ArticlePDF Available

Placenta growth factor: Potentiation of vascular endothelial growth factor bioactivity, in vitro and in vivo, and high affinity binding to Flt-1 but not to Flk-1/KDR

Authors:

Abstract

The recently identified placenta growth factor (PIGF) is a member of the vascular endothelial growth factor (VEGF) family of growth factors. PIGF displays a 53% identity with the platelet-derived growth factor-like region of VEGF. By alternative splicing of RNA, two PIGF isoforms are generated: PIGF131 (PIGF-1) and PIGF152 (PIGF-2). Relative to PIGF131, PIGF152 has a 21-amino acid insertion enriched in basic amino acids. Little is known at the present time about the significance and function of these proteins. To assess their potential role, we cloned the cDNAs coding for both isoforms, expressed them in mammalian cells, and purified to apparent homogeneity the recombinant proteins. Like VEGF, the PIGF isoforms are homodimeric glycoproteins. PIGF131 is a non-heparin binding protein, whereas PIGF152 strongly binds to heparin. We examined the ability of PIGF to bind to soluble VEGF receptors, Flt-1 and Flk-1/KDR, and characterized the binding of PIGF to endothelial cells. While the PIGF proteins bound with high affinity to Flt-1, they failed to bind to Flk-1/KDR. Binding of 125I-PIGF to human endothelial cells revealed two classes of sites, having high and low affinity. The high affinity site is consistent with Flt-1; the identity of the low affinity site remains to be determined. Purified PIGF isoforms had little or no direct mitogenic or permeability-enhancing activity. However, they were able to significantly potentiate the action of low concentrations of VEGF in vitro and, more strikingly, in vivo.
THE
JOURNAL
OF
BIOL~CICAL
CHEMISTRY
0
1994 by The American Society
for
Biochemistry and Molecular Biology, Inc. Vol. 269,
No.
41, Issue
of
October 14,
pp.
2564625654, 1994
Printed
in
U.S.A.
Placenta
Growth
Factor
POTENTIATION OF VASCULAR ENDOTHELIAL GROWTH FACTOR BIOACTIVITY,
IN VITRO
AND
IN VIVO,
AND
HIGH
AFFINITY BINDING TO
Flt-1
BUT NOT
TO
Flk-l/KDR*
(Received for publication, April 12, 1994, and in revised form, July 18, 1994)
John
E.
Park, Helen
H.
Chen, Jane Winer, Keith
A..
HouckS, and Napoleone FerraraO
From Genentech, Inc., South San Francisco, California
94080
The recently identified placenta growth factor (PlGF)
is a member
of
the vascular endothelial growth factor
(VEGF) family of growth factors. PlGF displays a
53%
identity with the platelet-derived growth factor-like re-
gion of VEGF.
By
alternative splicing of
RNA,
two PlGF
isoforms are generated PlGF,,, (PlGF-1) and PlGF,,,
(PlGF-2). Relative to PIGF,,,, PlGF,,, has a 21-amino acid
insertion enriched in basic amino acids. Little
is
known
at the present time about the significance and function
of these proteins.
To
assess their potential role, we
cloned the cDNAs coding for both isoforms, expressed
them in mammalian cells, and purified to apparent ho-
mogeneity the recombinant proteins. Like VEGF, the
PlGF isoforms are homodimeric glycoproteins. PlGF,,,
is a non-heparin binding protein, whereas PIGFIG,
strongly binds to heparin. We examined the ability of
PlGF to bind to soluble VEGF receptors, Flt-1 and
Flk-l/KDR, and characterized the binding of PlGF to en-
dothelial cells. While the PlGF proteins bound with high
affinity to Flt-1, they failed to bind to Flk-l/KDR. Bind-
ing of l2%P1GF to human endothelial cells revealed two
classes of sites, having high and low affinity. The high
affinity site is consistent with Flt-1; the identity of the
low affinity site remains to be determined. Purified
PlGF isoforms had little or no direct mitogenic or per-
meability-enhancing activity. However, they were able
to significantly potentiate the action of low concentra-
tions of VEGF
in vitro
and, more strikingly,
in vivo.
The vascular endothelium is one of the most versatile sys-
tems
in
the body, serving
a
variety of essential exchange and
regulatory functions. A fundamental property of vascular en-
dothelial cells
is
the ability to proliferate and form
a
network of
capillaries. This process, known as angiogenesis,
is
prominent
during embryonic development (1,2). In a normal adult, angio-
genesis occurs only following injury or, in
a
cyclical fashion, in
the endometrium and
in
the ovary (3). Angiogenesis, however,
is known to play
a
critical role
in
the pathogenesis of a variety
of disorders, most notably growth and metastasis of solid
tumors (4).
Several potential positive regulators of angiogenesis have
been described including acidic and basic fibroblast growth
factors, transforming growth factors
a
and
p,
tumor necrosis
*
The costs of publication
of
this
article
were
defrayed
in
part
by
the
payment
of
page charges.
This article must therefore
be
hereby marked
"a'aduertisement"
in
accordance
with
18
U.S.C. Section
1734
solely
to
indicate this
fact.
27717.
$
Present address: Sphinx Pharmaceuticals Corp., Durham,
NC
Dept.
of
Cardiovascular Research, Genentech Inc.,
460
Point San
Bruno
8
To whom correspondence and reprint requests should
be
addressed:
Blvd., South San Francisco,
CA
94080. Tel.: 415-225-2968; Fax: 415-
225-6327.
factor
a,
and angiogenin (5-14). The recently described vascu-
lar endothelial growth factor (VEGF)'
has
several attractive
features
as
a regulator of normal and pathological angiogenesis
(15).
VEGF is an endothelial cell-specific mitogen
in vitro
and an
angiogenic inducer
in uiuo.
In addition, VEGF is able to induce
vascular leakage in the Miles assay (7,8). On the basis of
this
activity, VEGF
has
been also implicated
as
a mediator of the
abnormal permeability properties of tumor blood vessels
(7,
8).
By alternative splicing of mRNA, VEGF may exist in four
different homodimeric molecular species, each monomer hav-
ing, respectively, 121,
165,
189, or 206 amino acids (VEGF,,,,
VEGF,,,,VEGF,,,, VEGF,,,). VEGF,,, and VEGF,,, are diffus-
ible proteins, whereas VEGF,,, and VEGF,,, are mostly bound
to heparin-containing proteoglycans in the extracellular matrix
(16, 17). Recent studies have suggested
that
VEGF
is
an im-
portant regulator of both developmental and ovarian angiogen-
esis (18-21). Furthermore, inhibition of VEGF action results in
marked suppression
of
the growth of several human tumor cell
lines
in
nude mice (22).
Two
tyrosine kinases have been identified
as
putative VEGF
receptors (23,241. The fms-like tyrosine kinase (Flt-1) (25) and
the kinase domain region
(KDR)
(26) proteins have been shown
to bind VEGF with high affinity. The murine homologue of
KDR, known
as
fetal liver kinase-1 (Flk-1) (271, has been also
shown to bind VEGF (28, 29). Flt-1 and Flk-l/KDR receptors
have
a
single signal sequence, one transmembrane domain,
seven immunoglobulin-like domains in their extracellular do-
main, and
a
consensus tyrosine kinase sequence domain. In
addition,
a
cDNAencoding
a
truncated form of Flt-1 lacking the
seventh immunoglobulin-like domain, the cytoplasmic domain,
and transmembrane sequence has been identified in human
umbilical vein endothelial cells (30).
Recently,
a
cDNA encoding
a
protein having a 53% identity
with the platelet-derived growth factor-like region of VEGF has
been isolated from a human placental cDNA library
(31).
The
encoded protein, named placenta growth factor (PlGF), was
expected to have 149 amino acids. Subsequently,
a
longer PlGF
cDNA was identified (32,331. Compared to the originally iden-
tified PlGF species, the long PlGF protein was expected to have
a
21-amino acid insertion highly enriched in basic residues.
The two isoforms, resulting from alternative splicing of RNA,
were named, respectively, PlGF-1 and PlGF-2 (32, 33). In con-
trast
to the widespread distribution of the VEGF mRNA
(151,
expression of PlGF mRNA appears to be restricted
to
placenta,
trophoblastic tumors, and cultured human endothelial cells
(31-33). Based on the homology with VEGF, PlGF was pro-
The abbreviations
used
are: VEGF, vascular endothelial growth fac-
tor;
Flt-1, fms-like tyrosine
kinase;
Flk-1, fetal liver
kinase
1;
KDR,
kinase
domain
region;
PlGF,
placenta
growth
factor;
PAGE,
polyacryl-
nant human;
b,
basic;
PBS, phosphate-buffered
saline;
ACCE, adrenal
amide
gel
electrophoresis; FGF, fibroblast growth factor;
rh,
recombi-
cortex-derived capillary
endothelial;
HUVE,
human
umbilical
vein
endothelial.
25646
PlGF Binds to Flt-1 and Potentiates VEGF Activity
25647
.~
Flt-1
IgG:
GAATTC C ATG GTC AGC TAC.
.
. . . .
.
. .
. . . . . .
.
EcoRI
MVSY4
BstBI
4
TCT AAT
TTC
Ge GAC
AAA
ACT.
. .
S
N
E
E
D
K
T761
Original
receptor:
MVSY4
E758
~
KDR
IgG:
ClaI
sper
ATCG ATG GAG AGC AAG..
.
.TCA CTA GTT.
. .
.TTG GAT CCG TTC G.
BamHI
BStBI
4
”-
“ESK4
S
L
v416
L
2
2
E
ID
K
T770
Original
receptor:
s
K2
s
L
v414
L
E
762
FIG.
1.
Partial amino acid and
cDNA
sequence
of
Flt-1
IgG
and
KDR
IgG. Arrows
point to the junction between the extracellular domain
of each receptor and the IgG. Below each receptor-IgG, sequence of the original receptor is shown for comparison. Underlined amino acids or
nucleotides indicate modifications of the original sequences.
IgG
residues are enclosed in boxes.
posed to be an angiogenic factor, and initial evidence suggested
that
conditioned media of transfected cells expressing PlGF
were weakly mitogenic to endothelial cells derived from bovine
pulmonary artery
or
aorta
(31,331.
To
date, PlGF has not been
reported to have been purified to homogeneity.
In
the present study, we purified
to
apparent homogeneity
the two PlGF isoforms and tested them
in
vitro
for mitogenic
activity
on
bovine and human vascular endothelial cells and in
an
in
vivo
system, the Miles vascular permeability assay. Fur-
thermore, we examined their ability
to
bind to soluble VEGF
receptors,
Flt-1
and Flk-l/KDR, and characterized the binding
of
lZ5I-PlGF to endothelial cells. PlGF was able to bind with
high affinity to Flt-1 but not to Flk-liKDR,
a
receptor thought
to be critically involved in mediating WGF biological actions.
PlGF had little
or
no direct mitogenic or permeability-enhanc-
ing activity. Intriguingly, however, it was able to significantly
potentiate the effects of low concentrations of WGF, both
in
vitro
and
in
vivo.
EXPERIMENTAL PROCEDURES
Materials-Tissue culture reagents and media were obtained from
Life Technologies, Inc. through the Genentech media facility. Restric-
tion enzymes were from New England Biolabs except for Pfu polymer-
ase which was from Stratagene. Q-Sepharose Resource fast protein
liquid chromatography column
(6
ml), protein A-Sepharose, wheat germ
agglutinin agarose, heparin-Sepharose, and S-Sepharose fast flow res-
ins were from Pharmacia Biotech. The C4 reversed phase high perform-
ance liquid chromatography column (4.6
x
100 mm) was purchased from
SynChrom (Lafayette, IN). The preparative TSK 3000 GS column (21
x
300 mm) was from Hewlett-Packard. Trifluoroacetic acid was from
Pierce. Acetonitrile was purchased from Fisher Scientific. Tissue cul-
ture plates were from Costar except for large scale Nunc plates (24.5
x
24.5 cm), which were from Applied Scientific. Molecular weight stand-
ards for gel chromatography and prestained low molecular weight
markers for SDSPAGE gel were purchased from Bio-Rad. Recombinant
human (rh) basic FGF (bFGF) was from R
&
D Systems (Minneapolis,
MN). rhVEGF,,, was purified from transfected Chinese hamster ovary
cells
as
described (34). Iodination of VEGFle, and PlGF proteins was
performed by the indirect IODOGEN method
(35).
Specific activities
ranged from
18,000
to 32,000 cpdng for lZ5I-P1GF and 48,000 to 97,000
cpdng for lZ5I-VEGF. Rabbit polyclonal antibodies to Flk-l/KDR and
Flt-1
were produced by immunizing New Zealand White rabbits
monthly with purified Flk-1 IgG or Flt-1
IgG.
GCCGGTCA TGAGGCTGT 3’ and
5’
GAATTCTCTAGAGGTTACCTC-
Cloning
of
PLGF-Oligonucleotide primers
(5’
GAATTCTCTAGAT-
CGGGGAACAGCATC 3’) were designed to amplify
a
full-length PlGF
cDNA clone
(31).
The polymerase chain reaction was performed using
human placenta cDNA as the template. The polymerase chain reaction
products were subjected
to
polyacrylamide gel electrophoresis.
Ethidium bromide staining of the gel revealed two reaction products
at
475 and 538 base pairs, respectively. The lower molecular weight prod-
uct is consistent with PlGF-1, the higher with the longer isoform,
PlGF-2 (31-33). The bands were cut, electroeluted, and subcloned into
the XbaI site
of
pHEB023 for expression in CEN4 cells. Such cells are
a derivative of the human embryonic kidney 293 cell line that stably
expresses the Epstein-Barr virus nuclear antigen-1, required for episo-
mal replication of pHEB023 vector (36). Transfections and selection of
transformants were performed
as
described previously
(16,
17). The
authenticity of all clones was verified by DNA sequencing.
Construction of VEGF Receptor-IgG Chimeras-The extracellular do-
mains of Flt-l and KDR were cloned by polymerase chain reaction using
Pfu polymerase. Human placental cDNA served
as
the template. Prim-
ers encompassed the entire extracellular domains, including signal pep-
tides (25, 26). The cDNAs for each receptor extracellular domain were
cloned in two pieces to facilitate sequencing.
Two
sets of primers (shown
below) were used, and the resulting bands
(-1
kilobase) were digested
with the appropriate enzymes and subcloned into pBluescript I1
or
pSL301. The cDNAs produced encoded the
first
758 amino acids
of
Flt-1
and the first
762
amino acids of KDR. Since the
5’
end of the published
KDR sequence
is
incomplete (26), the initiator methionine and gluta-
mate were added to enable secretion of the fusion protein, resulting in
a 764-amino acid extracellular domain. The KDR primers add a silent
internal
SpeI
site for cloning purposes (Fig.
1).
Full-length KDR extra-
cellular domain cDNA was created by ligating the two parts at the novel
SpeI site, while full-length Flt-1 extracellular domain cDNA was cre-
ated by ligating the two
Flt-1
polymerase chain reaction clones
at
a
unique natural MunI site.
Flt-1 Set
1:
5‘
TCTAGAGAATTCCATGGTCAGCTACTGGGACACC
3’
5’
CCAGGTCATTTGAACTCTCGTGTTC
3’
Flt-1 Set
2
5‘
TACTTAGAGGCCATACTCTTGTCCT
3’
5‘
GGATCCTTCGAAATTAGACTTGTCCGAGGTTC
3’
KDR
Set
1
5’
GAATTCATCGATGGAGAGCAAGGTGCTGCTGGCCGTC
3’
5’
ACACAACTAGTGAGACCACATGGCTCTGCTTCTC
3‘
KDR
Set
2
5: GGTCTCACTAGTTGTGTATGTCCCACCCCAGATT
3‘
5‘
GAATTCGGATCCAAGTTCGTCTTTTCCGGGCA
3‘
These primers changed amino acid 757
to
phenylalanine and introduced
a
BstBI site
at
the 3’ end of the
Flt-1
extracellular domain (Fig.
1).
The
KDR primers added
a
BamHI site to the
3’
end of the KDR extracellular
domain. A BstBI mutation which eliminated any linker sequences was
introduced
at
the
5’
end of CH,CH,, an IgGyl heavy chain cDNA clone
(37). Flt-1 extracellular domain sequences were fused to the coding
sequences for amino acids 216443
of
this IgGyl heavy chain clone via
the unique BstBI site
at
the 3’ end of the extracellular domain coding
region while KDR extracellular domain sequences were fused
to
the
upstream BamHI site (see Fig. 1). Both constructs were then subcloned
into pHEB023 for expression in CEN4 cells (17). The authenticity of all
clones was verified by DNA sequencing. The strategies for construction
KDR IgG.
and cloning of Flk-1
IgG
were essentially similar to those applied to
Purification
of
VEGF Receptor-ZgG Chimeras-Conditioned media of
transfected CEN4 cells expressing the three chimeric receptors were
concentrated
5-
to 10-fold by ultrafiltration and then applied onto pro-
tein A-Sepharose columns
(10
ml) that had been pre-equilibrated in
PBS. The flow
rate
was 2 mumin. Absorbance was monitored at 280 nm.
The columns were washed with PBS until the absorbance at 280 nm
became negligible and then eluted with 100 mM citric acid, pH
3.0.
Sufficient 2
M
Tris
base was added
to
tubes to immediately neutralize
the acid. A silver-stained SDSPAGE gel revealed the presence of a
single major band
at
>300 kDa in nonreducing and
-180
kDa in reduc-
ing conditions. The authenticity of the proteins was verified by NH,-
terminal amino acid sequence.
Purification
of
PZGF Isoforms-Serum-free conditioned media from
transfected CEN4 cells expressing each PlGF molecular species
(-2.5
liters) were concentrated
5-
to 10-fold by ultrafiltration using Amicon
stir cells
CYM
10
membrane). For purification of PlGF,,JPlGF-2 (pre-
dictably
a
basic protein), the concentrated conditioned medium was
extensively dialyzed against 25 mM sodium phosphate, pH
6.0,
and then
applied onto
a
cation exchange S-Sepharose fast flow resin packed in a
25648
PlGF
Binds to Flt-1 and Potentiates
VEGF
Activity
glass column (Omni,
10
x
100 mm) pre-equilibrated with the same
buffer. The flow rate was 2 mumin. After loading, the column was
washed with 22 ml of 25
m~
sodium phosphate, pH
6.0,
containing 0.4
M
NaC1. This wash resulted in a major peak of absorbance at 280 nm
that contained less than 10% of the activity able to compete with lZ5I-
VEGF,,, for binding to Flt-IgG. Such radioreceptor assay was used at
all steps to monitor the purification. Elution of bound PlGF was with a
linear gradient
(0.4-1.0
M)
of NaCl over 30 min. Fractions containing
the highest competing activity
(0.5-0.7
M
NaCl) were pooled, diluted
6-fold in water containing 0.1% trifluoroacetic acid, and applied by
multiple injections into a C4 reversed phase column that had been
pre-equilibrated with 20% acetonitrile/
0.1%
trifluoroacetic acid. The
flow rate was
0.6
mumin. The column was washed with
5
ml of equil-
ibration buffer and was then eluted with a linear gradient of acetonitrile
(2045%) in 95 min. The activity eluted as a broad and asymmetric peak
of absorbance at 210 nm between 30 and 32% acetonitrile.
A
silver-
stained SDSffAGE gel of the most active fractions revealed the pres-
ence of a single band at -32 kDa in reducing conditions. Such fractions
were pooled and subjected
to
microsequencing. The yield in purified
protein was approximately
0.5
mgfliter, as estimated by amino acid
analysis. The recovery, as assessed by radioreceptor assay, was -40%.
For purification of PlGF,,,/PlGF-l, concentrated conditioned medium
was dialyzed against 20
m~
Tris,
pH 8.0, and then applied onto a
Q-Sepharose Resource anion exchange column. The flow rate was 2
mumin. The column was eluted with a linear gradient of NaCl(0-0.5
M)
in 30 min. The activity capable of competing with lZ5I-VEGF for binding
to
Flt-IgG was eluted in the presence of -0.2
M
NaC1. The most active
fractions were pooled and then applied onto a TSK 3000 GS column
equilibrated with
100
mM KH,PO,, pH 6.8. The flow rate was 3 mumin.
The activity eluted with an apparent molecular weight of
-60,000.
Fractions containing such activity were further purified by reverse
phase chromatography, exactly as described for PlGF-2. Purified pro-
tein was quantified by amino acid analysis. Protein yield and recovery
were very similar to those described for PlGF-2.
Binding Assays with Soluble Receptors-Ninety-six-well breakaway
enzyme-linked immunosorbent assay plates (Nunc, Kampstrup, Den-
mark) were coated overnight at 4 "C with
2
pg/ml affinity-purified goat
anti-human Fc IgG (Organon-Teknika) in 50 mM Na,CO,, pH
9.6.
Plates
were blocked for
1
h
with 10% fetal bovine serum in PBS (buffer
B).
Blocking buffer was then removed.
To
each well, IgG fusion protein
(-1
ng), "'1-VEGF or 1251-P1GF, and cold competitor were added
to
a final
volume of 100
1.11
in buffer B. Unless noted otherwise, 12,000 cpm
of
radioligand were added to the assay. When noted, binding experiments
contained
1-10
pg/ml heparin. Binding was carried out at room tem-
perature for
3.5
h followed by 6 washes with buffer B. Binding was
determined by counting individual wells in a
y
counter. Data were
analyzed by a 4-parameter nonlinear curve-fitting program developed
at Genentech.
Endothelial Cell Culture and Mitogenic Assays-Bovine adrenal cor-
tex-derived capillary endothelial (ACCE) cells
or
human umbilical vein
endothelial (HUVE) cells were used for binding and mitogenic assays.
Stock plates of ACCE cells were maintained in the presence of low
glucose Dulbecco's modified Eagle's medium supplemented with 10%
calf serum, 2 mM glutamine (growth medium) plus bFGF at a final
concentration of 2 ng/ml (6, 17). For proliferation assays, ACCE cells
were seeded at 7,00O/well in 6-multiwell plates in the presence of
growth medium. HUVE cells were maintained in endothelial growth
medium (Clonetics) supplemented with
5%
fetal bovine serum and bo-
vine pituitary extract, according to the instruction of the manufacturer.
For proliferation assays,
HUVE
cells were seeded at 12,00O/well in
6-multiwell plates in the presence
of
endothelial growth medium con-
taining 2% serum, without pituitary extract. Heparin
(1
or 10 pg/ml)
was added when noted. After
5-8
days, cells were dissociated by trypsin,
and cell numbers were determined with a Coulter counter.
Binding to Endothelial Cells-ACCE or
HUVE
cells were grown
to
confluence in 6-well plates in the appropriate growth media. The media
were removed, and binding was carried out in low glucose Dulbecco's
modified Eagle's medium containing 10% fetal bovine serum, 0.2% gela-
tin, 10 &ml heparin, and 10 mM HEPES, pH 7.4 (binding medium) on
ice. One ml of binding medium containing
50,000
cpm of lZ5I-VEGF (or
'zsI-PIGF) and various competitive agents were added
to
each well.
When noted, amounts of radioligand were varied. Monolayers were
washed three times with binding medium and then extracted with
1
ml
of
1
N
NaOH
to
recover bound lZ5I-VEGF (or '261-P1GF). The entire
NaOH extract was counted in a
y
counter. Data were analyzed by the
method of Scatchard (38).
Cross-linking ofLigands-HUVE cells were grown
to
confluence in
10-cm dishes. All subsequent steps were performed at 4 "C
or
on ice.
Media were replaced with the binding medium (detailed above) contain-
ing 1261-VEGF (110
PM)
or '251-P1GF
(230
PM)
?
2 pgiml cold VEGF, as
indicated, for
2
h. Cells were washed twice with cold PBS and incubated
for 20 min at
4
"C with
0.5
mM BS3 cross-linking agent (Pierce) dissolved
in PBS. The reaction was quenched with one-tenth volume of 200 mM
glycine,
10
mM Tris-HC1, pH
8.
The dishes were washed twice with cold
PBS and extracted with 1.25 ml of lysis buffer containing
1%
Triton
X-100,
137 mM NaCl, 10% glycerol, 2
m~
EDTA,
1
mM vanadate,
1
m~
phenylmethylsulfonyl fluoride,
0.15
unit/ml aprotinin, 20
p~
leupeptin,
20 mM
Tris
HC1, pH 8.0 (39). The extracts were spun to remove debris,
and the supernatants were saved. The supernatants were split into
equal volumes and then immunoprecipitated with rabbit antisera to
Flt-1, Flk-l/KDR, or with control antisera at the final dilution of
150
for
1
h. Protein A-agarose was added and mixed for an additional hour at
4 "C. The beads were pelleted and washed twice with lysis buffer before
addition of double-strength Laemmli sample buffer to elute proteins
from the beads. The eluates were boiled for
1
min and electrophoresed
on 620% SDS-polyacrylamide gels. The gels were dried and exposed
to
a phosphor-imaging plate for
-36
h. The plate was read on a BAS-2000
image analyzer (Fuji, Japan).
nrosine Phosphorylation Assay-ACCE cells were grown to conflu-
ence in 60-mm dishes without added growth factors. Confluent cultures
were preincubated for 24 h in the presence of low glucose Dulbecco's
modified Eagle's medium supplemented with
0.5%
calf serum (deprived
medium). Media were then removed and cells were incubated for
5
min
at 37 "C in the presence of deprived medium containing 50 ng/ml
VEGF,,,
or
varying amounts of PlGF-2 (PIGF,,,). At the end of the
incubation, media were removed and cells were washed with cold PBS
containing
1
mM sodium vanadate and then extracted with 0.25 ml
of
lysis buffer (see above). This and all subsequent steps were performed
at 4 "C. The extracts were centrifuged briefly in a Microfuge
to
pellet
debris. The supernatants were mixed for
1
h with
50
1.11 of wheat germ
agglutinin agarose beads to collect membrane proteins. The beads were
pelleted and washed once with lysis buffer. Double-strength Laemmli
sample buffer was added to elute proteins, boiled for
1
min, and subjected
to electrophoresis on 4-20% SDS-polyacrylamide gels. The gels were
electroblotted to polyvinylidene difluoride membranes. Following incu-
bation with 4G10 anti-phosphotyrosine monoclonal antibody (UBI, Lake
Placid,
NY),
immunoreactive bands were visualized with an ABC kit,
according
to
the instructions of the manufacturer (Vector Laboratories).
Miles Vascular Permeability Assay-Induction
of
vascular permeabil-
ity was determined by the Miles assay essentially as described previ-
ously (8, 40). One ml of 0.5% (w/v) Evans blue was injected intracardi-
ally into anesthetized guinea pigs. One hour later, PBS alone, different
concentrations of VEGF or PlGF, individually or in combination, were
injected intradermally in 200-1.11 aliquots in the dorsal area. Leakage of
protein-bound dye was detected by
a
blue spot surrounding the injection
site.
RESULTS
Biochemical Characterization
of
PlGF-The two PlGF pro-
teins were purified to apparent homogeneity. Anion and cation
exchange chromatographies were used
as
initial purification
steps for PIGF,,l/PIGF-l and PlGF,,@GF-2, respectively. The
final purification step
for
both isoforms was reversed phase
chromatography on
a
C4
column. Analysis of the purified ma-
terials by silver-stained SDS-PAGE gel in reducing conditions
revealed major bands (Fig. 2)
at
-32 kDa (PlGF,,,, lane
4)
and
-28 kDa (PlGF,,,, lane
61,
respectively. In nonreducing condi-
tions, bands of approximately twice that size were evidenced
(PIGF,,,
(lune
1
)
and PlGF,,, (lane
3)).
This
is
consistent with
a
homodimeric structure. Microsequencing
of
the purified ma-
terial revealed in both cases
a
single NH,-terminal amino acid
sequence: LPAW. Therefore, the PlGF proteins are generated
following cleavage of an 18-amino acid signal sequence (31,321.
The mature PlGF monomers are expected
to
have, respectively,
131
and 152 amino acids. By analogy with the WGF polypep-
tides, they may be called PlGF,,, and PlGF,,,. Both PlGF
iso-
forms reveal higher apparent molecular weight than WGFlG5
(Fig. 2), in spite of the fact that their predicted sequence
is
shorter.
A
higher level
of
glycosylation probably accounts for
such differences. This
is
consistent with the presence
of
two
putative glycosylation sites in each PlGF monomer (31-33) uer-
sus
only one in VEGF
(6).
Furthermore, the finding
that
both
PlGF Binds to Flt-1 and Potentiates VEGF Activity
25649
PlGF isoforms are eluted
as
broad and asymmetric peaks by
reversed phase HPLC (data not shown)
is
consistent with heav-
ily glycosylated proteins. The PlGF proteins were then com-
pared for their ability to bind to heparin-Sepharose
or
S-Sepha-
rose.
As
shown in Fig. 3, P1GFl3, did not bind appreciably to
S-Sepharose
or
heparin-Sepharose. In contrast, PIGF,,, bound
strongly to both matrices and was eluted in the presence of
20.5
M
and
0.9
M
NaCl, respectively. The differential chromato-
graphic behavior of PlGF,,, and PlGF,,, is clearly reminiscent
of VEGF,,, and VEGF,,, (16) and is consistent with the primary
amino acid sequence of the PlGF proteins
(31-33).
PlGF Binds to Flt-1 IgG but Not to
KDR
ZgG-Flt-1 IgG
or
KDR IgG were
first
tested in competition binding assays using
I2,I-VEGF as the radioligand (Fig.
4,
A
and B). The IC,, for
VEGF,,, binding to
Flt-1
IgG was 22
-c
3
PM,
while that
for
VEGF,,, binding to KDR IgG was 125
2
20
PM.
In both cases,
123456
106
-
80
-
495-
M
I
32
5-
27
5-
18
5-
FIG.
2.
Silvepstained SDS-polyacrylamide gel
of
purified
PlGF
isoforms and
VEGF,,.
PIGF,,,, VEGF,,,, or PlGFIs2 (approximately
200
ng) were electrophoresed on a 10-20% polyacrylamide gradient gel
at
40
d.
Lunes
13
were under nonreducing conditions while
lanes
4-6
were under reducing conditions.
Lanes
1
and
4
contain PIGF,,,
lunes
2
and
5,
VEGF,.,, and
lanes
3
and
6,
PIGF,,,. Molecular sizes of
I
111"
prestained staiJards are in kilodaltons:
of
PlGF,,, (A
and
B)
or
PlGF,,,
(C
and
FIG.
3.
Chromatographic behavior
D)
on S-Sepharose
(A
and
C)
or
hep-
arin-Sepharose
(B
andD) columns.
In
A
and
C,
concentrated conditioned me-
dium from transfected cells expressing ei-
ther PlGF
isoform
was equilibrated in 25
mM sodium phosphate, pH
6.0.
The mate-
rial was loaded onto S-Sepharose columns
(1
ml), washed with starting buffer, and
eluted stepwise with increasing concen-
trations of NaCI, as indicated. In
B
and
D,
concentrated conditioned medium was
equilibrated in 10 mM
Tris
HCI, pH
7.2,
containing
50
mM NaCI. The material was
loaded onto heparin-Sepharose columns
(1
ml), washed with starting buffer, and
eluted with increasing concentrations
of
NaCI. PlGF was detected by inhibition of
'2sI-VEGF binding to Flt-1 IgG.
100
80
60
0
40
e
5
C
20
LL
Eo
?
100
s
80
60
'
40
(u
7
.-
.-
c
C
.-
20
0
*251-VEGF,,, bound to the soluble receptor with an affinity very
similar to that of the full-length receptor expressed in trans-
fected cells (23, 24, 28). Therefore, these soluble receptor-IgGs
provide suitable in vitro model systems to study each VEGF
receptor independently of the other.
In preliminary experiments, we determined that conditioned
medium derived from transfected CEN4 cells expressing either
PlGF,,,
or
PIGF,,, was able to compete for 1251-VEGF,,, binding
to
Flt-1
IgG. Such activity provided
a
convenient assay for the
purification of both PlGF species. Like the conditioned me-
dium, highly purified PlGF was able to compete with lZ5I-
VEGF,,, for binding to
Flt-1
IgG (Fig.
5A).
In contrast, PlGF
(up to
48
nM) was unable to displace 12,1-VEGF bound to KDR
IgG (Fig.
5A)
or
Flk-1 IgG (not shown). Both '251-P1GF,,2 and
1251-PIGF13, bound to Flt-1 IgG with high affinity (Fig.
5,
B-D).
The
ED,, values were 250
?
35
PM
and 186
2
40
PM,
respectively.
However, lZ5I-P1GF failed
to
bind to KDR IgG (Fig.
5E).
In
panels D and E, binding experiments were performed in the
presence of
1
pg/ml heparin, although experiments without
heparin
or
using 1251-P1GF,31, which fails to bind heparin, gave
similar results. Similar results were also obtained when we
tested the binding of lZ5I-P1GF on transfected COS cells ex-
pressing full-length Flt-1
or
KDR
receptors (data not shown).
Binding of VEGF
or
PlGF to soluble receptors, endothelial cells,
or
transfected cells was not affected by bFGF, hepatocyte
growth factor,
or
insulin (data not shown), indicating that the
binding was specific. Thus, Flt-1, but not Flk-l/KDR,
is
a po-
tential high affinity receptor for PlGF.
PlGF Binds to Endothelial Cells-Ligand binding to endo-
thelial cells was performed to determine whether binding sites
consistent with those observed in vitro were duplicated on en-
dothelial cells.
As
shown in Fig.
6A,
1251-VEGF,6, bound
to
ACCE and could be only partially displaced by PlGF, indicating
the presence of both PlGF-sensitive and -insensitive VEGF re-
ceptors on ACCE cells. Similar results were observed in HUVE
cells (data not shown). Since HUVE cells consistently displayed
A
B
DA
0
5
10
15
20
25
5
10
15
20
25
Fraction
25650 PlGF Binds to Flt-1 and Potentiates VEGF Activity
higher expression of PlGF-sensitive VEGF binding sites than
ACCE cells, we chose to characterize PlGF binding to
HUVE
cells in greater detail (Fig. 6B). Binding of 1251-PlGFl,, to
100
A
1
10
100 1000
100
B
"i
20",
n
-.
1
10
100
1000
10000
VEGF165
(pM)
FIG.
4.
Binding
of
VEGF
to
Flt-1 IgG
or
to
KDR
IgG.
Microtiter
plates were coated with goat anti-human Fc IgG overnight and blocked
VEGF,,, (-12,000 cpdwell) and either Flt-1
IgG
or
KDR
IgG
and
with 10% fetal bovine serum in PBS. Wells were incubated with
lZ5I-
various concentrations
of
cold VEGF as described under "Experimental
Procedures." Wells were washed and individually counted in a
y
counter
to
determine binding. Nonspecific binding (determined in the absence
of
receptor-IgG) was subtracted.
LL
W
i3
>
100
50
0
1
10
100
1000
10000
100000
PlGF
(pM)
$
oL
10
HUVE
cells demonstrated two classes of sites, having affinities
of
230
2
16
PM
and
>2
nM.
HWE
cells express approximately
15,000 high affinity and
>30,000
low affinity PlGF receptors
per cell. These results demonstrate that endothelial cells ex-
press displaceable cell surface PlGF receptors. The high affin-
ity PlGF binding site is consistent with
Flt-1,
while the identity
of the low af'finity site remains to be determined.
Cross-linking
of
1251-VEGF,,S or '251-P1GFl,, to
HUVE
cells,
followed by immunoprecipitation of these complexes with an-
tibodies to
Flt-1
or Flk-l/KDR, was performed in the attempt to
identify the receptors involved in the binding events. Immuno-
precipitation
of
'251-VEGF,,,-cross-linked
complexes with anti-
serum directed against
Flt-1
yielded a relatively faint band
at
-190,000 and lower molecular weight bands consistent with
un-cross-linked VEGF (Fig.
7,
lane
1).
The anti-Flt-1 antiserum
also immunoprecipitated 12,1-P1GF (Fig.
7,
lane
3).
However,
cross-linked 1251-P1GFl,2~Flt-1 complexes
at
-
190,000 could not
be clearly visualized. Lower molecular weight complexes con-
sistent with possible oligomers of PlGF were observed (Fig.
7,
lane
3). Immunoprecipitation of cross-linked 1251-VEGF1,, with
anti-Flk-1lKDR antiserum yielded
a
strong receptor-ligand
complex band
at
-210 kDa. Un-cross-linked VEGF was also
observed (Fig.
7,
lane
5).
In contrast, the anti-Flk-1/KDR anti-
serum failed to precipitate significant lZ5I-P1GF (Fig.
7,
lane
7).
All of the observed labeling events could be inhibited by an
excess of cold VEGF (Fig.
7,
lanes
2,4,
and
6).
Preimmune sera
failed to precipitate lZ5I-VEGF or lZ5I-P1GF (data not shown).
PlGF Fails to Induce Tyrosine Phosphorylation in Endothe-
lial Cells-To
determine whether PlGF might play
a
direct role
in intracellular signal transduction pathways, we tested VEGF
or PlGF for their ability
to
trigger tyrosine phosphorylation of
membrane proteins in ACCE cells. In agreement with previous
studies (41), VEGF,,, stimulated tyrosine phosphorylation of
an -200-kDa membrane protein. In contrast, P1GFl,2,
at
all
concentrations tested, failed
to
stimulate tyrosine phosphoryl-
ation (Fig.
8).
PlGF Potentiates VEGF Mitogenic Activity in Cultured En-
dothelial Cells-We
next examined the effects of PlGF alone, or
w
60-
0
1
10
100
1000
Radioligand
(pM)
1
1
10000
=
50
5
40
5
30
;?
20
-
a,
v
U
0
D
G
;
10
ln
-
'Io
1
10
100
1000
10000
100000
Competitor added
(pM)
1
10
100
1000
10000
Radioligand
(pM)
FIG.
5.
PlGF
binds
to Flt-1 IgG
but not
to
KDR
IgG.
In
AX,
wells were incubated with radioligand, as indicated (-12,000 cpdwell),
receptor-IgG, and various concentrations
of
cold VEGF
or
PlGF, as described under "Experimental Procedures."
Panel
A
shows that PlGF
(548
m)
can displace '261-VEGF,,, bound to Flt-1 IgG but not
to
KDR IgG. In
B,
'251-P1GFl,, binding
to
Flt-1 IgG is competed by native PlGF,,,. In
C,
'251-P1GF,,, binding
to
Flt-1 IgG is competed by cold PlGF,,,
or
by VEGF,,,. Flt-1 IgG shows concentration-dependent binding of both 1251-VEGF,,,
and '251-P1GF,,,
(panel D).
In contrast, KDR
IgG
binds only 1251-VEGF,6,
(panel
E).
PlGF Binds to Flt-1 and Potentiates VEGF Activity
25651
0
i
IO
100
1000
loooo
101
600,
h
-
400
\
u)
n
U
c
v
g
200
m
0
B
Bound
(ImoVWell)
/
1
100
1000
10
125l-PLGF152
(pM)
30
10
FIG.
6.
Identification
of
1261-VEGF,, binding sites displaceable
by PlGF on ACCE cells and concentration dependence
of
lz5I-
PlGF,,, binding
to
HUVE
cells.
In
A,
confluent ACCE cells were
incubated with 12sI-VEGF,65 and varying concentrations
of
cold PlGF,5p.
Binding conditions were as described under "Experimental Proce-
dures." In
B,
confluent
HUVE
cells were incubated with varying con-
centrations of 1251-P1GF15p. Data plotted by the method of Scatchard
(38)
are shown in the
inset.
Nonspecific binding was determined in each case
with
5
pg of VEGF,, and was always
515%
of the total binding.
in combination with VEGF, on cultured endothelial cells. Con-
ditioned media of transfected CEN4 cells
or
partially purified
PlGF isoforms had little
or
no mitogenic activity for ACCE
or
HUVE cells (data not shown). Likewise, highly purified
PlGF,,,, tested up
to
530
ng/ml, did not promote the growth of
ACCE cells
(Fig.
9A). PlGF,,, also failed to stimulate ACCE cell
growth, in the presence
or
in the absence of heparin (Fig.
9B).
HUVE cells showed
a
"20% growth stimulation
at
concen-
trations of PlGF
>
100
ng/ml. In contrast, rhVEGF,,,, at the
concentrations of
2.5
or
10
ng/ml, induced
a
3-
to
4-fold increase
in final cell count
(Fig.
9, A-D). Purified PlGF did not affect the
maximal mitogenic response (efficacy)
to
VEGF16, in ACCE
cells (Fig. 9C). However, when PlGF was added in the presence
of low, marginally efficacious concentrations of VEGF,,,, this
resulted in
a
dose-dependent potentiation of the mitogenic ac-
tivity of VEGF. In ACCE cells, addition of 190 ng/ml PlGF,,2 to
0.03
ng/ml VEGF resulted in
a
4040% increase in cell number
over that observed with VEGF alone. However, PIGF,,, failed
to
potentiate low doses of bFGF, demonstrating that the observed
potentiation
is
specific for VEGF. Fig. 9D illustrates
a
repre-
sentative experiment. A similar potentiation of VEGF bioactiv-
ity by PlGF was observed in
HUVE
cells (data not shown).
PlGF Does Not Directly Induce Vascular Permeability but
It
Dramatically Potentiates VEGF Action-We then tested PlGF
12
3
4 5
678
14.3
-
or
llbI-PIGF,,, bound to
HUVE
cells.
HUVE
cells were incubated in
FIG.
7.
Cross-linking and immunoprecipitation
of
1261-VEGF,,
medium containing 1""IVEGF,6,
(lanes
1,
2,
5,
and
6)
or 12sII-PIGF,,,
(lanes
3,4,
7,
and
8).
Lanes
2,4,6,
and
8
reflect the addition of
2
pg/ml
cold VEGF,,,. Cells were washed and incubated with
0.5
mM BE?' cross-
linking agent. After cross-linking, the reaction was quenched, and then
the dishes were washed and extracted. The extracts were equally di-
vided and immunoprecipitated with rabbit antisera
to
Flt-1
(lanes
14)
or
Flk-l/KDR
(lanes
5-8).
The immunoprecipitation products were sub-
jected to electrophoresis on
4-20%
SDS-polyacrylamide gels. The gels
were dried and exposed to
a
phosphor-imaging plate and read on BAS-
2000
image analyzer.
D
ng/d
PlGF
200
c
116.5
-
49.5
c
FIG.
8.
Effect
of
VEGF or PlGF on tyrosine phosphorylation
of
ACCE cell surface glycoproteins.
ACCE cells cultured in 6-cm
dishes were stimulated with VEGF,, or various concentrations of
PlGF,,,
for
5
min at
37
"C, as indicated. Cell extracts were prepared and
then precipitated with wheat germ agglutinin agarose, followed by elu-
tion with Laemmli sample buffer. The eluates were subjected
to
elec-
trophoresis, transferred to polyvinylidene difluoride membranes, and
immunoblotted with an anti-phosphotyrosine monoclonal antibody.
in an in vivo system, the Miles vascular permeability assay
(Fig. 10).
This
assay provides
a
rapid and reproducible measure
of
one of the
known
activities of VEGF and is suitable for the
screening
of
multiple samples in
a
single animal. Tested alone
up to
500
ng/site, PlGF,,, showed no activity (Fig. 10,
24,
while VEGF16, was able to induce Evans blue extravasation in
a
dose-dependent manner (Fig. 10,
8-12),
in agreement with
previous studies
(7,
8,
40). Once again, PlGF was able to po-
tentiate the activity of minimally effective doses of VEGF. In
the presence of
500
ng of PIGF,,,,
a
dose of VEGF,,, (10 ng) that
alone gave
a
barely detectable response induced
a
near-maxi-
mal increase in dye extravasation, comparable to that induced
by 250 ng of VEGF,,, (Fig. 10,
14-18).
This potentiation was
dose-dependent. However, PlGF did not increase the maximal
25652
PlGF
Binds
to Flt-1
and
Potentiates VEGF Activity
100
I
60
A
0 0
1
10
100
1000
PIGF,,, (ng/ml)
0
no addition
2.5
nglrnl
VEGF+PIGF
A
21
0
1
10
100
PIGF,,,
(ng/ml)
0
PIGF152
+10
pg/ml
heparin
PIGF152
no heparin
2
ng/ml
FGF
401
2ot
ok
I
I I
0
1
1
10
100
1000
PIGF,,, (ng/ml)
""
Dl
0
10
ng/ml
VEGF
10
0
1
10
100
1000
PIGF,,;! (ng/ml)
FIG.
9.
Effects
of
PlGF on endothelial cell growth.
ACCE
cells were seeded into 6-well plates at
7,000
cells per well in the presence
of
VEGF,
PlGF,
or
bFGF, as indicated. After
5-7
days, cells were trypsinized and cell counts were determined using a Coulter counter. In
A,
cells were
cultured in the presence
of
varying amounts
of
PIGFIB,. In
B,
varying amounts
of
PlGF,,, were added, in the presence
or
absence
of
10
pg/ml
heparin. In
C,
each well received
2.5
ng/ml VEGF,,, in the presence
or
in the absence
of
varying amounts
of
PlGF,,,. In
D,
each well received either
0.03
ng/ml VEGF,,,
or
0.015
ng/ml bFGF and varying amounts
of
PIGF,,,. VEGF,,, at
10
ng/ml and bFGF at
2
ng/ml served as positive controls.
response (efficacy) of VEGF nor could bFGF potentiate the
activity of VEGF in this assay (data not shown).
DISCUSSION
Angiogenesis
is
a
tightly regulated process, integral to nor-
mal and pathologic conditions, including wound healing, cycli-
cal growth of the corpus luteum, rheumatoid arthritis, and
growth and metastasis
of
solid tumors. Recent evidence impli-
cates VEGF as
a
major regulator of these events
(18-22, 42,
43).
In addition, VEGF has been proposed to play
a
role in the
regulation of microvascular permeability
(7,8).
Yet, the signals
which regulate or modulate the actions of VEGF on the vascu-
lar endothelium are largely unknown. The recently identified
PlGF may share some of the functions of the VEGF polypep-
tides since: (i)
it
belongs to the same gene family, (ii) displays
significant sequence homology to WGF, and (iii) the PlGF gene
undergoes alternative exon splicing events reminiscent of the
VEGF gene. Therefore, it
is
possible that VEGF and PlGF
interact with similar pathways. To better understand this po-
tential interaction, we chose to purify to homogeneity the two
PlGF isoforms and test whether they are able to bind to the
known WGF receptors. The construction of soluble chimeric
Flt-1, KDR, and Flk-1 receptors represented
a
very useful tool
as
it
allowed
us
to individually characterize the properties of
each receptor both
in vitro
and, potentially,
in vivo,
without
interference from the other receptor
or
other factors. The bind-
ing characteristics of such soluble receptors are very similar to
those of the full-length receptors, indicating that the extracel-
lular domain contains all the information required for high
affinity binding.
PlGF bound with high affinity to Flt-1 IgG. That Flt-1
is
truly expressed in HUVE cells and available to bind PlGF
is
demonstrated by the identification of a cross-linked receptor-
lz5I-VEGF band of the predicted size (-190 kDa) following im-
munoprecipitation with an anti-Flt-1 antiserum. Such anti-
serum also immunoprecipitated Flt-1-bound '251-P1GF. Our
inability to directly visualize PlGF-Flt-1 cross-linked com-
plexes
is
likely to reflect both the lower affinity of PlGF com-
pared to VEGF as well as a low cross-linking efficiency of both
ligands to Flt-1. That such a receptor may also be present in
ACCE cells is suggested by the finding that these cells express
high affinity VEGF binding sites consistent with
Flt-1
(44).
Also,
PlGF
is
able to compete about half of the lz5I-VEGF bind-
ing to such cells. Furthermore, recent
in
situ hybridization
studies have demonstrated the ubiquitous expression of Flt-1
mRNA in endothelial cells
(45).
However, a recent study
(44)
failed to detect expression
of
the Flt-1 gene in ACCE cells by
Northern blot analysis of total RNA using
a
heterologous cDNA
probe. The RNA blotting methods employed in that study may
be less sensitive
or
specific than the radioligand binding
or
in
situ
hybridization studies. Alternatively, it
is
possible that
ACCE cells express
a
novel Flt-1-like VEGFPIGF receptor.
Remarkably, the PlGF proteins did not demonstrate high
affinity binding to Flk-1 IgG
or
to
KDR
IgG.
In agreement with
these findings, an anti-Flk-l/KDR antiserum failed to precipi-
tate '251-P1GF bound to
HUVE
cells. InFontrast, the antiserum
PlGF Binds to Flt-1 and Potentiates
VEGF
Activity
25653
FIG.
10.
Potentiation
of
VEGF
action
by
PlGF
in
the
Miles
vas-
cular
permeability assay.
Evans Blue was injected intracardially into
guinea pigs. After
1
h, various doses
of
VEGF1,,, PlGF,,,
or
a combi-
nation
of
both, were administered intradermally in 0.2 ml
of
PBS. PBS
(sites
1,
7,
and
13)
was used as control.
Sites
2-6
reflect the response to
500-,
250-,
125-,
50-,
and 25-ng doses of PIGF,,,.
Sites
8-12
show the
response
to
250-,
125-,50-, 20-, and 10-ng doses
of
VEGF,6,.
Sites
14-18
show the effect
of
the administration
of
a constant dose
of
VEGF,,, (10
ng) in the presence
of
500-, 250-, 125-,
50-,
and 25-ng doses
of
PIGF,,,.
efficiently immunoprecipitated 1251-VEGF-receptor complexes.
Although we cannot rule out the possibility that PlGF has
a
very low affinity for Flk-l/KDR
(Kd
>
100 nM), such a low
affinity binding would unlikely be physiologically relevant. A
recent report suggests that heparin enhances cross-linking of
VEGF to Flk-1 receptors (46). We observed only a modest in-
crease in binding of VEGF
or
PlGF to soluble receptors in the
presence of heparin (data not shown). However, since all bind-
ing assays described here were performed in the presence of
serum (which may contain heparin sulfate proteoglycans), the
need for exogenous heparin may have been obviated.
Purified PlGF showed little
or
no mitogenic activity for
ACCE and HUVE cells. Likewise, PlGF failed to induce extrav-
asation in the Miles vascular permeability assay. These find-
ings suggest that binding to
Flt-1
(and/or other unidentified
receptor) is not sufficient to trigger an effective mitogenic re-
sponse
or
to induce vascular permeability. Unlike the Flk-1
receptor (28, 291, Flt-1 fails to demonstrate VEGF-dependent
tyrosine phosphorylation (23). PIGF, tested over
a
wide dose
range, did not induce tyrosine phosphorylation in ACCE cells.
Interaction with Flk-l/KDR may be a critical requirement to
induce the full spectrum of VEGF actions. In this context,
a
negative-dominant Flk-1 mutant has been recently shown to
suppress tumor angiogenesis
in vivo
(43).
Intriguingly, however, PlGF was able to potentiate the action
of low, marginally efficacious, concentrations of VEGF on en-
dothelial cell growth. At such concentrations, VEGF is expected
to preferentially occupy
its
higher affinity receptors
(i.e.
Flt-1).
However, PlGF did not affect the mitogenic response to VEGF
at concentrations where the latter would be expected
to
occupy
both Flt-1 and Flk-lKDR receptors. Potentiation of VEGF ac-
tion by PlGF was replicated
in vivo
in the Miles vascular per-
meability assay. In fact, the potentiation
of
such activity of
VEGF by PlGF was much more striking than the potentiation
of the mitogenic action. Such an effect required approximately
a 10-20-fold molar excess of PlGF over VEGF. Interestingly,
this
is
similar to the difference in the relative affinities of
VEGF and PlGF for Flt-1. As mentioned above, recent
in situ
hybridization studies have documented the widespread expres-
sion of the Flt-1 mRNA in endothelial cells in normal adult
tissues, including the skin (45). This suggests’that interaction
of PlGF with Flt-1 may take place
in
vivo.
It will be of signifi-
cant interest to determine whether PlGF is also able
to
enhance
major
in vivo
actions of VEGF such
as
the ability to promote
angiogenesis in tumors (22)
or
in ischemic limbs (47).
One possible explanation for the effects that we report here
is that Flt-1 (or other unidentified high affinity VEGFPIGF
receptor) behaves as
a
“decoy,” having little
or
no transducing
activity alone. Therefore, binding to this receptor would limit
the bioactivity of VEGF by preventing its binding to a signal
transducing receptor. According to this hypothesis, PlGF would
act
to
release VEGF from Flt-1 and increase its availability
to
the more relevant Flk-VKDR. Such
a
mechanism would be
similar to that recently described for IL-1 receptors type I and
type I1 (48). Type I receptor is responsible for mediating IL-1
biological activities, whereas type I1 receptor has been shown
to
play an inhibitory role by acting as a “decoy” target for IL-1.
Regulation of Flt-1 and/or PlGF expression would provide
a
means to adjust the sensitivity of the endothelium to VEGF. In
this context, recent studies have shown that the PlGF mRNAis
expressed in HUVE cells (33). Therefore, PlGF may modulate
VEGF action by an autocrine mechanism.
Alternatively, the formation of heterodimers between KDR
and Flt-1 might confer new properties
or
ligand specificities
upon these receptors. Both possibilities are currently being
examined. In addition, we attempted to ascertain whether KDR
(or
other receptor) could acquire high affinity binding for PlGF
following stimulation with VEGF.
So
far, we have found no
evidence for such
a
mechanism in HUVE
or
ACCE cells. Fur-
thermore, low concentrations of VEGF spiked into binding as-
says did not reveal cryptic, VEGF-dependent, binding sites for
1251-P1GF,,2 in KDR IgG (data not shown).
The lack of pronounced effects
of
PlGF on endothelial cell
growth and vascular permeability suggest that this protein
may not be responsible for the direct induction of angiogenesis/
permeability and thus differs from VEGF. In this respect, it is
interesting to point out that high PlGF expression has been
reported in trophoblastic hydatiform mole and choriocarcino-
mas (321, conditions characterized by paucity or absence of blood
vessels (49). Rather, PlGF may function to enhance the activity
ofVEGF in situations where the concentrations of the latter are
limiting. Interestingly, the PlGF mRNAis expressed in the pla-
centa at substantially higher levels than the VEGF mRNA.’
Therefore, the molar excess of PlGF
versus
VEGF required
to
elicit the effects that we describe here may occur
in vivo.
The PlGF isoforms share several biochemical properties with
the VEGF polypeptides. PIGF,,, lacks the highly basic domain
characteristic of PIGF,,,. Accordingly, PlGF,,, fails to bind to
heparin
or
to
cation-exchange resins and is
a
diffusible protein.
This behavior is very similar to that displayed by VEGF,,, (16).
Although the 21-amino acid insertion confers strong heparin-
J.
E.
Park, H. H. Chen,
J.
Winer,
K.
A.
Houck, and
N.
Ferrara,
unpublished observations.
25654
PlGF Binds to Flt-1 and Potentiates VEGF Activity
binding properties on PIGF,,,, this protein differs from
VEGF18, or VEGF,,, since
it
is
substantially diffusible. In this
respect, the behavior
of
PIGF,,, is reminiscent of that of
VEGF,,, (16), a diffusible protein that nevertheless displays
significant binding
to
heparin-containing proteoglycans. Fur-
ther studies are required to assess whether PIGF,,, binds to the
extracellular matrix or whether proteolysis may play
a
role in
regulating PlGF bioavailability, by analogy with the VEGF
polypeptides (16, 17).
It has been observed that PlGF sequences are present in the
genome
of
a variety of species including
Drosophila melano-
gaster,
suggesting important or even essential functions
(32).
The availability of highly purified recombinant PlGF should
facilitate addressing such questions and may also permit the
identification of novel receptors which could in turn shed fur-
ther light on the significance of this protein. Furthermore, by
virtue of its selective interaction with VEGF receptors, PlGF
may constitute an important molecular tool
to
analyze the dif-
ferential role of such receptors in mediating the various biolog-
ical actions of this family
of
proteins and to dissect critical
structural domains involved in receptor binding.
Acknowledgments-We thank William J. Henzel
for
protein microse-
quencing, Allan Padua
for
amino acid analysis, Brian Fendly
for
anti-
bodies, Eileen Soriano-Szatkowski
for
performing the Miles vascular
permeability assay, and Louis Tamayo for graphics. We are grateful
to
Joffre Baker
for
helpful discussions and advice.
REFERENCES
1.
Ramsey, E. M., and Donner, M. W.
(1980)
Placental Vasculature and Circula-
2.
Gilbert, S.
F.
(1988)
Developmental Biology, 2nd Ed, Sinauer, Sunderland,
MA
3.
Bassett, D. L.
(1943)Am.
J.
Anat.
73, 251
4.
Klagshrun, M., and DAmore,
P.
A.
(1991)
Annu.
Reu.
Physiol.
35,217-239
5.
Folkman, J., and Klagsbrun, M.
(1987)
Science
235,442447
6.
Leung,
D.
W., Cachianes, G., Kuang, W. J., Goeddel, D. V., and Ferrara, N.
(1989)
Science
246, 1306-1309
7.
Keck,
P.
J.,
Hauser,
S.
D.,
Krivi,
G.,
Sanzo, K., Warren, T., Warren, T., Feder,
J., and Connolly, D. T.
(1989)
Science
246, 1309-1312
8.
Connolly, D.
T.,
Heuvelman, D. M., Nelson, R., Olander,
J.
V., Eppley,
B.,
Inuest. 84,
1470-1478
Delfino,
J.
J., Siegel, N. R., Leimgruber,
R.
S.,
and Feder,
J.
(1989)
J.
Clin.
tion, Georg Thieme, Stuttgart
9.
Basilico, C., and Moscatelli, D.
(1992)
Ado. Cancer
Res.
59, 115-165
10.
Yang, E. Y., and Moses, H. L.
(1990)
J.
Cell
Biol.
111,731-741
11.
Massague,
J.
(1990)
J.
Biol.
Chem.
265,21393-21396
12.
Klagsbrun, M., and DAmore,
PA
(1991)
Annu.
Reu.
Physiol.
53, 217-239
13.
Amento, E. P., and Beck, L.
S.
(1991)
Ciba Found. Symp.
157,115-123
14.
Leibovich,
S.
J., Polverini,
P.
J., Shepard, H. M., Wiseman,
D.
M., Shively, V.,
15.
Ferrara, N., Houck, K.
A,,
Jakeman, L., and Leung, D. W.
(1992)
Endocr.
Reu.
16.
Houck,
K.
A,,
Leung, D. W., Rowland,
A.
M., Winer, J., and Ferrara, N.
(1992)
17.
Park, J. E., Keller, G. A., and Ferrara, N.
(1993)
Mol.
Bid.
Cell
4, 1317-1326
18.
Ravindrath, N., Little-Ihrig, L., Phillips, H.
S.,
Ferrara, N., and Zeleznick, A.
and Nuseir, N.
(1987)
Nature
329,630432
13, 18-32
J.
Bid.
Chem.
267,26031-26037
J.
(1992)
Endocrinology
94,
1192-1199
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
Phillips, H.
S.,
Hains, J.,
Leung,
D. W., and Ferrara, N.
(1990)
Endocrinology
127,965-967
Jakeman, L. B., Armanini, M., Phillips, H.
S.,
and Ferrara,
N.
(1993)
Endo-
crinology
133, 848459
Breier, G., Albrecht, U., Sterrer,
S.,
and Risau,
W.
(1992)
Development
114,
521-532
Kim, K. J., Li, B., Winer, J., Armanini, M., Gillett, N., Phillips, H.
S.,
and
Ferrara,
N.
(1993)
Nature
362,8414344
deVries, C., Escobedo,
J.
A,, Ueno, H., Houck, K., Ferrara, E, and Williams, L.
T.
(1992)
Science
256,989-991
Terman,
B.
I.,
Dougher-Vermazen, M., Carrion, M. E., Dimitrov, D., Armellino,
D. C., Gospodarowicz, D., and Bohlen,
P.
(1992)
Biochem. Biophys. Res.
Shibuya, M., Yamaguchi,
S.,
Yamane,
A,,
Ikeda, T.,
Tojo,
A,, Matsushime, H.,
Commun.
187,1579-1586
Terman,
B.
I.,
Carrion, M. E., Kovacs, E., Rasmussen, B. A., Eddy, R. L., and
and Sato, M.
(1990)
Oncogene
5,519-524
Matthews, W., Jordan, C. T., Gavin, M., Jenkins, N.
A,,
Copeland, N. G., and
Shows,
T.
B.
(1991)
Oncogene
6,1677-1683
Quinn, T. P., Peters, K. G., De Vries, C., Ferrara, N., and Williams, L.
T.
(1993)
Lemischka,
I.
R.
(1991)
Proc. Natl. Acad.
Sci.
U.
S.
A.
88,
902G9030
Millauer, B., Wizigmann-Voos,
S.,
Schnurch, H., Martinez,
R.,
Moller,
N.
P. H.,
Proc. Natl. Acad.
Sci.
U.
S.
A.
90, 7533-7537
Kendell, R. L., and Thomas, K.
A.
(1993)
Proc. Natl. Acad.
Sei.
U.
S.
A.
90,
Risau, W., and Ullrich,
A.
(1993)
Cell
72, 8354346
10705-10709
Maglione, D., Guerriero, V., Vigliette, G., Delli-Bovi, P., and Persico, M. G.
Maglione, D., Guerriero, V., Viglietto, G., Ferrara, M. G., Aprelikova,
O.,
Ali-
(1991)
Proc. Natl. Acad. Sci.
U.
S.
A.
88,9267-9271
talo,
K.,
Del Vecchio,
S.,
Lei, K. J., Chou,
J.
Y., and Persico, M. G.
(1993)
Oncogene
8,
925-931
Hauser,
S.,
and Weich, H. A.
(1993)
Growth Factors
9, 259-268
Ferrara,
N.,
Leung, D. W., Cachianes, G., Winer, J., and Henzel, W.
J.
(1991)
Jakeman,
L.
B., Winer,
J.,
Bennett, G. L., Altar, C. A., and Ferrara, N.
(1992)
Methods Enzymol.
198, 391405
Su,
W., Middleton, T., Sudgen,
B.,
and Echols, H.
(1991)
Proc. Natl. Acad. Sei.
J.
Clin. Inuest.
89,
244-253
Capon, D. J., Chamow,
S.
M., Mordenti,
J.,
Marsters,
S.
A.,
Gregory, T., Mit-
U.
S.
A.
88,
10870-10874
and Smith, D.
H.
(1989)
Nature
337,525-531
suya, H., Bym, R. A,, Lucas, C., Wurm,
F.
M., Groopman,
J.
E., Broder,
S.,
Scatchard, G.
(1949)
Ann. N.
I:
Acad. Sci.
61,660472
Fantl, W. J., Escobedo,
J.
A,,
Martin,
G.
A,,
Turck, C. W., del Rosario, M.,
Kim, K. J., Li, B., Houck, K., Winer, J., and Ferrara, N.
(1992)
Growth Factors
Myoken, Y., Kayada, Y., Okamoto,
T.,
Kan, M., Sato, G. H., and Sato,
J.
D.
Brown, L. F., Yeo, K. T., Berse,
B.,
Yeo,
T.
K., Senger, D. R., Dvorak, H. F., and
(1991)
Proc. Natl. Acad. Sci. U.
S.
A. 88,
5819-5823
Millauer, B., Shawver, L. K., Plate, K. H., Risau, W., and Ullrich, A.
(1994)
van de Water, L.
(1992)
J.
Exp. Med.
176, 1375-1379
Barleone, B., Hauser,
S.,
Schollmann,
C.,
Weindel, K., Marme, D., Yayon,
A.,
Nature
367,57&579
Peters, K. G., De Vries,
C.,
and Williams, L.
T.
(1993)
Proc. Natl. Acad. Sci.
and Weich, H.
A.
(1994)
J.
Cell.
Biochem.
54,5646
Tessler,
S.,
Rockwell, P., Hicklin, D., Cohen, T., Levi, B.
Z.,
Witte, L.,
U.
S.
A.
90,89154919
Takeshita,
S.,
Zheng, L. P., Brogi, E., Kearney,
M.,
Pu, L.-Q., Bunting,
S.,
Lemischka,
I.
R., and Neufeld, G.
(1994)
J.
Biol.
Chem.
269, 1245612461
Colotta, F., Re, F., Muzio, M., Bertini, R., Polentarutti, N., Sironi, M., Gin,
J.
Ferrara, N., Symes,
J.
F., and Isner,
J.
M.
(1994)
J.
Clin.
Inuest.
93,662470
G., Dowker,
S.
K., Sirns,
J.
E.,
and Mantovani,
A.
(1993)
Science
261,
Kraus,
F.
T. (1990jinAnderson’s Pathology (Kissane,
J.
M., ed) pp.
1706-1710,
472475
The C. V. Mosby Co., St. Louis, MO
McCormick, F., and Williams, L. T.
(1992)
Cell
69, 413423
7,5344
... KDR might be the most important receptor which is involved in VEGF-induced angiogenesis [70][71][72]. FLT1 can regulate the VEGF activity via interacting with VEGF and making it less available to KDR [73]. Decreased expression of VEGF, FLT1 and KDR might be related to follicle atresia in chickens [49]. ...
Preprint
Full-text available
Background: Changshun green-shell laying hen with strong broodiness is a Chinese indigenous chicken breed. Little is known about the mechanisms responsible for ovary development of Changshun green-shell laying hens from egg-laying period (LP) to incubation period (BP). Here, RNA sequencing (RNA-seq) of ovaries from Changshun hens in LP and BP was performed to identify candidate genes and pathways associated with broodiness. Results: We identified 1650 differently expressed genes (DEGs), including 429 up-regulated and 1221 down-regulated DEGs, in chicken ovaries between LP and BP groups. GO and KEGG analysis further revealed that these DEGs were mainly involved in the pathways related to follicle development in chicken ovaries, including focal adhesion, MAPK signaling pathway, and FoxO signaling pathway, and vascular smooth muscle contraction, ECM-receptor interaction, and GnRH signaling pathway were down-regulated in incubating ovaries. Eight candidate genes (EGFR, VEGFRKDRL, FLT1, KDR, PDGFRA, TEK, KIT and FGFR3) related to angiogenesis, folliculogenesis, steroidogenesis and oogenesis in ovaries were suggested to play important roles in the ovarian development of Changshun hens during the transition from LP to BP. Conclusions: We discovered critical genes and pathways which is closely associated with ovary development in incubating chickens, indicating the complexity of reproductive behaviour of different chicken breeds.
... As such, while VEGF-C/D and their receptor VEGFR3 are primarily involved in lymphangiogenesis, VEGF-A activates VEGFR1 and VEGFR2 and primarily exerts regulatory effects on the cell migration of macrophage and endothelial cell lineages, vasculogenesis, and vascular permeability-phenomena primarily observed in iMCD-TAFRO. Although VEGF-A binds to VEGFR1 with high affinity (10 pM), the induction of VEGFR1 phosphorylation is weak, and VEGFR1 is not primarily thought to be a receptor transmitting a mitogenic signal, but, rather, a 'decoy' receptor [27]; however, a motif in the juxtamembrane region of VEGFR1, but not VEGFR2, represses PI3K activation and EC migration [28]. As such, VEGFR2 is thought to act as the primary signaling VEGFR in blood vascular endothelial cells. ...
Article
Full-text available
TAFRO (thrombocytopenia (T), anasarca (A), fever (F), reticulin fibrosis (F/R), renal failure (R), and organomegaly (O)) is a heterogeneous clinical subtype of idiopathic multicentric Castleman disease (iMCD) associated with a significantly poorer prognosis than other subtypes of iMCD. TAFRO symptomatology can also be seen in pathological contexts outside of iMCD, but it is unclear if those cases should be considered representative of a different disease entity or simply a severe presentation of other infectious, malignant, and rheumatological diseases. While interleukin-6 (IL-6) is an established driver of iMCD-TAFRO pathogenesis in a subset of patients, the etiology is unknown. Recent case reports and literature reviews on TAFRO patients suggest that vascular endothelial growth factor (VEGF), and the interplay of VEGF and IL-6 in concert, rather than IL-6 as a single cytokine, may be drivers for iMCD-TAFRO pathophysiology, especially renal injury. In this review, we discuss the possible role of VEGF in the pathophysiology and clinical manifestations of iMCD-TAFRO. In particular, VEGF may be involved in iMCD-TAFRO pathology through its ability to activate RAS/RAF/MEK/ERK and PI3K/AKT/mTOR signaling pathways. Further elucidating a role for the VEGF-IL-6 axis and additional disease drivers may shed light on therapeutic options for the treatment of TAFRO patients who do not respond to, or otherwise relapse following, treatment with IL-6 targeting drugs. This review investigates the potential role of VEGF in the pathophysiology of iMCD-TAFRO and the potential for targeting related signaling pathways in the future.
... sFLT1 is able to bind and block the vascular endothelial growth factor (VEGF) and the placental growth factor (PGF), two proangiogenic proteins that are found in lower levels in this disease. The cause of this diminution is the increased presence of sFLT1, since it acts as an antagonist, preventing the proangiogenic signaling induced by these two molecules [20]. ...
Article
Full-text available
H2S, a gasotransmitter that can be produced both via the transsulfuration pathway and non-enzymatically, plays a key role in vasodilation and angiogenesis during pregnancy. In fact, the involvement of H2S production on plasma levels of sFLT1, PGF, and other molecules related to preeclampsia has been demonstrated. Interestingly, we have found that maternal fructose intake (a common component of the Western diet) affects tissular H2S production. However, its consumption is allowed during pregnancy. Thus, (1) to study whether maternal fructose intake affects placental production of H2S in the offspring, when pregnant; and (2) to study if fructose consumption during pregnancy can increase the risk of preeclampsia, pregnant rats from fructose-fed mothers (10% w/v) subjected (FF) or not (FC) to a fructose supplementation were studied and compared to pregnant control rats (CC). Placental gene expression, H2S production, plasma sFLT1, and PGF were determined. Descendants of fructose-fed mothers (FC) presented an increase in H2S production. However, if they consumed fructose during their own gestation (FF), this effect was reversed so that the increase disappeared. Curiously, placental synthesis of H2S was mainly non-enzymatic. Related to this, placental expression of Cys dioxygenase, an enzyme involved in Cys catabolism (a molecule required for non-enzymatic H2S synthesis), was significantly decreased in FC rats. Related to preeclampsia, gene expression of sFLT1 (a molecule with antiangiogenic properties) was augmented in both FF and FC dams, although these differences were not reflected in their plasma levels. Furthermore, placental expression of PGF (a molecule with angiogenic properties) was decreased in both FC and FF dams, becoming significantly diminished in plasma of FC versus control dams. Both fructose consumption and maternal fructose intake induce changes in molecules that contribute to increasing the risk of preeclampsia, and these effects are not always mediated by changes in H2S production.
... There are trophoblastic and non-trophoblastic sources of PLGF, but placental trophoblast is the main source of PLGF throughout the gestational period of pregnancy [8]. It shares structural as well as amino acid sequence similarity with VEGF, but PLGF has binding a nity only for VEGF receptor1(VEGFR-1) [9]. The inter-and intramolecular cross-talk between the VEGFR-1 and VEGFR-2 is regulated by PLGF. ...
Preprint
Full-text available
The hypoxia-inducible factors (HIF) are considered master regulators of oxygen homeostasis and are oxygen level sensitive. Currently, there is no information regarding the expression of HIF-1α in pregnant women with COVID-19 infection and its potential involvement in the placentation of this condition. The current study aims to detect the expression of HIF-1α and possible molecular link between the expression of HIF-1α and PLGF. A case-control study was conducted in a tertiary university hospital between January and September 2022. Placental tissue of post COVID-19 infected pregnant women (n = 34) and healthy control (n = 16) were collected and processed for gene expression of HIF-1α and PLGF by Quantitative real-time PCR. There was statistically significant higher median level of HIF-1α among cases compared to controls (P < 0.001). Moreover, there was a statistically significant higher median level of PLGF among cases compared to controls (P < 0.001). There was statistically significant moderate positive correlation between HIF-1α and PLGF gene (r = 0.636, P < 0.001) among studied samples, however no statistically significant correlation between the two genes among cases and controls groups separately. In conclusion, the higher levels of HIF-1α and PLGF in the placentas of post COVID-infected pregnant women compared to normal pregnant women indicate that COVID-19 hypoxia did not affect the process of placentation. This is confirmed also by the positive correlation between HIF-1α and PIGF in placental tissues among total studied samples.
... A study by Gilbert et al supports this study which found VEGFR-1 distribution were not increased in diabetes.36 Last possible explanation is that VEGF has the ability to increase PlGF regulation.37 Last finding in our study is the contradictory ratio between vitreous and plasma VEGF and PlGF. ...
Article
Full-text available
Purpose: To compare plasma and vitreous level of vascular endothelial growth factor (VEGF) and placental growth factor (PlGF) in diabetic rats with poor blood glucose (BG) control, reconstitution of good BG control, and nondiabetic rats, and to investigate the effect of reconstitution of good BG control to VEGF and PlGF plasma and vitreous level. Methods: This is an experimental study using eighteen Sprague Dawley rats which were divided into intervention group (n=14) and control group (n=4). Intervention group were given Streptozotocin (STZ) injection to induce diabetes. After 4 weeks, intervention group was randomly divided into group I for termination and group II for reconstitution of good BG control with insulin for the following 4 weeks, and so was the control group. Plasma and vitreous samples were taken. VEGF and PlGF levels were evaluated with enzyme-linked immunosorbent assay (ELISA). Results: Seventeen of 18 rats survived in intervention group. BG level of intervention group II decreased dramatically to normoglycemia. ELISA at month 1 showed that VEGF vitreous level tend to be higher in intervention group I compared to control I, 196.36 ± 65.24 pg/dL and 123.64 ± 44.99, respectively (p=0.20). ELISA at month 2 showed that PlGF vitreous level of intervention group I were significantly higher compared to control I, 59.04 ± 2.48 and 51.93 ± 3.15, respectively (p=0.01). Vitreous and plasma VEGF of intervention group I and II were not different, while vitreous and plasma PlGF were significantly higher in group II. Conclusions: Vitreous levels of VEGF and PlGF were increased in diabetic rats compared to nondiabetic, and reconstitution of good BG control for 1 month were unable to reduce VEGF and PlGF levels.
... VEGFR-1 is also present in uterine smooth muscle cells [70], and in monocytes, cells that may be responsible for pathological angiogenesis by expressing VEGF-A and other cytokines and growth factors [71]. Evidence suggests that some VEGF-A receptors, particularly VEGFR-1 and NRP-1, are expressed by some tumor cells, suggesting an autocrine loop that may stimulate tumor cell growth and migration [72]. Usually, the binding of VEGF-A triggers a dimerization of VEGFR-2 and activates crucial signaling pathways, including the phospholipase Cγ (PLCγ)-extracellular signal-regulated kinase 1/2 (ERK1/2) pathway with a role in vascular development and adult arteriogenesis; phosphatidyl-inositol 3kinase (PI3K)-protein kinase B (PKB/AKT)-mechanistic target of rapamycin (mTOR) pathway involved in cell survival, regulation of vasomotion and of barrier function; and SRC kinases (SRC proto-oncogene, nonreceptor tyrosine kinase) and small G-proteins (guanine nucleotide-binding proteins) that control cell migration, polarization, and endothelial junctions [56]. ...
Article
Full-text available
The complement component C1q plays a role as a pro-angiogenic factor in different contexts, acting in a complement-independent way. For example, this molecule is able to foster the remodeling of the spiral arteries for a physiological pregnancy and to promote the wound healing process. It is also involved in angiogenesis after post-stroke ischemia. Furthermore, it has a role in supporting the tumor vessel growth. Given its role in promoting angiogenesis both under physiological and pathological situations, other studies are needed to understand its potential therapeutic implications.
Article
Placental insufficiency is one of the major causes of fetal growth restriction (FGR), a significant pregnancy disorder in which the fetus fails to achieve its full growth potential in utero. As well as the acute consequences of being born too small, affected offspring are at increased risk of cardiovascular disease, diabetes and other chronic diseases in later life. The placenta and heart develop concurrently, therefore placental maldevelopment and function in FGR may have profound effect on the growth and differentiation of many organ systems, including the heart. Hence, understanding the key molecular players that are synergistically linked in the development of the placenta and heart is critical. This review highlights the key growth factors, angiogenic molecules and transcription factors that are common causes of defective placental and cardiovascular development.
Article
Full-text available
Tumor biomarkers, the substances which are produced by tumors or the body’s responses to tumors during tumorigenesis and progression, have been demonstrated to possess critical and encouraging value in screening and early diagnosis, prognosis prediction, recurrence detection, and therapeutic efficacy monitoring of cancers. Over the past decades, continuous progress has been made in exploring and discovering novel, sensitive, specific, and accurate tumor biomarkers, which has significantly promoted personalized medicine and improved the outcomes of cancer patients, especially advances in molecular biology technologies developed for the detection of tumor biomarkers. Herein, we summarize the discovery and development of tumor biomarkers, including the history of tumor biomarkers, the conventional and innovative technologies used for biomarker discovery and detection, the classification of tumor biomarkers based on tissue origins, and the application of tumor biomarkers in clinical cancer management. In particular, we highlight the recent advancements in biomarker-based anticancer-targeted therapies which are emerging as breakthroughs and promising cancer therapeutic strategies. We also discuss limitations and challenges that need to be addressed and provide insights and perspectives to turn challenges into opportunities in this field. Collectively, the discovery and application of multiple tumor biomarkers emphasized in this review may provide guidance on improved precision medicine, broaden horizons in future research directions, and expedite the clinical classification of cancer patients according to their molecular biomarkers rather than organs of origin.
Article
Islet transplantation holds significant promise as a curative approach for type 1 diabetes (T1D). However, the transition of islet transplantation from the experimental phase to widespread clinical implementation has not occurred yet. One major hurdle in this field is the challenge of insufficient vascularization and subsequent early loss of transplanted islets, especially in non-intraportal transplantation sites. The establishment of a fully functional vascular system following transplantation is crucial for the survival and secretion function of islet grafts. This vascular network not only ensures the delivery of oxygen and nutrients, but also plays a critical role in insulin release and the timely removal of metabolic waste from the grafts. This review summarizes recent advances in effective strategies to improve graft revascularization and enhance islet survival. These advancements include the local release and regulation of angiogenic factors (e.g., vascular endothelial growth factor, VEGF), co-transplantation of vascular fragments, and pre-vascularization of the graft site. These innovative approaches pave the way for the development of effective islet transplantation therapies for individuals with T1D.
Article
Full-text available
The vascular endothelial growth factor (VEGF) family encompasses four polypeptides that result from alternative splicing of mRNA. We have previously demonstrated differences in the secretion pattern of these polypeptides. Stable cell lines expressing VEGFs were established in human embryonic kidney CEN4 cells. VEGF121, the shortest form, was secreted and freely soluble in tissue culture medium. VEG189 was secreted, but was almost entirely bound to the cell surface or extracellular matrix. VEGF165 displayed an intermediary behavior. Suramin induced the release of VEGF189, permitting its characterization as a more basic protein with higher affinity for heparin than VEGF165 or VEGF121, but with similar endothelial cell mitogenic activity. Heparin, heparan sulfate, and heparinase all induced the release of VEGF165 and VEGF165 suggesting heparin-containing proteoglycans as candidate VEGF-binding sites. Finally, VEGF165 and VEGF189 were released from their bound states by treatment with plasmin. The released 34-kDa dimeric species are active as endothelial cell mitogens and as vascular permeability agents. We conclude that the bioavailability of VEGF may be regulated at the genetic level by alternative splicing that determines whether VEGF will be soluble or incorporated into a biological reservoir and also through proteolysis following plasminogen activation.
Article
Full-text available
The fms-like tyrosine kinase (Flt) is a transmembrane receptor in the tyrosine kinase family. Expression of flt complementary DNA in COS cells conferred specific, high-affinity binding of vascular endothelial growth factor, also known as vascular permeability factor (VEGF-VPF), a factor that induces vascular permeability when injected in the guinea pig skin and stimulates endothelial cell proliferation. Expression of Flt in Xenopus laevis oocytes caused the oocytes to release calcium in response to VEGF-VPF. These findings show that flt encodes a receptor for VEGF-VPF.
Article
Full-text available
Application of TGF beta 1 (10-100 ng) to the chicken chorioallantoic membrane (CAM) for 72 h resulted in a dose-dependent, gross angiogenic response. The vascular effects induced by TGF beta 1 were qualitatively different than those induced by maximal doses of basic FGF (bFGF) (500 ng). While TGF beta 1 induced the formation of large blood vessels by 72 h, bFGF induced primarily small blood vessels. Histologic analysis revealed that TGF beta 1 stimulated pleiotropic cellular responses in the CAM. Increases in fibroblast and epithelial cell density in the area of TGF beta 1 delivery were observed as early as 4 h after TGF beta 1 treatment. By 8 h, these cell types also demonstrated altered morphology and marked inhibition of proliferation as evidenced by 3H-thymidine labeling. Thus, the TGF beta 1-stimulated accumulation of these cell types was the result of cellular chemotaxis from peripheral areas into the area of TGF beta 1 delivery. Microscopic angiogenesis in the form of capillary sprouts and increased endothelial cell density first became evident at 16 h. By 24 h, capillary cords appeared within the mesenchyme of the CAM, extending towards the point of TGF beta 1 delivery. 3H-thymidine labeling revealed that the growth of these capillary cords was due to endothelial cell proliferation. Finally, perivascular mononuclear inflammation did not become evident until 48 h of treatment, and its presence correlated spatially and temporally with the gross and histological remodelling of newly formed capillary cords into larger blood vessels. In summary, these data suggest that, in the chicken CAM, TGF beta 1 initiates a sequence of cellular responses that results in growth inhibition, cellular accumulation through migration, and microvascular angiogenesis.
Article
Persistent microvascular hyperpermeability to plasma proteins even after the cessation of injury is a characteristic but poorly understood feature of normal wound healing. It results in extravasation of fibrinogen that clots to form fibrin, which serves as a provisional matrix and promotes angiogenesis and scar formation. We present evidence indicating that vascular permeability factor (VPF; also known as vascular endothelial growth factor) may be responsible for the hyperpermeable state, as well as the angiogenesis, that are characteristic of healing wounds. Hyperpermeable blood vessels were identified in healing split-thickness guinea pig and rat punch biopsy skin wounds by their capacity to extravasate circulating macromolecular tracers (colloidal carbon, fluoresceinated dextran). Vascular permeability was maximal at 2-3 d, but persisted as late as 7 d after wounding. Leaky vessels were found initially at the wound edges and later in the subepidermal granulation tissue as keratinocytes migrated to cover the denuded wound surface. Angiogenesis was also prominent within this 7-d interval. In situ hybridization revealed that greatly increased amounts of VPF mRNA were expressed by keratinocytes, initially those at the wound edge, and, at later intervals, keratinocytes that migrated to cover the wound surface; occasional mononuclear cells also expressed VPF mRNA. Secreted VPF was detected by immunofluoroassay of medium from cultured human keratinocytes. These data identify keratinocytes as an important source of VPF gene transcript and protein, correlate VPF expression with persistent vascular hyperpermeability and angiogenesis, and suggest that VPF is an important cytokine in wound healing.
Article
The receptor for platelet-derived growth factor (PDGF) binds two proteins containing SH2 domains, GTPase activating protein (GAP) and phosphatidylinositol 3-kinase (Pl3-kinase). The sites on the receptor that mediate this interaction were identified by using phosphotyrosine-containing peptides representing receptor sequences to block specifically binding of either Pl3-kinase or GAP. These results suggested that Pl3-kinase binds two phosphotyrosine residues, each located in a 5 aa motif with an essential methionine at the fourth position C-terminal to the tyrosine. Point mutations at these sites caused a selective elimination of Pl3-kinase binding and loss of PDGF-stimulated DNA synthesis. Mutation of the binding site for GAP prevented the receptor from associating with or phosphorylating GAP, but had no effect on Pl3-kinase binding and little effect on DNA synthesis. Therefore, GAP and Pl3-kinase interact with the receptor by binding to different phosphotyrosine-containing sequence motifs.
Article
I. Introduction THE establishment of a vascular supply is a critical requirement for cellular inflow of nutrients, outflow of waste products, and gas exchange in most tissues and organs (1). In endocrine glands, the vascularization not only serves such needs but also provides a pathway for the specific secretory products (2, 3). Furthermore, in the anterior pituitary (4–6) and in the adrenal medulla (7, 8), an unusual angioarchitecture, where a portal capillary plexus delivers venous blood originating from an adjacent gland, is intimately involved in the control of the secretory activity. In the adrenal medulla, this vascular design may even determine the ultimate secretory product (8). Not surprisingly, the cardiovascular system is the first organ system to develop and reach a functional state in an embryo (9–12). In the human, primitive blood vessels appear as early as day 15, and a circulation with a beating heart is already established by the end of the third week.
Article
Angiogenesis plays critical roles in organ development during embryonic and fetal life, wound healing and in a variety of pathological conditions. Vascular endothelial growth factor (VEGF) is a secreted growth factor specific for vascular endothelial cells which induces angiogenesis in vivo. To gain a better understanding of the physiological role of VEGF, we have generated and characterized four murine monoclonal antibodies (mAbs) using the 165 amino acid species of recombinant human VEGF as immunogen. These mAbs (A3.13.1, A4.6.1, B4.3.1 and B2.6.2) belong to IgG1 isotype and have high affinities for VEGF (dissociation constants range from 2.2 x 10(-9) to 4 x 10(-10) M). Two different epitopes were detected with these mAbs. One epitope is recognized by mAbs A3.13.1 and B2.6.2, and the other recognized by mAbs A4.6.1 and B4.3.1. The epitope recognized by mAb A4.6.1 appears to be continuous while mAb B2.6.2 recognizes a discontinuous epitope. MAb A4.6.1 recognized three species of VEGF generated by alternative splicing, VEGF121, VEGF165 and VEGF189 while mAb B2.6.2 binds only VEGF165 and VEGF189. Results using an in vitro bovine adrenal cortex endothelial cell proliferation assay, in in vivo vascular permeability assay and an in vivo embryonic chicken angiogenesis assay showed that mAb A4.6.1 has potent VEGF neutralizing activities. MAb A4.6.1 was shown to block the binding of VEGF to its receptor(s) suggesting the inhibitory mechanism for VEGF activities. These well-defined mAbs should be very powerful tools to understand the structure-function relationship of various domains of VEGF and may have therapeutic potential.
Article
Publisher Summary The family of fibroblast growth factors (FGF) is the largest family of growth factors involved in soft–tissue growth and regeneration. The FGF family includes seven members that share a varying degree of homology at the protein level and appear to have a similar broad mitogenic spectrum. They are angiogenic and promote the proliferation of a variety of cells of mesodermal and neuroectodermal origin . Three members of the family–namely, K–FGF/HST, FGF–5, INT–2 are identified originally as oncogenes, while two additions, FGF–6 and keratinocyte growth factor (KGF), are isolated by sequence homology or factor purification and cloning. FGF was identified as an activity in pituitary extracts that stimulated the proliferation of 3T3 cells in mice. The two prototypes of basic and acidic protein structure, the FGF genes, and the expressions of FGF are also discussed. FGF consist of three exons, separated by two introns of variable length and FGF genes map on several chromosomes. The molecular regulation of FGF expression, FGF receptors, and their interaction with extracellular matrix is described. The oncogenic potential of FGF members and their involvement in tumors is also discussed. All possible mechanisms operating on the expression of FGFs and their receptors: transcriptional controls, posttranscriptional regulation involving alternative splicing, alternative translation starts resulting in proteins with different properties, and control affecting the secretion of these proteins are discussed.
Article
Vascular endothelial cell growth factor (VEGF), also known as vascular permeability factor, is an endothelial cell mitogen which stimulates angiogenesis. Here we report that a previously identified receptor tyrosine kinase gene, KDR, encodes a receptor for VEGF. Expression of KDR in CMT-3 (cells which do not contain receptors for VEGF) allows for saturable 125I-VEGF binding with high affinity (KD = 75 pM). Affinity cross-linking of 125I-VEGF to KDR-transfected CMT-3 cells results in specific labeling of two proteins of M(r) = 195 and 235 kDa. The KDR receptor tyrosine kinase shares structural similarities with a recently reported receptor for VEGF, flt, in a manner reminiscent of the similarities between the alpha and beta forms of the PDGF receptors.
Article
Epstein-Barr nuclear antigen 1 (EBNA-1) is the only viral protein required to support replication of Epstein-Barr virus during the latent phase of its life cycle. The DNA segment required for latent replication, oriP, contains two essential binding regions for EBNA-1, termed FR and DS, that are separated by 1 kilobase pair. The FR site appears to function as a replicational enhancer providing for the start of replication at the DS site. We have used electron microscopy to visualize the interaction of EBNA-1 with its binding sites and to study the mechanism for communication between the FR and DS sites. We have found that DNA-bound EBNA-1 forms a DNA loop between the FR and DS sites. From these results, we suggest that EBNA-1 bound to the replicational enhancer acts by a DNA-looping mechanism to facilitate the initiation of DNA replication. Occupancy of the DS site alone is highly sensitive to competition with nonspecific DNA. In contrast, occupancy of the DS site by looping from FR is largely resistant to the competitor DNA. These experiments support the concept that enhancers act in cis from nearby sites to provide a high local concentration of regulatory proteins at their target sites and to stabilize regulatory interactions.