Structural Studies Revealed Active Site Distortions of Human Furin by

Apr 12, 2017 - Structural Studies Revealed Active Site Distortions of Human Furin by a ... starting point of structure guided inhibitor development fo...
0 downloads 0 Views 3MB Size
Letters pubs.acs.org/acschemicalbiology

Structural Studies Revealed Active Site Distortions of Human Furin by a Small Molecule Inhibitor Sven O. Dahms,†,§ Guan-Sheng Jiao,‡,∥ and Manuel E. Than*,† †

Protein Crystallography Group, Leibniz Institute on Aging - Fritz Lipmann Institute (FLI), Beutenbergstr. 11, 07745 Jena, Germany Department of Chemistry, Hawaii Biotech, Inc., Honolulu, Hawaii, United States § Department of Molecular Biology, University of Salzburg, Billrothstrasse 11, A-5020 Salzburg, Austria ∥ MedChem ShortCut, LLC, Pearl City, Hawaii United States ‡

S Supporting Information *

ABSTRACT: Proprotein convertases (PCs) represent highly selective serine proteases that activate their substrates upon proteolytic cleavage. Their inhibition is a promising strategy for the treatment of several pathologies including cancer, atherosclerosis, hypercholesterolaemia, and infectious diseases. Here, we present the first experimental complex of furin with a non-substrate-like small molecule inhibitor, and the X-ray structure of the enzyme complexed to the small molecule inhibitor 1 at 1.9 Å resolution. Two molecules of inhibitor 1 were found to interact with furin. One is anchored at the S4 pocket of the enzyme and interferes directly with the conformation and function of the catalytic triade; the other molecule shows weaker binding and interacts with a distant, less conserved region of furin. The observed binding modes represent a new inhibition strategy of furin and imply the possibility to attain specificity among the PCs providing an innovative starting point of structure guided inhibitor development for furin.

T

specific hydrogen bonds and tight contacts. The primary interactions occur at the P1/S1 and P4/S4 sites. A number of additional main-chain contacts stabilize binding and registry further. The structural study of unliganded furin further showed how this unusually tight interaction still allows for a rapid turnover by a substrate-induced conformational change and activation mechanism.20 This process involves conversion of an inactive low-energy state to an active high-energy state triggered by binding of substrates or canonical inhibitors. Molecular modeling approaches suggested a high conservation of substrate binding in the PC protease family.21 Potent nonpeptidic 2,5-dideoxystreptamine-based small molecule inhibitors against furin have been described in the literature.11 However, it is yet largely unclear how these molecules bind to furin and how they act on a molecular/ atomic basis. Typical substrate-like furin inhibitors often carry a net positive charge to induce the electrostatic interaction with the large negative charge within the active site cleft of furin.9 Inhibitors, which do not bind directly into the active site cleft of furin or do not interact in a substrate-like manner, however, do not need to fulfill this prerequisite. This is of special importance for developing inhibitors of preferential pharmacokinetic and

he highly specific, calcium-dependent proprotein/prohormone convertases (PCs) are subtilisin-related serine proteases, composed in their mature form of a subtilisin-like catalytic domain and a P or HomoB domain.1 They proteolytically mature and thus activate a large number of secreted homeostatic but also pathogenic peptides and proteins. In mammals, the PC family contains seven members that cleave after multiple basic residues (furin, PC1, PC2, PC4, PACE4, PC5/6, and PC7) recognizing the general cleavage site (R/ K)Xn(R/K)↓ (where “↓” represents the scissile peptide bond). Among them, furin2 represents the probably best studied member of this family and is often regarded as “prototype-PC,” which does preferentially recognize the motive R-X-K/R-R-↓.1 In addition to their homeostatic function, PCs are also involved in diverse pathological situations. These include bacterial and viral infections as well as cancer and metastasis that are associated with PC activity. Therefore, furin and other PCs are intensely investigated as pharmacological targets.3,4 Several inhibition strategies have been employed in the past including the use of substrate-like5−10 and nonpeptidic small molecule inhibitors, 11−14 but also of larger inhibitory proteins.15,16 Binding of canonical substrate-like inhibitors into the active site cleft of mouse furin,17 human furin,8,18 and the yeast homologue Kex2p19 was investigated by X-ray crystallography, which revealed that furin possesses an extended, negatively charged active site cleft that binds to the primarily positively charged substrate/inhibitor involving many © 2017 American Chemical Society

Received: December 15, 2016 Accepted: April 12, 2017 Published: April 12, 2017 1211

DOI: 10.1021/acschembio.6b01110 ACS Chem. Biol. 2017, 12, 1211−1216

Letters

ACS Chemical Biology

Figure 1. Structure of the furin:1 complex. (a) Chemical structures of inhibitors 1, 2, and 3, which were used for soaking of furin crystals. (b−d) Human furin is shown as a cartoon representation. The catalytic domain and the P domain are colored in gold and blue, respectively. Bound calcium and sodium ions are shown as green and magenta spheres, respectively. Inhibitor 1 is shown as a ball and stick model with 1−1 in magenta and 1−2 in green. (b) Overview of the complex structure. Side chains of catalytic residues are indicated as a stick model in cyan. (c) Closer view of the binding site of 1−1 and 1−2, showing the region indicated as a black rectangle in b. Side chains of the catalytic triade are indicated as a stick model in cyan. The Fo−Fc annealed omit electron density map is shown as a dark green mesh contoured at 3σ. (d) View as shown in c rotated by 45°.

physicochemical parameters. Furin adopts distinct conformational states, and small molecule noncanonical inhibitors might also target its inactive state or interfere with the conversion to the active state.20 In order to understand their mode of inhibition and to further develop their molecular architecture, we have analyzed the structure of 2,5-dideoxystreptamine derived inhibitors in complex with human furin. Here, we show how the small molecule inhibitor 1 interacts with furin different from a peptide-based substrate in an unusual binding mode. Complexation of Human Furin with 2,5-Dideoxystreptamine Derived Inhibitors. We have studied the binding of three representatives of described, noncovalent 2,5-dideoxystreptamine derived furin inhibitors by protein

crystallography. Following a soaking procedure, the previously established orthorhombic crystals of human furin18 were probed with inhibitors 1 (racemate), 2 (racemate), and 3 (meso compound; Figure 1a). These orthorhombic crystals were grown by cocrystallization of furin with a peptide based inhibitor. Although the peptidic inhibitor in the orthorhombic crystals can be replaced by soaking with other inhibitors at competing concentrations,18 we never observed binding of 1, 2, or 3. We then repeated the soaking trials with hexagonal crystals of unliganded furin (inhibitor or substrate free enzyme20). For 1, we observed binding to furin as indicated by the electron density. Binding of 2 or 3, however, was also not observed in the hexagonal crystals despite their lower Ki values (Table 111). This observation suggests a different 1212

DOI: 10.1021/acschembio.6b01110 ACS Chem. Biol. 2017, 12, 1211−1216

Letters

ACS Chemical Biology

assignment as the two O-linked moieties. The short “arm” of electron density (Figure 1c and d, for stereo view see Figure S1) was interpreted as the guanidino group but could also be the proximal ring atoms of the phenyl-ring of the single guanidino-phenyl-amino substituent in the other enantiomer, i.e., (1S,2S,4R,5R)-1−1. As a racemic mixture was used for soaking, this alternative stereoisomer cannot be safely excluded (Figure 1a). The single guanidino group and the single guanidino-phenyl-amino substituent might thus be exchanged in molecule 1−1. Both substituents are largely flexible and make no major contact with residues of the protein, however (see below). For the second molecule of inhibitor 1, 1−2, the assignment of stereochemistry is unequivocal. As a certain electron density is visible for all three large substituents of the central cyclohexane ring, only the enantiomer with the (1R,2R,4S,5S)- configuration is possible (Figure 1c and d). Molecule 1−1 Binds to the Catalytic Triade. The first molecule of inhibitor 1, 1−1, interferes directly with the catalytic competent conformation of the catalytic triade. Hereby one guanidino-phenoxyl group of 1−1 forms charged hydrogen bonds to Asp153 (Figure 2a). The guanidino group adopts a position that is usually occupied by His194 (see also below). As a consequence Ser368, which usually mediates a hydrogen bond to His194, also reorients (Figure 2a). We found two alternative orientations of Ser368, which locates at a loose hydrogen bonding distance either to Asn295 or to the carbonyl oxygen of Thr365. Inhibitor 1−1 is anchored with the other guanidio-phenoxyl group at the S4 site of furin. Here, the guanidino group forms charged hydrogen bonds to Asp264 and Glu236 as well as a hydrogen bond to Tyr308 (Figure 2a). In addition, the two phenyl rings of the guanidino-phenoxyl groups and the cyclohexane core of 1−1 contribute to hydrophobic interactions. The inhibitor covers a surface area composed of Val231, Leu227, and TRP254 at the bottom of furin’s active site cleft (Figure 2a). The guanidino group directly attached to the central cyclohexane ring and the guanidino-phenyl-amino group are not involved in any interaction, and both point into the solvent (Figure 1b−d, Figure 2a). Likely due to the lack of specific interactions, these parts are also flexible and hence are not well-defined in the electron density map (Figure 1c and d). Molecule 1−2 Blocks the S2 Pocket. The second molecule of inhibitor 1, 1−2, binds further “north” of 1−1. It is positioned on top of a planar peptide stretch including Asp228−Glu230 and mediates van der Waals interactions to this loop region (Figure 2b). Interestingly, 1−2 is bound directly side by side to 1−1, covering a hydrophobic surface region that is between the two inhibitor molecules. The only tight contact of 1−2 to furin is mediated by its guanidinophenyl-amino group forming hydrogen bonds to Asp228 and Asn192 at the “northern rim” of S2 (Figure 2b, Figure S2). The two guanidino-phenoxyl groups and the guanidino group on the central cyclohexane ring are defined only by weaker electron density (see above). The lack of specific interactions for these parts of 1−2 might allow for some degree of flexibility. Weaker binding of 1−2, compared to 1−1, is also suggested by the lower occupancy of the molecule which was refined to 0.6 (occupancy of 1−1 was refined to 1.0). Consequently, even though concentrated inhibitor 1 (5 mM) was used for the soaking experiment, the binding site of 1−2 might still not be saturated. We therefore expect a much weaker affinity at binding site 2. One can also assume cooperative binding at this site because it is in part formed by molecule 1−1.

Table 1. Overview of Soaked 2,5 DideoxystreptamineDerived Furin Inhibitors inhibitor 1 (racemate) 2 (racemate) 3 (meso compound)

Ki (μM)a

soaking concentration (mM)

bound in structure

0.046 0.006 0.012

5 5 5

+ − −

a

Data taken from Jiao et al.11 (inhibitors 1, 2, and 3 were labeled as 1n, 1e, and 1g, respectively, in ref 11).

binding mode of 1 compared to 2 and 3. Binding of the latter inhibitors might involve interactions with regions that are covered by crystal packing. In addition, the binding kinetics of 2 or 3 in the crystalline state might be slower compared to those of 1, resulting in a very low occupation of their binding sites within the time frame of the soaking experiments. Alternatively, these inhibitors might also bind in multiple orientations and thus are invisible in the electron density despite an apparent higher affinity. Structure of Human Furin in Complex with Inhibitor 1. Crystals of human furin in complex with inhibitor 1 diffracted up to 1.9 Å (Table 2) and contained one molecule of Table 2. Data Collection and Refinement Statistics Data Collection wavelength (Å) 0.918 unit cell parameters P6522 a (Å), b (Å), c (Å) 132.2, 132.2, 155.7 resolution rangea (Å) 50.0−1.9 (2.02−1.90) Rsymmb (%) 13.1 (79.1) I/sigIa 11.0 (2.2) completenessa (%) 99.2 (97.6) observations: total/unique 311,107/63,025 Refinement Statistics Rwork/Rfree nonhydrogen atoms protein/inhibitor/other B-factors (Å2) overall/Wilson plot protein/inhibitor/other RMSD bond length (Å) RMSD bonded B-factors (Å2)

16.5/18.4 4,275 3695/86/494 23.0/25.5 21.6/35.1/31.6 0.007 4.1

Highest resolution shell is given in parentheses. bRsymm = ΣhΣl|Ihl − ⟨Ih⟩|/ΣhΣl|⟨Ih⟩| a

furin and two molecules of 1 in the asymmetric unit (Figure 1b). The structure is well-defined from amino acid Val109 to amino acid Ala574 of the enzyme, comprising the catalytic domain and the P domain of human furin. The complex structure was refined to a final R/Rfree of 16.5/18.5% with good stereochemistry (Table 2). Two molecules of inhibitor 1 bind in a well-defined manner at the “northern rim” (based on standard orientation) of the catalytic cleft of furin (Figure 1b,c,d; Figure S1). Both binding sites are unaffected by any crystal contacts. The characteristic molecular features of inhibitor 1 allowed only one placement of the defined atoms into the observed electron density map. The electron density map clearly shows that one methylene group of the central cyclohexane ring must be between the two large substituents of enantiomer (1R,2R,4S,5S)-1−1 (either O- or Nlinked guanidino-phenyl moieties), which clearly gives their 1213

DOI: 10.1021/acschembio.6b01110 ACS Chem. Biol. 2017, 12, 1211−1216

Letters

ACS Chemical Biology

Figure 2. Inhibition mechanism of inhibitor 1. (a and b) Closer view of the interactions mediated by molecules 1−1 and 1−2, respectively. Inhibitor 1 is shown as a ball and stick model with 1−1 in magenta and 1−2 in green. Important amino acids are given as a stick model in cyan. Important interactions are highlighted as dashed lines in black. The Fo−Fc annealed omit electron density map is shown as a dark green mesh contoured at 3σ for the side chains of the catalytic triade in a. (c) Alignment of the modeled furin:substrate complex (gray cartoon representation) with the furin:1 complex (golden cartoon representation) in stereo view. The peptide substrate and 1−1 are shown as ball and stick models in gray and magenta, respectively. The active site residues of the furin:substrate model and of the furin:1 complex are shown as stick models in gray and cyan, respectively. Canonical interaction sites of the substrate peptide are indicated from P5 to P2′.

The Binding Mode of Inhibitor 1 Is Different from Peptide-Like Substrates. Superimposing the structure of a peptide substrate modeled into the catalytic cleft of human furin (PDB-ID: 5JXH20) with that of inhibitor 1 bound to human furin (Figure 2c) shows an interesting mode of inhibition. The P2 and, to an even larger degree, the P4 side chain of the substrate would spatially collide with the 1−1 molecule of the bound inhibitor. Whereas this should diminish proper processing of the substrate by a competitive mechanism, we identified a second, rather unusual way of inactivating the protease. One of the guanidino-phenoxyl groups of the inhibitor molecule 1−1 pushes His194 out of its position. This is achieved largely by a rotation of ∼120° around the Cα− Cβ bond of His194 and is visible from the positional comparison of the amino acid residues taking part in the catalytic cycle. Whereas Asp153 shows a largely comparable conformation (side chain RMSD value 0.2 Å), the spatial arrangement of His194 significantly deviates between the catalytically competent and the inhibitor 1-inhibited states of furin (side chain RMSD value of ∼3.5 Å). In addition, Ser368 becomes flexible upon the binding of inhibitor 1. This is evidenced by an increased B value and the occurrence of a second conformation of its Oγ that now would not be positioned for a nuceleophilic attack toward the scissile peptide bond of a bound substrate as previously observed for the inactive unliganded state of the enzyme.20 In this way, the strong H-bonds between Asp153, His194, and Ser368 of the catalytic triade of furin get disrupted, and the respective proton shuttle mechanism is no longer possible. Implications for Next Generation Furin Inhibitors Based on Inhibitor 1. The inhibition mode brought about by 1, together with previous results,18,20 suggests several ways to improve not only the potency but also the specificity and

pharmacokinetics of future inhibitors of furin and other PCs. The specific contacts with furin’s active site observed for the two molecules of inhibitor 1 might be realized in an impoved compound. Combining the weaker contacts of molecule 1−2 (bound to Asp228 and Asn192) with the main contacts of molecule 1−1 (bound to Asp153, Asp264, Tyr308, Asp236) will probably increase the overall affinity of inhibitor 1. At the same time, the “side-arms” of 1−1 that do not contribute to specific contacts with the enzyme surface could be omitted. In addition, the molecular contacts identified here for inhibitor 1 might also be combined with other strong molecular contacts observed previously from the peptidic inhibitors with the amidomethyl-benzamidine group binding into S1 and the guanidino-benzyl group at P5.18,20 As the structures of the PCfamily members differ dramatically from one another outside the catalytic cleft and the S1−S4 area,21 alternative contact sites in this region might be beneficial for specificity. For this purpose, 1−2 could serve as a starting point to explore the interactions with the less conserved loop region, Asp228− Glu230, and even with other nonconserved residues of the enzyme. These interactions usually do not require polar or charged groups, thus incorporating them into new inhibitors might also help alleviate the current tendency toward highly positively charged compounds and, thereby, improve the often limited cell penetration capabilities of current inhibitors. Another inhibition strategy for furin, but also for other PCs that originate from the binding pose of 1−1, could be based on the observed deterioration of the active site residues, largely reminiscent of the mode of action of allosteric inhibitors. The respectively deformed enzyme would no longer be able to catalyze its intended reaction, whereas the inhibitor would not have to bind to the negatively charged active site cleft of the 1214

DOI: 10.1021/acschembio.6b01110 ACS Chem. Biol. 2017, 12, 1211−1216

Letters

ACS Chemical Biology

glycoproteins by the subtilisin-like endoprotease furin and its inhibition by specific peptidylchloroalkylketones. Biochimie 76, 217− 225. (6) Kacprzak, M. M., Peinado, J. R., Than, M. E., Appel, J., Henrich, S., Lipkind, G., Houghten, R. A., Bode, W., and Lindberg, I. (2004) Inhibition of furin by polyarginine-containing peptides: nanomolar inhibition by nona-D-arginine. J. Biol. Chem. 279, 36788−36794. (7) Levesque, C., Couture, F., Kwiatkowska, A., Desjardins, R., Guerin, B., Neugebauer, W. A., and Day, R. (2015) PACE4 inhibitors and their peptidomimetic analogs block prostate cancer tumor progression through quiescence induction, increased apoptosis and impaired neovascularisation. Oncotarget 6, 3680−3693. (8) Hardes, K., Becker, G. L., Lu, Y., Dahms, S. O., Kohler, S., Beyer, W., Sandvig, K., Yamamoto, H., Lindberg, I., Walz, L., von Messling, V., Than, M. E., Garten, W., and Steinmetzer, T. (2015) Novel Furin Inhibitors with Potent Anti-infectious Activity. ChemMedChem 10, 1218−1231. (9) Becker, G. L., Sielaff, F., Than, M. E., Lindberg, I., Routhier, S., Day, R., Lu, Y., Garten, W., and Steinmetzer, T. (2010) Potent inhibitors of furin and furin-like proprotein convertases containing decarboxylated P1 arginine mimetics. J. Med. Chem. 53, 1067−1075. (10) Lu, Y., Hardes, K., Dahms, S. O., Bottcher-Friebertshauser, E., Steinmetzer, T., Than, M. E., Klenk, H. D., and Garten, W. (2015) Peptidomimetic furin inhibitor MI-701 in combination with oseltamivir and ribavirin efficiently blocks propagation of highly pathogenic avian influenza viruses and delays high level oseltamivir resistance in MDCK cells. Antiviral Res. 120, 89−100. (11) Jiao, G. S., Cregar, L., Wang, J., Millis, S. Z., Tang, C., O’Malley, S., Johnson, A. T., Sareth, S., Larson, J., and Thomas, G. (2006) Synthetic small molecule furin inhibitors derived from 2,5-dideoxystreptamine. Proc. Natl. Acad. Sci. U. S. A. 103, 19707−19712. (12) Komiyama, T., Coppola, J. M., Larsen, M. J., van Dort, M. E., Ross, B. D., Day, R., Rehemtulla, A., and Fuller, R. S. (2009) Inhibition of furin/proprotein convertase-catalyzed surface and intracellular processing by small molecules. J. Biol. Chem. 284, 15729−15738. (13) Ramos-Molina, B., Lick, A. N., Blanco, E. H., Posada-Salgado, J. A., Martinez-Mayorga, K., Johnson, A. T., Jiao, G. S., and Lindberg, I. (2015) Identification of potent and compartment-selective small molecule furin inhibitors using cell-based assays. Biochem. Pharmacol. 96, 107−118. (14) Sielaff, F., Than, M. E., Bevec, D., Lindberg, I., and Steinmetzer, T. (2011) New furin inhibitors based on weakly basic amidinohydrazones. Bioorg. Med. Chem. Lett. 21, 836−840. (15) Scamuffa, N., Sfaxi, F., Ma, J., Lalou, C., Seidah, N., Calvo, F., and Khatib, A. M. (2014) Prodomain of the proprotein convertase subtilisin/kexin Furin (ppFurin) protects from tumor progression and metastasis. Carcinogenesis 35, 528−536. (16) Zhu, J., Declercq, J., Roucourt, B., Ghassabeh, G. H., Meulemans, S., Kinne, J., David, G., Vermorken, A. J., Van de Ven, W. J., Lindberg, I., Muyldermans, S., and Creemers, J. W. (2012) Generation and characterization of non-competitive furin-inhibiting nanobodies. Biochem. J. 448, 73−82. (17) Henrich, S., Cameron, A., Bourenkov, G. P., Kiefersauer, R., Huber, R., Lindberg, I., Bode, W., and Than, M. E. (2003) The crystal structure of the proprotein processing proteinase furin explains its stringent specificity. Nat. Struct. Biol. 10, 520−526. (18) Dahms, S. O., Hardes, K., Becker, G. L., Steinmetzer, T., Brandstetter, H., and Than, M. E. (2014) X-ray Structures of Human Furin in Complex with Competitive Inhibitors. ACS Chem. Biol. 9, 1113−1118. (19) Holyoak, T., Kettner, C. A., Petsko, G. A., Fuller, R. S., and Ringe, D. (2004) Structural basis for differences in substrate selectivity in Kex2 and furin protein convertases. Biochemistry 43, 2412−2421. (20) Dahms, S. O., Arciniega, M., Steinmetzer, T., Huber, R., and Than, M. E. (2016) Structure of the unliganded form of the proprotein convertase furin suggests activation by a substrate-induced mechanism. Proc. Natl. Acad. Sci. U. S. A. 113, 11196−11201.

protease, again providing possibilities to gain specificity in nextgeneration furin inhibitors.



METHODS



ASSOCIATED CONTENT

Details about expression, purification, and crystallization of human furin were described previously.18,20 Inhibitors 1, 2, and 3 were prepared following the procedures reported previously.11 Hexagonal crystals of unliganded furin were soaked with a 5 mM solution of inhibitor 1 in 100 mM MES at pH 5.5, 200 mM K/NaH2PO4, 1 mM CaCl2, 10% DMSO, and 3.4 mM NaCl for 16 h. Diffraction data for the flash cooled furin crystals in complex with inhibitor 1 were collected at the BESSY-II beamline 14.1 of the Helmholtz-Zentrum Berlin (HZB).22 Data processing was performed with XDS23 (v.03/ 2013) and programs of the CCP4-suite24 (CCP4 v.6.3.0, CCP4 interface v.2.2.0). Model building was carried out in COOT25 (v.0.6.2), and PHENIX26 (v.1.9−1692) was used for refinement. Parameter files of inhibitor 1 for refinement were generated with the PRODRG-server.27 PYMOL (http://www.pymol.org) was used for molecular graphics. A model of furin bound to the peptide H-ARGARG-VAL-ARG-ARG(↓)-SER-VAL-OH was obtained by manual docking in MAIN28 and geometry optimization in CNS29 (v.1.3.). Further details of the preparation of furin, of the crystallographic work, and of the structural analysis are described in the Supporting Information. S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acschembio.6b01110. Supporting figures (PDF) Accession Codes

Coordinates and structure factors of human furin in complex with inhibitor 1 have been deposited to the Protein Databank (PDB) with the accession code 5MIM.



AUTHOR INFORMATION

Corresponding Author

*Phone: +49-3641-656170. Fax: +49-3641-656335. E-mail: manuel.than@leibniz-fli.de. ORCID

Manuel E. Than: 0000-0002-4707-9225 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We acknowledge the Helmholtz Zentrum Berlin BESSY II for provision of synchrotron radiation at the beamline BL 14.1 and the scientific staff for assistance. We thank A. Johnson for helpful discussion and M. Perry for proof reading of the manuscript.



REFERENCES

(1) Artenstein, A. W., and Opal, S. M. (2011) Proprotein convertases in health and disease. N. Engl. J. Med. 365, 2507−2518. (2) Thomas, G. (2002) Furin at the cutting edge: from protein traffic to embryogenesis and disease. Nat. Rev. Mol. Cell Biol. 3, 753−766. (3) Basak, A. (2005) Inhibitors of proprotein convertases. J. Mol. Med. 83, 844−855. (4) Seidah, N. G., and Prat, A. (2012) The biology and therapeutic targeting of the proprotein convertases. Nat. Rev. Drug Discovery 11, 367−383. (5) Garten, W., Hallenberger, S., Ortmann, D., Schafer, W., Vey, M., Angliker, H., Shaw, E., and Klenk, H. D. (1994) Processing of viral 1215

DOI: 10.1021/acschembio.6b01110 ACS Chem. Biol. 2017, 12, 1211−1216

Letters

ACS Chemical Biology (21) Henrich, S., Lindberg, I., Bode, W., and Than, M. E. (2005) Proprotein convertase models based on the crystal structures of furin and kexin: explanation of their specificity. J. Mol. Biol. 345, 211−227. (22) Mueller, U., Darowski, N., Fuchs, M. R., Forster, R., Hellmig, M., Paithankar, K. S., Puhringer, S., Steffien, M., Zocher, G., and Weiss, M. S. (2012) Facilities for macromolecular crystallography at the Helmholtz-Zentrum Berlin. J. Synchrotron Radiat. 19, 442−449. (23) Kabsch, W. (2010) Xds. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 125−132. (24) Winn, M. D., Ballard, C. C., Cowtan, K. D., Dodson, E. J., Emsley, P., Evans, P. R., Keegan, R. M., Krissinel, E. B., Leslie, A. G., McCoy, A., McNicholas, S. J., Murshudov, G. N., Pannu, N. S., Potterton, E. A., Powell, H. R., Read, R. J., Vagin, A., and Wilson, K. S. (2011) Overview of the CCP4 suite and current developments. Acta Crystallogr., Sect. D: Biol. Crystallogr. 67, 235−242. (25) Emsley, P., Lohkamp, B., Scott, W. G., and Cowtan, K. (2010) Features and development of Coot. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 486−501. (26) Adams, P. D., Afonine, P. V., Bunkoczi, G., Chen, V. B., Davis, I. W., Echols, N., Headd, J. J., Hung, L. W., Kapral, G. J., GrosseKunstleve, R. W., McCoy, A. J., Moriarty, N. W., Oeffner, R., Read, R. J., Richardson, D. C., Richardson, J. S., Terwilliger, T. C., and Zwart, P. H. (2010) PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr., Sect. D: Biol. Crystallogr. 66, 213−221. (27) Schuttelkopf, A. W., and van Aalten, D. M. (2004) PRODRG: a tool for high-throughput crystallography of protein-ligand complexes. Acta Crystallogr., Sect. D: Biol. Crystallogr. 60, 1355−1363. (28) Turk, D. (2013) MAIN software for density averaging, model building, structure refinement and validation. Acta Crystallogr., Sect. D: Biol. Crystallogr. 69, 1342−1357. (29) Brunger, A. T. (2007) Version 1.2 of the Crystallography and NMR system. Nat. Protoc. 2, 2728−2733.

1216

DOI: 10.1021/acschembio.6b01110 ACS Chem. Biol. 2017, 12, 1211−1216