You are on page 1of 6

Insight into autoproteolytic activation from the

structure of cephalosporin acylase: A protein


with two proteolytic chemistries
Jin Kwang Kim*†, In Seok Yang*†, Hye Jeong Shin‡, Ki Joon Cho*, Eui Kyung Ryu*§, Sun Hwa Kim*¶, Sung Soo Park*,
and Kyung Hyun Kim‡储**
*Department of Life Sciences and Biotechnology, School of Life Sciences and Biotechnology, and ‡Department of Bio-Microsystem Technology, Korea
University, Seoul 136-701, Korea; and 储Department of Biotechnology, College of Science and Technology, Korea University, Jochiwon 339-700, Korea

Edited by Earl W. Davie, University of Washington, Seattle, WA, and approved December 19, 2005 (received for review September 12, 2005)

Cephalosporin acylase (CA), a member of the N-terminal nucleo- between Gly-169 and Ser-170 generates the ␣⬘ subunit (the ␣
phile hydrolase family, is activated through sequential primary and subunit with a 9-aa pro segment) and the ␤ subunit. A subsequent
secondary autoproteolytic reactions with the release of a pro secondary autocleavage between Gly-160 and Asp-161 releases the
segment. We have determined crystal structures of four CA mu- pro segment from the ␣⬘ subunit, resulting in an active heterotet-
tants. Two mutants are trapped after the primary cleavage, and the ramer (␣␤)2. The release of the pro segment may be required for
other two undergo secondary cleavage slowly. These structures optimal CA activity. Ser-170, which becomes the N-terminal resi-
provide a look at pro-segment conformation during activation in due (Ser-1)†† of the ␤ subunit of CA is invariant and known to play
N-terminal nucleophile hydrolases. The highly strained helical pro a critical role in both autoproteolytic activation and enzyme catal-
segment of precursor is transformed into a relaxed loop in the ysis (23–25).
intermediates, suggesting that the relaxation of structural con- In an effort to gain a better understanding of the autoproteolyic
straints drives the primary cleavage reaction. The secondary auto- mechanism, we have designed constructs containing mutations that
proteolytic step has been proposed to be intermolecular. However, partially prevent or retard autocleavage and carried out the struc-
our analysis provides evidence that CA is processed in two sequen- tural determination of these mutants. Here, we report structural
tial steps of intramolecular autoproteolysis involving two distinct evidence that CA is processed in two sequential steps of intramo-
residues in the active site, the first a serine and the second a lecular autoproteolysis involving two distinct proteolytic mecha-
glutamate. nisms, the first mediated by a serine residue and the second by a
glutamate.
autoproteolysis 兩 precursor activation 兩 intermediate structure 兩
pro segment Results
Choice of Mutant Proteins. A single amino acid mutation of
Ser-170 to alanine (S170A) prevents primary cleavage, pro-
P osttranslational autoproteolysis is a mechanism used to activate
many proteins via self-catalyzed peptide bond rearrangements,
which play an essential role in a wide variety of biological processes.
ducing an enzymatically inactive single-polypeptide precursor
(Fig. 1B). The crystal structures of precursor (S170A) and
They include activation cascades such as blood coagulation and mature CA have been reported (11). In this study, we have
fibrinolysis, cell death, embryonic development, protein targeting focused mainly on four mutants. S170C and E159Q underwent
and degradation, viral protein processing, and zymogen activation primary autocleavage but no secondary autocleavage. Y202L
(1–6). Increasing evidence suggests that autoproteolysis may be and R226K, slow processing mutants, showed a delayed release
involved in more biological activation processes than previously of the pro segment in comparison with the mature enzyme.
appreciated. These processes are all mediated by intramolecular or Tyr-202 and Arg-226 are located in the substrate binding site
intermolecular autocleavage reactions. Four types of intramolecu- of CA. The Y202L and R226K mutants were initially analyzed
lar protein modifications have been studied extensively: hedgehog as altered-activity mutants and found to be capable of sequen-
proteins, inteins, N-terminal nucleophile (Ntn) hydrolases, and tially producing ␣⬘ and ␣ subunits during activation. In the case
pyruvoyl enzymes (4, 7–9). of Y202L, it took ⬇24 h at 37°C to make a complete transition
Members of the Ntn hydrolase family have divergent sequences, from the uncleaved to the cleaved form (Fig. 1C, Y202L
but share an ␣␤␤␣ sandwich structural motif and a common control). Thus, early (Y202L) and late (Y202L,L) stages of
enzyme mechanism (7, 8). In addition, they catalyze internal
peptide bond rearrangement through an N 3 O or N 3 S acyl shift
Conflict of interest statement: No conflicts declared.
by a nucleophilic Ntn amino acid created by autoproteolytic pro-
This paper was submitted directly (Track II) to the PNAS office.
cessing (4, 7). Structures of mature Ntn hydrolases, including
cephalosporin acylase (CA), penicillin G acylase (PA), penicillin V Abbreviations: CA, cephalosporin acylase; Ntn, N-terminal nucleophile.

acylase, glutamine 5-phosphoribosyl-1-pyrophosphate amidotrans- Data deposition: The atomic coordinates have been deposited in the Protein Data Bank,
ferase, 20S proteasome ␤ subunit (PRO), glycosylasparaginase www.rcsb.org (PDB ID codes 2ADV, 2AE5, 2AE4, and 2AE3 for Y202L, S170C, E159Q, and
R226K, respectively).
(GA), and L-aminopeptidase, have been determined (10–17). †J.K.K. and I.S.Y. contributed equally to this work.
Structures of the precursors of CA, PA, PRO, and GA have been §Present address: Department of Chemistry, Pohang University of Science and Technology,
reported, providing some insight into the activation mechanism (11, Pohang 790-784, Korea.
18–22). However, the understanding of mechanism cannot be ¶Present address: Department of Microbiology and Immunology, College of Medicine,
complete without structural information on the first cleavage Catholic University, Seoul 137-701, Korea.
intermediates. **To whom correspondence should be addressed. E-mail: khkim@korea.ac.kr.
Ntn hydrolases are of particular interest because they include ††The amino acid residue is numbered according to the precursor (residues 1– 691) com-
acylase enzymes used in the biosynthesis of the ␤-lactam antibiotics prising the ␣ subunit (residues 1–160), the pro-segment region (residues 161–169), and
cephalosporin and penicillin. We have focused on CA. This enzyme the ␤ subunit (residues 170 – 691). The amino acid residue of the ␤mature subunit (precursor
is expressed as a 76-kDa precursor that is activated by two sequen- number minus 169) is numbered in parentheses, if applicable.
tial autocleavage steps (11, 23) (Fig. 1A). A primary cleavage © 2006 by The National Academy of Sciences of the USA

1732–1737 兩 PNAS 兩 February 7, 2006 兩 vol. 103 兩 no. 6 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0507862103


Fig. 2. Structural comparison of CA proteins. (A) Superposition of the overall
structures of S170A, S170C, E159Q, Y202L, R226K, and the mature CA. Each
structure is shown in yellow, cyan, dark green, green, slate blue, and marine,
respectively. To emphasize the differences of the pro segments, the pro
segment of Y202L is shown in red. Residues Gly-160, Ser-170, Tyr-202, and
Arg-226 are shown in a ball-and-stick representation in magenta. (B) Stereo-
view of the pro-segment conformation of Y202L. The electron density map
corresponds to the 2 Fo ⫺ Fc map calculated at 0.5 ␴.

complete release of the pro segment, which was consistent with


the SDS兾PAGE result (Fig. 1B, lane Y202L,L).

Comparison of Mutant Structures and Pro Segments. The refined


overall structures of the mutants, Y202L, R226K, S170C, and
Fig. 1. Autoproteolytic cleavage of glutaryl 7-aminocephalosporanic acid E159Q (see Table 1, which is published as supporting information
acylase. (A) Schematic diagram of two-step autoproteolytic cleavage. The on the PNAS web site), are similar to the precursor and mature
precursor is activated through a primary cleavage to generate the ␣ subunit structures reported earlier (11) (Fig. 2A). In brief, they adopt the
containing the pro segment (␣⬘ subunit) and the ␤ subunit. A subsequent
␣␤␤␣ motif formed by two ␤-sheets packed against each other,
secondary autocleavage releases the pro segment. (B) SDS兾PAGE analyses of
the autoproteolytic cleavage of CA. Lane M, molecular weight markers; S170A
sandwiched by two layers of ␣-helices. After superposition of the C␣
is a precursor form. Y202L,L was obtained by incubating Y202L at 37°C for 24 h. atoms, the rms deviations of mutants against the precursor S170A
(C) Time-course experiments of secondary autocleavage. Y202L or S170C was range from 0.37 to 0.53 Å. The structural ␣␤␤␣ core is highly
incubated with mature enzyme (WT) or 1-ethyl-3-(3-dimethylaminopropyl- conserved in the Ntn hydrolase superfamily (8), as in structures of
)carbodiimide (EDC) or at different pH values. Incubation and reconstitution glutarylamidase from Pseudomonas sp. and CA from Pseudomonas
experiments (Inc兾Rec). Lane 1, freshly prepared Y202L; lane 2, crystals of fresh diminuta (10, 24). However, there are significant differences at the
Y202L; lane 3, crystals of Y202L incubated at 37°C; lane 4, reconstitution with entryway to the active site, which is a deep cleft at the tip of the
the combination of ␣mature and ␤mature subunits; lane 5, reconstitution with ␣␤␤␣ sandwich motif, where the pro segment is located.
␣⬘S170C and ␤mature subunits; and lane 6, reconstitution with ␣⬘E159Q and ␤mature
In precursor CA the pro segment is ␣-helical from Pro-163 to
subunits.
Ala-166 (11). Proline is not favored in helices, and significant
deviations from helical geometry suggest that this is a high-energy
conformation, generating a peptide flip for initiation of primary
Y202L during the time course of autoproteolysis could be autoproteolysis. A similar strained conformation, which was sug-
captured for crystallization (Fig. 1B, lanes Y202L and gested to regulate the activation of glycosylasparaginase (21) and a
Y202L,L). The presence or absence of the pro segment linked pyruvoyl enzyme (22), is believed to be the driving force for
to the ␣ subunit of the mutant (␣⬘␤)2 or (␣␤)2 was confirmed
BIOPHYSICS

autocleavage to break the scissile peptide bond (26). In the inter-


by mass spectrometric and crystallographic results (see Fig. 6, mediate mutants of CA, which have undergone the initial but not
which is published as supporting information on the PNAS web the secondary cleavage, the pro segments adopt strikingly different
site). The electron density of the pro segment in Y202L,L conformations from that of the precursor (Fig. 2 A). First, the pro
revealed a clear discontinuity after Gly-160, indicating a segments show a loop conformation without regular secondary

Kim et al. PNAS 兩 February 7, 2006 兩 vol. 103 兩 no. 6 兩 1733


structure. While their N termini are covalently attached to the ␣
subunit, the C termini upon primary cleavage are moved from
residue 170 and may be partly exposed to solvent. In addition, there
are no significant geometric deviations from ideal geometry in the
pro segments, indicating unstrained conformations. Second, they
are highly mobile with high B factors (68–83 vs. 18–35 Å2 in
precursor). The electron densities for the pro segments are visible
only at a low contour level, as shown in Fig. 2B (and see Fig. 7A,
which is published as supporting information on the PNAS web
site). The pro segment may become highly mobile upon primary
cleavage, a characteristic of a terminal end, which would probably
facilitate secondary autocatalysis and subsequent dissociation of
the pro segment. Thus the highly strained helical conformation
of the precursor pro segment is transformed into a highly flexible
and relaxed loop conformation in the intermediates, supporting
the notion that the relaxation of structural constraints drives the
primary autocleavage reaction.
A comparative analysis of the mutant structures reveals that
the pro segments exist in a variety of conformations (see Fig. 7B).
Although Y202L, R226K, and E159Q follow similar backbone
paths from Glu-159 to Pro-162, moving down to the deep cleft at
Glu-159, they extend away from each other at Pro-163. By contrast,
residues Glu-159 and Gly-160 in S170C have ␸ conformational
angles opposite to the signs of other mutants and the pro segment
moves downward at Gly-160 to the deep cleft. The differences in the
backbone path place Pro-163 of S170C and Y202L at positions ⬇10
Å apart from each other. Moreover, S170A precursor structure
reveals extensive contacts between the pro segment and protein
residues (see Tables 2 and 3, which are published as supporting
information on the PNAS web site), but there are significantly
fewer contacts in the intermediate mutants. Although only partial
models of the pro segments could be built, it is clearly shown that
the number of polar and van der Waals contacts is greatly dimin-
ished in the intermediate mutants (see below). It is likely that
conformational constraints in precursor may be confined within the
pro segment during protein folding with expenditure of the non-
covalent interaction energies, which are released upon primary
autocleavage.

Primary Autoproteolytic Site. The active sites of mutants formed


upon primary autocleavage display structural features similar to the
mature enzyme. First, an intact catalytic pseudotriad of Ser-His- Fig. 3. Secondary autocleavage site of CA proteins. The interactions be-
Glu becomes registered in all of the mutants, involving the free tween pro-segment and protein residues in Y202L, precursor S170A, R226K,
␣-amino group of (Ser-1), not the usual hydroxyl group, as in the S170C, and E159Q. Those involved in hydrophobic interactions in S170A are
wild-type structures (11, 24). The side chain of (His-23) in the superimposed in a transparent space-filling representation. Hydrogen bonds
wild-type contributes to a reduction in the pKa of the ␣-amino are indicated by dotted lines, and water molecules are represented as red
group of (Ser-1), which then acts as a general base to enhance the spheres. The distances between the side chain of Glu-159 and the backbone
nucleophilicity of its own hydroxyl group (data not shown). Second, carbon atom of Gly-160 via WAT1 in Y202L and R226K are highlighted by thick
Ser-170(Ser-1), which is buried in the precursor (11, 18), becomes dotted lines.
partially exposed to solvent in the mutants; it is completely open to
solvent in mature enzyme (11). The differential burial of Ser-170
vented secondary autoproteolysis (25). CA secondary autoprote-
may correspond to a structural activation signal for enzyme catal-
olysis might be correlated with catalytic activity. Nevertheless, the
ysis. Third, despite the partial enclosure of the active site by the pro
segment, a solvent molecule emerges near the hydroxyl group of question remains: how is secondary cleavage initiated in the ab-
(Ser-1) at the active site in Y202L and R226K. The solvent molecule sence of the active mature enzyme? Importantly, the release rate of
at this position plays a crucial role in the enzyme activity of mature the pro segment of Y202L was unchanged, and the autocleavage of
CA (10, 11, 24), as recent studies revealed the role of the bound S170C was not restored when incubated with the mature enzyme
water near the catalytic residue in Ntn hydrolases (18, 21, 26). (Fig. 1C, Y202L⫹WT and S170C⫹WT), providing experimental
evidence against intermolecular secondary cleavage. More con-
Secondary Autoproteolysis: Is It Intermolecular or Intramolecular? In vincingly, Y202L crystals were incubated at 37°C for 24 h, where
contrast to the primary autocleavage mechanisms for the Ntn intermolecular autocleavage is not allowed because of crystal
hydrolases (11, 19, 20, 22), the secondary cleavage mechanism of packing, and SDS兾PAGE analysis of the incubated Y202L crystals
CA is not clearly understood. An intermolecular cleavage mecha- demonstrated the generation of the cleaved ␣ subunit (Fig. 1C,
nism for CA activation was proposed based on reconstitution Y202L Inc兾Rec, lane 3).
experiments (23). CA takes glutaryl 7-aminocephalosporanic acid In the precursor structure, the carboxyl group of Glu-159 is
(7-ACA) as a primary substrate and catalyzes its hydrolysis into within hydrogen-bonding distance of Gln-75 (Fig. 3, S170A). The
glutaric acid and 7-ACA. It was reported that the mutations in the side chain of Asp-161 is hydrogen-bonded to those of Ser-152 and
substrate binding site that inhibited deacylation activity also pre- Arg-155. Moreover, Leu-165 is involved in hydrophobic contacts

1734 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0507862103 Kim et al.


with Leu-148, Tyr-199, and Phe-200. There are extensive contacts
between the pro segment and protein residues in the deep cleft (see
Table 2). In the intermediate mutants, however, there are fewer
contacts and no hydrophobic interaction is found between Leu-165
and protein residues. Notably, Gln-75 points away toward the
solvent and its hydrogen-bonding to Glu-159 is absent. In Y202L,
Glu-159 is linked to Thr-76 by a network of hydrogen bonds
involving a series of water molecules (Fig. 3, Y202L). A positive
density at the Fo ⫺ Fc difference Fourier map could be modeled as
solvent molecule WAT1, which is hydrogen-bonded to the carbonyl
oxygen of Asp-161 and another solvent molecule. Interestingly,
WAT1 is also close to the side chain of Glu-159 (4.0 Å) and is
positioned on top of the carbonyl carbon of Gly-160 (4.0 Å) at the
scissile peptide bond of secondary autocleavage. R226K has a very
similar backbone configuration to that of Y202L at the secondary
cleavage site (Fig. 3). Notably, a solvent molecule corresponding to
WAT1 of Y202L could be placed in hydrogen-bonding distance to
Glu-159 (3.5 Å) and close to the carbonyl carbon of Gly-160 (4.0 Å).
Both mutant structures thus reveal that WAT1 could be modeled
only 3.5–4.0 Å from Glu-159 adjacent to the conserved secondary
site, suggesting that it is well within the range to facilitate proton
abstraction.
Fig. 4. Close-up view of the active site in Y202L, R226K, S170C, and E159Q
In the S170C mutant structure, however, strong hydrogen bonds structures. Cation–␲ and aromatic interactions are shown in a ball-and-stick
could be made between a water molecule and the backbone oxygen representation. The electron densities from the 2 Fo ⫺ Fc maps were contoured
and nitrogen atoms of Glu-159 and Gly-160, respectively, instead of at 1.0 ␴. Hydrogen bonds are indicated by dotted lines, and water molecules
the side chain of Glu-159 (Fig. 3). S170C displays different back- are represented as red spheres.
bone conformation of the secondary cleavage site, in contrast to
other mutants (see Fig. 7B). In E159Q, where the local backbone
conformation is similar to that of Y202L or R226K, no significant structures as well as in precursor and mature CA (see Fig. 8, which
electron densities for solvent molecules could be observed at the is published as supporting information on the PNAS web site).
secondary site. F177P mutant remains a nonprocessed precursor with no catalytic
activity (25). Notably, Y202L, R226K, and S170C mutants have the
Carboxyl Proteolytic Autocleavage. In an effort to derive a detailed defects localized in this region, where Tyr-202(Tyr-33), Arg-
molecular mechanism of secondary autocleavage, we further car- 226(Arg-57), and (S1) are replaced by leucine, lysine, and cysteine,
ried out a number of biochemical and mutational experiments. At respectively. In Y202L, solvent water molecules are incorporated in
first, it appeared that the sequential arrangement of two carboxylic hydrogen-bonding networks, similar to precursor or mature CA.
acids, Glu-159 and Asp-161, together with WAT1 has some resem- The cation–␲ interaction in R226K may become weak because of
blance to aspartyl protease. A carboxyl protease inhibitor, a car- the mutation, resulting in a much slower processing mutant than
bodiimide-mediated modification of Y202L, blocked the secondary Y202L (data not shown). S170C and E159Q, on the other hand,
cleavage (Fig. 1C, Y202L⫹EDC). Monitoring the pH dependence lack the hydrogen-bonding networks, probably interrupted by an
of the secondary autocleavage also indicated that deprotonation of unexplained strong positive density found midway between (Arg-
the carboxyl group was required for autocatalysis (Fig. 1C, Y202L 57) and (Cys-1) or (Ser-1). Therefore, the perturbation in the
pH 5.0–7.0). However, the mutants D161L or D161N gave a fully integrity of this region caused by mutation may result in a serious
processed enzyme (Fig. 1B). Mutations in other candidate residues defect in the secondary autocatalytic activity, suggesting that the
that are close to the scissile peptide bond, S152A, R155K, and cation–␲ and hydrogen-bonding interactions may play an important
R155M, also produced a successfully processed enzyme. Only two structural role in supporting proper autocatalysis.
mutants, E159M and E159Q, lost their secondary autocleavage
activities. To test the possibility of general base catalysis by gluta- Discussion
mate, the ␣⬘ subunits from S170C and E159Q mutants were A glutamate residue acting as a general base was first discovered in
purified separately in the denatured condition (␣⬘S170C and ␣⬘E159Q, histone acetyltransferase, where the conserved glutamic acid is an
respectively) and reconstituted with the denatured ␤ subunit essential catalytic residue by deprotonating the histone substrate
(␤mature) of the mature enzyme by refolding (Fig. 1C, Y202L (30). N-myristoyl transferase also has a conserved glutamic acid
Inc兾Rec, lanes 4–6). The secondary cleavage occurred only when residue that is essential in catalysis (31). Interestingly, chloram-
the enzyme was reconstituted with the combination of ␣⬘S170C and phenicol acetyltransferase uses a histidine residue as a general base
␤mature subunits, but did not when it was reconstituted with the to abstract a proton from the primary hydroxyl group of chloram-
␣⬘E159Q and ␤mature subunits. Our results convincingly demonstrated phenicol. A glutamic acid replacement of this histidine can substi-
that Glu-159 is pivotal for this secondary processing. tute as the general base, albeit with lower catalytic activity (32).
Mutational studies in other homologous CA proteins support our
Cation–␲ Interaction. The oxyanion binding of Arg-155 could be notion that Glu-159 plays a critical role in secondary autoprote-
stabilized by the carbonyl oxygens of Tyr-149 and Val-150 (Fig. 3, olysis: The Glu-159 mutants, E159L, E159M, and E159Q, of CA
Y202L). Importantly, the side chains of Tyr-149 and (Phe-177) from P. diminuta also underwent primary autocleavage but lost the
form a near edge-face aromatic interaction and (Phe-177) is in a secondary autocleavage activity (18). In the crystal structure of
favorable arrangement for a cation–␲ interaction (27) with (Arg- E159Q, the backbone conformation from Gly-158 to Asp-161 is
BIOPHYSICS

57) in the substrate binding site (Fig. 4). Similar interactions are fairly similar to that of the Y202L and R226K forms (Fig. 3).
recognized as crucial for the structural stabilization of caspase BIR Based on the intermediate structures, we therefore propose that
domain (28) and neutrophil collagenase (29). Moreover, (Phe-177) CA may have evolved separate autoproteolytic mechanisms, the
is the only residue of 682 modeled amino acid residues in a first a serine proteolysis and the second a carboxyl proteolysis.
disallowed region of the Ramachandran plot in all mutant crystal Upon protein folding, precursor CA undergoes the primary auto-

Kim et al. PNAS 兩 February 7, 2006 兩 vol. 103 兩 no. 6 兩 1735


truly belongs to the Ntn hydrolase family of intramolecular self-
cleavage.
Two mutants, Y202L and R226K, showed significant decreases
in the rates of autocleavage reactions, thus being capable of
kinetically trapping the uncleaved and the cleaved forms of the pro
segment. We postulate that Y202L and R226K may represent
intermediate structures in transition from an immature to a mature
arrangement during activation, whereas S170C and E159Q may be
regarded as trapped intermediates. However, we cannot rule out
the possibility that the pro segment containing the Glu-159–WAT1
pair in Y202L or R226K may not represent a legitimate interme-
diate during autoproteolysis of CA or may adopt a slightly altered
conformation from that of a native intermediate. Nevertheless, our
results are strongly suggestive that the proposed mechanisms rep-
resent a common molecular mode of action for CA autoproteolysis.
In conclusion, our finding of dual catalysis at intervals of 9 aa,
both serine and a previously undiscovered carboxyl autoproteolysis,
may have novel implications for understanding the mechanisms
of autoproteolysis that are important in such diverse biological
processes.

Materials and Methods


Plasmids, Strains, and Site-Directed Mutagenesis. The pET23d plas-
mids harboring the cloned mature CA gene from Pseudomonas sp.
strain GK16 and the S170A mutant gene have been described (11).
The mutants S152A, R155K, R155M, E159M, E159Q, D161L,
D161N, S170C, Y202L, and R226K were constructed by site-
directed mutagenesis, using a PCR. Each PCR product was cloned
into vector pET22b or pET23d (Novagen), which carries a sequence
coded with a C-terminal His tag downstream from the T7 RNA
polymerase promoter site. The construct was then transformed into
an Escherichia coli DH5␣. The bacteria were grown at 37°C, and the
plasmids were purified by using a plasmid mini kit (Qiagen,
Valencia, CA). All mutations were verified by automated DNA
sequencing (data not shown). E. coli strains were grown in LB
Fig. 5. Proposed autoproteolytic activation mechanisms for CA. (A) Primary
medium.
serine autoproteolytic cleavage. A peptide flip (open arrow) under the influ-
ence of the strained conformation creates a hydrogen bond between WAT1 Protein Expression and Purification. The clones were inserted into E.
and the carbonyl group of Gln-168. WAT1, held by hydrogen bonds in coli BL21兾DE3 (Novagen) for protein overexpression, and bacteria
pseudotetrahedral geometry, acts as a general base and deprotonates the were grown at 37°C until the culture reached an absorbance of 0.6
hydroxyl group of Ser-170 for initiation of primary cleavage (curved arrow). at 600 nm. Protein expression was induced by addition of isopropyl
Dotted lines indicate hydrogen-bond interactions. WAT and WAT1 represent 1-thio-␤-D-galactopyranoside (Sigma) to a final concentration of
a water molecule. (B) Secondary carboxyl autoproteolytic cleavage. It is initi- 0.4 mM, which was followed by an additional 4 h of incubation. The
ated by a general-base-catalyzed attack of WAT1 on the Gly-160 –Asp-161
purification of all of the mutant CA proteins was performed
peptide bond, where Glu-159 acts as a general base (curved arrows). The
main-chain NH of Arg-155 stabilizes the oxyanion of Gly-160.
according to published procedures (11). All of the mutant proteins
yielded proteins of sufficient purity for crystallization.

proteolysis that uses a N 3 O acyl shift (Fig. 5A). Initially, a Analysis of Autocleavage. The extent of primary and secondary
high-energy state of the pro segment caused by conformational autocleavage for the purified mutants was analyzed by SDS兾PAGE.
constraints may trigger a peptide flip. WAT1, held by hydrogen Y202L showed an approximately equimolar mixture of cleaved and
bonds in pseudotetrahedral geometry, acts as a general base to uncleaved ␣ subunits subsequent to purification (Y202L early:
deprotonate the hydroxyl group of Ser-170 to enhance its nucleo- Y202L). Both the ␣ subunit linked with the pro segment (␣⬘
philicity. The hydroxyl group carries out the nucleophilic attack on subunit) and the ␣ subunit itself were subjected to MALDI-TOF
the peptide bond between Gly-169 and Ser-170. This primary mass spectrometry (Korea Basic Research Institute, Seoul). The
time-course monitoring of in vitro secondary processing of S170C
cleavage results in generating a helix-to-loop transition of the pro
or Y202L was carried out in 20 mM Tris (pH 8.0) at 37°C for 24 h.
segment, which consequently triggers an activation signal switch for
S170C remained in the uncleaved form because of the lack of
initiation of the secondary carboxyl autoproteolysis. Glu-159 acting secondary cleavage. For Y202L, it took ⬇24 h at 37°C, to make a
as a general base may accept a proton from WAT1 to enhance its complete transition from the uncleaved to the cleaved form (Y202L
nucleophilicity (Fig. 5B). The nucleophilic attack on the peptide late: Y202L,L). From the solution processed at 37°C, 5-␮l aliquots
bond between Gly-160 and Asp-161 is carried out by WAT1, were sampled at 3-h intervals. The addition of the mature enzyme
producing a tetrahedral transition state. Formation of the ester to each mutant solution or 0.1 M 1-ethyl-3-(3-dimethylaminopro-
intermediate (a N 3 O acyl shift) requires stabilization of the pyl) carbodiimide (EDC) to Y202L in 1.0 M glycine ethyl ester at
oxyanionic transition state, which may be undertaken by the pH 5.0 at 25°C was used to detect changes in secondary processing.
backbone nitrogen of Arg-155. The secondary autoproteolysis For pH analysis of the secondary autocleavage of Y202L, time-
results in cleavage and subsequent dissociation of the pro segment, course monitoring was carried out at different pH values, ranging
producing the ␣ subunit. Autoproteolytic cleavage in Ntn hydro- from 5.0 to 8.0. For reconstitution experiments, ␣, ␣⬘, or ␤ subunit
lases is mediated by intramolecular autocleavage reactions. CA then was separately purified from mature CA, S170C, and E159Q in the

1736 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0507862103 Kim et al.


denatured condition (␣mature and ␤mature, ␣⬘S170C, and ␣⬘E159Q, could be obtained at the 0.5- to 0.8-␴ level depending on mutants.
respectively). They were reconstituted with the combination of At 1.0 ␴, there was no electron density visible for the region. In the
␣mature and ␤mature, ␣⬘S170C and ␤mature, or ␣⬘E159Q and ␤mature by case of Y202L, the pro-segment atoms were modeled with half-
refolding. The results of reconstitution and time-course monitoring occupancy, based on the equimolar ratio of the cleaved and
experiments were analyzed by SDS兾PAGE. uncleaved ␣ subunits on SDS兾PAGE (Fig. 1C, Y202L Inc兾Rec,
lanes 1 and 2). Nevertheless, the electron density for the pro-
Crystallization and Data Collection. Unlike autocleavage analysis, segment backbone was quite clear in the mutant structures. The
only four mutants were crystallized for structural analysis at the residues up to Ala-166, Leu-165, Asp-164, and Asp-164 of the pro
cleavage sites. The crystals of the two Y202L forms (Y202L and segments could be modeled in the Y202L, R226K, S170C, and
Y202L,L), R226K, S170C, and E159Q were grown at 22°C, using E159Q structures, respectively. More importantly, the electron
the hanging-drop vapor diffusion method. One to two microliters of density map for the secondary cleavage site including residues from
protein solution (30 mg兾ml) was mixed with an equal volume of well Glu-159 to Gly-160 was readily interpretable. Water molecules were
solution and equilibrated against 1 ml of the well solution of 0.1 M added by using the Fo ⫺ Fc map peaks ⬎3.0 ␴, if the putative water
Na-cacodylate (pH 6.5), 0.2 M MgCl2, and PEG 6000. For Y202L, molecule made possible hydrogen-bonding contacts and had no
PEG 3000 was used. Crystals were produced with a well developed adverse van der Waals contacts. Some water molecules contacting
bipyramidal morphology within 3–4 days. Diffraction data were the pro segment were added by using the Fo ⫺ Fc map peaks ⬍3.0
collected with the crystals flash cooled at 100 K by using either a ␴, if the B factors were ⬍60 Å2 after refinement. The free R value
laboratory x-ray source or a synchrotron radiation source, beamline was used as an indicator to validate the water picking and refine-
4A at Pohang Light Source (Pohang, Korea). All intensity data ment procedure and to guard against possible overfitting of the data
were processed and scaled by using the programs DENZO and (37). The stereochemical analysis of all of the refined structures by
SCALEPACK (33). The unit cell dimensions of the mutant protein
using PROCHECK (38) showed four prolines [Pro-131 of the ␣
crystals of P41212 in this study were almost identical (see Table 1).
subunit and (Pro-253), (Pro-379), and (Pro-466) of the ␤ subunit]
in a cis-peptide conformation and one residue (Phe-177) of the ␤
Structure Solution and Refinement. The mutant crystal structures of
subunit in a disallowed region of the Ramachandran plot (see Fig.
Y202L, Y202L,L, R226K, S170C, and E159Q were solved by using
8). Fig. 3 was prepared with the programs MOLSCRIPT (39) and
CNS (34) for molecular replacement, using the mature structure as
RASTER3D (40). Figs. 2, 4, 5, 6B, and 7 were prepared with the
a search model (11). The initial solution was optimized by rigid body
refinement, which produced a clearly interpretable electron density program PYMOL (www.pymol.org; ref. 41).
for the overall ␣ and ␤ subunit structures and the mutation site.
We thank the staff at beamline 4A, Pohang Light Source for help with
Manual adjustment of the backbone and side chain and the manual
data collection and Prof. T. Blundell, Dr. Z. Dauter, and Dr. K. H. Han
building of the pro-segment residues was conducted in O (35). for comments on the manuscript. This work was supported by Ministry
Crystallographic refinement was carried out by using the program of Science and Technology Grant R05-2004-10651, Korea Research
REFMAC5 (36). The sigma A weighted 2 Fo ⫺ Fc maps were used to Foundation Grant KRF-2005-C00036, and Biogreen 21 Program 2005-
perform a visual inspection during the rebuilding of the pro 0301034369, Rural Development Administration, Suwon, Korea. J.K.K.,
segment with O (35). After a few rounds of model rebuilding, a I.S.Y., K.J.C., E.K.R., and S.H.K. were supported by the BK21 program
continuous electron density for the residues of the pro segment of the Ministry of Education, Seoul.

1. Davie, E. W. (1995) Thromb. Haemostasis 74, 1–6. 21. Xu, Q., Buckley, D., Guan, C. & Guo, H. (1999) Cell 98, 651–661.
2. Salvesen, G. S. & Dixit, V. M. (1999) Proc. Natl. Acad. Sci. USA 96, 10964–10967. 22. Albert, A., Dhanaraj, V., Genschel, U., Khan, G., Ramjee, M. K., Pulido, R., Sibanda, B. L.,
3. Dissing, M., Giordano, H. & DeLotto, R. (2001) EMBO J. 20, 2387–2393. von Delft, F., Witty, M. & Blundell, T. L., et al. (1998) Nat. Struct. Mol. Biol. 5, 289–293.
4. Perler, F. B. (1998) Cell 92, 1–4. 23. Lee, Y. S. & Park, S. S. (1998) J. Bacteriol. 180, 4576–4582.
5. Babe, L. M. & Craik, C. S. V. (1997) Cell 91, 427–430. 24. Fritz-Wolf, K., Koller, K. P., Lange, G., Liesum, A., Sauber, K., Schreuder, H., Arets, W.
6. Kahn, A. R. & James, M. N. G. (1998) Protein Sci. 7, 815–836. & Kabsch, W. (2002) Protein Sci. 11, 92–103.
7. Perler, F. B. (1998) Nat. Struct. Mol. Biol. 5, 249–252. 25. Kim, S. & Kim, Y. (2001) J. Biol. Chem. 276, 48376–48381.
8. Brannigan, J. A., Dodson, G., Duggleby, H. J., Moody, P. C., Smith, J. L., Tomchick, D. R. 26. Qian, X. Q., Guan, C. & Guo, H. (2003) Structure (London) 11, 997–1003.
& Murzin, A. G. (1995) Nature 378, 416–419. 27. Zacharia, N. & Dougherty, D. A. (2002) Trends Pharmacol. Sci. 23, 281–287.
9. van Poelje, P. D. & Snell, E. E. (1990) Biochemistry 29, 132–139. 28. Luque, L. E., Grape, K. P. & Junker, M. (2002) Biochemistry 41, 13663–13671.
10. Kim, Y., Yoon, K., Khang, Y., Turley, S. & Hol, W. G. (2000) Structure (London) 8, 29. Gavuzzo, E., Pochetti, G., Mazza, F., Gallina, C., Gorini, B., D’Alessio, S., Pieper, M.,
1059–1068. Tschesche, H. & Tucker, P. A. (2000) J. Med. Chem. 43, 3377–3385.
11. Kim, J. K., Yang, I. S., Rhee, S., Dauter, Z., Lee, Y. S., Park, S. S. & Kim, K. H. (2003) 30. Tanner, K. G., Trievel, R. C., Kuo, M. H., Howard, R. M., Berger, S. L., Allis, C. D.,
Biochemistry 42, 4084–4093. Marmorstein, R. & Denu, J. M. (1999) J. Biol. Chem. 274, 18157–18160.
12. Duggleby, H. J., Tolley, S. P., Hill, C. P., Dodson, E. J., Dodson, G. & Moody, P. C. (1995) 31. Weston, S. A., Camble, R., Colls, J., Rosenbrock, G., Taylor, I., Egerton, M., Tucker, A. D.,
Nature 373, 264–268. Tunnicliffe, A., Mistry, A., Mancia, F., et al. (1998) Nat. Struct. Mol. Biol. 5, 213–221.
13. Suresh, C. G., Pundle, A. V., SivaRaman, H., Rao, K. N., Brannigan, J. A., McVey, C. E., 32. Lewendon, A., Murray, I. A., Shaw, W. V., Gibbs, M. R. & Leslie, A. G. (1994) Biochemistry
Verma, C. S., Dauter, Z., Dodson, E. J. & Dodson, G. G. (1999) Nat. Struct. Mol. Biol. 6, 33, 1944–1950.
414–416. 33. Otwinowski, Z. & Minor, W. (1997) Methods Enzymol. 276, 307–326.
14. Smith, J. L., Zaluzec, E. J., Wery, J. P., Niu, L., Switzer, R. L., Zalkin, H. & Satow, Y. (1994) 34. Brunger, A. T., Adams, P. D., Clore, G. M., Delano, W. L., Gros, P., Grosse-Kunstleve,
Science 264, 1427–1433. R. W., Jiang, J. S., Kuszewski, J., Nilges, M., Pannu, N. S., et al. (1998) Acta Crystallogr. D
15. Lowe, J., Stock, D., Jap, B., Zwickl, P., Baumeister, W. & Huber, R. (1995) Science 268, 54, 905–921.
533–539. 35. Jones, T. A., Zou, J. Y., Cowan, S. W. & Kjeldgaard, M. (1991) Acta Crystallogr. A 47,
16. Oinonen, C., Tikkanen, R., Rouvinen, J. & Peltonen, L. (1995) Nat. Struct. Mol. Biol. 2, 110–119.
1102–1108. 36. Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997) Acta Crystallogr. D 53, 240–255.
17. Bompard-Gilles, C., Villeret, V., Davies, G. J., Fanuel, L., Joris, B., Frere, J. M. & Van 37. Brunger, A. T. (1992) Nature 355, 472–475.
Beeumen, J. (2000) Structure (London) 8, 153–162. 38. Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thorton, J. M. (1993) J. Appl.
18. Kim, Y., Kim, S., Earnest, T. N. & Hol, W. G. J. (2001) J. Biol. Chem. 277, 2823–2829. Crystallogr. 26, 283–291.
19. Hewitt, L., Kasche, V., Lummer, K., Lewis, R. J., Murshudov, G. N., Verma, C. S., Dodson, 39. Kraulis, P. J. (1991) J. Appl. Crystallogr. 24, 946–950.
G. G. & Wilson, K. S. (2000) J. Mol. Biol. 302, 887–896. 40. Merritt, E. A. & Murphy, M. E. P. (1994) Acta Crystallogr. D 50, 869–873.
20. Ditzel, L., Huber, R., Mann, K., Heinemeyer, W., Wolf, D. H. & Groll, M. (1998) J. Mol. 41. DeLano, W. L. (2002) The PYMOL Molecular Graphics System (DeLano Scientific, San
Biol. 279, 1187–1191. Carlos, CA).
BIOPHYSICS

Kim et al. PNAS 兩 February 7, 2006 兩 vol. 103 兩 no. 6 兩 1737

You might also like