12
Conformational investigation of designed short linear peptides able to fold into b-hairpin structures in aqueous solution Eva de Alba, M Angeles Jiménez, Manuel Rico and José L Nieto Background: Formation of secondary structure plays an important role in the early stages of protein folding. The conformational analysis of designed peptides has proved to be very useful for identifying the interactions responsible for the formation and stability of a-helices. However, very little is known about the factors leading to the formation of b-hairpins. In order to get a good b-hairpin- forming model peptide, two peptides were designed on the basis of b-sheet propensities and individual statistical probabilities in the turn sites, together with solubility criteria. The conformational properties of the two peptides were analyzed by two-dimensional NMR methods. Results: Long-range cross-correlations observed in NOE and ROE spectra, together with other NMR evidence, show that peptide IYSNPDG TWT forms a highly populated b-hairpin in aqueous solution with a type I b-turn plus a G1 b- bulge conformation in the chain-bend region. The analogous peptide with a Pro5 substituted by Ser forms, in addition to the previous conformation, a second b- hairpin with a standard type I b-turn conformation, and the two forms are in fast dynamic equilibrium with one another. The effect of pH demonstrates the existence of a stabilizing interaction between the Asn and Asp sidechains. The populations of b-hairpin conformations increase in the presence of trifluoroethanol (a structure-enhancing solvent). On the other hand, some residual structure persists at a high denaturant concentration (8 M urea). Conclusions: This work highlights the importance of the b-turn residue composition in determining the particular type of b-hairpin adopted by a peptide, though a role of interstrand sidechain interactions in the stabilization of the formed b-hairpin is not discarded. The fact that trifluoroethanol can stabilize a- helices or b-hairpins depending on the intrinsic properties of the peptide sequence is again shown. An additional example of the presence of residual structure under denaturing conditions is also presented. Introduction There exists a large body of evidence indicating that sec- ondary structure formation plays an important role in the early steps of protein folding [1–3]. As a result, the identi- fication of the factors governing the formation of these structural elements is of crucial importance in understand- ing the folding reaction. Because of the absence of tertiary interactions, the conformational study of designed pep- tides has become a convenient tool for identifying specific sidechain interactions involved in secondary structure for- mation. In fact, designed peptides have been shown to be remarkably useful for the study of short-range interactions involved in the formation and stability of a-helices [4,5]. In contrast, information on the interactions affecting b- sheet or b-hairpin formation is scarce, possibly because previous attempts to obtain monomeric, water-soluble, b- structure-forming peptides were hampered by their ten- dency to aggregate [2]. There exists an obvious interest in the conformational study of b-hairpin-forming peptides, as little is known about the folding of all-b proteins [6]. All these reasons stimulated us to search for model peptides that could fold into highly populated b-hairpin conforma- tions, and to identify factors responsible for b-hairpin sta- bility, by using amino acid substitutions or by modifying medium conditions. b-sheet propensities have been determined in model pro- teins [7–11], but monomeric b-sheet formation has been observed with only five linear peptides so far [12–16]. Even among them, only three fold into a b-hairpin confor- mation in aqueous solution: a protein fragment of protein G [14] and two partially designed peptides, one deriving from the N-terminal sequence of ubiquitin (published during the reviewing process of this paper [15]), and another with the sequence YQNPDG SQA (peptide 1) [16]. In peptide 1, the strand residues correspond to residues 15–23 of tendamistat, where they form a b- hairpin with a type I b-turn [17]. The underlined Address: Instituto de Estructura de la Materia, Consejo Superior de Investigaciones Científicas, Serrano 119, 28006 Madrid, Spain. Correspondence to: Manuel Rico E-mail address: [email protected] Key words: b-hairpin conformation, NMR, peptide design, peptide structure Received: 22 Nov 1995 Revisions requested: 20 Dec 1995 Revisions received: 28 Dec 1995 Accepted: 15 Jan 1996 Published: 26 Feb 1996 Electronic identifier: 1359-0278-001-00133 Folding & Design 26 Feb 1996, 1:133–144 © Current Biology Ltd ISSN 1359-0278 Research Paper 133

Conformational investigation of designed short linear ... · In addition, the motif Thr8-X-Thr10, a possible b-strand-stabilizing motif — as suggested by its occurrence in two of

  • Upload
    ngokiet

  • View
    221

  • Download
    1

Embed Size (px)

Citation preview

Conformational investigation of designed short linear peptidesable to fold into b-hairpin structures in aqueous solutionEva de Alba, M Angeles Jiménez, Manuel Rico and José L Nieto

Background: Formation of secondary structure plays an important role in theearly stages of protein folding. The conformational analysis of designed peptideshas proved to be very useful for identifying the interactions responsible for theformation and stability of a-helices. However, very little is known about thefactors leading to the formation of b-hairpins. In order to get a good b-hairpin-forming model peptide, two peptides were designed on the basis of b-sheetpropensities and individual statistical probabilities in the turn sites, together withsolubility criteria. The conformational properties of the two peptides wereanalyzed by two-dimensional NMR methods.

Results: Long-range cross-correlations observed in NOE and ROE spectra,together with other NMR evidence, show that peptide IYSNPDGTWT forms ahighly populated b-hairpin in aqueous solution with a type I b-turn plus a G1 b-bulge conformation in the chain-bend region. The analogous peptide with a Pro5substituted by Ser forms, in addition to the previous conformation, a second b-hairpin with a standard type I b-turn conformation, and the two forms are in fastdynamic equilibrium with one another. The effect of pH demonstrates theexistence of a stabilizing interaction between the Asn and Asp sidechains. Thepopulations of b-hairpin conformations increase in the presence oftrifluoroethanol (a structure-enhancing solvent). On the other hand, someresidual structure persists at a high denaturant concentration (8 M urea).

Conclusions: This work highlights the importance of the b-turn residuecomposition in determining the particular type of b-hairpin adopted by a peptide,though a role of interstrand sidechain interactions in the stabilization of theformed b-hairpin is not discarded. The fact that trifluoroethanol can stabilize a-helices or b-hairpins depending on the intrinsic properties of the peptidesequence is again shown. An additional example of the presence of residualstructure under denaturing conditions is also presented.

IntroductionThere exists a large body of evidence indicating that sec-ondary structure formation plays an important role in theearly steps of protein folding [1–3]. As a result, the identi-fication of the factors governing the formation of thesestructural elements is of crucial importance in understand-ing the folding reaction. Because of the absence of tertiaryinteractions, the conformational study of designed pep-tides has become a convenient tool for identifying specificsidechain interactions involved in secondary structure for-mation. In fact, designed peptides have been shown to beremarkably useful for the study of short-range interactionsinvolved in the formation and stability of a-helices [4,5].In contrast, information on the interactions affecting b-sheet or b-hairpin formation is scarce, possibly becauseprevious attempts to obtain monomeric, water-soluble, b-structure-forming peptides were hampered by their ten-dency to aggregate [2]. There exists an obvious interest inthe conformational study of b-hairpin-forming peptides, as

little is known about the folding of all-b proteins [6]. Allthese reasons stimulated us to search for model peptidesthat could fold into highly populated b-hairpin conforma-tions, and to identify factors responsible for b-hairpin sta-bility, by using amino acid substitutions or by modifyingmedium conditions.

b-sheet propensities have been determined in model pro-teins [7–11], but monomeric b-sheet formation has beenobserved with only five linear peptides so far [12–16].Even among them, only three fold into a b-hairpin confor-mation in aqueous solution: a protein fragment of proteinG [14] and two partially designed peptides, one derivingfrom the N-terminal sequence of ubiquitin (publishedduring the reviewing process of this paper [15]), andanother with the sequence YQNPDGSQA (peptide 1)[16]. In peptide 1, the strand residues correspond toresidues 15–23 of tendamistat, where they form a b-hairpin with a type I b-turn [17]. The underlined

Address: Instituto de Estructura de la Materia,Consejo Superior de Investigaciones Científicas,Serrano 119, 28006 Madrid, Spain.

Correspondence to: Manuel RicoE-mail address: [email protected]

Key words: b-hairpin conformation, NMR, peptide design, peptide structure

Received: 22 Nov 1995Revisions requested: 20 Dec 1995Revisions received: 28 Dec 1995Accepted: 15 Jan 1996

Published: 26 Feb 1996Electronic identifier: 1359-0278-001-00133

Folding & Design 26 Feb 1996, 1:133–144

© Current Biology Ltd ISSN 1359-0278

Research Paper 133

sequence was designed to have maximum individual siteprobability to form a type I b-turn [18,19]. To improve b-hairpin formation, peptide 2—with the sequence IYS-NPDGTWT—was designed to incorporate residues withhigher b-sheet propensities than those present in peptide1. Also, to get rid of the cis/trans equilibrium arising fromthe peptide bond previous to proline, we examinedpeptide 3, with the sequence IYSNSDGTWT (see Fig. 1).

On the basis of NMR data, NOE pattern and proton chem-ical shifts, it is shown in this paper that the two designedpeptides fold into b-hairpin structures in aqueous solution.Differences in the turn residue composition have directconsequences not only on the population of b-hairpin, butalso on the nature of the b-hairpin formed. Interactionsresponsible for the changes in b-hairpin population withpH have been analyzed and the effects of denaturants (8 Murea) and structure-stabilizing solvents (trifluoroethanol

[TFE]) on the b-hairpin population have been probed.New information has thus been gained on the factors gov-erning the formation of isolated b-hairpins.

ResultsPeptide designAs in peptide 1, the chain-bend of peptide 2 hasmaximum individual site probability to form a type I b-turn [18,19]. The flanking amino acids were chosenamong those with high propensities to adopt b-strand con-formations [7–11]. In addition, the motif Thr8-X-Thr10, apossible b-strand-stabilizing motif — as suggested by itsoccurrence in two of the four known b-hairpin-formingpeptides [12,14] — was also introduced to enhance thepopulation of b-hairpin conformation. Thehydrophilic/hydrophobic pattern characteristic of b-sheetstogether with solubility criteria were also taken intoaccount. As amino acids with high propensities to adopt b-strand conformations are usually very hydrophobic, solu-bility criteria are crucial in order to get a water-solublenon-aggregating peptide.

Peptide 2 exists in two conformational states as a result ofthe cis/trans X-Pro isomerization with the trans X-Pro pop-ulation predominating. In peptide 3, the Pro residue wassubstituted by Ser, which has the highest probability afterPro to be in position i+1 of the type I b-turn [18]. Thissingle residue substitution simplifies the spectrum andallows an investigation of the role of the turn rigidity inthe formation of the b-hairpin.

Secondary structure predictionsThe lowest energy backbone conformations predicted forthe designed peptides using the program Prelude [20] aregiven in Table 1. For peptide 2, a marked preference for b-strand or extended conformation is predicted for residues1–5 and 8–9, and turn-like conformations for residues 6–7.For peptide 3, residues 1–3 and 8–9 adopt b-strand confor-mations and residues 4–7 adopt turn-like states.

Aggregation stateOne-dimensional 1H NMR spectra were recorded inaqueous solution at peptide concentrations of 0.1 mM and15 mM in the case of peptide 2, and at 0.15 mM and15 mM in the case of peptide 3. The absence of anyappreciable change in either linewidths or chemical shiftssuggests the absence of aggregation. Nevertheless, inorder to rule out the possibility of aggregates that nolonger change their NMR properties in the concentrationrange of 0.1–15 mM, fluorescence spectra were recordedfor the single Trp present in peptides 2 and 3 at lowerconcentrations than those used in NMR experiments (i.e.10–100 mM). The intensity of the Trp fluorescence emis-sion at 350 nm (Trp excited at 280 nm) varied linearlywith concentration up to 80 mM. A slight curvature wasobserved for peptide concentrations between 80 and

134 Folding & Design Vol 1 No 2

Figure 1

Schematic representation of the peptide backbone conformationcorresponding to (a) the experimentally found b-hairpin (b-hairpin 3:5)for peptides 1, 2, and 3, and (b) the additional b-hairpin found inpeptide 3 (b-hairpin 4:4). The Ha–Ha(i, j) and the NN(i, j) NOEcrosspeaks experimentally observed for each b-hairpin are shown bythin black arrows and the hydrogen bonds expected for each b-hairpinby open arrows. The sequence of peptide 2 is given in b-hairpin 3:5 (a)and that of peptide 3 in b-hairpin 4:4 (b).

H3N

H

O

N

H

O

N

H

N

H

O

O

O

H

H

H H

N

O

H

O

N

O

H

H

N

H

H

H

N O

N

H

H

O

NO

H

H3N

H

O

N

H

O

N

H

O

N

H

N

H

O

N

H

O

N

H

O

O

O

NO

H

H

H

H

H

H

H

O HN

ON

+

T 10

N 4S 5

T 8_

Y 2

I 1 S 3

W 9 G 7

D 6

D 6

+ I 1

N 4

P 5

Y 2

T 8_

W 9

S 3

G 7

T 10

(a)

(b)

100 mM due to the inner filter effect (data not shown). Asanother test to rule out aggregation, circular dichroism(CD) spectra were acquired at peptide concentrations of50 and 100 mM. Identical CD spectra were obtained atboth concentrations, consistent with absence of aggrega-tion in this concentration range (data not shown). Oncethe absence of aggregation over a very wide range ofpeptide concentrations (10 mM to 15 mM) was confirmed,concentrations of peptide between 5 mM and 15 mMwere used in all the NMR experiments reported here.

Spectral assignmentThe 1H NMR spectra of peptides 2 and 3 in aqueous solu-tion at various pH values were completely assigned usingthe standard sequential assignment procedure [21,22], aswere the assignments of the 1H NMR spectra of peptides2 and 3 under denaturing conditions (6–8 M urea) and in

the presence of 30% TFE. The 1H chemical shifts of pep-tides 2 and 3 in aqueous solution at pH 4.3 and 5°C arelisted in Tables 2 and 3, respectively.

Conformational behavior in aqueous solution.The detection and identification of the folded conforma-tions adopted by peptides 2 and 3 were based on thestructural information provided by NOE data, Ha confor-mational shifts (deviation of the chemical shift values withrespect to those in random coil [RC] peptides) and NHshift-temperature coefficients. NOE data are the strongestevidence for confirming the existence of a folded struc-ture. In order to avoid spin diffusion problems, shortmixing times were used in the NOESY spectra. Further-more, the same pattern of NOE crosspeaks was found inthe ROESY spectra recorded under identical experimental

Research Paper Folding of designed peptides into b-hairpin structures de Alba et al. 135

Table 1

Predicted conformational states for peptides 2 and 3*.

Peptide 2 Peptide 3

Sequence Conformation number Sequence Conformation number

1 2 3 1 2 3

Ile1 B B B Ile1 B B BTyr2 B B B Tyr2 B B BSer3 P B P Ser3 B B CAsn4 B B B Asn4 G G GPro5 P P P Ser5 C C CAsp6 C C C Asp6 C C CGly7 G G G Gly7 G G GThr8 B B C Thr8 B B BTrp9 B B P Trp9 B P BThr10 X X X Thr10 X X X

Predicted energy(kcal mol–1) –5.53 –5.47 –5.46 –5.88 –5.78 –5.78

rms deviation† (Å) 0.82 1.25 0.60 2.38

*Predictions were made according to the method of Rooman et al.[20]. †rms deviations are calculated with respect to the lowest energyconformation predicted. B, b-strand; C, 310 helix; G, left-handed helix;

P, extended; X, not computed; c and f, torsional anglescorresponding to these conformational states are defined in Rooman et al. [20].

Table 2

Chemical shifts (ppm from TSP) of peptide 2 in aqueous solution at 5°C and pH 4.3.

NH CaH CbH Other

Ile1 3.92 1.96 CgH 1.48, 1.19; CgH3 0.98; CdH3 0.86Tyr2 8.86 4.70 3.00, 3.00 CdH 7.11; CeH 6.83Ser3 8.20 4.30 3.56, 3.52Asn4 8.53 4.87 2.84, 2.68 NdH2 7.61, 7.09Pro5 4.38 2.32, 1.98 CgH 2.04, 2.04; CdH 3.86, 3.79Asp6 8.24 4.61 2.87, 2.76Gly7 8.22 4.04, 3.84Thr8 7.97 4.36 4.20 CgH3 1.13Trp9 8.46 4.94 3.35, 3.20 Ne1H 10.14; Cd1H 7.24; Ce3H 7.58; Cz3H 7.12; Ch2H 7.23; Cz2H 7.49Thr10 8.08 4.22 4.18 CgH3 1.16

conditions, confirming the absence of spin diffusion.ROESY spectra are less sensitive to spin diffusion effects,and spin diffusion mediated crosspeaks can be distin-guished easily from true ROE signals by their oppositesign. The NOE patterns obtained for peptides 2 and 3from the combined analysis of ROESY and NOESYspectra are summarized in Figure 2. Selected regions ofNOESY and ROESY spectra of both peptides showing

some long-range, sequential, and intraresidue NOE cross-peaks are given in Figures 3 and 4.

Evidence for the presence of a chain-bend in both pep-tides comes from three points. First, there is the Ha–NH(i, i+2) NOE between residues Pro5 and Gly7 (Fig. 2) inpeptide 2, i.e. the non-sequential NOE expected to occurbetween residues i+1 and i+3 of a b-turn [22] (the

136 Folding & Design Vol 1 No 2

Table 3

Chemical shifts (ppm from TSP) of peptide 3 in aqueous solution at 5°C and pH 4.3.

NH CaH CbH Other

Ile1 3.88 1.94 CgH 1.45, 1.16; CgH3 0.96; CdH3 0.87Tyr2 8.83 4.68 2.97, 2.88 CdH 7.04; CeH 6.80Ser3 8.33 4.39 3.68, 3.62Asn4 8.58 4.70 2.84, 2.78 NdH2 7.62, 7.04Ser5 8.60 4.36 3.92, 3.85Asp6 8.23 4.66 2.86, 2.76Gly7 8.24 4.01, 3.84Thr8 7.93 4.38 4.17 CgH3 1.12Trp9 8.49 4.91 3.34, 3.17 Ne1H 10.13; Cd1H 7.22; Ce3H 7.52 Cz3H 7.10; Ch2H 7.21; Cz2H 7.46Thr10 8.04 4.20 4.17 CgH3 1.14

Figure 2

Ile1 Tyr2 Ser3Asn4Pro5 Asp6Gly7 Thr8 Trp9 Thr10

dαN

dNN

dαN(i,i+2)

d(i,i+2)

d(i,i+3)

d(i,i+4)

d(i.i+5)

d(i,i+6)

d(i,i+7)

d(i,i+8)

d(i,i+9)

Ile1 Tyr2 Ser3Asn4Ser5 Asp6Gly7 Thr8 Trp9 Thr10

dαN

dNN

dαN(i,i+2)

d(i,i+2)

d(i,i+3)

d(i,i+4)

d(i.i+5)

d(i,i+6)

d(i,i+7)

d(i,i+8)

d(i,i+9)

** * ** *

(a) (b)

Summary of NOE connectivities observed for (a) peptide 2 and (b)peptide 3 at 5°C and pH 4.3. The thickness of the lines reflects theintensity of the sequential NOE connectivities, i.e. weak, medium and

strong. An asterisk or a broken line indicates unobserved NOEconnectivity due to signal overlapping, closeness to diagonal oroverlap with solvent signal.

corresponding NOE between residues Ser5 and Gly7 inpeptide 3 is observed only at pH 6.3 due to spectral over-

lapping present at lower pH values). Second, the verysmall temperature coefficients of the amide proton

Research Paper Folding of designed peptides into b-hairpin structures de Alba et al. 137

Figure 3

Selected regions of ROESY and NOESYspectra of peptide 2. (a) ROESY regionshowing the long-range NOE crosspeaksconnecting the sidechains of residues Ile1and Trp9 (200 ms mixing time). (b) ROESYregion showing the HaSer3–HaTrp9 NOEcrosspeak (100 ms mixing time). Conditions:5 mM peptide in D2O, 2°C, pH 4.3. (c) NNregion of a NOESY spectrum, in which theNHTyr2–NHThr10 NOE is boxed. Conditions:15 mM peptide in H2O/D2O (9:1 by volume),5°C, pH 4.3, 200 ms mixing time.

Figure 4

ROESY spectrum of peptide 3. (a) Fingerprintregion. The asterisks indicate a small impurity.Conditions: 15 mM peptide in H2O/D2O (9:1by volume), 5°C, pH 4.3, 200 ms mixing time.(b) Spectral region showing Ha–Ha NOEcrosspeaks. Conditions: 5 mM peptide inD2O, 2°C, pH 6.3.

resonances of residues Asp6 and Gly7 (Table 4) indicateprotection against solvent exchange by their possibleinvolvement in intramolecular hydrogen bonds or solventinaccessibility [23]. Third, Ha chemical shifts, whose devi-ations from random coil values are diagnostic of secondarystructure. Ha conformational shifts, Dd = d – dRC ppm, arenegative for helical or turn regions and positive in b-con-formations or extended conformations [24,25]. Somecaution must be exercised in the application of the Hachemical shift criterion because of ring current effects ofthe aromatic residues Tyr2 and Trp9 on the chemicalshifts. The Ha conformational shifts of residues Asn4 toGly7 are negative, in agreement with the presence of turnstructures in this region, whereas some of the Ha confor-mational shifts of the residues flanking the turn are posi-tive, as expected for b-strand conformations (Figs 5,6).

The presence of a b-hairpin structure of the type shownschematically in Figure 1a (a b-hairpin 3:5 according tothe b-hairpin classification by Thornton and colleagues[26,27]) in peptide 2 in aqueous solution is strongly sup-ported by the pattern of long-range NOE crosspeaksobserved, in particular the NOE crosspeak connectingHaSer3 with HaTrp9, and the one between NHTyr2 andNHThr10 (Fig. 3). The presence of these two crosspeakscan be explained only if these residues (Ser3–Trp9 andTyr2–Thr10) face each other in the b-hairpin conforma-tion (as shown in Fig. 1a). Additional NOE cross-correla-tions observed between residues from different strandsand on the same side of the b-hairpin, including thosebetween sidechain protons of Ile1 and Trp9 (Fig. 3), arealso consistent with the formation of the b-hairpin 3:5shown in Figure 1a. The positive Ha conformational shiftsobserved for most of the residues in the b-strands, as indi-cated above, are also in agreement with the formation of

b-hairpin 3:5. This type of b-hairpin was also the oneformed by peptide 1. A very weak HaTyr2–HaTrp9 NOEconnectivity was also detected in peptide 2. This NOEconnectivity is incompatible with the b-hairpin 3:5 andcould be interpreted as resulting from the presence of avery low population of b-hairpin 4:4 (estimated to beabout 5%, whereas that of b-hairpin 3:5 is almost 30%), asdiscussed below. During the reviewing process of thispaper, a similar behavior was reported for a peptide deriv-ing from the N-terminal sequence of ubiquitin and withthe same turn residues, NPDG [15].

In peptide 3, the HaSer3–HaTrp9 and theNHTyr2–NHThr10 NOE cross-correlations characteristicof b-hairpin 3:5, together with most of the long-range NOEcrosspeaks involving sidechain protons observed in peptide2, are also present, which indicates that peptide 3 adopts ab-hairpin 3:5 as well. In addition, a NOE crosspeak con-necting HaTyr2 with HaTrp9 (stronger than in peptide 2)and another one between NHSer3 and NHThr8 werefound in the NOESY and ROESY spectra of peptide 3(Fig. 4). These two NOE connectivities are incompatiblewith b-hairpin 3:5. One plausible explanation for theseNOE connectivities is the existence of another b-hairpinstructure, b-hairpin 4:4 (Fig. 1b) [26,27], in dynamic con-formational equilibrium with the previous one.

The CD spectra of peptides 2 and 3 show a minimum atabout 201 nm, slightly shifted from the one below 200 nmexpected for random coil peptides, and a maximum atabout 230 nm, usually observed in peptides containing aro-matic residues in their sequence [28,29]. This fits with thepresence of some non-random structure in the peptides.Identification of the nature of such non-random structureby CD is not possible due to the shortness of the two b-

138 Folding & Design Vol 1 No 2

Table 4

Amide proton temperature coefficients (–Dd/DT, ppb.K–1) obtained for peptides 2 and 3 under various experimental conditions.

Peptide 2 Peptide 3

H2O 6M urea 8M urea 30% TFE H2O 8M urea 30% TFE

pH

Residue 3.0 4.3 5.3 5.3 5.3 4.3 3.0 4.3 5.3 5.3 4.3

Ile1Tyr2 8.3 8.7 12.7 9.0 8.3 9.7 7.7 8.7 9.0 8.7 9.3Ser3 3.7 4.3 4.7 6.7 6.7 5.0 4.7 5.0 4.7 6.7 5.3Asn4 7.3 7.6 7.3 8.0 7.7 7.7 6.3 7.0 7.3 7.7 7.0Xaa5* 6.7 9.0 9.0 8.7 8.7Asp6 5.6 3.2 2.3 3.7 3.0 1.3 4.0 1.0 0.3 3.0 0.7Gly7 4.7 4.7 4.3 5.0 4.7 3.7 4.3 3.0 2.7 4.0 2.0Thr8 5.0 4.2 4.0 4.7 4.7 2.3 4.0 3.3 3.0 4.7 2.3Trp9 8.1 8.7 9.0 8.3 8.0 8.7 8.3 10.0 10.0 8.0 9.7Thr10 6.3 9.0 9.7 9.0 9.0 12.0 6.7 8.7 9.0 8.3 10.3

*Xaa is Pro in peptide 2 and Ser in peptide 3.

strands (only three residues) and the variety of CD spectrafound experimentally for the different types of b-turns[30,31]. CD is not a suitable technique for identifyingsmall populations of b-turn and b-hairpin conformations.

Dependence of b-hairpin population on pHThe pH dependence of the NMR parameters observedfor peptides 2 and 3 was investigated. At low pH (3.0), rel-

ative to higher pH values (4.3 and 5.3), the medium-rangeand long-range NOE crosspeaks observed for peptide 2are smaller in number and in intensity, the absolute valuesof Ha conformational shifts decrease both for turn-like andb-strand residues (Fig. 5), and the absolute values of theamide proton temperature coefficients for residues Asp6and Gly7 increase (Table 4). The Ha conformational shiftsand amide proton temperature coefficients for residues

Research Paper Folding of designed peptides into b-hairpin structures de Alba et al. 139

Figure 5

0.25

0.20

0.15

0.10

0.05

0.00

-0.05

-0.10

-0.15

δ =

δ -

δrc,

ppm

0.25

0.20

0.15

0.10

0.05

0.00

-0.05

-0.10

-0.15

δ =

δ -

δrc,

ppm

(a)

I Y S N P D G T W T

(b)

I Y S N P D G T W T

(c)

I Y S N P D G T W T

(d)

I Y S N P D G T W T

(e)

I Y S N P D G T W T

Figure 6

Ha conformational shifts of peptide 3 at 5°Cand pH 4.3. The Ha conformational shiftsmeasured for peptide 3 in (a) aqueoussolution and (b) 30% TFE solution.

0.25

0.20

0.15

0.10

0.05

0.00

-0.05

-0.10

-0.15

δ =

δ -

δrc,

ppm

(a)

I Y S N P D G T W T

(b)

I Y S N P D G T W T

Ha conformational shifts of peptide 2 at5°C. The Ha conformational shifts wereobtained with reference to 6 M urea,denaturing conditions (Dd = d – dRC ppm),in aqueous solution at (a) pH 3.0,(b) pH 4.3 and (c) pH 5.3, and (d) in 30%TFE at pH 4.3. (e) The Ha conformationalshifts obtained for peptide 2 and pH 5.3with reference to 8 M urea.

Asp6 and Gly7 of peptide 3 (Table 4) present a similar pHdependence.

Conformational behavior under denaturing conditionsChemical shifts of Ha protons of amino acid X intetrapeptides Gly-Gly-X-Ala are frequently used as refer-ence for the disordered state [32]. The Ha chemical shiftsfor most peptides show small variations from theserandom coil values. To avoid sequence effects on randomcoil d values, the Ha conformational shifts of peptides 2and 3 were obtained, using as reference for the coil stateHa chemical shifts obtained for the same peptide underdenaturating conditions (6 M urea). This way of evaluat-ing conformational shifts allows the detection of smallpopulations of structure in partially folded peptides[33,34].

However, the NOESY and ROESY spectra of peptides 2and 3 in 6 M urea, performed in order to get the d valuesunder denaturing conditions, contain medium-range andlong-range NOE connectivities indicative of structure,though less numerous and weaker than in aqueous solu-tion at the same pH value. The long-range NOE connec-tivities include those involving sidechain protons of Ile1and Trp9. Furthermore, the absolute values of the amideproton temperature coefficient of Asp6 in both peptides 2and 3 are small (Table 4).

These findings indicate that some residual folded struc-ture persists in peptides 2 and 3 under these denaturingconditions and also in higher concentrations of urea (7 Mand 8 M; Table 4). The Ha conformational shifts with ref-erence to 6 M urea and to 8 M urea at pH 5.3 and 5°C areplotted as a function of the peptide sequence in Figure 5for peptide 2. The profiles of the Ha conformational shiftsobtained for peptide 2 (Fig. 5) and peptide 3 (data notshown) with respect to 6 M urea and with respect to 8 Murea are identical, except for the absolute values of the Haconformational shifts, being slightly larger when obtainedfrom 8 M urea, consistent with the presence of a smallerbut still persistent folded structure at the higher denatu-rant concentration.

Conformational behavior in 30% trifluoroethanol solutionTFE is known to increase the population of secondarystructure formed by protein fragments [35–37]. It is gen-erally accepted that TFE increases helical populations inpeptides of different sequences in a non-specific way.However, there are several cases in which the presence ofmixed solvent alcohol/water also enhances the formationof isolated b-hairpins [12,13,38]. Secondary structure sta-bilization by TFE probably occurs by increasing thestrength of intra-peptide hydrogen bonds whicheverstructure they belong to. We therefore investigated theeffect of TFE on the conformational behavior of peptides2 and 3 in the mixed solvent 30% TFE by NMR.

The absolute values of the Ha conformational shiftsobtained for peptides 2 and 3 in 30% TFE at pH 4.3 and5°C (Figs 5,6) are larger than those obtained for eachpeptide in aqueous solution at the same pH and tempera-ture. The increase in the absolute values of the Ha confor-mational shifts is particularly large in the b-strand regions.The profile of Ha conformational shifts obtained for pep-tides 2 and 3 in 30% TFE (Figs 5,6) is that expected for anideal b-hairpin. In addition, the small absolute values ofthe amide temperature coefficients measured for residuesAsp6, Gly7 and Thr8 in aqueous solution decrease to evensmaller absolute values for both peptides in 30% TFE atpH 4.3 (Table 4). The pattern of NOE connectivitiesobtained for each peptide in 30% TFE is also almost iden-tical to that previously obtained in aqueous solution at thesame pH. Both b-hairpin conformations found for peptide3 in aqueous solution are also present in 30% TFE, asindicated by the presence of the NOE connectivities char-acteristic of b-hairpin conformations 3:5 and 4:4. The onlydifference between the set of NOE connectivitiesobtained for peptides 2 and 3 in aqueous solution and in30% TFE is in the intensity of many NOE crosspeakswhich increase in 30% TFE, consistent with an increase inthe populations of b-hairpin conformations in TFE.

It can also be noted that in peptide 3, the HaSer3–HaTrp9NOE crosspeak (characteristic of b-hairpin 3:5) is slightlyweaker in aqueous solution and slightly stronger in 30%TFE than the HaTyr2–HaTrp9 NOE crosspeak (charac-teristic of b-hairpin 4:4), possibly suggesting that the sta-bilizing effect of TFE on b-hairpin 3:5 differs somewhatfrom that on b-hairpin 4:4. In conclusion, TFE increasesthe population of b-hairpin without changing the nature ofb-hairpin structure.

Estimation of b-hairpin populationIn contrast to a-helices, no methods have yet been estab-lished for quantifying either b-turn or b-hairpin conforma-tions adopted by peptides in solution. To estimate thepopulation of the b-hairpins formed by the model peptides2 and 3, we use the ratio of the intensities of the Ha–HaNOE characteristic of each b-hairpin (HaSer3–HaTrp9 forb-hairpin 3:5 and HaTyr2–HaTrp9 for b-hairpin 4:4) tothat of the Ha–Ha′Gly7 NOE (IntensityHa–Haobserved /IntensityHa–Ha′Gly7), because being the Ha–Ha′Gly7 dis-tance constant, the Ha–Ha′Gly7 NOE intensity can beused as reference. The intensity ratio corresponding to a100% b-hairpin was taken as [dHai–Haj]–6 / [dHa–Ha′Gly7]–6,where dHai–Haj is the averaged Hai–Haj distance valuefound for antiparallel b-sheets in proteins [22], being i andj facing residues with the Ha protons pointing towards theinterior in an antiparallel b-sheet, and dHa–Ha′Gly7 is theHa–Ha′ distance characteristic of Gly residues. It is clearthat this approach is approximate, because it implicitlyassumes equal correlation times for the peptide in any con-formational state and small variations in the actual Ha–Ha

140 Folding & Design Vol 1 No 2

distance with respect to the averaged value in antiparallelb-sheets of proteins can lead to large over- or under-esti-mations of the b-hairpin relative population. Nevertheless,given the absence of any other available method, we arebound to use this method to estimate the b-hairpin popula-tions formed by peptides 2 and 3.

This estimation yields a population for b-hairpin 3:5 ofabout 27% for peptide 2 in aqueous solution at pH 6.3 and2°C, a conformation which would be interconverting in afast dynamic equilibrium with about 5% of b-hairpin 4:4and a number of other conformations, including those cor-responding to the ‘random coil’. In peptide 3 under thesame experimental conditions, the conformational equilib-rium would be formed by about 27% of b-hairpin 3:5, 32%of b-hairpin 4:4, and other disordered forms. Among these,the presence of molecules containing the b-turn but notthe b-strands is highly probable.

Peptide three-dimensional structures in aqueous solutionDue to the usual conformational averaging that occurs inpeptides, NOE intensities cannot be rigorously inter-preted in terms of a unique structure. Nevertheless, it isuseful to calculate a limited number of structures compati-ble with NOE constraints, which helps to visualize theconformational properties of the ensemble, albeit in a verysimplified way.

Structure calculations were performed on the basis of thecomplete set of NOE restraints obtained for peptide 2 inaqueous solution at pH 4.3, and using a distance geometryprocedure [39]. A table, listing the NOE constraints usedfor the structure calculation, is available as Supplementarymaterial (published with this paper on the internet). fangles were also constrained to the range 0 to –180°,except for Gly. Starting from 250 randomized conforma-tions, 31 conformers which satisfy the constraints with noviolation greater than 0.13 Å were selected. The values ofthe pairwise root mean square deviations (RMSDs) for thebackbone atoms of these 31 conformers are 0.7 ± 0.2 Å.The superposition of six of these calculated structures forpeptide 2 are shown in Figure 7. The set of structures cal-culated without any angle constraint was very similar tothose shown in Figure 7, but less well defined (RMSD forthe backbone atoms 1.4 ± 0.5 Å).

The calculated structures for peptide 2 on the basis of theexperimental set of NOE restraints coincide with the b-hairpin 3:5 schematically represented in Figure 1a. Manyof the calculated structures present a hydrogen bondbetween the CO group of Asn4 and the NH of Gly7, and ahydrogen bond between the NH of Asn4 and the COgroup of Thr8. These data, together with the range ofvalues obtained for the angles f and c of residues Asn4,Pro5, Asp6, Gly7, and Thr8, indicate that the conforma-tion of the chain-bend region is not a standard type I b-

turn but a type I b-turn plus a G1 b-bulge [26,27], whichgives the pattern of alternating residues in the b-strandsrepresented in Figure 1a and observed in Figure 7. The b-hairpin formed by peptide 2 can be classified as a b-hairpin 3:5 [26,27].

A similar Distance Geometry calculation was performed forpeptide 3 in order to check the compatibility of all the NOEconnectivities observed with a unique conformation. Thecalculation performed using a set of NOE data including allthe NOE restraints observed leads to large NOE constraintviolations. This result evidences the presence of someincompatible distance constraints within the structures cor-responding to b-hairpin 3:5 (Fig. 1a). As expected, the NOErestraints with large violations include HaTyr2–HaTrp9and NHSer3–NHThr8 NOE connectivities.

Research Paper Folding of designed peptides into b-hairpin structures de Alba et al. 141

Figure 7

Stereoscopic view of the superposition of six calculated structures forpeptide 2. (a) Showing sidechain and backbone atoms. (b) Showingonly backbone atoms.

In order to calculate the additional conformation observedin peptide 3, b-hairpin 4:4 (Fig. 1b), a set of NOE dataexcluding all those NOE restraints present in b-hairpin3:5 were used. The calculated structures were not welldefined due to the smallness of the NOE data set. Never-theless, the range of f and c angles of the chain-bendregion coincide with the ones usually observed in a b-hairpin with a four-residue chain-bend, that is a standardtype I b-turn [26,27]. It is likely that the conformationalfreedom gained by the substitution of Pro5 by Ser allowsthe chain-bend to adopt two different conformations: atype I b-turn plus G1 b-bulge resulting in a five-residueloop b-hairpin (b-hairpin 3:5) and a standard type I b-turnresulting in a four-residue loop b-hairpin (b-hairpin 4:4).

DiscussionThe b-hairpin-forming propensities shown by peptide 1,reported previously [16], are conserved or even improvedby the two designed peptides analyzed in this work, indi-cating the usefulness of statistical data on site probabilitiesand intrinsic propensities determined in proteins indesigning b-strand-forming peptides. A next step wouldbe to use the information about pairwise sidechain interac-tions, recently determined by statistical methods for anti-parallel b-sheets [40], to experimentally confirm theexistence of these and other sidechain interactions respon-sible for the b-hairpin stability. Peptide 2 has many of thecharacteristics to serve as a convenient model peptide forthe study of these stabilizing interactions. It is watersoluble, it does not aggregate and forms a b-hairpin (type3:5, Fig. 1a) with a relatively high population (~30 %). Sta-bilizing interactions could be identified by analyzing thechanges in relative b-hairpin populations resulting fromspecific amino acid substitutions.

An example of the effects of amino acid substitution isseen in the different conformational properties of peptides2 and 3. Although the main purpose of substituting Pro bySer at position 5 was to remove the set of signals arisingfrom the Asn–Pro cis isomer, the introduction of a Serresidue in the turn composition also revealed the exis-tence of a different significantly populated b-hairpin 4:4,while conserving that of b-hairpin 3:5. The b-hairpin 4:4has a four-residue turn and an inside-out inversion of thefirst b-strand in contrast to b-hairpin 3:5 which has a five-residue turn (Fig. 1). A possible explanation for the differ-ences found between the conformational properties ofpeptides 2 and 3 may lie in the greater flexibility of theSer peptide in the turn region. As the b-hairpin 4:4 confor-mation also appears to be present in peptide 2, but popu-lated to a very small degree, the two conformationsprobably do not differ energetically by a significantamount. The fact that a single-residue substitution in theb-turn region leads to the appearance of a new b-hairpinconformation highlights the importance of the turn regionin defining the final b-hairpin conformation.

When the backbone conformations are examined usingthe program Prelude [20], residues 1–3 and 8–9 in bothpeptides predicted to be in a b-strand state (Table 1) arein a b-strand state for both b-hairpins 3:5 and 4:4. The fand c angles of residue Asn4 correspond to a b-strandstate in b-hairpin 3:5, consistent with the secondary struc-ture prediction for Asn4 in peptide 2 (Table 1). Theagreement between predicted and experimental sec-ondary structure is not so good for peptide 3, as residueAsn4 is predicted to be in a left-handed helix, but corre-sponds to a right-handed helix [26,27] in b-hairpin 4:4.Anyway, it is interesting to note that the different sec-ondary structure predictions for peptides 2 and 3 correlatewith different conformational properties.

The b-hairpin population in peptide 2 is pH dependent,decreasing at low pH values. This pH conformationaldependence has also been found in peptide 1, spanningidentical b-turn residues (NPDG), and originates in ahydrogen-bonding interaction between the sidechains ofAsn and Asp which is destabilized by the protonation ofthe Asp sidechain [33]. This same interaction can accountfor the pH conformational dependence existent in peptide2 and also in the two b-hairpin conformations adopted bypeptide 3. Thus, the interaction between the Asn and Aspsidechains within the sequence Asn-X-Asp is notrestricted to the presence of a Pro residue in the centralposition, X. Probably, such an interaction exists wheneverAsn, X and Asp residues have the necessary geometry (f,c) and proximity.

A striking result is that some residual non-random confor-mations persist in peptides 2 and 3 under normally denatur-ing conditions (8 M urea). A similar observation of residualnon-random structure in 7 M urea has been observed previ-ously for the N-terminal 63-residue domain of the 434-repressor protein and its peptide fragment (residues 44–63),where residual structure corresponded to a hydrophobiccluster [41,42]. In peptides 2 and 3, a hydrophobic interac-tion between residues Ile1 and Trp9, only partially affectedby urea, may be sufficiently strong to allow the persistenceof structure, at least in NMR-detectable populations in 8 Murea. An important consequence of these findings is thatextreme caution must be taken when using urea as a refer-ence for the fully random reference state of peptides andproteins, as indicated by Neri et al. [41].

The populations of the b-hairpin conformation formed bypeptide 2 and those of the two b-hairpins adopted bypeptide 3 increase significantly in the mixed solvent 30%TFE. The b-hairpin conformation formed by peptide 1 isalso stabilized in 30% TFE (unpublished data). Theseresults together with those found in other peptides [13,38]indicate that TFE is not only a helix-promoting solvent,but may promote b-hairpin conformations as well, at leastin short monomeric peptides. Thus, the effect of TFE on

142 Folding & Design Vol 1 No 2

the structure adopted by a determined peptide maydepend on the peptide sequence. It is interesting that therelative populations of the two b-hairpin conformationsadopted by peptide 3 appear to be slightly different inaqueous solution and in 30% TFE solution, suggestingthat TFE has different stabilizing effects for different b-hairpin conformations.

The results obtained in this work support the idea that thechain-bend composition is more important in determiningthe nature of the b-hairpin to be adopted than the strandresidues and their possible pairwise interactions. Pairwiseinteractions were not explicitly included in the design ofpeptides 2 and 3. For design purposes, it will be importantto determine the relative importance of the effect of par-ticular residues in the chain-bend and in the b-strands onthe nature of the b-hairpin formed. We are currentlyexploring the effects of both by changing the turnresidues, i.e. taking those residues to have maximal b-turnprobability in each position of a b-turn I′, instead of a b-turn I, and by optimizing interstrand pairwise interactionsfor b-hairpin 3:5 or for b-hairpin 4:4.

In conclusion, the results obtained on the basis of theNMR conformational study indicate that the compositionof the chain-bend region is important for the final confor-mation of the b-hairpin, although further investigation isrequired to clarify the influence of interstrand interactionsin b-hairpin formation. The early formation of backboneturns or bends is of primary importance in the foldingprocess because it reduces the conformational space andprovides the opportunity for tertiary interactions to occur.The use of the peptides designed here as referencemodels and the conformational analysis of designed sitemutant peptides appears to be a most promising way toobtain important information about the factors governingb-hairpin formation.

Materials and methodsMaterialsPeptides were chemically synthesized by Neosystem Laboratoire(Strasbourg, France). Urea was from Fluka Chemie, AG (Buchs,Switzerland). 2,2,2-Trifluoroethanol-d3 was purchased from CambridgeIsotopes Laboratories (Cambridge, USA).

Secondary structure predictionsThe program Prelude [20] developed to identify conformational prefer-ences of protein backbone segments was used to predict peptide sec-ondary structure.

Fluorescence dataFluorescence emission spectra were recorded in an AMINCOBOWMAN SERIES 2 Luminescence Spectrometer after excitation at280 nm on peptide samples at pH 4.3 and 4.5°C (peptide 2) and 5°C(peptide 3). Slitwidths were 4 nm.

CD measurementsCD spectra were acquired on an Aviv 60 DS Spectropolarimeter with aHewlett-Packard 89100A temperature control unit and cuvettes with

either 10-mm or 1-mm path lengths. CD spectra were the average ofthree scans obtained by collecting data at 1 nm intervals from250–186 nm. Peptide samples were at pH 4.3 and 5°C. Peptide con-centrations were determined by UV absorbance [43].

Proton NMR spectraPeptide solutions were 15 mM in 0.5 ml of H2O/D2O (9:1 ratio byvolume), except for experiments under denaturating conditions, in TFEor in pure D2O, where the peptide concentrations were around 5 mM.The pH, adjusted by adding minimal amounts of dilute DCl or NaOD,was measured with a glass microelectrode and the reported valueswere not corrected for isotope effects. Samples for experiments underdenaturating conditions were prepared by adding the appropriateamount of urea and the pH was readjusted. To 5 mM peptide samples,deuterated TFE was added to 30% by volume and the pH adjusted asrequired. The temperature of the NMR probe was calibrated using amethanol sample. Sodium 3-trimethylsilyl [2,2,3,3-2H] propionate (TSP)was used as an internal reference at 0.00 ppm. The 1H NMR spectrawere acquired on a Bruker AMX-600 pulse spectrometer operating at aproton frequency of 600.13 MHz. One-dimensional spectra wereacquired using 32 K data points, which were zero-filled to 64 K datapoints before performing the Fourier transformation. Phase-sensitivetwo-dimensional scalar coupling correlated spectroscopy (COSY)[44,45], total correlation spectroscopy (TOCSY) [45], nuclear Over-hauser effect spectroscopy (NOESY) [46,47] and rotating frameexchange spectroscopy (ROESY) [48] spectra were recorded by stan-dard techniques using presaturation of the water signal and the time-proportional phase incrementation mode. A mixing time of 200 ms wasused for NOESY and ROESY spectra (apart from NOESY experimentsin TFE which were recorded using a mixing time of 100 ms). TOCSYspectra were recorded using 80 ms MLEV 17 spin-lock sequence [49]at different temperatures to obtain amide proton temperature depen-dence. Acquisition data matrices were defined by 2048 × 512 points inthe frequency domain. Additional NOESY and ROESY spectra(100 ms mixing time) were recorded on peptide samples in pure D2Oto facilitate the observation of Ha–Ha NOE crosspeaks too close to thewater signal. Data were processed using the standard UXNMR Brukerprograms in an Aspect X32 computer. The two-dimensional data matrixwas multiplied by a phase-shifted square-sine-bell window function andzero-filled to 4 K × 2 K prior to Fourier transformation. Baseline correc-tion was applied in both dimensions.

Structure calculationsNOE intensities were evaluated in a qualitative way as strong, mediumand weak and were used to obtain upper limit distance constraints of3 Å, 4 Å and 5 Å, respectively. Pseudoatom corrections were addedwhere necessary. Structures were calculated on a Silicon GraphicsIndigo computer using the program DIANA [39]. As erroneous resultscan be obtained due to the contribution of the heterogeneous confor-mational behavior of the peptide to the intraresidue and sequentialNOE crosspeak intensities, only medium-range and long-range NOEcrosspeaks were considered to calculate the set of structures ofpeptide 2. Another set of structures was calculated for peptide 2 withthe φ angles constrained to the range 0° to –180°, except for Gly. Forpeptide 3, two calculations were performed: one with a complete set ofNOE data and the other with a set of NOE data excluding all thoseNOE constraints in the b-hairpin 3:5 (Fig. 1).

Supplementary material availableA tabulation listing the medium-range and long-range NOE connectivi-ties observed for peptides 2 and 3 in aqueous solution at pH 4.3 and5°C is published with this paper on the internet.

AcknowledgementsWe are grateful to Professor RL Baldwin and all his lab group for the use offluorescence and CD facilities at Stanford University and to Dr S Padman-abhan for his critical reading of the manuscript. This work was supported bythe Spanish DGYCT project no PB90-1020.

Research Paper Folding of designed peptides into b-hairpin structures de Alba et al. 143

References1. Kim, P.S. & Baldwin, R.L. (1990). Intermediates in the folding reactions

of small proteins. Annu. Rev. Biochem. 59, 631–660.2. Dyson, H.J. & Wright, P.E. (1991). Defining solution conformations of

small linear peptides. Annu. Rev. Biophys. Chem. 20, 519–538.3. Dyson, H.J. & Wright, P.E. (1993). Peptide conformation and protein

folding. Curr. Opin. Struct. Biol. 3, 60–65.4 Scholtz, J.M. & Baldwin, R.L. (1992) The mechanism of a-helix forma-

tion by peptides. Annu. Rev. Biophys. Biomol. Struct. 21, 95–118.5. Baldwin, R.L. (1995). a-Helix formation by peptides of defined

sequence. Biophys. Chem. 55, 127–135.6. Carlsson, U. & Jonsson, B.H. (1995). Folding of b-sheet proteins.

Curr. Opin. Struct. Biol. 5, 482–487.7. Kim, C.A. & Berg, J.M. (1993). Thermodynamic b-sheet propensities

measured using a zinc-finger host peptide. Nature 362, 267–270.8. Minor, D.E. Jr. & Kim, P.S. (1994). Measurement of the b-sheet-

forming propensities of amino acids. Nature 367, 660–663.9. Minor, D.E. Jr. & Kim, P.S. (1994). Context is a major determinant of b-

sheet propensity. Nature 371, 264–267.10. Smith, C.K., Withka, J.M. & Regan, L. (1994). A thermodynamic scale

for the b-sheet forming tendencies of the amino acids. Biochemistry33, 5510–5517.

11. Otzen, D.E. & Fersht, A.R. (1995). Side-chain determinants of b-sheetstability. Biochemistry 34, 5718–5724.

12. Cox, J.P.H., Evans, P.A., Packman, L.C., Williams, D.H, & Woolfson,D.N. (1993). Dissecting the structure of a partially folded protein: cir-cular dichroism and nuclear magnetic resonance studies of peptidesfrom ubiquitin. J. Mol. Biol. 234, 483–492.

13. Blanco, F.J., Jiménez, M.A., Pineda, A., Rico, M., Santoro, J. & Nieto,J.L. (1994). NMR solution structure of the isolated N-terminal fragmentof protein-G B1 domain. Evidence of trifluoroethanol induced native-like b-hairpin formation. Biochemistry 33, 6004–6014.

14. Blanco, F.J., Rivas, G. & Serrano, L. (1994). A short linear peptide thatfolds into a native stable b-hairpin in aqueous solution. Nature Struct.Biol. 1, 584–590.

15 Searle, M.S., Williams, D.H. & Packman, L.C. (1995). A short linearpeptide derived from the N-terminal sequence of ubiquitin folds into awater-stable non-native b-hairpin. Nature Struct. Biol. 2, 999–1006.

16. Blanco, F.J., Jiménez, M.A., Herranz, J., Rico. M., Santoro, J. & Nieto,J.L. (1993). NMR evidence of a short linear peptide that folds into a b-hairpin in aqueous solution. J. Am. Chem. Soc. 115, 5887–5888.

17. Pflugrath, J.W., Wiegand, G., Huber, R. & Vèrtesy, L. (1986). Crystalstructure determination, refinement and the molecular model of the a-amylase inhibitor Hoe-467A. J. Mol. Biol. 189, 383–386.

18. Wilmot, C.M. & Thornton, J.M. (1988). Analysis and prediction of thedifferent types of b-turn in proteins. J. Mol. Biol. 203, 221–232.

19. Hutchinson, E.G. & Thornton, J.M. (1994). A revised set of potentialsfor b-turn formation in proteins. Protein Sci. 3, 2207–2216.

20. Rooman, M.J., Kocher J.P. & Wodak, S.J. (1991). Prediction of proteinbackbone conformation based on seven structure assignments: influ-ence of local interactions. J. Mol. Biol. 221, 961–979.

21. Wüthrich, K., Billeter, M. & Braun, W. (1984). Polypeptide secondarystructure determination by nuclear magnetic resonance observation ofshort proton–proton distances. J. Mol. Biol. 180, 715–740.

22. Wüthrich, K. (1986). NMR of Proteins and Nucleic Acids. J. Wiley &Sons, New York.

23. Jiménez, M.A., Nieto, J.L., Rico. M., Santoro, J., Herranz, J. & Bermejo,F. (1986). A study of the NH NMR signals of Gly-Gly-X-Ala tetrapep-tides in H2O at low temperature. J. Mol. Struct. 143, 435–438.

24. Wishart, D.S. & Sykes, B.D. (1994). Chemical shifts as a tool for struc-ture determination. Methods Enzymol. 239, 363–392.

25. Case, D.A., Dyson, H.J. & Wright, P.E. (1994). Use of chemical shiftsand coupling constants in nuclear magnetic resonance structuralstudies on peptides and proteins. Methods Enzymol. 239, 392–416.

26. Sibanda, B.L., Blundell, T.L. & Thornton, J.M. (1989). Conformation ofb-hairpins in protein structures. A systematic classification with appli-cations to modelling by homology, electron density fitting and proteinengineering. J. Mol. Biol. 206, 759–783.

27. Sibanda, B.L. & Thornton, J.M. (1991). Conformation of b-hairpins inprotein structures. Classification and diversity in homologous struc-tures. Methods Enzymol. 202, 59–82.

28. Woody, R.W. (1985). Circular dichroism of peptides. In The Peptides,vol. 7. (Hruby, J.R., ed), pp. 15–114, Academic Press, Orlando.

29. Chakrabartty, A., Kortemme, T., Padmanabhan, S. & Baldwin, R.L.(1993). Aromatic side-chain contribution to far-ultraviolet circulardichroism of helical peptides and its effect on measurement of helixpropensities. Biochemistry 32, 5560–5565.

30. Rose, G.D., Gierasch, L.M. & Smith, J.A. (1985). Turns in peptides andproteins. Adv. Protein Chem. 17, 1–109.

31. Johnson, W.C. Jr. (1988). Secondary structure of proteins through cir-cular dichroism spectroscopy. Annu. Rev. Biophys. Biophys. Chem.17, 145–166.

32. Bundi, A. & Wüthrich, K. (1979). 1H NMR parameters of the commonamino acid residues measured in aqueous solution of linear tetrapep-tides Gly-Gly-X-Ala. Biopolymers 18, 299–311.

33. De Alba, E., Blanco, F.J., Jiménez, M.A., Rico, M. & Nieto, J.L. (1995).Interactions responsible for the pH dependence of the b-hairpin con-formational population formed by a designed linear peptide. Eur. J.Biochem. 233, 283–292.

34. Jiménez, M.A., et al., & Nieto, J.L. (1994). Helix formation by the phos-pholipase A2 38–59 fragment: influence of chain shortening anddimerization monitored by NMR chemical shifts. Biopolymers 34,647–661.

35. Nelson, J.W. & Kallenbach, N.R. (1989). Persistence of the a-helixstop signal in S-peptide in trifluoroethanol solutions. Biochemistry 28,5256–5261.

36. Dyson, H.J., Merutka, G., Waltho, J.P., Lerner, R.A. & Wright, P.E.(1992). Folding of protein fragments comprising the completesequence of proteins. Models for initiation of protein folding. I. Myohe-merythrin. J. Mol. Biol. 226, 201–217.

37. Jiménez, M.A., et al., & Rico, M. (1993). CD and 1H NMR studies onthe conformational properties of peptide fragments from the C-terminaldomain of thermolysin. Eur. J. Biochem. 211, 569–581.

38. Blanco, F.J. & Serrano, L. (1995). Folding of protein G B1 domainstudied by the conformational characterization of fragments compris-ing its secondary structure elements. Eur. J. Biochem. 230, 634–649.

39. Günter, P., Braun, W. & Wüthrich, K. (1991). Efficient computation ofthree-dimensional protein structures from NMR using the programDIANA and the supporting programs CALIBA, HABAS and GLOMSA.J. Mol. Biol. 217, 517–530.

40. Wouters, M.A. & Curmi, P.M. (1995). An analysis of side chain interac-tions and pair correlations within antiparallel b-sheets: the differencesbetween backbone hydrogen-bonded and non-hydrogen-bondedresidue pairs. Proteins 22, 119–131.

41. Neri, D., Billeter, M., Wider, G. & Wüthrich, K. (1992). NMR determina-tion of residual structure in a urea-denatured protein, the 434-repres-sor. Science 257, 1559–1563.

42. Neri, D., Wider, G. & Wüthrich, K. (1992). 1H, 15N and 13C NMRassignments of the 434 repressor fragments 1–63 and 44–63unfolded in 7M urea. FEBS Lett. 303, 129–135.

43. Gill, S.C. & Von Hippel, P.H. (1989). Calculation of protein extinctioncoefficients from amino acid sequence data. Analyt. Biochem. 182,319–326.

44. Aue, W.P., Bertholdi, E. & Ernst, R.R. (1976). Two-dimensional spec-troscopy. Application to nuclear magnetic resonance. J. Chem. Phys.64, 2229–2246.

45. Rance, M. (1987). Improved techniques for homonuclear rotating-frame and isotropic mixing experiments. J. Magn. Reson. 74,557–564.

46. Jeener, J., Meier, B.H., Bachmann, P. & Ernst, R.A. (1979). Investiga-tion of exchange processes by two-dimensional NMR spectroscopy. J.Chem. Phys. 71, 4546–4553.

47. Kumar, A., Ernst, R.R. & Wüthrich, K. (1980). A two-dimensionalnuclear Overhauser enhancement (2D NOE) experiment for the eluci-dation of complete proton–proton cross-relaxation networks in biologi-cal macromolecules. Biochem. Biophys. Res. Commun. 95, 1–6.

48. Bothner-By, A.A., Stephens, R.L., Lee, J.M., Warren, C.D. & Jeanloz,R.W. (1984). Structure determination of a tetrasaccharide: transientnuclear Overhauser effects in the rotating frame. J. Am. Chem. Soc.106, 811–813.

49. Braunschweiler, L. & Ernst, R.R. (1983). Coherence transfer byisotropic mixing: application to proton correlation spectroscopy. J.Magn. Reson. 53, 521–528.

144 Folding & Design Vol 1 No 2

Because Folding & Design operates a ‘Continuous PublicationSystem’ for Research Papers, this paper will have beenpublished via the internet before being printed. Forinformation on how to gain access via the internet, see theexplanation on the contents page.