ArticlePDF Available

Coexisting proinflammatory and antioxidative endothelial transcription profiles in a disturbed flow region of the adult porcine aorta

Authors:

Abstract and Figures

In the arterial circulation, regions of disturbed flow (DF), which are characterized by flow separation and transient vortices, are susceptible to atherogenesis, whereas regions of undisturbed laminar flow (UF) appear protected. Coordinated regulation of gene expression by endothelial cells (EC) may result in differing regional phenotypes that either favor or inhibit atherogenesis. Linearly amplified RNA from freshly isolated EC of DF (inner aortic arch) and UF (descending thoracic aorta) regions of normal adult pigs was used to profile differential gene expression reflecting the steady state in vivo. By using human cDNA arrays, approximately 2,000 putatively differentially expressed genes were identified through false-discovery-rate statistical methods. A sampling of these genes was validated by quantitative real-time PCR and/or immunostaining en face. Biological pathway analysis revealed that in DF there was up-regulation of several broad-acting inflammatory cytokines and receptors, in addition to elements of the NF-kappaB system, which is consistent with a proinflammatory phenotype. However, the NF-kappaB complex was predominantly cytoplasmic (inactive) in both regions, and no significant differences were observed in the expression of key adhesion molecules for inflammatory cells associated with early atherogenesis. Furthermore, there was no histological evidence of inflammation. Protective profiles were observed in DF regions, notably an enhanced antioxidative gene expression. This study provides a public database of regional EC gene expression in a normal animal, implicates hemodynamics as a contributory mechanism to athero-susceptibility, and reveals the coexistence of pro- and antiatherosclerotic transcript profiles in susceptible regions. The introduction of additional risk factors may shift this balance to favor lesion development.
Content may be subject to copyright.
Coexisting proinflammatory and antioxidative
endothelial transcription profiles in a disturbed flow
region of the adult porcine aorta
Anthony G. Passerini*
†‡
, Denise C. Polacek*
, Congzhu Shi*, Nadeene M. Francesco*, Elisabetta Manduchi
§
,
Gregory R. Grant
§
, William F. Pritchard
, Steven Powell
, Gary Y. Chang*
, Christian J. Stoeckert, Jr.
§
**,
and Peter F. Davies*
†,††‡‡
*Institute for Medicine and Engineering, Departments of
Bioengineering,
††
Pathology and Laboratory Medicine, and **Genetics, and
§
Center for
Bioinformatics, University of Pennsylvania, Philadelphia, PA 19104;
U.S. Food and Drug Administration, Rockville, MD 20852; and
AstraZeneca
Pharmaceuticals, Mereside Alderley Park, Macclesfield, Cheshire SK10 4TG, United Kingdom
Edited by Louis J. Ignarro, University of California School of Medicine, Los Angeles, CA, and approved December 22, 2003 (received for review
September 15, 2003)
In the arterial circulation, regions of disturbed flow (DF), which are
characterized by flow separation and transient vortices, are suscep-
tible to atherogenesis, whereas regions of undisturbed laminar flow
(UF) appear protected. Coordinated regulation of gene expression by
endothelial cells (EC) may result in differing regional phenotypes that
either favor or inhibit atherogenesis. Linearly amplified RNA from
freshly isolated EC of DF (inner aortic arch) and UF (descending
thoracic aorta) regions of normal adult pigs was used to profile
differential gene expression reflecting the steady state in vivo.By
using human cDNA arrays, 2,000 putatively differentially expressed
genes were identified through false-discovery-rate statistical meth-
ods. A sampling of these genes was validated by quantitative real-
time PCR andor immunostaining en face. Biological pathway analysis
revealed that in DF there was up-regulation of several broad-acting
inflammatory cytokines and receptors, in addition to elements of the
NF-
B system, which is consistent with a proinflammatory pheno-
type. However, the NF-
B complex was predominantly cytoplasmic
(inactive) in both regions, and no significant differences were ob-
served in the expression of key adhesion molecules for inflammatory
cells associated with early atherogenesis. Furthermore, there was no
histological evidence of inflammation. Protective profiles were ob-
served in DF regions, notably an enhanced antioxidative gene ex-
pression. This study provides a public database of regional EC gene
expression in a normal animal, implicates hemodynamics as a con-
tributory mechanism to athero-susceptibility, and reveals the coex-
istence of pro- and antiatherosclerotic transcript profiles in suscepti-
ble regions. The introduction of additional risk factors may shift this
balance to favor lesion development.
T
here is a strong correlation between flow characteristics and the
focal and regional nature of atherogenesis. Sites of disturbed
flow (DF) (e.g., the inner wall of curved vessels and the wall
opposite the flow divider at branches and bifurcations) are suscep-
tible to lesion development, whereas regions of undisturbed laminar
flow (UF) are relatively protected (1–3). Throughout the initiation
and development of atherosclerotic lesions, the endothelium is
retained as the interface between blood and arterial tissues (4, 5),
where its function both before and during lesion formation is critical
to the pathological outcome. It has been hypothesized that coor-
dinated regulation of endothelial gene expression in response to
local biomechanical forces results in differing regional phenotypes
that promote athero-protection or athero-susceptibility (6). This
hypothesis has been addressed in principle by gene expression
studies of cultured endothelium (7–20). However, it has only been
addressed in vivo for a limited number of candidate genes and
proteins (21–25).
Studies of candidate gene expression by endothelial cells (EC) in
culture (12, 18, 23) and in knock-out animals (22, 24, 25) suggest
that transcriptional activity is linked to flow disturbances. In EC
cultures, unidirectional laminar shear stress (LSS) was correlated
with protective profiles of gene expression (e.g., antioxidative,
antiinflammatory, andor antiproliferative) when compared to the
absence of flow (7–11, 13, 14, 17, 19) and selected examples of the
protective genes have been shown to be expressed in endothelium
in vivo (14, 19). A small number of microarray studies comparing
EC gene expression between DF and UF conditions in vitro have
been conducted (15, 16); however, transcriptional profiling in vivo
has been limited by the small sample size available within hemo-
dynamic regions of interest. The in vivo correlation between local
gene expression and athero-susceptibility has therefore been gen-
erated primarily by extrapolation of in vitro findings. Using RNA
amplification to overcome sampling limitations, we have profiled
regional EC gene expression in the normal adult pig aorta to
investigate athero-susceptibility as a function of differential hemo-
dynamics in vivo.
Materials and Methods
Sample Collection and Characterization. Endothelial cells were
freshly harvested from eight adult pig aortas (Landrace X York-
shire males, 230–250 pounds; Hatfield Industries, Hatfield, PA).
The ascending aorta, aortic arch, and descending thoracic aorta
were dissected from the surrounding tissue (Fig. 1A), flushed with
cold sterile PBS, and incised lengthwise. EC were gently scraped
from a 1-cm
2
region located at the inner-curve and lateral walls of
the aortic arch (DF) and separately from the dorsal descending
thoracic aorta (UF) as illustrated in Fig. 1B, with each sample
representing 10,000 cells. The cells were transferred directly to a
lysis buffer containing RNase inhibitors. Representative cell
scrapes were periodically transferred to glass microscope slides to
monitor cell purity by immunostaining with EC-specific anti-
platelet-endothelial cell adhesion molecule 1 and smooth muscle
cell (SMC)-specific anti-
-actin antibodies. Additional tissue sam-
ples were fixed in 4% paraformaldehyde for en face immunostain-
ing, regional EC nuclear staining (Hoechst 33258), and cross-
sectional vessel histology.
Microarray Procedures. These procedures have been described (26)
and are provided in full in Supporting Materials and Methods, which
is published as supporting information on the PNAS web site. Total
This paper was submitted directly (Track II) to the PNAS office.
Abbreviations: DF, disturbed flow; UF, undisturbed flow; EC, endothelial cells; LSS, laminar
shear stress; SMC, smooth muscle cell; QRT-PCR, quantitative real-time PCR; VCAM-1,
vascular cell adhesion molecule 1; ROS, reactive oxygen species; GPX3, glutathione perox-
idase 3; TNF, tumor necrosis factor.
A.G.P. and D.C.P. contributed equally to this work.
‡‡
To whom correspondence should be addressed at: Institute for Medicine and Engineer-
ing, University of Pennsylvania, 1010 Vagelos Laboratories, 3340 Smith Walk, Philadel-
phia, PA 19104. E-mail: pfd@pobox.upenn.edu.
© 2004 by The National Academy of Sciences of the USA
2482–2487
PNAS
February 24, 2004
vol. 101
no. 8 www.pnas.orgcgidoi10.1073pnas.0305938101
RNA (100 ng; 1 ng of mRNA) was linearly amplified (27), and 2
g of amplified antisense RNA was used to synthesize
33
P-labeled
DNA probes. These probes were hybridized to nylon microarray
filters, custom designed, and printed by AstraZeneca Pharmaceu-
ticals (Alderley Park, UK), which contained 13,824 3-biased,
sequence-verified human cDNA clones (1.52.0 kb) spotted in
duplicate. Matched DF and UF arrays for the same animal (n 8)
were processed simultaneously.
Data Analysis and Bioinformatics. Image files were quantified with
ARRAYVISION 6.3 (Imaging Research, St. Catherines, ON, Canada)
as described (26). A putative set of differentially expressed genes
was identified by taking the union of predictions made through the
methods of
ARRAYSTAT 1.2 (Imaging Research) (28) in ‘‘unpaired’’
mode and
SAM 1.15 (Significance Analysis of Microarrays) (29) in
‘‘one class’’ mode. Both methods used an expected false-discovery
rate (30) of 5%. The putative set of differentially expressed genes
was imported into
GENESPRING (Silicon Genetics, Redwood City,
CA) for annotation by using information available in public data-
bases and hierarchical classification according to a simple gene
ontology construction. A detailed description of the data analysis is
provided in Supporting Materials and Methods. Results obtained by
other analytical approaches can be accessed at www.cbil.upenn.edu/
RAD/normalpigstudy. In accordance with proposed standards of
the Microarray Gene Expression Data Society (www.mged.org),
the complete annotated study is publicly available in a
MIAME
compliant framework through the RNA Abundance Database
(www.cbil.upenn.edu/RAD) (31).
Validation Studies. Immunostaining on fixed tissue sections was
performed according to standard procedures for selected proteins.
Quantitative real-time PCR (QRT-PCR) was performed on se-
lected genes as described (26). Expression was measured in paired
DF and UF samples from six animals, and a ratio was calculated for
each animal based on a minimum of three replicate observations.
If the PCR ratios for at least four of six animals were 1.1 or 0.9,
then the PCR result was designated ‘‘up-regulation’’ or ‘‘down-
regulation,’’ respectively. If the PCR ratios for at least four of six
animals were between 0.9 and 1.1, then the designation was
‘‘unchanged.’’ In all other cases, no decision was made based on the
PCR results. Where the PCR ratios were 1.1 in three animals and
0.9 in the remaining animals, a note was made about this
interanimal discrepancy.
Results
Regional Characterization of EC. The presence of flow reversal in the
aortic arch of adult boars similar to that measured in humans (32)
was confirmed by ultrasound (data not shown). EC isolated from
DF and UF regions (Fig. 1 A and B) displayed differences in cell
shape and alignment that reflected the local hemodynamic condi-
tions (Fig. 4 A and B, which is published as supporting information
on the PNAS web site). In situ immunostaining with EC-specific
platelet-endothelial cell adhesion molecule 1 and SMC-specific
-actin antibodies confirmed the presence of an intact endothelium
(Fig. 4 C and D). There was no histological evidence of inflamma-
tion at either site (Fig. 5, which is published as supporting infor-
mation on the PNAS web site). To monitor the specificity of the
isolation procedure for EC, cell scrapes were transferred to glass
microscope slides where cell-specific staining confirmed EC purity
of 99% (Fig. 1 C and D). Thus near-pure isolates of EC showing
morphologies consistent with differential hemodynamic environ-
ments in vivo were rapidly obtained for transcript profiling.
Differential Expression. Putative sets of differentially expressed EC
genes (DF compared with UF) identified in the normal pig aorta
through the predictions of
ARRAYSTAT andor SAM are summa-
rized in Table 1, which also shows their annotation by biological
classifications of known relevance to the initiation and progression
of atherosclerosis.
SAM analysis identified a larger number of genes
(1,823) than did
ARRAYSTAT (1,048) at the same false-discovery
rate of 5%, while capturing 75% of the
ARRAYSTAT set. An
expanded version of Table 1 with links to fully annotated gene lists
is available online (www.cbil.upenn.edu/RAD/normalpigstudy).
QRT-PCR. Twenty-seven genes identified as significantly differen-
tially expressed were selected for validation by QRT-PCR. Fig. 2
summarizes the concordance of differential expression determined
Fig. 1. (A and B) Illustration of the endothelial cell isolation procedure. About
10,000 cells were scraped from precisely dened regions (1cm
2
) of the aortic
arch (DF) and the descending thoracic aorta (UF). (C and D) A representative eld
of EC isolated by mechanical scraping and stained with anti-platelet-endothelial
cell adhesion molecule 1 (PECAM-1) and anti-
-actin antibodies. EC purity (green)
was routinely 99% with only occasional contamination by isolated SMC (red;
arrow).
Table 1. Genes identied as differentially expressed (DF vs. UF)
in the normal pig aorta by biological classication
Biological classication
No. represented
on array
No. differentially
expressed
All genes 13824 2091
Adhesion 298 44
Apoptosis 244 47
Coagulation 63 8
Complement 71 19
Extracellular matrix 270 47
Growth factors 238 50
Immune response 136 28
Inammation 150 28
Lipidcholesterol metabolism 167 20
NF-
B5110
Oxidative mechanisms 166 27
Proliferation 391 73
Signal transduction 2430 412
TNF-
69 8
Transcription factors 557 94
Passerini et al. PNAS
February 24, 2004
vol. 101
no. 8
2483
MEDICAL SCIENCES
by QRT-PCR with that predicted by array analysis. The array
prediction was validated for 60% (1627) of the genes sampled, and
there was disagreement with the array prediction for 18% (527).
For 22% (627) of the gene sample, high interanimal variability
prevented a definitive conclusion (reported as ‘‘undetermined’’ in
Fig. 2) despite highly reproducible measurements for each animal
(n 3). The basis of the interanimal variability for these genes is
unclear, but it may be due to confounding factors other than the
influence of regional hemodynamics, and it may be genetic in origin.
The validation of about two of three genes after linear amplification
is consistent with our previous report, for which we used a model
system (26).
Mining of Biological Pathways. Table 2 shows selected differentially
expressed genes (DF vs. UF) in the biological classifications of
inflammation, cell adhesion, and oxidative mechanisms. Also
shown in Table 2 are the
ARRAYSTAT ratio (mean, n 8) and the
putative positive or negative impact, deduced from the literature,
on mechanisms contributing to atherogenesis (references are pro-
vided as Supporting References to Table 2).
Proinflammatory gene expression in DF.
The differential regulation of
inflammatory and immune-related genes revealed an endothe-
lium primed for inflammation in DF. Consistent with a proin-
flammatory response, there was up-regulation of several broad-
acting cytokines, chemokines, and receptors (interleukin 1
,
interleukin 1 receptor 1, interleukin 6, interleukin 8 receptor
,
advanced glycosylation end-product-specific receptor, monocyte
chemotactic protein 1) as well as down-regulation of the anti-
inflammatory interleukin 10 in DF when compared with UF.
Furthermore, up-regulation of elements of the NF-
B system
(nuclear factor of kappa light polypeptide gene enhancer in
B cells 1 and 2 and NF-
B inhibitor
) and simultaneous
down-regulation of transcription for the I
B kinase-complex-
associated protein involved in assembling an active kinase
complex (33), and the cellular zinc finger anti-NF-
B protein
(Cezanne) which down-regulates NF-
B (34) were noted (Table
2). However, NF-
B was inactive in both DF and UF regions as
shown by cytoplasmic localization and nuclear exclusion of the
NF-
B complex (p65 antibody) (Fig. 3A and B). These obser-
vations are consistent with an NF-
B system primed but inactive
in DF as described by Hajra et al. (22). Differentially expressed
genes that may mitigate an inflammatory tendency included the
up-regulation of endothelial protein C receptor and the down-
regulation of platelet-activating factor receptor, chemokine re-
ceptor 4, complement component C3, several cathepsins, and the
MHC class II antigen HLA-DRB4.
Absence of differences in adhesion molecule expression.
Several adhe-
sion-related molecules were differentially regulated by flow type,
including von Willebrand factor, CD44, fibronectin 1, several
integrins, cadherins, and junction plakoglobin. The up-regulation of
von Willebrand factor (24) and CD44 (25) in particular may have
implications for atherogenesis. However, VCAM-1, intercellular
adhesion molecule 1, E-selectin, and P-selectin, all associated with
early NF-
B-mediated inflammatory responses and the onset of
atherogenesis, were not detected as differentially expressed despite
the increased transcriptional levels of proinflammatory cytokines
and NF-
B pathway components noted above. Because the statis-
tical methods aimed at detecting putative differential expression do
not allow us to make any definitive statement regarding unchanged
genes, we also measured differential expression of these genes by
QRT-PCR (data not shown). There was no clear pattern of
differential expression, with the results indicating slight up-
regulation in some animals, but slight down-regulation or no change
in others. Immunostaining for VCAM-1 was weak and indistin-
guishable between the two regions (data not shown). Collectively,
these observations in the context of normal histology suggest that,
although the endothelium in DF regions may be primed for an
inflammatory response, the process is held in check, possibly at the
level of NF-
B.
Protective antioxidative expression profile in DF.
Reactive oxygen spe-
cies (ROS) and prooxidative pathways are implicated in the initi-
ation and progression of atherosclerosis (35). Furthermore, ROS is
a potent activator of NF-
B (36, 37). Data mining revealed an
antioxidative profile in endothelium located in regions of DF that
may protect against NF-
B-mediated inflammation. Significant
suppression of a component of superoxide-generating NADH
oxidase was noted. An antioxidative state was supported by con-
comitant up-regulation of key antioxidative enzymes, including
GPX3, extracellular superoxide dismutase, NADH quinone oxi-
doreductase, heme oxygenase, and two forms of GST (GST-
1 and
microsomal MGST2). Several of these genes have been reported to
be up-regulated by laminar shear stress in vitro in a response
mediated by a shear-responsive antioxidant response element (38).
Down-regulation of components of the cytochrome c oxidase and
NADH dehydrogenase complexes (intramitochondrial sources of
ROS through their involvement in electron transport), and of
endothelial nitric oxide synthase 3 were noted. Also protective was
the down-regulation of thioredoxin-interacting protein [i.e., vitamin
D
3
-up-regulated protein 1 (VDUP-1)], an inhibitor of thioredoxin
that, in turn, is an important regulator of cell redox balance (39).
Berk and colleagues have shown that shear stress influences
VDUP-1 gene and protein expression (B. Berk; personal commu-
nication). Down-regulation of VDUP-1 leads to increased thiore-
doxin activity and, ultimately, down-regulation of VCAM-1 expres-
sion through the mitogen-activated protein kinaseJun kinase
pathway. Few prooxidative genes were differentially expressed, and
the balance was decisively shifted in the direction of a protective
antioxidative state in DF regions in these healthy animals. The
expression of antioxidative enzymes GPX3 and heme oxygenase
were validated by QRT-PCR (Fig. 2). The prominent up-regulation
of GPX3 in DF was confirmed by immunostaining (Fig. 3 C and D).
The antioxidative expression profile observed here may be critical
to maintaining a delicate balance in gene expression in vulnerable
regions of normal animals.
Fig. 2. Comparison of QRT-PCR results with microarray predictions for selected
genes: red, up-regulated in DF; green, down-regulated in DF; gray, unchanged;
white, undetermined (high interanimal variability).
2484
www.pnas.orgcgidoi10.1073pnas.0305938101 Passerini et al.
Gene expression in other biological classifications.
Although it is not
within the scope of this paper to exhaustively analyze the entire
database of differentially expressed genes, this study resulted in a
very rich dataset, which is now provided as a public resource.
Reference to the online database also revealed a balance of gene
expression related to coagulation mechanisms with a shift toward
protective anticoagulant and profibrinolytic mechanisms in the DF
region. Notably, all components of the plasminogen-activator sys-
tem identified by our analyses were regulated in a manner consis-
tent with antithrombotic mechanisms in DF. Lipidcholesterol
metabolism-related gene expression reflected a balance in DF
regions between the expression of genes, which might promote
cholesterol deposition and reverse cholesterol transport. Numerous
growth factors and receptors were shown to be differentially
regulated. The differential expression of numerous transcription
factors may provide new insight into the modulation of the endo-
thelial response to hemodynamics through the identification of
common regulatory binding elements.
Table 2. Selected genes identied as differentially expressed (DF vs. UF) in the context of a putative role in atherosclerosis
Up-regulated Down-regulated
Gene GenBank ID Ratio Putative effect* Gene GenBank ID Ratio Putative effect*
Inammatoryimmune response
IL-1A X02851 2.01 IL-10 M57627 0.67
IL-1R1 M27492 1.46 IKBKAP AF153419 0.77
IL-6 X04430 1.68 CEZANNE AJ293573 0.66
MCP-1 M37719 1.57 IL-8 M28130 0.62
RAGE AB036432 4.02 CXCR4 AF005058 0.29
IL-8RB M99412 1.65 FOS BC004490 0.43
NFKB1 M58603 1.34 PAFR D10202 0.66
NFKB2 S76638 1.35 C3 J04763 0.27
NFKBIA BC004983 2.76 HLADRB4 M19556 0.55
EPCR L35545 1.90 TRAF6 H12612 0.62
PTGS2 D28235 1.55 IGFBP1 R81994 0.66
IL-14 L15344 0.67 ?
Oxidative mechanisms
GPX3 D00632 5.64 TXNIP S73591 0.44
SOD3 U10116 1.39 COX6B BC001015 0.61
NQO1 BC000474 1.50 COX7A2 AY007643 0.74
POR AF258341 1.36 NDUFC1 AK023115 0.69
HMOX1 X06985 1.66 NDUFS4 AF020351 0.64
GSTT1 AB057594 1.55 NDUFS6 AF044959 0.56
MGST2 U77604 2.83 NDUFS3 AL135819 0.52
NDUFA7 Y16007 1.40 NDUFA3 AF044955 0.42
NDUFB7 AF112200 1.35 NDUFV2 M22538 0.75
NDUFB10 AF088991 1.35 NDUFA6 BC002772 0.76
NDUFS8 U65579 0.70
CYBB X04011 0.46
NOS3 M93718 0.69
SOD1 K00065 0.77
Cell adhesion
VWF X04385 5.04 CTNND1 AF062343 0.59 ?
CD44 AJ251595 1.47 CDH1 Z13009 0.50 ?
CDH3 X63629 1.43 CDH13 L34058 0.65 ?
FN1 AJ276395 2.51 ? ITGA6 X53586 0.43 ?
ITGB4 X51841 1.43 ? ITGA5 BC008786 0.69 ?
MO1A J03270 1.34 ? JUP Z68228 0.19 ?
IL-1A, IL-1
; IL-1R1, IL-1 receptor 1; IL-6, IL-6 (interferon
2); MCP-1, monocyte chemotactic protein 1; RAGE, advanced glycosylation end product-specic
receptor; IL-8RB (CXCR2), IL-8 receptor
; NFKB1, nuclear factor of
light polypeptide gene enhancer in B cells 1 (p105); NFKB2, nuclear factor of
light
polypeptide gene enhancer in B cells 2 (p49p100); NFKBIA, NF-
B inhibitor
; EPCR, endothelial protein C receptor; PTGS2, prostaglandin endoperoxide
synthase-2 (COX-2); IKBKAP, inhibitor of
light polypeptide gene enhancer in B-cells, kinase complex-associated protein; CEZANNE, cellular zinc nger
anti-NF-
B; CXCR4, chemokine (C-X-C motif) receptor 4; FOS, v-fos FBJ murine osteosarcoma viral oncogene; PAFR, platelet-activating factor receptor; C3,
complement component C3; HLADRB4, MHC class II HLA-DRw53-beta (DR4,w4); TRAF6, tumor necrosis factor receptor-associated factor 6; IGFBP1, insulin-like
growth factor binding protein 1; GPX3, glutathione peroxidase 3; SOD3, superoxide dismutase 3; NQO1, NADH quinone oxidoreductase; POR, cytochrome P450
oxidoreductase; HMOX1, heme oxygenase (decycling) 1; GSTT1, GST
1; MGST2, microsomal GST 2; NDUFA7, NADH ubiquinone oxidoreductase; NDUFB7, NADH
dehydrogenase (ubiquinone) component B7; NDUFB10, NADH dehydrogenase (ubiquinone) component B10; TXNIP, thioredoxin interacting protein (VDUP1);
COX6B, cytochrome c oxidase subunit VIb; COX7A2, subunit VIIa; NDUFC1, NADH dehydrogenase (ubiquinone) component C1; NDUFS4, component S4; NDUFS6,
component S6; NDUFS3, component S3; NDUFA3, component A3; NDUFV2, component V2; NDUFA6, component A6; NDUFS8, component S8; CYBB, cytochrome
b-245 beta (NADPH oxidase component p91 phox); NOS3, endothelial nitric oxide synthase; SOD1, superoxide dismutase 1; VWF, von Willebrand factor; CD44,
cell adhesion molecule (CD44); CDH3, cadherin 3 (P-cadherin); FN1, bronectin 1; ITGB4, integrin
4; MO1A, leukocyte adhesion glycoprotein Mo1
; CTNND1,
catenin
1; CDH1, cadherin 1 (E-cadherin); CDH13, cadherin 13 (H-cadherin); ITGA6, integrin
6; ITGA5, integrin
5(bronectin receptor); JUP, junction
plakoglobin.
, pro-atherosclerotic; , anti-atherosclerotic; ?, unknown.
*References to the putative effect provided as Supporting References to Table 2, which is published as supporting information on the PNAS web site.
Vascular cell adhesion molecule 1 (VCAM-1), intercellular adhesion molecule 1, P-selectin, E-selectin were not signicantly different.
Passerini et al. PNAS
February 24, 2004
vol. 101
no. 8
2485
MEDICAL SCIENCES
Discussion
Although the localization of atherosclerotic lesions to predict-
able regions in mammalian arteries has been recognized for over
a century, compelling evidence implicating the local hemody-
namics is more recent (6). We have exploited RNA amplification
to demonstrate the heterogeneity of endothelial phenotypes
from hemodynamically distinct regions in vivo, demonstrating
that a delicate balance of pro- and antiatherosclerotic mecha-
nisms may exist simultaneously in endothelium of lesion-prone
sites of the adult porcine aorta to create a setting of vulnerability
to atherogenesis. Atherosclerosis has been described as a chronic
inflammatory disease involving many cell types and complex
cross-talk between them (40). Many interrelated physiological
mechanisms are recruited very early in the process, including cell
adhesion, oxidative metabolism, lipid metabolism, apoptosis,
innate and adaptive immunity, and coagulation (recently re-
viewed in refs. 4046). The predisposition for expression of key
molecules in multiple pathways is therefore likely to be impor-
tant in determining the susceptibility of a site to lesion initiation
when additional risk factors are prevalent, thereby influencing
the threshold or timing for atherogenesis.
NF-
B is a transcription factor proposed to play a central role as
a mediatorintegrator of atherogenesis (36). Its association with
inhibitory I
Bs masks the nuclear localization sequence, retaining
the complex in the cytoplasm. Activation of I
B kinase results in
phosphorylation of I
Bs and their targeting for degradation by the
proteosome, releasing NF-
B dimers for translocation to the nu-
cleus, where they activate transcription in target genes. These genes
include cytokines, chemokines, adhesion molecules, and genes
involved in cell proliferation and cell survival. NF-
B is controlled
by the redox state of the cell, which may be a common signaling
pathway leading to NF-
B activation through ROS (37). Activation
of NF-
B and NF-
B-regulated genes has been observed in ath-
erosclerotic plaques (47), and NF-
B-activated by shear stress has
been observed in cultured cells (48, 49) and regions of DF (22).
Collins et al. (36) have suggested that the survivalprotective genes
induced by NF-
B limit the inflammatory response at low levels,
although a strong challenge results in the expression of adhesion
moleculescytokines.
In one of the few in vivo regional arterial studies, Hajra et al. (22)
evaluated NF-
B regulation in regions of high and low probability
for atherosclerosis in mouse aorta by using en face immunostaining
and confocal microscopy. They demonstrated preferential activa-
tion of NF-
B (translocation to the nucleus) and induction of
NF-
B-responsive gene expression in high probability regions of
the aortas of lipopolysaccharide-treated or hypercholester-
olemic LDLR
/
mice. In normal mice, however, although p65 and
I
Bs were elevated in such regions, NF-
B activation remained
low, suggesting that the pathway was primed but not activated.
Our findings are consistent with and greatly extend such an
interpretation.
Several recent reports have used microarrays to assess gene
expression in EC exposed to unidirectional steady flow in vitro (9,
13, 20) and compared expression profiles to those of control cells
in no-flow conditions. The results, however, are not easily com-
pared with our study, which captures the profile of two distinct
regions long-adapted to differential hemodynamic environments in
vivo. More relevant are two in vitro microarray studies that com-
pared flow conditions analogous to DF and UF.
In the first, Garcia-Cardena et al. (16) compared differential
expression in cultured human umbilical vein endothelial cells
between turbulent shear stress (TSS) and LSS, analogous to DF vs.
UF regions in vivo in this study. One hundred genes were identified
as differentially expressed (TSS vs. LSS) after 24 h (68 up-regulated
and 32 down-regulated). The authors identified a set of genes that
was down-regulated in LSS but up-regulated in TSS as potentially
the most pathologically relevant. They considered highly regulated
genes of known or putative function in signaling, response to injury,
or atherogenesis and focused on linking the protective effects of
LSS to matrix biology and cell cycle. Brooks et al. (15) compared
gene expression in cultured human aortic EC under pulsatile DF
conditions with LSS by using microarrays (588 genes) and subtrac-
tion cloning. They identified 100 genes as differentially expressed
at 24 h, many of which were up-regulated genes associated
with mechanisms known to be proatherosclerotic, particularly
inflammatory molecules, adhesion factors, and oxidation-related
molecules.
Few genes identified as differentially expressed by these in vitro
studies were common to those identified by our in vivo study.
Furthermore, there was little agreement between the in vitro
studies. The reasons may be the different nature of the DF in each
study, different cell origins, sensitivities of microarray analysis, or a
myriad of technical considerations. Despite significantly different
outcomes, however, both of these in vitro studies demonstrated that
conditions of DF resulted in significant differential endothelial gene
expression compared with conditions of unidirectional steady flow.
The methods used here have refined endothelial transcriptional
profiling to regions small enough to represent exposure to discrete
flow characteristics in vivo, but these methods also provide suffi-
cient cells to amplify RNA and profile gene expression with
reasonable confidence. We have previously evaluated the linear
amplification protocol by using similar amounts of human matched
probetarget and found it to identify differential expression with
fidelity and enhanced sensitivity (26). In agreement with this
previous study, the QRT-PCR results reported in the current study
corroborate approximately two of three genes tested. When con-
sidering differential expression with cross-species hybridization,
differences in sequence homology are expected to lead to fewer
genes hybridizing than for the equivalent same-species hybridiza-
tion, thus contributing to false negatives. However, both samples
should hybridize in the same way so that false positives will not be
Fig. 3. En face immunostaining in DF and UF regions. (A and B) Subcellular
localization of NF-
B complex (p65 antibody) was visualized by 2D projection of
confocal imaging through the EC layer. The staining was predominantly cyto-
plasmic with nuclear exclusion in both aortic regions. Also clearly illustrated are
the EC morphologies characteristic of DF (A) and UF (B). (C and D) Epiuorescence
microscopy for the prominent differentially expressed antioxidative enzyme
GPX3 revealed strong expression in DF (C) in contrast to low expression in UF (D),
consistent with microarray predictions.
2486
www.pnas.orgcgidoi10.1073pnas.0305938101 Passerini et al.
affected. The linear amplification protocol minimizes bias possibly
introduced by other amplification methods, but it is clear that
certain transcripts may be reproducibly misrepresented. Although
we verified EC purity at 99%, we cannot rule out the possibility
that a small number of contaminating SMC could lead to some false
positives with the use of amplification. Given the increased sensi-
tivity of amplification, it is possible that strongly differentially
expressed genes in a small number of contaminating cells could be
detected.
Inflammatory responses leading to atherogenesis in DF regions
may share common signaling pathways as responses mediated by
cytokines, such as tumor necrosis factor (TNF)-
. In particular, the
two may share common mechanisms mediated by NF-
B. We have
previously profiled EC gene expression in response to TNF-
in
vitro (26), and we find the overlap with the present in vivo study to
be small. Up-regulated genes common to both studies included
monocyte chemotactic protein 1, interleukin 1
, NF-
B1, NF-
B2,
NF-
B inhibitor
, TNF-
-induced protein 6, and cycloxygenase-2,
whereas common down-regulated genes included TNF receptor-
associated factor 6 and the chemokine receptor CXCR4. An
important distinction between these studies is that differential
expression of VCAM-1, intercellular adhesion molecule 1, and
E-selectin was induced by TNF-
, but not by differential hemody-
namics in the normal animal.
Because this is the first in vivo study profiling regional endothelial
gene expression on a large scale, we have attempted to bridge the
findings to published candidate gene studies in vivo (2124) and
experiments of flow disturbance in vitro (8, 15, 16, 19, 23). Fur-
thermore, we have presented specific observations relating to some
pathways that are central to atherogenesis. Further mining of these
data with emphasis on different biological pathways (such as those
highlighted in Table 1) will provide additional insights into the
regional susceptibility to atherosclerosis.
In summary, a baseline is established for gene expression pro-
filing by using endothelium from DF and UF regions in the normal
adult pig aorta, which defines in vivo the existing relationships
between many classes of molecules responsible for vascular ho-
meostasis and implicated in early atherogenesis. This study revealed
distinctly different patterns of gene expression than reported by in
vitro studies to date. The hemodynamic characteristics of DF may
prime the endothelium toward inflammation (and by inference,
atherogenesis) but protective profiles in gene expression (e.g., a net
antioxidative profile) prevent disease initiation in the absence of
additional risk factors.
We thank Drs. Garret FitzGerald, Aron Fisher, and Paul Janmey of the
University of Pennsylvania for critical reading of the manuscript, Re-
becca Riley for expert technical assistance, and Dr. Richard Magid for
help with NF-
B imaging. This work was supported by National Insti-
tutes of Health Grants HL62250, HL70128, K25-HG-02296, and K25-
HG-00052; National Space Biomedical Research Institute (National
Aeronautics and Space Administration) Grant NSBRI-01-102; and a
Sponsored Research Award from AstraZeneca Pharmaceuticals.
1. Cornhill, J. F. & Roach, M. R. (1976) Atherosclerosis 23, 489501.
2. Glagov, S., Zarins, C., Giddens, D. P. & Ku, D. N. (1988) Arch. Pathol. Lab. Med.
112, 10181031.
3. Wissler, R. W. (1994) Atherosclerosis 108, S3S20.
4. Davies, P. F., Reidy, M. A., Goode, T. B. & Bowyer, D. E. (1976) Atherosclerosis
25, 125130.
5. Faggiotto, A., Ross, R. & Harker, L. (1984) Arteriosclerosis 4, 323340.
6. Gimbrone, M. A., Jr., Anderson, K. R., Topper, J. N., Langille, B. L., Clowes,
A. W., Bercel, S., Davies, M. G., Stenmark, K. R., Frid, M. G., Weiser-Evans, M. C.,
et al. (1999) J. Vasc. Surg. 29, 11041151.
7. Nagel, T., Resnick, N., Atkinson, W. J., Dewey, C. F., Jr., & Gimbrone, M. A., Jr.
(1994) J. Clin. Invest. 94, 885891.
8. Topper, J. N., Cai, J., Falb, D. & Gimbrone, M. A., Jr. (1996) Proc. Natl. Acad. Sci.
USA 93, 1041710422.
9. Chen, B. P., Li, Y. S., Zhao, Y., Chen, K. D., Li, S., Lao, J., Yuan, S., Shyy, J. Y.
& Chien, S. (2001) Physiol. Genomics 7, 5563.
10. De Keulenaer, G. W., Chappell, D. C., Ishizaka, N., Nerem, R. M., Alexander,
R. W. & Griendling, K. K. (1998) Circ. Res. 82, 10941101.
11. Bongrazio, M., Baumann, C., Zakrzewicz, A., Pries, A. R. & Gaehtgens, P. (2000)
Cardiovasc. Res. 47, 384393.
12. Passerini, A. G., Milsted, A. & Rittgers, S. E. (2003) J. Vasc. Surg. 37, 182190.
13. McCormick, S. M., Eskin, S. G., McIntire, L. V., Teng, C. L., Lu, C. M., Russell,
C. G. & Chittur, K. K. (2001) Proc. Natl. Acad. Sci. USA 98, 89558960.
14. Wasserman, S. M., Mehraban, F., Komuves, L. G., Yang, R. B., Tomlinson, J. E.,
Zhang, Y., Spriggs, F. & Topper, J. N. (2002) Physiol. Genomics 12, 1323.
15. Brooks, A. R., Lelkes, P. I. & Rubanyi, G. M. (2002) Physiol. Genomics 9, 2741.
16. Garcia-Cardena, G., Comander, J., Anderson, K. R., Blackman, B. R. & Gim-
brone, M. A., Jr. (2001) Proc. Natl. Acad. Sci. USA 98, 44784485.
17. Lin, K., Hsu, P. P., Chen, B. P., Yuan, S., Usami, S., Shyy, J. Y., Li, Y. S. & Chien,
S. (2000) Proc. Natl. Acad. Sci. USA 97, 93859389.
18. DePaola, N., Davies, P. F., Pritchard, W. F., Jr., Florez, L., Harbeck, N. & Polacek,
D. C. (1999) Proc. Natl. Acad. Sci. USA 96, 31543159.
19. Topper, J. N., Cai, J., Qiu, Y., Anderson, K. R., Xu, Y. Y., Deeds, J. D., Feeley,
R., Gimeno, C. J., Woolf, E. A., Tayber, O., et al. (1997) Proc. Natl. Acad. Sci. USA
94, 93149319.
20. Peters, D. G., Zhang, X. C., Benos, P. V., Heidrich-OHare, E. & Ferrell, R. E.
(2002) Physiol. Genomics 12, 2533.
21. Iiyama, K., Hajra, L., Iiyama, M., Li, H., DiChiara, M., Medoff, B. D. & Cybulsky,
M. I. (1999) Circ. Res. 85, 199207.
22. Hajra, L., Evans, A. I., Chen, M., Hyduk, S. J., Collins, T. & Cybulsky, M. I. (2000)
Proc. Natl. Acad. Sci. USA 97, 90529057.
23. De Nigris, F., Lerman, L. O., Ignarro, S. W., Sica, G., Lerman, A., Palinski, W.,
Ignarro, L. J. & Napoli, C. (2003) Proc. Natl. Acad. Sci. USA 100, 14201425.
24. Methia, N., Andre, P., Denis, C. V., Economopoulos, M. & Wagner, D. D. (2001)
Blood 98, 14241428.
25. Cuff, C. A., Kothapalli, D., Azonobi, I., Chun, S., Zhang, Y., Belkin, R., Yeh, C.,
Secreto, A., Assoian, R. K., Rader, D. J. & Pure, E. (2001) J. Clin. Invest. 108,
10311040.
26. Polacek, D. C., Passerini, A. G., Shi, C., Francesco, N. M., Manduchi, E., Grant,
G. R., Powell, S., Bischof, H., Winkler, H., Stoeckert, C. J., Jr., & Davies, P. F.
(2003) Physiol. Genomics 13, 147156.
27. Van Gelder, R. N., von Zastrow, M. E., Yool, A., Dement, W. C., Barchas, J. D.
& Eberwine, J. H. (1990) Proc. Natl. Acad. Sci. USA 87, 16631667.
28. Nadon, R. & Shoemaker, J. (2002) Trends Genet. 18, 265271.
29. Tusher, V. G., Tibshirani, R. & Chu, G. (2001) Proc. Natl. Acad. Sci. USA 98,
51165121.
30. Benjamini, Y. & Hockberg, Y. (1995) J. R. Stat. Soc. B 57, 289300.
31. Manduchi, E., Grant, G. R., He, H., Liu, J., Mailman, M. D., Pizarro, A. D.,
Whetzel, P. L. & Stoeckert, C. J., Jr. (2004) Bioinformatics, in press.
32. Kilner, P. J., Yang, G. Z., Mohiaddin, R. H., Firmin, D. N. & Longmore, D. B.
(1993) Circulation 88, 22352247.
33. Cohen, L., Henzel, W. J. & Baeuerle, P. A. (1998) Nature 395, 292296.
34. Evans, P. C., Taylor, E. R., Coadwell, J., Heyninck, K., Beyaert, R. & Kilshaw, P. J.
(2001) Biochem. J. 357, 617623.
35. Griendling, K. K., Sorescu, D. & Ushio-Fukai, M. (2000) Circ. Res. 86, 494501.
36. Collins, T. & Cybulsky, M. I. (2001) J. Clin. Invest. 107, 255264.
37. Li, N. & Karin, M. (1999) FASEB J. 13, 11371143.
38. Chen, X. L., Varner, S. E., Rao, A. S., Grey, J. Y., Thomas, S., Cook, C. K.,
Wasserman, M. A., Medford, R. M., Jaiswal, A. K. & Kunsch, C. (2003) J. Biol.
Chem. 278, 703711.
39. Schulze, P. C., De Keulenaer, G. W., Yoshioka, J., Kassik, K. A. & Lee, R. T.
(2002) Circ. Res. 91, 689695.
40. Steinberg, D. (2002) Nat. Med. 8, 12111217.
41. Binder, C. J., Chang, M. K., Shaw, P. X., Miller, Y. I., Hartvigsen, K., Dewan, A.
& Witztum, J. L. (2002) Nat. Med. 8, 12181226.
42. Ruggeri, Z. M. (2002) Nat. Med. 8, 12271234.
43. Li, A. C. & Glass, C. K. (2002) Nat. Med. 8, 12351242.
44. Repa, J. J. & Mangelsdorf, D. J. (2002) Nat. Med. 8, 12431248.
45. Dzau, V. J., Braun-Dullaeus, R. C. & Sedding, D. G. (2002) Nat. Med. 8,
12491256.
46. Libby, P. & Aikawa, M. (2002) Nat. Med. 8, 12571262.
47. Brand, K., Page, S., Rogler, G., Bartsch, A., Brandl, R., Knuechel, R., Page, M.,
Kaltschmidt, C., Baeuerle, P. A. & Neumeier, D. (1996) J. Clin. Invest. 97,
17151722.
48. Lan, Q., Mercurius, K. O. & Davies, P. F. (1994) Biochem. Biophys. Res. Commun.
201, 950956.
49. Khachigian, L. M., Resnick, N., Gimbrone, M. A., Jr., & Collins, T. (1995) J. Clin.
Invest. 96, 11691175.
Passerini et al. PNAS
February 24, 2004
vol. 101
no. 8
2487
MEDICAL SCIENCES
... Alignment is accompanied by the expression of anti-inflammatory genes and 'healthy' EC function [56]. In contrast, ECs exposed to turbulent 'atherogenic' flow do not align, show pro-inflammatory characteristics and 'unhealthy' function [56,57]. Regions of turbulent flow in vessels are the major sites of atherosclerosis, the pathological process that narrows blood vessels, reducing oxygen and nutrient delivery to tissues. ...
Article
Full-text available
Neuropilin-1 (NRP1) is a transmembrane glycoprotein expressed by several cell types including, neurons, endothelial cells (ECs), smooth muscle cells, cardiomyocytes and immune cells comprising macrophages, dendritic cells and T cell subsets. Since NRP1 discovery in 1987 as an adhesion molecule in the frog nervous system, more than 2300 publications on PubMed investigated the function of NRP1 in physiological and pathological contexts. NRP1 has been characterised as a coreceptor for class 3 semaphorins and several members of the vascular endothelial growth factor (VEGF) family. Because the VEGF family is the main regulator of blood and lymphatic vessel growth in addition to promoting neurogenesis, neuronal patterning, neuroprotection and glial growth, the role of NRP1 in these biological processes has been extensively investigated. It is now established that NRP1 promotes the physiological growth of new vessels from pre-existing ones in the process of angiogenesis. Furthermore, several studies have shown that NRP1 mediates signalling pathways regulating pathological vascular growth in ocular neovascular diseases and tumour development. Less defined are the roles of NRP1 in maintaining the function of the quiescent established vasculature in an adult organism. This review will focus on the opposite roles of NRP1 in regulating transforming growth factor β signalling pathways in different cell types, and on the emerging role of endothelial NRP1 as an atheroprotective, anti-inflammatory factor involved in the response of ECs to shear stress.
... Although at first surprising, this hypothesis is supported by several lines of evidence. First, accumulating evidence suggests that CD44 might be an early indicator of risk for atherogenesis [32][33][34] . CD44 alters gene expression in these areas prior to lesion development and even in the absence of any difference in the cholesterol levels 35 . ...
Article
Full-text available
The decline of endothelial autophagy is closely related to vascular senescence and disease, although the molecular mechanisms connecting these outcomes in vascular endothelial cells (VECs) remain unclear. Here, we identify a crucial role for CD44, a multifunctional adhesion molecule, in controlling autophagy and ageing in VECs. The CD44 intercellular domain (CD44ICD) negatively regulates autophagy by reducing PIK3R4 and PIK3C3 levels and disrupting STAT3-dependent PtdIns3K complexes. CD44 and its homologue clec-31 are increased in ageing vascular endothelium and Caenorhabditis elegans, respectively, suggesting that an age-dependent increase in CD44 induces autophagy decline and ageing phenotypes. Accordingly, CD44 knockdown ameliorates age-associated phenotypes in VECs. The endothelium-specific CD44ICD knock-in mouse is shorter-lived, with VECs exhibiting obvious premature ageing characteristics associated with decreased basal autophagy. Autophagy activation suppresses the premature ageing of human and mouse VECs overexpressing CD44ICD, function conserved in the CD44 homologue clec-31 in C. elegans. Our work describes a mechanism coordinated by CD44 function bridging autophagy decline and ageing.
... Steady and oscillatory shear stress affect the endothelial transcriptome, methylome, and proteome [1][2][3][4][5]. The context of previous studies interrogating the effects of shear stress exerted by disturbed flow is the pathogenesis of atherosclerosis; establishing a link between atheroprone regions of the vasculature and separated flow, which is characterized by low, oscillatory shear stress [6]. ...
Article
Full-text available
Cerebral aneurysms are more likely to form at bifurcations in the vasculature, where disturbed fluid is prevalent due to flow separation at sufficiently high Reynolds numbers. While previous studies have demonstrated that altered shear stress exerted by disturbed flow disrupts endothelial tight junctions, less is known about how these flow regimes alter gene expression in endothelial cells lining the blood–brain barrier. Specifically, the effect of disturbed flow on expression of genes associated with cell–cell and cell–matrix interaction, which likely mediate aneurysm formation, remains unclear. RNA sequencing of immortalized cerebral endothelial cells isolated from the lumen of a 3D blood–brain barrier model reveals distinct transcriptional changes in vessels exposed to fully developed and disturbed flow profiles applied by both steady and physiological waveforms. Differential gene expression, validated by qRT-PCR and western blotting, reveals that lumican, a small leucine-rich proteoglycan, is the most significantly downregulated gene in endothelial cells exposed to steady, disturbed flow. Knocking down lumican expression reduces barrier function in the presence of steady, fully developed flow. Moreover, adding purified lumican into the hydrogel of the 3D blood–brain barrier model recovers barrier function in the region exposed to fully developed flow. Overall, these findings emphasize the importance of flow regimes exhibiting spatial and temporal heterogeneous shear stress profiles on cell–matrix interaction in endothelial cells lining the blood–brain barrier, while also identifying lumican as a contributor to the formation and maintenance of an intact barrier.
Article
Full-text available
Ascending thoracic aortic aneurysms may be fatal upon rupture or dissection and remain a leading cause of death in the developed world. Understanding the pathophysiology of the development of ascending thoracic aortic aneurysms may help reduce the morbidity and mortality of this disease. In this review, we will discuss our current understanding of the protective relationship between ascending thoracic aortic aneurysms and the development of atherosclerosis, including decreased carotid intima–media thickness, low-density lipoprotein levels, coronary and aortic calcification, and incidence of myocardial infarction. We also propose several possible mechanisms driving this relationship, including matrix metalloproteinase proteins and transforming growth factor-β.
Article
Linear and disturbed flow differentially regulate gene expression, with disturbed flow priming endothelial cells (ECs) for a proinflammatory, atheroprone expression profile and phenotype. Here, we investigated the role of the transmembrane protein neuropilin-1 (NRP1) in ECs exposed to flow using cultured ECs, mice with an endothelium-specific knockout of NRP1, and a mouse model of atherosclerosis. We demonstrated that NRP1 was a constituent of adherens junctions that interacted with VE-cadherin and promoted its association with p120 catenin, stabilizing adherens junctions and inducing cytoskeletal remodeling in alignment with the direction of flow. We also showed that NRP1 interacted with transforming growth factor-β (TGF-β) receptor II (TGFBR2) and reduced the plasma membrane localization of TGFBR2 and TGF-β signaling. NRP1 knockdown increased the abundance of proinflammatory cytokines and adhesion molecules, resulting in increased leukocyte rolling and atherosclerotic plaque size. These findings describe a role for NRP1 in promoting endothelial function and reveal a mechanism by which NRP1 reduction in ECs may contribute to vascular disease by modulating adherens junction signaling and promoting TGF-β signaling and inflammation.
Article
Coronary artery disease is the leading cause of morbidity and mortality worldwide, especially in developed countries, with an increasing incidence in developing countries. Despite the ad-vances in cardiology, there are yet many unanswered questions about the natural history of cor-onary atherosclerosis. However, it has not been fully explained why some coronary artery plaques remain quiescent over time, whereas others evolve to a high-risk, "vulnerable" plaque with a predisposition to destabilize and induce a cardiac event. Furthermore, approximately half of the patients with acute coronary syndromes demonstrate no prior symptoms of ischemia or angiographically evident disease. Recent findings have indicated that apart from cardiovas-cular risk factors, genetics, and other unknown factors, local hemodynamic forces, such as en-dothelial shear stress, blood flow patterns, and endothelial dysfunction of the epicardial and microvascular coronary arteries, are associated with the progression of coronary plaque and the development of cardiovascular complications with complex interactions. In this review article, we summarize the mechanisms that affect coronary artery plaque progression, indicating the importance of endothelial shear stress, endothelial dysfunction of epicardial and microvascular vessels, inflammation, and their complex associations, underlying in parallel the clinical per-spectives of these findings.
Article
In dysfunctional arteriovenous fistulae (AVF) for hemodialysis access, neointimal hyperplasia (NH) is prone to occur in the region exposed to disturbed flow. We hypothesized that disturbed flow contributes to NH in AVF by inducing endothelial mesenchymal transition (EndMT) through activation of osteopontin/CD44 axis. In rat with aortocaval fistula, a rodent model of AVF, we demonstrated development of EndMT and expression of osteopontin and CD44 specifically in the vicinity of arteriovenous junction using immunostaining. Duplex scan confirmed this region was exposed to a disturbed flow. A mixed ultrastructural phenotype of endothelium and smooth muscle cells was found in luminal endothelial cells of the arteriovenous junction by electron microscopy ascertaining the presence of EndMT. Endothelial lineage tracing using Cdh5-Cre/ERT2;ROSA26-tdTomato transgenic mice showed that EndMT was involved in NH of AVF since early stage and that the endothelial-derived cells contributed to 24% of neointimal cells. In human umbilical vein endothelial cells (HUVECs) in culture, osteopontin treatment induced EndMT, which was suppressed by CD44 knockdown. Exposure to low oscillatory wall shear stress using a parallel-plate system induced EndMT in HUVECs; also suppressed by osteopontin or CD44 knockdown. In AVF of CD44 knockout mice, EndMT was mitigated and NH decreased by 35% compared to that in wild-type mice. In dysfunctional AVF of patients with uremia, expressions of osteopontin, CD44, and mesenchymal markers in endothelial cells overlying the neointima was also found by immunostaining. Thus, the osteopontin/CD44 axis regulates disturbed flow-induced EndMT, plays an important role in neointimal hyperplasia of AVF, and may act as a potential therapeutic target to prevent AVF dysfunction.
Article
Full-text available
Modulation of endothelial cell behavior and phenotype by hemodynamic forces involves many signaling components, including cell surface receptors, intracellular signaling intermediaries, transcription factors, and epigenetic elements. Many of the signaling mechanisms that underlie mechanotransduction by endothelial cells are inadequately defined. Here we sought to better understand how β‐arrestins, intracellular proteins that regulate agonist‐mediated desensitization and integration of signaling by transmembrane receptors, may be involved in the endothelial cell response to shear stress. We performed both in vitro studies with primary endothelial cells subjected to β‐arrestin knockdown, and in vivo studies using mice with endothelial specific deletion of β‐arrestin 1 and β‐arrestin 2. We found that β‐arrestins are localized to primary cilia in endothelial cells, which are present in subpopulations of endothelial cells in relatively low shear states. Recruitment of β‐arrestins to cilia involved its interaction with IFT81, a component of the flagellar transport protein complex in the cilia. β‐arrestin knockdown led to marked reduction in shear stress response, including induction of NOS3 expression. Within the cilia, β‐arrestins were found to associate with the type II bone morphogenetic protein receptor (BMPR‐II), whose disruption similarly led to an impaired endothelial shear response. β‐arrestins also regulated Smad transcription factor phosphorylation by BMPR‐II. Mice with endothelial specific deletion of β‐arrestin 1 and β‐arrestin 2 were found to have impaired retinal angiogenesis. In conclusion, we have identified a novel role for endothelial β‐arrestins as key transducers of ciliary mechanotransduction that play a central role in shear signaling by BMPR‐II and contribute to vascular development. This article is protected by copyright. All rights reserved.
Article
Full-text available
This study was designed to elucidate the mechanism underlying the inhibition of endothelial cell growth by laminar shear stress. Tumor suppressor gene p53 was increased in bovine aortic endothelial cells subjected to 24 h of laminar shear stress at 3 dynes (1 dyne = 10 μN)/cm2 or higher, but not at 1.5 dynes/cm2. One of the mechanisms of the shear-induced increase in p53 is its stabilization after phosphorylation by c-Jun N-terminal kinase. To investigate the consequence of the shear-induced p53 response, we found that prolonged laminar shear stress caused increases of the growth arrest proteins GADD45 (growth arrest and DNA damage inducible protein 45) and p21cip1, as well as a decrease in phosphorylation of the retinoblastoma gene product. Our results suggest that prolonged laminar shear stress causes a sustained p53 activation, which induces the up-regulation of GADD45 and p21cip1. The resulting inhibition of cyclin-dependent kinase and hypophosphorylation of retinoblastoma protein lead to endothelial cell cycle arrest. This inhibition of endothelial cell proliferation by laminar shear stress may serve an important homeostatic function by preventing atherogenesis in the straight part of the arterial tree that is constantly subjected to high levels of laminar shearing.
Article
Full-text available
The heterogeneity of neural gene expression and the spatially limited expression of many low-abundance messenger RNAs in the brain has made cloning and analysis of such messages difficult. To generate amounts of nucleic acids sufficient for use in standard cloning strategies, we have devised a method for producing amplified heterogeneous populations of RNA from limited quantities of cDNA. Whole cerebellar RNA was primed with a synthetic oligonucleotide containing the T7 RNA polymerase promoter sequence 5' to a polythymidylate region. After second-strand cDNA synthesis, T7 RNA polymerase was used to generate amplified antisense RNA (aRNA). Up to 80-fold molar amplification has been achieved from nanogram quantities of cDNA. The amplified material is similar in size distribution to the parent cDNA and shows sequence heterogeneity as assessed by Southern and Northern blot analysis. Specific messages for moderate-abundance mRNAs for actin and guanine nucleotide-binding protein (G-protein) alpha subunits have been detected in the amplified material. By using in situ transcription to generate cDNA, sequences for cyclophilin have been detected in aRNA derived from single cerebellar tissue sections. cDNA derived from a single cerebellar Purkinje cell also has been amplified and yields material that hybridizes to cognate whole RNA and mRNA but not to Escherichia coli RNA.
Article
The following extended abstracts were presented at the Research Initiatives in Vascular Disease Conference, Movers and Shakers in the Vascular Tree—Hemodynamic and Biomechanical Factors in Blood Vessel Pathology, sponsored by The Lifeline Foundation and the Cardiovascular & Interventional Radiology Research and Educational Foundation; jointly sponsored by the International Society for Cardiovascular Surgery, North American Chapter, The Society for Vascular Surgery, and The Society of Cardiovascular and Interventional Radiology; in cooperation with the National Institutes of Health–National Heart, Lung &Blood Institute on Mar 11–12, 1999, in Bethesda, Md.
Article
Objective: Shear stress is known to modulate gene expression. However, the molecular link between blood flow and long-time vessel adaptation is still unclear. In this study, the variations of gene expression by prolonged shear stress exposure was investigated in order to identify genes possibly involved in flow dependent vascular adaptation. Methods: Human umbilical vein endothelial cells (HUVECs) were exposed to laminar shear stress (6 dyn/cm(2); 24 h) and analyzed by differential display (DDRT-PCR). Flow-modulation of differentially expressed genes by different exposure times (4, 24, 48 h) and in human cardiac microvascular endothelial cells (HCMECs) (24 h exposure) was analyzed by RT-PCR and northern blotting. Results: DDRT-PCR analysis displayed 13 down- and 20 up-regulated products in response to flow. Four known genes were identified: Angiopoietin-2, a protein reported to reduce vessel stability, was progressively (4-48 h) down-regulated by shear stress. The induction of the anti-angiogenic metalloproteinase METH-1 was maximal after 4 h exposure and sustained over the time (24-48 h). Growth arrest-specific mRNA 3 (gas3) and calpactin 1 light chain (p11) were up-regulated only by prolonged exposure (24-48 h). Analysis of the expression of angiopoietin-2, METH-1, gas3, and p11 in shear stress exposed (24 h) HCMECs showed modulation patterns comparable to those observed in HUVECs. Conclusion: Since angiopoietin-2 and METH-1 are known to be involved in vessel regression/stabilization, the reported modulation of these genes by prolonged shear stress exposure strongly suggests their participation in flow-dependent vascular adaptation.
Article
A quantitative study of the size and location of early sudanophilic lesions has been carried out on the aortae of five rabbits. The atherosclerotic lesions were induced by feeding an average of 114 (+/- 3.7) egg yolks over an average period of 83 (+/- 1.7) days. The aortic lesions were visualuized by gross staining with Sudan III and measured by the polar coordinate method. The lesions were almost entirely around orifices: their size was directly proportional to the area of the associated ostium (P less than 0.005). The sudanophilic deposits were located downstream from the ostia in areas believed to experience high shear stresses. The area of the intercostal ostia increased as one proceeded down the thoracic aorta (P less than 0.005). A deviation from the distal distribution pattern was observed where local flow and shear stresses were disturbed. The coronary lesions completely surrounded the ostia, the coeliac lesions had significant proximal components and the left renal and inferior mesenteric lesions were skewed to the right. The study suggests that hemodynamic forces and particularly high shear stress is of considerable importance in both the initiation and localization of early atherosclerotic lesions.
Article
The luminal surface of fatty lesions of atherosclerosis was viewed by scanning electron microscopy (SEM). Endothelial cells were outlined by staining intercellular junctions with silver and the aortas were fixed in situ at physiological pressure. When aortas were dehydrated by passage through organic solvents followed by critical point drying from liquid CO2, there was considerable disruption of the luminal surface and it was not possible to correctly interpret the morphological integrity of the endothelium. In contrast, simple air-drying of aortas, without solvent dehydration after fixation, allowed the integrity of the cell layer overlying the lesion to be evaluated. The success of this technique was attributed to the retention of arterial lipids during dehydration of the tissue.
Article
Atherosclerosis affects the major elastic and muscular arteries, but some vessels are largely spared while others may be markedly diseased. The carotid bifurcation, the coronary arteries, the infrarenal abdominal aorta, and the vessels supplying the lower extremities are at highest risk. The propensity for plaque formation at bifurcations, branchings, and curvatures has led to conjectures that local mechanical factors such as wall shear stress and mural tensile stress potentiate atherogenesis. Recent studies of the human vessels at high risk, and of corresponding models, have provided quantitative evidence that plaques tend to occur where flow velocity and shear stress are reduced and flow departs from a laminar, unidirectional pattern. Such flow characteristics tend to increase the residence time of circulating particles in susceptible regions while particles are cleared rapidly from regions of relatively high wall shear stress and laminar unidirectional flow. The flow patterns associated with plaque localization are most prominent during systole. Long-term consequences are therefore likely to be greatly enhanced by elevated heart rate and may exert a selective effect on the coronary arteries. The point-by-point redistribution of wall tension at regions of geometric transition has not been quantitatively related to plaque localization. Enlargement of arteries as plaques increase in size and the associated modeling of plaque and wall configuration tend to preserve an adequate and regular lumen cross section. Hemodynamic forces appear to determine changes in vessel diameter so as to restore normal levels of wall shear stress, while wall thickness architecture, and composition are closely related to tensile stress. Hemodynamic forces may also be implicated in the symptom-producing destabilization of plaques, especially in relation to wall instabilities near stenoses. The relative roles of wall shear stress, tensile stress, and the metabolism of the artery wall in the progression and complication of atherosclerosis remain to be clarified. Development of clinical techniques for relating hemodynamic and tensile properties to plaque location, stenosis, and composition should permit pathologists to provide new insights into the bases for the topographic and individual differences in plaque progression and outcome.
Article
Morphologic studies resulting from events that occur during the development of the lesions of atherosclerosis were studied in chronic, diet-induced hypercholesterolemia in a series of nonhuman primates. Within 12 days of hypercholesterolemia in Macaca nemestrina, monocytes became adherent to the surface of the endothelium. These monocytes appeared to migrate subendothelially, accumulate lipid, and become lipid-laden macrophages (foam cells). Within a month, a "serofibrinous insudate" formed together with variable numbers of subendothelial lipid-laden macrophages. By the second month, foam cells increased in number, often in multilayers, to form a fatty streak. Concomitantly, the luminal surface of the arteries became increasingly irregular due to the subendothelial accumulation of foam cells. Numerous monocytes continued to attach to the endothelial surface over the fatty streaks, and many of them appeared to enter the intima and participate in the growth of the fatty streaks. Lipid-laden smooth muscle cells appeared in small numbers and formed two to four layers between the macrophages and the internal elastic lamella at 2 to 3 months. During the third month of hypercholesterolemia, endothelial cell continuity over the lipid-laden macrophages became interrupted, exposing the underlying foam cells to circulating blood. Foam cells were then readily observed in whole blood smears, suggesting that many of the lipid-laden macrophages leave the intima and enter the circulation. After 4 months, significant endothelial denudation was found in the iliac artery and many exposed macrophages were covered by adherent platelets in the form of a mural thrombus. Thus, the early components of atherosclerosis induced by chronic hypercholesterolemia centered around the monocyte-macrophage and its interaction with endothelium in the induction of the fatty streak. Subsequent changes that lead to macrophage-smooth muscle interactions, platelet-macrophage interactions, and platelet-endothelial interactions appeared to set the stage for the development of more advanced proliferative lesions.
Article
Hemodynamic forces induce various functional changes in vascular endothelium, many of which reflect alterations in gene expression. We have recently identified a cis-acting transcriptional regulatory element, the shear stress response element (SSRE), present in the promoters of several genes, that may represent a common pathway by which biomechanical forces influence gene expression. In this study, we have examined the effect of shear stress on endothelial expression of three adhesion molecules: intercellular adhesion molecule-1 (ICAM-1), which contains the SSRE in its promoter, and E-selectin (ELAM-1) and vascular cell adhesion molecule-1 (VCAM-1), both of which lack the SSRE. Cultured human umbilical vein endothelial cells, subjected to a physiologically relevant range of laminar shear stresses (2.5-46 dyn/cm2) in a cone and plate apparatus for up to 48 h, showed time-dependent but force-independent increases in surface immunoreactive ICAM-1. Upregulated ICAM-1 expression was correlated with increased adhesion of the JY lymphocytic cell line. Northern blot analysis revealed increased ICAM-1 transcript as early as 2 h after the onset of shear stress. In contrast, E-selectin and vascular cell adhesion molecule-1 transcript and cell-surface protein were not upregulated at any time point examined. This selective regulation of adhesion molecule expression in vascular endothelium suggests that biomechanical forces, in addition to humoral stimuli, may contribute to differential endothelial gene expression and thus represent pathophysiologically relevant stimuli in inflammation and atherosclerosis.