 8, andthis was observed in

  • 7/26/2019 Helix Conformations

    6/13

    Figure 4. Conformational behavior of the MW peptide. (A) Structural transitions for the 8 cutoff simulation shown as a ()conformational space plot for each residue of the peptide. Local -helical and -conformations are shown in cyan and in red,respectively, and are based on the corresponding () regions defined in the text. (B) Representative structures are shown every 2.5nsec for the 8 cutoff simulation. (C) Number of-, 310-, and-helical hydrogen bonds (O . . . H distance 2.6 ) for the 8 cutoffsimulation. (D) Structural transitions for the 10 cutoff simulation, as described earlier. (E) Representative structures from the 10 cutoff simulation. (F) Number of-, 310-, and -helical hydrogen bonds for the 10 cutoff simulation.

    Armen et al.

    1150 Protein Science, vol. 12

  • 7/26/2019 Helix Conformations

    7/13

    Sasisekharan 1968). In addition to-helical conformations,there were also a number of Schellman-like conformations(Schellman 1980; Gunasekaran et al. 1998) over the C-terminal residues 1116. These structures formed hydrogenbonds between residues 11 and 16 and 12 to 15, and residue16 was in the L-conformation. These Schellman-like con-formations were transient.

    Importance of main chainside chain

    hydrogen bonding

    In the 8 cutoff simulation, a hydrogen bond between theN-terminal cap, Ace, and Lys 5 was the dominant main

    chainside chain (mcsc) hydrogen bond. This hydrogenbond was intact before (9.0539.057 nsec), during (9.2659.267, 9.3769.393, 9.4789.634 nsec), and after (9.8019.829, 9.8469.856 nsec) the transition (9.2359.760nsec) shown in Figure 7. This interaction played a signifi-cant role in stabilizing the formation of-helical structure,as well as assisting in the formation of the first two -hy-drogen bonds by restricting the conformational space avail-able to the chain. After these-hydrogen bonds formed, the

    Ace/Lys 5 interaction facilitated the dihedral transition ofAla 1 out of the -region, resulting in the formation of thenext two -hydrogen bonds (Fig. 7).

    In a similar manner, a hydrogen bond between Ace andTrp 7 stabilized a structural transition from mixed /-helixinto -helix and turn structures around 18.5 nsec in the 8 cutoff simulation. This hydrogen bond stabilized the struc-tures both before (17.97917.98 nsec) and after (18.56218.582 nsec) the transition. In addition, this interaction sta-bilized the turns over residues 47 and residues 58. Similar

    Table 1. Percentage of simulation time in which turns

    were formeda

    Turnresiduesb

    8 Type Iturna

    (%)

    8 Type II

    turna

    (%)

    8 Most

    populatedc

    (nsec)

    10 Type Iturna

    (%)

    10 Type II

    turna

    (%)

    10 Most

    populated(nsec)

    14 60.8 35.6 19 88.9 26.8 15, 1025N(2,4)25 58.0 36.7 1020 10.9 14.7 162236 3.8 5.9 1618 0.3 9.6 162047 13.3 19.4 510, 1724 0.7 5.9 15N(5,7)58 13.5 11.6 1017 0.1 1.5 162069 0.7 1.7 48 0.2 1.0 1617N(7,9)710 0.7 1.3 1820 0 0811 0.1 0.1 2025 0 0912 0 0 0 01013 0 0 0 01114 0 0 0 01215 0 0 0 01316 0 0 0 0

    a The percentage of total simulation time (025 nsec) in which specificturns formed. Turn formation was calculated based on the distance sepa-rating i and i + 3 C

    atoms, d, and the angle formed by the vector

    connecting the i and i + 1 C

    atoms, Ri, i+1, and that for the i + 2 and i +3 atoms, Ri+2, i+3(Kuntz 1972). Using this approach, type I hairpin turnshaved5.5 and 135. Typically, type I -turns are defined by the() angles for the i + 1 and i + 2 residue in the turn: i + 1( 60, 30),i + 2( 90, 0). Type II hairpin turns have eitherd5.5 and 135 > 120 or 5.5 d6.5 and 135 . TypeII-turns are typically defined by i+ 1( 60, 120), i+ 2(80, 0).b Residues involved in turn formation. Turns that may contribute to ex-perimental N(i, i + 2) NOE cross-peaks are shown.c The time that the most populated turn was continuously intact.

    Figure 5. Diagram of possible 310-, -, and -hydrogen bonds for the MWpeptide. Above each possible hydrogen bond is the percentage of simula-tion time (averaged over 125 nsec) that it was intact (O . . . HN distance2.6 ) during the 8 (on the left) and 10 (on the right) cutoffsimulation. Each hydrogen bond reflects the percentage of the time that itsatisfies the distance criterion, regardless of bifurcation with other peptidedonors/acceptors. Using a more stringent hydrogen bonding definition in-cluding both distance (O . . . HN distance 2.6 ) and angular cutoffs(O . . . H-N angle within 45 of linearity) significantly lowered the calcu-lated population ofi i+3hydrogen bonds observed per structure. Whenthe angular constraint was applied, for the 8 cutoff simulation the av-

    erage number ofi i +3 hydrogen bonds decreased from 1.6 (using thedistance criterion alone) to

  • 7/26/2019 Helix Conformations

    8/13

    interactions and transitions were observed using the 10 cutoff (Fig. 7).

    The effect of changing the nonbonded cutoff

    on energetics

    Up to this point we have focused on the peptides structuralproperties. Although these two simulations are in reason-able agreement with the CD and NMR data, the 10 cutoffsimulation is in poor agreement with the calculated Zimm-Bragg helix-coil parameters, and it shows slower helix-coiltransitions than the 8 cutoff. We have investigated theeffects of changing the cutoff on the energetics, particularly

    because a priori one would expect the longer cutoff range of10 to be in better agreement with experiment.

    For the entire system, the average proteinprotein, pro-teinwater, and waterwater energies were compared for thefirst nanosecond of the 8 and 10 cutoff simulations(Table 2), because during this time period the two trajecto-ries were the most similar with respect to peptide confor-mation. The greatest relative changes in energy going froman 8 to a 10 cutoff were the proteinprotein and the

    proteinwater van der Waals interactions (Table 2), whichboth contribute to slower transitions in the 10 cutoffsimulation. The reason for this is that the repulsive van derWaals term is scaled as a function of cutoff distance tocompensate for the attractive interactions lost on truncation(Levitt et al. 1997). Values of 0.84 and 0.92 are used for 8 and 10 , respectively. In this way, sterically hinderedtransitions between conformational states are discouragedless with the 8 cutoff.

    A unique feature of our force-shifted potential is thathydrogen bond energies are invariant of cutoff within thecutoff range (Levitt et al. 1995). The potential energy pair-wise sum of all atoms involved in proteinprotein and pro-

    teinwater hydrogen bonds change

  • 7/26/2019 Helix Conformations

    9/13

    the shifted short (04 ) and medium-range (46 ) inter-actions. The shifts in the short-range repulsive van derWaals interactions were the most significant contributionsto the slower transitions in the 10 cutoff simulation.

    Discussion

    Recent temperature-jump experiments have shown that thekinetics of helix formation tend to be single exponential ormultiphasic with a fast phase of 1020 nsec, which isthought to represent helix propagation and fraying, and aslower phase ( 130300 nsec), which appears to repre-sent helix nucleation (Williams et al. 1996; Gilmanshin etal. 1997; Thompson et al. 1997; Huang et al. 2001a,b, 2002;Werner et al. 2002). The bulk of the experimental studiesshow an increase in the relaxation rate with temperatureindicative of an activation barrier for folding and unfolding.

    However, some recent simulation studies (Hummer et al.2000, 2001) have indicated that helix formation kinetics canbe modeled as a diffusive conformational search within thecoil state with no barrier for the transition to the helicalstate. Such a process predicts a departure from single-ex-ponential to nonexponential or stretched-exponential relax-ation kinetics in temperature-jump experiments, which israre but has been reported in some experimental studies(Sabelko et al 1999; Leeson et al. 2000; Huang et al. 2002).

    Figure 7. Four representative structures during the transition show the propagation of-helical hydrogen bonds for the 8 and 10 cutoff simulations. -Helical hydrogen bonds are shown in magenta and -helical hydrogen bonds are shown in green. Themain chainside chain hydrogen bond between Ala 1 and Lys 5 is shown in blue for the 8 cutoff simulation. The number of hydrogenbonds is given beloweach structure.

    Table 2. Average nonbonded interaction energies

    Interaction typesa

    ij

    Uij8

    (kcal/mole)

    ij

    Uij10

    (kcal/mole)

    P..P (vdw) 40.0 73.1

    P..P (els) 75.0 85.6P..W (vdw) 5.1 58.4P..W (els) 293.1 348.6W..W (vdw) 2870.3 3682.8W..W (els) 15888.3 23310.3Total (vdw) 2825.4 3551.5Total (els) 16256.0 23544.5

    a Van der Waals (vdw) and electrostatic (els) contributions to the potentialenergy of proteinprotein (P..P), proteinwater (P..W), and waterwater(W..W) interactions, and the total potential energy of the system.

    -, 310-, and -helixcoil transitions

    www.proteinscience.org 1153

  • 7/26/2019 Helix Conformations

    10/13

    However, Ferguson and Fersht (2003) recently pointed outthat these stretched-exponential kinetics occur on a time-scale that poses problems for the instrumentation. If this isthe case, such results could be artifacts, and they note thatnonexponential kinetics have disappeared when they haveremoved all artifacts. They also point out that the kineticdata can be fit within experimental error to a series of se-quential or parallel exponentials.

    Alternatively, rather than barrierless folding, such kinet-ics, if free from artifact, could be due to heterogeneity of thenonhelical/coil state. In this regard, our results for bothsimulations indicate that -helical intermediates may con-tribute to the roughness of the free energy landscape ofhelical peptides. Nevertheless, even though our moleculardynamics simulations are fairly long compared with othersimulation studies, the 25-nsec simulation period is an in-sufficient amount of time for crossover from the coil to thehelical state. Consequently, we focus on structural proper-ties of the helix coil process.

    Although our 8 and 10 cutoff simulations were simi-lar overall with respect to their ability to model the NMRresults, the peptide sampled many fewer nonhelical confor-mations using the longer cutoff. This leads to disparitiesbetween the s and parameters associated with helix-coiltheory. The values determined from the 8 simulation,however, are in good agreement with experiment. Using the10 cutoff, the peptide appears to be artificially stabilizedin the helical state at this timescale. It is possible, however,that this is nothing more than a sampling problem and thatthe simulations would converge with increasing simulationtime, and, in fact, the two simulations are similar if otherproperties are considered (such as hydrogen bonds). We

    note that the proper timescale and amplitude of motion wasobtained in simulations of barstar using the 10 cutoff asassessed by comparison with the results of NMR relaxationexperiments (Wong and Daggett 1998). The differences ob-served here are mainly due to shifts in short-range interac-tions, rather than the effects of long-range electrostatic in-teractions. For the purposes of seeing interesting conforma-tional behavior on a reasonable timescale by MD, the 8 cutoff simulations are preferred for this system. However,the longer time required for equilibration and the effect onthe barrier to conformational transitions for our methodsusing a 10 cutoff may in fact be correct. Future studieswill involve the calibration of our van der Waals screening

    term to reproduce timescales being determined by fast IRmeasurements on labeled peptides (W. DeGrado and F. Gai,pers. comm.). In addition, we note that unfolding times fromMD simulations of the engrailed homeodomain are in goodagreement with temperature-jump experiments (Mayoret al. 2003).

    Nonetheless, it is important to remember that force fieldsare empirically derived, contain many interrelated terms,and are parameterized with specific characteristics in mind.

    In this way, increasing a cutoff does not always result in thenave, expected effect. Furthermore, we show that shortercutoffs can not only be adequate, but can be better than alonger cutoff. The relative interplay of force-field terms andthe many ways in which nonbonded interactions can betruncated make generalizations like those popularized bySchreiber and Steinhauser (1992a,b), who assert that simu-lations using short cutoffs are unstable and ultimatelywrong, unsound. Although this may have been true for theirforce field and methods, we have repeated their work and donot obtain the same results with our force field and proto-cols (D.A.C. Beck, in prep.).

    Turns versus contiguous 310-helix: Interpretation of

    N(i,i+2)connectivities

    A low population ofi i+3hydrogen bonds (

  • 7/26/2019 Helix Conformations

    11/13

    transitions in the simulation were mediated through mcschydrogen bonding, which can explain how local side chaininteractions dominate local conformational preferences.Main chainside chain hydrogen bonding, such as helicalN-capping interactions, can restrict conformational spacebetween the residues in contact, and these restrictions havebeen proposed to assist in the nucleation of more stablestructure (Presta and Rose 1988; Karpen et al. 1992). In arecent example, mc-sc hydrogen bonding between a threo-nine and a carbonyl group four residues before it in thesequence leads to a shift from -helical structure to mixed310//-helical conformations, as assessed by MD simula-tions and NMR and IR experiments on a designed helicalbundle protein (Walsh et al. 2003).

    The experimental 3JHN coupling constants for the MWpeptide reflect systematically higher values (>6 Hz) for thelongerii+5spaced side chains. Millhauser and cowork-ers (1997) suggest that this is because local folding/unfold-ing originates at the positions of the longer side chains,

    which is consistent with transitions observed via MD (forexample, see Fig. 7). Both side chain mediated unfoldingand the existence of-helical intermediates can contributeto this i i+5 experimental trend in 3JHN.

    The MW peptide is a variant of canonical alanine-basedpeptides (AAAAK)n, and Shirley and Brooks (1997) havealso observed formation of -helical structure in(AAKAA)3and (AAQAA)3peptides. Related repeating se-quences (AAQA), (AAQAAA), and (AAQ) were also stud-ied, and they concluded that i i +5interactions betweenside chains promoted-helical structure formation. Similar transitions were observed in a more complicatedAla-based peptide by Lee and coworkers (2000). In this

    case, -helix seems to have been triggered by repulsiveinteractions between i i+3 spaced Glu residues, whichafter conversion show a 3 increase in the distance be-tween their carboxylates. Both of these other MD studieswere performed using CHARMM with a force-shifted po-tential, and a 10 to 12 cutoff range (Brooks et al. 1983).Despite differences in the integration method used, the cut-offs, the water models, and the sequences, both theCHARMM and ENCAD potential functions support the for-mation of -helical structure, and the -helical structurewas stabilized significantly by side chain interactions.

    Garcia and Sanbonmatsu (2002) have suggested that mcsc hydrogen bonds contribute to the stabilization of-heli-

    cal structure. Our simulations showed similar results forinteractions between Lys or Trp and the main chain of theC-terminal capping group, but, in general, rather than sta-bilizing -helical conformations, our results indicate thatthese interactions facilitate structural transitions to and fromnonhelical and -helical states. The main chain carbonylgroups of Met, Trp, and Lys showed less hydrogen bondingwith water compared with the Ala residues across the se-quence. Furthermore, there was a tendency for Ala to be

    less hydrated when following a Lys residue (Fig. 6). Thisresult is consistent with simulation studies of Scheraga andcoworkers (2002), who find that solvation of Lys sidechains decreases the solvation of the neighboring mainchain groups. Side chain shielding of the main chain fromhydrogen bonding with, and attack by, solvent may contrib-ute more to -helix stability than side chain-to-main chainhydrogen bonding, but water was still effective at reachingand interacting with carbonyl groups near lysine residues inour simulations and such interactions do not appear to behelix destabilizing per se in our simulations.

    Comparison of 310-hydrogen bonding in simulations of

    peptides and proteins

    Although we observed 310-hydrogen bonding in the simu-lations, these hydrogen bonds had lower populations than -and -hydrogen bonds and they did not play a dominantrole in unfolding/refolding events. Therefore, the question

    remains as to why there are so many 310- helices in experi-mentally determined protein structure. It is possible that oursimulations are underestimating the 310-helical content, orthat force fields in general may be energetically biasedagainst i i+3 hydrogen bonding arrangements. To testthis idea, G. Millhauser (pers. comm.) suggested that weperform a simulation of the leucine-rich repeat variant,which is a protein with a high percentage of 310-helix (Pe-ters et al. 1996). Only the repeat domain (residues 41231)was used, which is composed of a 24-residue tandem repeatpattern of-helix, turn, and 310-helix. In a 3-nsec simula-tion of this domain at 298 K, using the same 8 nonbondedcutoff, the protein retained the majority of its 310-helical

    hydrogen bonds. Thus, the low population of 310-helix inour simulations of the MW peptide does not appear to be anartifact (data not shown). Many of these 310-helical hydro-gen bonds were stabilized by side chain-to-main chain hy-drogen bonds that have been described as supporting 310-helical structure (Karpen et al. 1992). Recent quantum me-chanical studies indicate that 310-helical structure is morestable in vacuo than in water (Topol et al. 2001), indicatingthat there may be a shift from 310to -helical structure insolution.

    Conclusions

    We conclude that both the 8 and 10 cutoff simulationsare in reasonable agreement with the helix content of theMW peptide as assessed experimentally by CD, the 3JHNcoupling constants, and the NOEs. Both the 8 and 10 cutoff simulations underwent structural transitions between-helix, 310-helix, and-helix, and both simulations exhib-ited a large population of-helix. However, the 8 cutoffis in better agreement with calculated Zimm-Bragg param-eters and the CD data, and unfolds to more nonhelical con-

    -, 310-, and -helixcoil transitions

    www.proteinscience.org 1155

  • 7/26/2019 Helix Conformations

    12/13

    formers. Consequently, our results for the MW peptideshow that short cutoffs can be not only reasonable, but canalso be in better agreement with experiment than longercutoffs. Our results indicate that both N-terminal 310-helicalstructure and turns contribute to the observed N(i,i+2)NOEs. Furthermore, the simulations indicate that contigu-ous 310-helix is not populated appreciably without bifurca-tion involving simultaneous i i +4 interactions. 310-Hy-drogen bonds were involved in helix-coil transitions, as sug-gested by Millhauser and coworkers (1997), but the-hydrogen bonds were more important in this regard. Themajor structural transitions in the simulations were nucle-ated by transient mcsc hydrogen bonding and hydrophobicinteractions between side chains. This type of side chain-assisted helix formation and destruction may explain theapparent sequence dependence of-helix formation in pep-tides with i i+5 spaced side chains, as well as thespecific side chain interactions that stabilize 310-helicalstructure in proteins.

    Materials and methods

    All energy minimization and molecular dynamics was performedusing an in-house version of the program ENCAD (Levitt 1990).An all-atom representation was used for both the peptide and thewater, using a previously described macromolecular potentialfunction with a force-shifted atom-based truncation method (Levittet al. 1995), and the flexible F3C water model (Levitt et al. 1997).

    The 16-residue MW peptide was generated to have ideal helical and angles (57, 47, respectively), the N terminus wasacetylated, and the C terminus was amidated. The peptide wasminimized 500 steps in vacuo. Two different simulations wereperformed of the MW peptide using 8 and 10 cutoffs, each for25 nsec of simulation time. For the 8 cutoff simulation, the

    peptide was then solvated by adding waters extending at least 8 ,resulting in the addition of 1142 water molecules. For the 10 cutoff simulation, the peptide was solvated by adding waters ex-tending at least 10 , resulting in the addition of 1670 watermolecules. Structures were saved every 0.2 psec for analysis, re-sulting in 125,000 structures for each 25-nsec simulation.

    The box volume was adjusted to give the experimental densityof 0.997 g/mL for 298 K (Kell 1967). The solvent was prepared formolecular dynamics by performing 2000 cycles of minimization,followed by 2000 cycles of molecular dynamics, and another 2000steps of minimization. The peptide was then minimized for 500steps, followed by the minimization of the entire system for 500steps. The system was then brought to the target temperature of298 K. MD simulations were performed using the microcanonicalensemble (NVE) and periodic boundary conditions. The equations

    of motion were integrated using a modified Beeman algorithm anda 2-fsec time step (Levitt et al. 1995). The nonbonded lists wereupdated every 2 steps (4 fsec), and structures were saved every 0.2psec for analysis.

    Acknowledgments

    We are grateful for financial support provided by a research grantfrom the National Institutes of Health (GM 50789 to V.D.) and aMolecular Biophysics Training Grant (National Research Service

    Award T32 GM 08268 to R.A). We thank Drs. Glenn Millhauserand Bill DeGrado for stimulating discussions.

    The publication costs of this article were defrayed in part bypayment of page charges. This article must therefore be herebymarked advertisement in accordance with 18 USC section 1734solely to indicate this fact.

    References

    Barlow, D.J. and Thornton, J.M. 1988. Helix geometry in proteins. J. Mol. Biol.201:601619.

    Bradley, E.K., Thomason, J.F., Cohen, F.E., Kosen, P.A., and Kuntz, I.D. 1990.Studies of synthetic helical peptides using circular dichroism and nuclearmagnetic resonance. J. Mol. Biol. 215:607622.

    Brooks, B.R., Bruccoleri, R.E., Olafson, B.D., States, D.J., Swaminathan, S.,and Karplus, M. 1983. CHARMMA program for macromolecular energy,minimization, and dynamics calculations. J. Comput. Chem. 4: 187217.

    Chakrabartty, A., Kortemme, T., Padmanabhan, S., and Baldwin, R.L. 1993.Aromatic side-chain contribution to far-ultraviolet circular dichroism ofhelical peptides and its effect on measurements of helix propensities. Bio-chemistry 32: 55605565.

    Chandrasekhar, K., Profy, A.T., and Dyson, H.J. 1991. Solution conformationalpreferences of immunogenic peptides derived from the principal neutraliz-ing determinant of the HIV-1 envelope glycoprotein GP 120. Biochemistry30:91879194.

    Daggett, V. and Levitt, M. 1992. Molecular dynamics simulations of helixdenaturation. J. Mol. Biol. 233:11211138.

    Daggett, V., Kollman, P.A., and Kuntz, ID. 1991. A molecular dynamics simu-lation of polyalanineAn analysis of equilibrium motions and helix-coiltransitions. Biopolymers 31: 11151134.

    Ferguson, N. and Fersht, A.R. 2003. Early events in protein folding.Curr. Opin.Struct. Biol. 13: 7581.

    Fiori, W.R., Miick, S.M., and Millhauser, G.L. 1993. Increasing sequencelength favors -helix over 3(10)-helix in alanine-based peptides: Evidencefor a length-dependent structural transition. Biochemistry32:1195711962.

    Fiori, W.R., Lundberg, K.M., and Millhauser, G.L. 1994. A single carboxy-terminal arginine determines the amino-terminal helix conformation of analanine-based peptide. Nat. Struct. Biol. 1: 374377.

    Fodje, M.N. and Al-Karadaghi, S. 2002. Occurrence, conformational featuresand amino acid propensities for the -helix. Protein Eng. 15: 353358.

    Gans, P.J., Lyu, P.C., Manning, M.C., Woody, R.W., and Kallenbach, N.R.1991. The helix-coil transition in heterogeneous peptides with specific side-chain interactions: Theory and comparison with CD spectral data. Biopol-

    ymers31: 16051614.Garcia, A.E. and Sanbonmatsu, K.Y. 2002. -Helical stabilization by side chain

    shielding of backbone hydrogen bonds. Proc. Natl. Acad. Sci. 99:27822787.

    Gilmanshin, R., Williams, S., Callender, R.H., Woodruff, W.H., and Dyer, R.B.1997. Fast events in protein folding: Relaxation dynamics of secondary andtertiary structure in native apomyoglobin. Proc. Natl. Acad. Sci. 94: 37093713.

    Gratias, R., Konat, R., Kessler, H., Crisma, M., Valle, G., Polese, A., Formag-gio, F., Toniolo, C., Broxterman, Q.B., and Kamphuis, J. 1998. First steptoward the quantitative identification of peptide 3(10)-helix conformationwith NMR spectroscopy: NMR and X-ray diffraction structural analysis ofa fully developed 3(10)-helical peptide standard. J. Am. Chem. Soc. 120:47634770.

    Gunasekaran, K., Nagarajaram, H.A., Ramakrishnan, C., and Balaram, P. 1998.Stereochemical punctuation marks in protein structures: Glycine and prolinecontaining helix stop signals. J. Mol. Biol. 275:917932.

    Huang, C.Y., Klemke, J.W., Getahun, Z., DeGrado, W.F., and Gai, F. 2001a.Temperature-dependent helix-coil transition of an alanine based peptide. J.Am. Chem. Soc.123:92359238.

    Huang, C.Y., Getahun, Z., Wang, T., DeGrado, W.F., and Gai, F. 2001b. Time-resolved infrared study of the helix-coil transition using (13)C-labeled he-lical peptides. J. Am. Chem. Soc. 123:1211112112.

    Huang, C.Y., Getahun, Z., Zhu, Y., Klemke, J.W., DeGrado, W.F., and Gai, F.2002. Helix formation via conformation diffusion search. Proc. Natl. Acad.Sci. 99: 27882793.

    Hummer, G., Garcia, A.E., and Garde, S. 2000. Conformational diffusion andhelix formation kinetics. Physical Rev. Lett. 85: 26372640.

    . 2001. Helix nucleation kinetics from molecular simulations in explicitsolvent. Proteins 42: 7784.

    Armen et al.

    1156 Protein Science, vol. 12

  • 7/26/2019 Helix Conformations

    13/13

    Karle, I.L. and Balaram, P. 1990. Structural characteristics of-helical peptidemolecules containing AIB residues. Biochemistry 29: 67476756.

    Karpen, M.E., Haseth, P.L., and Neet, K.E. 1992. Differences in the amino-aciddistributions of 3(10) helices and -helices. Protein Sci. 1: 13331342.

    Karplus, M. 1959. Contact electron-spin coupling of nuclear magnetic moments.J. Chem. Phys. 30: 1115.

    Kell, G.S. 1967. Precise representation of volume properties of water at oneatmosphere. J. Chem. Eng. Data 12: 6669.

    Kuntz, I.D. 1972. Protein folding. J. Am. Chem. Soc. 94: 40094012.

    Lee, K.H., Benson, D.R., and Kuczera, K. 2000. Transitions from to helixobserved in molecular dynamics simulations of synthetic peptides. Bio-chemistry 39: 1373713747.

    Leeson, D.T., Gai, F., Rodriguez, H.M., Gregoret, L.M., and Dyer, R.B. 2000.Protein folding and unfolding on a complex energy landscape. Proc. Natl.

    Acad. Sci.97: 25272532.Levitt, M. 1990. ENCAD (computer program). Energy Calculations and Dy-

    namics, Molecular Applications group, Palo Alto, CA and Yeda, Rehovot,Israel.

    Levitt, M., Hirshberg, M., Sharon, R., and Daggett, V. 1995. Potential-energyfunction and parameters for simulations of the molecular dynamics of pro-teins and nucleic acids in solution. Comput. Phys. Commun. 91: 215231.

    Levitt, M., Hirshberg, M., Sharon, R., Laidig, K.E., and Daggett, V. 1997.Calibrating and testing a water model for simulation of the molecular dy-namics of proteins and nucleic acids in solution. J. Phys. Chem. B 101:50515061.

    Low, B.W. and Grenville-Wells, H.J. 1953. Generalized mathematical relationsfor polypeptide chain helixes. The coordinates for the helix. Proc. Natl.

    Acad. Sci.39: 785801.Lyu, P.C., Liff, M.I., Marky, L.A., and Kallenbach, N.R. 1990. Side chaincontributions to the stability of-helical structure in peptides. Science250:669673.

    Marqusee, S., Robbins, V.H., and Baldwin, R.L. 1989. Unusually stable helixformation in short alanine-based peptides. Proc. Natl. Acad. Sci. 86:52865290.

    Mayor, M., Guydosh, N.R., Johnson, C.M., Grossmann, J.G., Sato, S., Jas, G.S.,Freund, S.M.V., Alonso, D.O.V., Daggett, V., and Fersht, A.R. 2003. Thecomplete folding pathway of a protein from nanoseconds to microseconds.

    Nature 421:863867.Merutka, G., Morikis, D., Brushweiler, R., and Wright, P.E. 1993. NMR evi-

    dence for multiple conformations in a highly helical model peptide. Bio-chemistry 32: 1308913097.

    Miick, S.M., Casteel, K.M., and Millhauser, G.L. 1993. Experimental moleculardynamics of an alanine-based helical peptide determined by spin label elec-tron spin resonance. Biochemistry 31: 80148021.

    Millhauser, G.L. 1995. Views of helical peptides: A proposal for the position of

    3(10)-helix along the thermodynamic folding pathway. Biochemistry 34:38733877.Millhauser, G.L., Stenland, C.J., Bolin, K.A., and van de Ven, F.J.M. 1996.

    Local helix content in alanine-rich peptide as determined by the compete setof (3)J(HN ) coupling constants. J. Biomol. NMR 7: 331334.

    Millhauser, G.L., Stenland, C.J., Hanson, P., Bolin, K.A., and van de Ven F.J.M.1997. Estimating the relative populations of 3(10)-helix and -helix inAla-rich peptides: A hydrogen exchange and high field NMR study.J. Mol.

    Biol.267:963974.Osterhout, J.J., Baldwin, R.L., York, E.J., Stewart, J.M., Dyson, H.J., and

    Wright, P.E. 1989. 1-H NMR studies of the solution conformations of ananalog of the C-peptide of ribonuclease A. Biochemistry 28: 70597064.

    Peters, J.W., Stowell, M.H., and Rees, D.C. 1996. A leucine-rich repeat variantwith a novel repetitive protein structural motif. Nat. Struct. Biol. 3: 991994.

    Presta, L.G. and Rose, G.D. 1988. Helix signals in proteins. Science 240:16321641.

    Ramachandran, G.N. and Sasisekharan, V. 1968. Conformations of polypep-

    tides and proteins. Adv. Protein Chem. 23: 283438.

    Richardson, J.S. and Richardson, D.C. 1988. Amino-acid preferences for spe-cific locations at the ends of-helices. Science 240:16481652.

    Rohl C.A. and Doig A.J. 1996. Models for the 3(10)-helix/coil, pi-helix/coil,and-helix/3(10)-helix/coil transitions in isolated peptides. Protein Sci.5:16871696.

    Rohl, C.A., Scholtz, J.M., York, E.J., Stewart, J.M., and Baldwin, R.L. 1992.Kinetics of amide proton exchange in helical peptides of varing chainlengths. Interpretation by the Lifson-Roig equation. Biochemistry31:12631269.

    Sabelko, J., Ervin, J., and Gruebele, M. 1999. Observation of strange kinetics inprotein folding. Proc. Natl. Acad. Sci. 96: 60316036.Schellman, C. 1980. The

    L-conformation at the end of helices. In Protein

    folding(ed. R. Jaenicke), pp. 5361. Elsevier, New York.Scheraga, H.A., Vila, J.A., and Ripoll, D.R. 2002. Helix-coil transitions re-

    visited. Biophys. Chem. 101102:255265.Schreiber, H. and Steinhauser, O. 1992a. Cutoff size does strongly influence

    molecular dynamics results on solvated polypeptides. Biochemistry 31:58565860.

    . 1992b. Molecular dynamics studies of solvated polypeptides: Why thecutoff scheme does not work. Chem. Phys. 168:7589.

    Sheinerman, F.B. and Brooks, C.L. 1995. 3(10)-Helices in peptides and proteinsas studied by modified Zimm-Bragg Theory. J. Am. Chem. Soc. 117:1009810103.

    Shirley, W.A. and Brooks, C.L. 1997. Curious structure in canonical alaninebased peptides. Proteins 28: 5971.

    Smythe, M.L., Nakaie, C.R., and Marshall, G.R. 1995. -helical versus 3(10)-helical conformations of alanine-based peptides in aqueous solutionAn

    electron-spin-resonance investigation. J. Am. Chem. Soc. 117:1055510562.Soman, K.V., Karimi, A., and Case, D.A. 1991. Unfolding of an -helix inwater. Biopolymers 31: 13511361.

    Thompson, P.A., Eaton, W.A., and Hofrichter, J. 1997. Laser temperature jumpstudy of the helix-coil kinetics of an alanine peptide interpreted with akinetic zipper model. Biochemistry 36: 92009210.

    Tirado-Rives, J., Maxwell, D.S., and Jorgensen, W.L. 1993. Molecular dynam-ics and Monte Carlo simulations favor the -helical form for alanine-basedpeptides in water. J. Am. Chem. Soc. 115:1159011593.

    Todd, A.P. and Millhauser, G.L. 1991. ESR spectra reflect local and globalmobility in a short spin-labeled peptide throughout the -helixcoil transi-tion. Biochemistry 30: 55155523.

    Topol, I.A., Burt, S.K., Deretey, E., Tang, T.H., Perczel, A., Rashin, A., andCsizmadia, I.G. 2001. and 3(10)-helix interconversion: A quantum-chemical study on polyalanine systems in the gas phase and in aqueoussolvent. J. Am. Chem. Soc. 123:60546060.

    Walsh, S.T.R., Cheng, R.P., Wright, W.W., Alonso, D.O.V., Daggett, V.,Vanderkooi, J.M., and DeGrado, W.F. 2003. The hydration of amides in

    helices: A comprehensive picture from molecular dynamics, IR, and NMR.Protein Sci. 12: 520531.Weaver, T.M. 2000. The -helix translates structure into function. Protein Sci.

    9:201206.Werner, J.H., Dyer, R.B., Fesinmeyer, R.M., and Andersen, N.H. 2002. Dy-

    namics of the primary processes of protein folding: Helix nucleation. J.Phys. Chem. B 106:487494.

    Williams, S., Causgrove, T.P., Gilmanshin, R., Fang, K.S., Callender, R.H.,Woodruff, W.H., and Dyer, R.B. 1996. Fast events in protein folding: Helixmelting and formation in a small peptide. Biochemistry 35: 691697.

    Wong, K.B. and Daggett, V. 1998. Barstar has a highly dynamic hydrophobiccore: Evidence from molecular dynamics simulation and NMR relaxationdata. Biochemistry 37: 1118211192.

    Wuthrich, K. 1986. NMR of protein and nucleic acids. Wiley, New York.Zhang L. and Hermans, J. 1994. 3(10)-Helix versus -helixA molecular

    dynamics study of conformational preferences of AIB and alanine. J. Am.Chem. Soc. 116:1191511921.

    Zimm B.H. and Bragg, J.K 1959. Theory of the phase transition between helix

    and random coil in polypeptide chains. J. Chem. Phys. 31: 526535.

    -, 310-, and -helixcoil transitions

    www.proteinscience.org 1157