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

IUCrJ
Volume 6| Part 2| March 2019| Pages 238-247
ISSN: 2052-2525

Calixarene-mediated assembly of a small antifungal protein

CROSSMARK_Color_square_no_text.svg

aSchool of Chemistry, National University of Ireland, University Road, Galway, Ireland, and bInstitute of Chemistry, Centre of Arts, Humanities and Sciences, University of Debrecen, Hungary
*Correspondence e-mail: peter.crowley@nuigalway.ie

Edited by L. R. MacGillivray, University of Iowa, USA (Received 17 October 2018; accepted 8 January 2019; online 5 February 2019)

Synthetic macrocycles such as calixarenes and cucurbiturils are increasingly applied as mediators of protein assembly and crystallization. The macrocycle can facilitate assembly by providing a surface on which two or more proteins bind simultaneously. This work explores the capacity of the sulfonato-calix[n]arene (sclxn) series to effect crystallization of PAF, a small, cationic antifungal protein. Co-crystallization with sclx4, sclx6 or sclx8 led to high-resolution crystal structures. In the absence of sclxn, diffraction-quality crystals of PAF were not obtained. Interestingly, all three sclxn were bound to a similar patch on PAF. The largest and most flexible variant, sclx8, yielded a dimer of PAF. Complex formation was evident in solution via NMR and ITC experiments, showing more pronounced effects with increasing macrocycle size. In agreement with the crystal structure, the ITC data suggested that sclx8 acts as a bidentate ligand. The contributions of calixarene size/conformation to protein recognition and assembly are discussed. Finally, it is suggested that the conserved binding site for anionic calixarenes implicates this region of PAF in membrane binding, which is a prerequisite for antifungal activity.

1. Introduction

There is growing interest in the use of synthetic macrocycles as mediators of protein assembly (van Dun et al., 2017[Dun, S. van, Ottmann, C., Milroy, L. G. & Brunsveld, L. (2017). J. Am. Chem. Soc. 139, 13960-13968.]). The special case of protein crystallization (McPherson et al., 2011[McPherson, A., Nguyen, C., Cudney, R. & Larson, S. B. (2011). Cryst. Growth Des. 11, 1469-1474.]) has benefitted from `molecular glues' such as calixarenes and cucurbiturils that promote crystal packing (Guagnini et al., 2018[Guagnini, F., Antonik, P. M., Rennie, M. L., O'Byrne, P., Khan, A. R., Pinalli, R., Dalcanale, E. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 7126-7130.]; Rennie et al., 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]). The sulfonato-calix[n]arenes (sclxn, Fig. 1[link]) are highly water-soluble, anionic macrocycles with diverse applications in the biosciences (Baldini et al., 2017[Baldini, L., Casnati, A. & Sansone, F. (2017). Comprehensive Supramolecular Chemistry II, Vol. 4, edited by J. Atwood, G. W. Gokel & L. Barbour, pp. 371-408. Amsterdam: Elsevier.]; Giuliani et al., 2015[Giuliani, M., Morbioli, I., Sansone, F. & Casnati, A. (2015). Chem. Commun. 51, 14140-14159.]; Guo & Liu, 2014[Guo, D. S. & Liu, Y. (2014). Acc. Chem. Res. 47, 1925-1934.]). The hydrophobic core and the anionic rim of the calixarene can facilitate protein recognition, in particular, via the entrapment of arginine or lysine side chains (McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.], 2014[McGovern, R. E., McCarthy, A. A. & Crowley, P. B. (2014). Chem. Commun. 50, 10412-10415.], 2015[McGovern, R. E., Snarr, B. D., Lyons, J. A., McFarlane, J., Whiting, A. L., Paci, I., Hof, F. & Crowley, P. B. (2015). Chem. Sci. 6, 442-449.]; Wang et al., 2016[Wang, K., Guo, D. S., Zhao, M. Y. & Liu, Y. (2016). Chem. Eur. J. 22, 1475-1483.]; Mallon et al., 2016[Mallon, M., Dutt, S., Schrader, T. & Crowley, P. B. (2016). ChemBioChem, 17, 774-783.]; Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.], 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]; Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]). Consequently, sclx4 and related compounds readily co-crystallize with the highly cationic cytochrome c and lysozyme (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]; Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.], 2014[McGovern, R. E., McCarthy, A. A. & Crowley, P. B. (2014). Chem. Commun. 50, 10412-10415.], 2015[McGovern, R. E., Snarr, B. D., Lyons, J. A., McFarlane, J., Whiting, A. L., Paci, I., Hof, F. & Crowley, P. B. (2015). Chem. Sci. 6, 442-449.]). With increasing calixarene size there tends to be more pronounced effects; for example, phosphonato-calix[6]arene (pclx6) has an approximately tenfold increase in affinity (with respect to sclx4) and prompts dimerization of cytochrome c in solution (Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.]). Sulfonato-calix[8]arene (sclx8) on the other hand induces a tetramer of cytochrome c (Rennie et al., 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]). Furthermore, while calix[4]arene is locked in a bowl conformation, the larger calixarenes are flexible and adopt various conformations (Fig. 1[link]) (Atwood et al., 1992[Atwood, J. L., Clark, D. L., Juneja, R. K., Orr, G. W., Robinson, K. D. & Vincent, R. L. (1992). J. Am. Chem. Soc. 114, 7558-7559.]; Dalgarno et al., 2003[Dalgarno, S. J., Hardie, M. J., Makha, M. & Raston, C. L. (2003). Chem. Eur. J. 9, 2834-2839.]; Gutsche & Bauer, 1985[Gutsche, C. D. & Bauer, L. J. (1985). J. Am. Chem. Soc. 107, 6052-6059.]; Liu et al., 2009[Liu, Y., Liao, W., Bi, Y., Wang, M., Wu, Z., Wang, X., Su, Z. & Zhang, H. (2009). CrystEngComm, 11, 1803-1806.]; Perret et al., 2006[Perret, F., Bonnard, V., Danylyuk, O., Suwinska, K. & Coleman, A. W. (2006). New J. Chem. 30, 987-990.]; Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.], 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]; Smith et al., 2006[Smith, C. B., Barbour, L. J., Makha, M., Raston, C. L. & Sobolev, A. N. (2006). New J. Chem. 30, 991-996.]). Accordingly, sclx8 can bind to cytochrome c either via an extended `pleated loop' or a collapsed `double cone' conformation, as shown using X-ray crystallography (Rennie et al., 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]).

[Figure 1]
Figure 1
Sulfonato-calix[n]arenes. (a) Molecular structures and (b) cone (sclx4), double partial-cone (sclx6) and double cone (sclx8) conformations.

We were motivated to characterize the sclxn series with a single protein and thus investigate systematically how the calixarene size and flexibility influence protein recognition and assembly. Furthermore, we were interested in studying a protein for which a crystal structure was not available. Acknowledging the tendency of sclxn to complex cationic proteins we chose the Penicillium antifungal protein (PAF) (Marx et al., 1995[Marx, F., Haas, H., Reindl, M., Stöffler, G., Lottspeich, F. & Redl, B. (1995). Gene, 167, 167-171.], 2008[Marx, F., Binder, U., Leiter, É. & Pócsi, I. (2008). Cell. Mol. Life Sci. 65, 445-454.]) as a test case. PAF is a small (∼6.2 kDa, 55 residues) lysine-rich protein (13 × Lys, pI ≃ 9) and a potent agent against Aspergillus species and dermatophytes (Binder et al., 2010[Binder, U., Oberparleiter, C., Meyer, V. & Marx, F. (2010). Mol. Microbiol. 75, 294-307.]; Leiter et al., 2005[Leiter, É., Szappanos, H., Oberparleiter, C., Kaiserer, L., Csernoch, L., Pusztahelyi, T., Emri, T., Pócsi, I., Salvenmoser, W. & Marx, F. (2005). Antimicrob. Agents Chemother. 49, 2445-2453.]; Palicz et al., 2016[Palicz, Z., Gáll, T., Leiter, É., Kollár, S., Kovács, I., Miszti-Blasius, K., Pócsi, I., Csernoch, L. & Szentesi, P. (2016). Microbes Infect. 5, e114.]). The NMR structure is a twisted β-barrel composed of five antiparallel β-strands and stabilized by three disulfide bridges (Batta et al., 2009[Batta, G., Barna, T., Gáspári, Z., Sándor, S., Kövér, K. E., Binder, U., Sarg, B., Kaiserer, L., Chhillar, A. K., Eigentler, A., Leiter, É., Hegedüs, N., Pócsi, I., Lindner, H. & Marx, F. (2009). FEBS J. 276, 2875-2890.]; Fizil et al., 2015[Fizil, Á., Gáspári, Z., Barna, T., Marx, F. & Batta, G. (2015). Chem. Eur. J. 21, 5136-5144.], 2018[Fizil, Á., Sonderegger, C., Czajlik, A., Fekete, A., Komáromi, I., Hajdu, D., Marx, F. & Batta, G. (2018). PLoS One, 13, e0204825.]). Lys30, Phe31, Lys34, Lys35 and Lys38 (loop 3) belong to a conserved region of PAF that is important for antifungal activity (Batta et al., 2009[Batta, G., Barna, T., Gáspári, Z., Sándor, S., Kövér, K. E., Binder, U., Sarg, B., Kaiserer, L., Chhillar, A. K., Eigentler, A., Leiter, É., Hegedüs, N., Pócsi, I., Lindner, H. & Marx, F. (2009). FEBS J. 276, 2875-2890.]; Sonderegger et al., 2016[Sonderegger, C., Galgóczy, L., Garrigues, S., Fizil, Á., Borics, A., Manzanares, P., Hegedüs, N., Huber, A., Marcos, J. F., Batta, G. & Marx, F. (2016). Microb. Cell Fact. 15, 192.]; Garrigues et al., 2017[Garrigues, S., Gandía, M., Popa, C., Borics, A., Marx, F., Coca, M., Marcos, J. F. & Manzanares, P. (2017). Sci. Rep. 7, 14663.]). Similar to defensins, the mechanism of antifungal action is postulated to require interaction with anionic components on the cell membrane (Binder et al., 2010[Binder, U., Oberparleiter, C., Meyer, V. & Marx, F. (2010). Mol. Microbiol. 75, 294-307.]; Garrigues et al., 2017[Garrigues, S., Gandía, M., Popa, C., Borics, A., Marx, F., Coca, M., Marcos, J. F. & Manzanares, P. (2017). Sci. Rep. 7, 14663.]; Silva et al., 2014[Silva, P. M., Gonçalves, S. & Santos, N. C. (2014). Front. Microbiol. 5, 97.]). Recent X-ray crystal structures have revealed how defensin–phospholipid binding leads to oligomerization, suggesting a mechanism for membrane permeation (Poon et al., 2014[Poon, I. K. H., Kh, , Baxter, A. A., Lay, F. T., Mills, G. D., Adda, C. G., Payne, J. A., Phan, T. K., Ryan, G. F., White, J. A., Veneer, P. K., van der Weerden, N. L., Anderson, M. A., Kvansakul, M. & Hulett, M. D. (2014). eLife, 3, e01808.]; Kvansakul et al., 2016[Kvansakul, M., Lay, F. T., Adda, C. G., Veneer, P. K., Baxter, A. A., Phan, T. K., Poon, I. K. H. & Hulett, M. D. (2016). Proc. Natl Acad. Sci. USA, 113, 11202-11207.]; Cools et al., 2017[Cools, T. L., Vriens, K., Struyfs, C., Verbandt, S., Ramada, M. H. S., Brand, G. D., Bloch, C. Jr, Koch, B., Traven, A., Drijfhout, J. W., Demuyser, L., Kucharíková, S., Van Dijck, P., Spasic, D., Lammertyn, J., Cammue, B. P. A. & Thevissen, K. (2017). Front. Microbiol. 8, 2295.]; Järvå et al., 2018[Järvå, M., Lay, F. T., Phan, T. K., Humble, C., Poon, I. K. H., Bleackley, M. R., Anderson, M. A., Hulett, M. D. & Kvansakul, M. (2018). Nat. Commun. 9, 1962.]). These observations provided further motivation to characterize PAF binding with anionic receptors.

Here, we report three PAF–sclxn crystal structures, demonstrating the fitness of calixarenes as crystallization agents. Interestingly, all three calixarenes were bound to PAF, mainly at the conserved loop 3. A similar interaction site was determined by NMR studies; these results suggest that loop 3 is favoured for recognition by anionic receptors. The largest calixarene sclx8 mediated a PAF dimer that was observed both crystallographically and in solution. The thermodynamics of PAF–sclxn interactions were characterized by isothermal titration calorimetry, providing further evidence of PAF dimerization via sclx8. The results are discussed in the context of protein assembly and membrane binding. Finally, insights into protein complexation by flexible calixarenes are provided, including the role of PEG fragments at the protein–calixarene interface.

2. Experimental

2.1. Materials

PAF was produced as described (Batta et al., 2009[Batta, G., Barna, T., Gáspári, Z., Sándor, S., Kövér, K. E., Binder, U., Sarg, B., Kaiserer, L., Chhillar, A. K., Eigentler, A., Leiter, É., Hegedüs, N., Pócsi, I., Lindner, H. & Marx, F. (2009). FEBS J. 276, 2875-2890.]; Sonderegger et al., 2016[Sonderegger, C., Galgóczy, L., Garrigues, S., Fizil, Á., Borics, A., Manzanares, P., Hegedüs, N., Huber, A., Marcos, J. F., Batta, G. & Marx, F. (2016). Microb. Cell Fact. 15, 192.]). The calixarenes were purchased from TCI Chemicals. Stock solutions of sclx4, sclx6 and sclx8 were prepared in water and the pH was adjusted to 6.0.

2.2. Crystallization trials

Co-crystallization experiments were performed by the hanging-drop vapour-diffusion method at 20°C. The reservoir solution was 20–30% PEG 3350 and 50 mM sodium acetate, pH 5.6. A range of protein (0.7–7.0 mM PAF) and ligand (5–40 mM sclx4) concentrations were tested for PAF–sclx4 co-crystallization. Drops were prepared by combining sequentially 1 µl each of reservoir solution, protein and sclx4. Crystals grew at 7 mM PAF and 40 mM sclx4. In the case of PAF–sclx6 and PAF–sclx8, the protein–ligand solutions were premixed before combining with the reservoir solution. Co-crystals were obtained with 10 mM sclx6 and 40 mM sclx8. Crystals grew in 4–5 days (sclx4), 2–3 weeks (sclx6) or 6–8 weeks (sclx8).

The crystallization of ligand-free PAF (7 mM) was performed with an Oryx 8 Robot (Douglas Instruments) and a sparse matrix screen (JCSG++, Jena Bioscience). Spherulites were obtained in C6 (40% PEG 300, 100 mM potassium phosphate citrate pH 4.2) and needles grew in D7 (40% PEG 400, 100 mM Tris–HCl pH 8.5, 200 mM lithium sulfate). Manual crystallization trials under these conditions did not yield suitable crystals.

2.3. X-ray data collection

Crystals were cryo-protected in reservoir solution supplemented with 20% glycerol and cryo-cooled in liquid nitrogen. Diffraction data were collected at the SOLEIL synchrotron (France) to 1.30, 1.45 and 1.50 Å for PAF–sclx4, PAF–sclx6 and PAF–sclx8, respectively. Datasets were collected using φ scans of 0.1° over 200° (PAF–sclx4), 180° (PAF–sclx6) and 110° (PAF–sclx8) using an EIGER X 9M detector. In the case of pure PAF, a dataset extending to 3.0 Å was collected for the spherulites (condition C6), but was difficult to index/integrate in both XDS and iMOSFLM. The needle-like crystals (condition D7) did not diffract.

2.4. Structure determination

The observed reflections for PAF–sclx4 were processed with XDS (Kabsch, 2010[Kabsch, W. (2010). Acta Cryst. D66, 125-132.]), whereas iMOSFLM (Battye et al., 2011[Battye, T. G. G., Kontogiannis, L., Johnson, O., Powell, H. R. & Leslie, A. G. W. (2011). Acta Cryst. D67, 271-281.]) was used for the PAF–sclx6 and PAF–sclx8 datasets. In all cases, the data were scaled using POINTLESS (Evans, 2011[Evans, P. R. (2011). Acta Cryst. D67, 282-292.]) and AIMLESS (Evans & Murshudov, 2013[Evans, P. R. & Murshudov, G. N. (2013). Acta Cryst. D69, 1204-1214.]). Xtriage (PHENIX, Adams et al., 2010[Adams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213-221.]) suggested pseudo-merohedral twinning for the PAF–sclx4 data with twin lawh, −k, −hl, and estimated twin fractions of 0.025 (Britton analyses), 0.066 (H-test) and 0.022 (maximum-likelihood method). The structure was determined by molecular replacement in 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.]) by using the NMR structure (PDB reference 2mhv, conformer 1; Fizil et al., 2015[Fizil, Á., Gáspári, Z., Barna, T., Marx, F. & Batta, G. (2015). Chem. Eur. J. 21, 5136-5144.]) as the search model. A satisfactory solution (LLG, 134; TFZ, 7.4) was obtained with a search model in which residues 1–2, 17–24 and 47–49 were deleted and all six cysteines were replaced by alanine. The coordinates and restraints for sclx4 (ligand ID T3Y) were added in COOT. Twin refinement did not result in any significant improvement in the electron density. No twinning was indicated for the PAF–sclx6 or PAF–sclx8 data. The structures were solved by molecular replacement using the structure of PAF–sclx4 (devoid of sclx4) as the search model. The coordinates for sclx6 and sclx8 were built in JLigand (Lebedev et al., 2012[Lebedev, A. A., Young, P., Isupov, M. N., Moroz, O. V., Vagin, A. A. & Murshudov, G. N. (2012). Acta Cryst. D68, 431-440.]). High mosaic spread (0.3–0.9) in the PAF–sclx8 dataset made it difficult to obtain better R values. Truncating the images with high mosaicity did not help in this respect. Iterative cycles of manual model building in COOT (Emsley et al., 2010[Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486-501.]) and refinement in BUSTER (Smart et al., 2012[Smart, O. S., Womack, T. O., Flensburg, C., Keller, P., Paciorek, W., Sharff, A., Vonrhein, C. & Bricogne, G. (2012). Acta Cryst. D68, 368-380.]) were carried out until no further improvements in Rfree and electron density were observed. The final structures were validated with MolProbity (Chen et al., 2010[Chen, V. B., Arendall, W. B., Headd, J. J., Keedy, D. A., Immormino, R. M., Kapral, G. J., Murray, L. W., Richardson, J. S. & Richardson, D. C. (2010). Acta Cryst. D66, 12-21.]) and deposited in the Protein Data Bank as PAF–sclx4 (PDB reference 6ha4), PAF–sclx6 (PDB reference 6hah) and PAF–sclx8 (PDB reference 6haj).

2.5. Accessible surface area calculations

The effect of sclx4, sclx6 and sclx8 on the accessible surface area (ASA) of PAF residues in the crystal packing environments was determined in AreaIMol as described previously (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]).

2.6. NMR spectroscopy

The sample conditions were 0.3 or 0.5 mM 15N-PAF in 10 mM sodium phosphate buffer at pH 6.0. NMR titrations were performed at 298 K using 0.5–1 µl aliquots of 50 mM stocks of sclx4, sclx6 or sclx8. 1H-15N HSQC spectra were acquired with spectral widths of 12 p.p.m. (1H) and 19 p.p.m. (15N) using two scans and 128 increments on a Bruker Avance-II-500 NMR spectrometer. Ligand-induced chemical-shift perturbations were analysed in CCPN (Delaglio et al., 1995[Delaglio, F., Grzesiek, S., Vuister, G. W., Zhu, G., Pfeifer, J. & Bax, A. (1995). J. Biomol. NMR, 6, 277-293.]).

2.7. Isothermal titration calorimetry and data fitting

PAF samples were dissolved in 10 mM sodium phosphate pH 6.0. The same buffer was used to dilute stocks of sclx4 (7.1 mM, PAF 0.5 mM), sclx6 (3.6 mM, PAF 0.5 mM) and sclx8 (2.5 mM, PAF 0.3 mM) to the required concentration. Samples were degassed prior to the titration. Measurements were made at 25°C using a Microcal ITC-200 instrument. Titrations were performed in duplicate with similar trends between each replicate. A single replicate from each calixarene was used for model fitting. Separate titrations of each calixarene into buffer confirmed that the heats of dilution were small, exothermic and approximately constant.

NITPIC (Keller et al., 2012[Keller, S., Vargas, C., Zhao, H., Piszczek, G., Brautigam, C. A. & Schuck, P. (2012). Anal. Chem. 84, 5066-5073.]) was used for baseline correction and integration of the thermograms. Pytc (Duvvuri et al., 2018[Duvvuri, H., Wheeler, L. C. & Harms, M. J. (2018). Biochemistry, 57, 2578-2583.]) was used to perform model fitting and parameter estimation. The system of equations relating the independent variables of the model (total concentrations) to the experimental observations (heat generated during injections) for the single-site and bidentate-ligand models are as follows.

Single-site model,

[\eqalignno { &{[P_{\rm T}]_i = [P]_i + [PL]_i} \cr &{[L_{\rm T}]_i = [L]_i + [PL]_i}, & (1)}]

[[PL]_i = K[P]_i [L]_i, \eqno (2)]

[q_i = V_{{\rm cell}}\Delta H^\circ \left\{{[PL]}_i - {[PL]}_{i-1} (1 - v_i/V_{{\rm cell}})\right\} + q_{{\rm dil}}, \eqno (3)]

where [PT]i is the total cell concentration of protein at the ith injection (independent variable), [LT]i is the total cell concentration of ligand at the ith injection (independent variable), K1 is the equilibrium association constant (fit parameter), ΔH is the enthalpy (fit parameter) associated with K, Vcell is the volume of the cell, vi is the volume of the ith injection, qi is the heat generated from the ith injection (dependent variable) and qdil is the heat of dilution (fit parameter, assumed to be constant)

Bidentate-ligand model,

[\eqalignno { & { [P_{\rm T}]}_i = [P]_i + [PL]_i + 2[P_2L]_i \cr & {[L_{\rm T}]}_i = [L]_i + [PL]_i + [P_2L]_i, & (4)}]

[\eqalignno { & { [P}L]_i = 2{ [K]}_1 {[P]}_i [L]_i \cr & [P_2L]_i = K_1K_2[P]_i^2 [L]_i, & (5)}]

[\eqalign { q_i =&\ V_{{\rm cell}}\big(\Delta H^\circ_1 \left\{[PL]_i - [PL]_{i-1} (1 - v_i/V_{{\rm cell}})\right\} \cr &+ \left(\Delta H^\circ_1 + \Delta H^\circ_2\right) \left\{[P_2L]_i - [P_2L]_{i-1} (1 - v_i/V_{{\rm cell}})\right\}\big) + q_{{\rm dil}},} \eqno (6)]

where K1 and K2 are the microscopic equilibrium association constants (fit parameters), ΔH1 and ΔH2 are the enthalpies (fit parameters) associated with K1 and K2, respectively

The expressions for mass balance of the protein and ligand can be represented by equations (1) or (4). Equation (2) or (5) can be used to define the equilibrium constants. For the bidentate ligand model, equation (5)[link] was solved numerically (the Levenberg–Marquardt algorithm) to yield the free-protein ([P]i) and free-ligand ([L]i) concentrations. The free concentrations were used to compute the concentrations of the other states via the equilibrium equations. The heat generated from a given injection was determined using either equations (3) or (6). Parameters were constrained to physically reasonable bounds (e.g. K1 and K2 values between 102 and 1010M−1) and best-fits were obtained by maximum likelihood starting from a range of initial estimates. Parameter errors and correlations were estimated using a Bayesian approach (Markov chain Monte Carlo simulations). The error for each integrated heat was determined using NITPIC (Keller et al., 2012[Keller, S., Vargas, C., Zhao, H., Piszczek, G., Brautigam, C. A. & Schuck, P. (2012). Anal. Chem. 84, 5066-5073.]).

3. Results and discussion

3.1. PAF–sclxn co-crystallization

Pure PAF proved to be recalcitrant to crystallization. A sparse-matrix screen yielded spherulites or needle-like crystals only (see experimental[link]). In contrast, PAF–sclx4 mixtures were crystallized readily from solutions containing PEG and sodium acetate. PAF–sclx4, PAF–sclx6 and PAF–sclx8 co-crystals were obtained at 28–30% PEG 3350 and 50 mM sodium acetate pH 5.6 (Fig. S1 and Table S1 of the supporting information).

3.2. Data collection and model building

Datasets extending to 1.30, 1.45 and 1.50 Å resolution were collected from monoclinic (P1211) PAF–sclx4, PAF–sclx6 and hexagonal (P61) PAF–sclx8 co-crystals, respectively (Table S1). The PAF–sclx4 structure was determined using the NMR coordinates (PDB reference 2mhv; Fizil et al., 2015[Fizil, Á., Gáspári, Z., Barna, T., Marx, F. & Batta, G. (2015). Chem. Eur. J. 21, 5136-5144.]) as the search model. To obtain a satisfactory solution it was necessary to delete two loops and replace all six cysteines with alanines. After several rounds of model building and refinement a complete PAF structure was obtained. This model was used to solve the PAF–sclx6 and PAF–sclx8 structures. The PAF fold and the three disulfide bridges in the X-ray structures were consistent with the NMR model (Batta et al., 2009[Batta, G., Barna, T., Gáspári, Z., Sándor, S., Kövér, K. E., Binder, U., Sarg, B., Kaiserer, L., Chhillar, A. K., Eigentler, A., Leiter, É., Hegedüs, N., Pócsi, I., Lindner, H. & Marx, F. (2009). FEBS J. 276, 2875-2890.]; Fizil et al., 2015[Fizil, Á., Gáspári, Z., Barna, T., Marx, F. & Batta, G. (2015). Chem. Eur. J. 21, 5136-5144.], 2018[Fizil, Á., Sonderegger, C., Czajlik, A., Fekete, A., Komáromi, I., Hajdu, D., Marx, F. & Batta, G. (2018). PLoS One, 13, e0204825.]). Interestingly, the fold was altered slightly in response to sclxn binding (Fig. S2). Superposition of the three structures revealed a Cα r.m.s.d. of 0.54 Å (PAF–sclx6) and 0.78 Å (PAF–sclx8) relative to PAF–sclx4, with the largest differences at loops 2, 3 and 4. The calculated energies of the disulfide bonds (Schmidt et al., 2006[Schmidt, B., Ho, L. & Hogg, P. J. (2006). Biochemistry, 45, 7429-7433.]) were approximately threefold lower in the X-ray structures compared with the NMR structure (Table S2).

In contrast to the PAF–sclxn crystals, the spherulites and needles of pure PAF failed to provide a usable dataset. The needles did not diffract and the spherulites yielded a 3.0 Å resolution dataset which proved difficult to index and integrate. The difficulty in obtaining suitable crystals of pure PAF suggests that the calixarene facilitates protein assembly and crystallization (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]; Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.], 2014[McGovern, R. E., McCarthy, A. A. & Crowley, P. B. (2014). Chem. Commun. 50, 10412-10415.], 2015[McGovern, R. E., Snarr, B. D., Lyons, J. A., McFarlane, J., Whiting, A. L., Paci, I., Hof, F. & Crowley, P. B. (2015). Chem. Sci. 6, 442-449.]; Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.], 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]).

3.3. Different calixarene, similar binding site

The asymmetric unit of the PAF–sclxn complexes comprised one (in the case of PAF–sclx4 and PAF–sclx6) or two (PAF–sclx8) molecules of PAF. Each structure contained one calixarene, as shown by the 2FoFc electron-density maps (Figs. 2[link] and S1). Additional electron density adjacent to sclx6 and sclx8 was modelled as a PEG fragment equivalent to tetraethylene glycol (EG4) and heptaethylene glycol (EG7), respectively (Figs. 2[link] and 3[link]). Sclx4, locked in the cone conformation, encapsulates the side chain of a single lysine (Lys30), as observed previously in different protein-clx4 complexes (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]; Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.], 2014[McGovern, R. E., McCarthy, A. A. & Crowley, P. B. (2014). Chem. Commun. 50, 10412-10415.], 2015[McGovern, R. E., Snarr, B. D., Lyons, J. A., McFarlane, J., Whiting, A. L., Paci, I., Hof, F. & Crowley, P. B. (2015). Chem. Sci. 6, 442-449.]). The larger flexible sclx6 and sclx8 adopted distinct conformations and bound at least two lysines. Sclx6 was in the double partial-cone conformation (Atwood et al., 1992[Atwood, J. L., Clark, D. L., Juneja, R. K., Orr, G. W., Robinson, K. D. & Vincent, R. L. (1992). J. Am. Chem. Soc. 114, 7558-7559.]; Dalgarno et al., 2003[Dalgarno, S. J., Hardie, M. J., Makha, M. & Raston, C. L. (2003). Chem. Eur. J. 9, 2834-2839.]), with three sulfonates pointed upwards and three pointed downwards [Figs. 1[link](b) and 2[link](b)]. Sclx8 adopted the double cone conformation (Liu et al., 2009[Liu, Y., Liao, W., Bi, Y., Wang, M., Wu, Z., Wang, X., Su, Z. & Zhang, H. (2009). CrystEngComm, 11, 1803-1806.]; Perret et al., 2006[Perret, F., Bonnard, V., Danylyuk, O., Suwinska, K. & Coleman, A. W. (2006). New J. Chem. 30, 987-990.]; Smith et al., 2006[Smith, C. B., Barbour, L. J., Makha, M., Raston, C. L. & Sobolev, A. N. (2006). New J. Chem. 30, 991-996.]), with each half of the molecule acting like a calix[4]arene to bind one PAF molecule, thus mediating a crystallographic dimer [Fig. 2[link](c)].

[Figure 2]
Figure 2
Binding-site interactions in PAF–sclxn. (a) sclx4, (b) sclx6 and (c) sclx8 binding to PAF at Lys30. Note the altered conformations of Lys30 and Phe31 in each structure, while Pro29 provides a rigid hydrophobic surface for face-to-face interaction with sclx6 and sclx8. In PAF–sclx8, two protein chains interact with the calixarene. PEG fragments equivalent to tetraethylene glycol and heptaethylene glycol were bound to sclx6 and sclx8, respectively.
[Figure 3]
Figure 3
Protein–PEG–calixarene interfaces. The protein–calixarene interfaces are completed by a PEG fragment in (a) PAF–sclx6 and (b) PAF–sclx8. Lys9 Nζ simultaneously forms ion–dipole bonds to the PEG (crown-ether-like complex) and a salt bridge to one sulfonate. CH—π and lone-pair—π bonds also occur between PEG and the calixarene phenolic rings.

All three calixarenes bound to Lys30, while interacting also with neighbouring residues as well as other proteins (symmetry mates) in the crystal packing. Depending on the ligand size/conformation, the noncovalent contacts varied in their type and multiplicity. The PAF–sclx4 complex [Fig. 2[link](a)] was similar to cytochrome c–sclx4 (McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.]), involving a salt bridge and CH—π/cation—π bonds with the encapsulated lysine. Hydrogen bonds to the backbone amide NHs of Lys30, Phe31 and Asp32 were evident and the aromatic ring of Phe31 was in van der Waals contact with an sclx4 methylene bridge. Considering symmetry mates [Fig. 4[link](a)], sclx4 formed substantial interfaces (>150 Å2) with three proteins. Interestingly, a salt bridge was formed with the Nα of Ala1. Salt bridges also occurred with Lys2, Lys17, Lys22 and Lys35, emphasizing a substantial charge–charge component to complexation. In total, the protein–sclx4 interfaces buried ∼660 Å2 of protein.

[Figure 4]
Figure 4
Calixarenes as molecular glues. The crystal packing is dominated by PAF–sclxn interactions in (a) PAF–sclx4, (b) PAF–sclx6 and (c) PAF–sclx8. This observation suggests that the calixarene acts as a molecular glue in protein assembly. Proteins, calixarenes and unit-cell axes are depicted in grey, green and blue, respectively. The PEG fragments are depicted as sticks.

Sclx6 (1.5 times larger than sclx4) also completely encaged Lys30 [Fig. 2[link](b)]. However, one wall of the calixarene cage was composed of three phenolic groups. The phenolic oxygens were in van der Waals contact with the Cβ, Cγ and Cδ of Lys30, indicative of CH⋯O hydrogen bonding and the Lys30 Nα was hydrogen bonded to a phenolic OH (rather than to a sulfonate). Other differences, with respect to sclx4, were water-mediated salt bridges between Lys30 Nζ and two sulfonates and a weak ππ interaction with Phe31 [Fig. 2[link](b)]. The adjacent residue Pro29 was also important for calixarene binding (vide infra). In terms of crystal packing [Fig. 4[link](b)], the larger sclx6 was nestled between five proteins and formed numerous salt bridges (with Lys6, Lys9, Lys11, Lys27, Lys38, Lys42). The resulting protein–ligand contacts mask ∼970 Å2 of protein surface. Compared with sclx4, the more extensive interactions exhibited by sclx6 may explain why four times less ligand was required to achieve crystal growth (see experimental[link] and Table S1).

The interactions of sclx8 with PAF were similar to those observed with sclx6, though less extensive. At twice the size of sclx4 it might be expected that sclx8 would mask a larger protein surface; however, sclx8 formed a PAF dimer [Figs. 2[link](c) and 4[link](c)] resulting in a total protein surface coverage of ∼950 Å2. The double-cone conformation (compared with the `pleated loop', Rennie et al., 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]) adopted by sclx8 minimized its contact with protein surfaces. Salt-bridge interactions involved up to three lysines from each monomer. Here, again a hydrogen bond was formed between the Lys30 Nα and a phenolic OH. In one of the protein chains Phe31 formed an edge-to-face interaction with an sclx8 phenolic ring. In protein chain B, Phe31 was disordered [Fig. 2[link](c)].

In complex with PAF, sclx4, sclx6 and sclx8 contributed an additional surface of ∼550, ∼850 and ∼1290 Å2 to the protein, respectively (calculated for a single protein). The exposed calixarene surface is a relatively homogenous `mask' that is conducive to forming noncovalent bridges with other proteins. Apparently, the calixarene acts as molecular glue (Fig. 4[link]) by providing a patch that mediates protein assembly (subsequently driving protein crystallization) in a special case of the `patchy particle model' (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]; Fusco et al., 2014[Fusco, D., Headd, J. J., De Simone, A., Wang, J. & Charbonneau, P. (2014). Soft Matter, 10, 290-302.]; James et al., 2015[James, S., Quinn, M. K. & McManus, J. J. (2015). Phys. Chem. Chem. Phys. 17, 5413-5420.]; Staneva & Frenkel, 2015[Staneva, I. & Frenkel, D. (2015). J. Chem. Phys. 143, 194511.]; Derewenda & Godzik, 2017[Derewenda, Z. S. & Godzik, A. (2017). Methods Mol. Biol. 1607, 77-115.]).

The presence of PEG fragments (EG4 and EG7) markedly distinguished the PAF–sclx6 and PAF–sclx8 complexes (Fig. 3[link]). The PEG–calixarene interaction involved lone-pair–π (Jain et al., 2009[Jain, A., Ramanathan, V. & Sankararamakrishnan, R. (2009). Protein Sci. 18, 595-605.]) and CH–π bonds, while the PEG–protein contacts included hydrogen bonds between the oxygen lone pairs and Lys9 (Lys9 Nζ⋯O—PEG = 3.0–3.3 Å). This crown-ether like Lys9–PEG interaction resembles the binding of lysine to 18-crown-6 (PDB entry 3wur; Lee et al., 2014[Lee, C. C., Maestre-Reyna, M., Hsu, K. C., Wang, H. C., Liu, C. I., Jeng, W. Y., Lin, L. L., Wood, R., Chou, C. C., Yang, J. M. & Wang, A. H. (2014). Angew. Chem. Int. Ed. 53, 13054-13058.]). A heptaethylene glycol fragment has been observed bound to an antibody (PDB entry 2ajs; Zhu et al., 2006[Zhu, X., Dickerson, T. J., Rogers, C. J., Kaufmann, G. F., Mee, J. M., McKenzie, K. M., Janda, K. D. & Wilson, I. A. (2006). Structure, 14, 205-216.]), where it adopted a crown-ether like conformation, compared with the extended conformation in PAF–sclx8. In addition, a crystal structure of an SH3 domain (PDB entry 5xg9; Gautam et al., 2017[Gautam, G., Rehman, S. A. A., Pandey, P. & Gourinath, S. (2017). Acta Cryst. D73, 672-682.]) revealed various PEG fragments at protein–protein interfaces. These examples suggest that the role of PEG is as an interface `filler' and possibly the PEG fragments (Fig. 3[link]) contribute towards calixarene conformation selection/stability.

3.4. Selectivity of PAF–sclxn complexation, why Lys30?

Considering that PAF contains 13 lysines the question arises as to why Lys30 was selected by sclxn. ASA calculations were used to probe the selectivity of sclxn for the Pro29-Lys30-Phe31 patch over other possible binding sites (Fig. 4[link]). The calculations accounted for contributions from symmetry mates in the crystal packing (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]). The effect of ligand binding on the ASA of all Lys, Pro, Phe and Tyr residues is plotted in Fig. 5[link]. At least half of the lysines, including Lys30, are highly exposed (ASA ≥ 125 Å2) in each structure in the absence of sclxn. This observation suggests that steric accessibility (McGovern et al., 2014[McGovern, R. E., McCarthy, A. A. & Crowley, P. B. (2014). Chem. Commun. 50, 10412-10415.]) was not the determining factor in sclxn selectivity. For example, Lys2 (>150 Å2) was significantly masked (ΔASA ≥ 15%) by binding with sclx4 only. Perhaps a salt-bridge interaction with Asp46 reduced the availability of Lys2 in the other complexes. In contrast, Lys30 was strongly affected by all three calixarenes (ΔASA up to 80%). Adjacent residue Lys27 was also strongly affected in the complexes with sclx6 and sclx8. The differences in the degree of masking can be attributed to the calixarene sizes (small, sclx4) and conformations (`double cone', sclx8). However, sclx8 had more in common with sclx6 than sclx4. For example, Lys9, Lys11 and Lys38 were 30–50% buried by sclx6 or sclx8, while sclx4 had no effect on these residues. Overall, calixarene binding resulted in significant masking of five (sclx4), eight (sclx6) and six (sclx8) lysines.

[Figure 5]
Figure 5
ASA plots. Accessibility of Lys, Pro, Phe and Tyr residues in ligand-free (black) and ligand-bound (grey) PAF. The PAF–sclx8 data correspond to chain A.

PAF has five aromatic residues, Phe25, Phe31, Tyr3, Tyr16 and Tyr48 (Fig. 5[link]); the latter is highly solvent exposed (∼200 Å2) and might be expected to interact with sclxn. However, only minor contributions were evident (Fig. S3). Phe31 was the dominant aromatic residue for sclxn complexation. The adjacent Lys30, Lys34 and Lys35 may facilitate (via charge–charge interactions) calixarene binding here, compared with Tyr48, which is proximal to Lys2 only. The contribution of Pro29 merits special attention as it completes the binding site for both sclx6 and sclx8 via face-to-face hydrophobic stacks with a phenolic ring [Figs. 2[link](b) and 2[link](c)]. These interactions are reminiscent of polyphenol binding to proline-rich proteins (Baxter et al., 1997[Baxter, N. J., Lilley, T. H., Haslam, E. & Williamson, M. P. (1997). Biochemistry, 36, 5566-5577.]; Charlton et al., 2002[Charlton, A. J., Haslam, E. & Williamson, M. P. (2002). J. Am. Chem. Soc. 124, 9899-9905.]; Quideau et al., 2011[Quideau, S., Deffieux, D., Douat-Casassus, C. & Pouységu, L. (2011). Angew. Chem. Int. Ed. 50, 586-621.]). The rigid pyrrolidine ring appears to provide a stable platform for binding the `floppy' sclx6 or sclx8. Thus, it is perhaps unsurprising that the only proline residue in PAF was involved at the binding site.

As such, it appears to be the combination of the Pro29-Lys30-Phe31 motif and adjacent lysines (charge–charge interactions) that stabilize sclxn binding and impart selectivity. This region has been implicated in PAF function, with decreased antifungal activity when Phe31, Lys35 or Lys38 were mutated to Asn or Ala (Batta et al., 2009[Batta, G., Barna, T., Gáspári, Z., Sándor, S., Kövér, K. E., Binder, U., Sarg, B., Kaiserer, L., Chhillar, A. K., Eigentler, A., Leiter, É., Hegedüs, N., Pócsi, I., Lindner, H. & Marx, F. (2009). FEBS J. 276, 2875-2890.]; Sonderegger et al., 2016[Sonderegger, C., Galgóczy, L., Garrigues, S., Fizil, Á., Borics, A., Manzanares, P., Hegedüs, N., Huber, A., Marcos, J. F., Batta, G. & Marx, F. (2016). Microb. Cell Fact. 15, 192.]; Garrigues et al., 2017[Garrigues, S., Gandía, M., Popa, C., Borics, A., Marx, F., Coca, M., Marcos, J. F. & Manzanares, P. (2017). Sci. Rep. 7, 14663.]). The selectivity of the anionic calixarenes for this site suggests that it may be involved in cell membrane binding and permeation as required for antifungal activity.

3.5. NMR characterization and comparison with the solid state

PAF–calixarene binding in solution was assessed by NMR spectroscopy. Titrations were performed by the addition of microlitre aliquots of sclxn to 15N-labelled PAF, which was monitored by 1H-15N HSQC spectroscopy (Fizil et al., 2018[Fizil, Á., Sonderegger, C., Czajlik, A., Fekete, A., Komáromi, I., Hajdu, D., Marx, F. & Batta, G. (2018). PLoS One, 13, e0204825.]; McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.]). The overlaid spectra (Fig. 6[link]) revealed increasing chemical-shift perturbations (Δδ) as a function of sclx4 or sclx6 concentration, indicative of fast to intermediate exchange between the ligand-free and ligand-bound states. Some biphasic shifts were evident for sclx6. Severe broadening effects were observed with ≥0.3 eq sclx8, indicative of a slow-exchange process and suggesting the possibility of ligand-mediated oligomerization (Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; Fonseca-Ornelas et al., 2017[Fonseca-Ornelas, L., Schmidt, C., Camacho-Zarco, A. R., Fernandez, C. O., Becker, S. & Zweckstetter, M. (2017). Chem. Eur. J. 23, 13010-13014.]; Mallon et al., 2016[Mallon, M., Dutt, S., Schrader, T. & Crowley, P. B. (2016). ChemBioChem, 17, 774-783.]; Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.], 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]).

[Figure 6]
Figure 6
NMR characterization of PAF–sclxn­ complexation. (a) Region from overlaid 1H-15N HSQC spectra of pure PAF (black contours) and in the presence of 0.1–0.6 mM ligand (coloured scale). Biphasic shifts occurred for resonances Lys30, Lys34 and Cys36 in the presence of sclx6. Resonances Lys11, Cys28, Lys30, Lys34 and Cys36 were broadened at 0.3 mM sclx8, while resonances Thr8, Lys11, Asp32 and Thr37 were broadened beyond detection at 0.6 mM sclx8. (b) Plots of chemical-shift perturbations measured for PAF backbone amides in the presence of 0.6 mM sclx4, sclx6 or sclx8. Blanks correspond to Pro30 and undetectable resonances (due to broadening).

The Δδ plot (Fig. 6[link]) shows a clear selectivity for sclx4 binding to Lys30 and neighbouring residues 31–36. In the crystal structure, all of these residues occurred in the vicinity of sclx4. Significant Δδ were observed also for the C-terminal Val52 and Cys54, which are further from the crystallographic binding site. However, both of these residues are adjacent to Pro29, and Cys54 is hydrogen bonded to Lys34, suggesting a mechanism for how these resonances sense ligand binding. In the presence of sclx6, the Δδ plot again shows a preference for binding around Lys30 as well as effects at the C-terminus (Val52 Nα is hydrogen bonded to sclx6). However, compared with sclx4, the shifts are 2–4 times larger and other segments of the primary structure (residues 6–13 and 42–45) were also affected. These two regions correspond to additional sclx6 binding sites evident in the crystal packing. Therefore, the NMR data suggests that the PAF–sclx6 interaction fluctuates, with the calixarene exploring different patches on the protein surface, as observed previously for cytochrome c–sclx4 complexes (Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; McGovern et al., 2012[McGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527-533.]). Judging from the magnitude of the shifts, binding to Lys30 is preferred while a weaker interaction occurred at a patch involving Lys6 and Lys42.

The titrations with sclx8 resulted in different effects. In addition to pronounced perturbations of Lys30 and neighbours, substantial broadening effects occurred. Cys28, Lys30, Lys34 and Cys36 broadened at 0.3 mM, and Thr8, Lys11, Asp32 and Thr37 broadened beyond detection at 0.6 mM sclx8. These eight residues are located at the crystallographically defined binding site. Thus, the broadening effects may be indicative of PAF dimerization, consistent with the sclx8-mediated dimer in the crystal structure [Fig. 2[link](c)]. Previously, we observed a complete loss of the HSQC spectrum of cytochrome c in complex with pclx6, which also yielded a dimer in the solid state (Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.]).

3.6. Thermodynamics of PAF–sclxn complexation

Isothermal titration calorimetry was used to characterize the PAF–sclxn binding affinities and stoichiometries (Fig. 7[link]). The data were fitted to a single-site or a bidentate-ligand model. The latter model describes a bidentate ligand that can bind two protein molecules and was necessary to describe the obviously biphasic data for sclx8. The choice of this model is supported by the observation of a PAF–sclx8–PAF dimer in the crystal structure, and by the spectral broadening in the NMR experiments. All of the fit parameters were well determined by the data (Table 1[link]), with parameter errors assessed by Bayesian methods (Patil et al., 2010[Patil, A., Huard, D. & Fonnesbeck, C. J. (2010). J. Stat. Soft. 35, 1-81.]).

Table 1
Thermodynamics of PAF–sclxn complexation determined by ITC

Fit values are median (2.5% quantile, 97.5% quantile) from the Markov chain Monte Carlo method. In the case of sclx6, the fit parameters for both models are shown.

[Ligand] (µM) [PAF]M) KdM) ΔH (kJ mol−1) TΔS (kJ mol−1)
PAF–sclx4 (single-site model)
7143 (1248) 500 (412) 107.0 (0.0, 0.0) −16.9 (0.1, 0.1) −5.6 (0.2, 0.2)
         
PAF–sclx6 (single-site model)
3623 (633) 500 (412) 15.4 (0.0, 0.0) −28.2 (0.2, 0.1) −0.7 (0.2, 0.2)
         
PAF–sclx6 (bidentate-ligand model)
3623 (633) 500 (412) 47.8 (0.0, 0.0) −9.2 (0.1, 0.2) −15.4 (0.4, 0.3)
    45.8 (3.5, 4.1) −20.1 (0.4, 0.4) −4.6 (0.6, 0.6)
         
PAF–sclx8 (bidentate-ligand model)
2500 (437) 300 (247) 10.6 (1.3, 1.4) −3.5 (0.3, 0.3) −24.8 (0.3, 0.3)
    33.5 (4.8, 6.5) −36.0 (1.7, 1.4) 10.5 (1.8, 2.1)
†Calixarene concentrations in the syringe; the final concentrations are indicated in parentheses.
‡Protein concentrations in the cell; the final concentrations are indicated in parentheses.
[Figure 7]
Figure 7
ITC analysis of PAF–sclxn complexation. Top panels show the baseline-corrected thermograms for injections of sclx4, sclx6 or sclx8 into PAF. Bottom panels are the observed heats (data points) and the fits (solid line) for single-site (sclx4) and bidentate-ligand (sclx8) models.

The isotherms for sclx4 injected into PAF were fitted to a single-site binding model with Kd ∼110 µM. In contrast, the isotherms for sclx8 were biphasic (Brautigam, 2015[Brautigam, C. A. (2015). Methods, 76, 124-136.]) and fitted to a bidentate ligand model with Kd values of ∼10 and ∼30 µM, for binding the first and second molecule of PAF, respectively. The isotherms for sclx6 were intermediate between sclx4 and sclx8, suggesting that this ligand may exhibit weak bidentate binding. A satisfactory fit for this data was not obtained with either model. The ITC data demonstrate an increasing affinity for PAF as the calixarene size increases and a switch in binding mode from the small, rigid sclx4 (single site) to the large, flexible sclx8 (bidentate).

4. Conclusions

Using a combination of X-ray crystallography and NMR spectroscopy it was demonstrated that the sclxn series binds selectively to the highly cationic PAF. Despite the varying size and conformational flexibility, sclx4, sclx6 and sclx8 bound similarly the Pro29-Lys30-Phe31 motif in loop 3. The selectivity of the anionic calixarenes for this motif, and the role of loop 3 in antifungal activity, suggests that this region may be required for membrane binding. In addition to charge–charge interactions (showed by numerous lysine-to-sulfonate salt bridges), other noncovalent bonds including CH–π and ππ (via Pro29 and Phe31, respectively) participated in ligand stabilization. The presence of PEG fragments at the protein–sclx6 and protein–sclx8 interfaces suggests that PEG acts as a `filler' to complete the binding site, potentially reinforcing the calixarene conformation.

The structures of all three PAF–sclxn co-crystals highlight the potential of calixarenes as a `sticky patch' on the protein surface that facilitates assembly and crystallization. In the case of the sclx4 and sclx6 co-crystals (P1211), it is evident that the calixarene is a dominant contributor to the crystal packing (Fig. 4[link]). Similarly in the sclx8 structure (P61), the packing involves substantial protein–calixarene contacts, and the structure is interesting as sclx8 mediates a PAF dimer. Previously, we found that sclx8 mediates a tetramer of cytochrome c (Rennie et al., 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]). Generally, it seems that calixarene-mediated protein crystallization may be a special case of the patchy particle model for protein assembly (Alex et al., 2018[Alex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467-2473.]; Fusco et al., 2014[Fusco, D., Headd, J. J., De Simone, A., Wang, J. & Charbonneau, P. (2014). Soft Matter, 10, 290-302.]; James et al., 2015[James, S., Quinn, M. K. & McManus, J. J. (2015). Phys. Chem. Chem. Phys. 17, 5413-5420.]; Staneva & Frenkel, 2015[Staneva, I. & Frenkel, D. (2015). J. Chem. Phys. 143, 194511.]; Derewenda & Godzik, 2017[Derewenda, Z. S. & Godzik, A. (2017). Methods Mol. Biol. 1607, 77-115.]). Considering that PAF alone did not yield diffraction-quality crystals, we conclude that co-crystallization with sclxn was beneficial. Anionic calixarenes may generally facilitate crystallization and structure determination of small cationic proteins.

The binding surfaces observed in the NMR experiments were consistent with the X-ray data. However, the NMR effects were more pronounced with increasing calixarene size, suggesting that the larger calixarenes mask a greater portion of the protein surface and/or lead to assembly in solution. Similarly, the ITC experiments revealed tighter affinities and more complex effects with increasing calixarene size. In particular, sclx8 behaved as a bidentate ligand that facilitated PAF dimerization. These data add to the growing evidence of calixarene-mediated protein assembly in solution (Doolan et al., 2018[Doolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984-991.]; Rennie et al., 2017[Rennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517-5521.], 2018[Rennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764-13769.]). In terms of the biological relevance of these data it is noted that defensin oligomerization (upon phospholipid binding) has implications for antifungal activity (Poon et al., 2014[Poon, I. K. H., Kh, , Baxter, A. A., Lay, F. T., Mills, G. D., Adda, C. G., Payne, J. A., Phan, T. K., Ryan, G. F., White, J. A., Veneer, P. K., van der Weerden, N. L., Anderson, M. A., Kvansakul, M. & Hulett, M. D. (2014). eLife, 3, e01808.]; Järvå et al., 2018[Järvå, M., Lay, F. T., Phan, T. K., Humble, C., Poon, I. K. H., Bleackley, M. R., Anderson, M. A., Hulett, M. D. & Kvansakul, M. (2018). Nat. Commun. 9, 1962.]). Perhaps calixarenes can be used to modulate the activity of PAF and related proteins.

Acknowledgements

We acknowledge R. Pierattelli and the other organizers of the Chianti Workshop 2016 where this collaboration was initiated. Thanks also to the SOLEIL synchrotron for beam time allocation, and the staff at beamline PROXIMA 2A for their assistance with data collection.

Funding information

This research was supported by NUI Galway (Hardiman Scholarship to JMA), the Hungarian Science Fund (OKTA-ANN 110821 to GB), the European Regional Development Fund (GINOP-2.3.2–15-2016–00008 to GB and GINOP-2.3.3–15-2016–00004) and Science Foundation Ireland (13/ERC/B2912 and 13/CDA/2168 to PBC).

References

First citationAdams, P. D., Afonine, P. V., Bunkóczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L.-W., Kapral, G. J., Grosse-Kunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C. & Zwart, P. H. (2010). Acta Cryst. D66, 213–221.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAlex, J. M., Rennie, M. L., Volpi, S., Sansone, F., Casnati, A. & Crowley, P. B. (2018). Cryst. Growth Des. 18, 2467–2473.  CrossRef CAS Google Scholar
First citationAtwood, J. L., Clark, D. L., Juneja, R. K., Orr, G. W., Robinson, K. D. & Vincent, R. L. (1992). J. Am. Chem. Soc. 114, 7558–7559.  CrossRef CAS Google Scholar
First citationBaldini, L., Casnati, A. & Sansone, F. (2017). Comprehensive Supramolecular Chemistry II, Vol. 4, edited by J. Atwood, G. W. Gokel & L. Barbour, pp. 371–408. Amsterdam: Elsevier.  Google Scholar
First citationBatta, G., Barna, T., Gáspári, Z., Sándor, S., Kövér, K. E., Binder, U., Sarg, B., Kaiserer, L., Chhillar, A. K., Eigentler, A., Leiter, É., Hegedüs, N., Pócsi, I., Lindner, H. & Marx, F. (2009). FEBS J. 276, 2875–2890.  CrossRef CAS Google Scholar
First citationBattye, T. G. G., Kontogiannis, L., Johnson, O., Powell, H. R. & Leslie, A. G. W. (2011). Acta Cryst. D67, 271–281.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBaxter, N. J., Lilley, T. H., Haslam, E. & Williamson, M. P. (1997). Biochemistry, 36, 5566–5577.  CrossRef CAS PubMed Web of Science Google Scholar
First citationBinder, U., Oberparleiter, C., Meyer, V. & Marx, F. (2010). Mol. Microbiol. 75, 294–307.  CrossRef CAS Google Scholar
First citationBrautigam, C. A. (2015). Methods, 76, 124–136.  Web of Science CrossRef CAS Google Scholar
First citationCharlton, A. J., Haslam, E. & Williamson, M. P. (2002). J. Am. Chem. Soc. 124, 9899–9905.  CrossRef CAS Google Scholar
First citationChen, V. B., Arendall, W. B., Headd, J. J., Keedy, D. A., Immormino, R. M., Kapral, G. J., Murray, L. W., Richardson, J. S. & Richardson, D. C. (2010). Acta Cryst. D66, 12–21.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCools, T. L., Vriens, K., Struyfs, C., Verbandt, S., Ramada, M. H. S., Brand, G. D., Bloch, C. Jr, Koch, B., Traven, A., Drijfhout, J. W., Demuyser, L., Kucharíková, S., Van Dijck, P., Spasic, D., Lammertyn, J., Cammue, B. P. A. & Thevissen, K. (2017). Front. Microbiol. 8, 2295.  CrossRef Google Scholar
First citationDalgarno, S. J., Hardie, M. J., Makha, M. & Raston, C. L. (2003). Chem. Eur. J. 9, 2834–2839.  CrossRef CAS Google Scholar
First citationDelaglio, F., Grzesiek, S., Vuister, G. W., Zhu, G., Pfeifer, J. & Bax, A. (1995). J. Biomol. NMR, 6, 277–293.  CrossRef CAS PubMed Web of Science Google Scholar
First citationDerewenda, Z. S. & Godzik, A. (2017). Methods Mol. Biol. 1607, 77–115.  CrossRef CAS Google Scholar
First citationDoolan, A. M., Rennie, M. L. & Crowley, P. B. (2018). Chem. Eur. J. 24, 984–991.  CrossRef CAS Google Scholar
First citationDun, S. van, Ottmann, C., Milroy, L. G. & Brunsveld, L. (2017). J. Am. Chem. Soc. 139, 13960–13968.  Google Scholar
First citationDuvvuri, H., Wheeler, L. C. & Harms, M. J. (2018). Biochemistry, 57, 2578–2583.  CrossRef CAS Google Scholar
First citationEmsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486–501.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEvans, P. R. (2011). Acta Cryst. D67, 282–292.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationEvans, P. R. & Murshudov, G. N. (2013). Acta Cryst. D69, 1204–1214.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFizil, Á., Gáspári, Z., Barna, T., Marx, F. & Batta, G. (2015). Chem. Eur. J. 21, 5136–5144.  CrossRef CAS Google Scholar
First citationFizil, Á., Sonderegger, C., Czajlik, A., Fekete, A., Komáromi, I., Hajdu, D., Marx, F. & Batta, G. (2018). PLoS One, 13, e0204825.  CrossRef Google Scholar
First citationFonseca-Ornelas, L., Schmidt, C., Camacho-Zarco, A. R., Fernandez, C. O., Becker, S. & Zweckstetter, M. (2017). Chem. Eur. J. 23, 13010–13014.  CAS Google Scholar
First citationFusco, D., Headd, J. J., De Simone, A., Wang, J. & Charbonneau, P. (2014). Soft Matter, 10, 290–302.  CrossRef CAS Google Scholar
First citationGarrigues, S., Gandía, M., Popa, C., Borics, A., Marx, F., Coca, M., Marcos, J. F. & Manzanares, P. (2017). Sci. Rep. 7, 14663.  CrossRef Google Scholar
First citationGautam, G., Rehman, S. A. A., Pandey, P. & Gourinath, S. (2017). Acta Cryst. D73, 672–682.  CrossRef IUCr Journals Google Scholar
First citationGiuliani, M., Morbioli, I., Sansone, F. & Casnati, A. (2015). Chem. Commun. 51, 14140–14159.  CrossRef CAS Google Scholar
First citationGuagnini, F., Antonik, P. M., Rennie, M. L., O'Byrne, P., Khan, A. R., Pinalli, R., Dalcanale, E. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 7126–7130.  CrossRef CAS Google Scholar
First citationGuo, D. S. & Liu, Y. (2014). Acc. Chem. Res. 47, 1925–1934.  CrossRef CAS Google Scholar
First citationGutsche, C. D. & Bauer, L. J. (1985). J. Am. Chem. Soc. 107, 6052–6059.  CrossRef CAS Google Scholar
First citationJain, A., Ramanathan, V. & Sankararamakrishnan, R. (2009). Protein Sci. 18, 595–605.  Web of Science PubMed CAS Google Scholar
First citationJames, S., Quinn, M. K. & McManus, J. J. (2015). Phys. Chem. Chem. Phys. 17, 5413–5420.  CrossRef CAS Google Scholar
First citationJärvå, M., Lay, F. T., Phan, T. K., Humble, C., Poon, I. K. H., Bleackley, M. R., Anderson, M. A., Hulett, M. D. & Kvansakul, M. (2018). Nat. Commun. 9, 1962.  Google Scholar
First citationKabsch, W. (2010). Acta Cryst. D66, 125–132.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationKeller, S., Vargas, C., Zhao, H., Piszczek, G., Brautigam, C. A. & Schuck, P. (2012). Anal. Chem. 84, 5066–5073.  Web of Science CrossRef CAS PubMed Google Scholar
First citationKvansakul, M., Lay, F. T., Adda, C. G., Veneer, P. K., Baxter, A. A., Phan, T. K., Poon, I. K. H. & Hulett, M. D. (2016). Proc. Natl Acad. Sci. USA, 113, 11202–11207.  CrossRef CAS Google Scholar
First citationLebedev, A. A., Young, P., Isupov, M. N., Moroz, O. V., Vagin, A. A. & Murshudov, G. N. (2012). Acta Cryst. D68, 431–440.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationLee, C. C., Maestre-Reyna, M., Hsu, K. C., Wang, H. C., Liu, C. I., Jeng, W. Y., Lin, L. L., Wood, R., Chou, C. C., Yang, J. M. & Wang, A. H. (2014). Angew. Chem. Int. Ed. 53, 13054–13058.  CrossRef CAS Google Scholar
First citationLeiter, É., Szappanos, H., Oberparleiter, C., Kaiserer, L., Csernoch, L., Pusztahelyi, T., Emri, T., Pócsi, I., Salvenmoser, W. & Marx, F. (2005). Antimicrob. Agents Chemother. 49, 2445–2453.  CrossRef CAS Google Scholar
First citationLiu, Y., Liao, W., Bi, Y., Wang, M., Wu, Z., Wang, X., Su, Z. & Zhang, H. (2009). CrystEngComm, 11, 1803–1806.  CrossRef CAS Google Scholar
First citationMallon, M., Dutt, S., Schrader, T. & Crowley, P. B. (2016). ChemBioChem, 17, 774–783.  CrossRef CAS Google Scholar
First citationMarx, F., Binder, U., Leiter, É. & Pócsi, I. (2008). Cell. Mol. Life Sci. 65, 445–454.  CrossRef CAS Google Scholar
First citationMarx, F., Haas, H., Reindl, M., Stöffler, G., Lottspeich, F. & Redl, B. (1995). Gene, 167, 167–171.  CrossRef CAS 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 citationMcGovern, R. E., Fernandes, H., Khan, A. R., Power, N. P. & Crowley, P. B. (2012). Nat. Chem. 4, 527–533.  CrossRef CAS Google Scholar
First citationMcGovern, R. E., McCarthy, A. A. & Crowley, P. B. (2014). Chem. Commun. 50, 10412–10415.  CrossRef CAS Google Scholar
First citationMcGovern, R. E., Snarr, B. D., Lyons, J. A., McFarlane, J., Whiting, A. L., Paci, I., Hof, F. & Crowley, P. B. (2015). Chem. Sci. 6, 442–449.  CrossRef CAS Google Scholar
First citationMcPherson, A., Nguyen, C., Cudney, R. & Larson, S. B. (2011). Cryst. Growth Des. 11, 1469–1474.  Web of Science CrossRef CAS Google Scholar
First citationPalicz, Z., Gáll, T., Leiter, É., Kollár, S., Kovács, I., Miszti-Blasius, K., Pócsi, I., Csernoch, L. & Szentesi, P. (2016). Microbes Infect. 5, e114.  Google Scholar
First citationPatil, A., Huard, D. & Fonnesbeck, C. J. (2010). J. Stat. Soft. 35, 1–81.  CrossRef Google Scholar
First citationPerret, F., Bonnard, V., Danylyuk, O., Suwinska, K. & Coleman, A. W. (2006). New J. Chem. 30, 987–990.  CrossRef CAS Google Scholar
First citationPoon, I. K. H., Kh, , Baxter, A. A., Lay, F. T., Mills, G. D., Adda, C. G., Payne, J. A., Phan, T. K., Ryan, G. F., White, J. A., Veneer, P. K., van der Weerden, N. L., Anderson, M. A., Kvansakul, M. & Hulett, M. D. (2014). eLife, 3, e01808.  Google Scholar
First citationQuideau, S., Deffieux, D., Douat-Casassus, C. & Pouységu, L. (2011). Angew. Chem. Int. Ed. 50, 586–621.  CrossRef CAS Google Scholar
First citationRennie, M. L., Doolan, A. M., Raston, C. L. & Crowley, P. B. (2017). Angew. Chem. Int. Ed. 56, 5517–5521.  CrossRef CAS Google Scholar
First citationRennie, M. L., Fox, G. C., Pérez, J. & Crowley, P. B. (2018). Angew. Chem. Int. Ed. 57, 13764–13769.  CrossRef CAS Google Scholar
First citationSchmidt, B., Ho, L. & Hogg, P. J. (2006). Biochemistry, 45, 7429–7433.  Web of Science CrossRef PubMed CAS Google Scholar
First citationSilva, P. M., Gonçalves, S. & Santos, N. C. (2014). Front. Microbiol. 5, 97.  Google Scholar
First citationSmart, O. S., Womack, T. O., Flensburg, C., Keller, P., Paciorek, W., Sharff, A., Vonrhein, C. & Bricogne, G. (2012). Acta Cryst. D68, 368–380.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSmith, C. B., Barbour, L. J., Makha, M., Raston, C. L. & Sobolev, A. N. (2006). New J. Chem. 30, 991–996.  CrossRef CAS Google Scholar
First citationSonderegger, C., Galgóczy, L., Garrigues, S., Fizil, Á., Borics, A., Manzanares, P., Hegedüs, N., Huber, A., Marcos, J. F., Batta, G. & Marx, F. (2016). Microb. Cell Fact. 15, 192.  CrossRef Google Scholar
First citationStaneva, I. & Frenkel, D. (2015). J. Chem. Phys. 143, 194511.  CrossRef Google Scholar
First citationWang, K., Guo, D. S., Zhao, M. Y. & Liu, Y. (2016). Chem. Eur. J. 22, 1475–1483.  CrossRef CAS Google Scholar
First citationZhu, X., Dickerson, T. J., Rogers, C. J., Kaufmann, G. F., Mee, J. M., McKenzie, K. M., Janda, K. D. & Wilson, I. A. (2006). Structure, 14, 205–216.  Web of Science CrossRef PubMed CAS 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.

IUCrJ
Volume 6| Part 2| March 2019| Pages 238-247
ISSN: 2052-2525