conformational ensembles of neuromedin c reveal a progressive coil-helix transition … ·...

16
Vrije Universiteit Brussel Conformational ensembles of neuromedin C reveal a progressive coil-helix transition within a binding-induced folding mechanism Adrover, Miquel; Sanchis, Pilar; Vilanova, Bartolome; Pauwels, Kris; Martorell, Gabriel; Jesus Perez, Juan Published in: RSC Advances DOI: 10.1039/c5ra12753j Publication date: 2015 Document Version: Accepted author manuscript Link to publication Citation for published version (APA): Adrover, M., Sanchis, P., Vilanova, B., Pauwels, K., Martorell, G., & Jesus Perez, J. (2015). Conformational ensembles of neuromedin C reveal a progressive coil-helix transition within a binding-induced folding mechanism. RSC Advances, 5, 83074 - 83088. https://doi.org/10.1039/c5ra12753j General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Download date: 23. Jan. 2020

Upload: others

Post on 30-Dec-2019

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Vrije Universiteit Brussel

Conformational ensembles of neuromedin C reveal a progressive coil-helix transitionwithin a binding-induced folding mechanismAdrover, Miquel; Sanchis, Pilar; Vilanova, Bartolome; Pauwels, Kris; Martorell, Gabriel; JesusPerez, JuanPublished in:RSC Advances

DOI:10.1039/c5ra12753j

Publication date:2015

Document Version:Accepted author manuscript

Link to publication

Citation for published version (APA):Adrover, M., Sanchis, P., Vilanova, B., Pauwels, K., Martorell, G., & Jesus Perez, J. (2015). Conformationalensembles of neuromedin C reveal a progressive coil-helix transition within a binding-induced foldingmechanism. RSC Advances, 5, 83074 - 83088. https://doi.org/10.1039/c5ra12753j

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal

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

Download date: 23. Jan. 2020

Page 2: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

RSC Advances

PAPER

Conformational

aInstitut Universitari d'Investigacio en Cienc

Quımica, Universitat de les Illes Balears (U

Palma de Mallorca, Spain. E-mail: miquel.abInstituto de Investigacion Sanitaria de Pa

07010, Palma de Mallorca, SpaincStructural Biology Brussels, Vrije Univers

Brussels, BelgiumdStructural Biology Research Centre, Vlaa

Pleinlaan 2, 1050 Brussels, BelgiumeServeis Cientıco-Tecnics, Universitat de les7.5, E-07122, Palma de Mallorca, SpainfDepartament d'Enginyeria Quımica, Unive

ETSEIB, Av. Diagonal, 647, E-08028, Barcel

† Electronic supplementary informa10.1039/c5ra12753j

Cite this: RSC Adv., 2015, 5, 83074

Received 1st July 2015Accepted 15th September 2015

DOI: 10.1039/c5ra12753j

www.rsc.org/advances

83074 | RSC Adv., 2015, 5, 83074–830

ensembles of neuromedin C reveala progressive coil-helix transition within a binding-induced folding mechanism†

Miquel Adrover,*ab Pilar Sanchis,ab Bartolome Vilanova,ab Kris Pauwels,cd

Gabriel Martorelle and Juan Jesus Perezf

Neuromedin C (NMC) is a peptide that regulates various processes in the central nervous system and

gastrointestinal tract through its interaction with the bombesin receptor subtype-2 (BB2R). Hence, BB2R

antagonists hold the potential to treat disorders that occur as a result of NMC dysfunction or

misregulation. However, their efficient design requires a detailed understanding of the structural features

of NMC, which hitherto are unknown. Herein, we describe the conformational ensembles of NMC in an

aqueous solution, at five different TFE concentrations to decode its folding pathway and under its SDS

micelle bound state. NMC displays a disordered but well defined backbone architecture that undergoes a

progressive coil-helix transition with increasing TFE concentration, first at the C-terminus and then at

the N-terminus. NMC also adopts a C-terminal a-helical conformation upon binding to SDS micelles.

This micelle binding is directed by hydrophobic interactions that concur with the unfavourable

deprotonation of His8 and its further insertion into the micelle. Moreover, NMR relaxation data reveal

that the acquisition of the micelle bound a-helical conformation constrains the NMC flexibility more

than the confinement itself. This comprehensive study of the structural behaviour of NMC provides

essential mechanistic information that could be useful for the development of new therapeutics to treat

neurological, cancer-related or eating disorders.

1 Introduction

Neuromedin C (NMC) is an endogenous decapeptide(GNHWAVGHLM-NH2) that is highly conserved in mammals.1 Itexerts a variety of biological effects both on the central nervoussystem (CNS) and in the gastrointestinal tract.1 Together withgastrin-releasing peptide (GRP) and neuromedin B (NMB), NMCbelongs to the bombesin-like peptide family. Bombesin is a 14-residue peptide originally isolated from the amphibian Bombinabombina.2 In addition its two mammalian analogues (GRP and

ies de la Salut (IUNICS), Departament de

IB), Ctra. Valldemossa km 7.5, E-07122,

[email protected]

lma (IDISPA), Ctra. Valldemossa, 79, E-

iteit Brussels (VUB), Pleinlaan 2, 1050

ms Instituut voor Biotechnologie (VIB),

Illes Balears (UIB), Ctra. Valldemossa km

rsitat Politecnica de Catalunya (UPC),

ona, Spain

tion (ESI) available. See DOI:

88

NMB), NMC functions as a neurotransmitter, paracrinehormone, growth factor and it retains the full hormone activityof GRP.3

NMC exerts its physiological function mainly by its interac-tion with the subtype-2 bombesin receptor (BB2R), which is amember of the G-protein coupled receptor superfamily4 that islocated in the gut and in the CNS.5 For instance, NMC mediatesneurotransmission and neuromodulation,6 and it is able toexcite specic neurons by decreasing the resting potassiumconductance and increasing the non-specic conductance.7

NMC can also reduce the appetite, and therefore can act as aanorexia inducer,8 likely through its interaction with bombesinreceptors in the central amygdala.9 In addition, its intravenousadministration increases growth hormone levels in calves,10

while it can also act as an autocrine growth factor in humansmall-cell lung cancer.11 Moreover, NMC has been shown toregulate growth and/or differentiation of human tumors in awide range of tissues, including carcinomas of pancreas,stomach, breast, prostate and colon.12 Accordingly, a novelprotein vaccine comprising six covalently linked repeats of NMCwas successful in suppressing the proliferation of breast tumorscells.13

As a result of NMC's pharmacological prole, BB2R antago-nists are considered as prospective anticancer therapeutics14

and for the treatment of other illnesses.15 RC-3095, a

This journal is © The Royal Society of Chemistry 2015

Page 3: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Paper RSC Advances

peptidomimetic of NMC, was shown to produce long-lastingtumor regressions in different human models,16 as well as toshow benecial effects during the treatment of tumor necrosisfactor-dependent chronic inammatory conditions.17 Morerecently, a N-terminal modied NMC with acyclic tetraaminesfor binding of 99mTc ([99mTc]-demomedin C) was successfullytargeted in BB2R expressing tumor cells as a potent agonistinducing selective intracellular calcium release and triggeringGRP receptor mediated internalization of the radioligand.18

However, its tolerability, background radioactivity and reten-tion in tumor lesions warrant future studies as these pharma-cological aspects have led to the rejection of other99mTc-bombesin analogs.19

The efficient design of more potent antagonists of BB2Rrequires a detailed understanding of the structure–activity rela-tionships of NMC. However, very limited structural informationon NMC is currently available. In contrast to bombesin or NMB,NMC was predicted not to adopt a a-helical conformation uponbinding to non-polar sites, due to reduced hydrophobic interac-tions that arise from the replacement of Leu3 by His3 in NMB.20

Polverini et al. used CD spectroscopy to suggest that NMC couldadopt a helical-like conformation upon binding to lipids.21

Furthermore, NMR spectroscopy was applied to solve the solu-tion structure of the NMC–Ni2+ complex that consists of twoconnected turns,22 also likely adopted in the NMC–Cu2+ complexthat could be physiologically involved in metal transport alongthe CNS.23 More recently, we used replica exchange moleculardynamics (REMD) to demonstrate that NMC, in a simulatedaqueous environment, adopts different conformations resem-bling b-turns that are stabilized by different hydrogen bondsformed and broken along the trajectory.24

Although computer simulations can reveal the intrinsicconformational features of a peptide as encrypted in itssequence, caution should be taken about the thoroughness ofthe sampling. Therefore, we aim here to complement thesepreliminary computational results with further structuralevidences. We combined different biophysical techniques tostudy the conformational ensemble of NMC in an aqueoussolution, and at ve different 2,2,2-triuoroethanol (TFE)/waterconcentrations (i.e. at 10%, 25%, 40%, 60% and 90% TFE). AsTFE can typically induce a-helicity in polypeptides,25 we com-plemented these data by analyzing the a-helical structure ofNMC bound to SDS micelles, characterizing the NMC–SDSmicelle complex and evaluating independently the structuringand the binding effects on the peptide exibility. These dataconstitute a comprehensive overall picture about the confor-mational preferences of NMC under different environmentsand represents a new structural platform for the future devel-opment of BB2R antagonists.

2 Results2.1 NMC is an intrinsically disordered peptide thatundergoes a progressive coil-helix transition with increasingTFE concentration

It is already known that NMC displays a native random coilconformation in an aqueous solution.21 However, so far no

This journal is © The Royal Society of Chemistry 2015

experimental insight describes how the NMC conformationchanges upon varying the dielectric environment, as mightoccur when it binds to BB2R. Hence, we approached this foldingprocess by studying the effect of increasing TFE concentrationson the NMC conformation. TFE has been widely used as akosmotropic agent to study protein folding.25 Moreover, it has alow effect on the pH of acetate buffered solutions when it isadded up to 90% (DpH � 0.4),26 which makes it highly suitablefor our study.

The CD spectrum of NMC in an aqueous solution indicatesthat the peptide is disordered, as evidenced by the minimumlocated at 197 nm. A slight increase in the TFE percentage, from0 to 10%, scarcely modies the spectrum prole as well as thesecondary structure content. However, the addition of TFE atpercentages higher than 10% markedly enhances the intensityof the region between 190 and 202 nm and reduces that of theband located between 206 and 240 nm, which results in anotable increase of the a-helical content, while decreasing thepercentages of b-strands, turns and random coil regions (Fig. 1Aand Table S1†).

These results indicate that NMC adopts a predominant a-helix conformation upon addition of TFE, but only when thepercentages are higher than 10%.

2.2 The Trp4–Gly7 region holds the higher a-helicitytendency

Although CD spectroscopy provided general information on thefolding behaviour of NMC upon increasing the TFE/water ratio,it does not give any insight at the residue level on how thestructuring process occurs. Therefore, we carried out the NMRstudy of NMC at different TFE/water ratios. The 1H-, 15N- and13C-NMR assignments were obtained at 15 �C for different NMCaqueous solutions containing 0%, 10%, 25%, 40%, 60% or 90%TFE. Chemical shis of the backbone atoms (i.e. N, HN, Ha andCa) were obtained for all residues, except those correspondingto N and HN of Gly1 (that could not be achieved at any TFEpercentage) and Asn2 (which could only be accomplished at 0%,10% and 25% TFE). The Ca resonance of Asn

2 at 40% TFE couldnot be observed either. All the N, H and C atoms at the sidechains were assigned. The chemical assignments of NMC havebeen deposited to BMRB under the accessions codes 25519 (0%TFE), 25520 (10% TFE), 25521 (25% TFE), 25522 (40% TFE),25523 (60% TFE) and 25524 (90% TFE).

Chemical shis corresponding to HN, N, Ha, Hb, Ca and Cb

were used to determine the secondary structural propensity(SSP) scores (see Section 4.5) at different TFE/water ratios(Fig. 1B). The SSP plots obtained suggest that the structures ofNMC at 0% and 10% TFE are similar, except for the residueslocated at the C-terminus, which seem to slightly enhance theirhelicity at 10% TFE. While both the structures seem to berandomly coiled, the increase in the TFE percentage above 10%clearly induces an enlargement of the a-helix content, which ismore prominent in the central region of the peptide (Trp4–Gly7).On the other hand, most of the SSP values obtained at 90% TFEare higher than 0.5, which indicates that the correspondingresidues display a well-dened a-helical structure.

RSC Adv., 2015, 5, 83074–83088 | 83075

Page 4: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Fig. 1 Analysis of the NMC conformation under different experimental conditions. (A) CD spectra of NMC in the presence of differentpercentages of TFE in 10mM acetate buffer (pH 4.0) at 15 �C. (B) SSP values of NMC, as calculated from 1HN,

15N, 1Ha,1Hb,

13Ca and13Cb chemical

shift values at pH 4.0 and 15 �C at 0% ( ), 10% ( ), 25% ( ), 40% ( ), 60% ( ) and 90% ( ) of TFE, and in the presence of 150 mM SDS ( ). (C) NMRsolution ensembles of the ten lowest energy structures of NMC calculated at different TFE/water ratios in 10 mM acetate buffer (pH 4.0) at 15 �C.Backbone atoms are represented as ribbons, while the side chains are shown as atom coloured sticks. Images were generated using the UCSFChimera software.

RSC Advances Paper

These results are in agreement with the CD data and provethat the peptide central region (Trp4–Gly7) holds the higher a-helical tendency.

2.3 The folding pathway of NMC upon increasing the TFEconcentration

The 1H-, 15N- and 13C-NMR assignments were used to calculatethe solution structures of NMC at each TFE/water ratio. Thegeometrical restrains used during the calculations were takenfrom the NOEs intensities and automatically assigned using theCYANA soware27 (Fig. S1†). These assignments provided NMRensembles with all dihedral angles located in favoured andallowed regions and with low backbone RMSD, except for themost disordered NMC structures obtained at 0% and 10% TFE.The ensembles obtained have been deposited in the PDB underthe accessions codes 2n0b (0% TFE), 2n0c (10% TFE), 2n0d(25% TFE), 2n0e (40% TFE), 2n0f (60% TFE) and 2n0g (90%TFE), and all of them satisfy all convergence criteria forsuccessful structure calculations (Table 1).

The NMC ensemble calculated in an aqueous solutiondisplays a native random coil conformation. Nevertheless, thecorresponding backbone RMSD is lower than that expected for afully unstructured ensemble, mostly as a result of the shortrange NOEs detected between central amino acids, which createa backbone architecture resembling a distorted S (Fig. 1C).

83076 | RSC Adv., 2015, 5, 83074–83088

The NMC structure obtained at 10% TFE does not differsignicantly to that found in water, which is in perfectagreement with previously reported CD and SSP data (Fig. 1Aand B). However, while the depicted S-like conformationbetween His3–His8 is still preserved, the C-terminal regionappears to adopt a more extended structure (Fig. 1C).This subtle but remarkable difference was already revealedby the SSP data, but it can also be observed from the over-lapping of the 15N-HSQC spectra of NMC at 0% and 10%TFE, wherein the chemical shi perturbations of Val6, Gly7

and Leu9 are larger than those of Asn2, His3, Trp4 or Ala5

(Fig. 2A).CD and SSP data already suggested that an increase in the

TFE content from 10% to 25% implies a larger structural rear-rangement than that occurring when the TFE percentageincrease from 0% to 10%. Moreover, the overlapping of the 15N-HSQC spectra obtained at 10% and 25% TFE additionally showsthat this rearrangement mainly occurs at the C-terminus, as isevidenced by the chemical shi perturbations displayed by theVal6–Met10 stretch (Fig. 2B). The NMC structure obtained at25% TFE does not exhibit the S-like conformation observed at0% and 10% TFE. In contrast, residues at the C-terminus (up toTrp4) roll up and adopt a helical-like turn, whereas the N-terminal region bends back towards the central residues(Fig. 1C).

This journal is © The Royal Society of Chemistry 2015

Page 5: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Table 1 Structural statistics for the different conformers of NMC obtained at different d3-TFE/water ratios and in the presence of d25-SDSmicelles

0% TFE 10% TFE 25% TFE 40% TFE 60% TFE 90% TFE SDS

Structural computed conformers 18 17 20 20 19 20 20

Restrainsa

Short-range (|i � j| # 1) 47 71 91 121 80 79 99Medium-range (1 < |i � j| < 5) 3 2 23 49 24 35 39Long-range (|i � j| $ 5) 0 0 2 7 3 0 0NOE constrains per restrained residue 7.1 9.0 12.8 19.3 13.1 12.1 13.8Torsion angles restraints 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Restraints statisticsb

Distance violations > 0.0 �A 5 4 3 4 2 3 2Torsion angle violations > 0� 0 0 0 0 0 0 0

Target function value (�A2)Average/best 0.0 0.0 0.02 0.0 0.02 0.0 0.0

Pairwise RMSD of residues 3–8 in�Ac

Backbone N, CA, C0 0.73 � 0.26 0.63 � 0.29 0.08 � 0.07 0.01 � 0.00 0.06 � 0.04 0.25 � 0.1 0.17 � 0.7Heavy atoms 1.87 � 0.51 1.40 � 0.38 0.43 � 0.30 0.05 � 0.01 0.41 � 0.20 1.05 � 0.32 0.57 � 0.35

Ramachandran plotd

Most favoured regions (%) 54.8 47.1 47.9 55.0 75.7 82.9 60.7Additional allowed regions (%) 45.2 52.9 52.1 45.0 23.4 17.1 39.7Generously allowed regions (%) 0.0 0.0 0.0 0.0 0.0 0.0 0.0Disallowed regions (%) 0.0 0.0 0.0 0.0 0.0 0.0 0.0

a Restraint statistics reported for unique, unambiguous assigned NOEs. b Violations are only reported when present in six or more structures.c Coordinate precision is given as the average pair-wise Cartesian coordinate root mean square deviations over the ensemble. d Values obtainedfrom the PROCHECK-NMR analysis68 by using the Protein Structure Validation Server (PSV).69

Paper RSC Advances

The NMC solution structure obtained at 40% TFE shows theformation of a short but well dened a-helix at the C-terminus(Ala5–His8) as a result of an increased compactness of thestructure displayed at 25% TFE (Fig. 1C). On the other hand, theN-terminal region remains unstructured, but adopts a newlyextended conformation. Therefore, the increase in the TFEpercentage from 25% to 40% implies an overall structuralrearrangement that is also evident from the comparison of thecorresponding 15N-HSQC spectra, wherein the entire reso-nances shi (Fig. 2C).

At 60% TFE, the chemical shis of the residues between Val6

and Met10 do not show remarkable differences in comparisonwith those obtained at 40% TFE. However, more noticeable arethe shis of the cross peaks corresponding to His3 and Trp4

(Fig. 2D). These variations result from the preservation of the a-helical structure between Ala5 and His8 already formed at 40%TFE, whereas the residues at the N-terminus fold back alsoadopting a new a-helical conformation (Fig. 1C).

The increase in the TFE percentage mostly resulted in anoverall chemical shi variation of the 1H–15N cross peakstowards high eld, especially in the 1H dimension (Fig. 2),which can be attributed to the lower capacity of TFE to formhydrogen bonds relative to water.28 This was also the case whenthe TFE percentage increased from 60% to 90%, except for His3

and Trp4, whose cross peaks shied towards low eld, sug-gesting a further structural rearrangement at the N-terminus(Fig. 2E). The structure of NMC obtained at 90% TFE shows

This journal is © The Royal Society of Chemistry 2015

an enlargement of the Ala5–His8 a-helical stretch depicted at60% TFE towards Leu9, but also towards Trp4 and His3 (Fig. 1C).Hence, the more compact conformation adopted by theN-terminus when going from 40% to 60% TFE becomes fullya-helical at 90% TFE.

Thus, NMC is a disordered peptide in an aqueous solution,although its central region exhibits a constrained backbonearchitecture resembling a distorted S. Upon increasing TFE/water ratios, the C-terminal region rst stretches to fold backinto an a-helical structure, which also occurs at the N-terminusbut only at higher TFE concentrations.

2.4 NMC binds to SDS micelles

TFE is known to induce a-helicity in most polypeptides througha process that mimics their embedding into regions of lowdielectric constant and high viscosity.25 Hence, the use ofdifferent TFE/water ratios allowed the modelling of a foldingroute, which could mimic that occurring during its interactionwith BB2R.

In addition, we complemented these results by studying thebiophysical characteristics of NMC upon interaction with SDSmicelles, which were chosen as a model system.

We ran diffusion-oriented (DOSY) NMR experiments on asolution containing NMC alone or in the presence of SDSmicelles. The resulting diffusion coefficients (D) were inde-pendent of the NMC concentration, indicating that no self-

RSC Adv., 2015, 5, 83074–83088 | 83077

Page 6: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Fig. 2 Overlapping of the 15N-HSQC spectra of NMC obtained at different d3-TFE/water ratios. (A) 15N-HSQC spectra of NMC in the presence of0% (red) and 10% (purple) TFE. (B) 15N-HSQC spectra of NMC in the presence of 10% (purple) and 25% (green) TFE. (C) 15N-HSQC spectra of NMCin the presence of 25% (green) and 40% (orange) TFE. (D) 15N-HSQC spectra of NMC in the presence of 40% (orange) and 60% (black) TFE. (E)15N-HSQC spectra of NMC in the presence of 60% (black) and 90% (cyan) TFE. The arrows indicate the cross-peak shift from the lower to thehigher TFE percentage.

RSC Advances Paper

aggregation occurred in both the samples. Assuming that theslight viscosity change linked to the presence of SDS micellesequally affects the reference (i.e. acetate and DSS signals) andthe NMC signals, it is clear that NMC reduces its overallmobility in the presence of SDS micelles (Fig. 3A), whichpotentially suggests their binding.

Isothermal titration calorimetry (ITC) was then used tothermodynamically characterize the binding process. Thetitration curve of SDS in acetate buffer resulted in the appear-ance of initial exothermic peaks accounting for the low-temperature energy-favoured demicellization.29 These peaksbecame less exothermic as they approached the midpoint of theinection (critical micelle concentration; cmc),30 to nallybecome endothermic as a result of the micelle dilution effect29

(Fig. S2†).A similar trend was observed when titrations were carried

out on solutions containing NMC (Fig. S2†). However, severaldifferences that could be ascribed to the SDS–NMC interactionwere observed. Difference enthalpograms reveal an initial

83078 | RSC Adv., 2015, 5, 83074–83088

exothermic heat ow that rapidly levels off as the SDS concen-tration increases (Fig. 3B). This effect likely corresponds tospecic and cooperative electrostatic interactions occurringbetween the SDS sulfate group and cationic His.31 The amplitudeof this variation scales linearly with NMC concentration (Fig. 3C)and the slope obtained (aDH � �79 � 5 kJ mol�1 mM�1)indicates that there is a large change in the DH of the systemupon a small change in both protein and surfactant [aDH isproportional to (d2H)/(dnproteindnSDS) and measures the enthalpyof NMC–SDS interaction32].

Next, the enthalpy difference became slightly endothermic,which could be attributed to a conformational rearrangement ofNMC.33 Then, the curves began to deviate exothermically fromthe control curve at a SDS concentration of �3.3 mM, whichcorresponds to the onset for binding of NMC to SDS (criticalaggregation concentration; cac).34 The subsequent exothermicvariation is attributed to the association of SDS and NMC,33

whose saturation is at the inection point and corresponds tocmc, which slightly increases with the NMC concentration (3.9–

This journal is © The Royal Society of Chemistry 2015

Page 7: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Fig. 3 Study of the binding between NMC and SDS. (A) Overlap of the 2D-DOSY spectra of NMC in 10 mM acetate buffer (pH 4.0) alone (blue)and in the presence of 150 mM d25-SDS (red). The spectra were referenced to the acetate and DSS signals. (B) Difference curves arising from thesubtraction of the enthalpy curves obtained during the titration of a 35 mM SDS solution into a solution containing NMC at 10 mM ( ), 20.8 mM ( ),42 mM ( ) or 62.5 mM ( ) from that obtained when titration was carried out in free-NMC solution. All the titrations were carried out at 15 �C usingacetate buffer solution (10 mM) at pH 4.0. This panel only shows the data collected for the addition ongoing from 0 to 2 mM SDS concentration.(C) Difference enthalpy values for the initial exothermic enthalpy depicted in panel B (black) and for the exothermic enthalpy shown in thefollowing panel D (red) as a function of protein concentration. Points are the experimental data, while lines represent the fitting of this data to alinear correlation. (D) The same plot as depicted in panel B but showing the titration region ongoing from 2 to 9 mM SDS concentration.

Paper RSC Advances

4.2 mM). SDS injected beyond this point remains in the micellarform, having fewer NMC molecules to interact and leading tothe nal asymptotic curve (Fig. 3D). The amplitude of this curvelinearly scales with the NMC concentration and the obtainedaDH � �0.92 � 0.09 kJ mol�1 mM�1 (Fig. 3C) proves that thenature of the interaction between SDS micelles and NMC ismuch weaker than an electrostatic one.

The cac and cmc values were additionally used to calculatethe Gibbs free energy change of aggregation (DGmic) and theGibbs free energy change of aggregation in the presence of NMC(DGag) through the application of the charged phase separationand mass-action model.35 The obtained DGmic and DGag valueswere �24 � 2 and �25 � 1 kJ mol�1, respectively, which provesthat the micellar behaviour of SDS and the formation of NMC/SDS mixed micellar junctions are both similarly thermody-namically favoured. Moreover, DHmic and DHag, both <�0.1 kJmol�1 (Fig. 3D), are much smaller that the terms TDSmic orTDSag (�24 kJ mol�1), revealing that both the aggregation ofSDS in the absence and in the presence of NMC is entropydriven.

According to Lindman and Thalberg,35 the free energy todrive 1mol of monomeric SDS into NMC-boundmicelle (DGps¼DGag � DGmic) is indicative of the binding strength of SDS ontoNMC. The obtained DGps z �1 kJ mol�1 reveals that thebinding between NMC–SDS is only slightly thermodynamicallyfavoured in comparison to that occurring between SDS–SDSmolecules.

This journal is © The Royal Society of Chemistry 2015

2.5 Solution structure of NMC bound to SDS micelles

Upon showing that NMC binds to SDS micelles, we wanted toassess this binding effect on the peptide structure. In contrastto bombesin or NMB, NMC was predicted not to adopt amembrane-inserted a-helical conformation due to the reduc-tion of hydrophobic interactions upon replacement of Leu3 byHis3, which would be needed to effectively build the a-helix.20

However, the CD spectrum of NMC bound to SDS micellesindicated the contrary: the presence of micelles induced aredshi in the minimum, while the intensity of the regionsbetween 191 and 193 nm and 200 and 240 nm increased anddecreased, respectively (Fig. 4A). This indicates that NMCincreases its a-helicity upon binding to SDS micelles at a levelsimilar to that shown at 60% TFE content (Table S1†).

NMR was then used to obtain residue level insights on thishelical rearrangement. The comparison of the 15N-HSQCspectra of NMC obtained in the absence and in the presenceof d25-SDS show notable chemical shi perturbations as a resultof the binding. This occurs for all residues except for the crosspeaks of Asn2 (i.e. HN–N and HNd–Nd), which indicates that thechemical environment of Asn2 is only slightly modied uponmicelle interaction (Fig. 4B). All 1H, 15N, and 13C resonances(except those for N and HN of Gly1) were unambiguouslyassigned, deposited to BMRB (25525), and used to calculate theSSP values, which were also comparable to those obtained at60% TFE (Fig. 1B). Hence, CD and SSP data strongly suggest that

RSC Adv., 2015, 5, 83074–83088 | 83079

Page 8: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Fig. 4 Structural study of NMC bound to SDS micelles. (A) CD spectra profiles of NMC collected at 15 �C in the presence and in the absence ofSDS. (B) Overlapping of the 15N-HSQC spectra of NMC recorded at 15 �C in 10 mM acetate buffer (pH 4.0) alone (red) and in the presence of 150mM d25-SDS (purple). (C) NMR ensemble of the ten lowest energy structures of NMC calculated in the presence of SDS micelles. (D) Structuralalignment between His3 and Met10 of the mean NMC structure (averaged over the entire unfolded ensemble by using MOLMOL software)obtained in the presence of 60% TFE (blue) and when it is bound to SDSmicelles (orange). (E) Fluorescence emission spectra of NMC collected at15 �C in 10 mM acetate buffer (pH 4.0) alone and in the presence of SDS micelles.

RSC Advances Paper

the structure of NMC bound to SDS micelles must be similar tothat obtained at 60% TFE.

The geometrical restrains automatically obtained from theNMR assignment and the 1H–1H-NOE intensities (Fig. S1†) wereused to calculate the solution structure of NMC bound to d25-SDS micelles. The ensemble obtained, deposited to the PDBunder the accession code 2n0h, has an excellent Procheck-NMRscore satisfying all convergence criteria for structure calcula-tions (Table 1). The structure reveals that the binding processinduces the formation of an a-helical stretch between Ala5 andLeu9, whereas the N-terminus retains the native random coilconformation (Fig. 4C). In fact, this structure is very similar tothat obtained in 60% TFE, which is ascertained by the lowRMSD value arising from the alignment of the averaged struc-tures of both ensembles (0.52 A for the backbone atoms)(Fig. 4D).

Our results show that NMC binds to SDS micelles through aprocess that implies the formation of an a-helical stretch at theC-terminus, whereas the N-terminal segment remainsdisordered.

2.6 Trp4 is embedded into the SDS micelle

We then wanted to understand the molecular architecture ofthe complex formed between NMC and SDS micelles. Initially,we used uorescence spectroscopy to determine whether Trp4 isinserted into the micelles. The uorescence spectrum of NMCshows an emission maximum at 356 nm, typical of a solventexposed indole group. However, when NMC binds to SDS

83080 | RSC Adv., 2015, 5, 83074–83088

micelles, the uorescence maximum undergoes to a hyp-sochromic shi, from 356 to 345 nm, suggesting the incorpo-ration of the Trp4 side chain into a less polar environment(Fig. 4E). In contrast, the quantum yield of Trp4 also decreasesin agreement with what was found during other peptide–SDSmicelle interactions.36

The extent to which Trp4 was buried into the micelles wasdetermined using the acrylamide quenching experiments.Equal amounts of acrylamide were added to solutions con-taining either free NMC or NMC–SDS micelles complex. Thepresence of micelles decreased the Stern–Volmer constant (Ksv)of NMC threefold (�13 M�1 for the free form vs. �4 M�1 for thecomplex) (Fig. S3†), indicating a high degree of protection of theTrp4 side chain against the solvent. Hence, the Trp4 side chaininserts into the SDS micelles during the binding process.

2.7 Mapping the interactions between NMC and SDSmicelles

The atomic contact map between NMC and SDS micelles wasobtained from overlapping of the 1H–1H-NOESY spectraobtained in the presence of 150 mM d25-SDS and that obtainedin the presence of 50 mM SDS. The change in the SDSconcentration did not alter the NMC structure, as evidenced bythe comparison of both 15N-HSQC spectra (Fig. S4†).

The use of non-deuterated SDS resulted in the appearance ofnew NOEs that were unambiguously assigned to specic inter-molecular NMC–SDS contacts. The side chain of Trp4 is fullyembedded into the SDS micelles since its indole group exhibits

This journal is © The Royal Society of Chemistry 2015

Page 9: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Table 2 NOE connectivities found between NMC and SDS micellesa

NMC CH2 (1)b CH2 (2)

b CH2 (3–11)b

HN-V6 ++HN-L9 +HN-M10 +Ha-W4 ++Ha-A5 ++Ha-V6 ++Ha-L9 +Hb2-W4 +Hb3-W4 ++ ++Hb3-H8 +Hb3-M10 +Hg2-M10 +H31-W4 + + ++H33-W4 + ++Hz2-W4 + ++Hd1-W4 + + ++Hz3-W4 + +++Hd2-H8 ++Hg1-V6 ++Hg2-V6 ++

a The intensities of the NOE signals are divined as: “+++” high intensity;“++” medium intensity; “+” low intensity. b Atoms are numberedarbitrarily according the chemical structure of SDS.

Paper RSC Advances

strong NOEs with different SDS methylene groups. In addition,the HN of Val6, Leu9 and Met10, as well as the Ha of Trp4, Ala5,Val6 and Leu9 display different NOEs with the aliphatic tail ofSDS, proving that the backbone at the C-terminus is alsoinserted into the SDS micelles. NOEs signals connecting theVal6 and Met10 side chains with protons of C1 in SDS were alsofound, indicating that these regions protrude from the hydro-phobic core of the micelle (Fig. 5A and S5†) (Table 2).

When analyzing the structure of NMC upon binding to SDSmicelles (Fig. 4C), it is difficult to understand how this inter-action can occur since an amphipathic-like architecture islacking. For instance, His8 points towards the same face of thehelix than Trp4, while His3, Val6 and Met10 are in the oppositeone.

At pH 4.0 His side chains must be protonated, the pKa of theimidazole protons ranges from 4.9 to 6.6 in micellar media,37

and therefore highly unlikely to penetrate into the micelles.This has been the case for His containing peptides, wheremicelle insertion was only observed when the pH increasedabove the pKa of His.38 However, we unexpectedly detectedunambiguous NOEs between Hb/Hd2 of His8 and the aliphaticmethylene groups of SDS (Fig. 5A and S5†) (Table 2), whichproves that His8 inserts into the micelles even at pH 4.0. Thiscould only occur if the insertion occurs together with His8

deprotonation. In fact, the Hd2 and H32 chemical shi values(highly sensitive to imidazole protonation state) in His8 areshied upeld upon micelle binding in a range comparable tothat observed during the pH-induced deprotonation (i.e., �0.2ppm for Hd2 and�0.4 ppm for H32).38,39 This was not the case for

Fig. 5 Themapping of the interactions observed between NMC and SDSmicelles. (A) Overlapping of the 1H–1H NOESY spectra of NMC obtainedat 15 �C when it was bound to SDS micelles (red) and when it was bound to d25-SDS micelles (blue). The chemical structure of SDS is shownabove the NMR spectra. Carbon atoms are numbered arbitrarily. (B) Model representation of the interaction between NMC and SDS micelles.Aliphatic chains of SDS are coloured in grey, while the corresponding sulfate group is coloured in yellow (sulphur atoms) and red (oxygen atoms).NMC residues are coloured based on the previously determined free energy transfer of each amino acid from water to lipid bilayers (DG ¼ �1.8kcal mol�1, red; DG¼ 1.8 kcal mol�1, blue).40 His3 is coloured taking the DG value determined for protonated His, while His8 is coloured based onthe DG value determined for neutral His.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 83074–83088 | 83081

Page 10: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Table 3 Chemical shift values of imidazolinic protons of His3 and His8 and of the C-terminal and Asn2 amide group atoms when NMC is eitherfree or bound to SDS micelles

Hd2a H31

a Na H1a H2

a

His3 in free NMC 7.08 8.42 — — —His3 in micelle bound NMC 7.28 8.61 — — —His8 in free NMC 7.14 8.43 — — —His8 in micelle-bound NMC 7.02 8.03 — — —Ans2 side chain in free NMC — — 113.3 7.53 6.93Asn2 side chain in micelle-bound NMC — — 112.8 7.52 6.88C-terminal amide in free NMC — — 108.2 7.54 7.15C-terminal amide in micelle-bound NMC — — 105.6 7.20 7.03

a Values are given in ppm.

RSC Advances Paper

His3, since its Hd2 and H32 values only underwent a slightdowneld shi upon micelle binding (Table 3).

Our data indicate that NMC binds to SDS micelles onlythrough the insertion of its C-terminal region, while the N-terminal tail remains out of the micelle (Fig. 5B). This is addi-tionally supported by the upeld chemical shis of the C-terminal amide, whereas the chemical shis corresponding tothe amide side chain of Asn2 remain unaltered (Table 3)(Fig. 4B). The binding process occurs through the energeticallyfavoured insertion of Trp4, Val6, Gly7, Leu9 and Met10 have anegative DG for transfer form water to lipid bilayers,40 whichmust energetically compensate the deprotonation and thefurther insertion of His8 into the micelle (Fig. 5B).

2.8 The a-helical folding and the micelle binding effects onthe NMC exibility

NMC folds upon SDS micelle interaction adopting a structuresimilar to that displayed at 60% TFE. This scenario gives theunique opportunity to study separately the inuence of thefolding and the binding effect on the NMC dynamics. Hence, weacquired the 15N R1, R2, and HET-NOE relaxation data for NMCin an aqueous solution, in the presence of 60% TFE, and underits SDS micelle bound state.

In an aqueous solution, R1 and R2 constants are lower than1.3 s�1, whereas all HET-NOEs are negative, which indicatesthat both the features are typical of intrinsically disorderedpeptides (IDP).41 The addition of 60% TFE increases R1 and R2

constants by �2–3 times, while most of the HET-NOEs becomepositive as a result of the rigidity linked to the a-helix formation.These variations are evenmore pronounced when NMC binds toSDS micelles. In this case, R2 increases by �9 times and most ofthe HET-NOEs display values close to 0.5, thus proving that thebinding additionally constrains the a-helical NMC structure(Fig. 6A–C).

The R1 and R2 constants obtained in an aqueous solution forthe Trp4-indole group are similar to the backbone ones.However, HET-NOE is �0.9 units higher, likely a result of therigidity linked to the contacts of Trp4-Hd1, H33 with Ala5, whichmust reduce the side chain dynamics. The presence of 60% TFEslightly increases the R1 and R2 constants, revealing that themobility of the Trp4 side chain is slightly reduced upon folding.Moreover, the connement of the indole group into the SDS

83082 | RSC Adv., 2015, 5, 83074–83088

micelles enlarges by �4 times the R2, thus proving that thebinding also additionally constrains the mobility of the Trp4

side chain (Fig. 6A–C).The R1/R2 ratios within one standard deviation of the mean

were then used to determine the correlation time (sc) of eachstructural ensemble. Calculations carried out with the r2r1_tmand TENSOR 2.0 soware gave similar values, being 0.8� 0.1 nsin buffer, 2.1 � 0.1 in 60% TFE, and 6.9 � 0.1 ns under itsmicelle bound state. Although the NMC molecular size scarcelychanged in the presence of 60% TFE (Fig. 1C), sc increased by�3 times as a result of the increased viscosity of the TFE/watermixture.42 Furthermore, sc of the NMC–SDS micelle complex is�8 times bigger than that of free NMC, which can be ascribed tothe resulting high molecular weight complex. In addition, thissc value is also �1 ns bigger than that of free SDS micelles,43

which points to the formation of 1 : 1 NMC–SDS micellecomplex.

The backbone R1 and R2 values and the energy-minimizedrepresentative conformers of each NMR-derived solutionstructure were used to estimate the diffusion tensor (Dk/Dt)using the isotropic, axially symmetric and fully anisotropicdiffusion models in the soware Quadratic-Diffusion.44 TheDk/Dt values obtained for NMC were 1.2 in buffer, 0.98 in 60%TFE, and 0.8 under its micelle bound state. Hence, the diffusionmodel that best describes the NMC rotational behaviour underthese experimental conditions is the isotropic one (Dk/Dt < 1.3).

The 15N-relaxation parameters were then analyzed assumingan isotropic rotational diffusion model and according to theLipari–Szabo model-free formalism45 (Tables S2–S4†). The orderparameters (S2; indicative of the amplitude of internal ps–nstimescale motions) of NMC in water have an average value of0.45� 0.12, being within the typical range found in IDPs (Sav

2 �0.3–0.6).46,47 However, these values are not homogeneous alongthe NMC sequence. The Sav

2 of the residues between Trp4 andHis8 is notably higher than that arising from the three N-terminal and the two C-terminal residues (DSav

2 � 0.18)(Fig. 6D). This striking variation clearly proves that centralamino acids, although integrated within the fully disorderedNMC structure, display much slower motions than the terminalones, likely as result of their inter-residual interactions.

The acquisition of the a-helical structure in the presence of60% TFE reduced by�46% the exibility of the central residues

This journal is © The Royal Society of Chemistry 2015

Page 11: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Fig. 6 NMR dynamics of NMC in acetate buffer, in 60% TFE and under its micelle bound state at 14.1 T and 15 �C. (A–C) 15N relaxation dynamicsparameters (A) R1, (B) R2 and (C) heteronuclear NOE plotted as a function of the residue number. Error bars in plots indicate the curve-fit rootmean square deviation of each point. Mean values are plotted as columns. (D) Order parameters (S2) obtained from Liparizi–Szabo model-freeanalysis as a function of the residue number.

Paper RSC Advances

(His3–His8), as evidenced by DSav2 of �0.23. This was not the

case for the still unstructured two C-terminal residues, whosemobility was nearly unaltered upon folding of the Ala5–His8

stretch (DSav2 � 0.03). The insertion of the folded NMC into SDS

micelles additionally reduced the mobility of the central aminoacids (His3–His8) by �20% (DSav

2 � 0.1), whereas that corre-sponding to the C-terminal Leu9 and Met10 was reduced by�90% (DSav

2 � 0.36) as a result of their connement in the SDSmicelle (Fig. 6D).

Our results reveal that NMC in an aqueous solution displayshighly different conformational motions along its sequence.Moreover, the NMC a-helical folding markedly reduces themobility of the central residues, affecting the peptide exibilityto a larger extent than the subsequent structural connementinto SDS micelles.

3 Discussion

NMC modulates different physiological processes such asfeeding or tumor growth mostly through its interaction withBB2R. Although its physiological implications have been knownfor four decades,3,6,11 its structural features, either in its freeform or when bound to BB2R, have not yet been reported.Herein, we combined different biophysical techniques to study

This journal is © The Royal Society of Chemistry 2015

the structural and dynamical preferences of NMC in an aqueoussolution and under its SDS micelle bound state. In addition, wealso analyzed the structure of NMC at different TFE/water ratiosto decode the NMC folding pathway. TFE is known to inducepolypeptide folding in the sense that it deprives peptides ofestablishing hydrogen bonds with water thereby favouringintra-peptide hydrogen bonding.48

Most of the small peptides do not behave as pure randomcoils because their residues usually do not sample all stericallyaccessible regions, but rather exhibit local structural prefer-ences.49 Hence, we used NMR to gain residue insights on thestructural preferences of NMC in an aqueous solution. The 15N-HSQC spectrum only displays nine signals proving that eitherthere is a main conformational state or that the dynamicequilibrium between different conformers is a fast exchangeregime (Fig. 2A). This agrees with our REMD prediction of a lowenergy structure among the different NMC conformers.24 TheNMR ensemble possesses a well-dened backbone architectureresembling a distorted S (Fig. 1C), which is built through shortrange contacts between the central residues. We already pre-dicted these turns using REMD, since �20% of the samplingdisplayed the segment Trp4–Gly7 stabilized by a hydrogen bond.In addition, others turns, such as His3–Trp4, Trp4–Ala5 or Gly7–His8, were only �10% sampled.24 This S-like architecture is also

RSC Adv., 2015, 5, 83074–83088 | 83083

Page 12: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

RSC Advances Paper

adopted in the NMC–Ni2+ complex since two turns are formed,one involving the three rst residues coordinating the metaland the other linking Ala5 to His8.22

NMC folds into a helical structure upon increasing the TFEpercentage as result of the reduction of the dielectric constant,which favours the formation of intramolecular hydrogen bonds.However, the ease of adopting this a-helical structure is not thesame along the entire sequence. The presence of only 10% TFEinduces the stretching of the C-terminal region that folds backinto a helical-like structure only when the TFE percentageincreases to 25%. The C-terminal helical structuring is nallystrengthened at 40% TFE and it does not further change at 60 or90% TFE. The N-terminal segment is more resistant to undergothe coil-helix transition, since it only acquires a helical-likestructure when the TFE content is 60%, becoming fully a-helical at 90% TFE.

In addition, we studied the molecular complex formedbetween NMC and SDS micelles. ITC was used to characterizethe binding process. Initially, it appeared as a highly exothermicevent proportional to the NMC concentration and was attrib-uted to electrostatic interactions.31,33 This showed that NMC isthe limiting reactant in this region of the plot and that thebinding is notably weak. NMC started to bind SDS at a cac of�3.3 mM, consistent with what was observed when usinghydrophobically alkali-soluble emulsion polymers (HASE) (cac� 4 mM).50 The cac value was independent of the NMC or HASEconcentration, although it is concentration-dependent in foldedproteins.33 The binding saturation of NMC occurred at [SDS]� 4mM, which agrees with the cmc of SDS at 15 �C,29 and it did notchange with the NMC concentration, thus differing from whatoccurs in folded proteins.33 SDS added beyond cmc remains inmicellar form and leads to an asymptotic curve that is related tothe NMC hydrophobicity.34,50 Its aDH value is � 86 kJ mol�1

mM�1 lower than that determined for the electrostatic inter-actions between NMC and SDS monomers, which proves thatthe diving force leading to the NMC/SDS micelle binding isweaker than an electrostatic one and likely to be hydrophobic.This aDH is also �15 times lower than that determined for fol-ded proteins,33 which must account for the lower hydropho-bicity of NMC in comparison to larger polypeptides. Theformation of micelles alone or in the presence of NMC wasalways thermodynamically favoured through an entropy-drivenprocess. Moreover, the calculation of DGps (which compares thestability of the interactions between NMC–SDS and SDS–SDS)demonstrated that NMC–SDS was only ��1 kJ mol�1 morefavoured than the SDS–SDS interaction, being weaker than thatdetermined for HASE–SDS interactions (��4 kJ mol�1).50

Next, we calculated the solution structure of NMC under itsmicelle bound state and characterized the architecture of thecomplex. Although it was predicted that the replacement ofLeu3 in NMB by His3 in NMC would hinder the acquisition of ahelical membrane-bound structure,20 we have shown that the C-terminal region of NMC folds into an a-helix upon micelleinsertion, whereas the N-terminal segment remains unstruc-tured. This folding process must be directly related to theenergetically favoured insertion of Trp4, Val6, Gly7, Leu9 andMet10 (residues with a DG < 0 of transfer form water to lipid

83084 | RSC Adv., 2015, 5, 83074–83088

bilayers40) into the non-polar micelle, which was experimentallyobserved through intermolecular NOEs (Fig. 5B). However, thishydrophobic insertion cannot occur without the enclosure ofHis8 into the micelle, which is expected to be fully protonated atpH 4.0, and therefore highly unfavourable. Nevertheless, NMRdata reveal that His8, but not His3, deprotonates during theNMC insertion. These observations enabled us to hypothesizethat the favourable hydrophobic interactions of the residuesnear His8 would energetically compensate for its unfavourabledeprotonation and insertion. This idea is also supported by thesmall DGps value of the NMC–SDS complex in comparison toother peptide–SDS complexes.50

The micelle-bound NMC structure displays Trp4, His8 andLeu9 oriented toward the same face. These residues correlatewith Trp8, His12 and Leu13 in bombesin, which are essential forthe binding to the bombesin family of receptors.51 Hence, it islikely that Trp4, His8 and Leu9 also reorient as NMC approachesits receptor, which further validates the micelle-bound NMCstructure, also within a biologically relevant context.

The ability of NMC to form a short a-helix during its micelleinsertion may result in conformation and/or orientation selec-tive interactions. Hence, the wide spectrum of similar but notidentical biological activities of bombesin-related peptides rai-ses the possibility that uctuations of secondary structure canmodulate their accessibility to different receptors, a mechanismalready proven for neurokins and opioid peptides.52 Hence, wecompleted our structural data by analyzing the dynamics ofNMC in water, in the presence of 60% TFE and under its micellebound state. The fact that NMC at 60% TFE displays a structuresimilar to that adopted when it is embedded into micelles hasallowed us to discriminate the folding and the binding effectson the molecular tumbling and dynamics.

The R1/R2 ratios were used to determine the sc values, whichrepresent the time of the molecule to tumble as a function ofsize, shape and viscosity. The sc of NMC in water is similar tothat found for other peptides of similar size.53 The addition of60% TFE raised sc � 3 times, which is attributed to an increasedsolvent viscosity typical of the TFE/water mixtures42 and not tochanges in the peptide size.53,54 The sc of the NMC–SDS complexwas �1 ns bigger than that of free SDS micelles proving thatNMC slightly reduces the micellar tumbling rate as a result ofthe formation of a 1 : 1 complex; the stoichiometry mainlyobserved in peptide–micelle complexes.46,55

15N relaxation data were used to determine the diffusiontensor of NMC in water, at 60% TFE and under its micelle-bound state. All the Dk/Dt values were <1.3, suggesting anisotropic rotational behaviour. This model has already beenadopted to study the dynamics of other small peptides, either intheir free form56 or under their micelle-bound states.46 NMCdynamics were studied applying the model-free approach,45

which ts the relaxation data to one of the ve models charac-teristic of the complexity of the residue level dynamics. The nineresidues of NMC in water were described by model 2, indicatinginternal motions (se) on ps–ns timescales. This was also the casefor most of the residues of NMC at 60% TFE, except that His3

was tted to model 4, and Leu9/Met10 that were tted to model5, hence suggesting complex internal motions. Five out of the

This journal is © The Royal Society of Chemistry 2015

Page 13: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Paper RSC Advances

nine residues in the NMC–SDS complex were tted to model 1,proving their lack of exibility (Tables S2–S4†).

Residue level mobility was qualitatively compared within thesame NMC structure and across the three different NMCstructures through the analysis of S2. Folded proteins exhibitSav

2 of �0.8, whereas mobile terminal residues display a Sav2 of

�0.6 (ref. 57) similar to IDPs (Sav2 � 0.3–0.6).46,47 The Sav

2 ofNMC in water was within the typical range of IDPs. However, theS2 values notably changed between the central and the terminalresidues, pointing towards a constrained mobility of the centralresidues, a trend also observed at 60% TFE. S2 values alsorevealed that the formation of intramolecular hydrogen bondslinked to the a-helical folding reduces muchmore the backboneexibility than the intermolecular hydrophobic interactionsassociated to the insertion of NMC into the SDS micelles.Hence, the coil-helix transition has a higher impact on thedynamics than the NMC connement.

The 15N-relaxation data acquired at 60% TFE could beaffected by the increase in the viscosity (Dh � 1 cp at 25 �C (ref.42)). However, data comparing the NMR dynamics of theEscherichia coli orthologue of frataxin (CyaY) in water (h � 0.89cp (ref. 42)) and in hen egg white (h � 4 cp (ref. 58)) prove thatviscosity does not affect the CyaY fold nor the HET-NOE values,but notably decreased and increased the R1 and R2 values,respectively.59 The viscosity change scarcely affected the S2

values of CyaY, calculated assuming an axially symmetricdiffusion model (DSav

2 � 0.06) (Fig. S6†). In contrast, R1, R2,HET-NOE and S2 values of NMC were notably enhanced whengoing from pure water to 60% TFE, thus proving that thesechanges are associated with structural alterations rather thanviscosity modications.

4 Experimental4.1 Materials

NMC was purchased from Holzel Diagnostika Handels GmbH.Sodium dodecyl sulfate (SDS, 99%), deuterated sodium dodecylsulfate (d25-SDS, 98%), 3-(trimethylsilyl)-1-propanesulphonicacid (DSS), deuterated (d3-TFE, 99.5%) and non-deuterated2,2,2-triuoroethanol (TFE) were acquired from Sigma-Aldrich. All reagents were used without any further purication.

4.2 Circular dichroism (CD) studies

Circular dichroism (CD) spectra of NMCwere obtained on a Jasco-715 spectropolarimeter equipped with a thermostated cell holdercontrolled by a Jasco Peltier element. Far-UV CD spectra wereacquired from 260 to 190 nm at 15 �C in a 0.1 cm path lengthquartz cuvette at a NMC concentration of 30 mM in 10 mM acetatebuffer (pH 4.0) containing different percentages of TFE (0%, 10%,25%, 40%, 60% and 90%). The CD spectrum of NMC was alsoobtained in a buffered aqueous solution in the presence of 150mM SDS. The scan speed was 50 nm min�1 with a response timeof 1 s and a step resolution of 0.2 nm, whereas 15 scans wereaccumulated. Base-line spectra were subtracted for all spectra.The secondary structure content was derived from the far-UV CDspectra using the BeStSel on-line platform (http://bestsel.elte.hu/).

This journal is © The Royal Society of Chemistry 2015

4.3 Sample preparation for NMR studies

A 5 mM NMC solution was prepared in 10 mM acetate buffer atpH 4.0 containing 10% of D2O and 1.6 mM of DSS (added asinternal reference). This solution was additionally prepared inthe same acetate buffer but in the presence of 10%, 25%, 40%,60% or 90% d3-TFE (to study the NMC folding pathway uponaddition of a structuring solvent), in the presence of 150 mMd25-SDS (to determine the NMC structure bound to SDSmicelles), and in the presence of 50 mM SDS (to map theintermolecular interactions occurring between NMC and SDSmicelles).

4.4 NMR spectroscopy

NMR experiments were carried out at 15 �C on a Bruker AvanceIII spectrometer operating at 14.1 T (600 MHz) and equippedwith a 5 mm 13C, 15N, 1H triple resonance cryoprobe. Forsequence-specic assignments, 1H–1H-TOCSY experiments60

were performed with the MLEV-17 spin-mixing pulse using amixing time of 80 ms. The 1H–1H-NOESY experiments61 wereacquired with a mixing time of 500 ms for the samples con-taining different d3-TFE/water ratios, whereas a mixing time of200 ms was used for the samples containing SDS micelles. 2D15N-HSQC and 13C-HSQC spectra were also acquired at naturalabundance. Spectra were obtained with 2048 data points � 512increments and with a spectral width of 7184 Hz in bothdimensions. Water suppression was achieved by the eld-gradient method with the WATERGATE sequence.62 1H and13C chemical shis were measured relative to the methyl reso-nance of internal DSS at 0 ppm. 15N chemical shis werereferenced indirectly using the 1H, X frequency ratios of thezero-point. The relative diffusion coefficients (D) of NMC weremeasured by the pulse eld gradient spin echo (PGSE) using astandard ledbpgp2s experiment.63 D is a good indicator of themolecular mobility that is relative to the viscosity and to themolecular size. All the spectra were processed using NMRPipe/NMRDraw,64 analyzed by Xeasy/Cara65 and plotted using theSparky soware.66

4.5 NMR structure calculations

The NMR experiments permitted assignment of the resonancefrequencies of 13C, 1H and 15N for NMC. The assignments werethen used to calculate the secondary structure propensity (SSP)scores67 (http://pound.med.utoronto.ca/soware.html). The SSPscore is the weighted average of the chemical shis fromdifferent nuclei in a given residue with the relative weightingreecting the sensitivity of different secondary shis to struc-ture. An SSP score of 1.0 suggests a fully-formed a-helix, a �1.0value indicates a b-strand, whereas a 0 score species a randomcoil conformation. The chemical shi assignments were alsoused to obtain the geometrical restrains resulting from the NOEintensities, which were then used to calculate the NMC solutionstructures. NOE cross-peak assignment was done using theautomated NOE assignment of CYANA,27 a soware that wasfurther used to calculate the NMC ensembles. The standardprotocol was used with seven cycles of combined automated

RSC Adv., 2015, 5, 83074–83088 | 83085

Page 14: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

RSC Advances Paper

NOE assignment and structure calculation for 200 conformersin each cycle, of which the 20 structures with lowest targetfunction value were selected for further minimization andanalysis. PROCHECK-NMR68 was used to analyze the quality ofthe structures through the Protein Structure Validation Server(PSV) (http://psvs-1_4-dev.nesg.org/).69 The MOLMOL soware70

was used for visualization and UCSF Chimera71 was used forstructural representations.

4.6 NMR relaxation measurements15N longitudinal (R1) and transverse (R2) relaxation data, as wellas steady-state 15N HET-NOE data were acquired at 15 �C in anaqueous solution in the presence of 60% TFE and in the pres-ence of 150 mM d25-SDS. In all cases, R1 values were determinedusing a series of 11 experiments with relaxation delays rangingfrom 10 to 2000 ms, while 15N HET-NOE measurements wereperformed by 3 s high power pulse train saturation within a 5 srecycle delay. R1 and 15N HET-NOE data were acquired usingstandard pulse sequences,72 as well as R2 data of the NMCsolution containing d25-SDS, recorded using 11 different relax-ation delays ranging from 8 to 128 ms. R2 measurements of thesolutions prepared in water and in 60% TFE were carried outusing the pulse program recently developed by Yuwen andSkrynnikov with modications, which permits the increase ofthe relaxation delays thus avoiding cryoprobe heating.73 In thesecases, R2 values were measured using 9 relaxation delaysranging from 58 to 691 ms. Recycle delays were 3 s in both R1

and R2 experiments. Thirty two scans in R1 and R2, and 200scans in 15N HET-NOE spectra per t1 experiment were acquired.2048 � 128 complex points were obtained during R1 and R2

experiments, whereas 2048� 164 complex points were acquiredin the 15N HET-NOE experiments.

4.7 NMR relaxation analysis

R1 and R2 relaxation data were tted to a mono-exponentialdecay function, whereas 15N HET-NOE data were obtained asthe ratio of the peaks intensities from the saturated andunsaturated spectra. Relaxation constants and experimentalerrors were calculated using the Protein Dynamics Centersoware (Bruker, Germany). R1 and R2 relaxation constants werethen used to determine the NMC correlation times (sc) at eachexperimental condition using both the r2r1_tm (Palmer'sgroup, http://www.hhmi.umbc.edu/toolkit/analysis/palmer/r2r1_tm.html) and the TENSOR 2.0 soware.74 The magnitudeand orientation of the rotational diffusion tensor was deter-mined from the R1 and R2 relaxation constants and the energy-minimized representative conformers of the NMR-derivedsolution structures using the soware Quadratic-Diffusion.44

TENSOR 2.0 was also used to calculate the generalized orderparameters describing the amplitudes of internal motions (S2).The 15N relaxation constants and the energy-minimized solu-tion structures were analyzed according to the moleculardiffusion derived by Woessner in combination with the Lipari–Szabo model-free analysis of local exibility.45 The amide bondlength was xed at 1.02 �A. Five different models were tested tocharacterize the internal dynamics of the NH groups:75 model 1

83086 | RSC Adv., 2015, 5, 83074–83088

(S2), model 2 (S2, se), model 3 (S2, Rex), model 4 (S2, se, Rex) andmodel 5 (Sf

2, Ss2, se). se is the effective internal correlation time

(describes motions on a timescale > 20 ps), Rex is a chemicalexchange term (describes slow timescale motions on the orderof ms–ms), and Sf

2 and Ss2 are terms that result from splitting

the generalized order parameter into two order parametersreecting slower and faster motions, respectively. The con-dence levels were estimated using 100 Monte Carlo simulationsper run in combination with c2 and F-test criteria.

4.8 Fluorescence measurements

The intrinsic uorescence spectra of NMC were acquired at 15�C on a Varian Cary Eclipse uorimeter equipped with a Peltier-controlled cell holder. Emission spectra were obtained between300 and 400 nm using an excitation wavelength of 280 nm. Theuorescence spectra were collected using a 0.1 mM NMC solu-tion in 10 mM acetate buffer at pH 4.0, either alone or in thepresence of 50 mM of SDS. The change in the solvent accessiblesurface area of Trp4 was measured by the acrylamide quenchingexperiments, which permitted the calculation of the corre-sponding Stern–Volmer constants (Ksv). Aliquots of a 1 Macrylamide aqueous solution were added to the cuvette con-taining either the peptide alone or the peptide/SDS mixture.Emission spectra were obtained as previously described aereach addition of the quencher. The Ksv values were calculatedfrom the following equation,

F0/F ¼ 1 + Ksv[Q] (1)

where F0 is the initial uorescence of the peptide and F is theuorescence intensity following the addition of solublequencher, Q.

4.9 Isothermal titration calorimetry (ITC) measurements

ITC experiments were used to measure the thermodynamicparameters of the NMC–SDS interaction. ITC measurementswere carried out on a Nano-ITC (TA Instruments©) at 15 �C induplicate. In a typical experiment, the sample cell (190 ml) waslled with 10 mM acetate buffer (pH 4.0) in the absence of or inthe presence of different NMC concentrations (i.e. 10 mM, 20.8mM, 42 mM and 62.5 mM) and stirred at 250 rpm, while a 35 mMSDS buffered solution was loaded in the injection syringe. Bothsolutions were degassed before use. SDS solution was titratedinto the sample cell as a sequence of 32 injections (the rst fourinjections of 2 ml and the other injections of 1.5 ml). The timebetween successive injections was 400 s. Raw data corre-sponding to the heating rate (mJ s�1) was integrated to obtainthe observed molar enthalpy change (DH) at each SDS concen-tration. The data corresponding to the titration of SDS intoacetate buffer was subtracted from the data corresponding tothe titration of SDS into a NMC buffered solution before theanalysis of the DH variation. Titration of acetate buffer into aNMC buffered solution resulted in small DH variations thatwere not further considered in the data analysis. All dataacquisition and analysis were performed using the NanoAnalyzesoware.

This journal is © The Royal Society of Chemistry 2015

Page 15: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

Paper RSC Advances

The Gibbs free energy changes of aggregation (DGmic) andaggregation in the presence of NMC (DGag) were calculated fromthe ITC data through the application of the charged phaseseparation and mass-action models (eqn (2)).35

DGmic/ag ¼ (1 + K)RT ln[cac or cmc] (2)

A factor of (1 + K) is needed to calculate the free energy ofionic SDS, where K is the micellar charge fraction with a value of0.85.76 The enthalpy change and free energy changes were alsoused to calculate the entropy changes of micellization (DSmic)and aggregation in the presence of NMC (DSag).

5 Conclusions

The results reported in this study allow a rationalization of thestructural determinants of NMC in an aqueous solution anddescribe the different folding steps that potentially occur duringits interaction with its natural partners. The description of theNMC–SDS micelle complex revealed that the hydrophobicbinding must energetically compensate for the deprotonationand the micelle insertion of His8, which results in a low DG ofinteraction. The fact that the solution structure of NMC at 60%TFE mimics the micelle-bound structure provides an excellentopportunity to discriminate the folding and the binding effectson the NMC mobility. We observed that the folding constrainsthe NMC mobility twice as much as its micelle connement.Our data contribute to the general understanding of themechanisms involving peptide–lipid interactions and providefundamental to insights into the structure–exibility relation-ship of bombesin-like peptides. These novel insights into thebinding-induced folding mechanism of NMC also constitute anew molecular platform for the future design of antagonists ofthe bombesin family of receptors.

Acknowledgements

The authors are grateful for the excellent technical assistancefrom the Serveis Cienticotecnics at the UIB, especially to DrJoan Cifre for his generous assistance with ITC measurements.K. P. is supported by a FWO Pegasus long-term post-doctoralfellowship (1218713). P. S. is supported by a post-doctoralfellowship (PD/009/2013) from the Conselleria d'Educacio,Cultura i Universitats of the Balearic Government and theEuropean Social Fund through the ESF Operational Program ofthe Balearic Islands 2013–2017. This study is supported by theproject AAEE26/2014 from the Conselleria d'Educacio, Cultura iUniversitats of the Balearic Government and FEDER.

References

1 H. Ohki-Hamazaki, M. Iwabuchi and F. Maekawa, Int. J. Dev.Biol., 2005, 49, 293.

2 A. Anastasia, V. Erspamer and M. Bucci, Experientia, 1971,27, 166.

3 N. Minamino, K. Kangawa and H. Matsuo, Biochem. Biophys.Res. Commun., 1984, 119, 14.

This journal is © The Royal Society of Chemistry 2015

4 R. T. Jensen, J. F. Battey, E. R. Spindel and R. V. Benya,Pharmacol. Rev., 2008, 60, 1.

5 T. H. Moran, T. W. Moody, A. M. Hostetler, P. H. Robinson,M. Goldrich and P. R. McHugh, Peptides, 1988, 9, 643;E. E. Ladenheim, R. T. Jensen, S. A. Mantey andT. H. Moran, Brain Res., 1993, 593, 168.

6 A. Guglietta, C. L. Strunk, B. J. Irons and L. H. Lazarus,Peptides, 1985, 6(suppl. 3), 75.

7 T. Reynolds and R. D. Pinnock, Brain Res., 1997, 750, 67.8 T. Tachibana, K. Matsuda, H. Sawa, A. Mikami, H. Ueda andM. A. Cline, Gen. Comp. Endocrinol., 2010, 169, 144;T. Tachibana, K. Matsuda, S. I. Khan, H. Ueda andM. A. Cline, Comp. Biochem. Physiol., Part A: Mol. Integr.Physiol., 2010, 156, 394.

9 E. M. Fekete, E. E. Bagi, K. Toth and L. Lenard, Brain Res.Bull., 2007, 71, 386.

10 H. Zhao, S. Matsuda, S. Thanthan, S. Yannaing andH. Kuwayama, Peptides, 2012, 37, 194.

11 F. Cuttitta, D. N. Carney, J. Mulshine, T. W. Moody,J. Fedorko, A. Fischler and J. D. Minna, Nature, 1985, 316,823.

12 D. B. Cornelio, R. Roesler and G. Schwartsmann, Ann. Oncol.,2007, 18, 1457.

13 W. Guojun, G. Wei, O. Kedong, H. Yi, X. Yanfei, C. Qingmei,Z. Yankai, W. Jie, F. Hao, L. Taiming, L. Jingjing andC. Rongyue, Endocr.-Relat. Cancer, 2008, 15, 149.

14 M. Sioud and A. Mobergslien, Biochem. Pharmacol., 2012, 84,1123.

15 I. D. Majumdar and H. C. Weber, Curr. Opin. Endocrinol.,Diabetes Obes., 2012, 19, 3.

16 G. Schwartsmann, L. P. DiLeone, M. Horowitz,D. Schunemann, A. Cancella, A. S. Pereira, M. Richter,F. Souza, A. B. da Rocha, F. H. Souza, P. Pohlmann andG. de Nucci, Invest. New Drugs, 2006, 24, 403.

17 D. C. Damin, F. S. Santos, R. Heck, M. A. Rosito, L. Meurer,L. M. Kliemann, R. Roesler and G. Schwartsmann, Dig. Dis.Sci., 2010, 55, 2203.

18 B. A. Nock, R. Cescato, E. Ketani, B. Waser, J. C. Reubi andT. Maina, J. Med. Chem., 2012, 55, 8364.

19 T. Maina, B. A. Nock and S. Mather, Canc. Imag., 2006, 6, 153.20 D. Erne and R. Schwyzer, Biochemistry, 1987, 26, 6316.21 E. Polverini, P. Neyroz, P. Fariselli, R. Casadio and

L. Masotti, Biochem. Biophys. Res. Commun., 1995, 214,663.

22 G. Gasmi, A. Singer, J. Forman-Kay and B. Sarkar, J. Pept.Res., 1997, 49, 500.

23 C. Harford and S. Bibudhendra, Biochem. Biophys. Res.Commun., 1995, 209, 877.

24 P. Sharma, P. Singh, K. Bisetty, A. Rodriguez and J. J. Perez, J.Pept. Sci., 2011, 17, 174.

25 M. Buck, Q. Rev. Biophys., 1998, 31, 297; E. R. G. Main andS. E. Jackson, Nat. Struct. Biol., 1999, 6, 831.

26 S. Espinosa, E. Bosch, M. Roses and K. Valko, J. Chromatogr.A, 2002, 954, 77.

27 J. G. Jee and P. Guntert, J. Struct. Funct. Genomics, 2003, 4,179; T. Herrmann, P. Guntert and K. Wuthrich, J. Mol.Biol., 2002, 319, 209.

RSC Adv., 2015, 5, 83074–83088 | 83087

Page 16: Conformational ensembles of neuromedin C reveal a progressive coil-helix transition … · 2015-09-30 · Conformational ensembles of neuromedin C reveal a progressive coil-helix

RSC Advances Paper

28 G. Merutka, H. J. Dyson and P. E. Wright, J. Biomol. NMR,1995, 5, 14.

29 A. Wangsakan, P. Chinachoti and D. J. McClements, FoodHydrocolloids, 2006, 20, 461.

30 Z. Kiraly and I. Dekany, J. Colloid Interface Sci., 2001, 242,214.

31 M. D. Lad, V. M. Ledger, B. Briggs, R. J. Green andR. A. Frazier, Langmuir, 2003, 19, 5098.

32 C. Trandum, P. Westh, K. Jorgensen and O. G. Mouritsen,Biophys. J., 2000, 78, 2486.

33 M. M. Nielsen, K. K. Andersen, P. Westh and D. E. Otzen,Biophys. J., 2007, 92, 3674.

34 G. Wang and G. Olofsson, J. Phys. Chem., 1995, 99, 5588.35 B. Lindman and K. Thalberg, in Interactions of surfactants

with Polymers and Proteins, ed. E. D. Goddard and K. P.Ananthapadmanabhan, CRC Press, Boca Raton, FL, 1993,p. 203.

36 K. Polewski, Biochim. Biophys. Acta, 2000, 1523, 56;E. F. Haney, F. Lau and H. J. Vogel, Biochim. Biophys. Acta,2007, 1768, 2355.

37 B. Bechinger, J. Mol. Biol., 1996, 263, 768.38 J. Georgescu, V. H. Munhoz and B. Bechinger, Biophys. J.,

2010, 99, 2507; C. Aisenbrey, R. Kinder, E. Goormaghtigh,J. M. Ruysschaert and B. Bechinger, J. Biol. Chem., 2006,281, 7708.

39 D. H. Sachs, A. N. Schechter and J. S. Cohen, J. Biol. Chem.,1971, 246, 6576.

40 W. C. Wimley and S. H. White, Nat. Struct. Biol., 1996, 3, 842.41 A. T. Alexandrescu, K. Rathgeb-Szabo, K. Rumpel, W. Jahnke,

T. Schulthess and R. A. Kammerer, Protein Sci., 1998, 7, 389;T. Mittag, S. Orlicky, W. Y. Choy, X. Tang, H. Lin, F. Sicheri,L. E. Kay, M. Tyers and J. D. Forman-Kay, Proc. Natl. Acad. Sci.U. S. A., 2008, 105, 17772; R. Kiss, D. Kovacs, P. Tompa andA. Perczel, Biochemistry, 2008, 47, 6936.

42 G. Gente and C. la Mesa, J. Solution Chem., 2000, 29, 1159.43 K. Lycknert, T. Rundlof and G. Widmalm, J. Phys. Chem. B,

2002, 106, 5275.44 L. K. Lee, M. Rance, W. J. Chazin and A. G. Palmer, J. Biomol.

NMR, 1997, 9, 287.45 G. Lipari and A. Szabo, J. Am. Chem. Soc., 1982, 104, 4546.46 S. M. Patil, S. Xu, S. R. Sheic and A. T. Alexandrescu, J. Biol.

Chem., 2009, 284, 11982.47 J. B. Duvignaud, C. Savard, R. Fromentin, N. Majeau,

D. Leclerc and S. M. Gagne, Biochem. Biophys. Res.Commun., 2009, 378, 27; A. V. Buevich, U. P. Shinde,M. Inouye and J. Baum, J. Biomol. NMR, 2001, 20, 233;T. S. Ulmer and A. Bax, J. Biol. Chem., 2005, 280, 43179.

48 D. Roccatano, G. Colombo, M. Fioroni and A. E. Mark, Proc.Natl. Acad. Sci. U. S. A., 2002, 99, 12179.

49 R. Schweitzer-Stenner, Mol. BioSyst., 2012, 8, 122.50 W. P. Seng, K. C. Tam, R. D. Jenkins and D. R. Bassett,

Macromolecules, 2000, 33, 1727.51 D. C. Horwell, W. Howson, D. Naylor, S. Osborne,

R. D. Pinnock, G. S. Ratcliffe and N. Suman-Chauhan, Int.J. Pept. Protein Res., 1996, 48, 522.

83088 | RSC Adv., 2015, 5, 83074–83088

52 R. Schwyzer, Biochemistry, 1986, 25, 6335.53 M. Ramirez-Alvarado, V. A. Daragan, L. Serrano and

K. H. Mayo, Protein Sci., 1998, 7, 720; D. Idiyatullin,A. Krushelnitsky, I. Nesmelova, F. Blanco, V. A. Daragan,L. Serrano and K. H. Mayo, Protein Sci., 2000, 9, 2118.

54 P. Kaczka, M. Winiewska, I. Zhukov, B. Rempoła,K. Bolewska, T. Łozinski, A. Ejchart, A. Poznanska,K. L. Wierzchowski and J. Poznanski, Eur. Biophys. J., 2014,43, 581.

55 A. Perdih, A. R. Choudhury, S. Zuperl, E. Sikorska, I. Zhukov,T. Solmajer and M. Novic, PLoS One, 2012, 7, e38967.

56 M. Wang, M. Prorok and F. J. Castellino, Biophys. J., 2010, 99,302.

57 J. L. Goodman, M. D. Pagel and M. J. Stone, J. Mol. Biol.,2000, 295, 963.

58 E. Robinson-Lang and C. Rha, Int. J. Food Sci. Technol., 1982,17, 595.

59 G. Martorell, M. Adrover, G. Kelly, P. A. Temussi andA. Pastore, Proteins, 2011, 79, 1408.

60 A. Bax and D. G. Davis, J. Magn. Reson., 1985, 65, 355.61 A. Kumar, R. R. Ernst and K. Wuthrich, Biochem. Biophys.

Res. Commun., 1980, 95, 1.62 M. Piotto, V. Saudek and V. Sklenar, J. Biomol. NMR, 1992, 2,

661.63 D. Wu, A. Chen and C. S. Johnson Jr, J. Magn. Reson., Ser. A,

1995, 115, 260.64 F. Delaglio, S. Grzesiek, G. W. Vuister, G. Zhu, J. Pfeifer and

A. Bax, J. Biomol. NMR, 1995, 6, 277.65 C. Bartels, T. H. Xia, M. Billeter, P. Guntert and K. Wuthrich,

J. Biomol. NMR, 1995, 6, 1.66 T. D. Goddard and D. G. Kneller, SPARKY 3, University of

California, San Francisco, CA, USA, 2008.67 J. A. Marsh, V. K. Singh, Z. Jia and J. D. Forman-Kay, Protein

Sci., 2006, 15, 2795.68 R. A. Laskowski, J. A. Rullmann, M. W. MacArthur,

R. Kaptein and J. M. Thornton, J. Biomol. NMR, 1996, 8, 447.69 A. Bhattacharya, R. Tejero and G. T. Montelione, Proteins,

2007, 66, 778.70 R. Koradi, M. Billeter and K. Wuthrich, J. Mol. Graphics,

1996, 14, 51.71 E. F. Pettersen, T. D. Goddard, C. C. Huang, G. S. Couch,

D. M. Greenblatt, E. C. Meng and T. E. Ferrin, J. Comput.Chem., 2004, 25, 1605.

72 N. A. Farrow, R. Muhandiram, A. U. Singer, S. M. Pascal,C. M. Kay, G. Gish, S. E. Shoelson, T. Pawson,J. D. Forman-Kay and L. E. Kay, Biochemistry, 1994, 33, 5984.

73 T. Yuwen and N. R. Skrynnikov, J. Magn. Reson., 2014, 241,155.

74 P. Dosset, J. C. Hus, M. Blackledge and D. Marion, J. Biomol.NMR, 2000, 16, 23.

75 V. A. Jarymowycz and M. J. Stone, Chem. Rev., 2006, 106,1624.

76 J. R. Lu, A. Marrocco, T. J. Su, R. K. Thomas and J. Penfold, J.Colloid Interface Sci., 1996, 178, 614.

This journal is © The Royal Society of Chemistry 2015