research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL
BIOLOGY
ISSN: 2059-7983

Structures of designed armadillo-repeat proteins show propagation of inter-repeat interface effects

CROSSMARK_Color_square_no_text.svg

aDepartment of Biochemistry, University of Zürich, Winterthurerstrasse 190, 8057 Zürich, Switzerland
*Correspondence e-mail: plueckthun@bioc.uzh.ch, mittl@bioc.uzh.ch

Edited by Z. S. Derewenda, University of Virginia, USA (Received 18 September 2015; accepted 1 December 2015)

The armadillo repeat serves as a scaffold for the development of modular peptide-recognition modules. In order to develop such a system, three crystal structures of designed armadillo-repeat proteins with third-generation N-caps (YIII-type), four or five internal repeats (M-type) and second-generation C-caps (AII-type) were determined at 1.8 Å (His-YIIIM4AII), 2.0 Å (His-YIIIM5AII) and 1.95 Å (YIIIM5AII) resolution and compared with those of variants with third-generation C-caps. All constructs are full consensus designs in which the internal repeats have exactly the same sequence, and hence identical conformations of the internal repeats are expected. The N-cap and internal repeats M1 to M3 are indeed extremely similar, but the comparison reveals structural differences in internal repeats M4 and M5 and the C-cap. These differences are caused by long-range effects of the C-cap, contacting molecules in the crystal, and the intrinsic design of the repeat. Unfortunately, the rigid-body movement of the C-terminal part impairs the regular arrangement of internal repeats that forms the putative peptide-binding site. The second-generation C-cap improves the packing of buried residues and thereby the stability of the protein. These considerations are useful for future improvements of an armadillo-repeat-based peptide-recognition system.

1. Introduction

For the design of artificial peptide-binding modules, scaffolds with modular architectures are highly suitable. In particular, the armadillo repeat reveals structural properties that facilitate the design of peptide-binding modules on a rational basis (Andrade et al., 2001[Andrade, M. A., Perez-Iratxeta, C. & Ponting, C. P. (2001). J. Struct. Biol. 134, 117-131.]; Kajander et al., 2006[Kajander, T., Cortajarena, A. L. & Regan, L. (2006). Methods Mol. Biol. 340, 151-170.]; Boersma & Plückthun, 2011[Boersma, Y. L. & Plückthun, A. (2011). Curr. Opin. Biotechnol. 22, 849-857.]; Reichen, Hansen et al., 2014[Reichen, C., Hansen, S. & Plückthun, A. (2014). J. Struct. Biol. 185, 147-162.]). In natural armadillo-repeat proteins such as importin-α and β-catenin, each repeat comprises three α-helices that are assembled in a triangular spiral staircase arrangement. All repeats are fused into a single protein with an elongated hydrophobic core (Figs. 1[link]a and 1[link]b). They recognize their target peptides in extended β-sheet conformations with very regular binding topologies. The main chain of the peptide is bound in an antiparallel direction by conserved asparagine residues on the concave side of the armadillo-repeat protein (Huber et al., 1997[Huber, A. H., Nelson, W. J. & Weis, W. I. (1997). Cell, 90, 871-882.]; Conti et al., 1998[Conti, E., Uy, M., Leighton, L., Blobel, G. & Kuriyan, J. (1998). Cell, 94, 193-204.]; Kobe, 1999[Kobe, B. (1999). Nature Struct. Biol. 6, 388-397.]; Fontes et al., 2003[Fontes, M. R., Teh, T., Toth, G., John, A., Pavo, I., Jans, D. A. & Kobe, B. (2003). Biochem. J. 375, 339-349.]). Differences exist in side-chain preferences because the importin-α and β-catenin subfamilies recognize peptides with positively and negatively charged side chains, respectively (Conti & Kuriyan, 2000[Conti, E. & Kuriyan, J. (2000). Structure, 8, 329-338.]; Ishiyama et al., 2010[Ishiyama, N., Lee, S.-H., Liu, S., Li, G.-Y., Smith, M. J., Reichardt, L. F. & Ikura, M. (2010). Cell, 141, 117-128.]; Poy et al., 2001[Poy, F., Lepourcelet, M., Shivdasani, R. A. & Eck, M. J. (2001). Nature Struct. Biol. 8, 1053-1057.]).

[Figure 1]
Figure 1
(a) The triangular spiral staircase arrangement of helices indicative of the armadillo repeat. (b) Ribbon diagram of His-YIIIM5AII. The His6 tag, YIII-type, M-type and AII-type repeats are shown in magenta, green, blue and orange, respectively. (c) Sequence alignment of N-caps with and without a 3C protease cleavage site (the scissile bond is indicated by a grey arrow), internal repeats and C-­caps. Residues distinguishing different repeat versions are highlighted in red.

It is the goal of this protein-engineering project to develop a stable full-consensus armadillo-repeat scaffold. Internal repeats with identical sequences are characteristic of full-consensus designs. Later, the internal repeats will be functional­ized to recognize different amino-acid side chains. The modularity of the design, which is imposed by the repetitive architecture, should enable us to generate artificial peptide-binding proteins with properties that are precisely tailored according to the length and sequence of the target peptide (Parmeggiani et al., 2008[Parmeggiani, F., Pellarin, R., Larsen, A. P., Varadamsetty, G., Stumpp, M. T., Zerbe, O., Caflisch, A. & Plückthun, A. (2008). J. Mol. Biol. 376, 1282-1304.]; Reichen, Hansen et al., 2014[Reichen, C., Hansen, S. & Plückthun, A. (2014). J. Struct. Biol. 185, 147-162.]). Binding proteins with sequence-specific recognition properties for unstructured peptides should be of great interest in research and development because peptide–protein inter­actions represent 15–40% of all cellular interactions (Petsalaki et al., 2009[Petsalaki, E., Stark, A., García-Urdiales, E. & Russell, R. B. (2009). PLoS Comput. Biol. 5, e1000335.]). Here, many protein–protein interaction scaffolds are unsuitable because they recognize targets based on surface-complementarity properties and thus require a folded counterpart. Conversely, many recognition modules used in intracellular signalling recognize only very short sequences and thus have very low affinity (Pawson & Nash, 2003[Pawson, T. & Nash, P. (2003). Science, 300, 445-452.]). Indeed, specific peptide–protein interaction strategies are required to cope with the intrinsic flexibility of unstructured peptides (London et al., 2010[London, N., Movshovitz-Attias, D. & Schueler-Furman, O. (2010). Structure, 18, 188-199.]).

The first designed armadillo-repeat proteins (dArmRPs) were constructed using a consensus design approach based on 133 and 110 sequences from the importin-α and β-catenin subfamilies, respectively, in combination with structure-aided modifications of the hydrophobic core (Parmeggiani et al., 2008[Parmeggiani, F., Pellarin, R., Larsen, A. P., Varadamsetty, G., Stumpp, M. T., Zerbe, O., Caflisch, A. & Plückthun, A. (2008). J. Mol. Biol. 376, 1282-1304.]). They possess the overall composition YzMnAz, where Y, M and A represent the N-terminal, internal and C-terminal repeats, respectively. The subscripts denote the generation (version) count (z) and the number of internal repeats (n) in roman and arabic numbers, respectively. Since structure-based techniques are vital for this design approach, several structures of proteins from the YIIMnAII and YIIIMnAIII series have been determined. Initial crystal structures of dArmRPs with second-generation N- and C-caps revealed domain-swapped N-caps, suggesting that the YII-type N-cap was unstable in solution. To improve the thermodynamic stability of the caps, nine and six mutations were inserted in the N- and C-caps, respectively. These modifications had complementary effects on the thermodynamic stability of the proteins. Introduction of the third-generation N-cap (YIII-type) increased the melting temperature by 4.5°C, but the modifications in the C-cap (AIII-type) decreased it by 5.5°C (Madhurantakam et al., 2012[Madhurantakam, C., Varadamsetty, G., Grütter, M. G., Plückthun, A. & Mittl, P. R. (2012). Protein Sci. 21, 1015-1028.]). The thermodynamic stabilities of dArmRPs that have so far been designed in this project have been summarized in Reichen, Hansen et al. (2014[Reichen, C., Hansen, S. & Plückthun, A. (2014). J. Struct. Biol. 185, 147-162.]).

Although the initial crystal structures of His-YIIIM3AIII and His-YIIIM3AII revealed monomeric proteins (Reichen, Madhurantakam et al., 2014[Reichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572-1583.]), later studies on YIIIM5AIII (third-generation N-cap and C-cap) without an N-terminal His tag revealed domain-swapped N-caps and C-caps in the presence of calcium ions. However, domain swapping of YIIIM5AIII was not observed either in the absence of calcium ions or in the presence of the His tag because the His tag prevented the unfolding of the N-cap by binding to the neighbouring His-YIIIM5AIII molecule (Reichen, Madhurantakam et al., 2014[Reichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572-1583.]). To investigate the impact of the cap design on the structural parameters of dArmRPs, particularly in the absence of the His tag, we investigated the crystal structures of the more stable dArmRPs with third-generation N-caps and second-generation C-caps.

2. Materials and methods

2.1. Cloning, protein expression and purification

dArmRPs with cleavable and non­cleavable N-terminal His6 tags have been expressed and purified as described by Reichen, Madhurantakam et al. (2014[Reichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572-1583.]) with the following modifications: vectors pPank and p148_3C were used for the expression of proteins with and without a cleavable His6 tag, respectively. The initial designs had non­cleavable His6 tags, but in order to facilitate the elimination of the purification tag, a 3C protease cleavage site was inserted between the His6 tag and the N-terminus of the N-cap. The amino-acid sequences of the internal and capping repeats are depicted in Fig. 1[link](c).

The proteins comprise third-generation N-caps, second-generation C-caps and four or five internal repeats. All three constructs are full-consensus designs, with internal repeats derived from the [\overline{\rm M}]-type internal repeat described in Alfarano et al. (2012[Alfarano, P., Varadamsetty, G., Ewald, C., Parmeggiani, F., Pellarin, R., Zerbe, O., Plückthun, A. & Caflisch, A. (2012). Protein Sci. 21, 1298-1314.]). His-YIIIM4AII and YIIIM5AII contain M′-type internal repeats, whereas His-YIIIM5AII contains the M′′-type. In the M′′-type the aspartic acid at position 1, which was introduced to mimic a potential arginine-binding pocket, was mutated back to the consensus asparagine residue (for all sequences, see Fig. 1[link]c). To improve readability, we refer to M-type internal repeats throughout the text.

2.2. Crystallization and structure determination

A Phoenix crystallization robot (Art Robbins Instruments) was used to set up sitting-drop vapour-diffusion experiments in 96-well Corning plates (Corning, New York, USA). Initial crystallization conditions were identified by sparse-matrix screens from Hampton Research (California) and Molecular Dimensions (Suffolk, England), and were later refined by grid screens in which the pH and the precipitant concentrations were varied simultaneously. To confirm the expected peptide-binding site, (KR)5 peptide was added to YIIIM5AII in a 1.5:1 molar ratio prior to crystallization. (KR)5 peptide was used for this experiment because the designed molecular surface of YIIIM5AII resembled the most conserved importin-α peptide-binding site, which recognizes with its core repeats (major and minor binding sites) positive dipeptide motifs composed of lysine and arginine residues. The rationale for this experiment is discussed in Reichen, Hansen et al. (2014[Reichen, C., Hansen, S. & Plückthun, A. (2014). J. Struct. Biol. 185, 147-162.]). Protein solutions were mixed with reservoir solutions in 1:1, 1:2 or 2:1 ratios (200–300 nl final volume) and the mixtures were equilibrated against 50 µl reservoir solution at 4°C. Reservoir conditions are summarized in Table 1[link]. After washing, the crystals in reservoir solutions supplemented with glycerol were flash-cooled in liquid nitrogen.

Table 1
Data and refinement statistics

Values in parentheses are for the highest resolution shell.

Structure His-YIIIM4AII His-YIIIM5AII YIIIM5AII
PDB code 4v3q 4v3o 4v3r
Data statistics
 Crystallization condition 25% PEG 2000 MME, 0.2 M calcium acetate, 0.1 M sodium acetate pH 5.5 15% PEG 4000, 0.2 M calcium acetate, 0.1 M sodium acetate pH 5.5 30% PEG 4000, 0.2 M magnesium chloride, 0.1 M Tris–HCl pH 8.5
 Space group P32 P41 I4
 No. of molecules in asymmetric unit 4 4 2
 Unit-cell parameters
  a = b (Å) 96.50 102.59 129.91
  c (Å) 96.34 111.11 70.20
  α = β (°) 90 90 90
  γ (°) 120 90 90
 Resolution (Å) 1.80 (1.91–1.80) 2.00 (2.11–2.00) 1.95 (2.06–1.95)
Rmerge (%) 9.1 (88.6) 10.0 (75.0) 8.8 (47.6)
 No. of observations 744192 (120009) 601165 (75390) 107908 (15424)
 No. of unique reflections 93024 (15191) 76669 (10657) 41831 (6089)
 〈I/σ(I)〉 12.6 (2.3) 12.2 (2.6) 7.5 (2.0)
 Completeness (%) 100 (100) 94.3 (94.3) 98.1 (98.2)
Refinement statistics
 Resolution (Å) 96.34–1.80 111.11–2.00 91.86–1.95
Rcryst (%) 18.9 16.8 17.3
Rfree (%) 23.6 22.4 22.9
B factors
  Wilson B2) 27.0 28.7 21.4
  Mean B value (Å2) 35.5 35.2 23.2
 R.m.s.d. from ideal values
  Bond lengths (Å) 0.018 0.017 0.017
  Bond angles (°) 1.83 1.71 1.72
 Total No. of atoms
  Protein 7487 8618 4243
  Water 654 767 419
  Metal ions 16 12 3
  Ligands 2 1 0
 Ramachandran plot
  Favoured (%) 98.81 99.02 100.00
  Allowed (%) 1.19 0.98 0.00
  Outliers (%) 0.00 0.00 0.00

Data were collected on beamlines X06SA and X06DA at the Swiss Light Source (Paul Scherrer Institute, Villigen, Switzerland) using a Pilatus detector (Dectris, Baden, Switzerland) and a wavelength of 1.0 Å. Diffraction data were processed using MOSFLM (Leslie, 1992[Leslie, A. G. W. (1992). Jnt CCP4/ESF-EACBM Newsl. Protein Crystallogr. 26.]) and SCALA (Evans, 2006[Evans, P. (2006). Acta Cryst. D62, 72-82.]). Structures were solved by molecular replacement using Phaser (McCoy et al., 2007[McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658-674.]) together with the following search models. For His-YIIIM4AII we used the structure of YIIIM3AII (PDB entry 4db6; Madhurantakam et al., 2012[Madhurantakam, C., Varadamsetty, G., Grütter, M. G., Plückthun, A. & Mittl, P. R. (2012). Protein Sci. 21, 1015-1028.]). The refined His-YIIIM4AII structure was then used to solve the His-YIIIM5AII and finally the YIIIM5AII structures. The structures were refined using PHENIX (Adams et al., 2010[Adams, P. D. et al. (2010). Acta Cryst. D66, 213-221.]) and REFMAC5 (Murshudov et al., 2011[Murshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355-367.]). For manual model building we used the program Coot (Emsley & Cowtan, 2004[Emsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126-2132.]). The decrease in Rfree suggested the use of different refinement strategies for His-YIIIM4AII and His-YIIIM5AII. His-YIIIM4AII was refined without NCS restraints, whereas tight NCS restraints between chains A/B and C/D were applied for the refinement of His-YIIIM5AII. Figures were prepared using PyMOL (DeLano, 2002[DeLano, W. L. (2002). PyMOL. https://www.pymol.org.]). Metal ions were placed manually into strong difference electron-density peaks, taking into account the coordination geometry and the composition of the crystallization buffer. Calcium ions were validated by inspecting the anomalous difference map calculated with phases from the final structure. Water molecules were placed into well defined difference electron-density peaks at hydrogen-bond distance from the protein. No (KR)5 peptide was identified in the final electron-density map of YIIIM5AII. Side-chain conformations were assigned according to the rotamer library of Dunbrack & Cohen (1997[Dunbrack, R. L. Jr & Cohen, F. E. (1997). Protein Sci. 6, 1661-1681.]) as implemented in Coot.

3. Results and discussion

3.1. Structures of His-YIIIM4AII and His-YIIIM5AII

The crystal structures of His-YIIIM4AII and His-YIIIM5AII were refined at 1.8 and 2.0 Å resolution, respectively. In both cases the asymmetric units contain tetramers with 222 point symmetry and very similar topologies. The quaternary structures of His-YIIIM4AII and His-YIIIM5AII are governed by calcium ions that connect neighbouring chains in a zipper-like manner and the His6 tag that binds to the supposed peptide-binding site, albeit in different orientations (see below).

The His-YIIIM4AII tetramer contains 16 calcium ions. Five calcium ions connect two His-YIIIM4AII chains in an antiparallel orientation (Fig. 2[link]a). Considering the large size of this interface (average interface area of 1163 Å2) there are relatively few direct hydrogen bonds, and most interactions are made via calcium ions in the loops between helices H2 and H3. The coordination number of each calcium ion in His-YIIIM4AII is seven, which agrees very well with the statistical analysis of calcium-coordination geometry in protein and small-molecule complexes. Typically, the coordination number of calcium varies between six and eight, with an average length for coordination bonds of between 2.35 and 2.45 Å (Katz et al., 1996[Katz, A. K., Glusker, J. P., Beebe, S. A. & Bock, C. W. (1996). J. Am. Chem. Soc. 118, 5752-5763.]). In His-YIIIM4AII the coordination geometry of calcium differs among ions that are bound to internal or capping repeats.

[Figure 2]
Figure 2
(a) The subunits of the His-YIIIM4AII tetramer are connected via calcium ions. Two chains are sketched as ribbons and coloured as described in Fig. 1[link](b). Two chains are shown as grey surfaces. Calcium ions are indicated as spheres. Calcium ions binding only to internal repeats are in yellow, those involving the N-cap in light blue and those at the twofold axis in salmon. (b) Calcium-binding site between internal repeats viewed along the axial direction (from the direction of Pro23# O, which was omitted for clarity). Residues from different chains are shown as sticks with blue and salmon C atoms. Calcium ions and water molecules are depicted as grey and red spheres with reduced atomic radii, respectively. Polar interactions in the pentagonal plane involving the calcium ion are shown as dashed lines in orange. Additional interactions are in yellow. The 2FoFc and anomalous difference electron-density maps are contoured at 1.3σ (light blue) and 4σ (green), respectively. (c) Calcium-binding site involving the N-cap. (d) Calcium-binding site at the twofold axis. Colour coding is as described for (b).

Ca2+ ions that bind to internal repeats are contacted by Pro23 O and Glu25 OE1 from two symmetry-related chains (superscripts indicate the position in the repeat as indicated in Fig. 1[link]c) and three water molecules (Fig. 2[link]b). Here, Glu25 contributes one coordination bond (Glu25 OE1–Ca distance 2.5 Å). In contrast, calcium ions that bind between an internal repeat and the N-cap are contacted by two water molecules, two O atoms from Glu25 (Glu25 OE1–Ca distance of 2.5 Å and Glu25 OE2–Ca distance of 3.0 Å), Gln25 OE1 and Pro23 O (Fig. 2[link]c). Thus, the replacement of glutamic acid at position 25 by glutamine in the N-cap displaces one water molecule and allows Glu25 to serve as a bidentate ligand. This observation agrees well with previous data on the statistics of calcium binding, in which it was shown that bidentate binding of carboxylate groups to calcium is particularly prevalent if the coordination number is greater than six (Katz et al., 1996[Katz, A. K., Glusker, J. P., Beebe, S. A. & Bock, C. W. (1996). J. Am. Chem. Soc. 118, 5752-5763.]). In contrast to many natural calcium-binding sites, where all coordination bonds are approximately equal in length, the His-YIIIM4AII calcium-binding sites are distorted. In His-YIIIM4AII the axial calcium–ligand distances are shorter than the equatorial distances (axial distances 2.1–2.2 Å; equatorial distances 2.4–3.0 Å) and the Glu25 OE2–Ca bonding distances differ significantly from the average coordination bond length. The second coordination bond of Glu25 is longer, because the carboxylate group is rotated away from the Ca2+ ion. In contrast to natural calcium-binding sites that have evolved over time, the His-YIIIM4AII calcium-binding sites are distorted because they are artificial and are therefore less perfect. Besides these zipper-like Ca2+ ions bound to the N-termini of H3 helices, four well defined calcium ions additionally bind close to the twofold axes. These Ca2+ ions also show pentagonal-bipyramidal coordination spheres involving the Ser40 carbonyl O atom, the Glu2 side chain and five water molecules (Fig. 2[link]d). Furthermore, there are two weakly occupied calcium-binding sites involved in crystal contacts.

The His-YIIIM4AII tetramer is further stabilized by interactions between the N-terminal His6 tag and the supposed peptide-binding site. This contact is formed by His6, which interacts with Glu30 and Trp33 (Glu156 and Trp159) from the third internal repeat, and His8, which interacts with Trp33 (Trp201) from the fourth internal repeat and Glu33 (Glu243) from the C-cap (Fig. 3[link]a). Besides the salt bridges between histidine and glutamic acid side chains, the aromatic stacking interaction between His6 and Trp33 might contribute significant binding energy because the spatial orientation of side chains seen here is frequently found in protein structures (cluster 4 of His–Trp interactions in the atlas of protein side-chain interactions; Singh & Thornton, 1992[Singh, J. & Thornton, J. M. (1992). Atlas of Protein Side-Chain Interactions. Oxford: IRL Press.]). Since all four chains of His-YIIIM4AII are very similar (r.m.s.d. of 0.28 Å for residues 14–246) these interactions are equivalent in all four subunits of the crystallographic tetramer.

[Figure 3]
Figure 3
Interface between internal repeats M3 and M4 in chain C of His-YIIIM4AII (a) and His-YIIIM5AII (b). The dArmRPs are shown in blue and grey and the His6 tag with salmon C atoms. (c) Superposition based on the N-cap and internal repeats M1–M3 of His-YIIIM5AII chain A (dark blue), His-YIIIM5AII chain C (light blue) and YIIIM5AII (orange). Residues at the M3–M4 interface are labelled. (d) Cα trace of YIIIM5AII coloured in green (N-cap), blue (internal repeats) and orange (C-cap). The Leu32, Trp33 and Thr34 side chains are shown as sticks in blue, grey and green, respectively. Hydrogen bonds and general distances are shown as orange and grey dotted lines, respectively. Distances and conformations of Leu32 side chains are indicated (tg+, trans/gauche+; gt, gauche/trans).

In contrast to this, the crystallographic tetramer of His-YIIIM5AII is less symmetric. Here, chains A/B and C/D are pairwise identical (r.m.s.d. of 0.05 Å), whereas an r.m.s.d. of 0.85 Å for the comparison between pairs (e.g. chain A with D) suggests substantial differences. Furthermore, His-YIIIM5AII chains A/B are more similar to His-YIIIM4AII (r.m.s.d. of 0.72 Å for the superposition of residues 14–210 on the equivalent residues from His-YIIIM4AII) than chains C/D (r.m.s.d. of 1.17 Å). These differences are caused by different contacts within the tetramer. In chains C/D of His-YIIIM5AII the side chain of Glu198 interacts with His8 from chain D/C (Fig. 3[link]b), whereas in chains A/B the side chain of Glu198 intercalates between internal repeats 3 and 4 and forms a hydrogen bond to the side chain of Gln68 from chains B/A (similar to the interaction shown in Fig. 3[link]a for His-YIIIM4AII). As a consequence of this asymmetry, two calcium ions close to the twofold axis, which are present in all four chains of His-YIIIM4AII (Fig. 2[link]d), are only present in His-YIIIM5AII chains A/B and are absent from chains C/D.

3.2. Structure of YIIIM5AII without His tag

The structure of YIIIM5AII without His tag was determined in the absence of calcium ions and refined at 1.95 Å resolution. This structure is most similar to chains C/D of His-YIIIM5AII (r.m.s.d.s of 1.14 and 0.60 Å for Cα atoms of residues 14–288 of chains A/B and C/D, respectively). These differences are a consequence of a rigid-body movement of the C-terminal repeats (internal repeats M4 and M5 and the C-cap). A superposition of YIIIM5AII on His-YIIIM5AII based on the N-cap and internal repeats M1–M3 (residues 14–168) reveals that this part is very similar in all chains. However, in this superposition the C-terminal repeats of YIIIM5AII match nicely with the C-terminal repeats of His-YIIIM5AII chains C/D, but they are shifted towards M3 in chains A/B (1.4 Å shift of Trp201 CA towards Leu158 CA). This movement can be described as an 8° rotation around an axis that runs parallel to the stacking direction of the C-terminal part and is probably a consequence of different side-chain conformations of Leu158, Trp159, Glu198 and Trp201 at the interface between M3 and M4 (Fig. 3[link]c). The structures of His-YIIIM4AII and YIIIM5AII represent extreme cases that are most different. In His-YIIIM5AII these differences are combined into a single structure. His-YIIIM5AII chains A/B and C/D represent the conformations seen in His-YIIIM4AII (all chains) and YIIIM5AII (all chains), respectively. Similar structural plasticity has been observed previously for the comparison of β-catenin crystallized in two different crystal forms. For β-catenin the C-terminal repeats were rotated 11.5° around an axis that runs approximately parallel to the axis of the superhelix (Huber et al., 1997[Huber, A. H., Nelson, W. J. & Weis, W. I. (1997). Cell, 90, 871-882.]).

Thus, dArmRPs with second-generation C-caps and third-generation N-caps possess substantial flexibility, particularly for the side chains of Glu30, Leu32 and Trp33 (equivalent to Glu156, Leu158 and Trp159 in repeat M3 and Glu198, Leu200 and Trp201 in repeat M4). Experimental structural data for importin-α in complex with nuclear localization sequence (NLS) peptides (Conti et al., 1998[Conti, E., Uy, M., Leighton, L., Blobel, G. & Kuriyan, J. (1998). Cell, 94, 193-204.]) and modelling studies on dArmRPs–peptide complexes (Reichen, Hansen et al., 2014[Reichen, C., Hansen, S. & Plückthun, A. (2014). J. Struct. Biol. 185, 147-162.]) indicate that the superhelix parameters and the conformations of Glu30 and Trp33, which also participate in binding the His6 tag as outlined above, are important structural features for proper binding of the target peptide. In a first approximation, the curvature of the peptide-binding site can be described by the distances of Cα atoms at position 33. In the major NLS peptide-binding site of importin-α (PDB entry 1bk6; Conti et al., 1998[Conti, E., Uy, M., Leighton, L., Blobel, G. & Kuriyan, J. (1998). Cell, 94, 193-204.]) the distance between Cα atoms of adjacent Trp33 residues (e.g. Trp153, Trp195 and Trp237 in repeats 1–3) varies between 8.6 and 8.8 Å. In YIIIM5AII the average distance between these atoms is 8.82 ± 0.39 Å. However, in YIIIM5AII the spread between Trp33 Cα-atom distances is extremely large, with the largest distance observed between repeats M3 and M4 (the distances between Trp159 CA and Trp201 CA are 9.42 Å in chain A and 9.43 Å in chain B). This distance is probably too large for binding the target peptide in the desired conformation and this mismatch is located almost at the centre of the putative peptide-binding site. It is possible that this mismatch is responsible for the fact that the (KR)5 peptide was not observed in the electron-density map, although it was present during crystallization. Interestingly, the rigid-body movement of the C-terminal part as seen in His-YIIIM4AII (all chains) and His-YIIIM5AII (chains C/D) brings this value to the other extreme. Here, the distance of Trp33 Cα atoms between repeats M3 and M4 is 8.14 ± 0.06 Å, which might be too short for proper binding.

Although YIIIM5AII is considered to be a full consensus design regarding the sequence of internal repeats, the internal repeats are not identical in terms of structure. These differences can be exerted either by different lattice contacts (Figs. 3[link]a and 3[link]b) or by improper design, which prevents the internal repeats from obtaining a unique conformation throughout the protein. Different distances between adjacent repeats are probably the result of both effects. In particular, the side-chain conformations of buried residues in the hydrophobic core, such as Ile27, Leu32, Thr34, Gly36 and Ile38, mediate the contacts between adjacent repeats. In the structure of YIIIM5AII the side-chain conformations of Thr34, Ile38 and of course Gly36 are invariant. The side chain of Thr34 cross-links internal repeats by forming hydrogen bonds to the main-chain carbonyl groups of Leu32 and Glu30 from adjacent repeats. The side chain of Ile27 adopts mainly gauche/trans conformations, whereas the side chain of Leu32 alternates between trans/gauche+ and gauche/trans (Fig. 3[link]d).

This alternation suggests that a uniform conformation of Leu32 is impossible. In the interface between M3 and M4 of YIIIM5AII, where we observe the largest distance between Trp33 Cα atoms, Leu158 CD1 (Leu32 in M3) and Thr202 OG1 (Thr34 in M4) are at van der Waals distances (3.86 and 3.97 Å in chains A and B) because the Leu158 side chain adopts a trans/gauche+ conformation. Therefore, steric hindrance between Leu158 and Thr202 might be responsible for increasing the distance between Trp33 Cα atoms and for the failure to obtain a dArmRP–peptide complex structure. To adopt a Trp33 Cα distance which is similar to the values seen in the major binding site of importin-α, Thr202 OG1 would have to move closer to Leu158, but this approach would require a gauche/trans conformation of the Leu158 side chain. Of course, surface-exposed side chains (such as Trp33 and Glu30) also adopt different rotamers, but it can be assumed that these differences affect inter-repeat distances to a minor extent because the environments of surface-exposed side chains are usually less densely packed than the environments of buried side chains. However, some side-chain conformations of buried and surface-exposed residues are coupled. For example, the conformation of Trp33 is linked to the conformation of Leu32 in the preceding repeat. In repeats M1 and M3 Leu32 adopts trans/gauche+ conformations and Trp33 in repeats M2 and M4 is trans/+90°, whereas in repeats M2 and M4 Leu32 is gauche/trans and Trp33 adopts trans/−105° conformations in repeats M3 and M5 (Fig. 3[link]d). Only Trp243 in chain B deviates from this general observation.

3.3. Comparison of dArmRPs with second-generation and third-generation C-caps

The crystal structures of YIIIM5AIII with and without a His6 tag and third-generation C-caps have been published recently (Reichen, Madhurantakam et al., 2014[Reichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572-1583.]). YIIIM5AIII without a His6 tag but crystallized in the presence of calcium revealed domain-swapped N- and C-caps. Since YIIIM5AII without a His6 tag and a second-generation C-cap did not crystallize in the presence of calcium, it remains unclear whether the redesign of the C-cap was responsible for calcium-induced domain swapping.

Interestingly, YIIIM5AIII also shows an extended distance between Trp33 Cα atoms of internal repeats M3 and M4 (distance between Trp159 CA and Trp201 CA of 8.86 Å), a short distance between Thr202 OG1 and Leu158 CD2 of 3.91 Å and no electron density for the (KR)5 peptide, although it was present during crystallization (Reichen, Madhurantakam et al., 2014[Reichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572-1583.]). On the other hand, Leu158 shows the gauche/trans side-chain conformation, which is trans/gauche+ in YIIIM5AII, probably because Glu198 forms an additional hydrogen bond to Gln155 O (Fig. 4[link]a).

[Figure 4]
Figure 4
Superposition of YIIIM5AIII (third-generation C-cap; PDB entry 4plq; salmon) on YIIIM5AII (second-generation C-cap; blue). (a) Residues at the M3–M4 interface. General distances and hydrogen bonds are shown as grey and orange dotted lines, respectively. Distance values refer to YIIIM5AIII. The superposition is based on all Cα atoms from M3. (b) Residues at the M5–C-cap interface. Numbers refer to positions in the repeat (Fig. 1[link]c), with subscripts indicating the internal repeat number or the C-cap. Side chains of all residues that differ between YII and YIII and some residues from the hydrophobic core are shown in stick representation. The superposition is based on all Cα atoms from M5.

For dArmRPs with three internal repeats it was shown that the redesign of the C-cap (from AII to AIII) decreases the melting temperature by 5.5°C (Madhurantakam et al., 2012[Madhurantakam, C., Varadamsetty, G., Grütter, M. G., Plückthun, A. & Mittl, P. R. (2012). Protein Sci. 21, 1015-1028.]), and a domain-swapped C-cap was observed for YIIIM5AIII (Reichen, Madhurantakam et al., 2014[Reichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572-1583.]). Both observations suggest that YIIIM5AIII is less stable than YIIIM5AII. A superposition of YIIIM5AIII (PDB entry 4plq) and YIIIM5AII based on the last internal repeat suggests that this destabilization might be owing to subtle rearrangements in the hydrophobic core between internal repeats M4 and M5 and the C-cap. Three out of six mutations that were introduced at the C-cap are solvent-exposed and do not seem to have a significant effect on the structure. However, Lys15→Ala, His22→Ser and Leu38→Ile mutations cause a gentle rearrangement of the C-cap (Fig. 4[link]b). This rearrangement has implications for the packing of side chains in the hydrophobic core. In the more stable YIIIM5AII structure the side chains of Leu16, Leu20 and Val7 adopt a uniform distribution of side-chain rotamers in all repeats. Val7 adopts a trans conformation. Leu16 and Leu20 are always gauche/trans. In YIIIM5AIII this crystal-like arrangement is perturbed by the C-cap. In YIIIM5AIII the side chains of Leu16, Leu20 and Val7 adopt the same conformations as in YIIIM5AII only in the N-terminal part, whereas in the C-terminal part their conformations are clearly different. For Leu32 the situation is inverted. In YIIIM5AIII the rotamer distribution of Leu32 is uniform, whereas in YIIIM5AII alternating Leu32 conformations are observed (Fig. 3[link]d). Uniform distributions of rotamers are frequently observed in polypeptides with very high thermodynamic stabilities, such as amyloid fibrils (Nelson et al., 2005[Nelson, R., Sawaya, M. R., Balbirnie, M., Madsen, A. O., Riekel, C., Grothe, R. & Eisenberg, D. (2005). Nature (London), 435, 773-778.]) and β-helix proteins (Schulz & Ficner, 2011[Schulz, E. C. & Ficner, R. (2011). Curr. Opin. Struct. Biol. 21, 232-239.]). Therefore, it can be assumed that the uniform distribution of side-chain rotamers is related to the stability of dArmRPs and vice versa. On the other hand, the deterioration of uniformity, as caused by the third-generation C-cap, is linked to destabilization of the protein.

In conclusion, this detailed investigation of the different versions of dArmRPs has shown that small differences in packing between repeats, notably between internal repeats and the caps, can make the protein susceptible to perturbations caused by crystal contacts and ions used in crystallization, indicating a lack of rigidity. This leads to a surprising long-range effect of changes in the C-cap and helps to explain the astonishing observation that a full-consensus design does not necessarily generate a unique repeat conformation. Although the internal repeats are chemically absolutely identical, their conformations lack uniformity. The current analysis suggests that future improvements of an armadillo-repeat-based peptide-recognition system will have to take three considerations into account. (i) In particular, the deletion of the His tag seems to be crucial for liberating the presumed peptide-binding site. (ii) The second-generation C-cap presented here seems to be superior to the third-generation C-cap, which was initially believed to be more advanced. (iii) The choice of amino acids at the inter-repeat interface, particularly at positions 27, 32 and 34, should be reconsidered because the side chains at these positions show substantial conformational heterogeneity.

Supporting information


Footnotes

Present address: Molecular Partners AG, Wagistrasse 14, 8952 Zürich-Schlieren, Switzerland.

§Present address: Department of Biotechnology, TERI University, 10 Institutional Area, Vasant Kunj, New Delhi 110 070, India.

Acknowledgements

Beat Blatmann and Celine Stutz at the high-throughput crystallization centre and the staff of beamlines X06SA and X06DA at the Swiss Light Source are acknowledged for their skillful technical support. This work was financially supported by a Swiss National Science foundation grant to AP and PREM (Sinergia S-41105-06-01). SH is the recipient of Forschungskredit from the University of Zurich, grant No. FK-13-028.

References

First citationAdams, P. D. et al. (2010). Acta Cryst. D66, 213–221.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAlfarano, P., Varadamsetty, G., Ewald, C., Parmeggiani, F., Pellarin, R., Zerbe, O., Plückthun, A. & Caflisch, A. (2012). Protein Sci. 21, 1298–1314.  CrossRef CAS PubMed Google Scholar
First citationAndrade, M. A., Perez-Iratxeta, C. & Ponting, C. P. (2001). J. Struct. Biol. 134, 117–131.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBoersma, Y. L. & Plückthun, A. (2011). Curr. Opin. Biotechnol. 22, 849–857.  CrossRef CAS PubMed Google Scholar
First citationConti, E. & Kuriyan, J. (2000). Structure, 8, 329–338.  Web of Science CrossRef PubMed CAS Google Scholar
First citationConti, E., Uy, M., Leighton, L., Blobel, G. & Kuriyan, J. (1998). Cell, 94, 193–204.  Web of Science CrossRef CAS PubMed Google Scholar
First citationDeLano, W. L. (2002). PyMOL. https://www.pymol.orgGoogle Scholar
First citationDunbrack, R. L. Jr & Cohen, F. E. (1997). Protein Sci. 6, 1661–1681.  CrossRef CAS PubMed Google Scholar
First citationEmsley, P. & Cowtan, K. (2004). Acta Cryst. D60, 2126–2132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEvans, P. (2006). Acta Cryst. D62, 72–82.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFontes, M. R., Teh, T., Toth, G., John, A., Pavo, I., Jans, D. A. & Kobe, B. (2003). Biochem. J. 375, 339–349.  Web of Science CrossRef PubMed CAS Google Scholar
First citationHuber, A. H., Nelson, W. J. & Weis, W. I. (1997). Cell, 90, 871–882.  CrossRef CAS PubMed Web of Science Google Scholar
First citationIshiyama, N., Lee, S.-H., Liu, S., Li, G.-Y., Smith, M. J., Reichardt, L. F. & Ikura, M. (2010). Cell, 141, 117–128.  CrossRef CAS PubMed Google Scholar
First citationKajander, T., Cortajarena, A. L. & Regan, L. (2006). Methods Mol. Biol. 340, 151–170.  PubMed CAS Google Scholar
First citationKatz, A. K., Glusker, J. P., Beebe, S. A. & Bock, C. W. (1996). J. Am. Chem. Soc. 118, 5752–5763.  CrossRef CAS Web of Science Google Scholar
First citationKobe, B. (1999). Nature Struct. Biol. 6, 388–397.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLeslie, A. G. W. (1992). Jnt CCP4/ESF–EACBM Newsl. Protein Crystallogr. 26Google Scholar
First citationLondon, N., Movshovitz-Attias, D. & Schueler-Furman, O. (2010). Structure, 18, 188–199.  CrossRef CAS PubMed Google Scholar
First citationMadhurantakam, C., Varadamsetty, G., Grütter, M. G., Plückthun, A. & Mittl, P. R. (2012). Protein Sci. 21, 1015–1028.  CrossRef CAS PubMed Google Scholar
First citationMcCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C. & Read, R. J. (2007). J. Appl. Cryst. 40, 658–674.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMurshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355–367.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationNelson, R., Sawaya, M. R., Balbirnie, M., Madsen, A. O., Riekel, C., Grothe, R. & Eisenberg, D. (2005). Nature (London), 435, 773–778.  Web of Science CrossRef PubMed CAS Google Scholar
First citationParmeggiani, F., Pellarin, R., Larsen, A. P., Varadamsetty, G., Stumpp, M. T., Zerbe, O., Caflisch, A. & Plückthun, A. (2008). J. Mol. Biol. 376, 1282–1304.  CrossRef PubMed CAS Google Scholar
First citationPawson, T. & Nash, P. (2003). Science, 300, 445–452.  Web of Science CrossRef PubMed CAS Google Scholar
First citationPetsalaki, E., Stark, A., García-Urdiales, E. & Russell, R. B. (2009). PLoS Comput. Biol. 5, e1000335.  CrossRef PubMed Google Scholar
First citationPoy, F., Lepourcelet, M., Shivdasani, R. A. & Eck, M. J. (2001). Nature Struct. Biol. 8, 1053–1057.  Web of Science CrossRef PubMed CAS Google Scholar
First citationReichen, C., Hansen, S. & Plückthun, A. (2014). J. Struct. Biol. 185, 147–162.  CrossRef CAS PubMed Google Scholar
First citationReichen, C., Madhurantakam, C., Plückthun, A. & Mittl, P. R. (2014). Protein Sci. 23, 1572–1583.  CrossRef CAS PubMed Google Scholar
First citationSchulz, E. C. & Ficner, R. (2011). Curr. Opin. Struct. Biol. 21, 232–239.  CrossRef CAS PubMed Google Scholar
First citationSingh, J. & Thornton, J. M. (1992). Atlas of Protein Side-Chain Interactions. Oxford: IRL Press.  Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoSTRUCTURAL
BIOLOGY
ISSN: 2059-7983
Follow Acta Cryst. D
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds