11
proteins STRUCTURE O FUNCTION O BIOINFORMATICS The crystal structure of the secreted aspartic proteinase 3 from Candida albicans and its complex with pepstatin A Claudia Borelli, 1 * Elisabeth Ruge, 1 * Martin Schaller, 2 Michel Monod, 3 Hans Christian Korting, 1 Robert Huber, 4 and Klaus Maskos 4 1 Department of Dermatology and Allergy, Ludwig Maximilian University of Munich, Frauenlobstr. 9–11, 80337 Munich, Germany 2 Department of Dermatology, Eberhard Karl University of Tuebingen, Liebermeisterstr. 25, 72076 Tu ¨bingen, Germany 3 Service de Dermatologie (DHURDV), Laboratoire de Mycologie, BT422, Centre Hospitalier Universitaire Vaudois, 1011 Lausanne, Switzerland 4 Max-Planck-Institut fu ¨r Biochemie, Abteilung Strukturforschung, Am Klopferspitz 18a, 82152 Martinsried, Germany INTRODUCTION Aspartic proteinases are found in plants, retroviruses, mammals, and fungi. There are several distinct groups, including pepsins, chymosins, cathepsins, renins, retroviral pro- teinases, and fungal aspartic proteinases. 1, 2 The role of the secreted aspartic proteinases (Saps) of Candida albicans (C. albicans) encoded by 10 SAP genes has been intensively characterized in the last years by several groups. 3 The Saps are essential for the fungal nutrition process and are important virulence factors for localized and disseminated can- didal infections. 4,5 Sap2 has a very broad spectrum of activity degrading many human proteins, such as extracellular matrix, keratin, collagen, and many others. 6 These sub- strates can either be macromolecular or of low molecular weight. 7 Saps are differentially regulated during distinct stages of the infection process. Sap1–3 appear to play a role in the adherence and tissue damage of localized infection, whereas Sap4–6 may be of im- portance in systemic disease. 8–10 The different roles of the Sap1–3 and Sap4–6 isoen- zyme subgroups are explained by variations in amino acid sequences as well as by differ- ent enzymatic characteristics, such as the optimum pH and the net charge. The role of Sap9 and Sap10 for cell surface integrity, cell separation, and adhesion has been recently characterized, while almost nothing is known about Sap7 and Sap8. 11,12 Insights into the contributions of the Sap family to C. albicans pathogenicity have also been obtained studying Sap-deficient C. albicans strains; Sap1–3 have been implicated in mucosal infec- tions and Sap4–6 in systemic infections. 3 Based on the role of Saps as virulence factors, therapy or even better prevention of C. albicans infections in immunocompromised patients (e.g., patients with AIDS, organ transplants, and cancer) should be feasible using specific Sap inhibitors. 13 A crucial element in the process of Sap-specific inhibitor design is the determination of the exact enzyme structures, which is only known for Sap2. 14–16 Sap2 is the most abundantly secreted protein in vitro when grown in the presence of protein as the sole source of nitrogen, 17–19 and was the first member of the Sap gene family to be structurally studied as complexes with pepstatin A (pepA) and A- J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 1 ID: kannanb Date: 9/5/07 Time: 03:43 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114 Grant sponsor: The OPO-Stiftung, Zurich, Switzerland. Claudia Borelli and Elisabeth Ruge contributed equally to this work. M.S. was supported by the Deutsche Forschungsge- meinschaft (Sch 897/1, Sch 897/3). *Correspondence to: Claudia Borelli or Elisabeth Ruge, Department of Dermatology and Allergy, Ludwig Maximilian University of Munich, Frauenlobstr. 9-11, 80337 Munich, Germany. E-mail: [email protected] or [email protected]/[email protected] Received 29 September 2006; Revised 16 December 2006; Accepted 11 January 2007 Published online 00 Month 2007 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/prot.21425 ABSTRACT The family of secreted aspar- tic proteinases (Sap) encoded by 10 SAP genes is an impor- tant virulence factor during Candida albicans (C. albi- cans) infections. Antagonists to Saps could be envisioned to help prevent or treat candido- sis in immunocompromised patients. The knowledge of several Sap structures is cru- cial for inhibitor design; only the structure of Sap2 is known. We report the 1.9 and 2.2 A ˚ resolution X-ray crystal structures of Sap3 in a stable complex with pepstatin A and in the absence of an inhibitor, shedding further light on the enzyme inhibitor binding. In- hibitor binding causes active site closure by the movement of a flap segment. Comparison of the structures of Sap3 and Sap2 identifies elements re- sponsible for the specificity of each isoenzyme. Proteins 2007; 68:000–000. V V C 2007 Wiley-Liss, Inc. Key words: Candida albicans; inhibitor design; pepstatin A; secreted aspartic proteinase; X-ray crystal structure. V V C 2007 WILEY-LISS, INC. PROTEINS 1

The crystal structure of the secreted aspartic proteinase ... · portance in systemic disease.8–10 The different roles of the Sap1–3 and ... reaching full size within ... Synchrotron,

Embed Size (px)

Citation preview

proteinsSTRUCTURE O FUNCTION O BIOINFORMATICS

The crystal structure of the secretedaspartic proteinase 3 from Candida albicansand its complex with pepstatin AClaudia Borelli,1* Elisabeth Ruge,1* Martin Schaller,2 Michel Monod,3 Hans Christian Korting,1

Robert Huber,4 and Klaus Maskos4

1Department of Dermatology and Allergy, Ludwig Maximilian University of Munich, Frauenlobstr. 9–11, 80337 Munich,

Germany2

Department of Dermatology, Eberhard Karl University of Tuebingen, Liebermeisterstr. 25, 72076 Tubingen, Germany3

Service de Dermatologie (DHURDV), Laboratoire de Mycologie, BT422, Centre Hospitalier Universitaire Vaudois,

1011 Lausanne, Switzerland4

Max-Planck-Institut fur Biochemie, Abteilung Strukturforschung, Am Klopferspitz 18a, 82152 Martinsried, Germany

INTRODUCTION

Aspartic proteinases are found in plants, retroviruses, mammals, and fungi. There are

several distinct groups, including pepsins, chymosins, cathepsins, renins, retroviral pro-

teinases, and fungal aspartic proteinases.1, 2 The role of the secreted aspartic proteinases

(Saps) of Candida albicans (C. albicans) encoded by 10 SAP genes has been intensively

characterized in the last years by several groups.3 The Saps are essential for the fungal

nutrition process and are important virulence factors for localized and disseminated can-

didal infections.4,5 Sap2 has a very broad spectrum of activity degrading many human

proteins, such as extracellular matrix, keratin, collagen, and many others.6 These sub-

strates can either be macromolecular or of low molecular weight.7 Saps are differentially

regulated during distinct stages of the infection process. Sap1–3 appear to play a role in

the adherence and tissue damage of localized infection, whereas Sap4–6 may be of im-

portance in systemic disease.8–10 The different roles of the Sap1–3 and Sap4–6 isoen-

zyme subgroups are explained by variations in amino acid sequences as well as by differ-

ent enzymatic characteristics, such as the optimum pH and the net charge. The role of

Sap9 and Sap10 for cell surface integrity, cell separation, and adhesion has been recently

characterized, while almost nothing is known about Sap7 and Sap8.11,12 Insights into

the contributions of the Sap family to C. albicans pathogenicity have also been obtained

studying Sap-deficient C. albicans strains; Sap1–3 have been implicated in mucosal infec-

tions and Sap4–6 in systemic infections.3 Based on the role of Saps as virulence factors,

therapy or even better prevention of C. albicans infections in immunocompromised

patients (e.g., patients with AIDS, organ transplants, and cancer) should be feasible

using specific Sap inhibitors.13 A crucial element in the process of Sap-specific inhibitor

design is the determination of the exact enzyme structures, which is only known for

Sap2.14–16 Sap2 is the most abundantly secreted protein in vitro when grown in the

presence of protein as the sole source of nitrogen,17–19 and was the first member of the

Sap gene family to be structurally studied as complexes with pepstatin A (pepA) and A-

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 1

ID: kannanb Date: 9/5/07 Time: 03:43 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Grant sponsor: The OPO-Stiftung, Zurich, Switzerland.

Claudia Borelli and Elisabeth Ruge contributed equally to this work. M.S. was supported by the Deutsche Forschungsge-

meinschaft (Sch 897/1, Sch 897/3).

*Correspondence to: Claudia Borelli or Elisabeth Ruge, Department of Dermatology and Allergy, Ludwig Maximilian

University of Munich, Frauenlobstr. 9-11, 80337 Munich, Germany.

E-mail: [email protected] or [email protected]/[email protected]

Received 29 September 2006; Revised 16 December 2006; Accepted 11 January 2007

Published online 00 Month 2007 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/prot.21425

ABSTRACT

The family of secreted aspar-

tic proteinases (Sap) encoded

by 10 SAP genes is an impor-

tant virulence factor during

Candida albicans (C. albi-

cans) infections. Antagonists

to Saps could be envisioned to

help prevent or treat candido-

sis in immunocompromised

patients. The knowledge of

several Sap structures is cru-

cial for inhibitor design; only

the structure of Sap2 is

known. We report the 1.9 and

2.2 A resolution X-ray crystal

structures of Sap3 in a stable

complex with pepstatin A and

in the absence of an inhibitor,

shedding further light on the

enzyme inhibitor binding. In-

hibitor binding causes active

site closure by the movement

of a flap segment. Comparison

of the structures of Sap3 and

Sap2 identifies elements re-

sponsible for the specificity of

each isoenzyme.

Proteins 2007; 68:000–000.VVC 2007 Wiley-Liss, Inc.

Key words: Candida albicans;

inhibitor design; pepstatin A;

secreted aspartic proteinase;

X-ray crystal structure.

VVC 2007 WILEY-LISS, INC. PROTEINS 1

70450.14 A-70450 was designed by Abbott to act against

renin but is also able to inhibit C. albicans aspartic pro-

teinases.20 The classical aspartic proteinase inhibitor

pepA is a microbial pentapeptide produced by Streptomy-

ces.21 The crystal structure of the closely related clinical

isolate Sap2X with an amino acid sequence identity to

Sap2 of 96% in complex with A-70450 was reported

later.20 Knowledge of the structures of further Sap isoen-

zymes will facilitate specific structure-based inhibitor

design. Here, we present the X-ray crystal structure of

Sap3 at 1.9 and 2.2 A resolutions in a stable complex with

pepA and in the absence of the inhibitor. We also com-

pare the structures of Sap3 and Sap2.

MATERIALS AND METHODS

Expression of the Sap isoenzymesin Pichia pastoris

Sap3 was expressed as a recombinant protein using the

Pichia pastoris (P. pastoris) expression system.22 In short,

P. pastoris transformants were grown to cell densities

near saturation (OD600 ¼ 20) at 308C in 1000 mL of

glycerol-based yeast media [0.1M potassium phosphate

buffer at pH 6.0, containing 1% (w/v) yeast extract, 2%

(w/v) peptone, 1.34% (w/v) YNB without amino acids

(Difco, Sparks), 1% (w/v) glycerol and 4 3 10�5% (w/v)

biotin]. Cells were harvested and resuspended in 200 mL

of the same medium containing 0.5% (w/v) methanol,

instead of glycerol, and incubated for 2 days. Thereafter,

the supernatant was harvested by centrifugation.

Purification of recombinant proteinases

400 milliliters of P. pastoris culture supernatant of

Sap3 was concentrated to 100 mL using an Amicon

concentrating cell (10 kDa MWCO). For desalting, the

concentrate was dialysed against a 100-fold volume of

10 mM citric acid/NaOH pH 7.0. Sap3 was first purified

by anion-exchange. Therefore the dialysate was loaded on

a DEAE-column (DE 52, Whatman1), washed with

10 mM citric acid/NaOH pH 7.0, and eluted with a

gradient running from 10 to 100 mM citric acid/NaOH

at the same pH.

Pooled fractions were concentrated to a volume of

1.0 mL and further purified on a gel filtration-column

(Superdex 75 HiLoad 26/60 prep grade, Amersham

Pharmacia Biotech, Piscataway) using a 10 mM citric

acid/NaOH containing 150 mM NaCl, pH 7.0.

Finally, the protein was transferred into 10 mM MES/

NaOH, pH 6.5 via a NapTM10 column (Amersham

Pharmacia Biotech).

Protein extract analysis

Protein concentrations were calculated from the absorp-

tion of protein at 280 nm (OD280) by using the extinction

coefficient (Abs 0.1% at 280 nM ¼ 0.949 M�1 cm�1) as

given in Expasy—ProtParam Tool. Protein extracts were

analyzed by SDS-PAGE23 with a separation gel of 12%

polyacrylamide. Gels were stained with Coomassie

brilliant blue R-250 (Bio-Rad).

Peptide mass fingerprinting

The identity of Sap3 was confirmed by its peptide

mass fingerprint. The putative band in SDS-PAGE con-

taining the protein was analyzed by trypsin digestion

followed by MALDI-TOF mass spectrometry. Measured

peptide masses were compared with peptide masses from

an in silico digestion of the protein database using the

search engine MASCOT.

Activity assays

The proteolytic activity of Sap3 was measured with

0.02% resorufin-labeled casein as substrate at the pH

optimum in citric acid/NaOH (50 mM citric acid/NaOH,

pH 3.5)22 in a total volume of 0.1 mL. After incubation

at 378C for 30 min, the undigested substrate was precipi-

tated with trichloroacetic acid (5% final concentration)

and separated from the supernatant by centrifugation.

The absorbance of the supernatant was measured in the

alkaline range at 574 nm after addition of 100 lL 1M

Tris/HCl, pH 10.0. For practical purposes, one unit of

Sap activity was defined as that producing an absorbance

of 0.001/min.

Crystallization

Samples recovered from buffer exchange were concen-

trated to about 10 mg/mL using Centricon 10 filtration

units (Amicon, Charlotte). PepA (Sigma) was added to

samples in a 1.5-fold molar excess. Crystals appeared in

the presence of Zn2þ (10 mM zinc acetate). Sap3 without

inhibitor crystallized in the presence of 0.2M potassium

bromide, 8% PEG 20000, 8% PEG 550MME, and 0.1M

cacodylic acid/NaOH, pH 6.5. Sap3 in complex with

pepA formed crystals in the presence of reservoir buffer

containing 0.2M potassium bromide, 15% PEG 4000,

and 0.1M cacodylic acid/NaOH, pH 6.5.

The crystals grew within a few days after the addition

of 100 lL 5M NaCl at day 3 to the reservoir buffer (300 lL)

reaching full size within several weeks.

For data collections, crystals were shock frozen under

a stream of nitrogen at 100 K (Oxford Cryosystems

Cryostream).

Data collection, structure solution,and refinement

X-ray data were collected using synchrotron radiation

at the beamline BW6 of the DESY (Deutsches Elektronen

Synchrotron, Hamburg, Germany) using a Mar Research

CCD detector (Mar Research, Hamburg, Germany). Two

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 2

ID: kannanb Date: 9/5/07 Time: 03:43 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

C. Borelli et al.

2 PROTEINS DOI 10.1002/prot

complete native data sets to 1.8 and 2.2 A were collected

from the Sap3–pepA complex and the native enzyme,

respectively, using X-ray radiation with a wavelength of

1.05 A. Both crystals were of space group P3221 and con-

tained one molecule per asymmetric unit.

Diffraction data were processed and scaled with the

programs DENZO/SCALEPACK.24 Model building was

performed using the modeling program O.25 Statistics

for the data collection and refinement are summarized in

TableT1 I.

The Sap3–pepA complex was solved by molecular re-

placement with the program Molrep,26 where the struc-

ture of Sap2X served as a search model. One solution

was obtained resulting in a correlation coefficient of

50.4% and an R-factor of 45.2%.

The calculation of the electron density map and the

crystallographic refinement were performed using CNS.27

During the last steps of refinement, one zinc atom per

molecule was tentatively built into the density occurring

between residues His131Sap3 and Asp191Sap3 of one

Sap3 molecule and residues His197Sap3 and Asp214Sap3

of a symmetrically related neighbor molecule. Three hun-

dred eighteen water molecules were automatically built

with CNS27 and pepA was modeled into the extra

density extending along the active site cleft. The model

of the Sap3–pepA complex was refined to 1.9 A to a final

Rfree of 25.2% and an R-factor of 23.9%.

The final model of the Sap3–pepA complex was used

as a search model for the replacement by Molrep26 for

the dataset of the apoenzyme of Sap3. The replacement

resulted in one solution with a correlation factor of

64.8% and an R-factor of 38.6%. Refinement of this

model was done in the same way as for the pepA-com-

plexed structure. Two hundred fifty water molecules were

automatically built with CNS.27 Refinement was per-

formed to 2.2 A to a final Rfree of 26.1% and an R-factor

of 23.5%.

The quality of measured data did not allow for refine-

ment to lower R-factors despite the huge effort put into

structure refinement.

The main-chain angles of the refined structures were

analyzed by the program PROCHECK28 and plotted in a

Ramachandran diagram.29 Accordingly, 99.3% of all

main-chain angles in each structure exhibit most favored

or additionally allowed conformations, while residues

Ala133 and Asn160 exhibit a generously allowed confor-

mation in both structures.

Accession codes

The coordinates of the Sap3 apoenzyme and Sap3

complexed with pepA have been deposited at the RCSB

Protein Data Bank under the accession codes 2H6S and

2H6T, respectively.

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 3

ID: kannanb Date: 9/5/07 Time: 03:44 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Table IStatistics for Data Collection and Refinement

Data collectionCrystal content Sap3-pepstatin A complex Sap3 apoenzymeSpace group P3221 P3221Cell constants (�) a ¼ 61.10, b ¼ 61.10, c ¼170.74 a ¼ 61.35, b ¼ 61.35, c ¼171.98Resolution range (�) 20–1.8 20–2.2Reflections measured 133,783 71,472l/r 13.7 18.4Unique reflections 35,196 19,792Rmerge (%)a (last shell, 1.86–1.80/2.28–2.20) 5.0 (34.8) 6.0 (16.8)Completeness (%) (last shell, 1.86–1.80/2.28–2.20) 99.3 (100.0) 99.8 (99.7)

RefinementResolution 19.86–1.9 19.95–2.2Reflections used for refinement 28,626 18,790Solvent content (%, v/v) 51.13 51.88Residue number per asymmetric unit 340 340Water molecules per asymmetric unit 318 250Zinc ions per asymmetric unit 1 1Rfactor (%)b/Rfree (%)c 23.9/25.2 23.40/26.0rmsd bond length (�) 0.009 0.008rmsd bond angles (deg.) 1.7 1.7Average B-factors

Overall (�2) 28.949 33.303Protein atoms (�2) 27.996 32.834Water molecules (�2) 36.604 38.152Zinc ion (�2) 20.97 28.00Inhibitor atoms (�2) 29.378 n.d.

aRmerge ¼ Shkl jhIi � Ij/Shkl jIj.bRfactor ¼ Shkl jjFobsj � jFcalcjj//Shkl jFobsj.cRfree is the cross-validation R-factor computed for the test set of reflections (5.0% of total).

Crystal Structure of SAP3 and its complex with PepA

DOI 10.1002/prot PROTEINS 3

RESULTS AND DISCUSSION

With the establishment of the recombinant P. pastoris

expression system, we obtained a useful tool to produce

an isolated Sap3 protein in sufficient amounts for crystal-

lization trials.22 The protein could be produced with a

yield of up to 80 mg L�1.

The protein crystallizes in the trigonal space group

P3221 containing one molecule per asymmetric unit.

The superposition of the Sap3 apoenzyme (aSap3) and

the pepA–Sap3 complex (pSap3) exhibits an rms devia-

tion of all 340 Ca atoms of 0.70 A. The largest confor-

mational change of Sap3 on binding pepA occurs

between residues 81 and 91, that is, the active site flap

residues.

During the course of refinement, well-defined density

occurred between residues His131 and Asp191 of one

Sap3 molecule and residues His197 and Asp214 of a

symmetrically related neighboring molecule. The residues

build a tetrahedral coordination sphere with the liganded

atom. Sap3 forms ordered and well diffracting crystals

only in the presence of zinc. As no other cations were

added during crystallization, a zinc ion was tentatively

built into the density.

Overall structure of Sap3–Sap3 and Sap2,members of the aspartic proteinase family

Sap3 structurally appears as a kidney-shaped bilo-

bed globular protein mainly consisting of b-strands

[Fig. F11(A)] clearly divided into an N-terminal and a C-

terminal domain. Each domain contributes one catalytic

aspartic acid to the active site.32 The active site cleft is

partly covered by an antiparallel b-hairpin (s9,8) com-

prising residues 81–91, which is commonly known as the

active site flap in aspartic acid proteinases15,32,33 and

which plays an important role in inhibitor or substrate

binding.

Sap2/Sap2X—the closely related clinical isolate to

Sap2—and Sap3 show very high sequence identity [74%/

71%, Fig. 1(B)]. The total electrostatic charges of both

Sap2 and Sap3 are highly negative with a net charge of

�21 and �20, respectively.34 The optimal superposition

of Sap3 with Sap2 reveals topological equivalence of

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 4

ID: kannanb Date: 9/5/07 Time: 03:44 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Figure 1Structure of Sap3 complexed with pepstatin A (pepA). (A) Stereo overall view of Sap3 shown in ribbon structure. The side chains of the two catalytic aspartates are

shown in blue, the two disulphide bridges (Cys47–Cys59, Cys256–Cys294) are displayed in orange. Sap3 structurally appears as a kidney-shaped bilobed globular protein

mainly consisting of b-strands. The structure clearly divides into an N-terminal (bottom) and a C-terminal (top) domain. The figure was generated with Pymol.30 (B)

Sequence alignment of Sap3 and Sap2/Sap2X together with typical mammalian and fungal members of the aspartic proteinase family. b-Sheets, a-helices and loops found

in Sap3 are represented by arrows, cylinders, and lines, respectively. The catalytic aspartic acids are highlighted in orange. Residues that are identical within the sequences

of Sap3 and Sap2/Sap2X are colored in green. The amino acids highlighted in cyan exhibit 100% identity within the sequences of all aspartic proteinases listed. Marked

with black and white dots are those residues which build hydrogen bonds and hydrophobic interactions, respectively, with pepA when complexed with Sap3. The

numbering system is based on Sap2. The figure was created using ClustalW.31 (C) Superimposition of structures Sap2 and Sap3. Sap2 is shown in green whereas Sap3 is

colored in yellow. PepA bound to Sap3 is shown as sticks (purple). The side chains of the two catalytic aspartates of Sap3 are shown in blue. The salt bridge formation

between Lys129Sap3 and Asp37Sap3 subsequent to the loop movement that stabilizes this loop conformation is indicated as black dashed lines. In contrast to Sap3, the

loop in Sap2 is directed towards the active site cleft. While Asn131Sap2 and Ala133Sap2 are part of the S20 binding site in Sap2, Lys129Sap3, and Glu132Sap3 contribute

to S20 of Sap3. The loop movement of Sap3 may be supported by the observed Zn-binding.

COLOR

C. Borelli et al.

4 PROTEINS DOI 10.1002/prot

a-carbon atoms of 337 out of 340 with an rms deviation

of 1.594 A. Indeed, most secondary structural elements

are highly conserved within Sap2 and Sap3. With their

bilobed, mainly b-sheeted secondary structure, the two

enzymes Sap2 and Sap3 follow the classical aspartic pro-

teinase fold observed in several mammalian aspartic pro-

teinases (e.g., pepsin,33 renin,35 cathepsin D36) and fun-

gal aspartic proteinases (e.g., penicillopepsin,37 rhizopus-

pepsin38). Nevertheless, several features detected by

structure superposition make them unique among the

aspartic proteinase family.15, 16 The most striking dif-

ference from other aspartic proteinases is the disulfide

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 5

ID: kannanb Date: 9/5/07 Time: 03:44 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Figure 1(Continued)

COLOR

Crystal Structure of SAP3 and its complex with PepA

DOI 10.1002/prot PROTEINS 5

loop enclosed by Cys47 and Cys59, which builds up the

N-terminal loop (N-ent loop, i.e., s9–8 loop). Stewart

and Abad-Zapatero13 have specified this new flap as a

second active site flap. Together with the N-ent loop, the

C-terminal loops (C-ent loops, i.e., s19–20 loop and s24–

25 loop) form a funnel-like extended active site. Because

of high B-factors and weak density observed during

structure refinement, the N-terminal s9–8 loop and the

C-terminal s19–20 loop are likely highly flexible. A 7–8

residue deletion enlarges the substrate binding sites S1

and S3 because of the loss of a small helical segment.

Two short insertions corresponding to residues 247–249

and residues 286–290, respectively, enlarge the C-ent

loops. At the end of the carboxy terminus, an addition of

about 10 residues is special for Sap2 and Sap3.

As a result of the changes regarding the active site

cleft, both Sap2 and Sap3 comprise an enlarged and

more extended binding site compared with the other

members of the aspartic proteinase family. The conserva-

tion of these structural features in Sap2 and Sap3 identi-

fies them as signatures of a subfamily. Nevertheless, some

conformational changes between Sap2 and Sap3 can be

observed, which have functional significance.

Overall conformational differencesbetween Sap3 and Sap2

The amino acid sequence of Sap3 following the align-

ment of Stewart et al.34 shows two deletions (residues 51

and 212) compared with Sap2. At residue 212 (Sap2

numbering), the structural change causes a truncation of

a small loop in s17 in Sap3, whereas the secondary struc-

ture is retained. A more dramatic change in the N-ent

loop is caused by the second deletion of Tyr51 in Sap3

leading to a less restricted entrance to the active site cleft

of Sap3, and this partly contributed by residues with

smaller side chains such as Ala (Val in Sap2) or Gly (Thr

and Asp in Sap2).

Presumably the most striking difference in Sap3 is

caused by the replacement of the Sap2 residues Asn131-

Sap2 and Gly134Sap2 by His131Sap3 and Glu134Sap3 in

Sap3. More favored interactions with solvent molecules

of those residues apparently cause the movement of the

loop comprising residues 129–135 to the solvent exposed

area. The salt bridge formation between Lys129Sap3 and

Asp37Sap3 subsequent to the loop movement stabilizes

this loop conformation. In contrast to Sap3, the loop in

Sap2 is directed towards the active site cleft and contrib-

utes to S20 [Fig. 1(C)]. Consequences of the loop transla-

tion for the S20 binding site in Sap3 will be discussed in

detail in the next sections. Although the new loop con-

formation seems to be stabilized by the salt bridge for-

mation, it may also be supported by the Zn-binding,

which links neighboring molecules in the crystal and is

unlikely to be of physiological relevance.

Active site, substrate specificity,and inhibitor binding

Active site and substrate binding site

The active site of Sap3, which is located in a deep cleft

of the protein is highly negatively charged. Following

standard orientation, the substrate binding sites S4 to S20

are aligned from the upper left to the lower right side of

Sap3. The N-terminal front of the active site cleft is cov-

ered by the active site flap providing supplementary resi-

dues for substrate binding (Fig. F22).

The active site cleft of the Sap3 apoenzyme is filled

with water molecules mainly clustering around the cata-

lytic aspartates Asp32 and Asp218 (Fig. 2). Water mole-

cule W50 directly binds to both catalytic aspartates

Asp32 and Asp218 surrounded by other water molecules,

which make contacts to active site residues Gly34 (O),

Gly220 (O), Thr222 (OG1). Notably, just a few water

molecules are in contact with the residues of the active

site flap. The residues Tyr84 (OH) and Gly85 (N) which

are important hydrogen bonding partners for bound

inhibitors/substrates are ligated to the water molecules

W170 and W5.

Peptides of seven to nine residues might be accommo-

dated in the substrate binding region of Sap3. Relatively

large binding sites are found at the edges (S4, S3, S20),

while those lying in the center region appear to be

smaller (Fig. F33). Interestingly, the N-terminal binding

sites S4–S1 contain only hydrophobic and/or acidic

amino acids, while the only two basic amino acids, that is,

Arg195 and Lys129, belong to the primed site (Table T2II).

Altogether the active site of Sap3 retains a strong nega-

tive overall charge.

The substrate binding sites of Sap2 and Sap3 appear

quite similar in their character, sizes, and shapes. Sev-

enty-five percent of the amino acids that form the sub-

strate binding site are conserved within the sequences of

Sap2 and Sap3 [Table II, Fig. 1(B)]. A detailed analysis—

illustrated by the superposition of Sap2 and Sap3 in Fig-

ure 3—reveals some differences though. Compared with

Sap2, the polar character of the S4 binding site in Sap3 is

reduced by one polar side chain and its size is enlarged

with the exchange of Asp299Sap2 by Gly299Sap3.

S3 represents a notably large binding site compared

with the known mammalian and fungal aspartic protein-

ases, which are conserved within both enzyme structures.

Substitution of Thr13Sap2 by Ser13Sap3 and Ser88Sap2

by Thr88Sap3 can be detected.

Based on modeling studies of Sap isoenzymes Sap1–6,

Stewart et al.34 proposed that the exchange of Ala303-

Sap2 by Tyr303Sap3 limits the access to S2 and/or S10 to

residues with small side chains, such as Val, Ala, and Thr.

The crystal structure, however, shows that Tyr303Sap3

points away from the S2 site. It contributes to the forma-

tion of a ‘‘new’’ binding site—S30. Since S2 already con-

tains many polar and hydrophilic residues (Asp86,

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 6

ID: kannanb Date: 9/5/07 Time: 03:44 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

C. Borelli et al.

6 PROTEINS DOI 10.1002/prot

Thr221, and Tyr225) in both Sap2 and Sap3, the amino

acid change from Asn301Sap2 to Ser301Sap3 might have

little influence on the overall character of S2.

The S1 site shows most differences with three residues

out of six exchanged (Table II). The replacement of

Ile30Sap2 and Ile119Sap2 by the smaller Val30Sap3 and

Val119Sap3, respectively, enlarge it. Consequences of the

substitution of Ser88Sap2 by Thr88Sap3 for substrate

specificity will be discussed in the next section.

The modification of S20 subsequent to the loop turn

comprising residues 129–135 represents the most striking

difference between the two structures of Sap2 and Sap3.

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 7

ID: kannanb Date: 9/5/07 Time: 03:44 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Figure 3Stereo superimposition of the active site of Sap3 (yellow) and Sap2 (green). Residues that are identical in both sequences

are shown as lines while distinct residues are displayed as sticks. The catalytic aspartates of Sap3 are shown in cyan and

those of Sap2 are colored in orange. For clarity reasons, the bound inhibitors (pepA in Sap3, A-70450 in Sap2) are not

included. Figure made with Pymol.30

Figure 2Stereo plot of the substrate binding site of the Sap3 apoenzyme. Water-filled active site of Sap3. Waters are shown in

green, the catalytic aspartates are highlighted in cyan, and surrounding residues are shown in yellow. Black dashed lines

represent hydrogen bonding between active site residues and water molecules, while orange dashed lines indicate an

interaction within water molecules. Residues belonging to the active site loop are shown in grey. W50 is tightly bound to

both catalytic aspartic residues; W5 and W170 form hydrogen bonds with the active site flap residues Gly85 and Tyr84,

respectively. The figure was made with Pymol.30

COLOR

COLOR

Crystal Structure of SAP3 and its complex with PepA

DOI 10.1002/prot PROTEINS 7

While both Asn131Sap2 and Ala133Sap2 are part of the

S20 binding site in Sap2, the analog residues in Sap3

point toward the solvent exposed area, and Lys129Sap3

and Glu132Sap3 contribute to S20 of Sap3 instead [Fig.

1(C)]. With these ‘‘new’’ charged residues, the rather

neutral character of S20 in Sap2 changes to a more polar

one in Sap3. Strikingly, the loop turn does not only max-

imize the size of the S20 binding site but even enables

access to some deeper located aromatic amino acids, that

is, Phe80Sap3, Trp94Sap3, and Tyr137Sap3 (Fig. 3).

Thus, the cavity of S20 in Sap2 structure changes to a

channel in Sap3 with a polar entrance and an aromatic

base. The alteration of the S20 binding site was not

caused by substitution within the amino acid sequence

of the particular residues which arrange to the S20 site

but by changes in adjacent residues.

Substrate specificity

Aspartic proteinases usually display preference of

cleavage between large hydrophobic amino acids such as

Phe–Phe, but fungal aspartic proteinases have been

shown to be distinct from mammalian enzymes in that

they cleave substrates with polar residues such as His or

Lys in the P1 position.39 Only one study on substrate

specificities at the P1 and P10 sites for the different Sap

isoenzymes Sap1, 2, 3, and 6 exist.40 Koelsch et al.40

have observed that Sap3 significantly prefers Leu closely

followed by Arg at P1. By cleaving after 10 different

amino acids with the preference (Leu > Arg >> Phe >Glu > Asp > Tyr > Lys > Gln > Ala > Asn), Sap3 dis-

plays less specificity in its P1 position than Sap2. The

preferred residue for Sap2 in the P1 position is Phe or

Leu, but it also cleaves after Lys and Tyr in the order

(Phe > Leu > Lys > Tyr). The slightly larger Sap3-S1

site might be one reason for its broader range of sub-

strate specificity. Another reason might be the conforma-

tional alteration of residue 88. In both structures, residue

88 is located at the edge of S1 and S3, thus contributing

to both sites. While the hydroxyl group of Ser88Sap2

contributes to S3 in Sap2, the hydroxyl group of

Thr88Sap3 sticks into S1 in Sap3. Therefore, the addi-

tional hydroxyl group of Thr88Sap3 in the Sap3-S1 site

may contribute to the preference of Arg, which has its

highly polar and space filling guanidinium group flanked

by the carboxyl group of Asp86Sap3 from one side and

by the hydroxyl group of Thr88Sap3 from the other.

The P10 specificities of Sap2 and Sap3 are similar and

very broad comprising all amino acids except Asn, His,

Arg, Lys, and Pro. This is expected, as the residues form-

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 8

ID: kannanb Date: 9/5/07 Time: 03:45 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Table IISap3 and Sap2 Residues Forming the Substrate Binding Site Pockets

Pocket Residue Sap3 Sap2

S4 12a Val Val222a Thr Thr223 Ile Ile225 Tyr Tyr295 Gln Gln297 Leu Leu299 Gly Asp

S3 12 Val Val13 Ser Thr86 Asp Asp88 Thr Ser

118 Ser Ser120 Asp Asp220a Gly Gly

51 TyrS2 85a Gly Gly

86a Asp Asp221 Thr Thr225 Tyr Tyr301 Ser Asn303 Tyr Ala305a Ile Ile

S1 30a Val Ile84a Tyr Tyr86 Asp Asp88 Thr Ser

119 Val Ile123a Ile Ile

S10 34a Gly Gly193 Glu Glu195 Arg Arg216 Leu Leu303 Tyr Ala305 Ile Ile

S20 35a Ser Ser37 Asp Asp82 Ile Ile83a Glu Gly84 Tyr Tyr

129 Lys 131 Asn132 Glu 133 Ala80 Phe94 Trp137 Tyr

S30 303 Tyr

aResidues involved in pepstatin A binding.

Residues distinct in the sequence of Sap3 and Sap2 are given in italic.

Figure 4Chemical structure of pepstatin A (pepA), the classical aspartic proteinase

inhibitor. The hydroxyl group of the central statin moiety (��CHOH��CH2��)

forms two hydrogen bonds with the catalytic aspartic residues Asp32 and

Asp218.

COLOR

C. Borelli et al.

8 PROTEINS DOI 10.1002/prot

ing S10 are highly conserved within both enzymes. The

exclusion of Lys or Arg is explained by repulsion by

Arg195 in both enzymes.

Comparative data addressing Sap2 and Sap3 substrate

specificities at S4–S2 and S20 subsites do not exist. Resi-

due 83 may influence the substrate specificity of S2 in

different Sap enzymes of Candida albicans, Candida tro-

picalis, and Candida parapsilosis.7 As residue 83 (Sap

numbering based on Sap2, i.e., analog residue 74 in pep-

sin numbering) is at the margin of the substrate binding

sites S2 and S20, Fusek et al. stated that during the initial

fitting steps of substrate binding, residue 83 (residue 74,

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 9

ID: kannanb Date: 9/5/07 Time: 03:45 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Figure 5Pepstatin A (pepA) binding in the active site of Sap3. (A) Active site of the Sap3–pepA complex, superimposed with the difference electron density map of pepA. PepA is

shown as a stick model in green. Black dashed lines indicate polar interactions between pepA and active site residues. The active site loop which would cover the active

site cleft as seen in Figure 2 has been omitted for the sake of clarity, only residues 85 and 86 which take part in inhibitor binding are presented in grey. (B) Stereo ribbon

superimposition of aSap3 and pSap3. The active site loop of aSap3 (blue) is shifted towards the outside while the analogue loop of pSap3 (kept in yellow) is lowered down

onto the active site when the enzyme binds pepA (green). The active site flap movement significantly narrows the active site region and allows for polar interactions

between pepA and Gly85 and Asp86 in pSap3 (black dashed lines). The figures were created with Pymol.30

COLOR

COLOR

Crystal Structure of SAP3 and its complex with PepA

DOI 10.1002/prot PROTEINS 9

respectively) may come into close proximity with the P2

residue of substrate.7 With Glu83 in the C. tropicalis’

Sap, the enzyme was found to discriminate against acidic

residues in P2. Sap3 has also Glu83 and may share this

specificity. In contrast, the C. albicans’ Sap2 with Gly83

at this position tolerated a variety of side chains.

Inhibitor binding

Similar binding affinities for pepA (Fig.F4 4) to Sap2

and Sap3 indicated by Ki values being 17 nM (or IC50 27

nM, respectively41) for Sap2 and 60 nM for Sap340 have

been reported.

During the course of refinement of the pepA inhibited

Sap3 structure, well-defined density appeared along the

substrate binding groove so that pepA could easily be

built into the density. Only the nonbound primed site

end of the inhibitor, which sticks out toward the en-

trance loops of the cavity, exhibited partly weak density

indicating higher flexibility at this end.

The hydroxyl group of the central statin moiety of

pepA points towards the catalytic aspartic residues Asp32

and Asp218 [Fig.F5 5(A,B)], forming two hydrogen bonds

with these residues, characteristic for pepA-inhibited as-

partic proteinase structures,36-38, 42 where pepA acts as

a tetrahedral transition-state mimic. Further hydrogen

bonds are built between pepA and residues Gly34 (O),

Gly85 (N), Asp86 (N, OD2), Gly220 (O), Thr222 (N, OG1),

and Glu83 (O) but also many hydrophobic contacts con-

tribute to the interaction of pepA and Sap3 [Fig. 5(A), Table

II]. Residues that are responsible for pepA binding are

highly conserved within the structures of Sap2 and Sap3

(Table II). Within the residues taking part in inhibitor bind-

ing Ile30Sap2 is replaced by Val30Sap3 and Gly83Sap2 by

Glu83Sap3 in Sap3. At position 83, the hydrogen bond

between pepA and the enzyme is built via the main chain

peptide oxygen so that the alteration of the side chains of

residue 83 in Sap2 and Sap3 will probably not change the

pepA binding in both enzymes. The contact between pepA

and Val30Sap3 might be not as tight as with Ile30Sap2 in

Sap2, but still, Val30Sap3 is sufficient to interact hydro-

phobically with the inhibitor at P1 (3.99 A).

PepA in Sap3 fills suboptimally several binding sites,

especially S4, S3, and S10, which was also true for Sap2.

It was shown for the Sap2 inhibitor complex with A-

70450 containing a branched side chain at the S3 subsite

that exploiting the S3 site is critical for the high inhibi-

tion potency (IC50 ¼ 1.3 nM41) of A-70450 for Sap2.

Thus, optimal subsite filling offers an opportunity of

increasing binding strength and specificity.

In contrast to Sap2 where the voluminous inhibitor A-

70450 contacts Tyr51Sap2, which is part of the N-ent

loop, and contributes to S3, no participation of the N-

ent loop of Sap3 can be observed in the pepA binding.

As mentioned in the beginning, there is a conforma-

tional change of Sap3 on binding pepA with an rms devi-

ation of all 340 Ca atoms of 0.70 A after superposition

of aSap3 and pSap3, where the largest conformational

change occurs within the active site flap (residues 81–

91). The flap closes down onto the inhibitor by up to

5.9 A at its maximum and significantly narrows the

active site region [Fig. 5(B)]. We designate this as the

‘‘closed’’ conformation of the enzyme. The flap move-

ment allows the formation of hydrogen bonds and

hydrophobic interactions with pepA.

CONCLUSION

We have solved the structure of a second enzyme-Sap3-

out of the 10 member Sap family. The overall secondary

structure contains features similar to Sap2, confirming the

unique position of the Sap family among the aspartic pro-

teinases. Despite the highly conserved secondary structure

of Sap2 and Sap3, differences between both structures were

observed, which are consistent with particular substrate

specificities. The most striking alteration in the structure of

Sap3 consists of the S20 binding site becoming channel-

shaped subsequent to the turn of the loop with residues

129–135. This feature of S20 might be exploited best by a

more voluminous inhibitor than pepA, which would also

perhaps work better at subsites S4 and S3. The conforma-

tional move of the flexible active site flap as observed in the

structure of Sap3 when bound to pepA newly revealed the

importance of the active site flap for inhibitor binding.

With the determination of Sap3, we provide a broader ba-

sis for structure-based inhibitor design. Utilizing regions of

similarity within the structures of Sap2 and Sap3, inhibitors

acting on both enzymes can be designed. It will certainly be

of importance to solve the structure of Sap1 in order to

obtain the full set of the subgroup comprising Sap1–3 isoen-

zymes. Moreover, it will be interesting to see to what extent

the structures of Sap4-6 differ to those of Sap1–3.

ACKNOWLEDGMENTS

We thank I. Mathes and F. Lottspeich for performing

the MALDI-TOF peptide mass fingerprinting. We also

thank G. Bourenkov and H.D. Bartunik for help with

data collection at DESY.

REFERENCES

1. Szecsi PB. The aspartic proteases. Scand J Clin Lab Invest Suppl

1992;210:5–22.

2. Hoegl L, Korting HC, Klebe G. Inhibitors of aspartic proteases in

human diseases: molecular modeling comes of age. Pharmazie 1999;

54:319–329.

3. Naglik JR, Challacombe SJ, Hube B. Candida albicans secreted

aspartyl proteinases in virulence and pathogenesis. Microbiol Mol

Biol Rev 2003;67:400–428.

4. Schaller M, Korting HC, Borelli C, Hamm G, Hube B. Candida

albicans-secreted aspartic proteinases modify the epithelial cytokine

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 10

ID: kannanb Date: 9/5/07 Time: 03:45 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

C. Borelli et al.

10 PROTEINS DOI 10.1002/prot

response in an in vitro model of vaginal candidiasis. Infect Immun

2005;73:2758–2765.

5. Schaller M, Schackert C, Korting HC, Januschke E, Hube B. Inva-

sion of Candida albicans correlates with expression of secreted

aspartic proteinases during experimental infection of human epider-

mis. J Invest Dermatol 2000;114:712–717.

6. Naglik J, Albrecht A, Bader O, Hube B. Candida albicans protein-

ases and host/pathogen interactions. Cell Microbiol 2004;6:915–926.

7. Fusek M, Smith EA, Monod M, Dunn BM, Foundling SI. Extracel-

lular aspartic proteinases from Candida albicans, Candida tropicalis,

and Candida parapsilosis yeasts differ substantially in their specific-

ities. Biochemistry 1994;33:9791–9799.

8. De Bernardis F, Cassone A, Sturtevant J, Calderone R. Expression

of Candida albicans SAP1 and SAP2 in experimental vaginitis.

Infect Immun 1995;63:1887–1892.

9. Naglik JR, Newport G, White TC, Fernandes-Naglik LL, Greenspan

JS, Greenspan D, Sweet SP, Challacombe SJ, Agabian N. In vivo

analysis of secreted aspartyl proteinase expression in human oral

candidiasis. Infect Immun 1999;67:2482–2490.

10. Schaller M, Januschke E, Schackert C, Woerle B, Korting HC.

Different isoforms of secreted aspartyl proteinases (Sap) are

expressed by Candida albicans during oral and cutaneous candidosis

in vivo. J Med Microbiol 2001;50:743–747.

11. Albrecht A, Felk A, Pichova I, Naglik JR, Schaller M, de Groot P,

Maccallum D, Odds FC, Schafer W, Klis F, Monod M, Hube B. Gly-

cosylphosphatidylinositol-anchored proteases of Candida albicans

target proteins necessary for both cellular processes and host-patho-

gen interactions. J Biol Chem 2006;281:688–694.

12. Taylor BN, Hannemann H, Sehnal M, Biesemeier A, Schweizer A,

Rollinghoff M, Schroppel K. Induction of SAP7 correlates with viru-

lence in an intravenous infection model of candidiasis but not in a

vaginal infection model in mice. Infect Immun 2005;73:7061–7063.

13. Stewart K, Abad-Zapatero C. Candida proteases and their inhibition:

prospects for antifungal therapy. Curr Med Chem 2001;8:941–948.

14. Cutfield S, Marshall C, Moody P, Sullivan P, Cutfield J. Crystalliza-

tion of inhibited aspartic proteinase from Candida albicans. J Mol

Biol 1993;234:1266–1269.

15. Abad-Zapatero C, Goldman R, Muchmore SW, Hutchins C, Stewart

K, Navaza J, Payne CD, Ray TL. Structure of a secreted aspartic prote-

ase from C. albicans complexed with a potent inhibitor: implications

for the design of antifungal agents. Protein Sci 1996;5:640–652.

16. Cutfield SM, Dodson EJ, Anderson BF, Moody PC, Marshall CJ,

Sullivan PA, Cutfield JF. The crystal structure of a major secreted

aspartic proteinase from Candida albicans in complexes with two

inhibitors. Structure 1995;3:1261–1271.

17. Cannon RD, Timberlake WE, Gow NA, Bailey D, Brown A, Gooday

GW, Hube B, Monod M, Nombela C, Navarro F, Perez R, Sanchez

M, Pla J. Molecular biological and biochemical aspects of fungal

dimorphism. J Med Vet Mycol 1994;32 (Suppl 1):53–64.

18. Hube B, Monod M, Schofield DA, Brown AJ, Gow NA. Expression

of seven members of the gene family encoding secretory aspartyl

proteinases in Candida albicans. Mol Microbiol 1994;14:87–99.

19. White TC, Agabian N. Candida albicans secreted aspartyl protein-

ases: isoenzyme pattern is determined by cell type, and levels are

determined by environmental factors. J Bacteriol 1995;177:5215–

5221.

20. Abad-Zapatero C, Goldman R, Muchmore SW, Hutchins C, Oie T,

Stewart K, Cutfield SM, Cutfield JF, Foundling SI, Ray TL. Struc-

ture of secreted aspartic proteinases from Candida. Implications for

the design of antifungal agents. Adv Exp Med Biol 1998;436:297–313.

21. Umezawa H, Aoyagi T, Morishima H, Matsuzaki M, Hamada M.

Pepstatin, a new pepsin inhibitor produced by Actinomycetes.

J Antibiot (Tokyo) 1970;23:259–262.

22. Borg-von Zepelin M, Beggah S, Boggian K, Sanglard D, Monod M.

The expression of the secreted aspartyl proteinases Sap4 to Sap6

from Candida albicans in murine macrophages. Mol Microbiol

1998;28:543–554.

23. Laemmli UK. Cleavage of structural proteins during the assembly

of the head of bacteriophage T4. Nature 1970;15:680–685.

24. Otwinowski Z, Minor W. DENZO: a film processing program for

macromolecular crystallography. New Haven, CT: Yale University;

1993.

25. Kleywegt GJ, Jones TA. Efficient rebuilding of protein structures.

Acta CrystallogrSect D: Biol Crystallogr 1996;52:829–832.

26. Vagin A, Teplyakov A. An approach to multi-copy search in molec-

ular replacement. Acta CrystallogrSect D: Biol Crystallogr 2000;56:

1622–1624.

27. Brunger AT, Adams PD, Clore GM, DeLano WL, Gros P, Grosse-

Kunstleve RW. Crystallography and NMR system: a new software

suite for macromolecular structure determination. Acta Crystallogr-

Sect D: Biol Crystallogr 1998;54:905–921.

28. Laskowski R, Macarthur M, Moss D, Thornton J. PROCHEK—a

program to check the stereochemical quality of protein structures.

J Appl Crystallogr 1993;26:283–291.

29. Ramachandran GN, Sasisekharan V. Conformation of polypeptides

and proteins. Adv Protein Chem 1968;23:283–437.

30. DeLano WL. PyMOL reference manual. San Carlos, CA: DeLano

Scientific LLC; 2006.

31. Thompson JD, Higgins DG, Gibson TJ. CLUSTAL W: improving

the sensitivity of progressive multiple sequence alignment through

sequence weighting, position-specific gap penalties and weight

matrix choice. Nucleic Acids Res 1994;22:4673–4680.

32. Davies DR. The structure and function of the aspartic proteinases.

Annu Rev Biophys Biophys Chem 1990;19:189–215.

33. Sielecki AR, Fedorov AA, Boodhoo A, Andreeva NS, James MN.

Molecular and crystal structures of monoclinic porcine pepsin

refined at 1.8 A resolution. J Mol Biol 1990;214:143–170.

34. Stewart K, Goldman R, Abad-Zapatero C. The secreted proteinases

from Candida: challenges for structure-aided drug design. In: Dunn

B, editor. Proteases of infectious agents. San Diego, CA: Academic

Press; 1999. pp 117–138.

35. Dhanaraj V, Dealwis CG, Frazao C, Badasso M, Sibanda BL, Tickle

IJ, Cooper JB, Driessen HP, Newman M, Aguilar C, Wood SP, Blun-

dell TL, Hobert PM, Geoghegan KF, Ammirati MJ, Danley DE,

O’Connor BA, Hoover DJ. X-ray analyses of peptide-inhibitor com-

plexes define the structural basis of specificity for human and

mouse renins. Nature 1992;357:466–472.

36. Metcalf P, Fusek M. Two crystal structures for cathepsin D: the lysoso-

mal targeting signal and active site. EMBO J 1993;12:1293–1302.

37. James MN, Sielecki A, Salituro F, Rich DH, Hofmann T. Conforma-

tional flexibility in the active sites of aspartyl proteinases revealed

by a pepstatin fragment binding to penicillopepsin. Proc Natl Acad

Sci USA 1982;79:6137–6141.

38. Suguna K, Padlan EA, Bott R, Boger J, Parris KD, Davies DR.

Structures of complexes of rhizopuspepsin with pepstatin and other

statine-containing inhibitors. Proteins 1992;13:195–205.

39. Lowther WT, Majer P, Dunn BM. Engineering the substrate speci-

ficity of rhizopuspepsin: the role of Asp 77 of fungal aspartic pro-

teinases in facilitating the cleavage of oligopeptide substrates with

lysine in P1. Protein Sci 1995;4:689–702.

40. Koelsch G, Tang J, Loy JA, Monod M, Jackson K, Foundling SI, Lin

X. Enzymic characteristics of secreted aspartic proteases of Candida

albicans. Biochim Biophys Acta 2000;1480:117–131.

41. Capobianco JO, Lerner CG, Goldman RC. Application of a fluoro-

genic substrate in the assay of proteolytic activity and in the discov-

ery of a potent inhibitor of Candida albicans aspartic proteinase.

Anal Biochem 1992;204:96–102.

42. Fujinaga M, Chernaia MM, Tarasova NI, Mosimann SC, James

MN. Crystal structure of human pepsin and its complex with pep-

statin. Protein Sci 1995;4:960–972.

J_ID: Z7E Customer A_ID: 00512-2006.R1 Cadmus Art: PROT 21425 Date: 9-MAY-07 Stage: I Page: 11

ID: kannanb Date: 9/5/07 Time: 03:45 Path: J:/Production/PROT/Vol00000/070114/3B2/C2PROT070114

Crystal Structure of SAP3 and its complex with PepA

DOI 10.1002/prot PROTEINS 11