Cardiovasc. Res. 6311-21 (2004).doc

advertisement
Control of vascular cell proliferation and migration by cyclin-dependent kinase
signalling: new perspectives and therapeutic potential
Vicente Andrés
Laboratory of Vascular Biology, Department of Molecular and Cellular Pathology and
Therapy, Instituto de Biomedicina de Valencia-CSIC, 46010 Valencia, Spain
WORD COUNT: 8021
CORRESPONDENCE TO:
Vicente Andrés
Instituto de Biomedicina de Valencia
C/Jaime Roig 11, 46010 Valencia (Spain)
Tel: +34-96-3391752
FAX: +34-96-3391750
E-mail: vandres@ibv.csic.es
1
ABSTRACT
Neointimal lesion development is a chronic inflammatory process that involves excessive
cell proliferation and migration within the artery wall. Progression through the mammalian cell
cycle requires the sequential activation of holoenzymes composed of a catalytic cyclindependent protein kinase (CDK) and a regulatory subunit named cyclin. Members of the CDK
family of inhibitory proteins (CKIs) interact with and inhibit the activity of CDKs. Cell
migration occurs predominantly at the G1/S phase of the cell cycle, and both CDKs and CKIs are
among the molecular machines that co-ordinately regulate the cycling events that control cell
proliferation and locomotion. The purpose of this review is to discuss the role of CDK/cyclin
holoenzymes and CKIs in the regulation of vascular cell proliferation and migration and in the
control of neointimal thickening. Pharmacological and gene therapy strategies targeting these
cell cycle regulators for the treatment of cardiovascular disease will be also discussed.
2
INTRODUCTION
The initiation and growth of atherosclerotic lesions is a complex multifactorial process that
involves adaptative and innate immune mechanisms [1-4]. Endothelial cell (EC) dysfunction
induced by atherogenic stimuli is one of the earliest manifestations of atherosclerosis at sites of
predisposition to atheroma formation (Fig. 1). The damaged endothelium promotes the adhesion
and transendothelial migration of circulating leukocytes. Early fatty streaks contain mostly highly
proliferative macrophages that avidly uptake lipoproteins to become lipid-laden foam cells.
Activated intimal leukocytes produce a plethora of inflammatory chemokines and cytokines that
promote the proliferation of vascular smooth muscle cell (VSMC) and their migration towards the
atherosclerotic lesion, thus further contributing to atheroma growth [1,2,5,6]. Excessive cell
proliferation and migration are also involved in the growth of vascular obstructive lesions during
restenosis post-angioplasty, transplant vasculopathy, and graft atherosclerosis. Plaque rupture or
erosion at advanced disease stages can lead to acute occlusion due to thrombus formation, resulting
in myocardial infarction or stroke.
While several proliferation markers are expressed in human primary atheroma and restenotic
lesions [7-16], the relevance of proliferation during human atherosclerosis and restenosis has been
controversial, with some studies reporting very low proliferative rates [8,9,11,13,15] and others
reporting abundance of dividing cells [10,17]. Cell proliferation appeared more pronounced in
restenotic versus primary lesions [11,17,18], and primary VSMCs obtained from human advanced
primary stenosing displayed diminished proliferation compared with cells from fresh restenosing
lesions [19], suggesting that cell proliferation is maximal at early stages of neointimal lesion
growth. Consistent with this interpretation, atheroma size and cellular proliferation within the
atheromatous plaque of hyperlipidemic rabbits are inversely correlated [20-22], and experimental
angioplasty is characterized by the reestablishment of the quiescent phenotype after the initial
proliferative burst [23,24].
3
The mammalian cell cycle is controlled by holoenzymes composed of a catalytic cyclindependent protein kinase (CDK) and a regulatory subunit named cyclin [25,26]. Diffferent
CDK/cyclin complexes are orderly activated at specific phases of the cell cycle (Fig. 2).
CDK/cyclin-dependent hyperphosphorylation of the retinoblastoma protein (pRb) and the related
pocket proteins p107 and p130 from mid G1 to mitosis contributes to the transactivation of genes
with functional E2F-binding sites, including several growth and cell-cycle regulators (i.e., c-myc,
pRb, cdc2, cyclin E, cyclin A), and genes encoding proteins required for nucleotide and DNA
biosynthesis (i. e., DNA polymerase , histone H2A, proliferating cell nuclear antigen, thymidine
kinase) [27-30]. The identities of substrates of the yeast CDK1 (CDC2) have revealed that this
enzyme employs a global regulatory strategy involving phosphorylation of other regulatory
molecules as well as phosphorylation of the molecular machines that drive cell-cycle events [31].
Cyclin availability and phospo/dephosphorylation of CDKs and cyclins by specific kinases
and phosphatases regulate the activity of CDK/cyclin holoenzymes. Of central importance in cell
cycle regulation, CDK activity is attenuated by the interaction with CDK inhibitory proteins
(CKIs) of the Cip/Kip (for CDK interacting protein/Kinase inhibitory protein) and Ink4 (for
inhibitor of CDK4) families [32] (Fig. 2). Cip/Kip proteins (p21Cip1, p27Kip1, p57Kip2) bind to and
inhibit a wide spectrum of CDK/cyclin holoenzymes, while Ink4 proteins (p16 Ink4a, p15Ink4b,
p18Ink4c, p19Ink4d) are specific for cyclin D-associated CDKs. Mitogenic and antimitogenic stimuli
affect the rates of CKI synthesis and degradation, as well as their redistribution among different
CDK/cyclin heterodimers.
Control of vascular cell proliferation and neointimal lesion growth by CDKs and CKIs
VSMC proliferation in the balloon-injured rat carotid artery is associated with a temporally
and spatially coordinated expression of CDKs and cyclins [16,33], and augmented expression of
these proteins is associated with an increase in their kinase activity [16,34]. CDK and cyclin
4
expression has been also detected in human VSMCs within atherosclerotic and restenotic tissue
[10,16,35]. Collectively, these findings suggest that the assembly of functional CDK/cyclin
holoenzymes in the injured arterial wall is a hallmark of vascular proliferative disease.
p27Kip1 and p21Cip1 have been implicated in the mechanism of action of several
pharmacological agents that control vascular cell proliferation in vitro and neointimal thickening.
Treatment of VSMCs with salicylate prevented PDGF-induced downregulation of p27Kip1 and
p21Cip1 but not of p16Ink4a, and this was accompanied by reduced CDK2 activity and growth
arrest [36]. Likewise, beraprost sodium-dependent VSMC growth arrest and reduction of intimal
thickening in the balloon-injured dog coronary artery correlated with maintained p27Kip1
expression [37]. Upregulation of p27Kip1 and p21Cip1 may be one mechanism by which nonsteroidal anti-inflammatory drugs [38], nitric oxide donors [39] and gene transfer of endothelial
nitric oxide synthase induce VSMC growth arrest [40]. Similarly, induction of p21Cip1 was
associated with tranilast-dependent inhibition of CDK2 and CDK4 activities, VSMC growth
arrest in vitro and reduced intimal hyperplasia in the rat balloon-injured carotid artery [41].
Prevention of Rho GTPase-induced downregulation of p27Kip1 without changes in p21Cip1,
p16Ink4a, or p53 levels may mediate simvastatin-dependent inhibition of CDK2 activity and
VSMC proliferation [42]. Otterbein et al. have recently shown that exposure of VSMC cultures
to carbon monoxide (CO) transiently increases p21Cip1 expression and results in growth arrest
[43]. Notably, CO suppressed arteriosclerotic lesions associated with both chronic graft rejection
and with balloon injury in rats. However, although p21Cip1 was essential for CO-dependent
VSMC growth arrest in vitro, the therapeutic effect of CO in a mouse model of mechanical
arterial injury was not impaired in p21Cip1-null mice [43].
Hemodynamic forces are thought to play an important role in the initiation and
progression of atherosclerotic lesions [1,2]. Steady laminar stress induced EC growth arrest and
this correlated with p21Cip1 upregulation without changes in p27Kip1 protein levels [44]. On the
5
other hand, stretch-mediated inactivation of forkhead transcription factors and p27Kip1
downregulation in VSMCs was accompanied by activation of CDK2, pRb hyperphosphorylation
and proliferation, demonstrating that the earliest cell cycle events in VSMCs can occur in a
solely mechanosensitive fashion [45]. RhoA-dependent reduction of p27Kip1 expression,
mediated in part via phosphatidylinositol-3-kinase, induces VSMC proliferation and may
contribute to the enhanced vascular responsiveness associated to hypertension [46].
Cell cycle progression in the artery wall is regulated by specific components of the
extracellular matrix (ECM) and integrins [47]. Neointimal VSMCs synthesize novel ECM
components and induce the expression of matrix-degrading proteases that remodel the
surrounding ECM. Notably, matrix-degrading metalloproteinase expression is induced within
neointimal lesions [48-51], and metalloproteinase inhibitors repressed VSMC proliferation in
vitro and after angioplasty [52-54]. Significant changes in collagen content occur during
neointimal lesion development [55-57]. Because polymerized collagen may mimic the scenario
of a normal artery composed of quiescent VSMCs, and monomer collagen might resemble the
ECM surrounding proliferating VSMCs within atherosclerotic and restenotic plaques, Koyama et
al. studied the growth properties of VSMCs cultured on monomer collagen fibers and on
polymerized collagen [58]. Mitogen-stimulated VSMCs grown on monomer collagen disclosed
high proliferative activity, but underwent G1 arrest when seeded on polymerized collagen. This
inhibitory effect of polymerized collagen appeared to be mediated by 2 integrins, and correlated
with suppression of p70S6K and upregulation of p27Kip1 (and to a lesser extent p21Cip1). Thus,
regulation of CKIs in response to changes in specific ECM components might regulate the ability
of VSMCs to respond to growth signals in vitro. Interestingly, the quiescent phenotype of
nonadherent NRK fibroblasts correlated with an increased association of p27Kip1 and p21Cip1 to
cyclin E-containing holoenzymes [59]. Further studies are required to determine whether
6
changes in arterial CKI expression regulate cell proliferation in response to integrins and ECM
components in vivo.
The gradual increase in p21Cip1 and p27Kip1 observed in balloon-injured rat and porcine
arteries suggests that these factors may limit neointimal hyperplasia after the initial proliferative
wave [14,60,61]. Using a mouse model of transluminal femoral artery injury, Reis et al. showed
a rapid apoptotic response and downregulation of p27Kip1 in medial VSMCs, which was followed
by a gradual increase in cell proliferation that peaked at 2 weeks in both the media and neointima
and decreased thereafter [62]. Restoration of low proliferative activity during later phases of
vascular repair in this model correlated with increased p27Kip1 expression. Several studies have
suggested a role for CKIs in the regulation of cell proliferation during human neointimal
thickening: a) reduced p27Kip1expression was detected in primary atherosclerotic lesions
compared with that in aorta, internal mammary artery, and carotid artery thrombendarterectomy
specimens, and in lesions of in-stent restenosis patients [63]. These authors also found a
significant upregulation of p21Cip1 in estenosis compared with primary lesions and other vascular
regions; b) more frequent expression of p27Kip1 and p21Cip1 was found within regions of human
coronary atheromas not undergoing proliferation [14]; c) concordant expression of TGF-
receptors I and II in virtually all cells positive for p27Kip1 within human atherosclerotic plaques
suggests that the anti-mitogenic action of TGF-1 in these lesions may be mediated by p27Kip1
[35]; d) coexpression of p53 and p21Cip1 in human carotid atheromatous plaque cells that
revealed lack of proliferation markers suggests that induction of p21Cip1 may occur via p53dependent transcriptional activation [64]; e) attenuation of PDFGF-BB-induced p21Cip1 and
p27Kip1 expression by interleukin-1 may promote VSMC hyperplastic growth after vascular
injury and in atherosclerosis [65].
Collectively, the above studies suggest an important role of p21Cip1 and p27Kip1 in
neointimal lesion growth. We have established a causal link between decreased p27Kip1 protein
7
expression and atherogenesis in hypercholesterolemic apolipoprotein E (apoE)-null mice by
demonstrating that whole-body genetic inactivation of p27Kip1 increases arterial VSMC and
macrophage proliferation and accelerates atherosclerosis [66]. In another study, we have shown
that selective inactivation of p27Kip1 in hematopoietic progenitor cells increases neointimal
macrophage proliferation and accelerates atherosclerosis in fat-fed apoE-deficient mice [67],
consistent with previous studies demonstrating enhanced haematopoietic progenitor cell
proliferation upon p27Kip1 inactivation [68], and implicating p27Kip1 as a critical macrophage
growth suppressor [69,70]. Because macrophages were the most abundant neointimal cells in our
study [67], it seems reasonable to suggest that macrophage p27Kip1 safeguards against the
inflammatory/proliferative response induced by dietary cholesterol in apoE-null mice. Regarding
the consequences of CKI inactivation on neointimal thickening induced by mechanical injury,
Otterbein et al. reported 3 times more pronounced lesion size in p21Cip1-null versus wild-type
mice [43]. In contrast, neointimal hyperplasia after mechanical vascular damage was similar in
wild-type and p27Kip1-null mice [71]. Redundant roles between p21Cip1 and p27Kip1, or a
compensatory increase in p21Cip1 expression (or other CKIs) might account for the lack of
phenotype of p27Kip1-null mice in the setting of mechanical arterial injury.
Animal and human studies have recognized significant differences in the atherogenicity of
different segments of the arterial, which may be related to regional phenotypic variance of
VSMCs, both when comparing cells from different compartments of the same vessel or cells
isolated from vessels from different vascular beds [72-78]. Sustained p27Kip1 expression in spite
of growth stimuli may contribute to the resistance to growth of human VSMCs isolated from
internal mammary artery compared with saphenous vein VSMCs, and to the longer patency of
arterial versus venous grafts [77]. Likewise, distinct p15Ink4b and p27Kip1 expression correlated
with different proliferative potential of intimal and medial VSMCs stimulated with basic
fibroblast growth factor [78], and intrinsic regional differences in the proliferative and migratory
8
capacity of VSMCs due to distinct regulation of p27Kip1 may contribute to creating variability in
atherogenicity in different vascular beds [79].
CDK inhibitory approaches to reduce neointimal thickening
The importance of CDK activation for neointimal lesion growth has been demonstrated by
means of pharmacological (Table 1) and gene therapy (Table 2) CDK inhibitory strategies. CVT313 is a purine derivative that inhibits CDK activity by preventing the binding of ATP to the
adenine-binding pocket of CDKs [80-82]. The relative inhibitory potency of CVT-313 varies
from very high for CDK2, moderate for CDK1, and very low for CDK4 [83]. Flavopiridol (L868275) is a more potent CDK inhibitor that displays higher specificity towards CDK4 than
towards CDK1 and CDK2 [84,85]. Growth arrest in VSMCs treated with flavopiridol and CVT313 correlates with decreased pRb protein levels and/or inhibition of its phosphorylation [83,86].
It is noteworthy that CVT-313 [83] and flavopiridol [85,87] can cause blockade in different cell
cycle phases depending on both the concentration of the drug and the cell line analyzed. In the rat
carotid model of balloon angioplasty, a brief intraluminal exposure of CVT-313 [83] or oral
administration of flavopiridol for 5 days beginning at the day of injury [86] reduced neointima
formation by 80% and 39%, respectively.
Gene therapy approaches based on the use of antisense oligodeoxynucleotide (ODN)
targeting CDKs and cyclins have shown efficacy in reducing neointimal lesion formation in
animal models of balloon angioplasty, including ODN against CDK2 [34,88], CDC2 [34,89,90],
and cyclin B1 [90]. Likewise, downregulation of cyclin G1 expression by retrovirus-mediated
antisense gene transfer inhibited VSMC proliferation and neointima formation after balloon
angioplasty [91]. Antisense ODN to CDC2/PCNA [92] and CDK2 [93] also attenuated graft
atherosclerosis. Additional approaches based on the inactivation of positive cell cycle regulators
that do not directly target CDK/cyclin activity (i. e., E2F, c-myc, etc) have been reviewed
9
elsewhere [6,94].
Consistent with the notion that CKIs function as negative regulators of neointimal
thickening (see above), gene transfer of p21Cip1 [95-97], p27Kip1 [60,98], and p57Kip2 [99] reduced
neointima formation after angioplasty in normocholesterolemic animals. Likewise, p21Cip1
overexpression attenuated neointimal thickening after balloon injury in hypercholesterolemic
mice [100] and following vein grafting [101]. On the other hand, antisense ODN to p21Cip1
attenuated matrix protein secretion in VSMCs [102]. Lamphere et al. generated chimeric p16Ink4a
and p27Kip1 molecules, which were of comparable potency to the parental p27Kip1 in inhibiting the
activities of several CDKs in vitro [103]. Among these chimeras, W9 was the most potent
growth suppressor of human coronary artery VSMCs and ECs when compared to the parental
p16Ink4a and p27Kip1, p27Kip1 derivatives, or several alternative p27Kip1-p16Ink4a chimeras.
Moreover, W9 was more effective in inhibiting neointimal thickening after balloon angioplasty
in cholesterol-fed rabbits [104]. Thus, combining the activities of different CKIs might increase
the therapeutical activity in the treatment of neointimal thickening after angioplasty. Further
approaches based on the overexpression of growth suppressor that do not directly target
CDK/cyclin activity (i. e., pRb, p53, Gax, etc) are reviewed elsewhere [6,94].
Control of cell migration by CDKs and CKIs
Several cytostatic agents (eg, quercetin, mimosine, suramin, rapamycin, and troglitazone)
can reduce the migratory potential of VSMCs and tumor cells [105-110]. Likewise, 17-estradiol
and the transcription factors p53, AP-1 and c-myc regulate in a coordinated manner the
proliferative and migratory potential of ECs and VSMCs [111-114]. NBT-II rat bladder
carcinoma cells synchronized in G1 migrated simultaneously upon FGF-1 stimulation, and cells
arrested in G2/M did not respond to stimulation by this mitogen [115]. Moreover, maximal
migration of PDGF-BB-stimulated VSMCs occurred in late G1 [116]. These studies indicate that
10
the position in the cell cycle is a key determinant of a cell’s competence for migration. Of note in
this regard, diverse cytoskeletal reorganization genes and many genes involved in cell motility
and remodeling of the extracellular matrix exhibited cell-cycle dependent regulation in human
fibroblasts [117].
Overexpression of p27Kip1, p16INK4a and p21Cip1 inhibits cell spreading and migration in
human umbilical vein ECs, CS-1 3 melanoma cells, VSMCs and NIH-3T3 fibroblasts [79,118120]. Moreover, p27Kip1-null VSMCs were more resistant than wild-type cells to the
antimigratory properties of rapamycin [121]. Regarding the role of CDKs on cell locomotion,
CDK6 localized to the ruffling edge of spreading cells and suppressed p16INK4a-mediated
inhibition of cell spreading [119]. Likewise, CDK5 activity has been involved in the regulation
of specific components of neuronal migration at different developmental stages [122], as well as
in the modulation of actin cytoskeleton dynamics in cells [123]. Moreover, CDK5 plays a key
role in regulating morphology, cell adhesion, and apoptosis in the human astrocytoma cell line
U373 [124]
In view of the above connections between cell proliferation and migration, we
investigated whether the dual function of p27Kip1 as a cell-cycle and migration inhibitor is
achieved via common or independent molecular pathways [125]. We found that physiologically
high level of p27Kip1 expression inhibits CDK activity and attenuates both proliferation and
migration in VSMC and fibroblast cultures. Mutations that rendered p27Kip1 unable to abrogate
CDK activity also prevented p27Kip1-induced growth arrest and migration blockade. We also
showed that a constitutively active mutant of pRb insensitive to CDK-dependent
hyperphosphorylation inhibited both cell proliferation and migration. In contrast, pRb
inactivation by forced expression of the adenoviral oncogene E1A correlated with high
proliferative and migratory activity. Collectively, these results suggest that cellular proliferation
and migration are regulated in a coordinated manner by the p27 Kip1/CDK/pRb/E2F pathway (Fig.
11
3). Consistent with this notion, E2F-1-null keratinocytes exhibited delay in transit through both
G1 and S phases of the cell cycle and substantially impaired migration [126]. Future studies are
necessary to identify E2F-regulated genes implicated in cell locomotion.
Concluding remarks
Excessive cell proliferation and migration contribute to neointimal thickening. It has been
well established that CDKs, cyclins and CKIs are key regulators of these processes in vitro.
Moreover, changes in the expression and/or activity of these cell cycle regulators have been
documented in several animal models of vascular proliferative disease and in human
atherosclerotic and restenotic tissue. Importantly, synthetic CDK inhibitors (CVT-313,
flavopiridol), antisense ODN to CDK/cyclins, and CKI overexpression reduced neointimal
hyperplasia in the setting of experimental graft atherosclerosis and angioplasty. Although these
CDK inhibitory strategies have not been assessed in clinic, antiproliferative approaches that have
shown promising results for preventing human neointimal thickening are available [6,94]. The
bacterial macrolide rapamycin (sirolimus, rapamune) is the pharmacological agent with which
most experience has been gathered so far for the prevention of in-stent restenosis. Rapamycin is
a potent immunosuppressant that strongly inhibits VSMC proliferation and migration
[105,106,121,127] via both p27Kip1–dependent [121,127] and p27Kip1–independent [33,71]
mechanisms. Rapamycin potently inhibited neointimal thickening in animal models of
angioplasty, graft atherosclerosis, and diet-induced atherosclerosis [71,128-136], and recent
clinical trials using rapamycin-impregnated stents have shown promising results for the
prevention of neointimal proliferation, restenosis, and associated clinical events in patients
undergoing coronary angioplasty [137-139]. Activation of E2F is triggered by CDK-dependent
phosphorylation of pRb and pocket proteins, therefore blockade of E2F function may be a
common mechanism by which different CDK/cyclin inhibitory strategies reduce neointimal
12
thickening. E2F inactivation via transfection of synthetic ODN containing an E2F consensus
binding site reduced experimental hyperplasia after balloon angioplasty and vein grafting
[140,141], and application of this E2F ‘decoy’ strategy is safe and can achieve sequence-specific
inhibition of cell-cycle gene expression and DNA replication in patients receiving bypass vein
grafts [142,143]. Despite these encouraging results, significant effort in basic research is
warranted to identify additional target genes and strategies for the treatment of cardiovascular
disease.
ACKNOWLEDGEMENTS
I apologize to colleagues whose work has not been directly cited due to space limitations. I
thank María J. Andrés-Manzano for preparing the figures. Work in my laboratory is currently
supported by grants from the Spanish Ministry of Science and Technology and Fondo Europeo de
Desarrollo Regional (SAF2001-2358, SAF2002-1443), and from Instituto de Salud Carlos III
(Red de Centros C03/01).
13
Figure 1: Atheroma development is a multifactorial process. Endothelial dysfunction induced
by different atherogenic stimuli triggers a chronic inflammatory response within the artery wall
that results in excessive proliferation and migration of leukocytes and VSMCs. At advanced
disease states, plaque rupture or erosion can lead to thrombus formation and associated ischemic
events.
Figure 2: Cell cycle control in mammalian cells. Activation of specific CDK/cyclin complexes
drives progression through the cell cycle (CDK1=CDC2). CKIs interact with and inactivate
CDK/cyclin holoenzymes.
Figure 3: Coordinate control of cell proliferation and migration. In the presence of low level
of p27Kip1 protein, active CDK/cyclin holoenzymes trigger the hyperphosphorylation of pRb,
release of E2F and high proliferative and migratory activity. In contrast, CDK/cyclin inactivation
by high level of p27Kip1 leads to the accumulation of hypophosphorylated pRb, sequestration of
E2F and low proliferative and migratory activity [79,120,125]. High p21Cip1 protein level also
induces growth arrest and migration blockade [118].
14
Table 1. Pharmacological CDK inhibitors that limit neointimal hyperplasia in the rat
carotid artery model of balloon angioplasty
CDK inhibitor
IC50 for CDKs
CDK1 (CDC2)= 4 M
CVT-313
(Purine derivative) CDK2=0.5 M
Dose and route of administration
1.25 mg/kg, brief intraluminal
exposure immediately after
angioplasty
Ref.
[83]
CDK4=215 M
CDK1 (CDC2)=0.5 M
Flavopiridol
CDK2=0.1 M
(Flavonoid)
CDK4=0.065 M
5 mg/kg/day, oral administration
CDK6=0.06 M
CDK7=0.11-0.3 M
15
[84-86,144]
Table 2. Gene therapy strategies targeting CDK/cyclin activity with beneficial effects in
animal models of cardiovascular disease
Strategy
Antisense-mediated
CDK/cyclin inactivation
Targeted gene
Strategy
Animal model
CDK2
ODN
Balloon angioplasty (rat)
CDK2
ODN
Graft atherosclerosis (mouse) [93]
CDC2
ODN
Balloon angioplasty (rat)
[34,89,90]
cyclin B1
ODN
Balloon angioplasty (rat)
[90]
CDC2/PCNA
ODN
Graft atherosclerosis
(rabbit, rat)
[92]
Balloon angioplasty (rat)
[91]
cyclin G1
Retrovirus
p21Cip1
Adenovirus Balloon angioplasty
(rat, mouse, pig)
p21Cip1
Plasmid
Ref.
[34,88]
[95-98,100]
Graft atherosclerosis (rabbit) [101]
CKI overexpression
p27Kip1
Adenovirus Balloon angioplasty
(rat, pig)
p57Kip2
Adenovirus Balloon angioplasty (rabbit)
[60,98]
[99]
p27Kip1-p16Ink4a Adenovirus Balloon angioplasty (rabbit) [104]
chimera
16
REFERENCES
[1].
Ross R. Atherosclerosis: an inflammatory disease. N Engl J Med 1999;340:115-126.
[2].
Lusis AJ. Atherosclerosis. Nature 2000;407:233-241.
[3].
Binder CJ, Chang MK, Shaw PX, Miller YI, Hartvigsen K, Dewan A, Witztum JL. Innate
and acquired immunity in atherogenesis. Nat Med 2002;8:1218-1226.
[4].
Greaves DR, Channon KM. Inflammation and immune responses in atherosclerosis.
Trends Immunol 2002;23:535-541.
[5].
Rivard A, Andrés V. Vascular smooth muscle cell proliferation in the pathogenesis of
atherosclerotic cardiovascular diseases. Histol Histopathol 2000;15:557-571.
[6].
Dzau VJ, Braun-Dullaeus RC, Sedding DG. Vascular proliferation and atherosclerosis:
new perspectives and therapeutic strategies. Nat Med 2002;8:1249-1256.
[7].
Burrig KF. The endothelium of advanced arteriosclerotic plaques in humans. Arterioscler
Thromb 1991;11:1678-1689.
[8].
Gordon D, Reidy MA, Benditt EP, Schwartz SM. Cell proliferation in human coronary
arteries. Proc Natl Acad Sci U S A 1990;87:4600-4604.
[9].
Katsuda S, Coltrera MD, Ross R, Gown AM. Human atherosclerosis. IV.
Immunocytochemical analysis of cell activation and proliferation in lesions of young
adults. Am J Pathol 1993;142:1787-1793.
[10].
Kearney M, Pieczek A, Haley L, Losordo DW, Andrés V, Schainfield R, Rosenfield R,
Isner JM. Histopathology of in-stent restenosis in patients with peripheral artery disease.
Circulation 1997;95:1998-2002.
[11].
O'Brien ER, Alpers CE, Stewart DK, Ferguson M, Tran N, Gordon D, Benditt EP,
Hinohara T, Simpson JB, Schwartz SM. Proliferation in primary and restenotic coronary
atherectomy tissue. Implications for antiproliferative therapy. Circ Res 1993;73:223-231.
[12].
Orekhov AN, Andreeva ER, Mikhailova IA, Gordon D. Cell proliferation in normal and
17
atherosclerotic human aorta: proliferative splash in lipid-rich lesions. Atherosclerosis
1998;139:41-48.
[13].
Rekhter MD, Gordon D. Active proliferation of different cell types, including
lymphocytes, in human atherosclerotic plaques. Am J Pathol 1995;147:668-677.
[14].
Tanner FC, Yang Z-Y, Duckers E, Gordon D, Nabel GJ, Nabel EG. Expression of cyclindependent kinase inhibitors in vascular disease. Circ Res 1998;82:396-403.
[15].
Veinot JP, Ma X, Jelley J, O'Brien ER. Preliminary clinical experience with the pullback
atherectomy catheter and the study of proliferation in coronary plaques. Can J Cardiol
1998;14:1457-1463.
[16].
Wei GL, Krasinski K, Kearney M, Isner JM, Walsh K, Andrés V. Temporally and
spatially coordinated expression of cell cycle regulatory factors after angioplasty. Circ
Res 1997;80:418-426.
[17].
Pickering JG, Weir L, Jekanowski J, Kearney MA, Isner JM. Proliferative activity in
peripheral and coronary atherosclerotic plaque among patients undergoing percutaneous
revascularization. J Clin Invest 1993;91:1469-1480.
[18].
O'Brien ER, Urieli-Shoval S, Garvin MR, Stewart DK, Hinohara T, Simpson JB, Benditt
EP, Schwartz SM. Replication in restenotic atherectomy tissue. Atherosclerosis
2000;152:117-126.
[19].
Dartsch PC, Voisard R, Bauriedel G, Hofling B, Betz E. Growth characteristics and
cytoskeletal organization of cultured smooth muscle cells from human primary stenosing
and restenosing lesions. Arteriosclerosis 1990;10:62-75.
[20].
Spraragen SC, Bond VP, Dahl LK. Role of hyperplasia in vascular lesions of cholesterolfed rabbits studied with thymidine-3H autoradiography. Circ Res 1962;11:329-336.
[21].
McMillan GC, Stary HC. Preliminary experience with mitotic activity of cellular elements
in the atherosclerotic plaques of cholesterol-fed rabbits studied by labeling with tritiated
18
thymidine. Ann N Y Acad Sci 1968;149:699-709.
[22].
Rosenfeld ME, Ross R. Macrophage and smooth muscle cell proliferation in
atherosclerotic lesions of WHHL and comparably hypercholesterolemic fat-fed rabbits.
Arteriosclerosis 1990;10:680-687.
[23].
Andrés V. Control of vascular smooth muscle cell growth and its implication in
atherosclerosis and restenosis. Int J Molec Med 1998;2:81-89.
[24].
Libby P, Tanaka H. The molecular basis of restenosis. Prog Cardiovasc Dis 1997;40:97106.
[25].
Morgan DO. Principles of CDK regulation. Nature 1995;374:131-134.
[26].
Nurse P. Ordering S phase and M phase in the cell cycle. Cell 1994;79:547-550.
[27].
Weinberg RA. The retinoblastoma protein and cell cycle control. Cell 1995;81:323-330.
[28].
Lavia P, Jansen-Durr P. E2F target genes and cell-cycle checkpoint control. Bioessays
1999;21:221-230.
[29].
Dyson N. The regulation of E2F by pRB-family proteins. Genes Dev 1998;12:2245-2262.
[30].
Stevaux O, Dyson NJ. A revised picture of the E2F transcriptional network and RB
function. Curr Opin Cell Biol 2002;14:684-691.
[31].
Ubersax JA, Woodbury EL, Quang PN, Paraz M, Blethrow JD, Shah K, Shokat KM,
Morgan DO. Targets of the cyclin-dependent kinase Cdk1. Nature 2003;425:859-864.
[32].
Vidal A, Koff A. Cell-cycle inhibitors: three families united by a common cause. Gene
2000;247:1-15.
[33].
Braun-Dullaeus RC, Mann MJ, Seay U, Zhang L, von Der Leyen HE, Morris RE, Dzau
VJ. Cell cycle protein expression in vascular smooth muscle cells in vitro and in vivo is
regulated through phosphatidylinositol 3-kinase and mammalian target of rapamycin.
Arterioscler Thromb Vasc Biol 2001;21:1152-1158.
[34].
Abe J, Zhou W, Taguchi J, Takuwa N, Miki K, Okazaki H, Kurokawa K, Kumada M,
19
Takuwa Y. Suppression of neointimal smooth muscle cell accumulation in vivo by
antisense cdc2 and cdk2 oligonucleotides in rat carotid artery. Biochem Biophys Res
Commun 1994;198:16-24.
[35].
Ihling C, Technau K, Gross V, Schulte-Monting J, Zeiher AM, Schaefer HE. Concordant
upregulation of type II-TGF-beta-receptor, the cyclin- dependent kinases inhibitor p27Kip1
and cyclin E in human atherosclerotic tissue: implications for lesion cellularity.
Atherosclerosis 1999;144:7-14.
[36].
Marra DE, Simoncini T, Liao JK. Inhibition of vascular smooth muscle cell proliferation
by sodium salicylate mediated by upregulation of p21 Waf1 and p27Kip1. Circulation
2000;102:2124-2130.
[37].
Ii M, Hoshiga M, Fukui R, Negoro N, Nakakoji T, Nishiguchi F, Kohbayashi E, Ishihara
T, Hanafusa T. Beraprost sodium regulates cell cycle in vascular smooth muscle cells
through cAMP signaling by preventing down-regulation of p27(Kip1). Cardiovasc Res
2001;52:500-508.
[38].
Brooks G, Yu XM, Wang Y, Crabbe MJ, Shattock MJ, Harper JV. Non-steroidal antiinflammatory drugs (NSAIDs) inhibit vascular smooth muscle cell proliferation via
differential effects on the cell cycle. J Pharm Pharmacol 2003;55:519-526.
[39].
Bauer PM, Buga GM, Ignarro LJ. Role of p42/p44 mitogen-activated-protein kinase and
p21waf1cip1 in the regulation of vascular smooth muscle cell proliferation by nitric oxide.
Proc Natl Acad Sci U S A 2001;98:12802-12807.
[40].
Sato J, Nair K, Hiddinga J, Eberhardt NL, Fitzpatrick LA, Katusic ZS, O'Brien T. eNOS
gene transfer to vascular smooth muscle cells inhibits cell proliferation via upregulation of
p27 and p21 and not apoptosis. Cardiovasc Res 2000;47:697-706.
[41].
Takahashi A, Taniguchi T, Ishikawa Y, Yokoyama M. Tranilast inhibits vascular smooth
muscle cell growth and intimal hyperplasia by induction of p21(waf1/cip1/sdi1) and p53.
20
Circ Res 1999;84:543-550.
[42].
Laufs U, Marra D, Node K, Liao JK. 3-hydroxy-3-methylglutaryl-CoA reductase
inhibitors attenuate vascular smooth muscle proliferation by preventing Rho GTPaseinduced down-regulation of p27(Kip1). Journal of Biological Chemistry 1999;274:2192621931.
[43].
Otterbein LE, Zuckerbraun BS, Haga M, Liu F, Song R, Usheva A, Stachulak C, Bodyak
N, Smith RN, Csizmadia E, Tyagi S, Akamatsu Y, Flavell RJ, Billiar TR, Tzeng E, Bach
FH, Choi AM, Soares MP. Carbon monoxide suppresses arteriosclerotic lesions
associated with chronic graft rejection and with balloon injury. Nat Med 2003;9:183-190.
[44].
Akimoto S, Mitsumata M, Sasaguri T, Yoshida Y. Laminar shear stress inhibits vascular
endothelial
cell
proliferation
by
inducing
cyclin-dependent
kinase
inhibitor
p21Sdi1/Cip1/Waf1. Circ Res 2000;86:185-190.
[45].
Sedding DG, Seay U, Fink L, Heil M, Kummer W, Tillmanns H, Braun-Dullaeus RC.
Mechanosensitive p27Kip1 regulation and cell cycle entry in vascular smooth muscle cells.
2003 2003;108:616-622.
[46].
Seasholtz TM, Zhang T, Morissette MR, Howes AL, Yang AH, Brown JH. Increased
expression and activity of RhoA are associated with increased DNA synthesis and
reduced p27Kip1 expression in the vasculature of hypertensive rats. Circ Res 2001;89:488495.
[47].
Assoian RK, Marcantonio EE. The extracellular matrix as a cell cycle control element in
atherosclerosis and restenosis. J Clin Invest 1996;98:2436-2439.
[48].
Bendeck MP, Zempo N, Clowes AW, Gelardy RE, Reidy MA. Smooth muscle cell
migration and matrix metalloproteinase expression after arterial injury in the rat. Circ Res
1994;75:539-545.
[49].
Galis S, Sukhova GK, Lark MV, Libby P. Increased expression of matrix
21
metalloproteinases and matrix degrading activity in vulnerable regions of human
atherosclerotic plaques. J Clin Invest 1994;94:2493-2503.
[50].
Zempo N, Kenagy RD, Au T, Bendeck M, Clowes MM, Reidy MA, Clowes AW. Matrix
metalloproteinases of vascular wall cells are increased in balloon-injured rat carotid
artery. J Vasc Surg 1994;20:209-217.
[51].
Southgate KM, Fisher M, Banning AP, Thurston VJ, Baker AH, Fabunmi RP, Groves PH,
Davies M, Newby AC. Upregulation of basement membrane-degrading metalloproteinase
secretion after balloon injury of pig carotid artery. Circ Res 1996;79:1177-1187.
[52].
Southgate KM, Davies M, Booth RFG, Newby AC. Involvement of extracellular matrix
degrading metalloproteinases in rabbit aortic smooth muscle cell proliferation. Biochem J
1992;288:93-99.
[53].
Zempo N, Koyama M, Kenagy RD, Lea HJ, Clowes AW. Regulation of vascular smooth
muscle cell migration and proliferation in vitro and in injured rat arteries by a synthetic
matrix metalloproteinase inhibitor. Arterioscler Throm Vasc Biol 1996;16:28-33.
[54].
Cheng L, Mantile G, Pauly R, Nater C, Felici A, Monticone R, Bilato C, Gluzband YA,
Crow MT, Stetler-Stevenson W, Capogrossi MC. Adenovirus-mediated gene transfer of
the human tissue inhibitor of metalloproteinase-2 blocks vascular smooth muscle cell
invasiveness in vitro and modulates neointimal development in vivo. Circulation
1998;98:2195-2201.
[55].
Strauss BH, Chisholm RJ, Keeley FW, Gotlieb AI, Logan RA, Armstrong PW.
Extracellular matrix remodeling after balloon angioplasty injury in a rabbit model of
restenosis. Circ Res 1994;75:650-658.
[56].
Karim MA, Miller DD, Farrar MA, Elefheriades E, Reddy BH, Brelan CM, Samarel AM.
Histomorphometric and biochemical correlates of arterial procollagen gene expression
during vascular repair after experimental angioplasty. Circulation 1995;91:2049-2057.
22
[57].
Coats WD, Jr., Whittaker P, Cheung DT, Currier JW, Han B, Faxon DP. Collagen content
is significantly lower in restenotic versus nonrestenotic vessels after balloon angioplasty
in the atherosclerotic rabbit model. Circulation 1997;95:1293-1300.
[58].
Koyama H, Raines EW, Bornfeldt KE, Roberts JM, Ross R. Fibrillar collagen inhibits
arterial smooth muscle proliferation through regulation of cdk2 inhibitors. Cell
1996;87:1069-1078.
[59].
Zhu X, Ohtsubo M, Böhmer RM, Roberts JM, Assoian RK. Adhesion-dependent cell
cycle progression linked to the expression of cyclin D1, activation of cyclin E-cdk2, and
phosphorylation of the retinoblastoma protein. J Cell Biol 1996;133:391-403.
[60].
Chen D, Krasinski K, Chen D, Sylvester A, Chen J, Nisen PD, Andrés V. Downregulation
of cyclin-dependent kinase 2 activity and cyclin A promoter activity in vascular smooth
muscle cells by p27Kip1, an inhibitor of neointima formation in the rat carotid artery. J Clin
Invest 1997;99:2334-2341.
[61].
Roque M, Cordón-Cardo C, Fuster V, Reis ED, Drobnjak M, Badimón JJ. Modulation of
apoptosis, proliferation, and p27 expression in a porcine coronary angioplasty model.
Atherosclerosis 2000;153:315-322.
[62].
Reis ED, Roque M, Cordón-Cardo C, Drobnjak M, Fuster V, Badimón JJ. Apoptosis,
proliferation, and p27 expression during vessel wall healing: time course study in a mouse
model of transluminal femoral artery injury. J Vasc Surg 2000;32:1022-1029.
[63].
Braun-Dullaeus RC, Ziegler A, Bohle RM, Bauer E, Hein S, Tillmanns H, Haberbosch W.
Quantification of the cell-cycle inhibitors p27(Kip1) and p21(Cip1) in human atherectomy
specimens: primary stenosis versus restenosis. J Lab Clin Med 2003;141:179-189.
[64].
Ihling C, Menzel G, Wellens E, Monting JS, Schaefer HE, Zeiher AM. Topographical
association between the cyclin-dependent kinases inhibitor p21, p53 accumulation, and
cellular proliferation in human atherosclerotic tissue. Arterioscler Thromb Vasc Biol
23
1997;17:2218-2224.
[65].
Nathe TJ, Deou J, Walsh B, Bourns B, Clowes AW, Daum G. Interleukin-1beta inhibits
expression of p21(WAF1/CIP1) and p27(KIP1) and enhances proliferation in response to
platelet-derived growth factor-BB in smooth muscle cells. Arterioscler Thromb Vasc Biol
2002;22:1293-1298.
[66].
Díez-Juan A, Andrés V. The growth suppressor p27Kip1 protects against diet-induced
atherosclerosis. FASEB J 2001;15:1989-1995.
[67].
Diez-Juan A, Perez P, Aracil M, Sancho D, Bernad A, Sanchez-Madrid F, Andres V.
Selective inactivation of p27Kip1 in hematopoietic progenitor cells increases neointimal
macrophage proliferation and accelerates atherosclerosis. Blood 2004;In Press.
[68].
Cheng T, Rodrigues N, Dombkowski D, Stier S, Scadden DT. Stem cell repopulation
efficiency but not pool size is governed by p27kip1. Nat Med 2000;6:1235-1240.
[69].
Antonov AS, Munn DH, Kolodgie FD, Virmani R, Gerrity RG. Aortic endothelial cells
regulate proliferation of human monocytes in vitro via a mechanism synergistic with
macrophage colony-stimulating factor. Convergence at the cyclin E/p27Kip1 regulatory
checkpoint. J Clin Invest 1997;99:2867-2876.
[70].
Xaus J, Comalada M, Cardo M, Valledor AF, Celada A. Decorin inhibits macrophage
colony-stimulating factor proliferation of macrophages and enhances cell survival through
induction of p27Kip1 and p21Waf1. Blood 2001;98:2124-2133.
[71].
Roque M, Reis ED, Cordon-Cardo C, Taubman MB, Fallon JT, Fuster V, Badimon JJ.
Effect of p27 deficiency and rapamycin on intimal hyperplasia: in vivo and in vitro
studies using a p27 knockout mouse model. Lab Invest 2001;81:895-903.
[72].
Bochaton-Piallat ML, Ropraz P, Gabbiani F, Gabbiani G. Phenotypic heterogeneity of rat
arterial smooth muscle cell clones. Implications for the development of experimental
intimal thickening. Arterioscler Thromb Vasc Biol 1996;16:815-820.
24
[73].
Chamley-Campbell JH, Campbell GR, Ross R. Phenotype-dependent response of cultured
aortic smooth muscle to serum mitogens. J Cell Biol 1981;89:379-383.
[74].
Li S, Fan YS, Chow LH, Van Den Diepstraten C, van Der Veer E, Sims SM, Pickering
JG. Innate diversity of adult human arterial smooth muscle cells: cloning of distinct
subtypes from the internal thoracic artery. Circ Res 2001;89:517-525.
[75].
Majack RA, Grieshaber NA, Cook CL, Weiser MC, McFall RC, Grieshaber SS, Reidy
MA, Reilly CF. Smooth muscle cells isolated from the neointima after vascular injury
exhibit altered responses to platelet-derived growth factor and other stimuli. J Cell Physiol
1996;167:106-112.
[76].
Topouzis S, Majesky MW. Smooth muscle lineage diversity in the chick embryo. Two
types of aortic smooth muscle cell differ in growth and receptor-mediated transcriptional
responses to transforming growth factor-. Dev Biol 1996;178:430-445.
[77].
Yang Z, Oemar BS, Carrel T, Kipfer B, Julmy F, Lüscher TF. Different proliferative
properties of smooth muscle cells of human arterial and venous bypass vessels: role of
PDGF receptors, mitogen-activated protein kinase, and cyclin-dependent kinase
inhibitors. Circulation 1998;97:181-187.
[78].
Olson NE, Kozlowski J, Reidy MA. Proliferation of intimal smooth muscle cells.
Attenuation of basic fibroblast growth factor 2-stimulated proliferation is associated with
increased expression of cell cycle inhibitors. J Biol Chem 2000;275:11270-11277.
[79].
Castro C, Díez-Juan A, Cortés MJ, Andrés V. Distinct regulation of mitogen-activated
protein kinases and p27Kip1 in smooth muscle cells from different vascular beds. A
potential role in establishing regional phenotypic variance. J Biol Chem 2003;278:44824490.
[80].
De Azevedo WF, Leclerc S, Meijer L, Havlicek L, Strnad M, Kim SH. Inhibition of
cyclin-dependent kinases by purine analogues: crystal structure of human cdk2 complexed
25
with roscovitine. Eur J Biochem 1997;243:518-526.
[81].
Gray NS, Wodicka L, Thunnissen AM, Norman TC, Kwon S, Espinoza FH, Morgan DO,
Barnes G, LeClerc S, Meijer L, Kim SH, Lockhart DJ, Schultz PG. Exploiting chemical
libraries, structure, and genomics in the search for kinase inhibitors. Science
1998;281:533-538.
[82].
Schulze-Gahmen U, Brandsen J, Jones HD, Morgan DO, Meijer L, Vesely J, Kim SH.
Multiple modes of ligand recognition: crystal structures of cyclin- dependent protein
kinase 2 in complex with ATP and two inhibitors, olomoucine and isopentenyladenine.
Proteins 1995;22:378-391.
[83].
Brooks EE, Gray NS, Joly A, Kerwar SS, Lum R, Mackman RL, Norman TC, Rosete J,
Rowe M, Schow SR, Schultz PG, Wang X, Wick MM, Shiffman D. CVT-313, a specific
and potent inhibitor of CDK2 that prevents neointimal proliferation. J Biol Chem
1997;272:29207-29211.
[84].
Losiewicz MD, Carlson BA, Kaur G, Sausville EA, Worland PJ. Potent inhibition of
CDC2 kinase activity by the flavonoid L86-8275. Biochem Biophys Res Commun
1994;201:589-595.
[85].
Carlson BA, Dubay MM, Sausville EA, Brizuela L, Worland PJ. Flavopiridol induces G1
arrest with inhibition of cyclin-dependent kinase (CDK) 2 and CDK4 in human breast
carcinoma cells. Cancer Res 1996;56:2973-2978.
[86].
Ruef J, Meshel AS, Hu Z, Horaist C, Ballinger CA, Thompson LJ, Subbarao VD, Dumont
JA, Patterson C. Flavopiridol inhibits smooth muscle cell proliferation in vitro and
neointimal formation In vivo after carotid injury in the rat. Circulation 1999;100:659-665.
[87].
Kaur G, Stetler-Stevenson M, Sebers S, Worland P, Sedlacek H, Myers C, Czech J, Naik
R, Sausville E. Growth inhibition with reversible cell cycle arrest of carcinoma cells by
flavone L86-8275. J Natl Cancer Inst 1992;84:1736-1740.
26
[88].
Morishita R, Gibbons GH, Ellison KE, Nakajima M, von der Leyen H, Zhang L, Kaneda
Y, Ogihara T, Dzau VJ. Intimal hyperplasia after vascular injury is inhibited by antisense
cdk2 kinase oligonucleotides. J Clin Invest 1994;93:1458-1464.
[89].
Morishita R, Gibbons GH, Ellison KE, Nakajima M, Zhang L, Kaneda Y, Ogihara T,
Dzau VJ. Single intraluminal delivery of antisense cdc2 kinase and proliferating-cell
nuclear antigen oligonucleotides results in chronic inhibition of neointimal hyperplasia.
Proc Natl Acad Sci USA 1993;90:8474-8478.
[90].
Morishita R, Gibbons GH, Kaneda Y, Ogihara T, Dzau VJ. Pharmacokinetics of antisense
oligodeoxyribonucleotides (cyclin B1 and CDC 2 kinase) in the vessel wall in vivo:
enhanced therapeutic utility for restenosis by HVJ-liposome delivery. Gene 1994;149:1319.
[91].
Zhu NL, Wu L, Liu PX, Gordon EM, Anderson WF, Starnes VA, Hall FL.
Downregulation of cyclin G1 expression by retrovirus-mediated antisense gene transfer
inhibits vascular smooth muscle cell proliferation and neointima formation. Circulation
1997;96:628-635.
[92].
Mann M, Gibbons GH, Kernoff RS, Diet FP, Tsao PS, Cooke JP, Kaneda Y, Dzau VJ.
Genetic engineering of vein grafts resistant to atherosclerosis. Proc Natl Acad Sci USA
1995;92:4502-4506.
[93].
Suzuki J-I, Isobe M, Morishita R, Aoki M, Horie S, Okubo Y, Kaneda Y, Sawa Y,
Matsuda H, Ogihara T, Sekiguchi M. Prevention of graft coronary arteriosclerosis by
antisense cdk2 kinase oligonucleotide. Nat Med 1997;3:900-903.
[94].
Gascón-Irún M, Sanz-González SM, Andrés V. Gene therapy antiproliferative strategies
against cardiovascular disease. Gene Ther Mol Biol 2003;7:75-89.
[95].
Chang MW, Barr E, Lu MM, Barton K, Leiden JM. Adenovirus-mediated overexpression of the cyclin/cyclin-dependent kinase inhibitor, p21 inhibits vascular smooth
27
muscle cell proliferation and neointima formation in the rat carotid artery model of
balloon angioplasty. J Clin Invest 1995;96:2260-2268.
[96].
Ueno H, Masuda S, SNishio S, Li JJ, Yamamoto H, Takeshita A. Adenovirus-mediated
transfer of cyclin-dependent kinase inhibitor p21 suppresses neointimal formation in the
balloon-injured rat carotid arteries in vivo. Ann N Y Acad Sci 1997;811:401-411.
[97].
Yang Z-Y, Simari RD, Perkins ND, San H, Gordon D, Nabel GJ, Nabel EG. Role of p21
cyclin-dependent kinase inhibitor in limiting intimal cell proliferation in response to
arterial injury. Proc Natl Acad Sci USA 1996;93:7905-7910.
[98].
Tanner FC, Boehm M, Akyürek LM, San H, Yang Z-Y, Tashiro J, Nabel GJ, Nabel EG.
Differential effects of the cyclin-dependent kinase inhibitors p27Kip1, p21Cip1, and p16Ink4
on vascular smooth muscle cell proliferation. Circulation 2000;101:2022-2025.
[99].
Tagaki Y. Adenovirus-mediated overexpression of a cyclin-dependent kinase inhibitor,
p57Kip2, suppressed vascular smooth muscle cell proliferation. Hokkaido Igaku Zasshi
2002;77:221-230.
[100]. Condorelli G, Aycock JK, Frati G, Napoli C. Mutated p21/WAF/CIP transgene
overexpression reduces smooth muscle cell proliferation, macrophage deposition,
oxidation-sensitive mechanisms, and restenosis in hypercholesterolemic apolipoprotein E
knockout mice. FASEB J 2001;15:2162-2170.
[101]. Bai H, Morishita R, Kida I, Yamakawa T, Zhang W, Aoki M, Matsushita H, Noda A,
Nagai R, Kaneda I, Higaki J, Ogihara T, Sawa Y, Matsuda H. Inhibition of intimal
hyperplasia after vein grafting by in vivo transfer of human senescent cell-derived
inhibitor-1 gene. Gene Ther 1998;5:761-769.
[102]. Weiss RH, Randour CJ. Attenuation of matrix protein secretion by antisense
oligodeoxynucleotides to the cyclin kinase inhibitor p21Waf1/Cip1. Atherosclerosis
2002;161:105-112.
28
[103]. Lamphere L, Tsui L, Wick S, Nakano T, Kilinski L, Finer M, McArthur J, Gyuris J.
Novel chimeric p16 and p27 molecules with increased antiproliferative activity for
vascular disease gene therapy. J Mol Med 2000;78:451-459.
[104]. McArthur JG, Qian H, Citron D, Banik GG, Lamphere L, Gyuris J, Tsui L, George SE.
p27-p16 chimera: a superior antiproliferative for the prevention of neointimal hyperplasia.
Mol Ther 2001;3:8-13.
[105]. Marx SO, Jayaraman T, Go LO, Marks AR. Rapamycin-FKBP inhibits cell cycle
regulators of proliferation in vascular smooth muscle cells. Circ Res 1995;76:412-417.
[106]. Poon M, Marx SO, Gallo R, Badimon JJ, Taubman MB, Marks AR. Rapamycin inhibits
vascular smooth muscle cell migration. J Clin Invest 1996;98:2277-2283.
[107]. Alcocer F, Whitley D, Salazar-Gonzalez JF, Jordan WD, Sellers MT, Eckhoff DE, Suzuki
K, Macrae C, Bland KI. Quercetin inhibits human vascular smooth muscle cell
proliferation and migration. Surgery 2002;131:198-204.
[108]. Kubens BS, Niggemann B, Zanker KS. Prevention of entrance into G2 cell cycle phase by
mimosine decreases locomotion of cells from the tumor cell line SW480. Cancer Lett
2001;162 Suppl:S39-S47.
[109]. Hu Y, Zou Y, Dietrich H, Wick G, Xu Q. Inhibition of neointima hyperplasia of mouse
vein grafts by locally applied suramin. Circulation 1999;100:861-868.
[110]. Law RE, Meehan WP, Xi X-P, Graf K, Wuthrich DA, Coats W, Faxon D. Troglitazone
inhibits vascular smooth muscle cell growth and intimal hyperplasia. J Clin Invest
1996;98:1897-1905.
[111]. Geraldes P, Sirois MG, Bernatchez PN, Tanguay JF. Estrogen regulation of endothelial
and smooth muscle cell migration and proliferation: role of p38 and p42/44 mitogenactivated protein kinase. Arterioscler Thromb Vasc Biol 2002;22:1585-1590.
[112]. Biro
S,
Fu
Y-M,
Yu
Z-X,
Epstein
29
SE.
Inhibitory
effects
of
antisense
oligodeoxynucleotides targeting c-myc mRNA on smooth muscle cell proliferation and
migration. Proc Natl Acad Sci U S A 1993;90:654-658.
[113]. Ahn JD, Morishita R, Kaneda Y, Lee SJ, Kwon KY, Choi SY, Lee KU, Park JY, Moon IJ,
Park JG, Yoshizumi M, Ouchi Y, Lee IK. Inhibitory effects of novel AP-1 decoy
oligodeoxynucleotides on vascular smooth muscle cell proliferation in vitro and
neointimal formation in vivo. Circ Res 2002;90:1325-1332.
[114]. Mayr U, Mayr M, Li C, Wernig F, Dietrich H, Hu Y, Xu Q. Loss of p53 accelerates
neointimal lesions of vein bypass grafts in mice. Circ Res 2002;90:197-204.
[115]. Bonneton C, Sibarita JB, Thiery JP. Relationship between cell migration and cell cycle
during the initiation of epithelial to fibroblastoid transition. Cell Motil Cytoskeleton
1999;43:288-295.
[116]. Fukui R, Amakawa M, Hoshiga M, Shibata N, Kohbayashi E, Seto M, Sasaki Y, Ueno T,
Negoro N, Nakakoji T, Ii M, Nishiguchi F, Ishihara T, Ohsawa N. Increased migration in
late G(1) phase in cultured smooth muscle cells. Am J Physiol Cell Physiol
2000;279:C999-C1007.
[117]. Cho RJ, Huang M, Campbell MJ, Dong H, Steinmetz L, Sapinoso L, Hampton G, Elledge
SJ, Davis RW, Lockhart DJ. Transcriptional regulation and function during the human
cell cycle. Nat Genet 2001;27:48-54.
[118]. Fukui R, Shibata N, Kohbayashi E, Amakawa M, Furutama D, Hoshiga M, Negoro N,
Nakakouji T, Ii M, Ishihara T, Ohsawa N. Inhibition of smooth muscle cell migration by
the p21 cyclin-dependent kinase inhibitor (Cip1). Atherosclerosis 1997;132:53-59.
[119]. Fahraeus R, Lane DP. The p16INK4a tumour suppressor protein inhibits v3 integrinmediated cell spreading on vitronectin by blocking PKC-dependent localization of v3 to
focal contacts. EMBO J 1999;18:2106-2118.
[120]. Goukassian D, Díez-Juan A, Asahara T, Schratzberger P, Silver M, Murayama T, Isner
30
JM, Andrés V. Overexpression of p27Kip1 by doxycycline-regulated adenoviral vectors
inhibits endothelial cell proliferation and migration and impairs angiogenesis. FASEB J
2001;15:1877-1885.
[121]. Sun J, Marx SO, Chen H-J, Poon M, Marks AR, Rabbani LE. Role for p27Kip1 in vascular
smooth muscle cell migration. Circulation 2001;103:2967-2972.
[122]. Ohshima T, Gilmore EC, Longenecker G, Jacobowitz DM, Brady RO, Herrup K,
Kulkarni AB. Migration defects of cdk5(-/-) neurons in the developing cerebellum is cell
autonomous. J Neurosci 1999;19:6017-6026.
[123]. Humbert S, Dhavan R, Tsai L. p39 activates cdk5 in neurons, and is associated with the
actin cytoskeleton. J Cell Sci 2000;113:975-983.
[124]. Gao C, Negash S, Wang HS, Ledee D, Guo H, Russell P, Zelenka P. Cdk5 mediates
changes in morphology and promotes apoptosis of astrocytoma cells in response to heat
shock. J Cell Sci 2001;114:1145-1153.
[125]. Díez-Juan A, Andres V. Coordinate control of proliferation and migration by the
p27Kip1/cyclin-dependent kinase/retinoblastoma pathway in vascular smooth muscle
cells and fibroblasts. Circ Res 2003;92:402-410.
[126]. D'Souza SJ, Vespa A, Murkherjee S, Maher A, Pajak A, Dagnino L. E2F-1 is essential for
normal epidermal wound repair. J Biol Chem 2002;277:10626-10632.
[127]. Luo Y, Marx SO, Kiyokawa H, Koff A, Massague J, Marks AR. Rapamycin resistance
tied to defective regulation of p27Kip1. Mol Cell Biol 1996;16:6744-6751.
[128]. Burke SE, Lubbers NL, Chen YW, Hsieh GC, Mollison KW, Luly JR, Wegner CD.
Neointimal formation after balloon-induced vascular injury in Yucatan minipigs is
reduced by oral rapamycin. J Cardiovasc Pharmacol 1999;33:829-835.
[129]. Gallo R, Padurean A, Jayaraman T, Marx S, Rogue M, Adelman S, Chesebro J, Fallon J,
Fuster V, Marks A, Badimón JJ. Inhibition of intimal thickening after balloon angioplasty
31
in porcine coronary arteries by targeting regulators of the cell cycle. Circulation
1999;99:2164-2170.
[130]. Gregory CR, Huie P, Billingham ME, Morris RE. Rapamycin inhibits arterial intimal
thickening caused by both alloimmune and mechanical injury. Its effect on cellular,
growth factor, and cytokine response in injured vessels. Transplantation 1993;55:14091418.
[131]. Gregory CR, Huie P, Shorthouse R, Wang J, Rowan R, Billingham ME, Morris RE.
Treatment with rapamycin blocks arterial intimal thickening following mechanical and
alloimmune injury. Transplant Proc 1993;25:120-121.
[132]. Meiser BM, Billingham ME, Morris RE. Effects of cyclosporin, FK506, and rapamycin
on graft-vessel disease. Lancet 1991;338:1297-1298.
[133]. Poston RS, Billingham M, Hoyt EG, Pollard J, Shorthouse R, Morris RE, Robbins RC.
Rapamycin reverses chronic graft vascular disease in a novel cardiac allograft model.
Circulation 1999;100:67-74.
[134]. Suzuki T, Kopia G, Hayashi S, Bailey LR, Llanos G, Wilensky R, Klugherz BD,
Papandreou G, Narayan P, Leon MB, Yeung AC, Tio F, Tsao PS, Falotico R, Carter AJ.
Stent-based delivery of sirolimus reduces neointimal formation in a porcine coronary
model. Circulation 2001;104:1188-1193.
[135]. Castro C, Campistol JM, Sancho D, Sánchez-Madrid F, Casals E, Andrés V. Rapamycin
attenuates atherosclerosis induced by dietary cholesterol in apolipoprotein-deficient mice
through a p27Kip1-independent pathway. Atherosclerosis 2004;In Press.
[136]. Elloso MM, Azrolan N, Sehgal SN, Hsu PL, Phiel KL, Kopec CA, Basso MD, Adelman
SJ. Protective effect of the immunosuppressant sirolimus against aortic atherosclerosis in
apo E-deficient mice. Am J Transplant 2003;3:562-569.
[137]. Marks AR. Sirolimus for the prevention of in-stent restenosis in a coronary artery. N Engl
32
J Med 2003;349:1307-1309.
[138]. Sousa JE, Sousa AG, Costa MA, Abizaid AC, Feres F. Use of rapamycin-impregnated
stents in coronary arteries. Transplant Proc 2003;35 (Suppl. 3A):S165-S170.
[139]. Windecker S, Roffi M, Meier B. Sirolimus eluting stent: a new era in interventional
cardiology? Curr Pharm Des 2003;9:1077-1094.
[140]. Mann MJ, Gibbons GH, Tsao PS, von der Leyen HE, Cooke JP, Buitrago R, Kernoff R,
Dzau VJ. Cell cycle inhibition preserves endothelial function in genetically engineered
rabbit vein grafts. J Clin Invest 1997;99:1295-1301.
[141]. Morishita R, Gibbons GH, Horiuchi M, Ellison KE, Nakajima M, Zhang L, Kaneda Y,
Ogihara T, Dzau VJ. A gene therapy strategy using a transcription factor decoy of the E2F
binding site inhibits smooth muscle cell proliferation in vivo. Proc Natl Acad Sci USA
1995;92:5855-5859.
[142]. Mann MJ, Whittemore AD, Donaldson MC, Belkin M, Conte MS, Polak JF, Orav EJ,
Ehsan A, Dell'Acqua G, Dzau VJ. Ex-vivo gene therapy of human vascular bypass grafts
with E2F decoy: the PREVENT single-centre, randomised, controlled trial. Lancet
1999;354:1493-1498.
[143]. Mangi AA, Dzau VJ. Gene therapy for human bypass grafts. Ann Med 2001;33:153-155.
[144]. Sedlacek HH. Mechanisms of action of flavopiridol. Crit Rev Oncol Hematol
2001;38:139-170.
33
Download