to aggregate [2]. There exists an obvious interest in the conformational study of -hairpin-forming peptides, as

Similar documents
Sequential resonance assignments in (small) proteins: homonuclear method 2º structure determination

THE TANGO ALGORITHM: SECONDARY STRUCTURE PROPENSITIES, STATISTICAL MECHANICS APPROXIMATION

Supporting Information

Resonance assignments in proteins. Christina Redfield

Physiochemical Properties of Residues

NMR parameters intensity chemical shift coupling constants 1D 1 H spectra of nucleic acids and proteins

Protein Structure. W. M. Grogan, Ph.D. OBJECTIVES

A prevalent intraresidue hydrogen bond stabilizes proteins

Biochemistry 530 NMR Theory and Practice

An NMR study on the -hairpin region of barnase José L Neira and Alan R Fersht

antibodies, it is first necessary to understand the solution structure antigenic human epithelial mucin core peptide. The peptide EXPERIMENTAL

BMB/Bi/Ch 173 Winter 2018

Identification of Two Antiparallel-sheet Structure of Cobrotoxin in Aqueous Solution by'hnmr

Timescales of Protein Dynamics

Section Week 3. Junaid Malek, M.D.

Central Dogma. modifications genome transcriptome proteome

Sequential Assignment Strategies in Proteins

Timescales of Protein Dynamics

Principles of Physical Biochemistry

Residual helical and turn structure in the denatured state of staphylococcal nuclease: analysis of peptide fragments Yi Wang and David Shortle

Introduction to Comparative Protein Modeling. Chapter 4 Part I

Conformational Geometry of Peptides and Proteins:

Secondary Structure. Bioch/BIMS 503 Lecture 2. Structure and Function of Proteins. Further Reading. Φ, Ψ angles alone determine protein structure

Christopher Pavlik Bioanalytical Chemistry March 2, 2011

Peptides And Proteins

NMR Characterization of Partially Folded and Unfolded Conformational Ensembles of Proteins

I690/B680 Structural Bioinformatics Spring Protein Structure Determination by NMR Spectroscopy

Protein Dynamics. The space-filling structures of myoglobin and hemoglobin show that there are no pathways for O 2 to reach the heme iron.

Native-like -structure in a Trifluoroethanol-induced Partially Folded State of the All- -sheet Protein Tendamistat

Introduction to" Protein Structure

Details of Protein Structure

Supersecondary Structures (structural motifs)

Protein structure. Protein structure. Amino acid residue. Cell communication channel. Bioinformatics Methods

Polypeptide Folding Using Monte Carlo Sampling, Concerted Rotation, and Continuum Solvation

NH 2. Biochemistry I, Fall Term Sept 9, Lecture 5: Amino Acids & Peptides Assigned reading in Campbell: Chapter

Packing of Secondary Structures

Structural and mechanistic insight into the substrate. binding from the conformational dynamics in apo. and substrate-bound DapE enzyme

Molecular Modeling lecture 2

Protein Structures: Experiments and Modeling. Patrice Koehl

What makes a good graphene-binding peptide? Adsorption of amino acids and peptides at aqueous graphene interfaces: Electronic Supplementary

Supporting Information. Copyright Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2009

Biology Chemistry & Physics of Biomolecules. Examination #1. Proteins Module. September 29, Answer Key

Computer simulations of protein folding with a small number of distance restraints

Magnetic Resonance Lectures for Chem 341 James Aramini, PhD. CABM 014A

Presenter: She Zhang

NMR, X-ray Diffraction, Protein Structure, and RasMol

Contents. xiii. Preface v

Supplementary Materials for

Table S1. Primers used for the constructions of recombinant GAL1 and λ5 mutants. GAL1-E74A ccgagcagcgggcggctgtctttcc ggaaagacagccgcccgctgctcgg

CHRIS J. BOND*, KAM-BO WONG*, JANE CLARKE, ALAN R. FERSHT, AND VALERIE DAGGETT* METHODS

Lecture 10. Assignment and Structure Determination in Proteins.

Read more about Pauling and more scientists at: Profiles in Science, The National Library of Medicine, profiles.nlm.nih.gov

1) NMR is a method of chemical analysis. (Who uses NMR in this way?) 2) NMR is used as a method for medical imaging. (called MRI )

LS1a Fall 2014 Problem Set #2 Due Monday 10/6 at 6 pm in the drop boxes on the Science Center 2 nd Floor

1. What is an ångstrom unit, and why is it used to describe molecular structures?

Secondary and sidechain structures

CONFOCHECK. Innovation with Integrity. Infrared Protein Analysis FT-IR

Determining Protein Structure BIBC 100

Model Mélange. Physical Models of Peptides and Proteins

Exam I Answer Key: Summer 2006, Semester C

Dana Alsulaibi. Jaleel G.Sweis. Mamoon Ahram

NMR in Medicine and Biology

Lecture 26: Polymers: DNA Packing and Protein folding 26.1 Problem Set 4 due today. Reading for Lectures 22 24: PKT Chapter 8 [ ].

Lecture 2 and 3: Review of forces (ctd.) and elementary statistical mechanics. Contributions to protein stability

Major Types of Association of Proteins with Cell Membranes. From Alberts et al

Structurele Biologie NMR

Introduction to Computational Structural Biology

Biotechnology of Proteins. The Source of Stability in Proteins (III) Fall 2015

Protein Structure Basics

Useful background reading

Basics of protein structure

Outline. The ensemble folding kinetics of protein G from an all-atom Monte Carlo simulation. Unfolded Folded. What is protein folding?

NMR study of complexes between low molecular mass inhibitors and the West Nile virus NS2B-NS3 protease

Molecular dynamics simulations of anti-aggregation effect of ibuprofen. Wenling E. Chang, Takako Takeda, E. Prabhu Raman, and Dmitri Klimov

Characterization of the Unfolding of Ribonuclease A by a Pulsed Hydrogen Exchange Study: Evidence for Competing Pathways for Unfolding

NMR of Nucleic Acids. K.V.R. Chary Workshop on NMR and it s applications in Biological Systems November 26, 2009

Introduction solution NMR

CHAPTER 29 HW: AMINO ACIDS + PROTEINS

NMR studies of protein folding

Protein Structure Prediction II Lecturer: Serafim Batzoglou Scribe: Samy Hamdouche

BIOCHEMISTRY Course Outline (Fall, 2011)

Supplementary Material

B O C 4 H 2 O O. NOTE: The reaction proceeds with a carbonium ion stabilized on the C 1 of sugar A.

4 Proteins: Structure, Function, Folding W. H. Freeman and Company

NMR Spectroscopy of Polymers

Effects of Chemical Exchange on NMR Spectra

Biochemistry 530 NMR Theory and Practice

arxiv: v1 [cond-mat.soft] 22 Oct 2007

Sensitive NMR Approach for Determining the Binding Mode of Tightly Binding Ligand Molecules to Protein Targets

Protein Structure Analysis and Verification. Course S Basics for Biosystems of the Cell exercise work. Maija Nevala, BIO, 67485U 16.1.

Proton Acidity. (b) For the following reaction, draw the arrowhead properly to indicate the position of the equilibrium: HA + K + B -

Figure 1. Molecules geometries of 5021 and Each neutral group in CHARMM topology was grouped in dash circle.

1. 3-hour Open book exam. No discussion among yourselves.

Measuring Protein Concentrations by NMR Spectroscopy

Supporting information to: Time-resolved observation of protein allosteric communication. Sebastian Buchenberg, Florian Sittel and Gerhard Stock 1

Solid-state NMR and proteins : basic concepts (a pictorial introduction) Barth van Rossum,

Physical principles of IR and Raman. Infrared Spectroscopy

Trpzip-based beta hairpin temperature jump IR studies enhanced by sitespecific

Supporting Information

NMR in Structural Biology

Transcription:

Research Paper 133 Conformational investigation of designed short linear peptides able to fold into -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 -helices. owever, very little is known about the factors leading to the formation of -hairpins. In order to get a good -hairpinforming model peptide, two peptides were designed on the basis of -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 NE and RE spectra, together with other NMR evidence, show that peptide IYSNPDGTWT forms a highly populated -hairpin in aqueous solution with a type I -turn plus a G1 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 hairpin with a standard type I -turn conformation, and the two forms are in fast dynamic equilibrium with one another. The effect of p demonstrates the existence of a stabilizing interaction between the Asn and Asp sidechains. The populations of -hairpin conformations increase in the presence of trifluoroethanol (a structure-enhancing solvent). n the other hand, some residual structure persists at a high denaturant concentration (8 M urea). Address: Instituto de Estructura de la Materia, Consejo Superior de Investigaciones Científicas, Serrano 119, 28006 Madrid, Spain. Correspondence to: Manuel Rico E-mail address: emjimenez@iem.csic.es Key words: -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 Conclusions: This work highlights the importance of the -turn residue composition in determining the particular type of -hairpin adopted by a peptide, though a role of interstrand sidechain interactions in the stabilization of the formed -hairpin is not discarded. The fact that trifluoroethanol can stabilize helices or -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 secondary structure formation plays an important role in the early steps of protein folding [1 3]. As a result, the identification of the factors governing the formation of these structural elements is of crucial importance in understanding the folding reaction. Because of the absence of tertiary interactions, the conformational study of designed peptides has become a convenient tool for identifying specific sidechain interactions involved in secondary structure formation. 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 -helices [4,5]. In contrast, information on the interactions affecting sheet or -hairpin formation is scarce, possibly because previous attempts to obtain monomeric, water-soluble, structure-forming peptides were hampered by their tendency to aggregate [2]. There exists an obvious interest in the conformational study of -hairpin-forming peptides, as little is known about the folding of all- proteins [6]. All these reasons stimulated us to search for model peptides that could fold into highly populated -hairpin conformations, and to identify factors responsible for -hairpin stability, by using amino acid substitutions or by modifying medium conditions. -sheet propensities have been determined in model proteins [7 11], but monomeric -sheet formation has been observed with only five linear peptides so far [12 16]. Even among them, only three fold into a -hairpin conformation 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 YQNPDGSQA (peptide 1) [16]. In peptide 1, the strand residues correspond to residues 15 23 of tendamistat, where they form a hairpin with a type I -turn [17]. The underlined

134 Folding & Design Vol 1 No 2 sequence was designed to have maximum individual site probability to form a type I -turn [18,19]. To improve hairpin formation, peptide 2 with the sequence IYS- NPDGTWT was designed to incorporate residues with higher -sheet propensities than those present in peptide 1. Also, to get rid of the cis/trans equilibrium arising from the peptide bond previous to proline, we examined peptide 3, with the sequence IYSNSDGTWT (see Fig. 1). n the basis of NMR data, NE pattern and proton chemical shifts, it is shown in this paper that the two designed peptides fold into -hairpin structures in aqueous solution. Differences in the turn residue composition have direct consequences not only on the population of -hairpin, but also on the nature of the -hairpin formed. Interactions responsible for the changes in -hairpin population with p have been analyzed and the effects of denaturants (8 M urea) and structure-stabilizing solvents (trifluoroethanol Figure 1 (a) + 3N I 1 (b) _ + 3N _ N Y 2 I 1 N S 3 T 10 N N Y 2 T 10 N N S 3 Schematic representation of the peptide backbone conformation corresponding to (a) the experimentally found -hairpin ( -hairpin 3:5) for peptides 1, 2, and 3, and (b) the additional -hairpin found in peptide 3 ( -hairpin 4:4). The (i, j) and the NN(i, j) NE crosspeaks experimentally observed for each -hairpin are shown by thin black arrows and the hydrogen bonds expected for each -hairpin by open arrows. The sequence of peptide 2 is given in -hairpin 3:5 (a) and that of peptide 3 in -hairpin 4:4 (b). W 9 N N N 4 N N 4 T 8 N W 9 N G 7 N T 8 N N P 5 N N G 7 N N D 6 S 5 D 6 [TFE]) on the -hairpin population have been probed. New information has thus been gained on the factors governing the formation of isolated -hairpins. Results Peptide design As in peptide 1, the chain-bend of peptide 2 has maximum individual site probability to form a type I turn [18,19]. The flanking amino acids were chosen among those with high propensities to adopt -strand conformations [7 11]. In addition, the motif Thr8-X-Thr10, a possible -strand-stabilizing motif as suggested by its occurrence in two of the four known -hairpin-forming peptides [12,14] was also introduced to enhance the population of -hairpin conformation. The hydrophilic/hydrophobic pattern characteristic of -sheets together with solubility criteria were also taken into account. As amino acids with high propensities to adopt strand conformations are usually very hydrophobic, solubility criteria are crucial in order to get a water-soluble non-aggregating peptide. Peptide 2 exists in two conformational states as a result of the cis/trans X-Pro isomerization with the trans X-Pro population predominating. In peptide 3, the Pro residue was substituted by Ser, which has the highest probability after Pro to be in position i+1 of the type I -turn [18]. This single residue substitution simplifies the spectrum and allows an investigation of the role of the turn rigidity in the formation of the -hairpin. Secondary structure predictions The lowest energy backbone conformations predicted for the designed peptides using the program Prelude [20] are given in Table 1. For peptide 2, a marked preference for strand or extended conformation is predicted for residues 1 5 and 8 9, and turn-like conformations for residues 6 7. For peptide 3, residues 1 3 and 8 9 adopt -strand conformations and residues 4 7 adopt turn-like states. Aggregation state ne-dimensional 1 NMR spectra were recorded in aqueous solution at peptide concentrations of 0.1 mm and 15 mm in the case of peptide 2, and at 0.15 mm and 15 mm in the case of peptide 3. The absence of any appreciable change in either linewidths or chemical shifts suggests the absence of aggregation. Nevertheless, in order to rule out the possibility of aggregates that no longer change their NMR properties in the concentration range of 0.1 15 mm, fluorescence spectra were recorded for the single Trp present in peptides 2 and 3 at lower concentrations than those used in NMR experiments (i.e. 10 100 M). The intensity of the Trp fluorescence emission at 350 nm (Trp excited at 280 nm) varied linearly with concentration up to 80 M. A slight curvature was observed for peptide concentrations between 80 and

Research Paper Folding of designed peptides into -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 B Tyr2 B B B Tyr2 B B B Ser3 P B P Ser3 B B C Asn4 B B B Asn4 G G G Pro5 P P P Ser5 C C C Asp6 C C C Asp6 C C C Gly7 G G G Gly7 G G G Thr8 B B C Thr8 B B B Trp9 B B P Trp9 B P B Thr10 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 energy conformation predicted. B, -strand; C, 3 10 helix; G, left-handed helix; P, extended; X, not computed; and, torsional angles corresponding to these conformational states are defined in Rooman et al. [20]. 100 mm due to the inner filter effect (data not shown). As another test to rule out aggregation, circular dichroism (CD) spectra were acquired at peptide concentrations of 50 and 100 M. Identical CD spectra were obtained at both concentrations, consistent with absence of aggregation in this concentration range (data not shown). nce the absence of aggregation over a very wide range of peptide concentrations (10 M to 15 mm) was confirmed, concentrations of peptide between 5 mm and 15 mm were used in all the NMR experiments reported here. Spectral assignment The 1 NMR spectra of peptides 2 and 3 in aqueous solution at various p values were completely assigned using the standard sequential assignment procedure [21,22], as were the assignments of the 1 NMR spectra of peptides 2 and 3 under denaturing conditions (6 8 M urea) and in the presence of 30% TFE. The 1 chemical shifts of peptides 2 and 3 in aqueous solution at p 4.3 and 5 C are listed in Tables 2 and 3, respectively. Conformational behavior in aqueous solution. The detection and identification of the folded conformations adopted by peptides 2 and 3 were based on the structural information provided by NE data, conformational shifts (deviation of the chemical shift values with respect to those in random coil [RC] peptides) and N shift-temperature coefficients. NE data are the strongest evidence for confirming the existence of a folded structure. In order to avoid spin diffusion problems, short mixing times were used in the NESY spectra. Furthermore, the same pattern of NE crosspeaks was found in the RESY spectra recorded under identical experimental Table 2 Chemical shifts (ppm from TSP) of peptide 2 in aqueous solution at 5 C and p 4.3. N C C ther Ile1 3.92 1.96 C 1.48, 1.19; C 3 0.98; C 3 0.86 Tyr2 8.86 4.70 3.00, 3.00 C 7.11; C 6.83 Ser3 8.20 4.30 3.56, 3.52 Asn4 8.53 4.87 2.84, 2.68 N 2 7.61, 7.09 Pro5 4.38 2.32, 1.98 C 2.04, 2.04; C 3.86, 3.79 Asp6 8.24 4.61 2.87, 2.76 Gly7 8.22 4.04, 3.84 Thr8 7.97 4.36 4.20 C 3 1.13 Trp9 8.46 4.94 3.35, 3.20 N 1 10.14; C 1 7.24; C 3 7.58; C 3 7.12; C 2 7.23; C 2 7.49 Thr10 8.08 4.22 4.18 C 3 1.16

136 Folding & Design Vol 1 No 2 Table 3 Chemical shifts (ppm from TSP) of peptide 3 in aqueous solution at 5 C and p 4.3. N C C ther Ile1 3.88 1.94 C 1.45, 1.16; C 3 0.96; C 3 0.87 Tyr2 8.83 4.68 2.97, 2.88 C 7.04; C 6.80 Ser3 8.33 4.39 3.68, 3.62 Asn4 8.58 4.70 2.84, 2.78 N 2 7.62, 7.04 Ser5 8.60 4.36 3.92, 3.85 Asp6 8.23 4.66 2.86, 2.76 Gly7 8.24 4.01, 3.84 Thr8 7.93 4.38 4.17 C 3 1.12 Trp9 8.49 4.91 3.34, 3.17 N 1 10.13; C 1 7.22; C 3 7.52 C 3 7.10; C 2 7.21; C 2 7.46 Thr10 8.04 4.20 4.17 C 3 1.14 conditions, confirming the absence of spin diffusion. RESY spectra are less sensitive to spin diffusion effects, and spin diffusion mediated crosspeaks can be distinguished easily from true RE signals by their opposite sign. The NE patterns obtained for peptides 2 and 3 from the combined analysis of RESY and NESY spectra are summarized in Figure 2. Selected regions of NESY and RESY spectra of both peptides showing some long-range, sequential, and intraresidue NE crosspeaks are given in Figures 3 and 4. Evidence for the presence of a chain-bend in both peptides comes from three points. First, there is the N (i, i+2) NE between residues Pro5 and Gly7 (Fig. 2) in peptide 2, i.e. the non-sequential NE expected to occur between residues i+1 and i+3 of a -turn [22] (the Figure 2 Ile1 Tyr2 Ser3Asn4 Pro5 Asp6Gly7 Thr8 Trp9 Thr10 Ile1 Tyr2 Ser3Asn4 Ser5 Asp6Gly7 Thr8 Trp9 Thr10 dαn dnn * dαn dnn * * * * * dαn(i,i+2) dαn(i,i+2) d(i,i+2) d(i,i+2) d(i,i+3) d(i,i+3) d(i,i+4) d(i,i+4) d(i.i+5) d(i.i+5) d(i,i+6) d(i,i+6) d(i,i+7) d(i,i+7) d(i,i+8) d(i,i+8) d(i,i+9) d(i,i+9) (a) (b) Summary of NE connectivities observed for (a) peptide 2 and (b) peptide 3 at 5 C and p 4.3. The thickness of the lines reflects the intensity of the sequential NE connectivities, i.e. weak, medium and strong. An asterisk or a broken line indicates unobserved NE connectivity due to signal overlapping, closeness to diagonal or overlap with solvent signal.

Research Paper Folding of designed peptides into -hairpin structures de Alba et al. 137 Figure 3 Selected regions of RESY and NESY spectra of peptide 2. (a) RESY region showing the long-range NE crosspeaks connecting the sidechains of residues Ile1 and Trp9 (200 ms mixing time). (b) RESY region showing the Ser3 Trp9 NE crosspeak (100 ms mixing time). Conditions: 5 mm peptide in D 2, 2 C, p 4.3. (c) NN region of a NESY spectrum, in which the NTyr2 NThr10 NE is boxed. Conditions: 15 mm peptide in 2 /D 2 (9:1 by volume), 5 C, p 4.3, 200 ms mixing time. corresponding NE between residues Ser5 and Gly7 in peptide 3 is observed only at p 6.3 due to spectral overlapping present at lower p values). Second, the very small temperature coefficients of the amide proton Figure 4 RESY spectrum of peptide 3. (a) Fingerprint region. The asterisks indicate a small impurity. Conditions: 15 mm peptide in 2 /D 2 (9:1 by volume), 5 C, p 4.3, 200 ms mixing time. (b) Spectral region showing NE crosspeaks. Conditions: 5 mm peptide in D 2, 2 C, p 6.3.

138 Folding & Design Vol 1 No 2 resonances of residues Asp6 and Gly7 (Table 4) indicate protection against solvent exchange by their possible involvement in intramolecular hydrogen bonds or solvent inaccessibility [23]. Third, chemical shifts, whose deviations from random coil values are diagnostic of secondary structure. conformational shifts, = RC ppm, are negative for helical or turn regions and positive in -conformations or extended conformations [24,25]. Some caution must be exercised in the application of the chemical shift criterion because of ring current effects of the aromatic residues Tyr2 and Trp9 on the chemical shifts. The conformational shifts of residues Asn4 to Gly7 are negative, in agreement with the presence of turn structures in this region, whereas some of the conformational shifts of the residues flanking the turn are positive, as expected for -strand conformations (Figs 5,6). The presence of a -hairpin structure of the type shown schematically in Figure 1a (a -hairpin 3:5 according to the -hairpin classification by Thornton and colleagues [26,27]) in peptide 2 in aqueous solution is strongly supported by the pattern of long-range NE crosspeaks observed, in particular the NE crosspeak connecting Ser3 with Trp9, and the one between NTyr2 and NThr10 (Fig. 3). The presence of these two crosspeaks can be explained only if these residues (Ser3 Trp9 and Tyr2 Thr10) face each other in the -hairpin conformation (as shown in Fig. 1a). Additional NE cross-correlations observed between residues from different strands and on the same side of the -hairpin, including those between sidechain protons of Ile1 and Trp9 (Fig. 3), are also consistent with the formation of the -hairpin 3:5 shown in Figure 1a. The positive conformational shifts observed for most of the residues in the -strands, as indicated above, are also in agreement with the formation of -hairpin 3:5. This type of -hairpin was also the one formed by peptide 1. A very weak Tyr2 Trp9 NE connectivity was also detected in peptide 2. This NE connectivity is incompatible with the -hairpin 3:5 and could be interpreted as resulting from the presence of a very low population of -hairpin 4:4 (estimated to be about 5%, whereas that of -hairpin 3:5 is almost 30%), as discussed below. During the reviewing process of this paper, a similar behavior was reported for a peptide deriving from the N-terminal sequence of ubiquitin and with the same turn residues, NPDG [15]. In peptide 3, the Ser3 Trp9 and the NTyr2 NThr10 NE cross-correlations characteristic of -hairpin 3:5, together with most of the long-range NE crosspeaks involving sidechain protons observed in peptide 2, are also present, which indicates that peptide 3 adopts a -hairpin 3:5 as well. In addition, a NE crosspeak connecting Tyr2 with Trp9 (stronger than in peptide 2) and another one between NSer3 and NThr8 were found in the NESY and RESY spectra of peptide 3 (Fig. 4). These two NE connectivities are incompatible with -hairpin 3:5. ne plausible explanation for these NE connectivities is the existence of another -hairpin structure, -hairpin 4:4 (Fig. 1b) [26,27], in dynamic conformational equilibrium with the previous one. The CD spectra of peptides 2 and 3 show a minimum at about 201 nm, slightly shifted from the one below 200 nm expected for random coil peptides, and a maximum at about 230 nm, usually observed in peptides containing aromatic residues in their sequence [28,29]. This fits with the presence of some non-random structure in the peptides. Identification of the nature of such non-random structure by CD is not possible due to the shortness of the two - Table 4 Amide proton temperature coefficients ( / T, ppb.k 1 ) obtained for peptides 2 and 3 under various experimental conditions. Peptide 2 Peptide 3 2 6M urea 8M urea 30% TFE 2 8M urea 30% TFE Residue 3.0 4.3 5.3 5.3 5.3 4.3 3.0 4.3 5.3 5.3 4.3 Ile1 Tyr2 8.3 8.7 12.7 9.0 8.3 9.7 7.7 8.7 9.0 8.7 9.3 Ser3 3.7 4.3 4.7 6.7 6.7 5.0 4.7 5.0 4.7 6.7 5.3 Asn4 7.3 7.6 7.3 8.0 7.7 7.7 6.3 7.0 7.3 7.7 7.0 Xaa5* 6.7 9.0 9.0 8.7 8.7 Asp6 5.6 3.2 2.3 3.7 3.0 1.3 4.0 1.0 0.3 3.0 0.7 Gly7 4.7 4.7 4.3 5.0 4.7 3.7 4.3 3.0 2.7 4.0 2.0 Thr8 5.0 4.2 4.0 4.7 4.7 2.3 4.0 3.3 3.0 4.7 2.3 Trp9 8.1 8.7 9.0 8.3 8.0 8.7 8.3 10.0 10.0 8.0 9.7 Thr10 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. p

Research Paper Folding of designed peptides into -hairpin structures de Alba et al. 139 Figure 5 0.25 0.20 (a) (b) (c) δ = δ - δrc, ppm 0.15 0.10 0.05 0.00-0.05-0.10-0.15 0.25 0.20 (d) (e) conformational shifts of peptide 2 at 5 C. The conformational shifts were obtained with reference to 6 M urea, denaturing conditions ( = RC ppm), in aqueous solution at (a) p 3.0, (b) p 4.3 and (c) p 5.3, and (d) in 30% TFE at p 4.3. (e) The conformational shifts obtained for peptide 2 and p 5.3 with reference to 8 M urea. δ = δ - δrc, ppm 0.15 0.10 0.05 0.00-0.05-0.10-0.15 strands (only three residues) and the variety of CD spectra found experimentally for the different types of -turns [30,31]. CD is not a suitable technique for identifying small populations of -turn and -hairpin conformations. Dependence of -hairpin population on p The p dependence of the NMR parameters observed for peptides 2 and 3 was investigated. At low p (3.0), relative to higher p values (4.3 and 5.3), the medium-range and long-range NE crosspeaks observed for peptide 2 are smaller in number and in intensity, the absolute values of conformational shifts decrease both for turn-like and -strand residues (Fig. 5), and the absolute values of the amide proton temperature coefficients for residues Asp6 and Gly7 increase (Table 4). The conformational shifts and amide proton temperature coefficients for residues Figure 6 conformational shifts of peptide 3 at 5 C and p 4.3. The conformational shifts measured for peptide 3 in (a) aqueous solution and (b) 30% TFE solution. δ = δ - δrc, ppm 0.25 0.20 0.15 0.10 0.05 0.00-0.05 (a) (b) -0.10-0.15

140 Folding & Design Vol 1 No 2 Asp6 and Gly7 of peptide 3 (Table 4) present a similar p dependence. Conformational behavior under denaturing conditions Chemical shifts of protons of amino acid X in tetrapeptides Gly-Gly-X-Ala are frequently used as reference for the disordered state [32]. The chemical shifts for most peptides show small variations from these random coil values. To avoid sequence effects on random coil values, the conformational shifts of peptides 2 and 3 were obtained, using as reference for the coil state chemical shifts obtained for the same peptide under denaturating conditions (6 M urea). This way of evaluating conformational shifts allows the detection of small populations of structure in partially folded peptides [33,34]. owever, the NESY and RESY spectra of peptides 2 and 3 in 6 M urea, performed in order to get the values under denaturing conditions, contain medium-range and long-range NE connectivities indicative of structure, though less numerous and weaker than in aqueous solution at the same p value. The long-range NE connectivities include those involving sidechain protons of Ile1 and Trp9. Furthermore, the absolute values of the amide proton temperature coefficient of Asp6 in both peptides 2 and 3 are small (Table 4). These findings indicate that some residual folded structure persists in peptides 2 and 3 under these denaturing conditions and also in higher concentrations of urea (7 M and 8 M; Table 4). The conformational shifts with reference to 6 M urea and to 8 M urea at p 5.3 and 5 C are plotted as a function of the peptide sequence in Figure 5 for peptide 2. The profiles of the conformational shifts obtained for peptide 2 (Fig. 5) and peptide 3 (data not shown) with respect to 6 M urea and with respect to 8 M urea are identical, except for the absolute values of the conformational shifts, being slightly larger when obtained from 8 M urea, consistent with the presence of a smaller but still persistent folded structure at the higher denaturant concentration. Conformational behavior in 30% trifluoroethanol solution TFE is known to increase the population of secondary structure formed by protein fragments [35 37]. It is generally accepted that TFE increases helical populations in peptides of different sequences in a non-specific way. owever, there are several cases in which the presence of mixed solvent alcohol/water also enhances the formation of isolated -hairpins [12,13,38]. Secondary structure stabilization by TFE probably occurs by increasing the strength of intra-peptide hydrogen bonds whichever structure they belong to. We therefore investigated the effect of TFE on the conformational behavior of peptides 2 and 3 in the mixed solvent 30% TFE by NMR. The absolute values of the conformational shifts obtained for peptides 2 and 3 in 30% TFE at p 4.3 and 5 C (Figs 5,6) are larger than those obtained for each peptide in aqueous solution at the same p and temperature. The increase in the absolute values of the conformational shifts is particularly large in the -strand regions. The profile of conformational shifts obtained for peptides 2 and 3 in 30% TFE (Figs 5,6) is that expected for an ideal -hairpin. In addition, the small absolute values of the amide temperature coefficients measured for residues Asp6, Gly7 and Thr8 in aqueous solution decrease to even smaller absolute values for both peptides in 30% TFE at p 4.3 (Table 4). The pattern of NE connectivities obtained for each peptide in 30% TFE is also almost identical to that previously obtained in aqueous solution at the same p. Both -hairpin conformations found for peptide 3 in aqueous solution are also present in 30% TFE, as indicated by the presence of the NE connectivities characteristic of -hairpin conformations 3:5 and 4:4. The only difference between the set of NE connectivities obtained for peptides 2 and 3 in aqueous solution and in 30% TFE is in the intensity of many NE crosspeaks which increase in 30% TFE, consistent with an increase in the populations of -hairpin conformations in TFE. It can also be noted that in peptide 3, the Ser3 Trp9 NE crosspeak (characteristic of -hairpin 3:5) is slightly weaker in aqueous solution and slightly stronger in 30% TFE than the Tyr2 Trp9 NE crosspeak (characteristic of -hairpin 4:4), possibly suggesting that the stabilizing effect of TFE on -hairpin 3:5 differs somewhat from that on -hairpin 4:4. In conclusion, TFE increases the population of -hairpin without changing the nature of -hairpin structure. Estimation of -hairpin population In contrast to -helices, no methods have yet been established for quantifying either -turn or -hairpin conformations adopted by peptides in solution. To estimate the population of the -hairpins formed by the model peptides 2 and 3, we use the ratio of the intensities of the NE characteristic of each -hairpin ( Ser3 Trp9 for -hairpin 3:5 and Tyr2 Trp9 for -hairpin 4:4) to that of the Gly7 NE (Intensity observed / Intensity Gly7 ), because being the Gly7 distance constant, the Gly7 NE intensity can be used as reference. The intensity ratio corresponding to a 100% -hairpin was taken as [d i j ] 6 / [d Gly7 ] 6, where d i j is the averaged i j distance value found for antiparallel -sheets in proteins [22], being i and j facing residues with the protons pointing towards the interior in an antiparallel -sheet, and d Gly7 is the distance characteristic of Gly residues. It is clear that this approach is approximate, because it implicitly assumes equal correlation times for the peptide in any conformational state and small variations in the actual

Research Paper Folding of designed peptides into -hairpin structures de Alba et al. 141 distance with respect to the averaged value in antiparallel -sheets of proteins can lead to large over- or under-estimations of the -hairpin relative population. Nevertheless, given the absence of any other available method, we are bound to use this method to estimate the -hairpin populations formed by peptides 2 and 3. Figure 7 This estimation yields a population for -hairpin 3:5 of about 27% for peptide 2 in aqueous solution at p 6.3 and 2 C, a conformation which would be interconverting in a fast dynamic equilibrium with about 5% of -hairpin 4:4 and a number of other conformations, including those corresponding to the random coil. In peptide 3 under the same experimental conditions, the conformational equilibrium would be formed by about 27% of -hairpin 3:5, 32% of -hairpin 4:4, and other disordered forms. Among these, the presence of molecules containing the -turn but not the -strands is highly probable. Peptide three-dimensional structures in aqueous solution Due to the usual conformational averaging that occurs in peptides, NE intensities cannot be rigorously interpreted in terms of a unique structure. Nevertheless, it is useful to calculate a limited number of structures compatible with NE constraints, which helps to visualize the conformational properties of the ensemble, albeit in a very simplified way. Structure calculations were performed on the basis of the complete set of NE restraints obtained for peptide 2 in aqueous solution at p 4.3, and using a distance geometry procedure [39]. A table, listing the NE constraints used for the structure calculation, is available as Supplementary material (published with this paper on the internet). angles were also constrained to the range 0 to 180, except for Gly. Starting from 250 randomized conformations, 31 conformers which satisfy the constraints with no violation greater than 0.13 Å were selected. The values of the pairwise root mean square deviations (RMSDs) for the backbone atoms of these 31 conformers are 0.7 ± 0.2 Å. The superposition of six of these calculated structures for peptide 2 are shown in Figure 7. The set of structures calculated without any angle constraint was very similar to those shown in Figure 7, but less well defined (RMSD for the backbone atoms 1.4 ± 0.5 Å). Stereoscopic view of the superposition of six calculated structures for peptide 2. (a) Showing sidechain and backbone atoms. (b) Showing only backbone atoms. The calculated structures for peptide 2 on the basis of the experimental set of NE restraints coincide with the hairpin 3:5 schematically represented in Figure 1a. Many of the calculated structures present a hydrogen bond between the C group of Asn4 and the N of Gly7, and a hydrogen bond between the N of Asn4 and the C group of Thr8. These data, together with the range of values obtained for the angles and of residues Asn4, Pro5, Asp6, Gly7, and Thr8, indicate that the conformation of the chain-bend region is not a standard type I turn but a type I -turn plus a G1 -bulge [26,27], which gives the pattern of alternating residues in the -strands represented in Figure 1a and observed in Figure 7. The hairpin formed by peptide 2 can be classified as a hairpin 3:5 [26,27]. A similar Distance Geometry calculation was performed for peptide 3 in order to check the compatibility of all the NE connectivities observed with a unique conformation. The calculation performed using a set of NE data including all the NE restraints observed leads to large NE constraint violations. This result evidences the presence of some incompatible distance constraints within the structures corresponding to -hairpin 3:5 (Fig. 1a). As expected, the NE restraints with large violations include Tyr2 Trp9 and NSer3 NThr8 NE connectivities.

142 Folding & Design Vol 1 No 2 In order to calculate the additional conformation observed in peptide 3, -hairpin 4:4 (Fig. 1b), a set of NE data excluding all those NE restraints present in -hairpin 3:5 were used. The calculated structures were not well defined due to the smallness of the NE data set. Nevertheless, the range of and angles of the chain-bend region coincide with the ones usually observed in a hairpin with a four-residue chain-bend, that is a standard type I -turn [26,27]. It is likely that the conformational freedom gained by the substitution of Pro5 by Ser allows the chain-bend to adopt two different conformations: a type I -turn plus G1 -bulge resulting in a five-residue loop -hairpin ( -hairpin 3:5) and a standard type I -turn resulting in a four-residue loop -hairpin ( -hairpin 4:4). Discussion The -hairpin-forming propensities shown by peptide 1, reported previously [16], are conserved or even improved by the two designed peptides analyzed in this work, indicating the usefulness of statistical data on site probabilities and intrinsic propensities determined in proteins in designing -strand-forming peptides. A next step would be to use the information about pairwise sidechain interactions, recently determined by statistical methods for antiparallel -sheets [40], to experimentally confirm the existence of these and other sidechain interactions responsible for the -hairpin stability. Peptide 2 has many of the characteristics to serve as a convenient model peptide for the study of these stabilizing interactions. It is water soluble, it does not aggregate and forms a -hairpin (type 3:5, Fig. 1a) with a relatively high population (~30 %). Stabilizing interactions could be identified by analyzing the changes in relative -hairpin populations resulting from specific amino acid substitutions. An example of the effects of amino acid substitution is seen in the different conformational properties of peptides 2 and 3. Although the main purpose of substituting Pro by Ser at position 5 was to remove the set of signals arising from the Asn Pro cis isomer, the introduction of a Ser residue in the turn composition also revealed the existence of a different significantly populated -hairpin 4:4, while conserving that of -hairpin 3:5. The -hairpin 4:4 has a four-residue turn and an inside-out inversion of the first -strand in contrast to -hairpin 3:5 which has a fiveresidue turn (Fig. 1). A possible explanation for the differences found between the conformational properties of peptides 2 and 3 may lie in the greater flexibility of the Ser peptide in the turn region. As the -hairpin 4:4 conformation also appears to be present in peptide 2, but populated to a very small degree, the two conformations probably do not differ energetically by a significant amount. The fact that a single-residue substitution in the -turn region leads to the appearance of a new -hairpin conformation highlights the importance of the turn region in defining the final -hairpin conformation. When the backbone conformations are examined using the program Prelude [20], residues 1 3 and 8 9 in both peptides predicted to be in a -strand state (Table 1) are in a -strand state for both -hairpins 3:5 and 4:4. The and angles of residue Asn4 correspond to a -strand state in -hairpin 3:5, consistent with the secondary structure prediction for Asn4 in peptide 2 (Table 1). The agreement between predicted and experimental secondary structure is not so good for peptide 3, as residue Asn4 is predicted to be in a left-handed helix, but corresponds to a right-handed helix [26,27] in -hairpin 4:4. Anyway, it is interesting to note that the different secondary structure predictions for peptides 2 and 3 correlate with different conformational properties. The -hairpin population in peptide 2 is p dependent, decreasing at low p values. This p conformational dependence has also been found in peptide 1, spanning identical -turn residues (NPDG), and originates in a hydrogen-bonding interaction between the sidechains of Asn and Asp which is destabilized by the protonation of the Asp sidechain [33]. This same interaction can account for the p conformational dependence existent in peptide 2 and also in the two -hairpin conformations adopted by peptide 3. Thus, the interaction between the Asn and Asp sidechains within the sequence Asn-X-Asp is not restricted to the presence of a Pro residue in the central position, X. Probably, such an interaction exists whenever Asn, X and Asp residues have the necessary geometry (, ) and proximity. A striking result is that some residual non-random conformations persist in peptides 2 and 3 under normally denaturing conditions (8 M urea). A similar observation of residual non-random structure in 7 M urea has been observed previously 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 hydrophobic cluster [41,42]. In peptides 2 and 3, a hydrophobic interaction between residues Ile1 and Trp9, only partially affected by urea, may be sufficiently strong to allow the persistence of structure, at least in NMR-detectable populations in 8 M urea. An important consequence of these findings is that extreme caution must be taken when using urea as a reference for the fully random reference state of peptides and proteins, as indicated by Neri et al. [41]. The populations of the -hairpin conformation formed by peptide 2 and those of the two -hairpins adopted by peptide 3 increase significantly in the mixed solvent 30% TFE. The -hairpin conformation formed by peptide 1 is also stabilized in 30% TFE (unpublished data). These results together with those found in other peptides [13,38] indicate that TFE is not only a helix-promoting solvent, but may promote -hairpin conformations as well, at least in short monomeric peptides. Thus, the effect of TFE on

Research Paper Folding of designed peptides into -hairpin structures de Alba et al. 143 the structure adopted by a determined peptide may depend on the peptide sequence. It is interesting that the relative populations of the two -hairpin conformations adopted by peptide 3 appear to be slightly different in aqueous solution and in 30% TFE solution, suggesting that TFE has different stabilizing effects for different hairpin conformations. The results obtained in this work support the idea that the chain-bend composition is more important in determining the nature of the -hairpin to be adopted than the strand residues and their possible pairwise interactions. Pairwise interactions were not explicitly included in the design of peptides 2 and 3. For design purposes, it will be important to determine the relative importance of the effect of particular residues in the chain-bend and in the -strands on the nature of the -hairpin formed. We are currently exploring the effects of both by changing the turn residues, i.e. taking those residues to have maximal -turn probability in each position of a -turn I, instead of a turn I, and by optimizing interstrand pairwise interactions for -hairpin 3:5 or for -hairpin 4:4. In conclusion, the results obtained on the basis of the NMR conformational study indicate that the composition of the chain-bend region is important for the final conformation of the -hairpin, although further investigation is required to clarify the influence of interstrand interactions in -hairpin formation. The early formation of backbone turns or bends is of primary importance in the folding process because it reduces the conformational space and provides the opportunity for tertiary interactions to occur. The use of the peptides designed here as reference models and the conformational analysis of designed site mutant peptides appears to be a most promising way to obtain important information about the factors governing -hairpin formation. Materials and methods Materials Peptides were chemically synthesized by Neosystem Laboratoire (Strasbourg, France). Urea was from Fluka Chemie, AG (Buchs, Switzerland). 2,2,2-Trifluoroethanol-d 3 was purchased from Cambridge Isotopes Laboratories (Cambridge, USA). Secondary structure predictions The program Prelude [20] developed to identify conformational preferences of protein backbone segments was used to predict peptide secondary structure. Fluorescence data Fluorescence emission spectra were recorded in an AMINC BWMAN SERIES 2 Luminescence Spectrometer after excitation at 280 nm on peptide samples at p 4.3 and 4.5 C (peptide 2) and 5 C (peptide 3). Slitwidths were 4 nm. CD measurements CD spectra were acquired on an Aviv 60 DS Spectropolarimeter with a ewlett-packard 89100A temperature control unit and cuvettes with either 10-mm or 1-mm path lengths. CD spectra were the average of three scans obtained by collecting data at 1 nm intervals from 250 186 nm. Peptide samples were at p 4.3 and 5 C. Peptide concentrations were determined by UV absorbance [43]. Proton NMR spectra Peptide solutions were 15 mm in 0.5 ml of 2 /D 2 (9:1 ratio by volume), except for experiments under denaturating conditions, in TFE or in pure D 2, where the peptide concentrations were around 5 mm. The p, adjusted by adding minimal amounts of dilute DCl or NaD, was measured with a glass microelectrode and the reported values were not corrected for isotope effects. Samples for experiments under denaturating conditions were prepared by adding the appropriate amount of urea and the p was readjusted. To 5 mm peptide samples, deuterated TFE was added to 30% by volume and the p adjusted as required. The temperature of the NMR probe was calibrated using a methanol sample. Sodium 3-trimethylsilyl [2,2,3,3-2 ] propionate (TSP) was used as an internal reference at 0.00 ppm. The 1 NMR spectra were acquired on a Bruker AMX-600 pulse spectrometer operating at a proton frequency of 600.13 Mz. ne-dimensional spectra were acquired using 32 K data points, which were zero-filled to 64 K data points before performing the Fourier transformation. Phase-sensitive two-dimensional scalar coupling correlated spectroscopy (CSY) [44,45], total correlation spectroscopy (TCSY) [45], nuclear verhauser effect spectroscopy (NESY) [46,47] and rotating frame exchange spectroscopy (RESY) [48] spectra were recorded by standard techniques using presaturation of the water signal and the timeproportional phase incrementation mode. A mixing time of 200 ms was used for NESY and RESY spectra (apart from NESY experiments in TFE which were recorded using a mixing time of 100 ms). TCSY spectra were recorded using 80 ms MLEV 17 spin-lock sequence [49] at different temperatures to obtain amide proton temperature dependence. Acquisition data matrices were defined by 2048 512 points in the frequency domain. Additional NESY and RESY spectra (100 ms mixing time) were recorded on peptide samples in pure D 2 to facilitate the observation of NE crosspeaks too close to the water signal. Data were processed using the standard UXNMR Bruker programs in an Aspect X32 computer. The two-dimensional data matrix was multiplied by a phase-shifted square-sine-bell window function and zero-filled to 4 K 2 K prior to Fourier transformation. Baseline correction was applied in both dimensions. Structure calculations NE intensities were evaluated in a qualitative way as strong, medium and weak and were used to obtain upper limit distance constraints of 3 Å, 4 Å and 5 Å, respectively. Pseudoatom corrections were added where necessary. Structures were calculated on a Silicon Graphics Indigo computer using the program DIANA [39]. As erroneous results can be obtained due to the contribution of the heterogeneous conformational behavior of the peptide to the intraresidue and sequential NE crosspeak intensities, only medium-range and long-range NE crosspeaks were considered to calculate the set of structures of peptide 2. Another set of structures was calculated for peptide 2 with the φ angles constrained to the range 0 to 180, except for Gly. For peptide 3, two calculations were performed: one with a complete set of NE data and the other with a set of NE data excluding all those NE constraints in the -hairpin 3:5 (Fig. 1). Supplementary material available A tabulation listing the medium-range and long-range NE connectivities observed for peptides 2 and 3 in aqueous solution at p 4.3 and 5 C is published with this paper on the internet. Acknowledgements We are grateful to Professor RL Baldwin and all his lab group for the use of fluorescence and CD facilities at Stanford University and to Dr S Padmanabhan for his critical reading of the manuscript. This work was supported by the Spanish DGYCT project no PB90-1020.

144 Folding & Design Vol 1 No 2 References 1. Kim, P.S. & Baldwin, R.L. (1990). Intermediates in the folding reactions of small proteins. Annu. Rev. Biochem. 59, 631 660. 2. Dyson,.J. & Wright, P.E. (1991). Defining solution conformations of small linear peptides. Annu. Rev. Biophys. Chem. 20, 519 538. 3. Dyson,.J. & Wright, P.E. (1993). Peptide conformation and protein folding. Curr. pin. Struct. Biol. 3, 60 65. 4 Scholtz, J.M. & Baldwin, R.L. (1992) The mechanism of -helix formation by peptides. Annu. Rev. Biophys. Biomol. Struct. 21, 95 118. 5. Baldwin, R.L. (1995). -elix formation by peptides of defined sequence. Biophys. Chem. 55, 127 135. 6. Carlsson, U. & Jonsson, B.. (1995). Folding of -sheet proteins. Curr. pin. Struct. Biol. 5, 482 487. 7. Kim, C.A. & Berg, J.M. (1993). Thermodynamic -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 -sheetforming propensities of amino acids. Nature 367, 660 663. 9. Minor, D.E. Jr. & Kim, P.S. (1994). Context is a major determinant of sheet propensity. Nature 371, 264 267. 10. Smith, C.K., Withka, J.M. & Regan, L. (1994). A thermodynamic scale for the -sheet forming tendencies of the amino acids. Biochemistry 33, 5510 5517. 11. tzen, D.E. & Fersht, A.R. (1995). Side-chain determinants of -sheet stability. Biochemistry 34, 5718 5724. 12. Cox, J.P.., Evans, P.A., Packman, L.C., Williams, D., & Woolfson, D.N. (1993). Dissecting the structure of a partially folded protein: circular dichroism and nuclear magnetic resonance studies of peptides from 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 fragment of protein-g B 1 domain. Evidence of trifluoroethanol induced nativelike -hairpin formation. Biochemistry 33, 6004 6014. 14. Blanco, F.J., Rivas, G. & Serrano, L. (1994). A short linear peptide that folds into a native stable -hairpin in aqueous solution. Nature Struct. Biol. 1, 584 590. 15 Searle, M.S., Williams, D.. & Packman, L.C. (1995). A short linear peptide derived from the N-terminal sequence of ubiquitin folds into a water-stable non-native -hairpin. Nature Struct. Biol. 2, 999 1006. 16. Blanco, F.J., Jiménez, M.A., erranz, J., Rico. M., Santoro, J. & Nieto, J.L. (1993). NMR evidence of a short linear peptide that folds into a hairpin in aqueous solution. J. Am. Chem. Soc. 115, 5887 5888. 17. Pflugrath, J.W., Wiegand, G., uber, R. & Vèrtesy, L. (1986). Crystal structure determination, refinement and the molecular model of the amylase inhibitor oe-467a. J. Mol. Biol. 189, 383 386. 18. Wilmot, C.M. & Thornton, J.M. (1988). Analysis and prediction of the different types of -turn in proteins. J. Mol. Biol. 203, 221 232. 19. utchinson, E.G. & Thornton, J.M. (1994). A revised set of potentials for -turn formation in proteins. Protein Sci. 3, 2207 2216. 20. Rooman, M.J., Kocher J.P. & Wodak, S.J. (1991). Prediction of protein backbone conformation based on seven structure assignments: influence of local interactions. J. Mol. Biol. 221, 961 979. 21. Wüthrich, K., Billeter, M. & Braun, W. (1984). Polypeptide secondary structure determination by nuclear magnetic resonance observation of short 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., erranz, J. & Bermejo, F. (1986). A study of the N NMR signals of Gly-Gly-X-Ala tetrapeptides in 2 at low temperature. J. Mol. Struct. 143, 435 438. 24. Wishart, D.S. & Sykes, B.D. (1994). Chemical shifts as a tool for structure determination. Methods Enzymol. 239, 363 392. 25. Case, D.A., Dyson,.J. & Wright, P.E. (1994). Use of chemical shifts and coupling constants in nuclear magnetic resonance structural studies on peptides and proteins. Methods Enzymol. 239, 392 416. 26. Sibanda, B.L., Blundell, T.L. & Thornton, J.M. (1989). Conformation of -hairpins in protein structures. A systematic classification with applications to modelling by homology, electron density fitting and protein engineering. J. Mol. Biol. 206, 759 783. 27. Sibanda, B.L. & Thornton, J.M. (1991). Conformation of -hairpins in protein structures. Classification and diversity in homologous structures. Methods Enzymol. 202, 59 82. 28. Woody, R.W. (1985). Circular dichroism of peptides. In The Peptides, vol. 7. (ruby, J.R., ed), pp. 15 114, Academic Press, rlando. 29. Chakrabartty, A., Kortemme, T., Padmanabhan, S. & Baldwin, R.L. (1993). Aromatic side-chain contribution to far-ultraviolet circular dichroism of helical peptides and its effect on measurement of helix propensities. Biochemistry 32, 5560 5565. 30. Rose, G.D., Gierasch, L.M. & Smith, J.A. (1985). Turns in peptides and proteins. Adv. Protein Chem. 17, 1 109. 31. Johnson, W.C. Jr. (1988). Secondary structure of proteins through circular dichroism spectroscopy. Annu. Rev. Biophys. Biophys. Chem. 17, 145 166. 32. Bundi, A. & Wüthrich, K. (1979). 1 NMR parameters of the common amino acid residues measured in aqueous solution of linear tetrapeptides 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 p dependence of the -hairpin conformational population formed by a designed linear peptide. Eur. J. Biochem. 233, 283 292. 34. Jiménez, M.A., et al., & Nieto, J.L. (1994). elix formation by the phospholipase A 2 38 59 fragment: influence of chain shortening and dimerization monitored by NMR chemical shifts. Biopolymers 34, 647 661. 35. Nelson, J.W. & Kallenbach, N.R. (1989). Persistence of the -helix stop signal in S-peptide in trifluoroethanol solutions. Biochemistry 28, 5256 5261. 36. Dyson,.J., Merutka, G., Waltho, J.P., Lerner, R.A. & Wright, P.E. (1992). Folding of protein fragments comprising the complete sequence of proteins. Models for initiation of protein folding. I. Myohemerythrin. J. Mol. Biol. 226, 201 217. 37. Jiménez, M.A., et al., & Rico, M. (1993). CD and 1 NMR studies on the conformational properties of peptide fragments from the C-terminal domain of thermolysin. Eur. J. Biochem. 211, 569 581. 38. Blanco, F.J. & Serrano, L. (1995). Folding of protein G B1 domain studied by the conformational characterization of fragments comprising its secondary structure elements. Eur. J. Biochem. 230, 634 649. 39. Günter, P., Braun, W. & Wüthrich, K. (1991). Efficient computation of three-dimensional protein structures from NMR using the program DIANA and the supporting programs CALIBA, ABAS and GLMSA. J. Mol. Biol. 217, 517 530. 40. Wouters, M.A. & Curmi, P.M. (1995). An analysis of side chain interactions and pair correlations within antiparallel -sheets: the differences between backbone hydrogen-bonded and non-hydrogen-bonded residue pairs. Proteins 22, 119 131. 41. Neri, D., Billeter, M., Wider, G. & Wüthrich, K. (1992). NMR determination of residual structure in a urea-denatured protein, the 434-repressor. Science 257, 1559 1563. 42. Neri, D., Wider, G. & Wüthrich, K. (1992). 1, 15 N and 13 C NMR assignments of the 434 repressor fragments 1 63 and 44 63 unfolded in 7M urea. FEBS Lett. 303, 129 135. 43. Gill, S.C. & Von ippel, P.. (1989). Calculation of protein extinction coefficients from amino acid sequence data. Analyt. Biochem. 182, 319 326. 44. Aue, W.P., Bertholdi, E. & Ernst, R.R. (1976). Two-dimensional spectroscopy. Application to nuclear magnetic resonance. J. Chem. Phys. 64, 2229 2246. 45. Rance, M. (1987). Improved techniques for homonuclear rotatingframe and isotropic mixing experiments. J. Magn. Reson. 74, 557 564. 46. Jeener, J., Meier, B.., Bachmann, P. & Ernst, R.A. (1979). Investigation 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-dimensional nuclear verhauser enhancement (2D NE) experiment for the elucidation of complete proton proton cross-relaxation networks in biological 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: transient nuclear verhauser effects in the rotating frame. J. Am. Chem. Soc. 106, 811 813. 49. Braunschweiler, L. & Ernst, R.R. (1983). Coherence transfer by isotropic mixing: application to proton correlation spectroscopy. J. Magn. Reson. 53, 521 528. Because Folding & Design operates a Continuous Publication System for Research Papers, this paper will have been published via the internet before being printed. For information on how to gain access via the internet, see the explanation on the contents page.