Download Mutations in the Catalytic Domain of Prohormone Convertase 2

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Cellular differentiation wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Cell culture wikipedia , lookup

Tissue engineering wikipedia , lookup

Cell encapsulation wikipedia , lookup

JADE1 wikipedia , lookup

List of types of proteins wikipedia , lookup

Amitosis wikipedia , lookup

Pap test wikipedia , lookup

Transcript
THE JOURNAL OF BIOLOGICAL CHEMISTRY
© 2000 by The American Society for Biochemistry and Molecular Biology, Inc.
Vol. 275, No. 19, Issue of May 12, pp. 14667–14677, 2000
Printed in U.S.A.
Mutations in the Catalytic Domain of Prohormone Convertase 2
Result in Decreased Binding to 7B2 and Loss of Inhibition with
7B2 C-terminal Peptide*
Received for publication, December 3, 1999, and in revised form, February 29, 2000
Ekaterina V. Apletalina, Laurent Muller‡, and Iris Lindberg§
From the Department of Biochemistry and Molecular Biology, Louisiana State University Health Sciences Center,
New Orleans, Louisiana 70112
Prohormone convertases 1 (PC1) and 2 (PC2) are members of a family of subtilisin-like proprotein convertases
responsible for proteolytic maturation of a number of
different prohormones and proneuropeptides. Although
sharing more than 50% homology in their catalytic domains, PC1 and PC2 exhibit differences in substrate
specificity and susceptibility to inhibitors. In addition
to these differences, PC2, unlike PC1 and other members of the family, specifically binds the neuroendocrine
protein 7B2. In order to identify determinants responsible for the specific properties of the PC2 catalytic domain, we compared its primary sequence with that of
other PCs. This allowed us to distinguish a PC2-specific
sequence at positions 242–248. We constructed two PC2
mutants in which residues 242 and 243 and residues
242–248 were replaced with the corresponding residues
of PC1. Studies of in vivo cleavage of proenkephalin, in
vivo production of ␣-MSH from proopiomelanocortin,
and in vitro cleavage of a PC2-specific artificial substrate by mutant PC2s did not reveal profound alterations. On the other hand, both mutant pro-PC2s exhibited a considerably reduced ability to bind to 21-kDa
7B2. In addition, inhibition of mutant PC2-(242–248) by
the potent natural inhibitor 7B2 CT peptide was almost
completely abolished. Taken together, our results show
that residues 242–248 do not play a significant role in
defining the substrate specificity of PC2 but do contribute greatly to binding 7B2 and are critical for inhibition
with the 7B2 CT peptide.
A family of subtilisin-like proprotein convertases (SPCs)1
is responsible for the proteolytic maturation of a number of
* This work was supported by National Institutes of Health Grant
DK 49703 (to I. L.). 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.
‡ Present address: INSERM U36, College de France, Paris 75005,
France.
§ Supported by a National Institute on Drug Abuse Research Scientist Development Award. To whom correspondence should be addressed: Dept. of Biochemistry and Molecular Biology, Louisiana State
University Health Sciences Center, 1901 Perdido St., New Orleans, LA
70112. Tel.: 504-568-4799; Fax: 504-568-6598; E-mail: ilindb@
Lsumc.edu.
1
The abbreviations used are: SPC, subtilisin-like proprotein convertase; PC1 and -2, prohormone convertase 1 and 2, respectively; wtPC2,
wild-type PC2; PE, proenkephalin; POMC, proopiomelanocortin;
EDDnp, ethylenediamine-2,4-dinitrophenyl; Abz, o-aminobenzoyl;
MCA, 4-methylcoumaryl-7-amide; CT, C-terminal; 7B2 CT peptide, human 7B2-(155–185); ACTH, adrenocorticotropic hormone; ␣-MSH,
␣-melanocyte-stimulating hormone; CHO, Chinese hamster ovary;
HEK, human embryonic kidney; PAGE, polyacrylamide gel electrophoresis; HPGPC, high pressure gel permeation chromatography; BisThis paper is available on line at http://www.jbc.org
different prohormones and neuropeptide precursors (reviewed in Refs. 1–3). To date, seven members of this family
have been identified in mammals. All share a common basic
domain structure consisting of an N-terminal signal peptide,
a poorly conserved propeptide, a highly conserved catalytic
domain with the typical subtilisin catalytic triad, a well
conserved middle or P, domain, and a highly variable Cterminal region (2, 3). No crystal structure is available thus
far for the SPCs. Two members of the family, prohormone
convertase 1 (PC1) and 2 (PC2), are expressed exclusively in
neural and endocrine tissues (reviewed in Ref. 1). Synthesized as full-length inactive proproteins (pro-PCs), both PC1
and PC2 undergo autocatalytic propeptide cleavage to yield
mature active forms (4 –10). PC1, but not PC2, is further
slowly cleaved C-terminally to generate a PC1 form with
increased activity (6, 11–13). The PC1 propeptide has been
shown recently to represent a very potent inhibitor of this
enzyme (14), a finding characteristic of subtilases (1, 15). The
PC2 propeptide is correspondingly expected to inhibit PC2,
and recent data indicate that this is indeed the case.2,3
PC1 and PC2 have distinct, although overlapping, substrate
specificities. The two enzymes most often cleave precursors at
pairs of basic residues (2, 3), but both are also capable of
cleaving at a single Arg residue (16 –18). The presence of a
basic residue at position P4 is favored by both enzymes for
cleavage of an artificial substrate (19 –21). However, the two
enzymes show important differences in specificity as regards
natural substrates; processing of endogenous substrates such
as proenkephalin (PE), proopiomelanocortin (POMC), proglucagon, procholecystokinin, proneurotensin, and promelaninconcentrating hormone by PC1 or by PC2 results in differing
product patterns (12, 22–30). In the majority of the studies
performed to date, PC1 exhibits a more restricted substrate
specificity than PC2. Although models of the active sites of PC1
and PC2 have been proposed based on the known x-ray structure of subtilisin (31), the amino acid residues responsible for
the different substrate specificities of these enzymes have not
been identified thus far.
Unlike other members of the family, pro-PC2 requires interaction with the neuroendocrine-specific protein 7B2 for its maturation and activation (32–37). 27-kDa 7B2 is cleaved within
the trans-Golgi network into an amino-terminal 21-kDa domain (21-kDa 7B2; residues 1–135) and a C-terminal peptide of
31 residues (CT peptide) (38, 39). The CT peptide represents a
very potent in vitro inhibitor of PC2 as does the parent 27-kDa
Tris, 2-[bis(2-hydroxyethyl)amino]-2-(hydroxymethyl)propane-1,3-diol;
enk, enkephalin.
2
L. Muller, A. Cameron, Y. Fortenberry, E. Apletalina, and I. Lindberg, manuscript in preparation.
3
C. Lazure, unpublished results.
14667
14668
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
thus contribute to PC2-specific properties. In the present
study, we constructed two murine PC2-related mutants: PC2(242–243), in which residues QP were replaced with the corresponding G residue found in PC1, and PC2-(242–248), in which
the murine PC1 sequence GIVTDA was substituted for the
corresponding sequence QPFMTDI of PC2. We report below
that these mutations result in significant loss of binding of 7B2
to pro-PC2 as well as substantial loss of inhibition of PC2 with
the 7B2 CT peptide.
EXPERIMENTAL PROCEDURES
FIG. 1. Comparison of the amino acid sequences of PC2, PC1,
and furin enzymes. The region consisting of residues 240 –254 (numbering is that of mouse PC2) is shown. The PC1 sequence shown is
shared by human, pig, mouse, and rat PC1 enzymes. The furin sequence
shown is shared by human, mouse, rat, and hamster furins. The sequence shown for mouse PC2 is also shared by rat, human, and pig
PC2s. The sequence shown for D. melanogaster PC2 is also shared by C.
elegans PC2. The residue substitutions in PC2-(242–243) and PC2(242–248) are underlined and in boldface type.
7B2 molecule (40 – 42). Recent attempts to demonstrate the
ability of overexpressed CT peptide to inhibit PC2 in vivo were
not, however, successful due to intracellular internal cleavage
of this peptide (43). Inhibition of PC2 by the 7B2 CT peptide
strongly depends on the presence of a K-K site (40, 42); other
residues within the peptide are also critical to inhibitory potency (44). No information is available thus far on which PC2
residues confer the ability to be inhibited by the CT peptide;
presumably, these residues are not shared by PC1, which is not
inhibited by this peptide (45).
In contrast to the CT peptide, the 21-kDa domain of 7B2
appears to be the portion of 7B2 responsible for pro-PC2 maturation and activation (32). Within this domain, a fragment
encompassing residues 86 –121 contains all of the information
required for 7B2-mediated activation of pro-PC2 (46). Studies
aimed at defining the structural elements of PC2 required for
interaction with 7B2 have identified two residues in the catalytic domain of PC2, Tyr194 and Asp309, that are apparently
critical to the PC2–7B2 binding (47– 49). However, it is likely
that other residues also play an important role in this
interaction.
The present study is aimed at defining the residues in the
active site of PC2 that contribute to the unique characteristics
of this enzyme as opposed to PC1, such as its comparatively
broad substrate specificity, its ability to bind to 7B2, and its
potent inhibition by CT peptide. For this purpose, the primary
sequences of several PC1, PC2, and furin enzymes from different species were aligned, and the residues within the predicted
substrate binding sites (15, 31, 50) were compared. This analysis revealed that in positions 242–248 (numbering is that of
PC2), PC2s contain the sequence QP(Y/F)MTD(L/I), which differs considerably from the corresponding sequence in PC1s and
furins, G(I/E)VTDA (Fig. 1). It is noteworthy that both PC1s
and furins contain a deletion in this sequence as compared with
PC2s; the single G residue in PC1s and furins thus corresponds
to the QP residues of PC2s (positions 242 and 243). The residues in this region (positions 242–248) are located within or in
the vicinity of subsites S2–S5 of prohormone convertases, according to homology modeling of protein convertases based on
the crystal structure of subtilisin (50, 31). These residues could
Materials—The human 7B2 CT peptide (encompassing 7B2 residues
155–185) and a murine PC2 propeptide fragment (residues 57– 84,
SLHHKRQLERDPRIKMALQQEGFDRKKR) were synthesized by
LSUHSC Core Laboratories (New Orleans, LA). pERTKR-MCA was
purchased from Peptides International (Louisville, KY). Ac-LLRVKRNH2 (51) was kindly provided by J. Appel and R. A. Houghten (Torrey
Pines Institute for Molecular Studies, San Diego, CA). The internally
quenched fluorogenic substrate Abz-VPEMEKRYGGFMQ-EDDnp (20)
was kindly provided by L. Juliano (Department of Biophysics, Escola
Paulista de Medicina, San Paolo, Brazil).
Construction of PC2 Mutants—Mouse PC2 cDNA (obtained from Dr.
N. G. Seidah, Clinical Research Institute of Montreal, Montreal, Canada) was first subcloned through blunt-end ligation into the pcDNA3
vector cut at the KpnI site, and this plasmid was used in constructing
PC2 mutants. Mutants were generated by synthesizing mutation-containing fragments by polymerase chain reaction and subcloning the
polymerase chain reaction-generated fragments into mPC2/pcDNA3
between the HindIII and Bsu36I sites. The mutants were constructed
using the same amino-terminal and carboxyl-terminal primers, 5⬘-CGCGCAAGCTTTTTCACTCCCAAAGAAGG-3⬘ and 5⬘-GCCGGCCTCAGGATTCCTCTTCCTCCCATT-3⬘, respectively. The middle primers were
as follows: for mutant PC2-(242–248), 5⬘-GACGGGATTGTGACAGACGCCATCGAAGCCTCCTCCATC-3⬘ and 5⬘-GATGGCGTC TGTCACAATCCCGTCCAGCATCCGGATCCC-3⬘; for mutant PC2-(242–243), 5⬘-GACGGGTTTATGACAGATATCATCGAAGCCTCCTCCATC-3⬘ and 5⬘GATGATATCTG TCATAAACCCGTCCAGCATCCGGATCCC-3⬘. Expand High Fidelity Taq polymerase (Roche Molecular Biochemicals)
was used in all polymerase chain reactions. The fidelity of polymerase
chain reaction-generated fragments was verified by DNA sequencing.
The PC2, PC2-(242–243), and PC2-(242–248) constructs in pCEP4 were
obtained by recloning the corresponding PC2 cDNA from a pcDNA3
plasmid into pCEP4 (Invitrogen, Carlsbad, CA) at the HindIII and NotI
sites using standard procedures.
Cell Culture and Transfection—Methotrexate-amplified CHO/21kDa 7B2 cells that express large quantities of 21-kDa 7B2 were generated by the same procedure described earlier (52). The cells were
maintained in ␣-minimum essential medium without nucleosides, containing 10% dialyzed fetal bovine serum (Irvine Scientific, Santa Ana,
CA) and 50 ␮M methotrexate (Sigma). AtT-20/27-kDa 7B2, AtT-20/PE,
and AtT-20/PC2 cells, representing AtT-20 cell lines stably expressing
rat 27-kDa 7B2 (32), rat proenkephalin (22), and murine wild-type PC2
(12), respectively, were cultured in Dulbecco’s modified Eagle’s medium
(high glucose) (Life Technologies, Inc.) containing 10% NuSerum (Becton Dickinson, Mountain View, CA), 2.5% fetal bovine serum, and
either 200 ␮g/ml hygromycin (Sigma) (for AtT-20/7B2 cells) or 150
␮g/ml G418 (Life Technologies). CHO/21-kDa 7B2, AtT-20/27-kDa 7B2,
and AtT-20/PE cell lines served as hosts for stable transfection of the
PC2 constructs described above. All transfections were performed using
30 ␮g of plasmid and 30 ␮l of Lipofectin (Life Technologies) in a 10-cm
dish. Transfected cells were then cultured in the presence of the appropriate selection agent (250 or 450 ␮g/ml G418 (for CHO/21-kDa 7B2 and
AtT-20/27-kDa 7B2 cells, respectively) or 200 ␮g/ml hygromycin). G418or hygromycin-resistant colonies were picked and screened for PC2
expression by Western blotting using PC2 antiserum (53) or by enzyme
assay with pERTKR-MCA as the substrate.
Production and Partial Purification of PC2 Proteins—Partially purified wild-type PC2, PC2-(242–248), and PC2-(242–243) were obtained
from the conditioned media of the CHO/21-kDa 7B2 cells stably expressing the corresponding 75-kDa pro-PC2 protein. Sequential collections of the conditioned medium from the confluent roller bottles were
performed every morning, 5–7 days in a row, after placing the cells into
100 ml of Opti-MEM (Life Technologies Inc.) containing 100 ␮g/ml
aprotinin (Bayer) overnight (usually 18 –19 h); during the day, the cells
were kept in complete medium (␣-minimum essential medium without
nucleosides, containing 10% dialyzed fetal bovine serum, 50 ␮M meth-
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
otrexate, and 250 ␮g/ml G418). The conditioned media were centrifuged
immediately after collection at low speed to remove floating cells, and
the supernatants were stored at ⫺70 °C until needed. For purification,
100 –200 ml of the medium were used. The medium was thawed and
centrifuged at 15,000 ⫻ g for 15 min, and the supernatant was diluted
3-fold with buffer A (20 mM Bis-Tris buffer, pH 6.65, containing 0.1%
Brij 35). Diluted medium was applied onto a 5-ml Econo-PackTM Q
Cartridge (Bio-Rad) equilibrated with buffer A. The column was washed
with buffer A and eluted at a flow rate of 0.5 ml/min with buffer B (1 M
sodium acetate in buffer A) using two linear gradients, from 0 to 0.35 M
in 175 min and from 0.35 to 1 M in 100 min.
Enzyme Assays—Duplicate reactions were performed in a polypropylene microtiter plate in a total volume of 50 ␮l, containing 100 mM
sodium acetate buffer, pH 5.0, 5 mM CaCl2, 0.4% n-octyl glucoside, an
inhibitor mixture (0.14 mM N␣-p-tosyl-L-lysine chloromethyl ketone
(Sigma), 0.3 mM N-p-tosyl-L-phenylalanine ketone (Sigma), 10 ␮M
trans-epoxysuccinyl-L-leucylamido-(4-guanidino)butane (Sigma), and 1
␮M pepstatin (Sigma)), 0.2 mM fluorogenic substrate pERTKR-MCA,
and 5 ␮l of the partially purified enzyme. Reactions to test the peptide
inhibitors (human 7B2 CT peptide, the PC2 propeptide fragment, or the
hexapeptide Ac-LLRVKR-NH2) also contained varying amounts of
these peptides. Prior to the addition of the peptide inhibitor (if used)
and the substrate, the partially purified enzyme was preincubated in
the assay buffer at 37 °C for 30 – 40 min. After the peptide inhibitor was
added, the reaction mixtures were preincubated at room temperature
for 30 min, followed by the addition of substrate. All reactions were
conducted at 37 °C for 30 – 45 min. Released 7-amino-4-methylcoumarin
was measured with a Fluoroscan Ascent fluorometer (Labsystems) using an excitation and an emission wavelength of 380 and 460 nm,
respectively. The IC50 values for the peptide inhibitors were calculated
through nonlinear regression of the experimental data (fluorescence
versus peptide concentration) using the program GraphPad (GraphPad
Software, Inc., San Diego, CA).
pH Dependence of Enzymatic Activity of PC2s—The effect of varying
pH on the activity of partially purified PC2 enzymes was determined by
measuring the rate of hydrolysis of 0.2 mM pERTKR-MCA in 100 mM
Bis-Tris, 100 mM sodium acetate buffer set at different pH values
(ranging from 4.0 to 7.4) in the presence of 5 mM CaCl2, 0.4% octyl
glucoside, and an inhibitor mixture (see above). Prior to the addition of
the substrate to start the reaction, the assay mixtures were preincubated at 37 °C for 30 – 40 min for enzyme activation.
Time Course Conversion of Pro-PC2 Proteins—Partially purified
wild-type pro-PC2, pro-PC2-(242–248), and pro-PC2-(242–243) were
incubated at 30 °C in 100 mM sodium acetate buffer, pH 5.0, in the
presence of 5 mM CaCl2, 0.1% Brij 35, and an inhibitor mixture (see
above). At varying time intervals, aliquots were frozen for subsequent
analysis by SDS-PAGE and Western blotting using PC2 antiserum
(53).
Quantitation of PC2 Proteins in the Partially Purified Preparations—To determine the relative concentrations of PC2 protein in the
various preparations, wild-type PC2, PC2-(242–248), and PC2-(242–
243) were diluted into 100 mM sodium acetate buffer, pH 5.0, containing
5 mM CaCl2 and 0.1% Brij 35, and two amounts (the second twice as
much as the first) of each sample were applied onto a 10% gel. On the
same gel were applied varying amounts of highly purified 66-kDa PC2
(diluted in the same buffer) with a known protein content. The samples
were subjected to SDS-PAGE and Western blotting using PC2 antiserum (53). Images of the blots were then obtained with a ScanJet 4C
scanner (Hewlett Packard), and the intensities of the bands were compared using ImageQuant Software (Molecular Dynamics, Inc., Sunnyvale, CA). The PC2 concentration in each enzyme preparation was
averaged from 2– 4 independent experiments.
Determination of kcat/Km Values for an Internally Quenched Fluorogenic Substrate—The internally quenched substrate Abz-VPEMEKRYGGFMQ-EDDnp (final concentration of 2 ␮M) was subjected to
digestion by partially purified PC2 or PC2-(242–248) in a buffer containing 200 mM sodium acetate, pH 5.0, 5 mM CaCl2, 0.4% octyl glucoside, and inhibitor mixture (see above) in a total volume of 50 ␮l. The
reactions were initiated by the addition of substrate and conducted for
2 h at 37 °C in polypropylene microtiter plates. Abz fluorescence was
recorded with a Fluoroscan Ascent fluorometer (Labsystems) using
excitation and emission wavelengths of 320 and 420 nm, respectively.
kcat/Km values were calculated based on the curves obtained, as described previously (20).
Metabolic Labeling—Subconfluent wells of AtT-20/27-kDa 7B2 cells,
expressing either wild-type PC2, PC2-(242–243), PC2-(242–248), or no
exogenous PC2, and wells of AtT-20/PC2 cells, plated in six-well plates,
were washed once with 3 ml of PBS and once with 1 ml of methionine
14669
and cysteine-free medium (ICN Biomedicals, Inc., Irvine, CA) containing 10 mM HEPES, pH 7.4, and were starved for 30 min in 1 ml of this
medium. Cells were then incubated in 1 ml of methionine and cysteinefree medium containing 0.5 mCi/ml [35S]methionine/cysteine (Promix,
Amersham Pharmacia Biotech) for different time periods depending on
the particular experiment. In the experiments to study production of
␣-MSH, cells were labeled for either 20 min or for 6 h. Following the 6-h
labeling period, cells were homogenized in 1 ml of acid mix (1 M acetic
acid, 20 mM HCl, and 0.1% ␤-mercaptoethanol). Following the experiment in which cells were pulse-labeled for 20 min, cells were further
incubated in 1 ml of chase medium (Dulbecco’s modified Eagle’s medium containing 10 mM HEPES, pH 7.4, and 2% fetal bovine serum) for
1 h prior to homogenization in acid mix. Cell extracts in acid mix were
frozen, thawed, and centrifuged at 14,000 ⫻ g for 5 min. The supernatants were collected and lyophilized overnight. The dried cell extracts
were resuspended in 500 ␮l of AG buffer (100 mM sodium phosphate
buffer, pH 7.4, 1 mM EDTA, 0.1% Triton X-100, 0.5% Nonidet P-40, and
0.9% NaCl), centrifuged briefly to remove insoluble material, and stored
frozen at ⫺20 °C prior to immunoprecipitation using polyclonal rabbit
antiserum JH93 directed against the N-terminal portion of ACTH (54).
To study binding of 7B2 to PC2s, cells were pulsed for 20 min and
chased for 30 min, and cell extracts were prepared as described previously (46). 7B2 and PC2 were immunoprecipitated from clarified cell
extracts with antiserum LSU13 or LSU18 (directed against 7B2 and
PC2, respectively; Ref. 32). In order to characterize pro-PC2 autocatalytic conversion in vitro, cells were pulsed for 20 min, chased for 120
min in the presence of 1 ␮M bafilomycin A1, an inhibitor of vacuolar
ATPase, and extracted as described previously (34). Pro-PC2 was immunoprecipitated from clarified cell extracts prior to a 30-min incubation at 37 °C in 100 mM Bis-Tris, 100 mM sodium acetate buffer set at
either pH 5 or pH 7. In the experiments to study intracellular maturation of pro-PC2s, cells were pulsed for 20 min, chased for 0 or 120 min,
and then boiled for 5 min in 0.1 ml of boiling buffer (50 mM sodium
phosphate, pH 7.4, 1% SDS, 50 mM ␤-mercaptoethanol, and 2 mM
EDTA). Samples were then diluted with 0.9 ml of AG buffer for immunoprecipitation using antiserum LSU 18.
Immunoprecipitation—For immunoprecipitation using antiserum
JH93, a 200- or 100-␮l aliquot of cell extract was incubated in AG buffer
with 2.5 mg of prehydrated Protein A-Sepharose CL-4B beads (Amersham Pharmacia Biotech) in a final volume of 350 ␮l for 1 h at 4 °C with
constant rocking. The samples were centrifuged for 2 min, and the
supernatants were transferred to fresh tubes and incubated with antiserum JH93 (5– 6 ␮l) for 3 h at 4 °C with constant rocking. Prehydrated
Protein A-Sepharose CL-4B beads in AG buffer were then added into
the supernatant, and the samples were further incubated for 1 h at
4 °C. Samples were centrifuged for 2 min, and the beads were washed
three times with 0.5 ml of AG buffer, once with 0.5 ml of 0.5 M NaCl, and
once with 0.5 ml of PBS. Immunoprecipitated proteins were extracted
from the beads by adding 120 ␮l of 21% (v/v) acetonitrile, containing
0.07% trifluoroacetic acid, 6 M urea, and 0.33 M acetic acid, and incubating at room temperature for 15 min, followed by centrifugation for 2
min at 1,500 ⫻ g. The supernatants were removed and stored frozen at
⫺20 °C prior to analysis by high pressure gel permeation chromatography (HPGPC). Immunoprecipitation of pro-PC2s and 7B2 was performed using antiserum LSU18 or LSU13 as described previously (34).
Immune complexes were boiled in Laemmli sample buffer, and proteins
were separated by SDS-PAGE on 8.8 or 15% gels. Gels were exposed to
a phosphoscreen and analyzed with a Storm PhosphorImager and ImageQuant software (Molecular Dynamics).
Proenkephalin Processing in Vivo—AtT-20/PE cells and AtT20/PE
cells stably transfected with PC2, PC2-(242–243), or PC2-(242–248)
were seeded in 10-cm dishes and grown until ⬃70 – 80% confluent and
then washed once with PBS and homogenized in 1 ml of acid mix on ice.
The homogenates were frozen, thawed, and centrifuged for 5 min at
14,000 ⫻ g. The clear supernatants were removed and lyophilized
overnight. The dried cell extracts were resuspended in 250 ␮l of 32%
acetonitrile containing 0.1% trifluoroacetic acid and centrifuged at
14,000 ⫻ g for 5 min to remove any insoluble material. The clear
supernatants were removed and stored at ⫺20 °C. 100-␮l aliquots were
size-fractionated by HPGPC, and the fractions obtained were subjected
to radioimmunoassay using the Met-enk-Arg-Gly-Leu antiserum Cass
(55). Three different clones of each cell line were used in these
experiments.
HPGPC—Samples were size-fractionated by HPGPC as described
previously (23). The system consisted of a TSK-GEL SW guard column
(7.5 ⫻ 7.5 mm; TosoHaas, Montgomeryville, PA) and two HPGPC
columns connected in series, a Protein Pak SW 300 column (300 ⫻ 7.8
mm; Waters, Milford, MA) and a Bio-Sil TSK-125 column (300 ⫻ 7.5
14670
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
mm; Bio-Rad). When radiolabeled proteins were separated by HPGPC,
the radioactivity was detected on-line using a ␤-RAMR Flow-Through
System, model 2 (IN/US Systems, Inc., Tampa, FL). Otherwise, 1-ml
fractions were collected into polypropylene tubes each containing 50 ␮g
of bovine serum albumin to prevent absorption to tube walls and stored
at ⫺20 °C. Aliquots of the fractions were subjected to radioimmunoassay using the Met-enk-Arg-Gly-Leu antiserum Cass.
Radioimmunoassay of Proenkephalin Digestion Products—Radioimmunoassay was carried out on HPGPC fractions obtained from cellular
extracts; duplicate aliquots of the fractions either were directly diluted
in 100 ␮l of radioimmunoassay buffer (100 mM sodium phosphate, pH
7.4, containing 0.1% heat-treated bovine serum albumin, 50 mM NaCl,
and 0.1% sodium azide) or in some assays were vacuum-dried in the
presence of bovine serum albumin and then resuspended in 100 ␮l of
the same buffer (when the volume of an aliquot exceeded 6 ␮l). The
samples were then subjected to Met-enk-Arg-Gly-Leu radioimmunoassay as described previously (55). All radioimmunoassays were performed twice with similar results.
Collection of Conditioned Medium from AtT-20 Cells—AtT-20/27kDa 7B2 cells expressing either no PC2, wild-type PC2, PC2-(242–243),
or PC2-(242–248) were seeded into a six-well plate (500,000 cells/well)
and grown for 1 or 2 days until 60 – 80% confluent. The wells were
washed once in 3 ml of PBS and then incubated with 1 ml of Opti-MEM
containing 100 ␮g/ml aprotinin for 16 –17 h. The conditioned medium
was removed from each well, centrifuged briefly to remove any floating
cells, and stored at ⫺20 °C. The cells were washed twice with 2 ml of
PBS and extracted with 350 – 400 ␮l of acid mix (see above) on ice, and
acid suspensions were subjected to protein assay to correct for variations in cell growth. Ten ␮l of duplicate aliquots of the conditioned
medium were used for the enzyme assay, and 100- or 200-␮l aliquots
were subjected to trichloroacetic acid precipitation and subsequent
SDS-PAGE and Western blotting to compare the amounts of PC2s. The
mixtures for the enzyme assay contained (in a 50-␮l volume) 100 mM
sodium acetate, pH 5.0, 5 mM CaCl2, 0.4% octyl glucoside, an inhibitor
mixture (see above), 10 ␮l of conditioned medium, and 200 ␮M fluorogenic substrate, pERTKR-MCA. The mixtures were incubated in a
microtiter plate at 37 °C for up to 16 h, and released 7-amino-4-methylcoumarin was measured with a Fluoroscan Ascent fluorometer using
excitation/emission wavelengths of 380/460 nm.
Transient Transfection of HEK 293 Cells—HEK 293 cells were maintained in Dulbecco’s modified Eagle’s medium (high glucose), containing
10% fetal bovine serum. The cells were plated into six-well plates and
grown until approximately 70 – 80% confluent. The cells were then
transiently transfected using 2 ␮g of pro-PC2 (wild-type or mutant)
expression vector and 5 ␮l of LipofectAMINE (Life Technologies Inc.) or
using 2 ␮g of pro-PC2 (wild-type or mutant) expression vector along
with 2 ␮g of 21-kDa 7B2 cDNA and 10 ␮l of LipofectAMINE in 1 ml of
Opti-MEM per well. In control transfections, empty pcDNA3 vector was
used instead of pro-PC2-encoding cDNAs. Approximately 24 h posttransfection 1 ml/well of Opti-MEM containing 100 ␮g/ml aprotinin was
placed on the cells for 16 –17 h. The conditioned medium was removed,
centrifuged briefly to remove nonadherent cells, and concentrated 10 –
16-fold at 4 °C using a 30K Microcon or Centricon concentrating unit
(Amicon). 10-␮l duplicate aliquots of the concentrated conditioned medium were used for the PC2 enzyme assay immediately following concentration. The assay conditions were identical to those used with
AtT-20 cells (see above); reaction mixtures were incubated at 37 °C for
up to 24 h.
SDS-PAGE and Western Blotting—Samples for SDS-PAGE analysis
were boiled for 2 min after the addition of 1⁄5 volume of 5⫻ Laemmli
sample buffer. After electrophoresis on either 8.8 or 10% gels, the
separated proteins were electroblotted onto nitrocellulose and subjected
to Western blotting using PC2 antiserum directed against the last 13
amino acids of PC2 (53).
Protein Assay—Protein concentrations were determined using the
Bio-Rad Protein Assay (Bio-Rad), with bovine serum albumin as a
standard.
RESULTS
In the present study, wild-type and mutant PC2s were characterized both in vitro and in vivo. For in vitro experiments,
PC2s were partially purified from the conditioned media of the
CHO/21-kDa 7B2 cells stably expressing the corresponding
75-kDa pro-PC2 protein. Using this expression system, we
were able to obtain all three PC2s in relatively large quantities.
In vivo experiments were performed using AtT-20/27-kDa 7B2
FIG. 2. Mutant PC2s display typical PC2-like pH dependence
of enzymatic activity. The rate of hydrolysis of 200 ␮M pERTKRMCA with wild type PC2, PC2-(242–243), or PC2-(242–248) obtained
from the conditioned media of CHO/21-kDa 7B2 cells expressing the
corresponding pro-PC2 was measured in 100 mM Bis-Tris/100 mM sodium acetate buffer set at different pH levels. Activity levels at the
different pH levels are expressed as a percentage of the maximum
activity for each enzyme.
or AtT-20/PE cells stably expressing wild-type and mutant
pro-PC2s. These experiments allowed us to assess the intracellular effects of mutations in the context of a neuroendocrine cell
type.
pH Dependence of Mutant PC2s—PC1 and PC2 have been
shown to possess slightly different pH optima for fluorogenic
substrate cleavage, namely a pH of about 5.0 for PC2 and pH
5.5– 6.0 for PC1 (6, 13, 40, 56). To determine if replacement of
residues in the vicinity of the PC2 catalytic site with PC1
residues shifts the pH optimum toward the higher PC1 optimum, we directly compared the pH dependences of partially
purified wild-type PC2, PC2-(242–248), and PC2-(242–243) toward the cleavage of the fluorogenic substrate pERTKR-MCA.
As is evident in Fig. 2, all three enzymes have similar pH
dependence profiles, indicating that the mutated residues do
not contribute to the lower pH optimum of PC2.
Maturation of Mutant Pro-PC2s—To determine if the mutations introduced affect pro-PC2 maturation, we first compared
the intracellular processing of mutant and wild-type pro-PC2s
in AtT-20/27-kDa 7B2 cells. Fig. 3A shows the results of an
experiment in which cells were pulsed with [35S]methionine/
cysteine and then either immediately extracted or chased for
120 min prior to extraction. It is evident that maturation of
pro-PC2-(242–248) as well as pro-PC2-(242–243) in these cells
proceeds at a slower rate as compared with that of wild-type
pro-PC2. In fact, the rates of maturation of mutant pro-PC2s
are closer to the rate of maturation of wild-type pro-PC2 in
AtT-20 cells expressing no exogenous 7B2.
To obtain insight into the nature of the slower intracellular
maturation reaction of the mutant pro-PC2s, we immunoprecipitated radiolabeled wild-type pro-PC2 and pro-PC2-(242–
248) from the same cells and incubated them at either pH 7.0
or pH 5.0 to assess in vitro maturation capability (Fig. 3B). The
results show that mutant pro-PC2 and wild-type pro-PC2 from
AtT-20 cells co-expressing 7B2 considerably differ in their ability to autocatalytically convert at acidic pH. In contrast to
wild-type PC2, virtually no mature PC2-(242–248) was detected after a 30-min incubation of the pro form at pH 5.0,
indicating a highly reduced ability of the mutant to autoconvert. In this respect, this mutant pro-PC2 resembles wild-type
pro-PC2 contained in AtT-20/PC2 cells expressing no exogenous 7B2. Conversion of wild-type pro-PC2 from AtT-20/PC2
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
14671
FIG. 3. Maturation of mutant pro-PC2s expressed in AtT-20/27-kDa 7B2 cells. A, intracellular processing of wild-type and the mutant
pro-PC2s. AtT-20/wild-type PC2 cells expressing no endogenous 7B2 and AtT-20/27-kDa 7B2 cells expressing either wild-type PC2, PC2-(242–243),
or PC2-(242–248) were pulsed for 20 min and chased for 0 or 120 min; extracts were subjected to immunoprecipitation with PC2 antiserum; and
immunoprecipitates were analyzed by SDS-PAGE. Similar results were obtained in two independent experiments employing three independent
clones for each mutant. B, autocatalytic conversion in vitro of wild-type pro-PC2 and pro-PC2-(242–248). AtT-20/wild-type PC2 cells expressing no
endogenous 7B2 and AtT-20/27-kDa 7B2 cells expressing either wild-type PC2 or PC2-(242–248) were pulsed for 20 min and then chased for 120
min in the presence of bafilomycin A1. Proteins were immunoprecipitated using PC2 antiserum and incubated for 30 min at 37 °C in 100 mM
Bis-Tris, 100 mM sodium acetate buffer, set at either pH 7 or pH 5. The samples were then analyzed by SDS-PAGE. Similar results were obtained
with three independent clones for PC2-(242–248).
cells proceeded at a significantly slower rate than wild-type
pro-PC2 obtained from AtT-20/27-kDa 7B2/PC2 cells (Fig. 3B)
but was still higher than the rate of conversion of
pro-PC2-(242–248).
We further compared the rates of autocatalytic conversion of
wild-type and mutant pro-PC2s expressed in CHO/21-kDa 7B2
cells and partially purified from the conditioned media (Fig. 4).
Rates of maturation of the three pro-PC2s examined were more
similar in this case than when the proteins were expressed in
AtT-20/27-kDa 7B2 cells. With both wild-type PC2 and PC2(242–248), 55–58% of the enzymes were represented by the
mature 66-kDa form after a 5-min incubation. With PC2-(242–
243), a slightly longer incubation (7.5 min) was required to
obtain 52% of the enzyme as the mature form. It should be
noted that some variability in the rate of autocatalytic conversion was observed between the two different preparations of
pro-PC2-(242–248) we used in this experiment. With the second enzyme preparations, 50% conversion (ratio of PC2/(proPC2 ⫹ PC2)) was observed after approximately 10 min of
incubation (data not shown). Nevertheless, even taking into
account this variability, we can conclude that when expressed
in CHO/21-kDa 7B2 cells, the two mutant pro-PC2s do not
exhibit profound alterations in the rate of autocatalytic conversion as compared with wild-type pro-PC2.
Activity of Mutant PC2s toward the Fluorogenic Substrate—We next examined whether the introduced mutations
affect the catalytic activity of PC2 against the fluorogenic substrate pERTKR-MCA. In the conditioned medium of AtT-20/27kDa 7B2 cells expressing different PC2s, the activity of both
mutants proved to be considerably lower (by 5–10 times) than
that of wild-type PC2, although PC2-(242–243) was more active
than PC2-(242–248) (Fig. 5).
FIG. 4. Autocatalytic conversion of mutant pro-PC2s expressed in CHO/21-kDa 7B2 cells. Wild-type pro-PC2, pro-PC2-(242–
243), and pro-PC2-(242–248) partially purified from conditioned media
of CHO/21-kDa 7B2 cells expressing the corresponding pro-PC2 were
incubated at 37 °C in 100 mM sodium acetate, pH 5.0, containing 5 mM
CaCl2, 0.1% Brij, and an inhibitor mixture. At the time points indicated,
aliquots were frozen for analysis by SDS-PAGE and Western blotting.
The pro-PC2 and PC2 bands were quantitated, and the ratio of PC2/
(pro-PC2 ⫹ PC2) was calculated and shown at the bottom of the Western blot.
Comparison of the specific activities of the PC2 enzymes
partially purified from the conditioned medium of CHO/21-kDa
7B2/PC2 (wild-type or mutant) cells was performed using
14672
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
FIG. 5. Activity of mutant PC2s in the conditioned medium of
AtT-20/27-kDa 7B2 cells. AtT-20/27-kDa 7B2 cells expressing either
wild-type PC2, PC2-(242–243), PC2-(242–248), or no PC2 (control) were
plated at 500,000 cells/well in a six-well plate; 2 days later, serum-free
medium was placed on the cells for 16 –17 h, and the conditioned
medium was then assayed in duplicate for PC2 enzymatic activity.
Western blotting indicated roughly comparable expression levels of the
three PC2s in the media (not shown), and protein assay indicated
similar growth levels of all cell lines used (not shown). The experiment
was repeated once with similar results.
quantitative Western blotting as a measure of PC2 protein. The
specific activity of the wild-type PC2 preparation was estimated at 35 nmol/h/mg of PC2. The activities of both mutant
PC2s varied from 40 to 80% of that of wild-type PC2 in several
preparations tested. The reason for this variability is not clear;
one possible explanation is differential loss of the enzyme activity during the purification process. It should be noted, however, that if even this is the case, the purification by no means
produced unstable PC2 forms, since all three PC2 enzymes
retained at least 90% of their initial activity against the fluorogenic substrate after being incubated at 37 °C in the reaction buffer for 2 h.
Taken together, these results show that the final activity of
the mutant PC2s against the fluorogenic substrate depends
heavily on the host cells used for expression. When expressed
in AtT-20/27-kDa 7B2 cells, both mutant PC2s exhibited considerably lower activities as compared with wild-type PC2.
When PC2s are expressed in CHO/21-kDa 7B2 cells, the activities of the mutant and wild-type PC2s were relatively close.
Binding of 7B2 to Mutant Pro-PC2s Is Greatly Reduced—As
was shown above, the rates of maturation of mutant pro-PC2s
in AtT-20/27-kDa 7B2/PC2 cells resembled that of wild-type
pro-PC2 in AtT-20/PC2 cells expressing no exogenous 7B2. This
result could potentially indicate impaired interaction of mutant
PC2s with 7B2, since 7B2 is known to play a significant role in
the intracellular maturation of pro-PC2 (32, 34, 35). Therefore,
we next studied binding of 7B2 to the mutant PC2s. Fig. 6A
shows the results of an experiment in which radiolabeled extracts obtained from AtT-20/27-kDa 7B2 cells co-expressing
either wild-type PC2, PC2-(242–243), or PC2-(242–248) were
immunoprecipitated with antiserum directed against 7B2. It is
evident that neither pro-PC2-(242–243) nor pro-PC2-(242–248)
were effectively co-immunoprecipitated with 7B2. 21-kDa 7B2
was not also co-immunoprecipitated with pro-PC2-(242–248)
(Fig. 6B) or pro-PC2-(242–243) (data not shown) using an antiserum against PC2. We conclude that the ability to bind 7B2
is greatly reduced in both mutants.
Co-expression of 7B2 Is Still Required for the Generation of
Enzymatically Active Mutant PC2s in a Constitutive Cell
Line—The intracellular presence of 7B2 is required for the
activation of wild-type pro-PC2 in CHO cells (32). We next
performed experiments to investigate whether mutant proPC2s continue to require co-expression of 7B2 to become enzymatically competent. For this purpose, CHO cells not express-
FIG. 6. The ability of both mutant pro-PC2s to bind to 7B2 is
highly reduced. AtT-20/27-kDa 7B2 cells expressing either wild-type
PC2, PC2-(242–243), or PC2-(242–248) were pulsed for 20 min and
chased for 30 min. 7B2 and PC2 were then subjected to co-immunoprecipitation using antiserum against 7B2 (A) or PC2 (B). Similar results
were obtained in two experiments and using three independent clones
for each mutant.
ing 7B2 were transiently transfected with the plasmids
encoding either wild-type or mutant pro-PC2s. No enzymatic
activity was detected in the concentrated (10 –16-fold) conditioned medium from the cells, whether the plasmids encoding
wild-type or mutant pro-PC2s were used for transfection (not
shown). However, the amount of PC2 forms in the medium was
very low, as detected by Western blotting (not shown). We
therefore performed transient transfection of another constitutive cell line, HEK 293, with the same vectors encoding wildtype or mutant pro-PC2s. Transfection of this cell line resulted
in higher expression levels of PC2s, as detected by Western
blotting (not shown). However, medium from HEK 293 cells
was enzymatically inactive, whether the plasmids encoding
wild-type or mutant pro-PC2s were used for transfection (Table
I). On the other hand, media obtained from HEK 293 cells that
were transiently transfected with mutant pro-PC2 expression
vectors along with a 21-kDa 7B2 plasmid exhibited readily
detectable levels of PC2 activity (Table I). We thus conclude
that mutant pro-PC2s, similar to wild-type pro-PC2, continue
to require co-expression of 7B2 in a constitutive cell line, such
as HEK 293, for expression of enzymatic activity.
Inhibition of Mutant PC2s with 7B2 CT Peptide Is Diminished—PC2 is the only prohormone convertase inhibited by the
7B2 CT peptide. This peptide represents a nanomolar inhibitor
of PC2 (40). To examine whether the introduced mutations
have any effect on this unique PC2 characteristic, we determined the IC50 values for the inhibition of wild-type and mutant PC2s with the CT peptide. For this experiment, enzymes
were partially purified from the conditioned medium of the
CHO/21-kDa 7B2 cells expressing these proteins. The data in
Table II show that the mutant PC2s exhibit considerably reduced inhibition with the CT peptide. The IC50 value of PC2(242–243) was about 50 times higher than that of wild-type
PC2. PC2-(242–248), in its turn, was only weakly inhibited
with the CT peptide. Even in the presence of as high a CT
peptide concentration as 100 ␮M, PC2-(242–248) retained approximately 50% of its activity. These experiments were performed using two different enzyme preparations of each type of
PC2 with identical results. Similar results were also obtained
with the conditioned media from CHO/21-kDa 7B2/PC2 (wildtype or mutant) cells as well as with conditioned media from
the CHO/21-kDa 7B2 cells that had been transiently transfected with plasmids encoding wild-type pro-PC2 or pro-PC2(242–248). These data clearly demonstrate that residues 242–
248 play a critical role in the inhibitory interaction of PC2 with
the 7B2 CT peptide.
Inhibition of Mutant PC2s with a PC2 Propeptide Fragment
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
TABLE I
Enzymatic activity of mutant PC2s in the concentrated conditioned
medium from the transiently transfected HEK 293 cells
HEK 293 cells were transiently transfected with the pro-PC2 (wildtype or mutant) expression vectors or with the pro-PC2 expression
vectors (or empty vector) together with a 21-kDa 7B2 expression plasmid. The conditioned media collected from cells were concentrated
10 –16-fold, and 10-␮l aliquots were used in the PC2 enzymatic assay.
AMC generated
14673
TABLE III
Specificity constants for wild-type and mutant PC2s for an internally
quenched PC2 substrate
Partially purified PC2 and PC2-(242–248) were obtained from the
conditioned media of CHO/21-kDa 7B2 cells expressing the corresponding pro-PC2 protein. The hydrolysis of an internally quenched substrate
was conducted under pseudo-first order conditions, and the progress
curves of Abz fluorescence were used to calculate kcat/Km values (see
“Experimental Procedures”). Each value represents the mean ⫾ S.D.,
determined from four tests.
pmol/5 h
a
⬍15a
2080
⬍15a
230
⬍15a
340
⬍15
Less than 15 pmol of AMC was detected after a 24-h incubation.
TABLE II
Inhibition constants against wild-type and mutant PC2s
Partially purified PC2s were obtained from the conditioned medium
of CHO/21-kDa 7B2 cells expressing the corresponding pro-PC2 protein. The rate of hydrolysis of pERTKR-MCA by PC2s was determined
in the presence of either the CT peptide, the hexapeptide Ac-LLRVKRNH2, or a PC2 propeptide fragment, as described under “Experimental
Procedures.” The results obtained were then used to calculate the IC50
values using nonlinear regression. Each value is the mean ⫾ S.D.,
determined from 2– 4 experiments.
IC50
Protein
Wild-type PC2
PC2-(242–243)
PC2-(242–248)
kcat/Km ⫻ 103
Peptide sequence
HEK 293 cells
Wild-type PC2
Wild-type PC2 ⫹ 7B2
PC2-(242–243)
PC2-(242–243) ⫹ 7B2
PC2-(242–248)
PC2-(242–248) ⫹ 7B2
Empty vector ⫹ 7B2
CT peptide
Ac-LLRVKR-NH2
0.028 ⫾ 0.008
1.7 ⫾ 0.3
⬎30
2.7 ⫾ 0.5
7.8 ⫾ 1.0
2.5 ⫾ 1.0
␮M
PC2 propeptide
fragment
5.0 ⫾ 0.8
⬎50
⬎100
and the Hexapeptide Ac-LLRVKR-NH2—In order to determine
whether the alteration of inhibitory susceptibility extended to
other inhibitory sequences, we next examined the inhibition of
mutant PC2s with two peptides: a fragment of the murine PC2
propeptide encompassing residues 57– 84 and the synthetic
hexapeptide Ac-LLRVKR-NH2 (obtained during a combinatorial library screen; Ref. 51). The PC2 propeptide represents a
potent inhibitor of PC2.2,3 The hexapeptide Ac-LLRVKR-NH2
represents an inhibitor of both PC2 and PC1 but inhibits PC1
with a much higher potency than PC2 (Ki values of 3.2 and 360
nM, respectively; Ref. 51). As is evident from Table II, inhibition
of both mutant PC2s, but especially of PC2-(242–248), with the
PC2 propeptide fragment is substantially reduced as compared
with that of wild-type PC2. In contrast, the hexapeptide exhibits approximately the same potency against the mutant PC2s
as it does against wild-type PC2. These data (obtained with two
different preparations of each PC2) show that PC2 residues
242–248 are essential for the inhibition of PC2 with the PC2
propeptide fragment and the CT peptide but do not significantly affect its inhibition with the synthetic hexapeptide.
In Vitro Cleavage of the Internally Quenched Fluorogenic
Substrate with PC2-(242–248)—To explore whether the introduced mutations confer PC1-like specificity to PC2, we examined in vitro cleavage of the internally quenched substrate
Abz-VPEMEKRYGGFMQ-EDDnp with partially purified PC2
and PC2-(242–248). This peptide represents a good in vitro
substrate for PC2 but is not well cleaved by PC1 (20). The
kcat/Km values obtained are presented in Table III. The value
for the mutant PC2 is approximately half of that for wild-type
PC2. The activity of PC2-(242–248) against the fluorogenic
substrate pERTKR-MCA in the same enzyme preparations was
decreased to a somewhat lesser degree (20%). Thus, PC2-(242–
PC2
PC2-(242–248)
M
Abz-VPEMEKR 2 YGGFMQ-EDDnp 5.7 ⫾ 0.7
⫺1
s⫺1
2.7 ⫾ 0.4
248) exhibits a decrease in its activity and/or affinity toward
the internally quenched substrate as compared with wild-type
PC2; however, this difference is not profound, since PC1 is
totally unable to cleave this internally quenched substrate. We
conclude that the introduction of the PC1 residues does not
substantially affect the specificity of PC2 toward this
substrate.
Proenkephalin Cleavage with Mutant PC2s in Vivo—In a
further attempt to determine if mutant PC2s acquire PC1-like
specificity, we studied the cleavage of PE, an opioid peptide
precursor. PE has been shown to be processed differentially by
PC1 and PC2. While PC1 cleaves this protein predominantly to
intermediates of 3–10 kDa in size, cleavage with PC2 yields
mainly mature small enkephalins (22, 23).
We compared the processing of PE by wild-type and mutant
PC2s in vivo by assessing the steady-state levels of two PE
cleavage products in AtT-20 cells stably expressing PE and
either wild-type PC2, PC2-(242–248), or PC2-(242–243). Fig. 7
shows the profiles of Met-enk-Arg-Gly-Leu immunoreactivity
for each cell line. In AtT-20/PE cells not transfected with PC2,
the major Met-enk-Arg-Gly-Leu-immunoreactive peak is represented by the 5.3-kDa intermediate (which terminates in
Met-enk-Arg-Gly-Leu). Transfection of the cells with either
wild-type PC2, PC-(242–248), or PC2-(242–243) resulted in the
elevated production of Met-enk-Arg-Gly-Leu-immunoreactive
peptides eluting in the position of the mature octapeptide.
However, no significant differences in the production of this
peptide were observed between cells expressing wild-type or
mutant PC2s. The molar ratios of the immunoreactivity associated with 5.3-kDa peptide to that of Met-enk-Arg-Gly-Leu
were similar in all three cell lines. We performed these experiments with three different clones for each type of PC2, and the
following ratios were observed: 1:1, 1:1.5, and 1:2 in cells expressing wild-type PC2; 1:1, 1:2, 1:2 in cells expressing PC2(242–243); and 1:1, 1:1.4, and 1:2 in cells expressing PC2-(242–
248). Thus, these experiments support the idea that the
specificity of the mutant enzymes is not shifted to the more
limited PC1-like profile, but that the mutants retain PC2-like
specificity.
␣-MSH Production in AtT-20/7B2 Cells Expressing PC2(242–243) or PC2-(242–248)—It has been previously shown
that in AtT-20 cells cleavage of POMC is mediated by PC1 and
PC2 in a strict temporal order (12, 24, 25); these studies have
supported the idea that PC1 cleaves this precursor predominantly during early steps of processing, while PC2 is involved
in later cleavages. ␣-MSH is among the products whose yield is
strongly enhanced by PC2 expression in AtT-20 cells (24, 25).
To determine whether the mutant PC2s lose the PC2-specific
ability to generate ␣-MSH, POMC processing was compared in
AtT-20/27 kDa-7B2 cells stably expressing either wild-type
PC2, PC2- (242–243), or PC2-(242–248). Fig. 8 shows the results of a pulse-chase experiment in which cell extracts were
14674
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
processing by mutant PC2s and wild-type PC2, these differences apparently did not result in the accumulation of any
changes to the final POMC processing pattern.
DISCUSSION
FIG. 7. Steady-state levels of proenkephalin processing products in AtT-20 cells. Cells extracts were prepared from subconfluent
10-cm dishes of AtT-20/PE cells expressing either wild-type PC2, PC2(242–248), PC2-(242–243), or no exogenous PC2. 100-␮l aliquots of
these cell extracts were then size-fractionated by HPGPC. Duplicate
aliquots of fractions were assayed for Met-enk-Arg-Gly-Leu-immunoreactivity by radioimmunoassay. Immunoreactivity is expressed as pmol
per fraction. The arrows indicate the positions of the size standards:
[35S]proenkephalin (PE), the 5.3-kDa proenkephalin-derived peptide
(5.3), and Met-enk-Arg-Gly-Leu.
immunoprecipitated with an antiserum directed against the
N-terminal portion of ACTH. Whereas in nontransfected AtT20/7B2 cells, POMC was cleaved predominantly to ACTH with
no detectable trace of ␣-MSH (Fig. 8A), in cells expressing
wild-type PC2, ␣-MSH represented the major labeled peptide
(B), and the ␣-MSH/ACTH ratio was approximately 2:1. However, decreased ability to produce ␣-MSH was observed for both
mutant PC2s. In cells expressing PC2-(242–248), the peak corresponding to ␣-MSH was 4 –5 times smaller than the ACTH
peak (Fig. 8C). In cells expressing the mutant PC2-(242–243),
production of ␣-MSH was elevated, indicating increased activity of this mutant compared with PC2-(242–248), but this peak
was still somewhat reduced compared with wild-type PC2 (Fig.
8D). These experiments were reproduced with two independent
clones of each mutant, and quantitatively similar results were
obtained. Thus, the PC2 mutants appear to be impaired with
regard to the rate of processing of POMC to ␣-MSH.
By contrast, using a steady labeling technique in which cells
are labeled for 6 h, no significant differences were observed in
the steady-state level of POMC products between cells expressing either wild-type PC2, PC2-(242–243), or PC2-(242–248);
␣-MSH represented the major radioactive peak in all three
cases (Fig. 9). Steady labeling results were confirmed using two
independent clones of wild-type PC2 and of each mutant PC2.
Although the rates of ␣-MSH production obtained in pulsechase experiments showed kinetic differences between POMC
The subtilisin-like enzymes PC2 and PC1, although sharing
over 50% homology in the catalytic domain, exhibit certain
differences in their substrate specificities (12, 22–27, 29, 30,
57). The two neuroendocrine-specific SPCs also exhibit dramatic differences in their profile of inhibition by peptide inhibitors such as the CT peptide (40, 41, 51). In addition to these
differences, PC2, but not PC1, specifically binds to the neuroendocrine protein 7B2 (32, 33, 58). Comparison of the primary
sequences of PC2s, PC1s, and furins within their proposed
substrate binding pockets (15, 31) has shown that the sequence
in positions 242–248 in PC2s differs considerably from the
corresponding sequences in PC1s and furins. Remarkably, all
PC1 and furins contain a single amino acid deletion in the
amino-terminal portion of this sequence. The only G residue in
the sequence, which is invariable in all PC1s and furins, corresponds to the invariable QP residues (positions 242 and 243)
found in PC2s. In the present study, we asked whether certain
PC2-specific properties are at least in part due to the specific
residues occurring in these positions. For this purpose, we
constructed two mutants of murine PC2 by replacing residues
in either positions 242 and 243 or positions 242–248 with the
corresponding residues found in murine PC1.
Our results indicate that while the insertion of PC1 sequence
into the PC2 backbone results in the synthesis of catalytically
active chimeric enzymes, the mutant PC2s seem not to exhibit
profound PC1-like changes in specificity. Processing of proenkephalin in AtT-20/PE cells expressing either wild-type or mutant PC2s resulted in a similar ratio of the PC1 and PC2
cleavage products, and PC2-(242–248) did not exhibit considerable changes in the kinetic constant toward the PC2-specific
internally quenched substrate. Some differences were observed
in the rate of production of ␣-MSH by wild-type and mutant
PC2s. The production of ␣-MSH by PC2-(242–243) and, especially, PC2-(242–248) proceeded at a slower rate as compared
with wild-type PC2. This change in rate, however, may result
not only from alterations in the substrate specificity of mutant
PC2s but also from a somewhat lower presence of the 66-kDa
PC2 form (due to the lower PC2/(PC2 ⫹ pro-PC2) ratio) found
in the cells expressing mutant PC2s as compared with the cells
expressing wild-type PC2 (data not shown). In any case, the
observed kinetic differences between POMC processing by
wild-type and mutant PC2s were apparently not large enough
to result in significant differences in the steady-state levels of
POMC-derived products in cells expressing the three PC2s.
Taken together, these data suggest that the distinct differences
in the substrate specificities of PC2 and PC1 do not significantly depend on sequence variations at residues 242–248.
In contrast to the lack of substantial effect on substrate
specificity, the ability of pro-PC2-(242–243) as well as pro-PC2(242–248) to bind to 7B2 (at least to its 21-kDa form) was
decreased to a large degree by the mutations introduced. The
latter finding indicates that the residue Gln242 and/or the residue Pro243 play a critical role in the interaction of PC2 with
7B2. Based on the three-dimensional structural model proposed for furin (50), Gln242 and Pro243 are located within a loop
on the surface of PC2, which would render them accessible to
7B2. Two other residues in the catalytic domain of PC2, Tyr194
and Asp309, have been previously shown to be involved in
PC2–7B2 binding (47– 49). According to the three-dimensional
model cited above, Tyr194 and Asp309 reside on the same side of
the PC2 surface as do Gln242 and Pro243. However, these two
residues are not located in close vicinity to Gln242 and Pro243; it
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
14675
FIG. 8. POMC processing by mutant
and wild-type PC2s in AtT-20/27-kDa
7B2 cells: pulse/chase labeling. AtT20/27-kDa 7B2 cells expressing either
wild-type PC2 (B), PC2-(242–248) (C),
PC2-(242–243) (D), or no exogenous PC2
(A) were labeled for 20 min and chased for
60 min before extraction in acid mix. Cell
extracts were immunoprecipitated with
an antibody directed against the N-terminal portion of ACTH. 50-␮l aliquots of the
immunoprecipitated proteins were sizefractionated by HPGPC, and radioactivity
was detected on-line. Radioactivity is expressed per amount applied onto the
HPGPC column (50 ␮l). The arrows indicate expected elution positions of POMC,
ACTH, and ␣-MSH. Similar results were
obtained in two experiments and using
two independent clones for each mutant.
FIG. 9. POMC processing by mutant
and wild-type PC2s in AtT-20/27-kDa
7B2 cells: steady-state labeling. AtT20/27-kDa 7B2 cells expressing either
wild-type PC2 (B), PC2-(242–248) (C),
PC2-(242–243) (D), or no exogenous PC2
(A) were labeled for 6 h, and cell extracts
were immunoprecipitated with an antibody directed against the N-terminal portion of ACTH. 30-␮l aliquots of the immunoprecipitated proteins were sizefractionated by HPGPC for on-line
detection of radioactivity (expressed per
amount applied onto the HPGPC column;
30 ␮l). For arrows, see legend to Fig. 8.
Similar results were obtained in two experiments and using two independent
clones for each mutant.
therefore seems unlikely that replacement of the Gln242 and
Pro243 residues could directly affect Tyr194 and Asp309. We
suggest that the PC2 region containing Gln242 and Pro243 probably represents a 7B2 binding determinant distinct from that
of Tyr194 and Asp309. Our data support the idea that binding to
7B2 involves multiple independent determinants within PC2.
Both mutant PC2s, when expressed in AtT-20/27-kDa 7B2
cells, exhibit a slow rate of intracellular maturation and substantially reduced enzymatic activity in the conditioned medium as compared with wild-type PC2. Pro-PC2-(242–248) immunopurified from these cells also displays a highly reduced
ability for autocatalytic conversion at acidic pH. Since interac-
tion with 7B2 is required for pro-PC2 maturation and activation (32–37, 45), we suggest that these properties of PC2-(242–
243) and PC2-(242–248) are at least partially due to impaired
interaction of mutants with 7B2. When expressed in CHO/21kDa 7B2 cells, the mutant PC2s and wild-type PC2 appear to
exhibit more similar rates of autocatalytic conversion and more
comparable levels of activity against the fluorogenic substrate.
One possible explanation for this discrepancy is that CHO/21kDa 7B2 cells express extremely high quantities of 7B2; thus,
mutant pro-PC2s, whose ability to bind 7B2 is greatly impaired, have a much greater potential for interaction with 7B2
in CHO/21-kDa 7B2 cells than do the same mutants expressed
14676
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
in AtT-20/27-kDa 7B2 cells. While their affinity for 7B2 is
severely reduced, mutant pro-PC2s most likely do not totally
lose their ability to interact with 7B2 (although this interaction
was not detectable by co-immunoprecipitation), since mutant
pro-PC2s transiently expressed in a constitutive cell line HEK
293 continue to absolutely require co-expression of 7B2 to become enzymatically competent.
Our data clearly indicate that the PC2 residues in positions
242–248 play a crucial role for inhibition of PC2 with its specific inhibitor, the 7B2 CT peptide (40, 41). Replacement of
residues 242–243 with the corresponding PC1 residues resulted in a 50-fold increase in the IC50 value for the CT peptide,
and replacement of the whole stretch resulted in a more than
1000-fold increase. Thus, PC2-(242–248) not only exhibits a
highly decreased ability to bind to 7B2 (at least to its 21-kDa
form) but also almost totally loses its ability to be inhibited by
the 7B2 CT peptide. These data indicate that mutations in the
region 242–248 affect both the binding site for 21-kDa 7B2 and
the binding site for the CT peptide on the PC2 molecule.
In addition to loss of inhibition by 7B2 CT peptide, the
mutant PC2s also exhibited a loss of inhibition with another
long PC2 inhibitor, a PC2 propeptide fragment. These data
imply that the CT peptide and this PC2 propeptide fragment
bind to the same site (or to overlapping sites) on PC2. However,
this dramatic alteration in inhibitory susceptibility does not
extend to small peptide inhibitors; inhibition of mutant PC2s
by the synthetic hexapeptide Ac-LLRVKR-NH2 (a peptide originally identified during a combinatorial library screen of PC1;
Ref. 51) was not significantly altered as compared with wildtype PC2. This hexapeptide represents a much more potent
inhibitor of PC1 than PC2 (51). Because of the competitive
nature of this inhibition the hexapeptide most probably binds
directly to the substrate binding pockets of PC2 and PC1. The
fact that similar inhibition constants were obtained against
wild type and the mutant PC2s suggests that mutation of
residues 242–248 does not significantly affect the PC2
hexapeptide binding site within the substrate binding pocket.
We therefore speculate that inhibition of PC2 with the CT
peptide/PC2 propeptide and with the synthetic hexapeptide
depends on interaction with separate sites within the PC2
binding pocket. The PC2 site required for inhibition with CT
peptide or PC2 propeptide may potentially be located close to,
but outside of, the hexapeptide binding region of PC2. Indeed,
our previous data support the idea that the inhibitory potency
of the CT peptide is at least partially attributable to residue
interactions more distal to the hexapeptide binding site, since
a synthetic hexapeptide based upon the CT peptide sequence
(and including its inhibitory K-K pair) is completely inactive
against PC2 (51). In fact, one of the most conserved portions of
7B2 is the VNPYLQG heptapeptide located 10 –17 residues
away from the K-K pair. Taken together, our results imply that
residues 242–248 most likely contribute to these distal inhibitory interactions with longer peptide inhibitors.
In conclusion, our data using PC1-substituted mutant PC2s
demonstrate that residues 242–248 do not play a significant
role in defining the substrate specificity of PC2 but do contribute greatly to binding of 7B2 and are critical for the inhibition
of PC2 with the 7B2 CT peptide. Further mutation of catalytic
pocket residues will help to identify additional residues that
differentiate the cellular biochemistry of these two neuroendocrine SPCs.
Acknowledgments—We thank Joelle Finley for assistance with tissue
culture and Dr. Dick Mains (Johns Hopkins University School of Medicine, Baltimore, MD) for providing antiserum JH93 as well as AtT-20/
PC2 cells. We also thank Dr. Luiz Juliano (Escola Paulista de Medicine,
Sao Paolo, Brazil) for providing the internally quenched fluorogenic
substrate and Jon Appel and Dr. Richard Houghten (Torrey Pines
Institute for Molecular Studies, San Diego, CA) for providing the synthetic hexapeptide.
REFERENCES
1. Rouille, Y., Duguay, S. J., Lund, K., Furuta, M., Gong, Q., Lipkind, G., Oliva,
A. A. Jr, Chan, S. J., and Steiner, D. F. (1995) Front. Neuroendocrinol. 16,
322–361
2. Seidah, N. G., and Chretien, M. (1997) Curr. Opin. Biotechnol. 8, 602– 607
3. Zhou, A., Webb, G., Zhu, X., and Steiner, D. F. (1999) J. Biol. Chem. 274,
20745–20748
4. Christie, D., Batchelor, D. C., and Palmer, D. J. (1991) J. Biol. Chem. 266,
15679 –15683
5. Benjannet, S., Reudelhuber, T., Mercure, C., Rondeau, N., Chretien, M., and
Seidah, N. G. (1992) J. Biol. Chem. 267, 11417–11423
6. Zhou, Y., and Lindberg, I. (1993) J. Biol. Chem. 268, 5615–5623
7. Shennan, K. I. J., Taylor, N. A., Jermany, J. L., Matthews, G., and Docherty,
K. (1995) J. Biol. Chem. 270, 1402–1407
8. Matthews, G., Shennan, K. I. J., Seal, A. J., Taylor, N. A., Colman, A., and
Docherty, K. (1994) J. Biol. Chem. 269, 588 –592
9. Scougall, K., Taylor, N. A., Jermany, J. L., Docherty, K., and Shennan, K. I. J.
(1998) Biochem. J. 334, 531–537
10. Lamango, N. S., Apletalina E., Liu J, and Lindberg I (1999) Arch. Biochem.
Biophys. 362, 275–282
11. Benjannet, S., Rondeau, N., Paquet, L., Boudreault, A., Lazure, C., Chretien,
M., and Seidah, N. G. (1993) Biochem. J. 294, 735–743
12. Zhou, A., and Mains, R. E. (1994) J. Biol. Chem. 269, 17440 –17447
13. Coates, L. C., and Birch, N. P. (1997) J. Neurochem. 68, 828 – 836
14. Boudreault, A., Gauthier, D., and Lazure, C. (1998) J. Biol. Chem. 273,
31574 –31580
15. Siezen, R., and Leunissen, J. A. (1997) Protein Sci. 6, 501–523
16. Dupuy, A., Lindberg, I., Zhou, Y., Akil, H., Lazure, C., Chretien, M., Seidah,
N. G., and Day, R. (1994) FEBS Lett. 337, 60 – 65
17. Dhanvantari, S., Seidah, N. G., and Brubaker, P. L. (1996) Mol. Endocrinol.
10, 342–355
18. Day, R., Lazure, C., Basak, A., Boudreault, A., Limperis, P., Dong, W., and
Lindberg, I. (1998) J. Biol. Chem. 273, 829 – 836
19. Jean, F., Boudreault, A., Basak, A., Seidah, N. G., and Lazure, C. (1995)
J. Biol. Chem. 270, 19225–19231
20. Johanning, K., Juliano M. A., Juliano, L., Lazure, C., Lamango, N., Steiner,
D. F., and Lindberg, I. (1998) J. Biol. Chem. 273, 22672–22680
21. Lazure, C., Gauthier, D., Jean, F., Boudreault, A., Bennett, H. P. J., and
Hendy, G. (1998) J. Biol. Chem. 273, 8572– 8580
22. Mathis, J., and Lindberg, I. (1992) Endocrinology 131, 2287–2296
23. Johanning, K., Mathis, J. P., and Lindberg, I. (1996) J. Neurochem. 66,
898 –907
24. Benjannet, S., Rondeau, N., Day, R., Chretien, M., and Seidah, N. G. (1991)
Proc. Natl. Acad. Sci. U. S. A. 88, 3564 –3568
25. Zhou, A., Bloomquist, B. T., and Mains, R. E. (1993) J. Biol. Chem. 268,
1763–1769
26. Rouille, Y., Martin, S., and Steiner, D. F. (1995) J. Biol. Chem. 270,
26488 –26496
27. Rouille, Y., Kantengwa, S., Irminger, J. C., and Halban, P. A. (1997) J. Biol.
Chem. 272, 32810 –32816
28. Yoon, J., and Beinfeld, M. C. (1997) Endocrinology 138, 3620 –3623
29. Rovere, C., Barbero, P., and Kitabgi, P. (1996) J. Biol. Chem. 271, 11368 –75
30. Viale, A., Ortola, C., Hervieu, G., Furuta, M., Barbero, P., Steiner, D. F.,
Seidah, N. G., and Nahon, J.-L. (1999) J. Biol. Chem. 274, 6536 – 6545
31. Lipkind, G., Gong, Q., and Steiner, D. F. (1995) J. Biol. Chem. 270,
13277–13284
32. Zhu, X., and Lindberg, I. (1995) J. Cell Biol. 129, 1641–1650
33. Benjannet, S., Savaria, D., Chretien, M., and Seidah, N. G. (1995) J. Neurochem. 64, 2303–2311
34. Muller L., Zhu X., and Lindberg I. (1997) J. Cell Biol. 139, 625– 638
35. Seidel, B., Dong, W., Savaria, D., Zheng, M., Pintar, J. E., and Day, R. (1998)
DNA Cell Biol. 17, 1017–1029
36. Barbero, P., and Kitabgi, P. (1999) Biochem. Biophys. Res. Commun. 257,
473– 479
37. Westphal, C. H., Muller, L., Zhou, A., Bonner-Weir, S., Schambelan, M.,
Steiner, D. F., Lindberg, I., and Leder, P. (1999) Cell 96, 689 –700
38. Ayoubi, T. A. Y., van Duijnhoven, H. L. P., van de Ven, W. J. M., Jenks, B. G.,
Roubos, E. W., and Martens, G. J. M. (1990) J. Biol. Chem. 265,
15644 –15647
39. Paquet, L., Bergeron, F., Boudreault, A., Seidah, N. G., Chretien, M., Mbikay,
M., and Lazure, C. (1994) J. Biol. Chem. 269, 19279 –19285
40. Lindberg, I., Van den Hurk, W. H., Bui, C., and Batie, C. J. (1995) Biochemistry
34, 5486 –5493
41. Van Horssen, A. M., Van den Hurk, W. H., Bailyes, E. M., Hutton, J. C.,
Martens, G. J. M., and Lindberg, I. (1995) J. Biol. Chem. 270, 14292–14296
42. Zhu, X., Rouille, Y., Lamango, N. S., Steiner, D. F., and Lindberg, I. (1996)
Proc. Natl. Acad. Sci. U. S. A. 93, 4919 – 4924
43. Fortenberry, Y., Liu, J., and Lindberg, I. (1999) J. Neurochem. 73, 994 –1003
44. Apletalina, E. V., Juliano, M. A., Juliano, L., Lindberg, I. (2000) Biochem.
Biophys. Res. Commun. 267, 940 –942
45. Martens, G. J. M., Braks, J. A. M., Eib, D. W., Zhou, Y., and Lindberg, I. (1994)
Proc. Natl. Acad. Sci. U. S. A. 91, 5784 –5785
46. Muller, L., Zhu, P., Juliano, M. A., Juliano, L., and Lindberg, I. (1999) J. Biol.
Chem. 274, 21471–21477
47. Zhu, X., Muller, L., Mains, R. E., and Lindberg, I. (1998) J. Biol. Chem. 273,
1158 –1164
48. Benjannet, S., Lusson, J., Savaria, D., Chretien, M., and Seidah, N. G. (1995)
FEBS Lett. 362, 151–155
49. Benjannet, S., Mamarbachi, A. M., Hamelin, J., Savaria, D., Munzer, J. S.,
Chretien, M., and Seidah, N. S. (1998) FEBS Lett. 428, 37– 42
PC2 Mutation Causes Loss of Inhibition with 7B2 CT Peptide
50. Siezen, R. J., Creemers, J. W. M., and Van de Ven, W. J. M. (1994) Eur.
J. Biochem. 222, 255–266
51. Apletalina, E., Appel, J., Lamango, N. S., Houghten, R. A., and Lindberg, I.
(1998) J. Biol. Chem. 273, 26589 –26595
52. Lindberg, I., and Zhou, Y. (1995) in Peptidases and Neuropeptide Processing
(Smith, A. I., ed) pp. 94 –108, Academic Press, Inc., San Diego
53. Shen, F-S, Seidah, N. G. and Lindberg, I. (1993) J. Biol. Chem. 268,
24910 –24915
14677
54. Mains, R. E., Dickerson, I. M., May, V., Stoffers, D. A., Perkins, S. N., Ouafik,
L., Husten, E. J., and Eipper, B. (1990) Front. Neuroendocrinol. 11, 52– 89
55. Lindberg, I., and White, L. (1986) Biochem. Biophys. Res. Commun. 139,
1024 –1032
56. Lamango, N. S., Zhu, X., and Lindberg, I. (1996) Arch. Biochem. Biophys. 330,
238 –250
57. Yoon, J., and Beinfeld, M. C. (1997) J. Biol. Chem. 272, 9450 –9456
58. Braks, J. A. M., and Martens, G. J. M. (1994) Cell 78, 263–273