ArticlePDF Available

Crystal Structure of Neuropsin, a Hippocampal Protease Involved in Kindling Epileptogenesis

Authors:

Abstract

Neuropsin is a novel serine protease, the expression of which is highly localized in the limbic areas of the mouse brain and which is suggested to be involved in kindling epileptogenesis and hippocampal plasticity. The 2.1-Å resolution crystal structure of neuropsin provides the first three-dimensional view of one of the serine proteases highly expressed in the nervous system, and reveals a serine protease fold that exhibits chimeric features between trypsin and nerve growth factor-γ (NGFγ), a member of the kallikrein family. Neuropsin possesses anN-glycosylated “kallikrein loop” but forms six disulfide bonds corresponding to those of trypsin. The ordered kallikrein loop projects proline toward the active site to restrict smaller residues or proline at the P2 position of substrates. Loop F, which participates in forming the S3/S4 sites, is similar to trypsin rather than NGFγ. The unique conformations of loops G and H form an S1 pocket specific for both arginine and lysine. These characteristic loop structures forming the substrate-binding site suggest the novel substrate specificity of neuropsin and give a clue to the design of its specific inhibitors.
Crystal Structure of Neuropsin, a Hippocampal Protease
Involved in Kindling Epileptogenesis*
(Received for publication, September 11, 1998)
Tadaaki Kishi‡, Masato Kato‡, Toshiyuki Shimizu, Keiko Kato§, Kazumasa Matsumoto§,
Shigetaka Yoshida§, Sadao Shiosaka§, and Toshio Hakoshima
From the Department of Molecular Biology and the §Department of Cell Biology, Nara Institute of Science and
Technology, 8916-5 Takayama, Ikoma, Nara 630-0101, Japan
Neuropsin is a novel serine protease, the expression of
which is highly localized in the limbic areas of the
mouse brain and which is suggested to be involved in
kindling epileptogenesis and hippocampal plasticity.
The 2.1-Å resolution crystal structure of neuropsin pro-
vides the first three-dimensional view of one of the ser-
ine proteases highly expressed in the nervous system,
and reveals a serine protease fold that exhibits chimeric
features between trypsin and nerve growth factor-
g
(NGF
g
), a member of the kallikrein family. Neuropsin
possesses an N-glycosylated “kallikrein loop” but forms
six disulfide bonds corresponding to those of trypsin.
The ordered kallikrein loop projects proline toward the
active site to restrict smaller residues or proline at the
P2 position of substrates. Loop F, which participates in
forming the S3/S4 sites, is similar to trypsin rather than
NGF
g
. The unique conformations of loops G and H form
an S1 pocket specific for both arginine and lysine. These
characteristic loop structures forming the substrate-
binding site suggest the novel substrate specificity of
neuropsin and give a clue to the design of its specific
inhibitors.
Proteases have been shown to play essential roles in the
nervous system, including those of neurite outgrowth (1), neu-
ral degeneration (2), and synaptic plasticity (3). These actions
are thought to be mediated by the proteolytic cleavage of zy-
mogen precursors, the activation of specific cell surface recep-
tors, or the degradation of extracellular matrix proteins (4).
Neuropsin was cloned from a mouse hippocampal cDNA library
using sequences for key regions of the serine protease domain
of nerve growth factor (NGF)-
g
1
(5). In the brain, no NGF
g
has
been identified so far, though NGF
b
is present. Neuropsin is
one of the serine proteases highly expressed in the nervous
system (68). The expression of neuropsin is localized at high-
est concentration in the hippocampus and the amygdala, which
are important for acquisition of memory and emotional mem-
ory, respectively. This localization is in contrast to that of
tissue plasminogen activator (tPA), which is well documented
to play a crucial role in the nervous system by mediating
plasticity but is distributed more uniformly across the other
brain regions and throughout other organs (9). Activity-de-
pendent changes in expression of neuropsin have been ob-
served upon direct hippocampal stimulation and induction of
kindling, which is a model for epilepsy and neuronal plasticity
characterized by the progressive development of electrographic
and behavioral seizures (5, 10). A single intraventricular injec-
tion of monoclonal antibodies specific to neuropsin reduces or
eliminates the epileptic pattern (11). Moreover, oxidative stress
is shown to effect the expression of neuropsin in the limbic
areas, which might be related to the disturbance in shock-
avoidance learning of mice (12). These activity-dependent
changes and the specific localization of neuropsin indicate the
involvement of this protease in hippocampal plasticity and its
pathogenesis. Knowledge of the three-dimensional structure of
neuropsin provides clues to the biological activity of this pro-
tease and also is important to the design of inhibitors that
might be useful in treatment of pathological conditions such as
epilepsy.
EXPERIMENTAL PROCEDURES
Protein Preparation and Crystallization—Neuropsin was over-ex-
pressed in baculovirus-infected High Five insect cells, purified, and
crystallized as described previously (13). The resulting sample and the
crystallized protein were verified with N-terminal analysis using an
Applied Biosystem automatic analyzer 476A. Neuropsin has a putative
glycosylation site at Asn
95
of the kallikrein loop. Time-of-flight mass
spectroscopy with PerSeptive JMS-ELITE matrix-associated laser des-
orption/ionization time-of-flight indicated its heterogeneous glycosyla-
tion. Two-dimensional high performance liquid chromatography map-
ping (14) revealed that the N-glycans contained 89% paucimannosidic
structures with and without attached fucose residue(s) at the innermost
GlcNAc residues but the glycosylation pattern exhibited high heteroge-
neity as found on many glycoproteins (15). The detailed procedures and
obtained structures will be described elsewhere. The crystals belong to
space group P1(a 5 38.15 Å, b 5 54.95 Å, c 5 64.29 Å,
a
5 95.72°,
b
5
90.03°, and
g
5 110.29°). X-ray diffraction data were collected with
Rigaku imaging plate area-detectors, R-Axis IV and R-Axis IIc, using
Cu-K
a
radiation and also with a Weissenberg camera at the BL-18B
beamline station of the Photon Factory, Tsukuba using 1-Å radiation.
Intensities were evaluated with the program DENZO/SCALEPAK (16),
which yielded 25,778 independent combined reflections corresponding
to 90.7% completeness at 2.1-Å resolution (74.5% in the highest reso-
lution bin), an R
merge
of 6.0% (19.8%), a mean ratio of intensity, and
s
of 8.5 (3.1).
Structure Determination—Initial phases were calculated by molecu-
lar-replacement with the program AMoRe (17) using a search model
based on the structure of bovine pancreatic
b
-trypsin (PDB code 4PTP)
(18). Rigid body refinements of the searched model were performed with
the program X-PLOR (19), followed by density averaging/histogram
* This work was supported by Grants-in-Aid for Scientific Research
(09308025, 10359003), on Priority Areas (10179104, 09277102), Bio-
metalics (09235220) (to T. H.) from the Ministry of Education, Science,
Sports and Culture, Japan. The costs of publication of this article were
defrayed in part by the payment of page charges. This article must
therefore be hereby marked advertisement in accordance with 18
U.S.C. Section 1734 solely to indicate this fact.
The atomic coordinates and structure factors (code 1NPM) have been
deposited in the Protein Data Bank, Brookhaven National Laboratory,
Upton, NY.
‡Supported by a research fellowship for young scientists from the
Japan Society for the Promotion of Science.
A member of the TARA project of Tsukuba University. To whom
correspondence should be addressed. Tel.: 0743-72-5570; Fax: 0743-72-
5579; E-mail: hakosima@bs.aist-nara.ac.jp.
1
The abbreviations used are: NGF, nerve growth factor; tPA, tissue
plasminogen activator; GlcNAc, N-acetylglucosamine; STI, soybean
trypsin inhibitor; MCA, 4-methylcoumaryl-7-amide; MSP, myelenceph-
alon-specific protease; r.m.s., root mean square.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 274, No. 7, Issue of February 12, pp. 4220 –4224, 1999
© 1999 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org4220
by guest on December 27, 2015http://www.jbc.org/Downloaded from
matching with the program DM (20). Six regions of insertions and
deletions were inspected on the resulting 2F
o
2 F
c
map, which was
generated with the program O (21). The structure was built and refined
through alternating cycles using the programs O and X-PLOR, respec-
tively. The kallikrein loop had large but mostly poor density connects to
the side chain of Asn
95
. After several cycles of refinements incorporat-
ing solvent water molecules located at the regions other than the
kallikrein loop, we defined one residue of N-acetylglucosamine (Glc-
NAc) residue bonded to Asn
95
, as well as weaker density for additional
sugar residues that we have been unable to identify definitively. The
residual weak density is extending toward a large solvent channel in
the crystal.
Current Structure—Two regions were poorly defined in the map. The
first is at the loop residues, Arg
74
and Asp
75
, and the second is at the
three C-terminal residues. These have uninterpretable densities imply-
ing complex disorder. The current structure contains 194 water mole-
cules. The R-factor is 18.6% (an R
free
of 22.7%) for all reflections to 2.1-Å
resolution. The root-mean-square (r.m.s.) deviations from target values
are 0.008 Å for bond lengths, 1.535° for bond angles, and 1.139° for the
peptide torsion angles. The averaged B-factor is 30.8 Å
2
. There is no
residue in disallowed regions as defined in PROCHECK (22), but 89.4%
residues in the most favorable regions and 10.6% residues in the addi-
tional allowed regions.
RESULTS
Overall Structure—Neuropsin consists of fourteen
b
-strands
(designated as
b
1–
b
14) that are extensively twisted, two
a
-hel-
ices (designated as
a
1 and
a
2), and one short 3
10
-helix (Fig. 1a).
Each seven-
b
-strand forms an antiparallel
b
-sheet folded in a
b
-sandwich with a cleft where the catalytic triad (Asp
57
, His
102
,
and Ser
195
) is located (Fig. 1, a and b). This overall structure is
homologous to those of the chymotrypsin-type serine proteases,
which share an identical catalytic mechanism, but among
which the substrate specificity varies. Many known structures
of these proteases delineate a clear framework, demonstrating
that this variety is a function of evolved diversity in the struc-
tures of surface loops that surround the substrate-binding site.
Because the loops of neuropsin, which contains eight promi-
nent loops (AH in Fig. 1a), are conserved in their relative
positions with respect to the active site, general themes for
their individual functions can be derived.
One of the characteristic features of neuropsin is the N-
glycosylated loop D that corresponds to the so-called “kallikrein
loop.” This loop, having an Asn-X-Ser sequence, is typical for
members of the kallikrein family that contains NGF
g
, which
exhibits relatively high (46%) sequence identity to neuropsin
(Fig. 2). Neuropsin, however, forms six disulfide bonds corre-
sponding to those of trypsin with an additional disulfide bond
(SS3 between Cys
128
and Cys
232
in Fig. 1a) that is missing in
members of the kallikrein family. Large differences exist in the
loop regions surrounding the substrate-binding site, whereas
the core region contains only minor variations. Excluding the
insertion and deletion residues, the main-chain atoms of neu-
ropsin superimpose on the corresponding atoms of bovine pan-
creatic trypsin (18), mouse submaxillary gland NGF
g
in7S
NGF (23), an
a
2
b
2
g
2
complex of NGF, and pancreatic porcine
kallikrein (24) with r.m.s. deviations of 1.26, 1.43, and 1.84 Å,
respectively. The geometry of the catalytic triad is highly sim-
ilar to those of the serine proteases with r.m.s. deviations in a
range of 0.2–0.24 Å. Neuropsin has no prominent structural
similarity to tPA, showing a high r.m.s. deviation of 2.74 Å for
79 identical residues (37% sequence identity).
S1 Site—Enzyme assay using several 4-methylcoumaryl-7-
amide (MCA) derivatives of oligopeptides (25) has shown that
neuropsin cleaves peptide bonds C-terminal to Arg or Lys. This
primary specificity is well interpreted by the S1 pocket, a deep
cylindrical pocket that is formed by two loops, G and H, and
punctuated at its base by the side chain of Asp
189
. However,
neuropsin has large conformational changes of loop G with
maximum displacements of 4.8–5.8 Å compared with those of
NGF
g
, kallikrein, and trypsin (Fig. 3). The conformational
changes in neuropsin seem to be caused by the one-residue
(Gly
186B
) deletion in loop G. In addition, loop H of neuropsin
also displays large displacements from these proteases because
FIG.1.Overall structure of neuropsin. a, ribbon representation of neuropsin. Seven-stranded
b
-sheets (the top and bottom halves) are
sandwiched with the catalytic triad at the cleft of the
b
-sandwich. The surface loops (A–H) forming the substrate-binding site are colored with
labels. Six disulfide bonds are shown by bridges in white, and the disulfide bond (SS3), which is conserved in trypsin but not in kallikrein, was
labeled. Loop D is the kallikrein loop that has an N-glycosylated Asn
95
with one visible GlcNAc residue. The side chains of the catalytic triad, Asp
189
at the S1-specific pocket, Lys
175
at the S3/4 site, Glu
149
and Asp
218
at the rim of the S1 pocket, and Leu
40
and Ile
41
at the S19 site, are also shown
with stick representations with one-letter amino acid labels. b, molecular surfaces of neuropsin viewed from nearly the same direction of panel a.
Surface electrostatic potentials calculated and rendered using GRASP (negative potentials are in red and positive in blue). The S1–S4 and S19 sites
and characteristic surface residues are labeled.
Crystal Structure of Brain Neuropsin 4221
by guest on December 27, 2015http://www.jbc.org/Downloaded from
loop H is heavily interacted with loop G in all the proteases. It
is noteworthy that loop H of trypsin has a one-residue deletion
of the cis proline, Pro
219
, that is conserved in neuropsin, NGF
g
,
and kallikrein. This deletion induces a larger displacement of
neuropsin loop H from trypsin (4.3 Å) than from NGF
g
(3.4 Å)
or kallikrein (3.2 Å). Interestingly, the P1 specificity of neurop-
sin for arginine is comparable with that for lysine. This is in
sharp contrast to NGF
g
, kallikrein, and trypsin, in which a
significant preference for arginine exists. Among key residues
of the S1 pocket, Gly
226
of neuropsin has relatively large dis-
placements from NGF
g
(0.7 Å) and kallikrein (1.2 Å). Alterna-
tively, Ser
217
of neuropsin has a relatively large displacement
from trypsin (1.3 Å). Compared with NGF
g
, the changes of the
neuropsin loop structures result in positional displacements
(0.40.8 Å) of Asp
189
, Thr
190
, and His
217
, which have been
reported to form hydrogen bonds to the P1 arginine of NGF
b
in
7 S NGF (23). These local differences may be responsible for the
unique P1 specificity of neuropsin. Like other serine proteases,
which are activated by cleavage of the bond between Arg
15
and
Ile
16
, neuropsin tucks the newly formed amino group of Ile
16
into the pocket to form multiple hydrogen bonds with the main
chains of loop E and to form an ion pair with Asp
194
of loop G.
Kallikrein Loop—The kallikrein loop of neuropsin differs
radically from those of NGF
g
and kallikrein. In these members
of the kallikrein family, the loop was cleaved into the highly
mobile nicked chains. In contrast, the loop of neuropsin is
packed into an ordered and relatively compact conformation
without any nicked site: neuropsin has no arginine or lysine
residue in the loop. Recently determined crystal structure (26)
of mouse glandular kallikrein-13 shows an ordered kallikrein
loop, but no conformational similarity with the loop of neurop-
sin. It seems unlikely that the N-glycan bound to Asn
95
partic-
ipates directly in the substrate binding because the GlcNAc
residue and the residual density are oriented away from the
active site as in members of the kallikrein family.
S2 Site—The kallikrein loop of neuropsin overhangs toward
the active site cleft with a prominent Pro
95D
residue, which
suggests its role in substrate binding. Interestingly, superpo-
sition of neuropsin on porcine pancreatic trypsin complexed
with soybean trypsin inhibitor (STI) (27) revealed steric clashes
between the kallikrein loop of neuropsin and the two STI loops
facing toward neuropsin. This was borne out by biochemical
experiments in which high molecular weight inhibitors, such as
STI or
a
-antitrypsin, were found to have little effect on the
neuropsin activity, whereas low molecular weight inhibitors,
such as leupeptin, markedly inhibited the activity. The over-
hanging kallikrein loop forms a narrow pocket (the S2 site) in
which Asp
102
is positioned at the base and would restrict the
size of the side chain in the P2 position of substrate peptides.
This is consistent with the results of a previous enzyme assay
(25), in which high activities of neuropsin were observed for
peptide substrates having smaller residues or proline in the P2
positions. It has been well demonstrated by the crystal struc-
ture of thrombin complexed with
D-Phe-Pro-Arg that loop B of
thrombin compresses the S2 site with the inserted residues,
Tyr
60A
and Trp
60D
, to deduce the P2 specificity of the enzyme
for proline (28). Superposition of neuropsin on thrombin shows
Pro
95D
of neuropsin located nearby Tyr
60A
and Trp
60D
of throm-
bin, but no contact between Pro
95D
and the proline residue of
D-Phe-Pro-Arg bound to thrombin, which suggests that the P2
preference for proline may be mediated by the kallikrein loop of
neuropsin, instead of loop B of thrombin, but rather weaker
than that of thrombin (Fig. 4). Phenylalanine at the P2 position
remarkably reduces the neuropsin activity, which is one of the
major differences from kallikrein and NGF
g
.
S3/S4 Site—Most striking is the structure of loop F that is
similar to that of trypsin rather than NGF
g
and kallikrein,
where significant conformational changes of loop F from neu-
ropsin occur with large displacements, 5.2 and 6 Å, respec-
tively, accompanied by movements of helix
a
1 (Fig. 3). Loop F
is one of the main elements forming the S3/S4 site. It is notable
that this loop has Tyr
172
, which forms hydrogen bonds with the
main chains of loop H and which is conserved in trypsin but is
replaced by histidine in NGF
g
and kallikrein. This residue has
been elucidated to be one of the distal determinant residues for
the substrate specificity (29). Moreover, Gly
174
, which is con-
served in neuropsin and trypsin, is one of the key residues for
the loop F structure because this residue is an essential com-
ponent of the type-II reverse turn, YPGK at 172–175. It should
be pointed out that the disulfide bond SS3, which is conserved
in trypsin as already mentioned, may contribute to the confor-
mational resemblance of loop F with that of trypsin through
contacts with strand
b
13 that associated with the C-terminal
b
-strand of loop F,
b
11.
FIG.2. Secondary structural elements and sequence align-
ment of neuropsin and the related proteases, mouse submaxil-
lary gland NGF
g
, porcine pancreatic kallikrein A, and bovine
pancreatic
b
-trypsin. The sequential numbering of neuropsin is
shown at the top, and the chymotrypsinogen based used throughout this
paper is at the bottom. The secondary structural elements of neuropsin
are shown at the top with arrows for
b
-strands (
b
1
b
14), cylinders for
a
-helices (
a
1,
a
2), and 3
10
-helix. The loops (A–H) forming the substrate-
binding site are marked with colored bars with labels. Insertion or
deletion sites are marked by heavy line boxes. The disulfide-bond-
forming cysteines are boxed and marked by labels (SS1–SS6), which
correspond to six disulfide-bonds, respectively. The catalytic triad is
marked by red circles. Asp
189
in the S1-specific pocket is marked by a
blue star. Gly
216
and Gly
226
of the rim of the S1 pocket are marked by
black stars. The putative N-glycosylation sequence of Asn-X-Ser, which
is conserved in members of the kallikrein family, two cis prolines of
neuropsin, Pro
147
and Pro
219
, and Tyr
172
-Pro
173
-Gly
174
of loop F are
boxed.
Crystal Structure of Brain Neuropsin4222
by guest on December 27, 2015http://www.jbc.org/Downloaded from
It is remarkable that the above differences in the loop F may
also be correlated with the conformational differences in the
kallikrein loops. In NGF
g
and kallikrein, the highly mobile
kallikrein loops heavily interact with loop F although the kal-
likrein loop of neuropsin has no direct interaction with loop F,
as it does with loop D of trypsin. One of the interesting conse-
quences of these conformational characteristics in loops D and
F is that the S3/S4 site of neuropsin is similar to that of trypsin
rather than NGF
g
. The aromatic rings of Trp
215
, Tyr
172
, and
His
99
provide a shallow but wide hydrophobic depression for
the S3/S4 site as in trypsin and could explain the high activities
of neuropsin observed for synthetic tripeptide substrates hav-
ing hydrophobic residues in the P3 position. Moreover, in neu-
ropsin, Lys
175
of loop F is projected toward the S3/S4 site (Fig.
1b), whereas Lys
175
of NGF
g
is projected away from the S3/S4
site. It is interesting that Lys
175
may play a role in the P3/P4
interaction with the substrates. Actually, one of the cleavage
sites of fibronectin (25), which is an extracellular matrix pro-
tein exhibiting strong proteolytic sensitivity for neuropsin, had
the N-terminal sequence of Asp-Val-Arg, whose acidic residue
at the P3 position may interact with Lys
175
.
DISCUSSION
A tripeptide substrate preferred for thrombin, Val-Pro-Arg-
MCA, has been found to exhibit the highest sensitivity for
neuropsin to date. However, poor structural homology between
neuropsin and thrombin is evident from the large r.m.s. devi-
ation of 2.7 Å for 78 identical residues. Moreover, thrombin
cleaves Val-Pro-Arg-MCA much faster (33-fold) than Phe-Ser-
Arg-MCA, which is one of the preferred substrates of trypsin,
whereas the activities of neuropsin for these substrates are
comparable. In addition, thrombin also exhibits a significant
preference for arginine at the P1 position, but no such prefer-
ence was observed for neuropsin, as described above. These
results, together with structural differences of several loops,
FIG.3.Comparison of the surface loops forming the substrate-binding site between neuropsin (colored) and the related proteases
(gray): NGF
g
-NGF
b
in 7 S NGF (a) and trypsin-leupeptin complexes (b). The C
a
-carbon atom tracings of loops B, D, and F–H of neuropsin
are colored as in Fig. 1a with both the N- and C-terminal residue numbers in black and superimposed on those of the proteases with the bound
substrate or inhibitor (green); the C-terminal Ala-Thr-Arg of NGF
b
in panel a, or leupeptin in panel b. The side chains of the key residues of the
loops and the catalytic triad of neuropsin are shown with yellow labels, but the others are not shown for the clarity. Some of the side chains and
the disulfide bond SS6 of NGF
g
and trypsin are also shown with white labels. These are Trp
215
, His
172
, Lys
175
, Asp
189
, and the disulfide bond SS6
of NGF
g
in panel a, and Trp
215
, Tyr
172
, Asp
189
, and the disulfide bond SS6 of trypsin. The hydrogen bonds between Tyr
172
(His
172
of NGF
g
) and
loop H are indicated by white (black for His
172
) broken lines.
FIG.4. Comparison of the surface loops forming the S2 site
between neuropsin (colored) and
a
-thrombin (gray). The C
a
-car-
bon atom tracings of loops of neuropsin, colored as in Fig. 1a, are
superimposed on
a
-thrombin complexed with D-Phe-Pro-Arg-chloro-
methylketone (green). The van der Waals surfaces of the inhibitor
peptide and P95D are shown with dot-surface representations. Labels
are as in Fig. 3.
Crystal Structure of Brain Neuropsin 4223
by guest on December 27, 2015http://www.jbc.org/Downloaded from
verify the unique substrate specificity of neuropsin, even if the
P2 preference for proline is analogous to thrombin.
Compared with the subsite preferences on the N-terminal
side of the scissile bond, little is known about subsite prefer-
ences on the C-terminal side at present. However, a shallow
bowl formed by Cys
42
, Ile
41
, and Leu
40
of strand
b
3 seems to
provide a hydrophobic S19 site (Fig. 1b). The shape of the
substrate-binding surface and the surface electrostatic distri-
bution of neuropsin display several differences in details from
other serine proteases. One of the pronounced characteristics is
the rim of the S1 pocket, where acidic residues, Asp
218
and
Glu
149
, expose the side chains to the solvent region. Glu
97
of the
kallikrein loop also is projected from the surface. These char-
acteristics may be related to the specificity that are distinct
from those of other proteases. Neuropsin exhibits weak proteo-
lytic activities against gelatin and collagen but effectively
cleaves fibronectin, as already mentioned. By changing the
extracellular environment, neuropsin may exert its limbic
effects.
It is an interesting question whether neuropsin could process
NGF
b
precursor and form 7 S NGF instead of NGF
g
.In7S
NGF, the active site of NGF
g
was occupied by the C-terminal
Arg
118
of the mature NGF
b
, as a cleaved product, with exten-
sive interactions of the large nicked kallikrein loop with the
C-terminal regions and
b
-strand of NGF
b
. Docking studies
suggest that neuropsin would lose most of these interactions
although the smaller residues, Thr
117
and Ala
116
,ofNGF
b
could fit to the S2 and S3 sites of neuropsin. In addition, the
small cavity of NGF
g
for Ala
116
of NGF
b
is missing in neurop-
sin. Moreover, Lys
192
, which is located at the rim of the active
site of NGF
g
and forms hydrogen bonds to main chain carbon-
yls of Thr
117
and Lys
74
of NGF
b
, is replaced by Gln
192
in
neuropsin. In 7 S NGF, a zinc ion was located at the interface
between NGF
g
and NGF
a
to stabilize the complex. This coor-
dination is also lost when NGF
g
is replaced by neuropsin
because zinc-coordinated His
217
and Glu
222
of NGF
g
are re-
placed by serine and lysine in neuropsin. Taken together, these
results suggest that neuropsin is incapable of forming 7 S NGF,
even if neuropsin could process the NGF
b
precursor. However,
this will require further investigation.
In addition to neuropsin, three other serine proteases have
been reported to be more highly expressed in the central nerv-
ous system than in most peripheral tissues. These include
myelencephalon (MSP)-specific protease (7), neurosin (8), and
neurotrypsin (6). MSP and neurosin exhibit sequence identities
to neuropsin of 48 and 46%, respectively. Neurotrypsin, which
is a multidomain serine protease whose expression is most
prominent in the cerebral cortex, hippocampus, and amygdala,
has a protease domain exhibiting 33% sequence identity to
neuropsin. Sequence alignments with neuropsin indicate that
these proteases would have different structures of surface loops
surrounding the substrate-binding site. Remarkably, loop D of
either of these proteases has no N-glycosylation site and no
inserted residues. This lack of a kallikrein loop would result in
their P2 specificities differing from that of neuropsin. More-
over, loop G of MSP and neurosin has no one-residue deletion,
which causes significant structural changes of loops G and H
forming the S1 pocket. Alternatively, compared with neurop-
sin, neurotrypsin has a three-residue insertion in loop G and a
one-residue deletion in loop H. These differences would endow
these other proteases with substrate specificities different from
that of neuropsin.
In conclusion, many aspects of neuropsin structure and func-
tion reveal that this hippocampal serine protease displays chi-
meric structural features of trypsin and NGF
g
with novel sub-
strate specificity. These findings could give a clue to the
structure-based drug design useful in treatment of pathological
conditions such as epilepsy and also useful in analyzing the
processes of synaptic plasticity.
Acknowledgments—We thank Drs. Ben Bax (Birkbeck College) and
S. E. Won Suh (Seoul National University) for providing the coordinates
of 7 S NGF and soybean trypsin inhibitor-trypsin complex, respectively;
M. Suzuki for data collection at PF; and S. Takayama and J. Tsukamoto
for help with the mass spectroscopy and N-terminal analysis.
REFERENCES
1. Monard, D. (1988) Trends Neurosci. 11, 541–544
2. Tsirka, S. E., Gualandris, A., Amaral, D. G., and Strickland, S. (1995) Nature
377, 340–344
3. Liu, Y., Fields, R. D., Festoff, B. W., and Nelson, P. G. (1994) Proc. Natl. Acad.
Sci. U. S. A. 91, 10300 –10304
4. McGuire, P. G., and Seeds, N. W. (1990) Neuron 4, 633–642
5. Chen, Z.-L., Yoshida, S., Kato, K., Momota, Y., Suzuki, J., Tanaka, T., Ito, J.,
Nishino, H., Aimoto, S., Kiyama, H., and Shiosaka, S. (1995) J. Neurosci.
15, 5088–5097
6. Gschwend, T. P., Krueger, S. R., Kozlov, S. V., Wolfer, D. P., and Sonderegger,
P. (1997) Mol. Cell. Neurosci. 9, 207–219
7. Scarisbrick, I. A., Towner, M. D., and Isackson, P. J. (1997) J. Neurosci. 17,
81568168
8. Yamashiro, K., Tsuruoka, N., Kodama, S., Tsujimoto, M., Yamamura, Y.,
Tanaka, T., Nakazato, H., and Yamagichi, N. (1997) Biochim. Biophys. Acta
1350, 11–14
9. Qian, Z., Gilbert, M. E., Colicos, M. A., Kandel, E. R., and Kuhl, D. (1993)
Nature 361, 453–457
10. Okabe, A., Momota, Y., Yoshida, S., Hirata, A., Ito, J., Nishino, H., and
Shiosaka, S. (1996) Brain Res. 728, 116–120
11. Momota, Y., Yoshida, S., Ito, J., Shibata, M., Kato, K., Sakurai, K., Matsumoto,
K., and Shiosaka, S. (1998) Eur. J. Neurosci. 10, 760–764
12. Akita, H., Matsuyama, T., Iso, H., Sugita, M., and Yoshida, S. (1997) Brain
Res. 769, 86–96
13. Kishi, T., Kato, M., Shimizu, T., Kato, K., Matsumoto, K., Yoshida, Y.,
Shiosaka, S., and Hakoshima, T. (1997) J. Struct. Biol. 118, 248–251
14. Tomiya, N., Awaya, J., Kurono, M., Endo, S., Arata, Y., and Takahashi, N.
(1988) Anal. Biochem. 171, 73–90
15. Friedrich, A. (1996) Trends Glycosci. Glycotech. 40, 101–114
16. Otwinowski, Z., and Minor, W. (1996) Methods Enzymol. 276, 307–326
17. Navaza, J. (1994) Acta Crystallogr. Sec. A 50, 157–163
18. Finer-Moore, J. S., Kossiakoff, A. A., Hurley, J. H., Earnest, T., and Stroud,
R. M. (1992) Proteins Struct. Finct. 12, 203–222
19. Brunger, A. T., Krukowski, A., and Erickson, J. W. (1990) Acta Crystallogr.
Sec. A 46, 585–593
20. Cowtan, K. D., and Main, P. (1996) Acta Crystallogr. Sec. D 52, 43–48
21. Jones, T. A., Zou, J.-Y., Cowan, S. W., and Kjeldgaard, M. (1991) Acta
Crystallogr. Sec. A 47, 110–119
22. Laskowski, R. A., MacArthur, M. W., Moss, D. S., and Thornton, J. M. (1993)
J. Appl. Crystallogr. 26, 283–291
23. Bax, B., Blundell, T. L., Murray-Rust, J., and McDonald, N. Q. (1997)
Structure (Lond.) 5, 1275–1285
24. Bode, W., Chen, Z., Bartels, K., Kutzbach, C., Schmidt-Kastner, G., and
Bartunik, H. (1983) J. Mol. Biol. 164, 237–282
25. Shimizu, C., Yoshida, S., Shibata, M., Kato, K., Momota, Y., Matsumoto, K.,
Shiosaka, T., Midorikawa, R., Kamachi, T., Kawabe, A., and Shiosaka, S.
(1998) J. Biol. Chem. 273, 11189–11196
26. Timm, D. E. (1997) Protein Sci. 6, 1418–1425
27. Song, H. K., and Suh, S. W. (1998) J. Mol. Biol. 275, 347–363
28. Bode, W., Mayr, I., Baumann, U., Huber, R., Stone, S. R., and Hofsteenge, J.
(1989) EMBO J. 8, 3467–3475
29. Hedstrom, L., Perona, J. J., and Rutter, W. J. (1994) Biochemistry 33,
8757–8763
Crystal Structure of Brain Neuropsin4224
by guest on December 27, 2015http://www.jbc.org/Downloaded from
Toshio Hakoshima
Shigetaka Yoshida, Sadao Shiosaka and
Shimizu, Keiko Kato, Kazumasa Matsumoto,
Tadaaki Kishi, Masato Kato, Toshiyuki
Kindling Epileptogenesis
Hippocampal Protease Involved in
Crystal Structure of Neuropsin, a
STRUCTURE:
PROTEIN CHEMISTRY AND
doi: 10.1074/jbc.274.7.4220
1999, 274:4220-4224.J. Biol. Chem.
http://www.jbc.org/content/274/7/4220Access the most updated version of this article at
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted
When this article is cited
to choose from all of JBC's e-mail alertsClick here
http://www.jbc.org/content/274/7/4220.full.html#ref-list-1
This article cites 29 references, 4 of which can be accessed free at
by guest on December 27, 2015http://www.jbc.org/Downloaded from
... Figure 1(a) shows the human neuropsin structure. Neuropsin has fourteen β-strands, three α-helices and a catalytic triad (His57, Asp102 and Ser195) which is located at the cleft formed by the seven β-strands [15]. The substrate pocket, where the catalytic triad is located at its bottom, is surrounded by the eight loops, Loops A ~ H as shown in Figure 1 (a). ...
... We built the three mouse neuropsin-substrate complex model structures for each substrate based on the crystal structure of the mouse hippocampus neuropsin (PDB ID: 1NPM, UniProtKB: Q61955) [15], since the three-dimensional (3D) neuropsin-substrate complex structures with the five substrates in question were not available in the Protein Data Bank. The sequence identity and similarity between human and mouse neuropsins were 75% and 91%, respectively. ...
Article
Full-text available
Neuropsin is one of serine proteases mainly found at the hippocampus and the amygdala, where it contributes to the long-term potentiation and memory acquisition by rebuilding of synaptic connections. Despite of the importance of neuropsin, the substrate specificity and regulation mechanisms of neuropsin have been unclear. Thus, we investigated the substrate specificity and the catalytic activity of neuropsin by the protein-ligand docking and molecular dynamics (MD) simulations and succeeded to reproduce the trend of the experimental results. Our study revealed that the substrate specificity and the activity of neuropsin depended on multiple factors: the substrate charge, the substrate orientation, the hydrogen bond network within the catalytic triad and the substrate, and the formation of the oxyanion hole. The apo neuropsin was not reactive without proper alignment of catalytic triad. The substrate binding induced the reactive alignment of catalytic triad. Then the substrate-neuropsin interaction forms the oxyanion hole that stabilizes the transition state and reduces the free-energy barrier of the following scission reaction. Fullsize Image
... In order to better assess the ability of the basis to reproduce the general dynamics of chymotrypsin-like proteases, we performed MD simulations of four proteins belonging to the same family and compared the per-residue fluctuations emerging from the simulations with those obtained by filtering the trajectory along the vectors of the basis; a good agreement would be indicative of the ability of the basis to describe the large-scale dynamics of the protein. Two of the proteins used as a test-case belong to the dataset; these are 1EKB [84] and 1NPM [85], eukaryotic proteases belonging to the S1A subfamily. The other two proteins, 4YOG [86] and 3W94 [87], are external to the dataset and, as such, have not been used to define the basis. ...
Article
Full-text available
The paradigmatic sequence–structure–dynamics–function relation in proteins is currently well established in the scientific community; in particular, a large effort has been made to probe the first connection, indeed providing convincing evidence of its strength and rationalizing it in a quantitative and general framework. In contrast, however, the role of dynamics as a link between structure and function has eluded a similarly clear-cut verification and description. In this work, we propose a pipeline aimed at building a basis for the quantitative characterization of the large-scale dynamics of a set of proteins, starting from the sole knowledge of their native structures. The method hinges on a dynamics-based clusterization, which allows a straightforward comparison with structural and functional protein classifications. The resulting basis set, obtained through the application to a group of related proteins, is shown to reproduce the salient large-scale dynamical features of the dataset. Most interestingly, the basis set is shown to encode the fluctuation patterns of homologous proteins not belonging to the initial dataset, thus highlighting the general applicability of the pipeline used to build it.
... The presence of additional P 4 Val, optimal P 3 Arg and, most importantly, a large P 2 Phe residue in our substrate libraries could affect the cooperativity of subsites and result in the observed varied selectivity of the enzyme with different substrate design. Indeed, KLK-family members are characterized by similar structure and organization of the substrate binding cleft [23,24], yet differ significantly in specificity, a phenomenon partially explained by the apparent cooperativity of the substrate binding subsites (where small differences at particular subsites may result in significant differences in the overall preference). For example, prior analysis of KLK14 specificity indicated strong interaction between S3 and S4 binding pockets, allowing identification of a preferred substrate, which was hydrolyzed significantly better than substrates previously identified using PC-SCL and phage display [25]. ...
Article
Full-text available
Kallikrein 13 (KLK13) was first identified as an enzyme that is downregulated in a subset of breast tumors. This serine protease has since been implicated in a number of pathological processes including ovarian, lung and gastric cancers. Here we report the design, synthesis and deconvolution of libraries of internally quenched fluorogenic peptide substrates to determine the specificity of substrate binding subsites of KLK13 in prime and non-prime regions (according to the Schechter and Berger convention). The substrate with the consensus sequential motive ABZ-Val-Arg-Phe-Arg-ANB-NH2 demonstrated selectivity towards KLK13 and was successfully converted into an activity-based probe by the incorporation of a chloromethylketone warhead and biotin bait. The compounds described may serve as suitable tools to detect KLK13 activity in diverse biological samples, as exemplified by overexpression experiments and targeted labeling of KLK13 in cell lysates and saliva. In addition, we describe the development of selective activity-based probes targeting KLK13, to our knowledge the first tool to analyze the presence of the active enzyme in biological samples.
... The 99-loop of KLK8 exhibits an intermediate length between those of KLKs 4, 5, 6 and 7 and the long, exposed "kallikrein-loops" of KLKs 1 to 3 with an insertion of 11 residues 23 . Guided by the crystal structures of murine neuropsin, KLK5 and other KLKs with confirmed Zn 2+ inhibition we generated mutants in order to investigate the functional role of critical residues located in the 75-and 99-loops 23,24 . The substitution of His99Ala resulted in an active KLK8 variant that exhibited a 13-fold higher IC 50 (46.7 ± 3.2 µM) for Zn 2+ inhibition with respect to the wild type (Fig. 4D). ...
Article
Full-text available
Human KLK8/neuropsin, a kallikrein-related serine peptidase, is mostly expressed in skin and the hippocampus regions of the brain, where it regulates memory formation by synaptic remodeling. Substrate profiles of recombinant KLK8 were analyzed with positional scanning using fluorogenic tetrapeptides and the proteomic PICS approach, which revealed the prime side specificity. Enzyme kinetics with optimized substrates showed stimulation by Ca2+ and inhibition by Zn2+, which are physiological regulators. Crystal structures of KLK8 with a ligand-free active site and with the inhibitor leupeptin explain the subsite specificity and display Ca2+ bound to the 75-loop. The variants D70K and H99A confirmed the antagonistic role of the cation binding sites. Molecular docking and dynamics calculations provided insights in substrate binding and the dual regulation of activity by Ca2+ and Zn2+, which are important in neuron and skin physiology. Both cations participate in the allosteric surface loop network present in related serine proteases. A comparison of the positional scanning data with substrates from brain suggests an adaptive recognition by KLK8, based on the tertiary structures of its targets. These combined findings provide a comprehensive picture of the molecular mechanisms underlying the enzyme activity of KLK8.
... The 99-loop of KLK8 exhibits an intermediate length between those of KLKs 4, 5, 6 and 7 and the long, exposed "kallikrein-loops" of KLKs 1 to 3 with an insertion of 11 residues 23 . Guided by the crystal structures of murine neuropsin, KLK5 and other KLKs with confirmed Zn 2+ inhibition we generated mutants in order to investigate the functional role of critical residues located in the 75-and 99-loops 23,24 . The substitution of His99Ala resulted in an active KLK8 variant that exhibited a 13-fold higher IC 50 (46.7 ± 3.2 µM) for Zn 2+ inhibition with respect to the wild type (Fig. 4D). ...
Article
Full-text available
In humans, three different trypsin-isoenzymes have been described. Of these, trypsin-3 appears to be functionally different from the others. In order to systematically study the specificity of the trypsin-isoenzymes, we utilized proteome-derived peptide libraries and quantitative proteomics. We found similar specificity profiles dominated by the wellcharacterized preference for cleavage after lysine and arginine. Especially trypsin-1 slightly favored lysine over arginine in this position, while trypsin-3 did not discriminate between them. In P1´ position, which is the residue C-terminal to the cleavage site, we noticed a subtle enrichment of alanine and glycine for all three trypsins and for trypsin-3 there were additional minor P1´ and P2´ preferences for threonine and aspartic acid, respectively. These findings were confirmed by FRET peptide substrates showing different susceptibility to cleavage by different trypsins. The preference of trypsin-3 for aspartic acid in P2´ is explained by salt bridge formation with the unique Arg193. This salt bridge enables and stabilizes a canonical oxyanion conformation by the amides of Ser195 and Arg193, thus manifesting a selective substrate-assisted catalysis. As trypsin-3 has been proposed to be a therapeutic target and marker for cancers, our results may aid the development of specific inhibitors for cancer therapy and diagnostic probes.
... From a structural point of view, the 3D crystal structures of six human KLKs have been resolved to date (namely, KLK1 and KLK3-KLK7) 16,[42][43][44][45][46][47][48][49] . As shown in FIG. ...
Article
Tissue kallikreins are a family of fifteen secreted serine proteases encoded by the largest protease gene cluster in the human genome. In the past decade, substantial progress has been made in characterizing the natural substrates, endogenous inhibitors and in vivo functions of kallikreins, and studies have delineated important pathophysiological roles for these proteases in a variety of tissues. Thus, kallikreins are now considered attractive targets for the development of novel therapeutics for airway, cardiovascular, tooth, brain, skin and neoplastic diseases. In this Review, we discuss recent advances in our understanding of the physiological functions and pathological implications of kallikrein proteases, and highlight progress in the identification of kallikrein inhibitors, which together are bringing us closer to therapeutically targeting kallikreins in selected disease settings.
Article
Neural networks are modified and reorganized throughout life, even in the matured brain. Synapses in the networks form, change, or disappear dynamically in the plasticity state. The pre- and postsynaptic signaling, transmission, and structural dynamics have been studied considerably well. However, not many studies have shed light on the events in the synaptic cleft and intercellular space. Neural activity-dependent protein shedding is a phenomenon in which (1) presynaptic excitation evokes secretion or activation of sheddases, (2) sheddases are involved not only in cleavage of membrane- or matrix-bound proteins but also in mechanical modulation of cell-to-cell connectivity, and (3) freed activity domains of protein factors play a role in receptor-mediated or non-mediated biological actions. Kallikrein 8/neuropsin (KLK8) is a kallikrein family serine protease rich in the mammalian limbic brain. Accumulated evidence has suggested that KLK8 is an important modulator of neural plasticity and consequently, cognition. Insufficiency, as well as excess of KLK8 may have detrimental effects on limbic functions.
Article
Full-text available
Kallikrein‐related peptidase 6 (KLK6) is a secreted serine protease that belongs to the family of tissue kallikreins. Aberrant expression of KLK6 has been found in different cancers and neurodegenerative diseases, and KLK6 is currently studied as a potential target in these pathologies. We report a novel series of KLK6 inhibitors discovered in a high‐throughput screen within the European Lead Factory program. Structure‐guided design based on docking studies enabled rapid progression of a hit cluster to inhibitors with improved potency, selectivity and pharmacokinetic properties. In particular, inhibitors 32 ((5R)‐3‐(4‐carbamimidoylphenyl)‐N‐((S)‐1‐(naphthalen‐1‐yl)propyl)‐2‐oxooxazolidine‐5‐carboxamide) and 34 ((5R)‐3‐(6‐carbamimidoylpyridin‐3‐yl)‐N‐((1S)‐1‐(naphthalen‐1‐yl)propyl)‐2‐oxooxazolidine‐5‐carboxamide) have single‐digit nanomolar potency against KLK6, with over 25‐fold and 100‐fold selectivities against the closely related enzyme trypsin, respectively. The most potent compound, 32, effectively reduces KLK6‐dependent invasion of HCT116 cells. The high potency in combination with good solubility and low clearance of 32 make it a good chemical probe for KLK6 target validation in vitro and potentially in vivo.
Article
Full-text available
Trypsin and chymotrypsin-like serine proteases from family S1 (clan PA) constitute the largest protease group in humans and more generally in vertebrates. The prototypes chymotrypsin, trypsin and elastase represent simple digestive proteases in the gut, where they cleave nearly any protein. Multidomain trypsin-like proteases are key players in the tightly controlled blood coagulation and complement systems, as well as related proteases that are secreted from diverse immune cells. Some serine proteases are expressed in nearly all tissues and fluids of the human body, such as the human kallikreins and kallikrein-related peptidases with specialization for often unique substrates and accurate timing of activity. HtrA and membrane-anchored serine proteases fulfill important physiological tasks with emerging roles in cancer. The high diversity of all family members, which share the tandem β-barrel architecture of the chymotrypsin-fold in the catalytic domain, is conferred by the large differences of eight surface loops, surrounding the active site. The length of these loops alters with insertions and deletions, resulting in remarkably different three-dimensional arrangements. In addition, metal binding sites for Na+, Ca2+ and Zn2+ serve as regulatory elements, as do N-glycosylation sites. Depending on the individual tasks of the protease, the surface loops determine substrate specificity, control the turnover and allow regulation of activation, activity and degradation by other proteins, which are often serine proteases themselves. Most intriguingly, in some serine proteases, the surface loops interact as allosteric network, partially tuned by protein co-factors. Knowledge of these subtle and complicated molecular motions may allow nowadays for new and specific pharmaceutical or medical approaches.
Article
Full-text available
Biological significance: Forebrain cortex of rats exposed to morphine for 10days is severely altered as far as the overall protein composition is involved. Depending on the method used for protein detection and quantification, 28 (MALDI-TOF MS/MS) or 113 (MaxLFQ) altered proteins were identified. Importantly, in rats sacrificed 20days after the last dose of morphine, the number of altered proteins was decreased to 14 (MALDI-TOF MS/MS) and 19 (MaxLFQ), respectively. Our data indicate the high ability of living organism to oppose the drastic, morphine-induced change of the target tissue protein composition with the aim to return to the physiological norm after complete removal of the drug.
Article
Full-text available
Publisher Summary X-ray data can be collected with zero-, one-, and two-dimensional detectors, zero-dimensional (single counter) being the simplest and two-dimensional the most efficient in terms of measuring diffracted X-rays in all directions. To analyze the single-crystal diffraction data collected with these detectors, several computer programs have been developed. Two-dimensional detectors and related software are now predominantly used to measure and integrate diffraction from single crystals of biological macromolecules. Macromolecular crystallography is an iterative process. To monitor the progress, the HKL package provides two tools: (1) statistics, both weighted (χ 2 ) and unweighted (R-merge), where the Bayesian reasoning and multicomponent error model helps obtain proper error estimates and (2) visualization of the process, which helps an operator to confirm that the process of data reduction, including the resulting statistics, is correct and allows the evaluation of the problems for which there are no good statistical criteria. Visualization also provides confidence that the point of diminishing returns in data collection and reduction has been reached. At that point, the effort should be directed to solving the structure. The methods presented in the chapter have been applied to solve a large variety of problems, from inorganic molecules with 5 A unit cell to rotavirus of 700 A diameters crystallized in 700 × 1000 × 1400 A cell.
Article
Full-text available
The PROCHECK suite of programs provides a detailed check on the stereochemistry of a protein structure. Its outputs comprise a number of plots in PostScript format and a comprehensive residue-by-residue listing. These give an assessment of the overall quality of the structure as compared with well refined structures of the same resolution and also highlight regions that may need further investigation. The PROCHECK programs are useful for assessing the quality not only of protein structures in the process of being solved but also of existing structures and of those being modelled on known structures.
Article
The Kunitz-type trypsin inhibitor from soybean (STI) consists of 181 amino acid residues with two disulfide bridges. Its crystal structures have been determined in complex with porcine pancreatic trypsin in two crystal forms (an orthorhombic form at 1.75 Å resolution and a tetragonal form at 1.9 Å) and in the free state at 2.3 Å resolution. They have been refined to crystallographic R-values of 18.9%, 21.6% and 19.8%, respectively. The three models of STI reported here represent a significant improvement over the partial inhibitor structure in the complex, which was previously determined at a nominal resolution of 2.6 Å by the multiple isomorphous replacement method. This study provides the first high-resolution picture of the complex between a Kunitz-type proteinase inhibitor with its cognate proteinase. Many of the external loops of STI show high B-factors, both in the free and the complexed states, except the reactive site loop whose B-factors are dramatically reduced upon complexation. The reactive site loop of STI adopts a canonical conformation similar to those in other substrate-like inhibitors. The P1 carbonyl group displays no out-of-plane displacement and thus retains a nominal trigonal planar geometry. Modeling studies on the complex between a homologous Kunitz-type trypsin inhibitor DE-3 from Erythrina caffra and the human tissue-type plasminogen activator reveal a new insight into the specific interactions which could play a crucial role in their binding.
Article
The solvent structure in orthorhombic crystals of bovine trypsin has been independently determined by X-ray diffraction to 1.35 Šresolution and by neutron diffraction to 2.1 Šresolution. A consensus model of the water molecule positions was obtained using oxygen positions identified in the electron density map determined by X-ray diffraction, which were verified by comparison to D2OH2O difference neutron scattering density. Six of 184 water molecules in the X-ray structure, all with B-factors greater than 50 Ų, were found to be spurious after comparison with neutron results. Roughly two-thirds of the water of hydration expected from thermodynamic data for proteins was localized by neutron diffraction; approximately one-half of the water of hydration was located by X-ray diffraction. Polar regions of the protein are well hydrated, and significant D2OH2O difference density is seen for a small number of water molecules in a second shell of hydration. Hydrogen bond lengths and angles calculated from unconstrained refinement of water positions are distributed about values typically seen in small molecule structures.
Article
Map interpretation remains a critical step in solving the structure of a macromolecule. Errors introduced at this early stage may persist throughout crystallographic refinement and result in an incorrect structure. The normally quoted crystallographic residual is often a poor description for the quality of the model. Strategies and tools are described that help to alleviate this problem. These simplify the model-building process, quantify the goodness of fit of the model on a per-residue basis and locate possible errors in peptide and side-chain conformations.
Article
The ability of differentiating sensory neurons to remodel a fibronectin substratum was examined. During the early stages of neurite outgrowth, fibronectin was cleared from areas beneath the neuronal soma and processes. The removal of fibronectin occurred in the presence and absence of plasminogen and was associated with the release of fibronectin fragments into the culture medium. The degradation of fibronectin was dependent upon neuronal contact with the substratum. Extraction of cells with the nonionic detergent Triton X-114 identified plasminogen activator and plasmin associated with the cell surface. These findings suggest that the plasminogen activator/plasmin system may play an important role in the interaction of differentiating sensory neurons with the extracellular matrix during axonal outgrowth.
Article
An improved protocol for crystallographic refinement by simulated annealing is presented. It consists of slow cooling starting at high temperatures. Tests of refinements of aspartate aminotransferase and procin pepsin show that the slow-cooling protocol produces lower R factors and better geometry than other protocols previously published. The influence of the temperature-control method, weighting, cooling rate and duration of the heating stage on the success of the slow-cooling protocol is studied. Analysis of the time course of the potential-energy fluctuations indicates no global changes in the state of order of the system. Fluctuations of the potential energy are interpreted as localized conformational changes during the course of the refinement.
Article
Proteolytic enzymes are localized at the growth cone where they are believed to sustain the motility and the penetration of the tip of the growing neurite. A glia derived neurite-promoting factor has been characterized as a potent protease inhibitor. These results seem at first glance paradoxical. However, they suggest that a delicate balance between cell-derived proteases and protease inhibitors modulates neurite elongation and regeneration.
Article
A stoichiometric complex formed between human alpha-thrombin and D-Phe-Pro-Arg chloromethylketone was crystallized in an orthorhombic crystal form. Orientation and position of a starting model derived from homologous modelling were determined by Patterson search methods. The thrombin model was completed in a cyclic modelling-crystallographic refinement procedure to a final R-value of 0.171 for X-ray data to 1.92 A. The structure is in full agreement with published cDNA sequence data. The A-chain, ordered only in its central part, is positioned along the molecular surface opposite to the active site. The B-chain exhibits the characteristic polypeptide fold of trypsin-like proteinases. Several extended insertions form, however, large protuberances; most important for interaction with macromolecular substrates is the characteristic thrombin loop around Tyr60A-Pro60B-Pro60C-Trp60D (chymotrypsinogen numbering) and the enlarged loop around the unique Trp148. The former considerably restricts the active site cleft and seems likely to be responsible for poor binding of most natural proteinase inhibitors to thrombin. The exceptional specificity of D-Phe-Pro-Arg chloromethylketone can be explained by a hydrophobic cage formed by Ile174, Trp215, Leu99, His57, Tyr60A and Trp60D. The narrow active site cleft, with a more polar base and hydrophobic rims, extends towards the arginine-rich surface of loop Lys70-Glu80 that probably represents part of the anionic binding region for hirudin and fibrinogen.