SlideShare a Scribd company logo
1 of 14
Download to read offline
A rt i c l e s
628	 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine
Atherosclerosis is a disease of chronic inflammation and is the leading
cause of morbidity and mortality worldwide. There is a general con-
sensus that the majority of coronary syndromes result from the rupture
of unstable plaques and associated thrombotic events1–3. Plaque
instability has been associated with disruption of the fibrous cap, an
atheroprotective layer of smooth muscle α-actin (ACTA2)-positive
cells that cover the atherosclerotic plaque3–7; large numbers of cells
positive for macrophage markers such as LGALS3; and the presence
of a large foam cell–laden necrotic core within the plaque2–4,8. Indeed,
these data have been interpreted as evidence that plaques that contain
a high ratio of macrophages relative to SMCs are less stable, particu-
larly those that have a thin ACTA2+ fibrous cap that is presumed to be
composed primarily of phenotypically modulated SMCs9,10.
Although ACTA2+ and LGALS3+ cells are assumed to be of SMC
and myeloid lineage respectively, there is extensive ambiguity about
the lineal origin of cells within atherosclerotic plaques11. These ambi-
guities were originally based on in vitro studies showing that SMCs
downregulate SMC markers and activate macrophage markers after
cholesterol loading12, and that macrophages activate SMC genes after
treatment with factors known to be present within lesions, including
thrombin13. However, the most compelling evidence that SMCs and
macrophages are often misidentified within human advanced coronary
lesions comes from studies of cross-gender bone marrow transplant
subjects. These studies have shown that >10% of ACTA2+ cells within
lesions are of hematopoietic stem cell (HSC) and not SMC origin14.
Consistent with these human data, a substantial fraction of the cells
that express SMC markers (including ACTA2 but not MYH11) within
lesions of Apoe−/− mice are of HSC and not SMC origin15.
Conversely, there is extensive evidence suggesting that many
SMC-derived cells within advanced lesions of Apoe−/− mice lack
detectable expression of conventional SMC markers such as ACTA2
(ref. 16), and/or activate expression of macrophage markers17. Notably,
in vivo studies from our laboratory showed that large numbers of
ACTA2−, myosin, heavy polypeptide 11, smooth muscle (MYH11)−,
and transgelin (TAGLN)− cells within advanced lesions of Apoe−/−
Western diet–fed mice retain expression of a mutant Tagln-lacZ trans-
gene that is resistant to downregulation compared to the unmutated
transgene16. Unfortunately, these studies are not definitive because
we could not rule out the possibility that non-SMCs present within
lesions may activate the mutant transgene.
1Robert M. Berne Cardiovascular Research Center, University of Virginia, Charlottesville, Virginia, USA. 2Department of Molecular Physiology and Biological Physics,
University of Virginia, Charlottesville, Virginia, USA. 3Department of Surgery, University of Virginia, Charlottesville, Virginia, USA. 4Department of Biochemistry
and Molecular Genetics, University of Virginia, Charlottesville, Virginia, USA. 5Department of Pathology, University of Virginia, Charlottesville, Virginia, USA.
6Intercollegiate Faculty of Biotechnology, University of Gdansk, Gdansk, Poland. 7Department of Pharmacology and Chemical Biology, University of Pittsburgh,
Pittsburgh, Pennsylvania, USA. 8Division of Biology and Biomedical Sciences, Washington University, St. Louis, Missouri, USA. Correspondence should be addressed
to G.K.O. (gko@virginia.edu).
Received 5 February; accepted 22 April; published online 18 May 2015; corrected after print 12 August 2015; doi:10.1038/nm.3866
KLF4-dependent phenotypic modulation of smooth
muscle cells has a key role in atherosclerotic plaque
pathogenesis
Laura S Shankman1,2, Delphine Gomez1, Olga A Cherepanova1, Morgan Salmon3, Gabriel F Alencar1,4,
Ryan M Haskins1,5, Pamela Swiatlowska1,6, Alexandra A C Newman1,4, Elizabeth S Greene1, Adam C Straub7,
Brant Isakson1,2, Gwendalyn J Randolph8 & Gary K Owens1,2
Previous studies investigating the role of smooth muscle cells (SMCs) and macrophages in the pathogenesis of atherosclerosis
have provided controversial results owing to the use of unreliable methods for clearly identifying each of these cell types. Here,
using Myh11-CreERT2 ROSA floxed STOP eYFP Apoe−/− mice to perform SMC lineage tracing, we find that traditional methods
for detecting SMCs based on immunostaining for SMC markers fail to detect >80% of SMC-derived cells within advanced
atherosclerotic lesions. These unidentified SMC-derived cells exhibit phenotypes of other cell lineages, including macrophages
and mesenchymal stem cells (MSCs). SMC-specific conditional knockout of Krüppel-like factor 4 (Klf4) resulted in reduced
numbers of SMC-derived MSC- and macrophage-like cells, a marked reduction in lesion size, and increases in multiple indices
of plaque stability, including an increase in fibrous cap thickness as compared to wild-type controls. On the basis of in vivo KLF4
chromatin immunoprecipitation–sequencing (ChIP-seq) analyses and studies of cholesterol-treated cultured SMCs, we identified
>800 KLF4 target genes, including many that regulate pro-inflammatory responses of SMCs. Our findings indicate that the
contribution of SMCs to atherosclerotic plaques has been greatly underestimated, and that KLF4-dependent transitions in SMC
phenotype are critical in lesion pathogenesis.
npg©2015NatureAmerica,Inc.Allrightsreserved.
a rt i c l e s
nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015	 629
Thus, despite decades of atherosclerosis research, we still do not
know which cells within lesions are SMC-derived and to what extent
they contribute to lesion pathogenesis. A recent SMC lineage tracing
study using a Tagln-ERT2Cre-lacZ Apoe−/− mouse model provided
evidence that SMC-derived cells within advanced lesions activate
some macrophage markers, including LGALS3 and CD68 (ref. 17).
Unfortunately, as highlighted in an editorial on this paper18, the
labeling efficiency of SMCs in this study was only 11%, preclud-
ing a determination of the fraction of macrophage-like cells within
lesions that are derived from SMCs or, most importantly, how these
cells might contribute to lesion pathogenesis. Another recent study19
showed that 50% of foam cells within advanced human coronary
artery lesions express the SMC marker ACTA2, highlighting the
magnitude of the ‘SMC/macrophage misidentification problem’ with
respect to our understanding of human disease. However, the majority
of these ACTA2+ foam cells also expressed the macrophage marker
CD68 and represented 40% of all CD68+ lesional cells19. Given clear
evidence that macrophages can activate SMC markers and, vice versa,
that SMCs can activate macrophage markers, it is unclear whether
these CD68+ lesional cells are derived from SMCs, macrophages, or
another cell type.
In view of the major ambiguities in identifying which cells within
atherosclerotic lesions are SMC-derived versus macrophage-derived,
the most crucial questions are: (i) how are the phenotypic transi-
tions of SMCs, macrophages, and other cell types regulated within
lesions; (ii) what is the function of these phenotypically modulated
cells; and (iii) how do these phenotypic transitions affect overall
disease pathogenesis? To begin to address these questions, we gener-
ated atherosclerosis-prone Apoe−/− mice with which we could lineage-
trace SMCs and study the effects of SMC-specific conditional
knockout of the stem cell pluripotency gene, KLF4. We chose to
study the role of KLF4 as we and others have previously shown that
this transcription factor plays a key role in regulating phenotypic
transitions of SMCs in vivo during development20 and after carotid
ligation injury21, as well as in vitro in cultured SMCs treated with
platelet-derived growth factor (PDGF)-BB22,23, PDGF-DD24 or
oxidized phospholipids25.
RESULTS
Most atherosclerotic plaque SMCs are not identified by ACTA2
SMCs are distinguished from other cell types by expression of a
unique repertoire of genes including Acta2, Tagln, and Myh11. These
genes are coordinately downregulated, at least in vitro, during SMC
phenotypic switching such that they may be undetectable using tra-
ditional immunohistochemical staining methods11,26. Therefore, to
rigorously analyze the overall contributions of SMCs to lesion patho-
genesis, we used a previously described27Myh11-CreERT2 ROSA
floxed STOP eYFP Apoe−/− (SMC YFP+/+Apoe−/−) mouse model
in which >95% of medial SMCs within large arteries were labeled
with YFP (Supplementary Fig. 1a). Because Cre excision is per-
manent, this SMC lineage-tracing model provides permanent YFP
lineage tagging of virtually all mature (MYH11+) arterial SMCs that
exist at the time of tamoxifen injection, allowing study of the fur-
ther differentiation of these cells or of their progeny, irrespective of
continued expression of Acta2, Myh11, or other SMC marker genes.
To ensure the fidelity and SMC specificity of this lineage tracing
model, we completed a number of further validation studies beyond
those shown in our previous studies27. We demonstrated (i) SMC-
specific YFP labeling within all tissue specimens examined, with
no detectable expression of the YFP indicator gene in the absence
of tamoxifen (Supplementary Figs. 1a,b and 2a); (ii) no detectable
YFP+ cells by flow cytometry within blood or bone marrow prepara-
tions (Supplementary Fig. 2a); (iii) no evidence of YFP+ cells within
lesions of Western diet–fed mice not given tamoxifen (Supplementary
Fig. 2b); (iv) no detectable YFP+ cells in the blood of the mice
when fed a high-fat diet for 18 weeks (Supplementary Fig. 2c); and
(v) YFP+ labeling of approximately 60% of freshly enzymatically
dissociated cells from the aorta (the ascending and descending
thoracic aorta plus the abdominal aorta to the iliac bifurcation,
including adventitial and intimal cells) on the basis of flow cytometric
analysis (Supplementary Fig. 2d).
We harvested brachiocephalic arteries (BCAs) from SMC YFP+/+
Apoe−/− mice that had been injected with tamoxifen between 6–8 weeks
of age, after an additional 18 weeks of Western diet. BCAs were
immunostained for YFP, ACTA2, and LGALS3. Confocal micros-
copy z-stacks of the collected tissues were acquired and analyzed to
accurately profile individual cells (Fig. 1a–e). Notably, ~82% of SMCs
within atherosclerotic lesions (YFP+DAPI+ cells) were ACTA2 nega-
tive (Fig. 1a and Supplementary Table 1), indicating that the majority
of SMCs within the lesion cannot be identified using traditional SMC
markers. These results also showed that phenotypically modulated
SMCs (YFP+ACTA2−DAPI+) comprised approximately 30% of the
total cells within lesions (Fig. 1a and Supplementary Table 1), which
far exceeds previous estimates based on ACTA2 immunostaining.
SMCs within atherosclerotic plaques express markers of
macrophages, MSCs, and myofibroblasts
We found that phenotypically modulated (YFP+) SMCs within lesions
expressed markers of macrophages (LGALS3) (Fig. 1c), MSCs (SCA1)
(Fig. 1d), and myofibroblasts (ACTA2 and PDGFβR) (Fig. 1e). From
these data we estimate the following distribution of SMC-derived cells
within lesions: 30% macrophage-like cells (YFP+ACTA2−LGALS3+),
7% MSC-like cells (YFP+ACTA2−SCA1+), 12% myofibroblast-like
cells (YFP+ACTA2+PDGFβR+), and 32–51% an indeterminate cell
phenotype (YFP+ACTA2−LGALS3−SCA1−) (Supplementary Table 1).
In addition, these analyses showed that 36% of LGALS3+ cells
within advanced atherosclerotic lesions were YFP+, indicating that
approximately one-third of cells that would normally be classified as
macrophages in most previous studies in the field actually originated
from SMCs, rather than from myeloid cells as previously assumed.
These initial studies were performed in paraffin-embedded sam-
ples to maintain the ultrastructure of the plaque and to determine
the location of various SMC-derived cells within lesions. However,
this technique limits the number of markers that can be simultane-
ously examined. To further characterize SMC-derived plaque cells,
we performed flow cytometric analyses of freshly dissociated cells
from the aorta, from the aortic root through the iliac bifurcation.
We found that substantial numbers of SMC-derived cells expressed
multiple additional macrophage and hematopoietic markers (Fig. 1f).
In particular, we identified YFP+ cells that co-expressed the monocyte/
macrophage marker ITGAM (CD11b)) and the mature macrophage
marker F4/80, as well as YFP+ cells that co-expressed ITGAM and
the dendritic cell marker ITGAX (CD11c). In addition, we found
that 13% of the YFP+ cells co-expressed the MSC markers SCA1 and
ENG (CD105) (Fig. 1f). When we analyzed MSC-like cells using
traditional negative-gating strategies (CD45−CD34−CDH5−),
we found that up to 45% of MSC-like cells within the aortas of
18-week Western diet–fed SMC YFP+/+Apoe−/− mice were YFP+
(Supplementary Fig. 3a), although it is unclear whether all of these
cells were located within lesions or whether they also contributed to
npg©2015NatureAmerica,Inc.Allrightsreserved.
A rt i c l e s
630	 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine
the population of adventitial SCA1+ cells that have previously been
described by several groups28–32. Gating strategies for all flow cytom-
etry experiments were determined on the basis of fluorescence minus
one (FMO) controls (Supplementary Fig. 3b,c).
To further assess the morphological and possible functional prop-
erties of SMC-derived macrophage-like cells in vivo, we analyzed
BCA lesions from SMC YFP+/+Apoe−/− mice by transmission elec-
tron microscopy, combined with detection of YFP expression by
immunogold labeling. As shown in Figure 2a and Supplementary
Figure 4, we identified YFP+ cells containing multiple large lipid
vacuoles that seem to be phagocytic. However, these cells were rela-
tively rare, possibly reflecting the low frequency of phagocytic YFP+
cells at any given instant in time, the transient nature of this process,
and/or technical difficulties in detecting these cells by immuno-
electron microscopy. In addition, we flow-sorted SMC-derived
MSC-like cells (YFP+ENG+SCA1+), non-SMC-derived MSC-like
cells (YFP−ENG+SCA1+) and SMC-derived non-MSC-like cells
(YFP+ENG−SCA1−) from 18-week Western diet–fed lineage-tracing
mice to test the ability of these cell populations to differentiate into
multiple lineages, including adipocytes and osteoblasts. After two
passages in MSC maintenance medium, the SMC non-MSC-like cells
became unhealthy in appearance and died (data not shown). The
SMC-derived MSC-like cells survived but seemed to be senescent
(Fig. 2b) and grew slowly (data not shown). These cells also failed
to differentiate into either adipocytes (Fig. 2c,d) or osteoblasts (data
not shown) when exposed to the appropriate differentiation culture
medium. In contrast, the YFP− (non-SMC-derived) MSC-like cells
grew well (data not shown) and showed high-efficiency differentiation
into adipocytes (Fig. 2c,d) and osteoblasts (data not shown). These
data indicate that although a subset of SMC-derived cells within
atherosclerotic lesions express multiple markers of MSCs, these cells
do not appear to be pluripotent and thus may not have the functional
properties of MSCs.
SMCs within human atheromas express the macrophage
marker CD68
To independently detect phenotypically modulated SMCs that express
LGALS3 and to validate a method for detecting these cells in human
lesions, we used an in situ hybridization proximity ligation assay (ISH-
PLA) recently developed by our laboratory27. This technique permits
the identification of phenotypically modulated SMCs within fixed
tissues based on the detection of the histone H3K4diMe in the Myh11
promoter (PLA+), a SMC-specific epigenetic signature that persists in
cells that have no detectable expression of SMC markers27,33. We first
validated this method by showing that YFP+LGALS3+ SMCs within
the lineage-tracing mice retained this SMC-specific epigenetic signa-
ture (Supplementary Fig. 5a). We also showed that neither cultured
RAW 264-7 mouse macrophage cells (Supplementary Fig. 5b) nor
human monocytes (Supplementary Fig. 5c) exposed to POVPC, an
oxidative product of LDL that activates monocytes/macrophages34,
exhibited H3K4diMe in the Myh11 promoter.
To determine whether SMC transition to a macrophage-like state
occurs in human lesions, we stained human coronary artery atheroscle-
rotic lesions for CD68 and ACTA2 and performed ISH-PLA detection
of MYH11 H3K4diMe. Multiple human coronary artery lesion sections
from 12 human subjects were analyzed (Supplementary Fig. 5d). In
these lesions, 18% of CD68+ cells were PLA+ (Fig. 3a–c), indicating that
they are of SMC origin. To further validate these findings, we performed
the ISH-PLA analysis in coronary artery samples from a male who
had received a cross-gender heart transplant (Supplementary Fig. 6).
Merge + DAPI
YFPACTA2
LGALS3
c
Merge + DAPISCA1
YFPACTA2
d
YFP
Merge + DAPIPDGFβR
ACTA2
e
ACTA2 YFP
MergeDAPI
ba
60K
40K
20K
0
10
0
10
1
10
2
YFP
SSarea
60K
40K
20K
0
0 20K 40K 60K
FSlin
FS area
60K
40K
20K
0
0 20K 40K 60K
SSarea
F4/80
10
1
10
0
10
2
10
3
10
4
10
1
10
0
10
2
10
3
10
4
10
1
10
0
10
2
10
3
10
4
ITGAM
ITGAX
ITGAM
ENG
SCA1
CD34
10
1
10
0
10
2
10
3
10
4
PTPRC
10
3
10
4
10
0
10
1
10
2
10
3
10
4
10
0
10
1
10
2
10
3
10
4
10
0
10
1
10
2
10
3
10
4
10
0
10
1
10
2
10
3
10
4
10
0
10
1
10
2
10
3
10
4
10
0
10
1
10
2
10
3
10
4
60K
40K
20K
0
LGALS3
SSArea
60K
40K
20K
0
CDH5
SSArea
f
FS area
Figure 1  Lineage tracing provides evidence for large populations of
phenotypically modulated SMCs within lesions. (a–e) Representative
immunofluorescence staining of BCAs from SMC YFP+/+Apoe−/− mice
fed a Western diet for 18 weeks. The boxed region in a is shown at higher
magnification in b. Shown are phenotypically modulated SMCs
(YFP+ACTA2− cells highlighted in the white rectangle, b), macrophage-like
SMCs (YFP+ACTA2−LGALS3+, c), mesenchymal stem cell-like SMCs
(YFP+ACTA2−SCA1+, d), and myofibroblast-like cells (YFP+ACTA2+PDGFβR+, e).
Yellow arrows in c–e indicate de-differentiated (YFP+ACTA2−) SMCs; white
arrows indicate differentiated (YFP+ACTA2+) SMCs. Samples were either fixed
and embedded in paraffin (a–c, e) or in Neg-50 frozen section medium (d).
Scale bars, 50 µm in a and 10 µm in b–e. (f) Flow cytometry of single cell
suspensions from aortas of SMC YFP+/+Apoe−/− mice treated with tamoxifen
between 6–8 weeks and then fed a western diet for 18 weeks. Top, cells were
gated for forward scatter (FS) area versus side scatter (SS) area to eliminate
debris; for ‘singlets’ (single cells) based on a plot of forward scatter linear
(FS lin) versus FS area; and for YFP. Middle, sub-populations of YFP+ cells
were found to be double-positive for ITGAM (CD11b) and F4/80, for ITGAM
(CD11b) and ITGAX (CD11c) and single-positive for LGALS3 (mac2). Bottom,
after negative gating for CD34−PTPRC− and CDH5 cells, YFP+ cells were
found to be double-positive for SCA1 and ENG. n = 6 mice per group.
npg©2015NatureAmerica,Inc.Allrightsreserved.
a rt i c l e s
nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015	 631
WefoundMYH11H3K4diMePLA+CD68+ cellsthatwereYchromosome–
negative (Fig.3d), consistent with the notion that these macrophage-like
cells are of SMC and not hematopoietic origin. Notably, we never
found MYH11 H3K4diMe PLA+ cells that were Y chromosome–
positive (Fig. 3d and data not shown), demonstrating that myeloid
cells do not acquire the MYH11 H3K4diMe SMC epigenetic signature
in human atherosclerotic lesions.
KLF4 has a critical role in regulating SMC phenotype and
plaque pathogenesis
We have previously shown that KLF4, a cell pluripotency factor35, is
required for SMC phenotypic switching in several in vitro models25,36.
However, there is as yet no evidence that SMC phenotypic transitions
within atherosclerotic lesions are KLF4 dependent, and, if so, what
role these transitions have in lesion pathogenesis. Consistent with our
hypothesis that KLF4 regulates phenotypic transitions of SMCs within
atherosclerotic lesions, we observed large numbers of YFP+ cells that
expressed KLF4 within BCA lesions from 18-week Western diet–fed
SMC YFP+/+Apoe−/− mice (Supplementary Fig. 7a). To determine
whether KLF4 regulates SMC phenotypic transitions and overall lesion
pathogenesis, we generated SMC YFP+/+Apoe−/− mice with SMC-
specific deficiency of KLF4 (SMC YFP+/+Klf4∆/∆Apoe−/− mice) by cross-
ing SMC YFP+/+Apoe−/− mice with Klf4fl/fl mice. We observed a high
level of recombination of the floxed Klf4 alleles (Klf4∆/∆) tamoxifen
treatment (Supplementary Fig. 7b,c), including what we estimate to
be nearly 100% recombination within SMCs in the aorta when the
data are corrected for the >40% of non-SMC DNA present in these
samples, based on flow cytometric analysis of aortic cell populations
(Supplementary Fig. 2b). Figure 4a shows representative confocal
immunofluorescence images of the BCA lesions of control mice (SMC
YFP+/+Klf4+/+Apoe−/−) and mice with loss of KLF4 in SMCs (SMC
YFP+/+Klf4∆/∆Apoe−/−) after being fed a Western diet for 18 weeks.
SMC YFP+/+Klf4∆/∆Apoe−/− mice had a nearly 50% reduction in
plaque size (Fig. 4b) and multiple changes consistent with increased
plaque stability, including a more-than-twofold increase in fibrous cap
area (Fig. 4c), an increase in the percentage of ACTA2+ cells within
the fibrous cap (Fig. 4d), and a reduced percentage of LGALS3+ cells
(Fig. 4e) as compared to control SMC YFP+/+Klf4+/+Apoe−/− mice.
SMC lineage-tracing analyses showed that loss of KLF4 within SMCs
did not result in a change in the overall number of SMCs (YFP+ cells)
within the lesions (Supplementary Fig. 8a), but it had major effects
on SMC phenotypic transitions. These effects included a 53% decrease
in the percentage of macrophage-like SMCs (YFP+LGALS3+/YFP+)
within a lesion (Fig. 4e), a 70% decrease in the percentage of MSC-like
SMCs (YFP+SCA1+/YFP+) within the medial area underlying lesions
(Supplementary Fig. 8b), but no change in the percentage of MSC-like
SMCsinthelesionitself(SupplementaryFig.8c)inSMCYFP+/+Klf4∆/∆
Apoe−/− mice as compared to control SMC YFP+/+Klf4+/+Apoe−/−
mice. Consistent with these results, flow cytometric analyses showed
a decrease in the percentage of SMC-derived MSC-like cells (YFP+
SCA1+ENG+CDH5−PTPRC−CD34−) (Supplementary Fig. 8d), but
no change in either the overall percentage of MSCs (Supplementary
Fig. 8e) or the percentage of YFP+ cells (Supplementary Fig. 8f)
inSMCYFP+/+Klf4∆/∆Apoe−/− miceascomparedtocontrolSMC YFP+/+
Klf4+/+Apoe−/− mice. As noted above, SMC specific Klf4-knockout
mice showed an increase in the percentage of ACTA2+ cells within the
fibrous cap (Fig. 4d) as compared to wild-type control mice. Similarly,
SMC YFP+/+Klf4∆/∆Apoe−/− mice showed an increase in the percentage
of ACTA2+ cells within lesions (Fig. 4f), but reduced proliferation of
SMC-derived cells (Fig. 4g) and a marked reduction in YFP+ SMC
apoptosis as compared to control YFP+/+Klf4+/+Apoe−/− mice (Fig. 4h).
These effects were not associated with changes in medial or luminal
area (Supplementary Fig. 8g), or in the percentage of cells that were
YFP+PDGFβR+ (Supplementary Fig. 8h) or YFP+ACTA2+ (Fig. 4f).
In addition, we did not observe changes in cholesterol or triglyceride
levels (Supplementary Fig. 8i).
KLF4 modulates phenotypic transitions and functional
properties of SMCs
We have previously presented evidence that Klf4 is induced in cul-
tured SMCs by treatment with oxidized phospholipids36 and that
KLF4 suppresses expression of SMC marker genes through several
mechanisms, including binding to the G/C repressor element found
in most SMC marker gene promoters (including Acta2, Tagln, and
Myh11) and inhibiting binding of the transcription factor SRF to
CArG elements16,37,38. To determine whether similar mechanisms
contribute to suppression of SMC marker gene expression within
atherosclerotic lesions in vivo, we performed chromatin immuno-
precipitation (ChIP) assays on chromatin extracted from the BCA
regions of Apoe−/− mice carrying a transgene containing either an
YFP
+
ENG
+
SCA1
+
YFP
–
ENG
+
SCA1
+
YFP
YFP
–
ENG
+
SCA1
+
YFP+
ENG+
SCA1+
DAPI
FABP4
b c
60
d
YFP–
YFP
+
%Differentiation
50
40
30
20
10
0
Adipocyte
differentiation
*
aFigure 2  Ultrastructural and functional characteristics of phenotypically
modulated SMCs. (a) Immuno-transmission election microscopy of a
BCA from a SMC YFP+/+Apoe−/− mouse fed a Western diet. A 10-nm gold
bead–conjugated secondary antibody was used to reveal lipid laden YFP+
cells engulfing neighboring cells. The boxed area in the left micrograph
is progressively enlarged in the center and right micrographs to enable
visualization of the immunogold beads, highlighted with yellow arrows.
6–10 electron microscopic sections from three SMC YFP+/+Apoe−/−
lineage-tracing mice were examined in these analyses. Additional
micrographs from this experiment are shown in Supplementary Figure 4.
Scale bars (from left to right): 2 µm, 0.5 µm, 0.2 µm. (b) Isolated YFP+
and YFP− MSCs (SCA1+ENG+) were cultured in mesenchymal stem cell
medium. Representative images 10 d after plating are shown (green, YFP).
Results are representative of two independent experiments using cells
derived from the BCA region of a total of 10 Myh11-eYFP mice and sorted
into the respective populations indicated. Scale bars, 50 µm. (c) YFP+ and
YFP− MSCs (SCA1+ENG+) were incubated in adipogenesis differentiation
medium and stained for the adipocyte marker FABP4. Scale bar, 200 µm.
(d) Quantification of adipogenesis from five fields of view per group.
Results are representative of two independent experiments. Error bars
are means ± s.d. *P < 0.05 by Student’s t-test.
npg©2015NatureAmerica,Inc.Allrightsreserved.
A rt i c l e s
632	 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine
unmutated Tagln promoter driving lacZ expression, or a Tagln pro-
moter with a mutation of the GC repressor element driving lacZ
expression. We found that compared to chow-fed Apoe−/− mice,
Western diet–fed Apoe−/− mice showed marked enrichment of
KLF4 binding to the Acta2, Tagln, Myh11, and Cnn1 endogenous
promoters (Fig. 5a). Using the two transgenic strains described
above, we showed that enhanced KLF4 binding to the Tagln pro-
moter under Western diet conditions was dependent on the G/C
repressor element (Fig. 5b). Finally, using the ISH-PLA assay, we
showed that KLF4 bound to the Tagln promoter in individual phe-
notypically modulated SMCs (YFP+ACTA2− cells) within lesions
of Apoe−/− mice (Fig. 5c). Taken together, results provide compel-
ling evidence that coordinated suppression of SMC marker gene
expression is mediated by direct binding of KLF4 to the promoters
of SMC marker genes.
We21,23 and others39–42 have shown that KLF4 can act as either
a transcriptional repressor or activator, depending on the cell type
and gene locus. To more fully define the repertoire of KLF4 target
genes that mediate SMC phenotypic switching, we performed KLF4
ChIP-seq analysis on chromatin samples derived from BCAs isolated
from SMC YFP+/+Klf4+/+Apoe−/− versus SMC YFP+/+Klf4∆/∆Apoe−/−
mice. This analysis required pooling of BCA samples from 16 mice
per group to obtain sufficient DNA. Notably, we identified 869 KLF4
target genes that were selectively enriched in Klf4+/+Apoe−/− versus
Klf4∆/∆Apoe−/− Western diet–fed mice (Fig. 5d and Supplementary
Table 2). These putative KLF4 target genes that were selectively bound
by KLF4 in SMCs include the SMC marker genes Acta2 and Tagln,
thereby validating the fidelity of our ChIP-seq analysis. In addition,
we found evidence of enriched KLF4 binding in genes within regula-
tory pathways that are likely to be important in the pathogenesis of
lesions, including gene families associated with phagocytosis, apop-
tosis, cell migration, and inflammation (Fig. 5d and Supplementary
Table 2). These putative KLF4 target genes probably contributed to
the beneficial effects of SMC-specific loss of KLF4 on lesion size and
plaque pathogenesis. KLF4 also bound regions near Itgal (CD11a),
Itgax (CD11b), Itgam (CD11c) (Fig. 5d and Supplementary Table 2),
and Arg1 (data not shown) in Klf4+/+ but not Klf4∆/∆ mice. We
also noted 459 KLF4 targets enriched in Klf4∆/∆Apoe−/− samples as
compared to Klf4+/+Apoe−/− samples (Supplementary Fig. 9 and
Supplementary Table 3), presumably representing KLF4 target genes
in non-SMCs that were altered as a secondary consequence of the loss
of KLF4 in SMCs.
iii
a b c
i
ii
iii
i DAPI
DAPI/ACTA2/PLA DAPI/ACTA2/PLA/CD68
PLA
CD68
d
DAPI
DAPI/PLA/Y-Chr/CD68
DAPI/PLA/Y-Chr DAPI/PLA/Y-Chr/CD68
DAPI/PLA/Y-Chr
Y-Chr DAPI/PLA/Y-Chr
DAPI/PLA/Y-Chr/CD68
PLA
CD68
ACTA2
ii
50
40
30
%ofpopulation
20
10
C
D
68+
PLA+
/C
D
68+
0
Actual count
Corrected count
C
D
68+
PLA+
AC
TA2+
C
D
68+
PLA+
Figure 3  SMCs within human coronary artery lesions express
the macrophage marker CD68. SMCs within advanced
atherosclerotic lesion specimens were identified by PLA
detection of the SMC-specific stable epigenetic signature
H3K4diMe in the MYH11 promoter. MYH11 H3K4diMe
PLA+ cells exhibit a punctate red dot within the nucleus;
the non-nuclear amorphous red staining is autofluorescence
or nonspecific background. (a) Samples treated for PLA (red)
were immunostained for CD68 (green), and DAPI (blue).
Three distinct cell populations were identified, as shown
at higher magnification and indicated with white arrows:
(i) PLA+CD68− SMCs, (ii) PLA−CD68+ cells (macrophages
of hematopoietic origin), and (iii) PLA+CD68+ SMC-derived
macrophage-like cells. Scale bar, 100 µm. (b) Samples
treated for PLA (red) were immunostained for ACTA2 (green),
CD68 (cyan) and DAPI (blue). Shoulder regions within plaques
showed a high incidence of SMC-derived macrophage-like cells
(PLA+CD68+) (yellow arrows). Phenotypically modulated SMCs negative for
CD68 (PLA+ACTA2−CD68−) were also observed in these regions (white arrows).
Scale bars, 50 µm. (c) Quantitative analysis of SMC-derived macrophage-like cells
within human coronary lesions without or with correction for the efficiency of PLA
detection27. n = 12 human atherosclerotic right coronary arteries. Error bars are
means ± s.e.m. (d) Combined epigenetic SMC and genetic HSC lineage tracing
analysis of a cross-gender human heart transplant sample. Coronary artery specimens
from a male patient who had received a female heart were processed for PLA (red), Y chromosome (Y-Chr) FISH (green), and CD68 staining (yellow).
A PLA+ Y chromosome− CD68+ cell (yellow arrows) represents an SMC-derived macrophage-like cell not of hematopoietic origin. In contrast, a PLA−Y
chromosome+CD68+ cell (red arrows) represents a macrophage of hematopoietic origin. A large number of cells of hematopoietic origin that are negative
for CD68 (PLA−Y chromosome+CD68−) were also observed (white arrows). Scale bars, 50 µm.
npg©2015NatureAmerica,Inc.Allrightsreserved.
a rt i c l e s
nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015	 633
Given our observation that SMC-specific loss of Klf4 is associ-
ated with a marked reduction in the percentage of SMC-derived
macrophage-like cells in Klf4∆/∆ mice as compared to wild-type
Klf4+/+ control mice (Fig. 4f), we tested whether KLF4 is required
for transition of cultured SMCs to a macrophage-like state in vitro
following cholesterol loading12. Owing to recent controversies
concerning the origin and purity of SMCs cultured from mouse
vessels43,44, we used SMCs obtained by sorting freshly isolated YFP+
aortic cells from our SMC lineage-tracing mice. Primary cultures of
aortic SMCs harvested from 9-week-old tamoxifen-injected SMC
YFP+/+ Apoe−/− mice that had been injected with tamoxifen between
6–8 weeks of age were >98% YFP+ (Supplementary Fig. 10a), indi-
cating that these cultured SMCs are derived from mature differen-
tiated SMCs in vivo and not from a stem cell source, as has been
speculated43,44. Cholesterol loading of cultured SMCs resulted in
increased expression of Lgals3 (Fig. 6a) and Klf4 (Supplementary
Fig. 10b) as compared to vehicle-loaded control cells; the increase in
Lgals3 expression was abolished in aortic SMCs derived from SMC
YFP+/+Klf4∆/∆ mice (Fig. 6a). Cholesterol loading was also associ-
ated with increased phagocytic behavior of the cells as compared to
vehicle-loaded control cells that was KLF4 dependent (Fig. 6b,c).
Finally, cholesterol loading resulted in KLF4-dependent increases in
the expression of the pro-inflammatory cytokines MCP1, CXCR1,
STNFR1 (Supplementary Fig. 10c–e), Lgals3 mRNA (Fig. 6a),
and the MSC markers Sca1 and Eng (Supplementary Fig. 10f,g) as
compared to vehicle-loaded control cells. In contrast, the increase
in Abca1 mRNA expression caused by cholesterol loading was not
KLF4 dependent (Supplementary Fig. 10h).
Global heterozygous knockout of Klf4 alters plaque pathogenesis
Previous studies using VE-cadherin-Cre Klf4fl/flApoe−/− and LysMCre/Cre
Klf4fl/flApoe−/− mouse models provided evidence that KLF4 has an
atheroprotective role in endothelial cells and myeloid cells, respec-
tively; KLF4 deficiency in these strains resulted in increased lesion size
and changes consistent with enhanced inflammation41,42. Indeed, we
confirmed the effects of myeloid-specific KLF4 deficiency by show-
ing that LysM-Cre-dependent knockout of Klf4 was associated with
increased lesion size (Supplementary Fig. 11a) and Sudan IV lipid
staining (Supplementary Fig. 11b) in LysMCre/Cre ROSA STOP floxed
eYFP Klf4fl/flApoe−/− mice. However, unlike previous studies41,42,
we did not observe changes in triglyceride or total cholesterol levels
(Supplementary Fig. 11c). We also found a reduced number of lesional
LGALS3+ cells derived from LysM-Cre–expressing cells in LysMCre/Cre
Klf4∆/∆Apoe−/− miceascomparedtowild-typecontrol LysMCre/CreKlf4+/+
Apoe−/− mice, based on YFP expression (Supplementary Fig. 11d).
However, these results are equivocal regarding the identity of the YFP+
cell types affected by KLF4 deficiency because cells other than myeloid
cells, including SMCs, may express the LysM-Cre transgene.
a
DAPI
SMCKIf4+/+
eYFP
+/+
Apoe
–/–
SMCKIf4∆/∆
eYFP+/+
Apoe
–/–
ACTA2 eYFP LGALS3 Merge
e 40
¥
NS*35
30
25
20
15
10
5
0
LGALS3
+
YFP
–
LGALS3
+
YFP
+
LGALS3
+
YFP
–
YFP
+
DAPI
+
%ofpopulationwithinthe
lesion
f NS
*
*
30
25
20
15
10
5
0
ACTA2
+
YFP
+
ACTA2
+
YFP
–
ACTA2
+
YFP
–
YFP
+
DAPI
+
%ofpopulationwithinthe
lesion
g
MKi67
+
YFP
+
MKi67
+
YFP
+
MKi67
+
YFP
+
YFP
–
DAPI
+
14
12
10
8
6
4
NS
*
2
0
%ofpopulationwithinthe
lesion
NS
h
30
25
20
15
10
5
0
CASP3+
YFP
+
CASP3
+
YFP
+
CASP3
+
YFP
+
YFP
–
DAPI
+
NS
*
*
%ofpopulationwithinthe
lesion
250,000
b
*
200,000
150,000
100,000
50,000
Lesionarea(µm2
)
0
c
*45
40
35
30
25
20
15
10
5
0
Fibrouscaparea/lesionarea
d
*
30
25
20
15
10
5
0
%ACTA2+
cellsinfibrouscap
KIf4
+/+
KIf4
∆/∆
Figure 4  SMC-specific Klf4 deficiency results in
decreased lesion size and increased indices of plaque
stability. SMC lineage tracing mice with or without
conditional SMC-specific Klf4 deletion induced by
treatment with tamoxifen at 6–8 weeks of age were fed
a Western diet for 18 weeks before analysis.
(a) Representative immunofluorescence staining of
BCA lesions. Scale bars, 50 µm. (b–h) Quantitative
data for BCA lesions: total lesion area (b), fibrous
cap area (defined as the region of the lesion within
30 µm of the luminal surface) relative to the size of
the lesion (c), the percentage of ACTA2+ cells within
the fibrous cap (d), the percentages of the indicated
populations of LGALS3+ cells within the lesion (e),
the percentages of the indicated populations of
ACTA2+ cells within the lesion (f), the percentages
of the indicated populations of proliferating (MKi67+)
cells within the lesion (g), and the percentages of the
indicated populations of cells undergoing cell death
(CASP3+) within the lesion (h). *P < 0.05, ¥P = 0.07,
analysis completed by two-way analysis of variance
(ANOVA) comparing genotype and distance from start
of the BCA with a Tukey post-test, error bars show
means ± s.e.m. Quantification is based on analysis
of five 14,283-µm2 regions in each of three sections (d–g), or two sections (h) per mouse, from locations 150 µm, 450 µm and 750 µm from the
start of the branch of the BCA from the aortic arch. SMC YFP+/+Klf4+/+Apoe−/−, n = 11; SMC YFP+/+Klf4∆/∆Apoe−/−, n = 8. CASP3, caspase-3; MKI67,
marker of proliferation Ki67; NS, nonsignificant.
npg©2015NatureAmerica,Inc.Allrightsreserved.
A rt i c l e s
634	 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine
Because the loss of Klf4 in SMCs had opposite effects on plaques to
Klf4 loss in endothelial or myeloid cells, a key unresolved question is
whether global conditional loss of Klf4 would be beneficial or detri-
mental. To test this, we generated tamoxifen-inducible homozygous
Klf4-knockout mice (ERT-Cre+ Klf4fl/flApoe−/−) and heterozygous
Klf4-knockout mice (ERT-Cre+ Klf4fl//+Apoe−/−); the latter represent
a model of partial inhibition of KLF4 across all cell types and thus
mimic potential therapeutic approaches that would achieve partial
suppression of KLF4. Unfortunately, conditional global homozygous
Klf4 mice had to be euthanized at 8–10 weeks after Western diet
feeding owing to excessive weight loss and the development of skin
lesions; these effects are presumably due to the role of KLF4 in reg-
ulating the proliferation and differentiation of epithelial cells45–48.
Tamoxifen-treated ERT-Cre+Klf4fl/+Apoe−/− mice (ERT-Cre+Klf4∆/+
Apoe−/−) demonstrated 50% recombination (the maximum possible
recombination for a heterozygous floxed allele) in the aorta, liver, and
colon (Supplementary Fig. 12a,b), but exhibited no changes in body
weight, heart weight, or cholesterol and triglyceride levels when com-
pared to tamoxifen-treated ERT-Cre−Klf4fl/+Apoe−/− littermate control
mice (Supplementary Fig. 12c).
Notably, mice with conditional global heterozygous Klf4 loss had
similar effects with respect to atherosclerotic lesions as those observed
for SMC-specific conditional Klf4 loss, including a 30% decrease in
lesion size (Supplementary Fig. 13a), and signs of increased plaque
stability, namely increased ACTA2+ cap coverage (Supplementary
Fig. 13b) and decreased LGALS3+ area (Supplementary Fig. 13c).
In addition, conditional global heterozygous Klf4-knockout mice
exhibited decreased intraplaque hemorrhage (Supplementary
Fig. 13d) and decreased apoptosis and cell proliferation
(Supplementary Fig. 13e,f) as compared to ERT-Cre−Klf4+/+Apoe−/−
control mice. Taken together, these results indicate that global loss of
one Klf4 allele has beneficial overall effects on plaque development,
leading to smaller and possibly more stable lesions.
DISCUSSION
Despite numerous reports showing that cultured SMCs downregulate
the expression of SMC differentiation marker genes after exposure to
environmental cues present in atherosclerotic lesions, including cho-
lesterol12, POVPC25, PDGF-BB49, and interleukin (IL)-1β (ref. 50),
the field has relied almost entirely on the detection of these markers
¥7
6
5
4
3
2
1
0
Acta2 Cnn1Myh11Tagln
*
*
#
KLF4binding/Input
a Chow diet
Western diet
KLF4binding/input
Endogenous EndogenousTransgene Transgene
Tagln WT Tagln G/C repressor mutant
¥
*
*
#
3.0
2.5
2.0
1.5
1.0
0.5
0
b
Total genes: 869
Binding
Apoptotic process
Biological adhesion
Biological regulation
Cellular component organization
Cellular process
Developmental process
Immune system process
Localization
Metabolic process
Multicellular organismal process
Reproduction
Response to stimulus
Macrophage activation
Immune response
Antigen processing and
presentation
Catalytic activity
Enzyme regulator activity
Nucleic acid binding transcription factor activity
Protein binding transcription factor activity
Receptor activity
Structural molecule activity
Translation regulator activity
Transporter activity
Total molecular function hits: 1033
Total genes: 869
Total biological process hits: 1798
Total genes: 72
Total immune process hits: 34
Total genes: 869
Cell junction
Cell part
Extracellular matrix
Extracellular region
Macromolecular complex
Membrane
Organelle
Total cellular component hits: 363
d
DAPI
DAPI/eYFP/PLA ACTA2 Merge
eYFP PLAc
Figure 5  KLF4 binds to >800 genes
within SMCs in advanced
atherosclerotic lesions. (a) In vivo
KLF4 binding to the indicated genes,
as determined by ChIP, in BCAs
obtained from Apoe−/− mice fed
either a chow or Western diet for
18 weeks beginning at 8–9 weeks
of age. P values are based on
one-way ANOVA with a Tukey
post hoc test. ¥P = 0.07, #P = 0.11,
*P < 0.05. Error bars show means
± s.e.m. n = 3 independent pooled
groups of five mice per treatment
group. (b) Demonstration of binding
of KLF4 to the Tagln promoter in
BCAs in vivo, as determined by ChIP.
Apoe−/− mice carrying a rat Tagln
wild-type (WT, unmutated) transgene
or Tagln G/C repressor mutant
transgene were fed a chow or
Western diet for 18 weeks beginning at 8–9 weeks of age. Quantitative ChIP assays measured KLF4 binding to the endogenous mouse Tagln gene
relative to its binding to the WT or G/C repressor mutant rat transgenes. P values are based on two-way ANOVA with Tukey post-test. ¥P = 0.06,
#P = 0.43, *P < 0.05. Error bars are means ± s.e.m. n = 3 independent pooled groups of five mice per treatment group. (c) KLF4 binding to the
G/C repressor element of the Tagln promoter in vivo was detected by KLF4-Tagln ISH-PLA. The white arrow indicates a KLF4-PLA+ (red dot)
YFP+ACTA2+ SMC (top enlarged image); the yellow arrow shows a KLF4−PLA+YFP+ACTA2− phenotypically modulated SMC (bottom enlarged image).
Scale bars, 10 µm. (d) Aortic segments from the aortic root and aortic arch up to the carotid artery bifurcation from SMC Klf4+/+YFP+/+Apoe−/− (n = 16)
or SMC Klf4∆/∆eYFP+/+Apoe−/− (n = 16) mice treated with tamoxifen between 6–8 weeks of age and then fed a Western diet for 18 weeks beginning at
8 weeks of age (pooled samples of chromatin from 16 mice per group) were used for ChIP-seq analysis of KLF4 binding targets. 869 targets were
enriched in Western diet–fed SMC Klf4+/+eYFP+/+Apoe−/− mice as compared to Western diet–fed SMC Klf4∆/∆YFP+/+Apoe−/− mice, and thus they
represent putative SMC KLF4 target genes. Pie charts generated by the PANTHER (Protein ANalysis THrough Evolutionary Relationships) classification
system show functional annotation analyses divided into Gene Ontology (GO) terms corresponding to molecular function (top left), cellular component
(top right), biological processes (bottom left) and immune processes (bottom right).
npg©2015NatureAmerica,Inc.Allrightsreserved.
a rt i c l e s
nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015	 635
to ascertain whether a given cell within a lesion is a SMC. Indeed, this
practice has contributed to the well-established dogma in the field that
the role of SMCs within plaques is rather limited, albeit presumed
to be beneficial by virtue of the role of phenotypically modulated
SMCs in producing extracellular matrix and thereby contributing to
fibrous cap formation. Herein we show that >80% of SMCs within
BCA lesions are phenotypically modulated and thus would have been
undetected by conventional techniques. These cells comprise ~30%
of the total cellular composition of the lesion. Moreover, we show
that phenotypically modulated SMCs transition to cells with multiple
phenotypes within lesions, including cells that express markers of
macrophages, MSCs, and/or myofibroblasts. Most notably, we show
that these transitions are functionally important in that selective
loss of Klf4 within SMCs results in reduced lesion size, increased
fibrous cap thickness, and major reductions in the fraction of SMC-
derived macrophage- and MSC-like cells, but causes an increase
in the fraction of ACTA2+ cells within the fibrous cap. In addition,
we show that cholesterol loading of cultured SMCs induces KLF4-
dependent activation of macrophage and MSC markers, expression
of pro-inflammatory cytokines, and increased phagocytic behavior.
Finally, our in vivo KLF4 ChIP-seq analyses identified >800 putative
KLF4 regulated genes within SMCs, including many associated with
pro-inflammatory processes. Taken together, these results provide
compelling evidence that transitions in SMCs phenotype have a
crucial role in lesion development, plaque composition, and stabil-
ity, and suggest that therapeutic approaches aimed at promoting
beneficial changes in SMC phenotype may be a viable means of
treating advanced atherosclerosis.
A key question is how the loss of Klf4 within SMCs results in a
marked reduction in overall lesion size as well as in multiple changes
consistent with increased plaque stability. Our data indicate that these
effects are not due to a change in the number of SMC-derived cells
within lesions. Recent studies by the Fisher laboratory51 found that
although some SMCs express macrophage markers after cholesterol
loading in vitro, principal component analysis of microarray
data revealed that these cells are distinctly different from classi-
cal monocytes, macrophages, and dendritic cells. Moreover, these
cells have reduced phagocytic capacity51. Notably, we found that
SMC derived MSC-like lesional cells seem to be dysfunctional.
Accordingly, we postulate that the loss of Klf4 within SMCs results
in phenotypic transitions that have favorable effects in inhibiting
plaque pathogenesis, including the loss of SMC-derived cells with
‘pro-inflammatory’ macrophage-like properties, and the gain of
SMC-derived cells that contribute to plaque stabilization through
mechanisms and functions yet to be defined. Although it has long
been postulated that SMCs within lesions have a beneficial role,
as indicated in many review articles9–11,26,52,53, our findings show
that this is an oversimplification and that the effects of SMCs
can vary dramatically depending on the nature of their pheno-
typic transitions. A critical challenge for future studies will be to
identify the environmental cues within advanced atherosclerotic
lesions that regulate phenotypic transitions of SMCs, as well as
the other major cell types within lesions, and to determine how
these might be manipulated therapeutically to reduce plaque burden
and increase plaque stability.
Methods
Methods and any associated references are available in the online
version of the paper.
Accession codes. The GEO accession number for KLF4 ChIP-seq
data is GSE65812.
Note: Any Supplementary Information and Source Data files are available in the
online version of the paper.
Acknowledgments
We would like to acknowledge the valuable contributions of technicians
M. McCanna and M. Bevard for their assistance in histological cutting and
staining; R. Tripathi for assistance with cell culture work; J. Lannigan and
c
Nocholesterol80µg/mlcholesteroltreatment
Bright field
Klf4∆/∆
cellsKlf4+/+
cells
eYFP LGALS3 Beads MergeDark field
3
a
Lgals3/18S
2
1
0
0 80
Cholesterol (µg)
*
b 5
4
3
2
1
0
0
Cholesterol (µg)
80
*
Klf4+/+
Klf4∆/∆
%YEP+
LGALS3+
cells
containingbeads/total
population
Figure 6  KLF4-dependent effects in cholesterol-loaded,
cultured SMCs. Aortic SMCs were isolated from SMC
YFP+/+Klf4+/+ and SMC YFP+/+Klf4∆/∆ mice and sorted
using a FACSVantage SE DIVA(Becton Dickinson) instrument
to ensure a pure SMC (YFP+) cell population. (a) Lgals3
mRNA expression after cholesterol loading (80 µg/ml
cholesterol for 72 h). P values are based on two-way
ANOVA with a Tukey post-test. *P < 0.05. The data are
normalized to Klf4+/+, 0 µg/ml cholesterol (vehicle control
only). Error bars are means ± s.e.m. n = 3 independent
experiments. (b,c) Klf4+/+ and Klf4∆/∆ cells were incubated
with 0.8-µm polystyrene beads for 1.5 h after 72 h of
cholesterol loading to induce a macrophage-like state. (b) Quantification of bead uptake in YFP+LGALS3+ cells. Fisher’s exact test was run on data
normalized to the total number of viable cells collected by the cytometer in the Klf4+/+ group treated with 0 µg. P values are based on Fisher’s exact
test (*P < 0.05). One representative experiment from two is shown. (c) Representative images showing staining for YFP (green), LGALS3 (yellow),
beads (red), and dark field (magenta; this parameter provides an index of cell granularity or density). Scale bar, 5 µm.
npg©2015NatureAmerica,Inc.Allrightsreserved.
A rt i c l e s
636	 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine
M. Solga from the Flow Cytometry Core at the University of Virginia for their help
designing flow cytometry panels and help running the cytometers; S. Guilot at the
Advanced Microscopy Facility at University of Virginia for her help running the
transmission electron microscope; J. Roithmayr, N. Hendley, M. Quetsch and
M. Goodwin for their assistance in immunofluorescence image analysis; Y. Babiy
for designing the cartoon in Supplementary Figure 6, and W. Evans for assistance
in processing statistical analyses. We would also like to thank N. McGinn from
the Department of Pathology, University of Virginia Hospital for the de-identified
coronary arteries specimens; C. Murray from the University of Washington for
the coronary artery heart transplant specimen; S. Offermanns from the Max
Planck Institute for the Myh11-CreERT2 mice; K. Kaestner for the Klf4fl/fl mice;
G. Randolph for the LysMCre/Cre mice; and A. Berns from the Netherlands Cancer
Institute, Amsterdam, Netherlands, for the transgenic tamoxifen-inducible Cre
(ERT-Cre) recombinase mice. This work was supported by US National Institutes
of Health R01 grants HL057353, HL098538 and HL087867 to G.K.O., HL112904
to A.C.S., as well as a pilot grant from AstraZeneca as part of a University of
Virginia–AstraZeneca Research Alliance to G.K.O.; and Mid-Atlantic American
Heart Association fellowship grants 11PRE7170008 and 13POST17080043 to
L.S.S. and D.G., respectively.
AUTHOR CONTRIBUTIONS
L.S.S. conducted experiments; performed data analysis; generated most of
the experimental mice; performed immunostaining, image analysis and flow
cytometry; and was primary writer of the manuscript. D.G. performed in vitro
ChIP analysis and conceived and performed all the ISH-PLA experiments.
M.S. generated Tagln wild-type lacZ Apoe−/− mice, performed in vivo ChIP assays,
and performed immunohistochemistry and data analysis. O.A.C. was involved in
designing experiments and data analysis; assisted in image analysis, cell culture
and animal experiments throughout the project. A.C.S. and B.I. participated in
designing the immuno-transmission electron microscopy protocol and analysis
of images. R.M.H. performed the MSC immunostaining and image analysis,
and the Klf4 ChIP-seq; conducted MSC differentiation experiments; and
helped design cartoons. G.F.A. conducted data analysis of ChIP-seq
experiments. P.S. assisted in cell culture experiments and analysis of data.
A.A.C.N. performed PDGFβR staining and analysis. E.S.G. assisted in animal
experiments, conducted statistical analyses and performed immunohistochemistry
and data analysis. L.S.S., D.G., M.S., O.A.C., E.S.G., B.I., G.J.R. and G.K.O.
participated in making final manuscript revisions. G.K.O. supervised the
entire project and had a major role in experimental design, data interpretation,
and writing the manuscript.
COMPETING FINANCIAL INTERESTS
The authors declare no competing financial interests.
Reprints and permissions information is available online at http://www.nature.com/
reprints/index.html.
1.	 Saffitz, J.E. & Schwartz, C.J. Coronary atherosclerosis and thrombosis underlying
acute myocardial infarction. Cardiol. Clin. 5, 21–30 (1987).
2.	 Libby, P. & Aikawa, M. Stabilization of atherosclerotic plaques: new mechanisms
and clinical targets. Nat. Med. 8, 1257–1262 (2002).
3.	 Falk, E., Nakano, M., Benton, J.F., Finn, A.V. & Virmani, R. Update on acute
coronary syndromes: the pathologists′ view. Eur. Heart J. 34, 719–728 (20123).
4.	 Falk, E., Shah, P.K. & Fuster, V. Coronary plaque disruption. Circulation 92,
657–671 (1995).
5.	 Virmani, R., Kolodgie, F.D., Burke, A.P., Farb, A. & Schwartz, S.M. Lessons from
sudden coronary death: a comprehensive morphological classification scheme for
atherosclerotic lesions. Arterioscler. Thromb. Vasc. Biol. 20, 1262–1275 (2000).
6.	 Lee, R.T. & Libby, P. The unstable atheroma. Arterioscler. Thromb. Vasc. Biol. 17,
1859–1867 (1997).
7.	 Ross, R. Atherosclerosis–an inflammatory disease. N. Engl. J. Med. 340, 115–126
(1999).
8.	 Libby, P. Inflammation in atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 32,
2045–2051 (2012).
9.	 Glass, C.K. & Witztum, J.L. Atherosclerosis: the road ahead. Cell 104, 503–516
(2001).
10.	Libby, P., Ridker, P.M. & Hansson, G.K. Progress and challenges in translating the
biology of atherosclerosis. Nature 473, 317–325 (2011).
11.	Gomez, D. & Owens, G.K. Smooth muscle cell phenotypic switching in atherosclerosis.
Cardiovasc. Res. 95, 156–164 (2012).
12.	Rong, J.X., Shapiro, M., Trogan, E. & Fisher, E.A. Transdifferentiation of mouse
aortic smooth muscle cells to a macrophage-like state after cholesterol loading.
Proc. Natl. Acad. Sci. USA 100, 13531–13536 (2003).
13.	Martin, K. et al. Thrombin stimulates smooth muscle cell differentiation from
peripheral blood mononuclear cells via protease-activated receptor-1, RhoA, and
myocardin. Circ. Res. 105, 214–218 (2009).
14.	Caplice, N.M. et al. Smooth muscle cells in human coronary atherosclerosis can
originate from cells administered at marrow transplantation. Proc. Natl. Acad. Sci.
USA 100, 4754–4759 (2003).
15.	Iwata, H. et al. Bone marrow-derived cells contribute to vascular inflammation but
do not differentiate into smooth muscle cell lineages. Circulation 122, 2048–2057
(2010).
16.	Wamhoff, B.R. et al. A G/C element mediates repression of the SM22α promoter
within phenotypically modulated smooth muscle cells in experimental atherosclerosis.
Circ. Res. 95, 981–988 (2004).
17.	Feil, S. et al. Transdifferentiation of vascular smooth muscle cells to macrophage-
like cells during atherogenesis. Circ. Res. 115, 662–667 (2014).
18.	Swirski, F.K. & Nahrendorf, M. Do vascular smooth muscle cells differentiate to
macrophages in atherosclerotic lesions? Circ. Res. 115, 605–606 (2014).
19.	Allahverdian, S., Chehroudi, A.C., McManus, B.M., Abraham, T. & Francis, G.A.
Contribution of intimal smooth muscle cells to cholesterol accumulation and
macrophage-like cells in human atherosclerosis. Circulation 129, 1551–1559
(2014).
20.	Cordes, K.R. et al. miR-145 and miR-143 regulate smooth muscle cell fate and
plasticity. Nature 460, 705–710 (2009).
21.	Yoshida, T., Kaestner, K.H. & Owens, G.K. Conditional deletion of Krüppel-like
factor 4 delays downregulation of smooth muscle cell differentiation markers but
accelerates neointimal formation following vascular injury. Circ. Res. 102,
1548–1557 (2008).
22.	Deaton, R.A., Gan, Q. & Owens, G.K. Sp1-dependent activation of KLF4 is required
for PDGF-BB–induced phenotypic modulation of smooth muscle. Am. J. Physiol.
Heart Circ. Physiol. 296, H1027–H1037 (2009).
23.	Yoshida, T., Gan, Q. & Owens, G.K. Kruppel-like factor 4, Elk-1, and histone
deacetylases cooperatively suppress smooth muscle cell differentiation markers in
response to oxidized phospholipids. Am. J. Physiol. Cell Physiol. 295,
C1175–C1182 (2008).
24.	Thomas, J.A. et al. PDGF-DD, a novel mediator of smooth muscle cell
phenotypic modulation, is upregulated in endothelial cells exposed to atherosclerosis-
prone flow patterns. Am. J. Physiol. Heart Circ. Physiol. 296, H442–H452 (2009).
25.	Pidkovka, N.A. et al. Oxidized phospholipids induce phenotypic switching of vascular
smooth muscle cells in vivo and in vitro. Circ. Res. 101, 792–801 (2007).
26.	Alexander, M.R. & Owens, G.K. Epigenetic control of smooth muscle cell
differentiation and phenotypic switching in vascular development and disease.
Annu. Rev. Physiol. 74, 13–40 (2012).
27.	Gomez, D., Shankman, L.S., Nguyen, A.T. & Owens, G.K. Detection of histone
modifications at specific gene loci in single cells in histological sections.
Nat. Methods 10, 171–177 (2013).
28.	Klein, D., Benchellal, M., Kleff, V., Jakob, H.G. & Ergun, S. Hox genes are involved
in vascular wall-resident multipotent stem cell differentiation into smooth muscle
cells. Sci. Rep. 3, 2178 (2013).
29.	Klein, D. et al. Vascular wall-resident CD44+ multipotent stem cells give rise to
pericytes and smooth muscle cells and contribute to new vessel maturation.
PLoS ONE 6, e20540 (2011).
30.	Xiao, Q. et al. Sca-1+ progenitors derived from embryonic stem cells differentiate
into endothelial cells capable of vascular repair after arterial injury. Arterioscler.
Thromb. Vasc. Biol. 26, 2244–2251 (2006).
31.	Xiao, Q., Zeng, L., Zhang, Z., Hu, Y. & Xu, Q. Stem cell-derived Sca-1+ progenitors
differentiate into smooth muscle cells, which is mediated by collagen IV-integrin
α1/β1/αv and PDGF receptor pathways. Am. J. Physiol. Cell Physiol. 292,
C342–C352 (2007).
32.	Passman, J.N. et al. A sonic hedgehog signaling domain in the arterial adventitia
supports resident Sca1+ smooth muscle progenitor cells. Proc. Natl. Acad. Sci.
USA 105, 9349–9354 (2008).
33.	McDonald, O.G., Wamhoff, B.R., Hoofnagle, M.H. & Owens, G.K. Control of SRF
binding to CArG box chromatin regulates smooth muscle gene expression in vivo.
J. Clin. Invest. 116, 36–48 (2006).
34.	Vladykovskaya, E. et al. Reductive metabolism increases the proinflammatory activity
of aldehyde phospholipids. J. Lipid Res. 52, 2209–2225 (2011).
35.	Takahashi, K. et al. Induction of pluripotent stem cells from adult human fibroblasts
by defined factors. Cell 131, 861–872 (2007).
36.	Cherepanova, O.A. et al. Oxidized phospholipids induce type VIII collagen expression
and vascular smooth muscle cell migration. Circ. Res. 104, 609–618 (2009).
37.	Salmon, M., Gomez, D., Greene, E., Shankman, L. & Owens, G.K. Cooperative
binding of KLF4, pELK-1, and HDAC2 to a G/C repressor element in the SM22α
promoter mediates transcriptional silencing during SMC phenotypic switching
in vivo. Circ. Res. 111, 685–696 (2012).
38.	Regan, C.P., Adam, P.J., Madsen, C.S. & Owens, G.K. Molecular mechanisms of
decreased smooth muscle differentiation marker expression after vascular injury.
J. Clin. Invest. 106, 1139–1147 (2000).
39.	Feinberg, M.W. et al. The Kruppel-like factor KLF4 is a critical regulator of monocyte
differentiation. EMBO J. 26, 4138–4148 (2007).
40.	Liao, X. et al. Kruppel-like factor 4 regulates macrophage polarization. J. Clin.
Invest. 121, 2736–2749 (2011).
41.	Sharma, N. et al. Myeloid Krüppel-like factor 4 deficiency augments atherogenesis
in Apoe−/− mice–brief report. Arterioscler. Thromb. Vasc. Biol. 32, 2836–2838
(2012).
42.	Zhou, G. et al. Endothelial Krüppel-like factor 4 protects against atherothrombosis
in mice. J. Clin. Invest. 122, 4727–4731 (2012).
npg©2015NatureAmerica,Inc.Allrightsreserved.
a rt i c l e s
nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015	 637
43.	Nguyen, A.T. et al. Smooth muscle cell plasticity: fact or fiction? Circ. Res. 112,
17–22 (2013).
44.	Tang, Z. et al. Differentiation of multipotent vascular stem cells contributes to
vascular diseases. Nat. Commun. 3, 875 (2012).
45.	Foster, K.W. et al. Induction of KLF4 in basal keratinocytes blocks the proliferation-
differentiation switch and initiates squamous epithelial dysplasia. Oncogene 24,
1491–1500 (2005).
46.	Jaubert, J., Cheng, J. & Segre, J.A. Ectopic expression of Krüppel-like factor 4
(Klf4) accelerates formation of the epidermal permeability barrier. Development
130, 2767–2777 (2003).
47.	Katz, J.P. et al. The zinc-finger transcription factor Klf4 is required for terminal
differentiation of goblet cells in the colon. Development 129, 2619–2628 (2002).
48.	Katz, J.P. et al. Loss of Klf4 in mice causes altered proliferation and differentiation
and precancerous changes in the adult stomach. Gastroenterology 128, 935–945
(2005).
49.	Dandré, F. & Owens, G.K. Platelet-derived growth factor-BB and Ets-1 transcription
factor negatively regulate transcription of multiple smooth muscle cell differentiation
marker genes. Am. J. Physiol. Heart Circ. Physiol. 286, H2042–H2051
(2004).
50.	Clément, N. et al. Notch3 and IL-1β exert opposing effects on a vascular smooth
muscle cell inflammatory pathway in which NF-κB drives crosstalk. J. Cell Sci.
120, 3352–3361 (2007).
51.	Vengrenyuk, Y. et al. Cholesterol loading reprograms the microRNA-143/145-
myocardin axis to convert aortic smooth muscle cells to a dysfunctional macrophage-
like phenotype. Arterioscler. Thromb. Vasc. Biol. 35, 535–546 (2015).
52.	Owens, G.K. Regulation of differentiation of vascular smooth muscle cells. Physiol.
Rev. 75, 487–517 (1995).
53.	Owens, G.K., Kumar, M.S. & Wamhoff, B.R. Molecular regulation of vascular smooth
muscle cell differentiation in development and disease. Physiol. Rev. 84, 767–801
(2004).
npg©2015NatureAmerica,Inc.Allrightsreserved.
nature medicine doi:10.1038/nm.3866
ONLINE METHODS
Mice. Animal protocols were approved by the University of Virginia Animal
Care and Use Committee. Male Myh11-CreERT2 (ref. 54), LysMCre/Cre, Apoe−/−,
ROSA26 STOP-flox eYFP+/+, Klf4fl/fl (ref. 47), and ERT-Cre (ref. 55) mice were
used in this study. We also used male Apoe−/− transgenic mice carrying either
an unmutated Tagln promoter-lacZ transgene, or a G/C repressor-mutated
version of the Tagln promoter-lacZ transgene16. ROSA26 STOP-flox eYFP+/+
and Apoe−/− mice were obtained from Jackson Laboratories. Myh11-CreERT2,
ROSA26 STOP-flox eYFP, Klf4fl/fl and ERT-Cre mice were genotyped by PCR as
previously described27,47,54,55. For both the Myh11-CreERT2 and ERT-Cre mouse
models, Cre recombinase was activated in male mice with a series of ten 1-mg
tamoxifen (Sigma, cat. no. T-5648) intraperitoneal injections from 6 to 8 weeks
of age, for a total of 10 mg of tamoxifen per mouse, which averaged 25 g of body
weight during this 2-week period. Male littermate controls were used for all
studies. Mice were fed a Western diet containing 21% milk fat and 0.15% choles-
terol (Western diet, Harlan Teklad) for 18 weeks starting at the end of tamoxifen
treatment. Irradiated mouse standard chow diet was purchased through Harlan
(cat. no. 7012). Mice were euthanized by CO2 asphyxiation and then perfused
via the left ventricle as follows: 5 ml PBS, 10 ml 4% paraformaldehyde, 5 ml PBS.
Brachiocephalic arteries were carefully dissected and fixed for an additional
hour in 4% paraformaldehyde before they were embedded in paraffin with the
exception of SCA1-stained tissues, which were snap-frozen in Richard-Allan
Scientific Neg-50 Frozen Section Medium (Thermo Scientific) before sectioning.
Assays for determining total plasma cholesterol and triglyceride levels (Abbott
Laboratories) were performed by the University of Virginia Clinical Pathology
Laboratory. Mice were allocated to experimental groups based on genotyping
and then randomized for the various experimental measurements. Any animal
with triglyceride or cholesterol levels beyond 3 s.d. of the mean level for all mice
were excluded from all future analyses.
Analysis of atherosclerotic plaques. Paraffin-embedded brachiocephalic arter-
ies (BCAs) were serially sectioned at a 10-µm thickness from the aortic arch to
the right subclavian artery. For immunofluorescence analysis of LGALS3+ SMCs
within the BCA, three sections were analyzed from a location 150 µm, 450 µm,
and 750 µm from the start of the branch of the BCA from the aortic arch. Slides
were stained with antibodies specific to GFP (Abcam ab6673), ACTA2 (Sigma
F3777),LGALS3(CedarlaneCL8942AP),MKI67(Abcamab15580),CASP3(Cell
Signaling 9661S), KLF4 (R&D Systems AF3158), MYH11 (Kamiya Biomedical
Company MC-352), PDGFβR (Abcam ab32570), and SCA1 (Ly6A/E) (Abcam
ab51317). Using a Zeiss LSM700 confocal microscope, a series of eight z-stack
images of 1 µm in thickness were acquired for further analysis. Owing to
variation in cellular composition in different regions of the lesion, cells from
five standardized fields (two near the shoulder, one near the fibrous cap, and
two near the media), each 14,283 µm2 in area, were counted to determine the
cellular composition within each lesion. Close examination of each plane of the
z-stack was conducted using Zen 2009 Light Edition Software (Zeiss) to ensure
the presence of immunofluorescence staining coinciding with a single DAPI+
nucleus. The region of the lesion within 30 µm of the luminal boundary, as
determined using Zen 2009 Light Edition Software, was analyzed to determine
the cellular composition of the lesion cap, and the area of this region was com-
pared to the entire area of the atherosclerotic lesion to determine cap area/lesion
area. Morphometric analyses of lesion size were completed using ImagePro Plus
(Media Cybernetics) as described in Alexander et al.56. Researchers were blinded
to the genotype of the animals until the end of the analysis.
Immuno-transmission electron microscopy. SMC YFP+/+Apoe−/− mice fed a
Western diet for 18 weeks were euthanized by CO2 asphyxiation and perfusion
fixed with 0.5% glutaraldehyde (Electron Microscopy Sciences 16300), 4% para-
formaldehyde (Electron Microscopy Sciences 15700) in 1× PBS. Brachiocephalic
arteries were isolated, frozen in liquid nitrogen, and sent to the University of
Virginia Advanced Microscopy Core for processing. Grids were stained with
an antibody specific to GFP (Abcam ab6673), followed by staining with a rab-
bit anti-goat secondary conjugated to 10-nm gold beads (Electron Microscopy
Sciences 25229). Images were captured using a JEOL 1230 transmission electron
microscope with an ultra-high resolution capture camera (Gatan UltraScan 1000
2k × 2k CCD digital imaging camera).
Flow cytometry. SMC YFP+/+Apoe−/− mice were euthanized by CO2 asphyxi-
ation after 18 weeks of Western diet treatment. Mice were then perfused with
10 ml of PBS, and the aorta from the iliac bifurcation to the aortic root was
gently cleaned of fat and fascia before removal from the animal. After removal
from the animal, the heart was dissected away from the aortic root and the tis-
sue was placed into an enzyme cocktail containing 4 U/ml Liberase TM (Roche
05401119001), 0.1 mg/ml DNase I, and 60 U/ml hyaluronidase in RPMI-1640.
Once immersed in the digestion cocktail, the tissue was cut longitudinally,
minced, and placed in a 37 °C incubator for 1.5 h. Cells were run through a
70-µm strainer and spun down at 500g for 5 min. Cells were resuspended in
red blood cell lysis buffer (BD PharmLyse 555899) for two minutes and then
inactivated using serum containing media, spun down again, and resuspended
in 200 µl of 1× PBS. SMC-derived macrophage-like cells were identified using
antibodies specific to F4/80 (eBioscience 17-4801), PTPRC (eBioscience 47-
0451), ITGAM (eBioscience 45-0112), DAPI (Invitrogen), LGALS3 (BioLegend
125405), and ITGAX (eBioscience 25-0114). SMC-derived MSC-like cells were
identified using negative gating for PTPRC (eBioscience 12-0451-82), CDH5
(eBioscience 17-1441), and CD34 (BioLegend 128611), and using positive
gating for ENG (eBioscience 48-1051-82) and SCA1 (eBioscience 45-5981-82).
All samples were run on a Beckman Coulter CyAn ADP LX flow cytometer
equipped with 405-nm, 488-nm, and 633-nm lasers.
Human specimens. De-identified atherosclerotic coronary artery specimens
from patients (n = 12) were collected during autopsy. These specimens were
processed, fixed in paraformaldehyde, and paraffin-embedded blocks were cut
into 5-µm sections. One coronary artery specimen was from a male patient who
had received a heart from a female donor (n = 1). The Institutional Review Board
at University of Virginia approved the use of all autopsy specimens.
In situ hybridization proximity ligation assay (ISH-PLA). ISH-PLA was per-
formed as previously described27. Briefly, human MYH11, mouse Myh11, and
mouse Tagln probes were generated by nick translation (using Nick Translation
kit no. 10976776001, Roche) using biotin-14-dATP (Invitrogen). Biotin labeled
probes (40 ng per slide) underwent denaturation in hybridization buffer (2× SSC,
50% high-grade formamide, 10% dextran sulfate, 1 µg of human or mouse Cot-1
DNA) for 5 min at 80 °C. First, sections were stained for ACTA2 (Sigma F3777)
and CD68 (Santa Cruz sc20060 KP1 clone) or LGALS3 (Cedarlane CL8942AP)
(Fig. 3 and Supplementary Fig. 5) or GFP (Abcam ab6673) and ACTA2 (Sigma
F3777) (Fig. 5). Briefly, slides were de-paraffinized and rehydrated in a xylene
and ethanol series. After antigen retrieval (Vector no. H-3300), sections were
blocked with fish skin gelatin oil in PBS (6 grams per liter) containing 10% horse
serum for 1 h at room temperature (21 °C). Slides were incubated with primary
antibodies for 1 h at room temperature, followed by incubation with donkey
anti-mouse antibody conjugated with Alexa Fluor 647 (4 µg/ml, Invitrogen).
Slides were then dehydrated in an ethanol series and incubated in 1mM EDTA
(pH 8.0) for 20 min. Then, samples were incubated with pepsin (0.5%) in buffer
(0.05 M Tris, 2 mM CaCl2, 0.01 M EDTA, 0.01 M NaCl) at 37 °C for 20 min,
as previously described14. The hybridization mixture containing biotin-labeled
probes (mMyh11, hMYH11 or mTagln) or a 5-TAMRA-dUTP–labeled Y chromo-
some probe (clone RP11-88F4, Empire Genomics) was applied on the sections.
Sections were then incubated at 80 °C for 5 min, followed by 16–24-h incubation
at 37 °C. Hybridization was followed by multiple washes in 2× SSC, 0.1% NP-40
buffer. PLA was performed directly after ISH according to the manufacturer’s
instructions (Olink) and as previously described27. Sections were incubated com-
bining mouse H3K4dime (5 µg/ml, clone CMA303 Millipore) and rabbit biotin
(5 µg/ml, no. ab53494, Abcam) antibodies (Fig. 3 and Supplementary Fig. 5)
or mouse KLF4 [56CT5.1.6] (2.5 µg/ml, no. ab75486, Abcam) and rabbit biotin
(5 µg/ml, no. ab53494, Abcam) antibodies (Fig. 5) overnight at 4 °C. The PLA
amplification step was performed using the Duolink detection kit (emission wave-
length: 555 nm). Finally, mounting medium with DAPI was used to coverslip the
slides. Images were acquired with an Olympus BX41 microscope fitted with a Q
imaging Retiga 2000R camera. Image acquisition was performed using Q Capture
Pro software (Media Cybernetics & Q Imaging, Inc.). Settings were fixed at
the beginning of both acquisition and analysis steps and were unchanged.
Brightness and contrast were equally adjusted after merging. Image analysis was
performed with ImageJ software. To estimate the percentage of SMC-derived
npg©2015NatureAmerica,Inc.Allrightsreserved.
nature medicinedoi:10.1038/nm.3866
macrophage-like cells, we counted the number of CD68+PLA+ cells in CD68
positive staining areas of human coronary lesions (n = 12). We previously
estimated that the efficiency of the ISH-PLA method in mouse and human
tissues is 65% (ref. 27). Considering the incomplete efficiency of ISH-PLA,
we applied a correction to the percentage of CD68+PLA+ cells to total CD68+
cells identified in human lesions. Both uncorrected and corrected percentages
of CD68+PLA+ cells to total CD68+ cells are shown in Figure 3. The corrected
values were proportionally calculated by dividing the uncorrected values by
0.65, the efficiency of the in situ hybridization portion of the PLA method
(see Gomez et al.27).
ChIP assays. Cell culture ChIP was performed as previously described57. Cells
were fixed with 1% paraformaldehyde for 10 min at room temperature. Cross-
linked chromatin was sheared by sonication into fragments of 200–600 base pairs.
The sheared chromatin was immunoprecipitated with 2 µg of anti-H3K4diMe
(clone CMA303, Millipore), or anti-KLF4 (Santa Cruz sc20691); negative control
samples were incubated with mouse or rabbit IgG (Jackson ImmunoResearch
Laboratories). Immune complexes were captured with magnetic bead-coupled
protein G (Millipore). After elution and purification of the genomic DNA
(gDNA), real-time PCR was performed on immunoprecipitated (IP) and non-
immunoprecipitated (INPUT) gDNA. Primer sets used for the Myh11, Arg1
and Cox2 promoters were as previously described27,40. Results are expressed as
the ratio of IP/INPUT. In vivo KLF4 ChIP assays were performed as previously
described22,37 usingflash-frozenBCAsfrommicethathadbeenfedeither18weeks
of high-fat or standard chow (Harlan 7012) diets beginning at 8 weeks of age.
KLF4 chromatin immunoprecipitation–sequencing (ChIP-seq). Segments of
the aorta from the arch to the aortic root and up to the carotid bifurcation isolated
from 18 week western diet fed SMC Klf4+/+eYFP+/+Apoe−/− (n = 16), and SMC
Klf4∆/∆eYFP+/+ Apoe−/− (n=16)miceweresnap-frozeninliquidnitrogenandthen
processed as indicated above for ChIP assays using the KLF4-specific antibody.
DNA library preparation and deep sequencing was performed by HudsonAlpha
Institute for Biotechnology using the Illumina TruSeq Chip Library Kit according
to the manufacturer’s protocol. Quality control and quantification of DNA and
library were performed using an Agilent 2100 Bioanalyzer and a Kapa Library
Quantification Kit (Kapa Biosystems) according to the manufacturer’s protocol.
ChIP-seqdataprocessing.SequencingreadsfromanIlluminaHiSeqSequencing
System were aligned to the mouse genome (mm10) using the BOWTIE align-
ment tool58. These aligned reads were then processed and converted into
bam/bai format (http://genome.ucsc.edu/goldenPath/help/bam.html), and
then loaded in the Integrative Genomics Viewer (http://www.broadinstitute.
org/igv/) for visualization. The processing steps involved removing duplicate
reads and format conversions using the SAMtools59 suite. The reads were also
converted to BED format (http://genome.ucsc.edu/FAQ/FAQformat#format1)
for further data analysis processes such as peak calling. KLF4 peaks were identi-
fied using MACS14 (ref. 60) with ChIP-seq BED files as input files and default
settings with a P value for significant peak calling ≤ 1 × 10−5. Once peaks were
obtained for both ChIP-seq data sets (SMC Klf4+/+eYFP+/+Apoe−/− and SMC
Klf4∆/∆eYFP+/+Apoe−/−), BEDtools61 was used to remove peaks that were present
in both data sets. Furthermore, BEDtools was used to identify the closest genes
to each peak. If a gene was present in both data sets, the gene was removed from
the analysis. Functional annotation was performed using PANTHER62 and Gene
Ontology Consortium (http://geneontology.org/) software; a statistical over-
representation test was performed using PANTHER62 (http://pantherdb.org/).
The GEO accession number for these data is GSE65812.
Cell culture studies. Primary mouse aortic SMCs were isolated using a previ-
ously described protocol63 from 8-week-old SMC-lineage tracing mice after
a series of ten tamoxifen injections. Cells isolated from SMC YFP+/+Klf4+/+
and SMC YFP+/+Klf4∆/∆ mice were passaged three times before undergoing
cell sorting for YFP+ cells using a FACSVantage SE DIVA (Becton Dickinson).
Cholesterol assays were performed using water-soluble cholesterol from Sigma
(C4951-30MG) as previously published12 with a minor modification: cholesterol
was reconstituted in DMEM-F12 media (Gibco) containing 0.2% FBS (Gibco),
100 U/ml penicillin/streptomycin (Gibco) and 1.6 mmol/liter l-glutamine,
(Gibco). Cells were allowed to grow to ~70% confluency in DF10 (DMEM-F12
medium (Gibco) containing 10% FBS (Gibco), 100 U/ml penicillin/streptomycin
(Gibco) and 1.6 mmol/liter l-glutamine (Gibco)) before being switched to 0, 20,
40, or 80 µg/ml cholesterol-containing medium. After 72 h, cells were harvested
for mRNA, protein, or ChIP analysis. Total RNA was extracted from cultured
cells using Trizol reagent (Invitrogen) according to the respective manufac-
turer’s instructions. One microgram of RNA was used to perform a reverse
transcription with iScript cDNA Synthesis Kit (Bio-Rad). A SensiFAST SYBR
NO-ROX Mix (Bioline) was used to carry out RT-PCR on a C1000 Thermal
Cycler CFX96TM (Bio-Rad) (Supplementary Table 4). Cell culture medium
from 72-h cholesterol-treated Klf4+/+ or Klf4∆/∆ cells was filtered through
0.22-µm Millex low-protein-binding Durapore filters (Merck/Millipore) to
remove traces of cholesterol. Then medium was concentrated ten times using
Amicon Ultra centrifugal filter devices (Millipore). Cellular proteins for cytokine
assay were isolated using the RIPA buffer protocol. Briefly, cells were incubated
in RIPA buffer (Pierce) with Halt Protease Inhibitor cocktail (Thermo Scientific)
at 4 °C for 15 min, centrifuged, and supernatants were used for analysis.
Protein concentrations were measured using the Bio-Rad DC Protein Assay
kit. Human coronary SMCs were purchased from Lonza and cultured in SMC
maintaining media (Lonza). RAW264.7 mouse macrophages (American Type
Culture Collection) and human monocytes (American Type Culture Collection)
were cultured per the company’s recommendations.
Bead uptake assays. Primary mouse aortic SMCs were treated with cholesterol
as indicated above. After 72 h of cholesterol treatment, cells were switched back
to DF10 culture medium containing 1.5% vol/vol 0.84-µm polystyrene beads
(Spherotech FP-0870-2) for 1.5 h. Cells were then harvested, stained with LGALS3
(BioLegend 125405) and run on an Amnis ImageStreamX Mark II flow cytometer
using a 60× objective. Data was analyzed using Amnis IDEAS software.
Mesenchymal stem cell differentiation. Cells were isolated from SMC
YFP+/+Apoe−/− mice fed a Western diet for 12 or 18 weeks as described in
the flow cytometry section. Next, cells were stained and negatively gated for
lineage markers (CD34 BioLegend 128611, PTPRC eBioscience 56-0451-80,
CDH5 eBioscience 17-1441-80), and positively gated for ENG (eBioscience
48-1051-82) and SCA1 (eBioscience 45-5981-82). YFP was detected by native
fluorescence. Cells were then sorted into four populations (YFP+SCA1+ENG+,
YFP−SCA1+ENG+, YFP+SCA1−ENG−, and YFP−SCA1−ENG−) using a Becton
Dickinson Influx Cell Sorter. Cells were then placed in StemXVivo MSC medium
(R&D Systems CCM004) media at a density of 30,000 cells/mm2 and were pas-
saged at 70% confluency (changing the medium every 2–3 d). YFP−SCA1+ENG+
and YFP+SCA1+ENG+ cells were expanded prior to plating for the differentia-
tion experiments. The differentiation experiment was conducted using a kit
from R&D Systems (mouse mesenchymal stem cell functional differentiation kit,
SC010). Cells were differentiated in the appropriate media for 14 d (changing the
medium every 2–3 d). The cells were then fixed and immunocytochemistry was
performed using antibodies from the kit according to the protocol provided.
Statistics. Fisher’s exact test was used for categorical data. Two-way ANOVA
with Tukey post hoc tests were used for multiple group comparisons determined
to have normal distribution by the Kolmogorov-Smirnov test. For experiments
in which multiple sections of the BCA were assessed, two-way ANOVAs were
conducted to determine if there were statistical differences between the dif-
ferent regions of the BCA evaluated (i.e., 150 µm, 450 µm, and 750 µm from
the branch of the BCA from the aortic arch). Because no statistical differences
were seen between BCA regions, we calculated an average across all regions.
Nonparametric data were analyzed using the Wilcoxon rank-sum test. There
were no significant interactions between genotype and region (from the start
of the BCA) for any of the endpoints analyzed by two-way ANOVA. P < 0.05
was considered significant. SAS v9.3 with Enterprise Guide v5.1 software (SAS
Institute Inc.) was used for all statistical analyses.
54.	Wirth, A. et al. G12–G13-LARG–mediated signaling in vascular smooth muscle is
required for salt-induced hypertension. Nat. Med. 14, 64–68 (2008).
npg©2015NatureAmerica,Inc.Allrightsreserved.
nature medicine doi:10.1038/nm.3866
59.	Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics
25, 2078–2079 (2009).
60.	Zhang, Y. et al. Model-based analysis of ChIP-seq (MACS). Genome Biol. 9, R137
(2008).
61.	Quinlan, A.R. & Hall, I.M. BEDTools: a flexible suite of utilities for comparing
genomic features. Bioinformatics 26, 841–842 (2010).
62.	Mi, H., Muruganujan, A., Casagrande, J.T. & Thomas, P.D. Large-scale gene function
analysis with the PANTHER classification system. Nat. Protoc. 8, 1551–1566 (2013).
63.	Geisterfer, A.A., Peach, M.J. & Owens, G.K. Angiotensin II induces hypertrophy,
not hyperplasia, of cultured rat aortic smooth muscle cells. Circ. Res. 62, 749–756
(1988).
55.	Vooijs, M., Jonkers, J. & Berns, A. A highly efficient ligand-regulated Cre recombinase
mouse line shows that loxP recombination is position dependent. EMBO Rep. 2,
292–297 (2001).
56.	Alexander, M.R. et al. Genetic inactivation of IL-1 signaling enhances atherosclerotic
plaque instability and reduces outward vessel remodeling in advanced atherosclerosis
in mice. J. Clin. Invest. 122, 70–79 (2012).
57.	McDonald, O.G. & Owens, G.K. Programming smooth muscle plasticity with
chromatin dynamics. Circ. Res. 100, 1428–1441 (2007).
58.	Langmead, B., Trapnell, C., Pop, M. & Salzberg, S.L. Ultrafast and memory-efficient
alignment of short DNA sequences to the human genome. Genome Biol. 10, R25
(2009).
npg©2015NatureAmerica,Inc.Allrightsreserved.
Corrigendum: KLF4-dependent phenotypic modulation of smooth muscle
cells has a key role in atherosclerotic plaque pathogenesis
Laura S Shankman, Delphine Gomez, Olga A Cherepanova, Morgan Salmon, Gabriel F Alencar, Ryan M Haskins, Pamela Swiatlowska,
Alexandra A C Newman, Elizabeth S Greene, Adam C Straub, Brant Isakson, Gwendalyn J Randolph & Gary K Owens
Nat. Med. 21, 628–637 (2015); published online 18 May 2015; corrected after print 12 August 2015
In the version of this article initially published, the labels to the left of the two micrographs in Figure 2c are reversed. Also, in Figure 4g, MKi67,
used as a cell proliferation marker, is misspelled. The errors have been corrected in the HTML and PDF versions of the article.
Corrigendum: Metformin activates a duodenal Ampk–dependent pathway to
lower hepatic glucose production in rats
Frank A Duca, Clémence D Côté, Brittany A Rasmussen, Melika Zadeh-Tahmasebi, Guy A Rutter, Beatrice M Filippi & Tony K T Lam
Nat. Med. 21, 506–11 (2015); doi:10.1038/nm.3787; published online 06 April 2015; corrected after print 7 May 2015
In the version of this article initially published, we incorrectly reported the value for the particles per milliliter of Ad-dn-AMPK (D157A) used in
the study. It was 3.1 × 10–9 PFU ml–1 and not 1.1 × 10–13 PFU ml–1 as originally reported. The errors have been corrected in the HTML and PDF
versions of the article.
Erratum: Cardiac RKIP induces a beneficial β-adrenoceptor–dependent
positive inotropy
Evelyn Schmid, Stefan Neef, Christopher Berlin, Angela Tomasovic, Katrin Kahlert, Peter Nordbeck, Katharina Deiss, Sabrina Denzinger,
Sebastian Herrmann, Erich Wettwer, Markus Weidendorfer, Daniel Becker, Florian Schäfer, Nicole Wagner, Süleyman Ergün,
Joachim P Schmitt, Hugo A Katus, Frank Weidemann, Ursula Ravens, Christoph Maack, Lutz Hein, Georg Ertl, Oliver J Müller,
Lars S Maier, Martin J Lohse & Kristina Lorenz
Nat. Med. 21, 1298–1306 (2015); published online 19 October 2015; corrected after print 29 October 2015
Intheversionofthisarticleinitiallypublished,therearetypographicalerrorsinthelabelstoFig.3e–h:‘WT+AAV9-eGFP’and‘WT+AAV9-RKIPWT’
areincorrect,andshouldbe‘RKIP– +AAV9-eGFP’and‘RKIP– +AAV9-RKIPWT’,respectively.TheerrorshavebeencorrectedintheHTMLandPDF
versions of the article.
Erratum: Snail1-induced partial epithelial-to-mesenchymal transition drives
renal fibrosis in mice and can be targeted to reverse established disease.
M Teresa Grande, Berta Sánchez-Laorden, Cristina López-Blau, Cristina A De Frutos, Agnès Boutet, Miguel Arévalo, R Grant Rowe,
Stephen J Weiss, José M López-Novoa & M Angela Nieto
Nat. Med. 21, 989–997 (2015); published online 03 August 2015; corrected after print 4 September 2015
In the version of this article initially published, the word ‘genetically’ was added by the editor to the penultimate sentence of the abstract during
proofing of the manuscript. However, the authors used not only a genetic knockout approach but also morpholino-induced inhibition of proper
mRNA processing to target Snail1 expression. Thus, the word ‘genetically’ has been deleted from the sentence to better convey the findings of the
report. The error has been corrected in the HTML and PDF versions of the article.
Erratum: Genomic landscape of carcinogen-induced and genetically
induced mouse skin squamous cell carcinoma
Dany Nassar, Mathilde Latil, Bram Boeckx, Diether Lambrechts & Cédric Blanpain
Nat. Med. 21, 946–954 (2015); published online 13 July 2015, corrected after print 6 August 2015
In the published article, the format was listed as Article, but this is a Resource. The error has been corrected in the HTML and PDF versions of
the article.
CO R R I G E N DA AN D E R R ATA
NATURE MEDICINE VOLUME 22 | NUMBER 2 | FEBRUARY 2016	 217
npg©2016NatureAmerica,Inc.Allrightsreserved.

More Related Content

What's hot

Proteomics Exploration of Chronic Lymphocytic Leukemia_Crimson Publishers
Proteomics Exploration of Chronic Lymphocytic Leukemia_Crimson PublishersProteomics Exploration of Chronic Lymphocytic Leukemia_Crimson Publishers
Proteomics Exploration of Chronic Lymphocytic Leukemia_Crimson PublishersCrimsonpublishersCancer
 
J. Biol. Chem.-2008-Lin-15003-14
J. Biol. Chem.-2008-Lin-15003-14J. Biol. Chem.-2008-Lin-15003-14
J. Biol. Chem.-2008-Lin-15003-14Da-Wei Lin
 
Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...
Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...
Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...Enrique Moreno Gonzalez
 
On predicting mutation status from transcriptome sequencing data
On predicting mutation status from transcriptome sequencing dataOn predicting mutation status from transcriptome sequencing data
On predicting mutation status from transcriptome sequencing dataShiraishi Yuichi
 
The Effects of Genetic Alteration on Reprogramming of Fibroblasts into Induc...
The Effects of Genetic Alteration on Reprogramming of  Fibroblasts into Induc...The Effects of Genetic Alteration on Reprogramming of  Fibroblasts into Induc...
The Effects of Genetic Alteration on Reprogramming of Fibroblasts into Induc...remedypublications2
 
Malmo Cancer retreat 2012 (2)
Malmo Cancer retreat 2012 (2)Malmo Cancer retreat 2012 (2)
Malmo Cancer retreat 2012 (2)Amr Al-Haidari
 
Research Paper - Naushad Moti
Research Paper - Naushad MotiResearch Paper - Naushad Moti
Research Paper - Naushad MotiNaushad Moti
 
Nuclear FABP7 immunoreactivity is preferentially expressed in infiltrated glioma
Nuclear FABP7 immunoreactivity is preferentially expressed in infiltrated gliomaNuclear FABP7 immunoreactivity is preferentially expressed in infiltrated glioma
Nuclear FABP7 immunoreactivity is preferentially expressed in infiltrated gliomaYu Liang
 
Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...
Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...
Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...Enrique Moreno Gonzalez
 
Austin Journal of Molecular and Cellular Biology research
Austin Journal of Molecular and Cellular Biology research Austin Journal of Molecular and Cellular Biology research
Austin Journal of Molecular and Cellular Biology research Austin Publishing Group
 
CBS patent_US20110105382
CBS patent_US20110105382CBS patent_US20110105382
CBS patent_US20110105382Jaehyun Choi
 
Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...
Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...
Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...Marion Hartmann
 
microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...
microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...
microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...Paul Schoenhagen
 

What's hot (20)

Proteomics Exploration of Chronic Lymphocytic Leukemia_Crimson Publishers
Proteomics Exploration of Chronic Lymphocytic Leukemia_Crimson PublishersProteomics Exploration of Chronic Lymphocytic Leukemia_Crimson Publishers
Proteomics Exploration of Chronic Lymphocytic Leukemia_Crimson Publishers
 
J. Biol. Chem.-2008-Lin-15003-14
J. Biol. Chem.-2008-Lin-15003-14J. Biol. Chem.-2008-Lin-15003-14
J. Biol. Chem.-2008-Lin-15003-14
 
Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...
Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...
Sox2 suppresses the invasiveness of breast cancer cells via a mechanism that ...
 
On predicting mutation status from transcriptome sequencing data
On predicting mutation status from transcriptome sequencing dataOn predicting mutation status from transcriptome sequencing data
On predicting mutation status from transcriptome sequencing data
 
4401899a
4401899a4401899a
4401899a
 
The Effects of Genetic Alteration on Reprogramming of Fibroblasts into Induc...
The Effects of Genetic Alteration on Reprogramming of  Fibroblasts into Induc...The Effects of Genetic Alteration on Reprogramming of  Fibroblasts into Induc...
The Effects of Genetic Alteration on Reprogramming of Fibroblasts into Induc...
 
Malmo Cancer retreat 2012 (2)
Malmo Cancer retreat 2012 (2)Malmo Cancer retreat 2012 (2)
Malmo Cancer retreat 2012 (2)
 
Final UROP poster 2013
Final UROP poster 2013Final UROP poster 2013
Final UROP poster 2013
 
440
440440
440
 
Nrgastro.2012.114
Nrgastro.2012.114Nrgastro.2012.114
Nrgastro.2012.114
 
news and views
news and viewsnews and views
news and views
 
Research Paper - Naushad Moti
Research Paper - Naushad MotiResearch Paper - Naushad Moti
Research Paper - Naushad Moti
 
Nuclear FABP7 immunoreactivity is preferentially expressed in infiltrated glioma
Nuclear FABP7 immunoreactivity is preferentially expressed in infiltrated gliomaNuclear FABP7 immunoreactivity is preferentially expressed in infiltrated glioma
Nuclear FABP7 immunoreactivity is preferentially expressed in infiltrated glioma
 
Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...
Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...
Variant G6PD levels promote tumor cell proliferation or apoptosis via the STA...
 
Austin Journal of Molecular and Cellular Biology research
Austin Journal of Molecular and Cellular Biology research Austin Journal of Molecular and Cellular Biology research
Austin Journal of Molecular and Cellular Biology research
 
2012-JCI
2012-JCI2012-JCI
2012-JCI
 
StntShwcs2015Abstract
StntShwcs2015AbstractStntShwcs2015Abstract
StntShwcs2015Abstract
 
CBS patent_US20110105382
CBS patent_US20110105382CBS patent_US20110105382
CBS patent_US20110105382
 
Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...
Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...
Relationship between CCL5 and TGFβ1 in breast cancer patients at both systemi...
 
microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...
microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...
microRNA-based diagnostics and therapy in cardiovascular disease—Summing up t...
 

Viewers also liked (19)

Curriculum_vitea_for_M_&_E[1]s
Curriculum_vitea_for_M_&_E[1]sCurriculum_vitea_for_M_&_E[1]s
Curriculum_vitea_for_M_&_E[1]s
 
Agasi lucas parcial domiciliario
Agasi lucas   parcial domiciliarioAgasi lucas   parcial domiciliario
Agasi lucas parcial domiciliario
 
Sheraton Sand Key Resort
Sheraton Sand Key ResortSheraton Sand Key Resort
Sheraton Sand Key Resort
 
Los simpson
Los simpsonLos simpson
Los simpson
 
Sebastian serna higuita.
Sebastian serna higuita.Sebastian serna higuita.
Sebastian serna higuita.
 
What Makes a Great Photo
What Makes a Great PhotoWhat Makes a Great Photo
What Makes a Great Photo
 
Segunda parte trabajo comu visualunlz 1
Segunda parte trabajo comu visualunlz 1Segunda parte trabajo comu visualunlz 1
Segunda parte trabajo comu visualunlz 1
 
ўқув амалиёти аппарат 22 март
ўқув амалиёти аппарат 22 мартўқув амалиёти аппарат 22 март
ўқув амалиёти аппарат 22 март
 
engr.Eslam
engr.Eslamengr.Eslam
engr.Eslam
 
NIIF
NIIFNIIF
NIIF
 
Juan rafael arroyo g. 22151892
Juan rafael arroyo g. 22151892Juan rafael arroyo g. 22151892
Juan rafael arroyo g. 22151892
 
Segunda parte trabajo comunicación visual unlz
Segunda parte trabajo comunicación  visual unlzSegunda parte trabajo comunicación  visual unlz
Segunda parte trabajo comunicación visual unlz
 
Trabajo práctico 3 Agasi- Arreguez-Gargiulo-Rodriguez
Trabajo práctico 3 Agasi- Arreguez-Gargiulo-RodriguezTrabajo práctico 3 Agasi- Arreguez-Gargiulo-Rodriguez
Trabajo práctico 3 Agasi- Arreguez-Gargiulo-Rodriguez
 
Trabajo Barthes
Trabajo BarthesTrabajo Barthes
Trabajo Barthes
 
Trabajo Práctico Número 4 Eliseo Verón
Trabajo Práctico Número 4 Eliseo VerónTrabajo Práctico Número 4 Eliseo Verón
Trabajo Práctico Número 4 Eliseo Verón
 
Tp5 (2)
Tp5 (2)Tp5 (2)
Tp5 (2)
 
Tipografia
TipografiaTipografia
Tipografia
 
Tp4 Color
Tp4 Color Tp4 Color
Tp4 Color
 
Agasi lucas parcial domiciliario
Agasi lucas   parcial domiciliarioAgasi lucas   parcial domiciliario
Agasi lucas parcial domiciliario
 

Similar to KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis

2016_Scientific Reports_Article
2016_Scientific Reports_Article2016_Scientific Reports_Article
2016_Scientific Reports_ArticleMauricio Rosenfeld
 
overexpression of mrps18 2 in cancer cell lines results
overexpression of mrps18 2 in cancer cell lines resultsoverexpression of mrps18 2 in cancer cell lines results
overexpression of mrps18 2 in cancer cell lines resultsAnimatedWorld
 
Final Dissertation
Final DissertationFinal Dissertation
Final DissertationRachel Li
 
8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients
8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients
8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patientscrea-autoinmunidad
 
Abstract piis0022202 x1831114x
Abstract piis0022202 x1831114xAbstract piis0022202 x1831114x
Abstract piis0022202 x1831114xWaddah Moghram
 
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...AnonIshanvi
 
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...NainaAnon
 
Stat3 protein & immunocompetent cells in psoriasis by yousry a mawla
Stat3   protein & immunocompetent  cells in psoriasis by yousry a mawlaStat3   protein & immunocompetent  cells in psoriasis by yousry a mawla
Stat3 protein & immunocompetent cells in psoriasis by yousry a mawlaM.YOUSRY Abdel-Mawla
 
Mastócitos
MastócitosMastócitos
Mastócitosgumaaa
 
Stat3protein immunocompetent cells in psoriasisb pathogenesis.ppt
Stat3protein  immunocompetent  cells  in  psoriasisb pathogenesis.pptStat3protein  immunocompetent  cells  in  psoriasisb pathogenesis.ppt
Stat3protein immunocompetent cells in psoriasisb pathogenesis.pptM.YOUSRY Abdel-Mawla
 
Stat3protein & immunocompetent cells in psoriasis pathogenesis
Stat3protein & immunocompetent  cells  in  psoriasis pathogenesisStat3protein & immunocompetent  cells  in  psoriasis pathogenesis
Stat3protein & immunocompetent cells in psoriasis pathogenesisM.YOUSRY Abdel-Mawla
 
Gari et al bmc medical genetics
Gari et al bmc medical geneticsGari et al bmc medical genetics
Gari et al bmc medical geneticsTariq Mohammed
 
Stat3 protein & immunocompetent cells cross talks in psoriasis by yousr...
Stat3   protein & immunocompetent  cells  cross talks   in psoriasis by yousr...Stat3   protein & immunocompetent  cells  cross talks   in psoriasis by yousr...
Stat3 protein & immunocompetent cells cross talks in psoriasis by yousr...M.YOUSRY Abdel-Mawla
 
Menon BB et al Mucosal Immunol 2015
Menon BB et al Mucosal Immunol 2015Menon BB et al Mucosal Immunol 2015
Menon BB et al Mucosal Immunol 2015Balaraj Menon, Ph.D.
 

Similar to KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis (20)

2016_Scientific Reports_Article
2016_Scientific Reports_Article2016_Scientific Reports_Article
2016_Scientific Reports_Article
 
overexpression of mrps18 2 in cancer cell lines results
overexpression of mrps18 2 in cancer cell lines resultsoverexpression of mrps18 2 in cancer cell lines results
overexpression of mrps18 2 in cancer cell lines results
 
ijms-17-00518
ijms-17-00518ijms-17-00518
ijms-17-00518
 
nature
naturenature
nature
 
Final Dissertation
Final DissertationFinal Dissertation
Final Dissertation
 
Ojchd.000534
Ojchd.000534Ojchd.000534
Ojchd.000534
 
8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients
8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients
8. Julieta Gonzalez. Bogota epithelial cells in lsg from ss patients
 
Abstract piis0022202 x1831114x
Abstract piis0022202 x1831114xAbstract piis0022202 x1831114x
Abstract piis0022202 x1831114x
 
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
 
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
A 43-Year-Old Male with PCM1-JAK2 Gene Fusion Experienced T-Lymphoblastic Lym...
 
blood
bloodblood
blood
 
Stat3 protein & immunocompetent cells in psoriasis by yousry a mawla
Stat3   protein & immunocompetent  cells in psoriasis by yousry a mawlaStat3   protein & immunocompetent  cells in psoriasis by yousry a mawla
Stat3 protein & immunocompetent cells in psoriasis by yousry a mawla
 
Mastócitos
MastócitosMastócitos
Mastócitos
 
Stat3protein immunocompetent cells in psoriasisb pathogenesis.ppt
Stat3protein  immunocompetent  cells  in  psoriasisb pathogenesis.pptStat3protein  immunocompetent  cells  in  psoriasisb pathogenesis.ppt
Stat3protein immunocompetent cells in psoriasisb pathogenesis.ppt
 
Stat3protein & immunocompetent cells in psoriasis pathogenesis
Stat3protein & immunocompetent  cells  in  psoriasis pathogenesisStat3protein & immunocompetent  cells  in  psoriasis pathogenesis
Stat3protein & immunocompetent cells in psoriasis pathogenesis
 
Nrneph.2014.170
Nrneph.2014.170Nrneph.2014.170
Nrneph.2014.170
 
Gari et al bmc medical genetics
Gari et al bmc medical geneticsGari et al bmc medical genetics
Gari et al bmc medical genetics
 
SAM
SAMSAM
SAM
 
Stat3 protein & immunocompetent cells cross talks in psoriasis by yousr...
Stat3   protein & immunocompetent  cells  cross talks   in psoriasis by yousr...Stat3   protein & immunocompetent  cells  cross talks   in psoriasis by yousr...
Stat3 protein & immunocompetent cells cross talks in psoriasis by yousr...
 
Menon BB et al Mucosal Immunol 2015
Menon BB et al Mucosal Immunol 2015Menon BB et al Mucosal Immunol 2015
Menon BB et al Mucosal Immunol 2015
 

KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis

  • 1. A rt i c l e s 628 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine Atherosclerosis is a disease of chronic inflammation and is the leading cause of morbidity and mortality worldwide. There is a general con- sensus that the majority of coronary syndromes result from the rupture of unstable plaques and associated thrombotic events1–3. Plaque instability has been associated with disruption of the fibrous cap, an atheroprotective layer of smooth muscle α-actin (ACTA2)-positive cells that cover the atherosclerotic plaque3–7; large numbers of cells positive for macrophage markers such as LGALS3; and the presence of a large foam cell–laden necrotic core within the plaque2–4,8. Indeed, these data have been interpreted as evidence that plaques that contain a high ratio of macrophages relative to SMCs are less stable, particu- larly those that have a thin ACTA2+ fibrous cap that is presumed to be composed primarily of phenotypically modulated SMCs9,10. Although ACTA2+ and LGALS3+ cells are assumed to be of SMC and myeloid lineage respectively, there is extensive ambiguity about the lineal origin of cells within atherosclerotic plaques11. These ambi- guities were originally based on in vitro studies showing that SMCs downregulate SMC markers and activate macrophage markers after cholesterol loading12, and that macrophages activate SMC genes after treatment with factors known to be present within lesions, including thrombin13. However, the most compelling evidence that SMCs and macrophages are often misidentified within human advanced coronary lesions comes from studies of cross-gender bone marrow transplant subjects. These studies have shown that >10% of ACTA2+ cells within lesions are of hematopoietic stem cell (HSC) and not SMC origin14. Consistent with these human data, a substantial fraction of the cells that express SMC markers (including ACTA2 but not MYH11) within lesions of Apoe−/− mice are of HSC and not SMC origin15. Conversely, there is extensive evidence suggesting that many SMC-derived cells within advanced lesions of Apoe−/− mice lack detectable expression of conventional SMC markers such as ACTA2 (ref. 16), and/or activate expression of macrophage markers17. Notably, in vivo studies from our laboratory showed that large numbers of ACTA2−, myosin, heavy polypeptide 11, smooth muscle (MYH11)−, and transgelin (TAGLN)− cells within advanced lesions of Apoe−/− Western diet–fed mice retain expression of a mutant Tagln-lacZ trans- gene that is resistant to downregulation compared to the unmutated transgene16. Unfortunately, these studies are not definitive because we could not rule out the possibility that non-SMCs present within lesions may activate the mutant transgene. 1Robert M. Berne Cardiovascular Research Center, University of Virginia, Charlottesville, Virginia, USA. 2Department of Molecular Physiology and Biological Physics, University of Virginia, Charlottesville, Virginia, USA. 3Department of Surgery, University of Virginia, Charlottesville, Virginia, USA. 4Department of Biochemistry and Molecular Genetics, University of Virginia, Charlottesville, Virginia, USA. 5Department of Pathology, University of Virginia, Charlottesville, Virginia, USA. 6Intercollegiate Faculty of Biotechnology, University of Gdansk, Gdansk, Poland. 7Department of Pharmacology and Chemical Biology, University of Pittsburgh, Pittsburgh, Pennsylvania, USA. 8Division of Biology and Biomedical Sciences, Washington University, St. Louis, Missouri, USA. Correspondence should be addressed to G.K.O. (gko@virginia.edu). Received 5 February; accepted 22 April; published online 18 May 2015; corrected after print 12 August 2015; doi:10.1038/nm.3866 KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis Laura S Shankman1,2, Delphine Gomez1, Olga A Cherepanova1, Morgan Salmon3, Gabriel F Alencar1,4, Ryan M Haskins1,5, Pamela Swiatlowska1,6, Alexandra A C Newman1,4, Elizabeth S Greene1, Adam C Straub7, Brant Isakson1,2, Gwendalyn J Randolph8 & Gary K Owens1,2 Previous studies investigating the role of smooth muscle cells (SMCs) and macrophages in the pathogenesis of atherosclerosis have provided controversial results owing to the use of unreliable methods for clearly identifying each of these cell types. Here, using Myh11-CreERT2 ROSA floxed STOP eYFP Apoe−/− mice to perform SMC lineage tracing, we find that traditional methods for detecting SMCs based on immunostaining for SMC markers fail to detect >80% of SMC-derived cells within advanced atherosclerotic lesions. These unidentified SMC-derived cells exhibit phenotypes of other cell lineages, including macrophages and mesenchymal stem cells (MSCs). SMC-specific conditional knockout of Krüppel-like factor 4 (Klf4) resulted in reduced numbers of SMC-derived MSC- and macrophage-like cells, a marked reduction in lesion size, and increases in multiple indices of plaque stability, including an increase in fibrous cap thickness as compared to wild-type controls. On the basis of in vivo KLF4 chromatin immunoprecipitation–sequencing (ChIP-seq) analyses and studies of cholesterol-treated cultured SMCs, we identified >800 KLF4 target genes, including many that regulate pro-inflammatory responses of SMCs. Our findings indicate that the contribution of SMCs to atherosclerotic plaques has been greatly underestimated, and that KLF4-dependent transitions in SMC phenotype are critical in lesion pathogenesis. npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 2. a rt i c l e s nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015 629 Thus, despite decades of atherosclerosis research, we still do not know which cells within lesions are SMC-derived and to what extent they contribute to lesion pathogenesis. A recent SMC lineage tracing study using a Tagln-ERT2Cre-lacZ Apoe−/− mouse model provided evidence that SMC-derived cells within advanced lesions activate some macrophage markers, including LGALS3 and CD68 (ref. 17). Unfortunately, as highlighted in an editorial on this paper18, the labeling efficiency of SMCs in this study was only 11%, preclud- ing a determination of the fraction of macrophage-like cells within lesions that are derived from SMCs or, most importantly, how these cells might contribute to lesion pathogenesis. Another recent study19 showed that 50% of foam cells within advanced human coronary artery lesions express the SMC marker ACTA2, highlighting the magnitude of the ‘SMC/macrophage misidentification problem’ with respect to our understanding of human disease. However, the majority of these ACTA2+ foam cells also expressed the macrophage marker CD68 and represented 40% of all CD68+ lesional cells19. Given clear evidence that macrophages can activate SMC markers and, vice versa, that SMCs can activate macrophage markers, it is unclear whether these CD68+ lesional cells are derived from SMCs, macrophages, or another cell type. In view of the major ambiguities in identifying which cells within atherosclerotic lesions are SMC-derived versus macrophage-derived, the most crucial questions are: (i) how are the phenotypic transi- tions of SMCs, macrophages, and other cell types regulated within lesions; (ii) what is the function of these phenotypically modulated cells; and (iii) how do these phenotypic transitions affect overall disease pathogenesis? To begin to address these questions, we gener- ated atherosclerosis-prone Apoe−/− mice with which we could lineage- trace SMCs and study the effects of SMC-specific conditional knockout of the stem cell pluripotency gene, KLF4. We chose to study the role of KLF4 as we and others have previously shown that this transcription factor plays a key role in regulating phenotypic transitions of SMCs in vivo during development20 and after carotid ligation injury21, as well as in vitro in cultured SMCs treated with platelet-derived growth factor (PDGF)-BB22,23, PDGF-DD24 or oxidized phospholipids25. RESULTS Most atherosclerotic plaque SMCs are not identified by ACTA2 SMCs are distinguished from other cell types by expression of a unique repertoire of genes including Acta2, Tagln, and Myh11. These genes are coordinately downregulated, at least in vitro, during SMC phenotypic switching such that they may be undetectable using tra- ditional immunohistochemical staining methods11,26. Therefore, to rigorously analyze the overall contributions of SMCs to lesion patho- genesis, we used a previously described27Myh11-CreERT2 ROSA floxed STOP eYFP Apoe−/− (SMC YFP+/+Apoe−/−) mouse model in which >95% of medial SMCs within large arteries were labeled with YFP (Supplementary Fig. 1a). Because Cre excision is per- manent, this SMC lineage-tracing model provides permanent YFP lineage tagging of virtually all mature (MYH11+) arterial SMCs that exist at the time of tamoxifen injection, allowing study of the fur- ther differentiation of these cells or of their progeny, irrespective of continued expression of Acta2, Myh11, or other SMC marker genes. To ensure the fidelity and SMC specificity of this lineage tracing model, we completed a number of further validation studies beyond those shown in our previous studies27. We demonstrated (i) SMC- specific YFP labeling within all tissue specimens examined, with no detectable expression of the YFP indicator gene in the absence of tamoxifen (Supplementary Figs. 1a,b and 2a); (ii) no detectable YFP+ cells by flow cytometry within blood or bone marrow prepara- tions (Supplementary Fig. 2a); (iii) no evidence of YFP+ cells within lesions of Western diet–fed mice not given tamoxifen (Supplementary Fig. 2b); (iv) no detectable YFP+ cells in the blood of the mice when fed a high-fat diet for 18 weeks (Supplementary Fig. 2c); and (v) YFP+ labeling of approximately 60% of freshly enzymatically dissociated cells from the aorta (the ascending and descending thoracic aorta plus the abdominal aorta to the iliac bifurcation, including adventitial and intimal cells) on the basis of flow cytometric analysis (Supplementary Fig. 2d). We harvested brachiocephalic arteries (BCAs) from SMC YFP+/+ Apoe−/− mice that had been injected with tamoxifen between 6–8 weeks of age, after an additional 18 weeks of Western diet. BCAs were immunostained for YFP, ACTA2, and LGALS3. Confocal micros- copy z-stacks of the collected tissues were acquired and analyzed to accurately profile individual cells (Fig. 1a–e). Notably, ~82% of SMCs within atherosclerotic lesions (YFP+DAPI+ cells) were ACTA2 nega- tive (Fig. 1a and Supplementary Table 1), indicating that the majority of SMCs within the lesion cannot be identified using traditional SMC markers. These results also showed that phenotypically modulated SMCs (YFP+ACTA2−DAPI+) comprised approximately 30% of the total cells within lesions (Fig. 1a and Supplementary Table 1), which far exceeds previous estimates based on ACTA2 immunostaining. SMCs within atherosclerotic plaques express markers of macrophages, MSCs, and myofibroblasts We found that phenotypically modulated (YFP+) SMCs within lesions expressed markers of macrophages (LGALS3) (Fig. 1c), MSCs (SCA1) (Fig. 1d), and myofibroblasts (ACTA2 and PDGFβR) (Fig. 1e). From these data we estimate the following distribution of SMC-derived cells within lesions: 30% macrophage-like cells (YFP+ACTA2−LGALS3+), 7% MSC-like cells (YFP+ACTA2−SCA1+), 12% myofibroblast-like cells (YFP+ACTA2+PDGFβR+), and 32–51% an indeterminate cell phenotype (YFP+ACTA2−LGALS3−SCA1−) (Supplementary Table 1). In addition, these analyses showed that 36% of LGALS3+ cells within advanced atherosclerotic lesions were YFP+, indicating that approximately one-third of cells that would normally be classified as macrophages in most previous studies in the field actually originated from SMCs, rather than from myeloid cells as previously assumed. These initial studies were performed in paraffin-embedded sam- ples to maintain the ultrastructure of the plaque and to determine the location of various SMC-derived cells within lesions. However, this technique limits the number of markers that can be simultane- ously examined. To further characterize SMC-derived plaque cells, we performed flow cytometric analyses of freshly dissociated cells from the aorta, from the aortic root through the iliac bifurcation. We found that substantial numbers of SMC-derived cells expressed multiple additional macrophage and hematopoietic markers (Fig. 1f). In particular, we identified YFP+ cells that co-expressed the monocyte/ macrophage marker ITGAM (CD11b)) and the mature macrophage marker F4/80, as well as YFP+ cells that co-expressed ITGAM and the dendritic cell marker ITGAX (CD11c). In addition, we found that 13% of the YFP+ cells co-expressed the MSC markers SCA1 and ENG (CD105) (Fig. 1f). When we analyzed MSC-like cells using traditional negative-gating strategies (CD45−CD34−CDH5−), we found that up to 45% of MSC-like cells within the aortas of 18-week Western diet–fed SMC YFP+/+Apoe−/− mice were YFP+ (Supplementary Fig. 3a), although it is unclear whether all of these cells were located within lesions or whether they also contributed to npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 3. A rt i c l e s 630 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine the population of adventitial SCA1+ cells that have previously been described by several groups28–32. Gating strategies for all flow cytom- etry experiments were determined on the basis of fluorescence minus one (FMO) controls (Supplementary Fig. 3b,c). To further assess the morphological and possible functional prop- erties of SMC-derived macrophage-like cells in vivo, we analyzed BCA lesions from SMC YFP+/+Apoe−/− mice by transmission elec- tron microscopy, combined with detection of YFP expression by immunogold labeling. As shown in Figure 2a and Supplementary Figure 4, we identified YFP+ cells containing multiple large lipid vacuoles that seem to be phagocytic. However, these cells were rela- tively rare, possibly reflecting the low frequency of phagocytic YFP+ cells at any given instant in time, the transient nature of this process, and/or technical difficulties in detecting these cells by immuno- electron microscopy. In addition, we flow-sorted SMC-derived MSC-like cells (YFP+ENG+SCA1+), non-SMC-derived MSC-like cells (YFP−ENG+SCA1+) and SMC-derived non-MSC-like cells (YFP+ENG−SCA1−) from 18-week Western diet–fed lineage-tracing mice to test the ability of these cell populations to differentiate into multiple lineages, including adipocytes and osteoblasts. After two passages in MSC maintenance medium, the SMC non-MSC-like cells became unhealthy in appearance and died (data not shown). The SMC-derived MSC-like cells survived but seemed to be senescent (Fig. 2b) and grew slowly (data not shown). These cells also failed to differentiate into either adipocytes (Fig. 2c,d) or osteoblasts (data not shown) when exposed to the appropriate differentiation culture medium. In contrast, the YFP− (non-SMC-derived) MSC-like cells grew well (data not shown) and showed high-efficiency differentiation into adipocytes (Fig. 2c,d) and osteoblasts (data not shown). These data indicate that although a subset of SMC-derived cells within atherosclerotic lesions express multiple markers of MSCs, these cells do not appear to be pluripotent and thus may not have the functional properties of MSCs. SMCs within human atheromas express the macrophage marker CD68 To independently detect phenotypically modulated SMCs that express LGALS3 and to validate a method for detecting these cells in human lesions, we used an in situ hybridization proximity ligation assay (ISH- PLA) recently developed by our laboratory27. This technique permits the identification of phenotypically modulated SMCs within fixed tissues based on the detection of the histone H3K4diMe in the Myh11 promoter (PLA+), a SMC-specific epigenetic signature that persists in cells that have no detectable expression of SMC markers27,33. We first validated this method by showing that YFP+LGALS3+ SMCs within the lineage-tracing mice retained this SMC-specific epigenetic signa- ture (Supplementary Fig. 5a). We also showed that neither cultured RAW 264-7 mouse macrophage cells (Supplementary Fig. 5b) nor human monocytes (Supplementary Fig. 5c) exposed to POVPC, an oxidative product of LDL that activates monocytes/macrophages34, exhibited H3K4diMe in the Myh11 promoter. To determine whether SMC transition to a macrophage-like state occurs in human lesions, we stained human coronary artery atheroscle- rotic lesions for CD68 and ACTA2 and performed ISH-PLA detection of MYH11 H3K4diMe. Multiple human coronary artery lesion sections from 12 human subjects were analyzed (Supplementary Fig. 5d). In these lesions, 18% of CD68+ cells were PLA+ (Fig. 3a–c), indicating that they are of SMC origin. To further validate these findings, we performed the ISH-PLA analysis in coronary artery samples from a male who had received a cross-gender heart transplant (Supplementary Fig. 6). Merge + DAPI YFPACTA2 LGALS3 c Merge + DAPISCA1 YFPACTA2 d YFP Merge + DAPIPDGFβR ACTA2 e ACTA2 YFP MergeDAPI ba 60K 40K 20K 0 10 0 10 1 10 2 YFP SSarea 60K 40K 20K 0 0 20K 40K 60K FSlin FS area 60K 40K 20K 0 0 20K 40K 60K SSarea F4/80 10 1 10 0 10 2 10 3 10 4 10 1 10 0 10 2 10 3 10 4 10 1 10 0 10 2 10 3 10 4 ITGAM ITGAX ITGAM ENG SCA1 CD34 10 1 10 0 10 2 10 3 10 4 PTPRC 10 3 10 4 10 0 10 1 10 2 10 3 10 4 10 0 10 1 10 2 10 3 10 4 10 0 10 1 10 2 10 3 10 4 10 0 10 1 10 2 10 3 10 4 10 0 10 1 10 2 10 3 10 4 10 0 10 1 10 2 10 3 10 4 60K 40K 20K 0 LGALS3 SSArea 60K 40K 20K 0 CDH5 SSArea f FS area Figure 1  Lineage tracing provides evidence for large populations of phenotypically modulated SMCs within lesions. (a–e) Representative immunofluorescence staining of BCAs from SMC YFP+/+Apoe−/− mice fed a Western diet for 18 weeks. The boxed region in a is shown at higher magnification in b. Shown are phenotypically modulated SMCs (YFP+ACTA2− cells highlighted in the white rectangle, b), macrophage-like SMCs (YFP+ACTA2−LGALS3+, c), mesenchymal stem cell-like SMCs (YFP+ACTA2−SCA1+, d), and myofibroblast-like cells (YFP+ACTA2+PDGFβR+, e). Yellow arrows in c–e indicate de-differentiated (YFP+ACTA2−) SMCs; white arrows indicate differentiated (YFP+ACTA2+) SMCs. Samples were either fixed and embedded in paraffin (a–c, e) or in Neg-50 frozen section medium (d). Scale bars, 50 µm in a and 10 µm in b–e. (f) Flow cytometry of single cell suspensions from aortas of SMC YFP+/+Apoe−/− mice treated with tamoxifen between 6–8 weeks and then fed a western diet for 18 weeks. Top, cells were gated for forward scatter (FS) area versus side scatter (SS) area to eliminate debris; for ‘singlets’ (single cells) based on a plot of forward scatter linear (FS lin) versus FS area; and for YFP. Middle, sub-populations of YFP+ cells were found to be double-positive for ITGAM (CD11b) and F4/80, for ITGAM (CD11b) and ITGAX (CD11c) and single-positive for LGALS3 (mac2). Bottom, after negative gating for CD34−PTPRC− and CDH5 cells, YFP+ cells were found to be double-positive for SCA1 and ENG. n = 6 mice per group. npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 4. a rt i c l e s nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015 631 WefoundMYH11H3K4diMePLA+CD68+ cellsthatwereYchromosome– negative (Fig.3d), consistent with the notion that these macrophage-like cells are of SMC and not hematopoietic origin. Notably, we never found MYH11 H3K4diMe PLA+ cells that were Y chromosome– positive (Fig. 3d and data not shown), demonstrating that myeloid cells do not acquire the MYH11 H3K4diMe SMC epigenetic signature in human atherosclerotic lesions. KLF4 has a critical role in regulating SMC phenotype and plaque pathogenesis We have previously shown that KLF4, a cell pluripotency factor35, is required for SMC phenotypic switching in several in vitro models25,36. However, there is as yet no evidence that SMC phenotypic transitions within atherosclerotic lesions are KLF4 dependent, and, if so, what role these transitions have in lesion pathogenesis. Consistent with our hypothesis that KLF4 regulates phenotypic transitions of SMCs within atherosclerotic lesions, we observed large numbers of YFP+ cells that expressed KLF4 within BCA lesions from 18-week Western diet–fed SMC YFP+/+Apoe−/− mice (Supplementary Fig. 7a). To determine whether KLF4 regulates SMC phenotypic transitions and overall lesion pathogenesis, we generated SMC YFP+/+Apoe−/− mice with SMC- specific deficiency of KLF4 (SMC YFP+/+Klf4∆/∆Apoe−/− mice) by cross- ing SMC YFP+/+Apoe−/− mice with Klf4fl/fl mice. We observed a high level of recombination of the floxed Klf4 alleles (Klf4∆/∆) tamoxifen treatment (Supplementary Fig. 7b,c), including what we estimate to be nearly 100% recombination within SMCs in the aorta when the data are corrected for the >40% of non-SMC DNA present in these samples, based on flow cytometric analysis of aortic cell populations (Supplementary Fig. 2b). Figure 4a shows representative confocal immunofluorescence images of the BCA lesions of control mice (SMC YFP+/+Klf4+/+Apoe−/−) and mice with loss of KLF4 in SMCs (SMC YFP+/+Klf4∆/∆Apoe−/−) after being fed a Western diet for 18 weeks. SMC YFP+/+Klf4∆/∆Apoe−/− mice had a nearly 50% reduction in plaque size (Fig. 4b) and multiple changes consistent with increased plaque stability, including a more-than-twofold increase in fibrous cap area (Fig. 4c), an increase in the percentage of ACTA2+ cells within the fibrous cap (Fig. 4d), and a reduced percentage of LGALS3+ cells (Fig. 4e) as compared to control SMC YFP+/+Klf4+/+Apoe−/− mice. SMC lineage-tracing analyses showed that loss of KLF4 within SMCs did not result in a change in the overall number of SMCs (YFP+ cells) within the lesions (Supplementary Fig. 8a), but it had major effects on SMC phenotypic transitions. These effects included a 53% decrease in the percentage of macrophage-like SMCs (YFP+LGALS3+/YFP+) within a lesion (Fig. 4e), a 70% decrease in the percentage of MSC-like SMCs (YFP+SCA1+/YFP+) within the medial area underlying lesions (Supplementary Fig. 8b), but no change in the percentage of MSC-like SMCsinthelesionitself(SupplementaryFig.8c)inSMCYFP+/+Klf4∆/∆ Apoe−/− mice as compared to control SMC YFP+/+Klf4+/+Apoe−/− mice. Consistent with these results, flow cytometric analyses showed a decrease in the percentage of SMC-derived MSC-like cells (YFP+ SCA1+ENG+CDH5−PTPRC−CD34−) (Supplementary Fig. 8d), but no change in either the overall percentage of MSCs (Supplementary Fig. 8e) or the percentage of YFP+ cells (Supplementary Fig. 8f) inSMCYFP+/+Klf4∆/∆Apoe−/− miceascomparedtocontrolSMC YFP+/+ Klf4+/+Apoe−/− mice. As noted above, SMC specific Klf4-knockout mice showed an increase in the percentage of ACTA2+ cells within the fibrous cap (Fig. 4d) as compared to wild-type control mice. Similarly, SMC YFP+/+Klf4∆/∆Apoe−/− mice showed an increase in the percentage of ACTA2+ cells within lesions (Fig. 4f), but reduced proliferation of SMC-derived cells (Fig. 4g) and a marked reduction in YFP+ SMC apoptosis as compared to control YFP+/+Klf4+/+Apoe−/− mice (Fig. 4h). These effects were not associated with changes in medial or luminal area (Supplementary Fig. 8g), or in the percentage of cells that were YFP+PDGFβR+ (Supplementary Fig. 8h) or YFP+ACTA2+ (Fig. 4f). In addition, we did not observe changes in cholesterol or triglyceride levels (Supplementary Fig. 8i). KLF4 modulates phenotypic transitions and functional properties of SMCs We have previously presented evidence that Klf4 is induced in cul- tured SMCs by treatment with oxidized phospholipids36 and that KLF4 suppresses expression of SMC marker genes through several mechanisms, including binding to the G/C repressor element found in most SMC marker gene promoters (including Acta2, Tagln, and Myh11) and inhibiting binding of the transcription factor SRF to CArG elements16,37,38. To determine whether similar mechanisms contribute to suppression of SMC marker gene expression within atherosclerotic lesions in vivo, we performed chromatin immuno- precipitation (ChIP) assays on chromatin extracted from the BCA regions of Apoe−/− mice carrying a transgene containing either an YFP + ENG + SCA1 + YFP – ENG + SCA1 + YFP YFP – ENG + SCA1 + YFP+ ENG+ SCA1+ DAPI FABP4 b c 60 d YFP– YFP + %Differentiation 50 40 30 20 10 0 Adipocyte differentiation * aFigure 2  Ultrastructural and functional characteristics of phenotypically modulated SMCs. (a) Immuno-transmission election microscopy of a BCA from a SMC YFP+/+Apoe−/− mouse fed a Western diet. A 10-nm gold bead–conjugated secondary antibody was used to reveal lipid laden YFP+ cells engulfing neighboring cells. The boxed area in the left micrograph is progressively enlarged in the center and right micrographs to enable visualization of the immunogold beads, highlighted with yellow arrows. 6–10 electron microscopic sections from three SMC YFP+/+Apoe−/− lineage-tracing mice were examined in these analyses. Additional micrographs from this experiment are shown in Supplementary Figure 4. Scale bars (from left to right): 2 µm, 0.5 µm, 0.2 µm. (b) Isolated YFP+ and YFP− MSCs (SCA1+ENG+) were cultured in mesenchymal stem cell medium. Representative images 10 d after plating are shown (green, YFP). Results are representative of two independent experiments using cells derived from the BCA region of a total of 10 Myh11-eYFP mice and sorted into the respective populations indicated. Scale bars, 50 µm. (c) YFP+ and YFP− MSCs (SCA1+ENG+) were incubated in adipogenesis differentiation medium and stained for the adipocyte marker FABP4. Scale bar, 200 µm. (d) Quantification of adipogenesis from five fields of view per group. Results are representative of two independent experiments. Error bars are means ± s.d. *P < 0.05 by Student’s t-test. npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 5. A rt i c l e s 632 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine unmutated Tagln promoter driving lacZ expression, or a Tagln pro- moter with a mutation of the GC repressor element driving lacZ expression. We found that compared to chow-fed Apoe−/− mice, Western diet–fed Apoe−/− mice showed marked enrichment of KLF4 binding to the Acta2, Tagln, Myh11, and Cnn1 endogenous promoters (Fig. 5a). Using the two transgenic strains described above, we showed that enhanced KLF4 binding to the Tagln pro- moter under Western diet conditions was dependent on the G/C repressor element (Fig. 5b). Finally, using the ISH-PLA assay, we showed that KLF4 bound to the Tagln promoter in individual phe- notypically modulated SMCs (YFP+ACTA2− cells) within lesions of Apoe−/− mice (Fig. 5c). Taken together, results provide compel- ling evidence that coordinated suppression of SMC marker gene expression is mediated by direct binding of KLF4 to the promoters of SMC marker genes. We21,23 and others39–42 have shown that KLF4 can act as either a transcriptional repressor or activator, depending on the cell type and gene locus. To more fully define the repertoire of KLF4 target genes that mediate SMC phenotypic switching, we performed KLF4 ChIP-seq analysis on chromatin samples derived from BCAs isolated from SMC YFP+/+Klf4+/+Apoe−/− versus SMC YFP+/+Klf4∆/∆Apoe−/− mice. This analysis required pooling of BCA samples from 16 mice per group to obtain sufficient DNA. Notably, we identified 869 KLF4 target genes that were selectively enriched in Klf4+/+Apoe−/− versus Klf4∆/∆Apoe−/− Western diet–fed mice (Fig. 5d and Supplementary Table 2). These putative KLF4 target genes that were selectively bound by KLF4 in SMCs include the SMC marker genes Acta2 and Tagln, thereby validating the fidelity of our ChIP-seq analysis. In addition, we found evidence of enriched KLF4 binding in genes within regula- tory pathways that are likely to be important in the pathogenesis of lesions, including gene families associated with phagocytosis, apop- tosis, cell migration, and inflammation (Fig. 5d and Supplementary Table 2). These putative KLF4 target genes probably contributed to the beneficial effects of SMC-specific loss of KLF4 on lesion size and plaque pathogenesis. KLF4 also bound regions near Itgal (CD11a), Itgax (CD11b), Itgam (CD11c) (Fig. 5d and Supplementary Table 2), and Arg1 (data not shown) in Klf4+/+ but not Klf4∆/∆ mice. We also noted 459 KLF4 targets enriched in Klf4∆/∆Apoe−/− samples as compared to Klf4+/+Apoe−/− samples (Supplementary Fig. 9 and Supplementary Table 3), presumably representing KLF4 target genes in non-SMCs that were altered as a secondary consequence of the loss of KLF4 in SMCs. iii a b c i ii iii i DAPI DAPI/ACTA2/PLA DAPI/ACTA2/PLA/CD68 PLA CD68 d DAPI DAPI/PLA/Y-Chr/CD68 DAPI/PLA/Y-Chr DAPI/PLA/Y-Chr/CD68 DAPI/PLA/Y-Chr Y-Chr DAPI/PLA/Y-Chr DAPI/PLA/Y-Chr/CD68 PLA CD68 ACTA2 ii 50 40 30 %ofpopulation 20 10 C D 68+ PLA+ /C D 68+ 0 Actual count Corrected count C D 68+ PLA+ AC TA2+ C D 68+ PLA+ Figure 3  SMCs within human coronary artery lesions express the macrophage marker CD68. SMCs within advanced atherosclerotic lesion specimens were identified by PLA detection of the SMC-specific stable epigenetic signature H3K4diMe in the MYH11 promoter. MYH11 H3K4diMe PLA+ cells exhibit a punctate red dot within the nucleus; the non-nuclear amorphous red staining is autofluorescence or nonspecific background. (a) Samples treated for PLA (red) were immunostained for CD68 (green), and DAPI (blue). Three distinct cell populations were identified, as shown at higher magnification and indicated with white arrows: (i) PLA+CD68− SMCs, (ii) PLA−CD68+ cells (macrophages of hematopoietic origin), and (iii) PLA+CD68+ SMC-derived macrophage-like cells. Scale bar, 100 µm. (b) Samples treated for PLA (red) were immunostained for ACTA2 (green), CD68 (cyan) and DAPI (blue). Shoulder regions within plaques showed a high incidence of SMC-derived macrophage-like cells (PLA+CD68+) (yellow arrows). Phenotypically modulated SMCs negative for CD68 (PLA+ACTA2−CD68−) were also observed in these regions (white arrows). Scale bars, 50 µm. (c) Quantitative analysis of SMC-derived macrophage-like cells within human coronary lesions without or with correction for the efficiency of PLA detection27. n = 12 human atherosclerotic right coronary arteries. Error bars are means ± s.e.m. (d) Combined epigenetic SMC and genetic HSC lineage tracing analysis of a cross-gender human heart transplant sample. Coronary artery specimens from a male patient who had received a female heart were processed for PLA (red), Y chromosome (Y-Chr) FISH (green), and CD68 staining (yellow). A PLA+ Y chromosome− CD68+ cell (yellow arrows) represents an SMC-derived macrophage-like cell not of hematopoietic origin. In contrast, a PLA−Y chromosome+CD68+ cell (red arrows) represents a macrophage of hematopoietic origin. A large number of cells of hematopoietic origin that are negative for CD68 (PLA−Y chromosome+CD68−) were also observed (white arrows). Scale bars, 50 µm. npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 6. a rt i c l e s nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015 633 Given our observation that SMC-specific loss of Klf4 is associ- ated with a marked reduction in the percentage of SMC-derived macrophage-like cells in Klf4∆/∆ mice as compared to wild-type Klf4+/+ control mice (Fig. 4f), we tested whether KLF4 is required for transition of cultured SMCs to a macrophage-like state in vitro following cholesterol loading12. Owing to recent controversies concerning the origin and purity of SMCs cultured from mouse vessels43,44, we used SMCs obtained by sorting freshly isolated YFP+ aortic cells from our SMC lineage-tracing mice. Primary cultures of aortic SMCs harvested from 9-week-old tamoxifen-injected SMC YFP+/+ Apoe−/− mice that had been injected with tamoxifen between 6–8 weeks of age were >98% YFP+ (Supplementary Fig. 10a), indi- cating that these cultured SMCs are derived from mature differen- tiated SMCs in vivo and not from a stem cell source, as has been speculated43,44. Cholesterol loading of cultured SMCs resulted in increased expression of Lgals3 (Fig. 6a) and Klf4 (Supplementary Fig. 10b) as compared to vehicle-loaded control cells; the increase in Lgals3 expression was abolished in aortic SMCs derived from SMC YFP+/+Klf4∆/∆ mice (Fig. 6a). Cholesterol loading was also associ- ated with increased phagocytic behavior of the cells as compared to vehicle-loaded control cells that was KLF4 dependent (Fig. 6b,c). Finally, cholesterol loading resulted in KLF4-dependent increases in the expression of the pro-inflammatory cytokines MCP1, CXCR1, STNFR1 (Supplementary Fig. 10c–e), Lgals3 mRNA (Fig. 6a), and the MSC markers Sca1 and Eng (Supplementary Fig. 10f,g) as compared to vehicle-loaded control cells. In contrast, the increase in Abca1 mRNA expression caused by cholesterol loading was not KLF4 dependent (Supplementary Fig. 10h). Global heterozygous knockout of Klf4 alters plaque pathogenesis Previous studies using VE-cadherin-Cre Klf4fl/flApoe−/− and LysMCre/Cre Klf4fl/flApoe−/− mouse models provided evidence that KLF4 has an atheroprotective role in endothelial cells and myeloid cells, respec- tively; KLF4 deficiency in these strains resulted in increased lesion size and changes consistent with enhanced inflammation41,42. Indeed, we confirmed the effects of myeloid-specific KLF4 deficiency by show- ing that LysM-Cre-dependent knockout of Klf4 was associated with increased lesion size (Supplementary Fig. 11a) and Sudan IV lipid staining (Supplementary Fig. 11b) in LysMCre/Cre ROSA STOP floxed eYFP Klf4fl/flApoe−/− mice. However, unlike previous studies41,42, we did not observe changes in triglyceride or total cholesterol levels (Supplementary Fig. 11c). We also found a reduced number of lesional LGALS3+ cells derived from LysM-Cre–expressing cells in LysMCre/Cre Klf4∆/∆Apoe−/− miceascomparedtowild-typecontrol LysMCre/CreKlf4+/+ Apoe−/− mice, based on YFP expression (Supplementary Fig. 11d). However, these results are equivocal regarding the identity of the YFP+ cell types affected by KLF4 deficiency because cells other than myeloid cells, including SMCs, may express the LysM-Cre transgene. a DAPI SMCKIf4+/+ eYFP +/+ Apoe –/– SMCKIf4∆/∆ eYFP+/+ Apoe –/– ACTA2 eYFP LGALS3 Merge e 40 ¥ NS*35 30 25 20 15 10 5 0 LGALS3 + YFP – LGALS3 + YFP + LGALS3 + YFP – YFP + DAPI + %ofpopulationwithinthe lesion f NS * * 30 25 20 15 10 5 0 ACTA2 + YFP + ACTA2 + YFP – ACTA2 + YFP – YFP + DAPI + %ofpopulationwithinthe lesion g MKi67 + YFP + MKi67 + YFP + MKi67 + YFP + YFP – DAPI + 14 12 10 8 6 4 NS * 2 0 %ofpopulationwithinthe lesion NS h 30 25 20 15 10 5 0 CASP3+ YFP + CASP3 + YFP + CASP3 + YFP + YFP – DAPI + NS * * %ofpopulationwithinthe lesion 250,000 b * 200,000 150,000 100,000 50,000 Lesionarea(µm2 ) 0 c *45 40 35 30 25 20 15 10 5 0 Fibrouscaparea/lesionarea d * 30 25 20 15 10 5 0 %ACTA2+ cellsinfibrouscap KIf4 +/+ KIf4 ∆/∆ Figure 4  SMC-specific Klf4 deficiency results in decreased lesion size and increased indices of plaque stability. SMC lineage tracing mice with or without conditional SMC-specific Klf4 deletion induced by treatment with tamoxifen at 6–8 weeks of age were fed a Western diet for 18 weeks before analysis. (a) Representative immunofluorescence staining of BCA lesions. Scale bars, 50 µm. (b–h) Quantitative data for BCA lesions: total lesion area (b), fibrous cap area (defined as the region of the lesion within 30 µm of the luminal surface) relative to the size of the lesion (c), the percentage of ACTA2+ cells within the fibrous cap (d), the percentages of the indicated populations of LGALS3+ cells within the lesion (e), the percentages of the indicated populations of ACTA2+ cells within the lesion (f), the percentages of the indicated populations of proliferating (MKi67+) cells within the lesion (g), and the percentages of the indicated populations of cells undergoing cell death (CASP3+) within the lesion (h). *P < 0.05, ¥P = 0.07, analysis completed by two-way analysis of variance (ANOVA) comparing genotype and distance from start of the BCA with a Tukey post-test, error bars show means ± s.e.m. Quantification is based on analysis of five 14,283-µm2 regions in each of three sections (d–g), or two sections (h) per mouse, from locations 150 µm, 450 µm and 750 µm from the start of the branch of the BCA from the aortic arch. SMC YFP+/+Klf4+/+Apoe−/−, n = 11; SMC YFP+/+Klf4∆/∆Apoe−/−, n = 8. CASP3, caspase-3; MKI67, marker of proliferation Ki67; NS, nonsignificant. npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 7. A rt i c l e s 634 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine Because the loss of Klf4 in SMCs had opposite effects on plaques to Klf4 loss in endothelial or myeloid cells, a key unresolved question is whether global conditional loss of Klf4 would be beneficial or detri- mental. To test this, we generated tamoxifen-inducible homozygous Klf4-knockout mice (ERT-Cre+ Klf4fl/flApoe−/−) and heterozygous Klf4-knockout mice (ERT-Cre+ Klf4fl//+Apoe−/−); the latter represent a model of partial inhibition of KLF4 across all cell types and thus mimic potential therapeutic approaches that would achieve partial suppression of KLF4. Unfortunately, conditional global homozygous Klf4 mice had to be euthanized at 8–10 weeks after Western diet feeding owing to excessive weight loss and the development of skin lesions; these effects are presumably due to the role of KLF4 in reg- ulating the proliferation and differentiation of epithelial cells45–48. Tamoxifen-treated ERT-Cre+Klf4fl/+Apoe−/− mice (ERT-Cre+Klf4∆/+ Apoe−/−) demonstrated 50% recombination (the maximum possible recombination for a heterozygous floxed allele) in the aorta, liver, and colon (Supplementary Fig. 12a,b), but exhibited no changes in body weight, heart weight, or cholesterol and triglyceride levels when com- pared to tamoxifen-treated ERT-Cre−Klf4fl/+Apoe−/− littermate control mice (Supplementary Fig. 12c). Notably, mice with conditional global heterozygous Klf4 loss had similar effects with respect to atherosclerotic lesions as those observed for SMC-specific conditional Klf4 loss, including a 30% decrease in lesion size (Supplementary Fig. 13a), and signs of increased plaque stability, namely increased ACTA2+ cap coverage (Supplementary Fig. 13b) and decreased LGALS3+ area (Supplementary Fig. 13c). In addition, conditional global heterozygous Klf4-knockout mice exhibited decreased intraplaque hemorrhage (Supplementary Fig. 13d) and decreased apoptosis and cell proliferation (Supplementary Fig. 13e,f) as compared to ERT-Cre−Klf4+/+Apoe−/− control mice. Taken together, these results indicate that global loss of one Klf4 allele has beneficial overall effects on plaque development, leading to smaller and possibly more stable lesions. DISCUSSION Despite numerous reports showing that cultured SMCs downregulate the expression of SMC differentiation marker genes after exposure to environmental cues present in atherosclerotic lesions, including cho- lesterol12, POVPC25, PDGF-BB49, and interleukin (IL)-1β (ref. 50), the field has relied almost entirely on the detection of these markers ¥7 6 5 4 3 2 1 0 Acta2 Cnn1Myh11Tagln * * # KLF4binding/Input a Chow diet Western diet KLF4binding/input Endogenous EndogenousTransgene Transgene Tagln WT Tagln G/C repressor mutant ¥ * * # 3.0 2.5 2.0 1.5 1.0 0.5 0 b Total genes: 869 Binding Apoptotic process Biological adhesion Biological regulation Cellular component organization Cellular process Developmental process Immune system process Localization Metabolic process Multicellular organismal process Reproduction Response to stimulus Macrophage activation Immune response Antigen processing and presentation Catalytic activity Enzyme regulator activity Nucleic acid binding transcription factor activity Protein binding transcription factor activity Receptor activity Structural molecule activity Translation regulator activity Transporter activity Total molecular function hits: 1033 Total genes: 869 Total biological process hits: 1798 Total genes: 72 Total immune process hits: 34 Total genes: 869 Cell junction Cell part Extracellular matrix Extracellular region Macromolecular complex Membrane Organelle Total cellular component hits: 363 d DAPI DAPI/eYFP/PLA ACTA2 Merge eYFP PLAc Figure 5  KLF4 binds to >800 genes within SMCs in advanced atherosclerotic lesions. (a) In vivo KLF4 binding to the indicated genes, as determined by ChIP, in BCAs obtained from Apoe−/− mice fed either a chow or Western diet for 18 weeks beginning at 8–9 weeks of age. P values are based on one-way ANOVA with a Tukey post hoc test. ¥P = 0.07, #P = 0.11, *P < 0.05. Error bars show means ± s.e.m. n = 3 independent pooled groups of five mice per treatment group. (b) Demonstration of binding of KLF4 to the Tagln promoter in BCAs in vivo, as determined by ChIP. Apoe−/− mice carrying a rat Tagln wild-type (WT, unmutated) transgene or Tagln G/C repressor mutant transgene were fed a chow or Western diet for 18 weeks beginning at 8–9 weeks of age. Quantitative ChIP assays measured KLF4 binding to the endogenous mouse Tagln gene relative to its binding to the WT or G/C repressor mutant rat transgenes. P values are based on two-way ANOVA with Tukey post-test. ¥P = 0.06, #P = 0.43, *P < 0.05. Error bars are means ± s.e.m. n = 3 independent pooled groups of five mice per treatment group. (c) KLF4 binding to the G/C repressor element of the Tagln promoter in vivo was detected by KLF4-Tagln ISH-PLA. The white arrow indicates a KLF4-PLA+ (red dot) YFP+ACTA2+ SMC (top enlarged image); the yellow arrow shows a KLF4−PLA+YFP+ACTA2− phenotypically modulated SMC (bottom enlarged image). Scale bars, 10 µm. (d) Aortic segments from the aortic root and aortic arch up to the carotid artery bifurcation from SMC Klf4+/+YFP+/+Apoe−/− (n = 16) or SMC Klf4∆/∆eYFP+/+Apoe−/− (n = 16) mice treated with tamoxifen between 6–8 weeks of age and then fed a Western diet for 18 weeks beginning at 8 weeks of age (pooled samples of chromatin from 16 mice per group) were used for ChIP-seq analysis of KLF4 binding targets. 869 targets were enriched in Western diet–fed SMC Klf4+/+eYFP+/+Apoe−/− mice as compared to Western diet–fed SMC Klf4∆/∆YFP+/+Apoe−/− mice, and thus they represent putative SMC KLF4 target genes. Pie charts generated by the PANTHER (Protein ANalysis THrough Evolutionary Relationships) classification system show functional annotation analyses divided into Gene Ontology (GO) terms corresponding to molecular function (top left), cellular component (top right), biological processes (bottom left) and immune processes (bottom right). npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 8. a rt i c l e s nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015 635 to ascertain whether a given cell within a lesion is a SMC. Indeed, this practice has contributed to the well-established dogma in the field that the role of SMCs within plaques is rather limited, albeit presumed to be beneficial by virtue of the role of phenotypically modulated SMCs in producing extracellular matrix and thereby contributing to fibrous cap formation. Herein we show that >80% of SMCs within BCA lesions are phenotypically modulated and thus would have been undetected by conventional techniques. These cells comprise ~30% of the total cellular composition of the lesion. Moreover, we show that phenotypically modulated SMCs transition to cells with multiple phenotypes within lesions, including cells that express markers of macrophages, MSCs, and/or myofibroblasts. Most notably, we show that these transitions are functionally important in that selective loss of Klf4 within SMCs results in reduced lesion size, increased fibrous cap thickness, and major reductions in the fraction of SMC- derived macrophage- and MSC-like cells, but causes an increase in the fraction of ACTA2+ cells within the fibrous cap. In addition, we show that cholesterol loading of cultured SMCs induces KLF4- dependent activation of macrophage and MSC markers, expression of pro-inflammatory cytokines, and increased phagocytic behavior. Finally, our in vivo KLF4 ChIP-seq analyses identified >800 putative KLF4 regulated genes within SMCs, including many associated with pro-inflammatory processes. Taken together, these results provide compelling evidence that transitions in SMCs phenotype have a crucial role in lesion development, plaque composition, and stabil- ity, and suggest that therapeutic approaches aimed at promoting beneficial changes in SMC phenotype may be a viable means of treating advanced atherosclerosis. A key question is how the loss of Klf4 within SMCs results in a marked reduction in overall lesion size as well as in multiple changes consistent with increased plaque stability. Our data indicate that these effects are not due to a change in the number of SMC-derived cells within lesions. Recent studies by the Fisher laboratory51 found that although some SMCs express macrophage markers after cholesterol loading in vitro, principal component analysis of microarray data revealed that these cells are distinctly different from classi- cal monocytes, macrophages, and dendritic cells. Moreover, these cells have reduced phagocytic capacity51. Notably, we found that SMC derived MSC-like lesional cells seem to be dysfunctional. Accordingly, we postulate that the loss of Klf4 within SMCs results in phenotypic transitions that have favorable effects in inhibiting plaque pathogenesis, including the loss of SMC-derived cells with ‘pro-inflammatory’ macrophage-like properties, and the gain of SMC-derived cells that contribute to plaque stabilization through mechanisms and functions yet to be defined. Although it has long been postulated that SMCs within lesions have a beneficial role, as indicated in many review articles9–11,26,52,53, our findings show that this is an oversimplification and that the effects of SMCs can vary dramatically depending on the nature of their pheno- typic transitions. A critical challenge for future studies will be to identify the environmental cues within advanced atherosclerotic lesions that regulate phenotypic transitions of SMCs, as well as the other major cell types within lesions, and to determine how these might be manipulated therapeutically to reduce plaque burden and increase plaque stability. Methods Methods and any associated references are available in the online version of the paper. Accession codes. The GEO accession number for KLF4 ChIP-seq data is GSE65812. Note: Any Supplementary Information and Source Data files are available in the online version of the paper. Acknowledgments We would like to acknowledge the valuable contributions of technicians M. McCanna and M. Bevard for their assistance in histological cutting and staining; R. Tripathi for assistance with cell culture work; J. Lannigan and c Nocholesterol80µg/mlcholesteroltreatment Bright field Klf4∆/∆ cellsKlf4+/+ cells eYFP LGALS3 Beads MergeDark field 3 a Lgals3/18S 2 1 0 0 80 Cholesterol (µg) * b 5 4 3 2 1 0 0 Cholesterol (µg) 80 * Klf4+/+ Klf4∆/∆ %YEP+ LGALS3+ cells containingbeads/total population Figure 6  KLF4-dependent effects in cholesterol-loaded, cultured SMCs. Aortic SMCs were isolated from SMC YFP+/+Klf4+/+ and SMC YFP+/+Klf4∆/∆ mice and sorted using a FACSVantage SE DIVA(Becton Dickinson) instrument to ensure a pure SMC (YFP+) cell population. (a) Lgals3 mRNA expression after cholesterol loading (80 µg/ml cholesterol for 72 h). P values are based on two-way ANOVA with a Tukey post-test. *P < 0.05. The data are normalized to Klf4+/+, 0 µg/ml cholesterol (vehicle control only). Error bars are means ± s.e.m. n = 3 independent experiments. (b,c) Klf4+/+ and Klf4∆/∆ cells were incubated with 0.8-µm polystyrene beads for 1.5 h after 72 h of cholesterol loading to induce a macrophage-like state. (b) Quantification of bead uptake in YFP+LGALS3+ cells. Fisher’s exact test was run on data normalized to the total number of viable cells collected by the cytometer in the Klf4+/+ group treated with 0 µg. P values are based on Fisher’s exact test (*P < 0.05). One representative experiment from two is shown. (c) Representative images showing staining for YFP (green), LGALS3 (yellow), beads (red), and dark field (magenta; this parameter provides an index of cell granularity or density). Scale bar, 5 µm. npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 9. A rt i c l e s 636 VOLUME 21 | NUMBER 6 | JUNE 2015  nature medicine M. Solga from the Flow Cytometry Core at the University of Virginia for their help designing flow cytometry panels and help running the cytometers; S. Guilot at the Advanced Microscopy Facility at University of Virginia for her help running the transmission electron microscope; J. Roithmayr, N. Hendley, M. Quetsch and M. Goodwin for their assistance in immunofluorescence image analysis; Y. Babiy for designing the cartoon in Supplementary Figure 6, and W. Evans for assistance in processing statistical analyses. We would also like to thank N. McGinn from the Department of Pathology, University of Virginia Hospital for the de-identified coronary arteries specimens; C. Murray from the University of Washington for the coronary artery heart transplant specimen; S. Offermanns from the Max Planck Institute for the Myh11-CreERT2 mice; K. Kaestner for the Klf4fl/fl mice; G. Randolph for the LysMCre/Cre mice; and A. Berns from the Netherlands Cancer Institute, Amsterdam, Netherlands, for the transgenic tamoxifen-inducible Cre (ERT-Cre) recombinase mice. This work was supported by US National Institutes of Health R01 grants HL057353, HL098538 and HL087867 to G.K.O., HL112904 to A.C.S., as well as a pilot grant from AstraZeneca as part of a University of Virginia–AstraZeneca Research Alliance to G.K.O.; and Mid-Atlantic American Heart Association fellowship grants 11PRE7170008 and 13POST17080043 to L.S.S. and D.G., respectively. AUTHOR CONTRIBUTIONS L.S.S. conducted experiments; performed data analysis; generated most of the experimental mice; performed immunostaining, image analysis and flow cytometry; and was primary writer of the manuscript. D.G. performed in vitro ChIP analysis and conceived and performed all the ISH-PLA experiments. M.S. generated Tagln wild-type lacZ Apoe−/− mice, performed in vivo ChIP assays, and performed immunohistochemistry and data analysis. O.A.C. was involved in designing experiments and data analysis; assisted in image analysis, cell culture and animal experiments throughout the project. A.C.S. and B.I. participated in designing the immuno-transmission electron microscopy protocol and analysis of images. R.M.H. performed the MSC immunostaining and image analysis, and the Klf4 ChIP-seq; conducted MSC differentiation experiments; and helped design cartoons. G.F.A. conducted data analysis of ChIP-seq experiments. P.S. assisted in cell culture experiments and analysis of data. A.A.C.N. performed PDGFβR staining and analysis. E.S.G. assisted in animal experiments, conducted statistical analyses and performed immunohistochemistry and data analysis. L.S.S., D.G., M.S., O.A.C., E.S.G., B.I., G.J.R. and G.K.O. participated in making final manuscript revisions. G.K.O. supervised the entire project and had a major role in experimental design, data interpretation, and writing the manuscript. COMPETING FINANCIAL INTERESTS The authors declare no competing financial interests. Reprints and permissions information is available online at http://www.nature.com/ reprints/index.html. 1. Saffitz, J.E. & Schwartz, C.J. Coronary atherosclerosis and thrombosis underlying acute myocardial infarction. Cardiol. Clin. 5, 21–30 (1987). 2. Libby, P. & Aikawa, M. Stabilization of atherosclerotic plaques: new mechanisms and clinical targets. Nat. Med. 8, 1257–1262 (2002). 3. Falk, E., Nakano, M., Benton, J.F., Finn, A.V. & Virmani, R. Update on acute coronary syndromes: the pathologists′ view. Eur. Heart J. 34, 719–728 (20123). 4. Falk, E., Shah, P.K. & Fuster, V. Coronary plaque disruption. Circulation 92, 657–671 (1995). 5. Virmani, R., Kolodgie, F.D., Burke, A.P., Farb, A. & Schwartz, S.M. Lessons from sudden coronary death: a comprehensive morphological classification scheme for atherosclerotic lesions. Arterioscler. Thromb. Vasc. Biol. 20, 1262–1275 (2000). 6. Lee, R.T. & Libby, P. The unstable atheroma. Arterioscler. Thromb. Vasc. Biol. 17, 1859–1867 (1997). 7. Ross, R. Atherosclerosis–an inflammatory disease. N. Engl. J. Med. 340, 115–126 (1999). 8. Libby, P. Inflammation in atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 32, 2045–2051 (2012). 9. Glass, C.K. & Witztum, J.L. Atherosclerosis: the road ahead. Cell 104, 503–516 (2001). 10. Libby, P., Ridker, P.M. & Hansson, G.K. Progress and challenges in translating the biology of atherosclerosis. Nature 473, 317–325 (2011). 11. Gomez, D. & Owens, G.K. Smooth muscle cell phenotypic switching in atherosclerosis. Cardiovasc. Res. 95, 156–164 (2012). 12. Rong, J.X., Shapiro, M., Trogan, E. & Fisher, E.A. Transdifferentiation of mouse aortic smooth muscle cells to a macrophage-like state after cholesterol loading. Proc. Natl. Acad. Sci. USA 100, 13531–13536 (2003). 13. Martin, K. et al. Thrombin stimulates smooth muscle cell differentiation from peripheral blood mononuclear cells via protease-activated receptor-1, RhoA, and myocardin. Circ. Res. 105, 214–218 (2009). 14. Caplice, N.M. et al. Smooth muscle cells in human coronary atherosclerosis can originate from cells administered at marrow transplantation. Proc. Natl. Acad. Sci. USA 100, 4754–4759 (2003). 15. Iwata, H. et al. Bone marrow-derived cells contribute to vascular inflammation but do not differentiate into smooth muscle cell lineages. Circulation 122, 2048–2057 (2010). 16. Wamhoff, B.R. et al. A G/C element mediates repression of the SM22α promoter within phenotypically modulated smooth muscle cells in experimental atherosclerosis. Circ. Res. 95, 981–988 (2004). 17. Feil, S. et al. Transdifferentiation of vascular smooth muscle cells to macrophage- like cells during atherogenesis. Circ. Res. 115, 662–667 (2014). 18. Swirski, F.K. & Nahrendorf, M. Do vascular smooth muscle cells differentiate to macrophages in atherosclerotic lesions? Circ. Res. 115, 605–606 (2014). 19. Allahverdian, S., Chehroudi, A.C., McManus, B.M., Abraham, T. & Francis, G.A. Contribution of intimal smooth muscle cells to cholesterol accumulation and macrophage-like cells in human atherosclerosis. Circulation 129, 1551–1559 (2014). 20. Cordes, K.R. et al. miR-145 and miR-143 regulate smooth muscle cell fate and plasticity. Nature 460, 705–710 (2009). 21. Yoshida, T., Kaestner, K.H. & Owens, G.K. Conditional deletion of Krüppel-like factor 4 delays downregulation of smooth muscle cell differentiation markers but accelerates neointimal formation following vascular injury. Circ. Res. 102, 1548–1557 (2008). 22. Deaton, R.A., Gan, Q. & Owens, G.K. Sp1-dependent activation of KLF4 is required for PDGF-BB–induced phenotypic modulation of smooth muscle. Am. J. Physiol. Heart Circ. Physiol. 296, H1027–H1037 (2009). 23. Yoshida, T., Gan, Q. & Owens, G.K. Kruppel-like factor 4, Elk-1, and histone deacetylases cooperatively suppress smooth muscle cell differentiation markers in response to oxidized phospholipids. Am. J. Physiol. Cell Physiol. 295, C1175–C1182 (2008). 24. Thomas, J.A. et al. PDGF-DD, a novel mediator of smooth muscle cell phenotypic modulation, is upregulated in endothelial cells exposed to atherosclerosis- prone flow patterns. Am. J. Physiol. Heart Circ. Physiol. 296, H442–H452 (2009). 25. Pidkovka, N.A. et al. Oxidized phospholipids induce phenotypic switching of vascular smooth muscle cells in vivo and in vitro. Circ. Res. 101, 792–801 (2007). 26. Alexander, M.R. & Owens, G.K. Epigenetic control of smooth muscle cell differentiation and phenotypic switching in vascular development and disease. Annu. Rev. Physiol. 74, 13–40 (2012). 27. Gomez, D., Shankman, L.S., Nguyen, A.T. & Owens, G.K. Detection of histone modifications at specific gene loci in single cells in histological sections. Nat. Methods 10, 171–177 (2013). 28. Klein, D., Benchellal, M., Kleff, V., Jakob, H.G. & Ergun, S. Hox genes are involved in vascular wall-resident multipotent stem cell differentiation into smooth muscle cells. Sci. Rep. 3, 2178 (2013). 29. Klein, D. et al. Vascular wall-resident CD44+ multipotent stem cells give rise to pericytes and smooth muscle cells and contribute to new vessel maturation. PLoS ONE 6, e20540 (2011). 30. Xiao, Q. et al. Sca-1+ progenitors derived from embryonic stem cells differentiate into endothelial cells capable of vascular repair after arterial injury. Arterioscler. Thromb. Vasc. Biol. 26, 2244–2251 (2006). 31. Xiao, Q., Zeng, L., Zhang, Z., Hu, Y. & Xu, Q. Stem cell-derived Sca-1+ progenitors differentiate into smooth muscle cells, which is mediated by collagen IV-integrin α1/β1/αv and PDGF receptor pathways. Am. J. Physiol. Cell Physiol. 292, C342–C352 (2007). 32. Passman, J.N. et al. A sonic hedgehog signaling domain in the arterial adventitia supports resident Sca1+ smooth muscle progenitor cells. Proc. Natl. Acad. Sci. USA 105, 9349–9354 (2008). 33. McDonald, O.G., Wamhoff, B.R., Hoofnagle, M.H. & Owens, G.K. Control of SRF binding to CArG box chromatin regulates smooth muscle gene expression in vivo. J. Clin. Invest. 116, 36–48 (2006). 34. Vladykovskaya, E. et al. Reductive metabolism increases the proinflammatory activity of aldehyde phospholipids. J. Lipid Res. 52, 2209–2225 (2011). 35. Takahashi, K. et al. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 131, 861–872 (2007). 36. Cherepanova, O.A. et al. Oxidized phospholipids induce type VIII collagen expression and vascular smooth muscle cell migration. Circ. Res. 104, 609–618 (2009). 37. Salmon, M., Gomez, D., Greene, E., Shankman, L. & Owens, G.K. Cooperative binding of KLF4, pELK-1, and HDAC2 to a G/C repressor element in the SM22α promoter mediates transcriptional silencing during SMC phenotypic switching in vivo. Circ. Res. 111, 685–696 (2012). 38. Regan, C.P., Adam, P.J., Madsen, C.S. & Owens, G.K. Molecular mechanisms of decreased smooth muscle differentiation marker expression after vascular injury. J. Clin. Invest. 106, 1139–1147 (2000). 39. Feinberg, M.W. et al. The Kruppel-like factor KLF4 is a critical regulator of monocyte differentiation. EMBO J. 26, 4138–4148 (2007). 40. Liao, X. et al. Kruppel-like factor 4 regulates macrophage polarization. J. Clin. Invest. 121, 2736–2749 (2011). 41. Sharma, N. et al. Myeloid Krüppel-like factor 4 deficiency augments atherogenesis in Apoe−/− mice–brief report. Arterioscler. Thromb. Vasc. Biol. 32, 2836–2838 (2012). 42. Zhou, G. et al. Endothelial Krüppel-like factor 4 protects against atherothrombosis in mice. J. Clin. Invest. 122, 4727–4731 (2012). npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 10. a rt i c l e s nature medicine  VOLUME 21 | NUMBER 6 | JUNE 2015 637 43. Nguyen, A.T. et al. Smooth muscle cell plasticity: fact or fiction? Circ. Res. 112, 17–22 (2013). 44. Tang, Z. et al. Differentiation of multipotent vascular stem cells contributes to vascular diseases. Nat. Commun. 3, 875 (2012). 45. Foster, K.W. et al. Induction of KLF4 in basal keratinocytes blocks the proliferation- differentiation switch and initiates squamous epithelial dysplasia. Oncogene 24, 1491–1500 (2005). 46. Jaubert, J., Cheng, J. & Segre, J.A. Ectopic expression of Krüppel-like factor 4 (Klf4) accelerates formation of the epidermal permeability barrier. Development 130, 2767–2777 (2003). 47. Katz, J.P. et al. The zinc-finger transcription factor Klf4 is required for terminal differentiation of goblet cells in the colon. Development 129, 2619–2628 (2002). 48. Katz, J.P. et al. Loss of Klf4 in mice causes altered proliferation and differentiation and precancerous changes in the adult stomach. Gastroenterology 128, 935–945 (2005). 49. Dandré, F. & Owens, G.K. Platelet-derived growth factor-BB and Ets-1 transcription factor negatively regulate transcription of multiple smooth muscle cell differentiation marker genes. Am. J. Physiol. Heart Circ. Physiol. 286, H2042–H2051 (2004). 50. Clément, N. et al. Notch3 and IL-1β exert opposing effects on a vascular smooth muscle cell inflammatory pathway in which NF-κB drives crosstalk. J. Cell Sci. 120, 3352–3361 (2007). 51. Vengrenyuk, Y. et al. Cholesterol loading reprograms the microRNA-143/145- myocardin axis to convert aortic smooth muscle cells to a dysfunctional macrophage- like phenotype. Arterioscler. Thromb. Vasc. Biol. 35, 535–546 (2015). 52. Owens, G.K. Regulation of differentiation of vascular smooth muscle cells. Physiol. Rev. 75, 487–517 (1995). 53. Owens, G.K., Kumar, M.S. & Wamhoff, B.R. Molecular regulation of vascular smooth muscle cell differentiation in development and disease. Physiol. Rev. 84, 767–801 (2004). npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 11. nature medicine doi:10.1038/nm.3866 ONLINE METHODS Mice. Animal protocols were approved by the University of Virginia Animal Care and Use Committee. Male Myh11-CreERT2 (ref. 54), LysMCre/Cre, Apoe−/−, ROSA26 STOP-flox eYFP+/+, Klf4fl/fl (ref. 47), and ERT-Cre (ref. 55) mice were used in this study. We also used male Apoe−/− transgenic mice carrying either an unmutated Tagln promoter-lacZ transgene, or a G/C repressor-mutated version of the Tagln promoter-lacZ transgene16. ROSA26 STOP-flox eYFP+/+ and Apoe−/− mice were obtained from Jackson Laboratories. Myh11-CreERT2, ROSA26 STOP-flox eYFP, Klf4fl/fl and ERT-Cre mice were genotyped by PCR as previously described27,47,54,55. For both the Myh11-CreERT2 and ERT-Cre mouse models, Cre recombinase was activated in male mice with a series of ten 1-mg tamoxifen (Sigma, cat. no. T-5648) intraperitoneal injections from 6 to 8 weeks of age, for a total of 10 mg of tamoxifen per mouse, which averaged 25 g of body weight during this 2-week period. Male littermate controls were used for all studies. Mice were fed a Western diet containing 21% milk fat and 0.15% choles- terol (Western diet, Harlan Teklad) for 18 weeks starting at the end of tamoxifen treatment. Irradiated mouse standard chow diet was purchased through Harlan (cat. no. 7012). Mice were euthanized by CO2 asphyxiation and then perfused via the left ventricle as follows: 5 ml PBS, 10 ml 4% paraformaldehyde, 5 ml PBS. Brachiocephalic arteries were carefully dissected and fixed for an additional hour in 4% paraformaldehyde before they were embedded in paraffin with the exception of SCA1-stained tissues, which were snap-frozen in Richard-Allan Scientific Neg-50 Frozen Section Medium (Thermo Scientific) before sectioning. Assays for determining total plasma cholesterol and triglyceride levels (Abbott Laboratories) were performed by the University of Virginia Clinical Pathology Laboratory. Mice were allocated to experimental groups based on genotyping and then randomized for the various experimental measurements. Any animal with triglyceride or cholesterol levels beyond 3 s.d. of the mean level for all mice were excluded from all future analyses. Analysis of atherosclerotic plaques. Paraffin-embedded brachiocephalic arter- ies (BCAs) were serially sectioned at a 10-µm thickness from the aortic arch to the right subclavian artery. For immunofluorescence analysis of LGALS3+ SMCs within the BCA, three sections were analyzed from a location 150 µm, 450 µm, and 750 µm from the start of the branch of the BCA from the aortic arch. Slides were stained with antibodies specific to GFP (Abcam ab6673), ACTA2 (Sigma F3777),LGALS3(CedarlaneCL8942AP),MKI67(Abcamab15580),CASP3(Cell Signaling 9661S), KLF4 (R&D Systems AF3158), MYH11 (Kamiya Biomedical Company MC-352), PDGFβR (Abcam ab32570), and SCA1 (Ly6A/E) (Abcam ab51317). Using a Zeiss LSM700 confocal microscope, a series of eight z-stack images of 1 µm in thickness were acquired for further analysis. Owing to variation in cellular composition in different regions of the lesion, cells from five standardized fields (two near the shoulder, one near the fibrous cap, and two near the media), each 14,283 µm2 in area, were counted to determine the cellular composition within each lesion. Close examination of each plane of the z-stack was conducted using Zen 2009 Light Edition Software (Zeiss) to ensure the presence of immunofluorescence staining coinciding with a single DAPI+ nucleus. The region of the lesion within 30 µm of the luminal boundary, as determined using Zen 2009 Light Edition Software, was analyzed to determine the cellular composition of the lesion cap, and the area of this region was com- pared to the entire area of the atherosclerotic lesion to determine cap area/lesion area. Morphometric analyses of lesion size were completed using ImagePro Plus (Media Cybernetics) as described in Alexander et al.56. Researchers were blinded to the genotype of the animals until the end of the analysis. Immuno-transmission electron microscopy. SMC YFP+/+Apoe−/− mice fed a Western diet for 18 weeks were euthanized by CO2 asphyxiation and perfusion fixed with 0.5% glutaraldehyde (Electron Microscopy Sciences 16300), 4% para- formaldehyde (Electron Microscopy Sciences 15700) in 1× PBS. Brachiocephalic arteries were isolated, frozen in liquid nitrogen, and sent to the University of Virginia Advanced Microscopy Core for processing. Grids were stained with an antibody specific to GFP (Abcam ab6673), followed by staining with a rab- bit anti-goat secondary conjugated to 10-nm gold beads (Electron Microscopy Sciences 25229). Images were captured using a JEOL 1230 transmission electron microscope with an ultra-high resolution capture camera (Gatan UltraScan 1000 2k × 2k CCD digital imaging camera). Flow cytometry. SMC YFP+/+Apoe−/− mice were euthanized by CO2 asphyxi- ation after 18 weeks of Western diet treatment. Mice were then perfused with 10 ml of PBS, and the aorta from the iliac bifurcation to the aortic root was gently cleaned of fat and fascia before removal from the animal. After removal from the animal, the heart was dissected away from the aortic root and the tis- sue was placed into an enzyme cocktail containing 4 U/ml Liberase TM (Roche 05401119001), 0.1 mg/ml DNase I, and 60 U/ml hyaluronidase in RPMI-1640. Once immersed in the digestion cocktail, the tissue was cut longitudinally, minced, and placed in a 37 °C incubator for 1.5 h. Cells were run through a 70-µm strainer and spun down at 500g for 5 min. Cells were resuspended in red blood cell lysis buffer (BD PharmLyse 555899) for two minutes and then inactivated using serum containing media, spun down again, and resuspended in 200 µl of 1× PBS. SMC-derived macrophage-like cells were identified using antibodies specific to F4/80 (eBioscience 17-4801), PTPRC (eBioscience 47- 0451), ITGAM (eBioscience 45-0112), DAPI (Invitrogen), LGALS3 (BioLegend 125405), and ITGAX (eBioscience 25-0114). SMC-derived MSC-like cells were identified using negative gating for PTPRC (eBioscience 12-0451-82), CDH5 (eBioscience 17-1441), and CD34 (BioLegend 128611), and using positive gating for ENG (eBioscience 48-1051-82) and SCA1 (eBioscience 45-5981-82). All samples were run on a Beckman Coulter CyAn ADP LX flow cytometer equipped with 405-nm, 488-nm, and 633-nm lasers. Human specimens. De-identified atherosclerotic coronary artery specimens from patients (n = 12) were collected during autopsy. These specimens were processed, fixed in paraformaldehyde, and paraffin-embedded blocks were cut into 5-µm sections. One coronary artery specimen was from a male patient who had received a heart from a female donor (n = 1). The Institutional Review Board at University of Virginia approved the use of all autopsy specimens. In situ hybridization proximity ligation assay (ISH-PLA). ISH-PLA was per- formed as previously described27. Briefly, human MYH11, mouse Myh11, and mouse Tagln probes were generated by nick translation (using Nick Translation kit no. 10976776001, Roche) using biotin-14-dATP (Invitrogen). Biotin labeled probes (40 ng per slide) underwent denaturation in hybridization buffer (2× SSC, 50% high-grade formamide, 10% dextran sulfate, 1 µg of human or mouse Cot-1 DNA) for 5 min at 80 °C. First, sections were stained for ACTA2 (Sigma F3777) and CD68 (Santa Cruz sc20060 KP1 clone) or LGALS3 (Cedarlane CL8942AP) (Fig. 3 and Supplementary Fig. 5) or GFP (Abcam ab6673) and ACTA2 (Sigma F3777) (Fig. 5). Briefly, slides were de-paraffinized and rehydrated in a xylene and ethanol series. After antigen retrieval (Vector no. H-3300), sections were blocked with fish skin gelatin oil in PBS (6 grams per liter) containing 10% horse serum for 1 h at room temperature (21 °C). Slides were incubated with primary antibodies for 1 h at room temperature, followed by incubation with donkey anti-mouse antibody conjugated with Alexa Fluor 647 (4 µg/ml, Invitrogen). Slides were then dehydrated in an ethanol series and incubated in 1mM EDTA (pH 8.0) for 20 min. Then, samples were incubated with pepsin (0.5%) in buffer (0.05 M Tris, 2 mM CaCl2, 0.01 M EDTA, 0.01 M NaCl) at 37 °C for 20 min, as previously described14. The hybridization mixture containing biotin-labeled probes (mMyh11, hMYH11 or mTagln) or a 5-TAMRA-dUTP–labeled Y chromo- some probe (clone RP11-88F4, Empire Genomics) was applied on the sections. Sections were then incubated at 80 °C for 5 min, followed by 16–24-h incubation at 37 °C. Hybridization was followed by multiple washes in 2× SSC, 0.1% NP-40 buffer. PLA was performed directly after ISH according to the manufacturer’s instructions (Olink) and as previously described27. Sections were incubated com- bining mouse H3K4dime (5 µg/ml, clone CMA303 Millipore) and rabbit biotin (5 µg/ml, no. ab53494, Abcam) antibodies (Fig. 3 and Supplementary Fig. 5) or mouse KLF4 [56CT5.1.6] (2.5 µg/ml, no. ab75486, Abcam) and rabbit biotin (5 µg/ml, no. ab53494, Abcam) antibodies (Fig. 5) overnight at 4 °C. The PLA amplification step was performed using the Duolink detection kit (emission wave- length: 555 nm). Finally, mounting medium with DAPI was used to coverslip the slides. Images were acquired with an Olympus BX41 microscope fitted with a Q imaging Retiga 2000R camera. Image acquisition was performed using Q Capture Pro software (Media Cybernetics & Q Imaging, Inc.). Settings were fixed at the beginning of both acquisition and analysis steps and were unchanged. Brightness and contrast were equally adjusted after merging. Image analysis was performed with ImageJ software. To estimate the percentage of SMC-derived npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 12. nature medicinedoi:10.1038/nm.3866 macrophage-like cells, we counted the number of CD68+PLA+ cells in CD68 positive staining areas of human coronary lesions (n = 12). We previously estimated that the efficiency of the ISH-PLA method in mouse and human tissues is 65% (ref. 27). Considering the incomplete efficiency of ISH-PLA, we applied a correction to the percentage of CD68+PLA+ cells to total CD68+ cells identified in human lesions. Both uncorrected and corrected percentages of CD68+PLA+ cells to total CD68+ cells are shown in Figure 3. The corrected values were proportionally calculated by dividing the uncorrected values by 0.65, the efficiency of the in situ hybridization portion of the PLA method (see Gomez et al.27). ChIP assays. Cell culture ChIP was performed as previously described57. Cells were fixed with 1% paraformaldehyde for 10 min at room temperature. Cross- linked chromatin was sheared by sonication into fragments of 200–600 base pairs. The sheared chromatin was immunoprecipitated with 2 µg of anti-H3K4diMe (clone CMA303, Millipore), or anti-KLF4 (Santa Cruz sc20691); negative control samples were incubated with mouse or rabbit IgG (Jackson ImmunoResearch Laboratories). Immune complexes were captured with magnetic bead-coupled protein G (Millipore). After elution and purification of the genomic DNA (gDNA), real-time PCR was performed on immunoprecipitated (IP) and non- immunoprecipitated (INPUT) gDNA. Primer sets used for the Myh11, Arg1 and Cox2 promoters were as previously described27,40. Results are expressed as the ratio of IP/INPUT. In vivo KLF4 ChIP assays were performed as previously described22,37 usingflash-frozenBCAsfrommicethathadbeenfedeither18weeks of high-fat or standard chow (Harlan 7012) diets beginning at 8 weeks of age. KLF4 chromatin immunoprecipitation–sequencing (ChIP-seq). Segments of the aorta from the arch to the aortic root and up to the carotid bifurcation isolated from 18 week western diet fed SMC Klf4+/+eYFP+/+Apoe−/− (n = 16), and SMC Klf4∆/∆eYFP+/+ Apoe−/− (n=16)miceweresnap-frozeninliquidnitrogenandthen processed as indicated above for ChIP assays using the KLF4-specific antibody. DNA library preparation and deep sequencing was performed by HudsonAlpha Institute for Biotechnology using the Illumina TruSeq Chip Library Kit according to the manufacturer’s protocol. Quality control and quantification of DNA and library were performed using an Agilent 2100 Bioanalyzer and a Kapa Library Quantification Kit (Kapa Biosystems) according to the manufacturer’s protocol. ChIP-seqdataprocessing.SequencingreadsfromanIlluminaHiSeqSequencing System were aligned to the mouse genome (mm10) using the BOWTIE align- ment tool58. These aligned reads were then processed and converted into bam/bai format (http://genome.ucsc.edu/goldenPath/help/bam.html), and then loaded in the Integrative Genomics Viewer (http://www.broadinstitute. org/igv/) for visualization. The processing steps involved removing duplicate reads and format conversions using the SAMtools59 suite. The reads were also converted to BED format (http://genome.ucsc.edu/FAQ/FAQformat#format1) for further data analysis processes such as peak calling. KLF4 peaks were identi- fied using MACS14 (ref. 60) with ChIP-seq BED files as input files and default settings with a P value for significant peak calling ≤ 1 × 10−5. Once peaks were obtained for both ChIP-seq data sets (SMC Klf4+/+eYFP+/+Apoe−/− and SMC Klf4∆/∆eYFP+/+Apoe−/−), BEDtools61 was used to remove peaks that were present in both data sets. Furthermore, BEDtools was used to identify the closest genes to each peak. If a gene was present in both data sets, the gene was removed from the analysis. Functional annotation was performed using PANTHER62 and Gene Ontology Consortium (http://geneontology.org/) software; a statistical over- representation test was performed using PANTHER62 (http://pantherdb.org/). The GEO accession number for these data is GSE65812. Cell culture studies. Primary mouse aortic SMCs were isolated using a previ- ously described protocol63 from 8-week-old SMC-lineage tracing mice after a series of ten tamoxifen injections. Cells isolated from SMC YFP+/+Klf4+/+ and SMC YFP+/+Klf4∆/∆ mice were passaged three times before undergoing cell sorting for YFP+ cells using a FACSVantage SE DIVA (Becton Dickinson). Cholesterol assays were performed using water-soluble cholesterol from Sigma (C4951-30MG) as previously published12 with a minor modification: cholesterol was reconstituted in DMEM-F12 media (Gibco) containing 0.2% FBS (Gibco), 100 U/ml penicillin/streptomycin (Gibco) and 1.6 mmol/liter l-glutamine, (Gibco). Cells were allowed to grow to ~70% confluency in DF10 (DMEM-F12 medium (Gibco) containing 10% FBS (Gibco), 100 U/ml penicillin/streptomycin (Gibco) and 1.6 mmol/liter l-glutamine (Gibco)) before being switched to 0, 20, 40, or 80 µg/ml cholesterol-containing medium. After 72 h, cells were harvested for mRNA, protein, or ChIP analysis. Total RNA was extracted from cultured cells using Trizol reagent (Invitrogen) according to the respective manufac- turer’s instructions. One microgram of RNA was used to perform a reverse transcription with iScript cDNA Synthesis Kit (Bio-Rad). A SensiFAST SYBR NO-ROX Mix (Bioline) was used to carry out RT-PCR on a C1000 Thermal Cycler CFX96TM (Bio-Rad) (Supplementary Table 4). Cell culture medium from 72-h cholesterol-treated Klf4+/+ or Klf4∆/∆ cells was filtered through 0.22-µm Millex low-protein-binding Durapore filters (Merck/Millipore) to remove traces of cholesterol. Then medium was concentrated ten times using Amicon Ultra centrifugal filter devices (Millipore). Cellular proteins for cytokine assay were isolated using the RIPA buffer protocol. Briefly, cells were incubated in RIPA buffer (Pierce) with Halt Protease Inhibitor cocktail (Thermo Scientific) at 4 °C for 15 min, centrifuged, and supernatants were used for analysis. Protein concentrations were measured using the Bio-Rad DC Protein Assay kit. Human coronary SMCs were purchased from Lonza and cultured in SMC maintaining media (Lonza). RAW264.7 mouse macrophages (American Type Culture Collection) and human monocytes (American Type Culture Collection) were cultured per the company’s recommendations. Bead uptake assays. Primary mouse aortic SMCs were treated with cholesterol as indicated above. After 72 h of cholesterol treatment, cells were switched back to DF10 culture medium containing 1.5% vol/vol 0.84-µm polystyrene beads (Spherotech FP-0870-2) for 1.5 h. Cells were then harvested, stained with LGALS3 (BioLegend 125405) and run on an Amnis ImageStreamX Mark II flow cytometer using a 60× objective. Data was analyzed using Amnis IDEAS software. Mesenchymal stem cell differentiation. Cells were isolated from SMC YFP+/+Apoe−/− mice fed a Western diet for 12 or 18 weeks as described in the flow cytometry section. Next, cells were stained and negatively gated for lineage markers (CD34 BioLegend 128611, PTPRC eBioscience 56-0451-80, CDH5 eBioscience 17-1441-80), and positively gated for ENG (eBioscience 48-1051-82) and SCA1 (eBioscience 45-5981-82). YFP was detected by native fluorescence. Cells were then sorted into four populations (YFP+SCA1+ENG+, YFP−SCA1+ENG+, YFP+SCA1−ENG−, and YFP−SCA1−ENG−) using a Becton Dickinson Influx Cell Sorter. Cells were then placed in StemXVivo MSC medium (R&D Systems CCM004) media at a density of 30,000 cells/mm2 and were pas- saged at 70% confluency (changing the medium every 2–3 d). YFP−SCA1+ENG+ and YFP+SCA1+ENG+ cells were expanded prior to plating for the differentia- tion experiments. The differentiation experiment was conducted using a kit from R&D Systems (mouse mesenchymal stem cell functional differentiation kit, SC010). Cells were differentiated in the appropriate media for 14 d (changing the medium every 2–3 d). The cells were then fixed and immunocytochemistry was performed using antibodies from the kit according to the protocol provided. Statistics. Fisher’s exact test was used for categorical data. Two-way ANOVA with Tukey post hoc tests were used for multiple group comparisons determined to have normal distribution by the Kolmogorov-Smirnov test. For experiments in which multiple sections of the BCA were assessed, two-way ANOVAs were conducted to determine if there were statistical differences between the dif- ferent regions of the BCA evaluated (i.e., 150 µm, 450 µm, and 750 µm from the branch of the BCA from the aortic arch). Because no statistical differences were seen between BCA regions, we calculated an average across all regions. Nonparametric data were analyzed using the Wilcoxon rank-sum test. There were no significant interactions between genotype and region (from the start of the BCA) for any of the endpoints analyzed by two-way ANOVA. P < 0.05 was considered significant. SAS v9.3 with Enterprise Guide v5.1 software (SAS Institute Inc.) was used for all statistical analyses. 54. Wirth, A. et al. G12–G13-LARG–mediated signaling in vascular smooth muscle is required for salt-induced hypertension. Nat. Med. 14, 64–68 (2008). npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 13. nature medicine doi:10.1038/nm.3866 59. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009). 60. Zhang, Y. et al. Model-based analysis of ChIP-seq (MACS). Genome Biol. 9, R137 (2008). 61. Quinlan, A.R. & Hall, I.M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010). 62. Mi, H., Muruganujan, A., Casagrande, J.T. & Thomas, P.D. Large-scale gene function analysis with the PANTHER classification system. Nat. Protoc. 8, 1551–1566 (2013). 63. Geisterfer, A.A., Peach, M.J. & Owens, G.K. Angiotensin II induces hypertrophy, not hyperplasia, of cultured rat aortic smooth muscle cells. Circ. Res. 62, 749–756 (1988). 55. Vooijs, M., Jonkers, J. & Berns, A. A highly efficient ligand-regulated Cre recombinase mouse line shows that loxP recombination is position dependent. EMBO Rep. 2, 292–297 (2001). 56. Alexander, M.R. et al. Genetic inactivation of IL-1 signaling enhances atherosclerotic plaque instability and reduces outward vessel remodeling in advanced atherosclerosis in mice. J. Clin. Invest. 122, 70–79 (2012). 57. McDonald, O.G. & Owens, G.K. Programming smooth muscle plasticity with chromatin dynamics. Circ. Res. 100, 1428–1441 (2007). 58. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S.L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009). npg©2015NatureAmerica,Inc.Allrightsreserved.
  • 14. Corrigendum: KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis Laura S Shankman, Delphine Gomez, Olga A Cherepanova, Morgan Salmon, Gabriel F Alencar, Ryan M Haskins, Pamela Swiatlowska, Alexandra A C Newman, Elizabeth S Greene, Adam C Straub, Brant Isakson, Gwendalyn J Randolph & Gary K Owens Nat. Med. 21, 628–637 (2015); published online 18 May 2015; corrected after print 12 August 2015 In the version of this article initially published, the labels to the left of the two micrographs in Figure 2c are reversed. Also, in Figure 4g, MKi67, used as a cell proliferation marker, is misspelled. The errors have been corrected in the HTML and PDF versions of the article. Corrigendum: Metformin activates a duodenal Ampk–dependent pathway to lower hepatic glucose production in rats Frank A Duca, Clémence D Côté, Brittany A Rasmussen, Melika Zadeh-Tahmasebi, Guy A Rutter, Beatrice M Filippi & Tony K T Lam Nat. Med. 21, 506–11 (2015); doi:10.1038/nm.3787; published online 06 April 2015; corrected after print 7 May 2015 In the version of this article initially published, we incorrectly reported the value for the particles per milliliter of Ad-dn-AMPK (D157A) used in the study. It was 3.1 × 10–9 PFU ml–1 and not 1.1 × 10–13 PFU ml–1 as originally reported. The errors have been corrected in the HTML and PDF versions of the article. Erratum: Cardiac RKIP induces a beneficial β-adrenoceptor–dependent positive inotropy Evelyn Schmid, Stefan Neef, Christopher Berlin, Angela Tomasovic, Katrin Kahlert, Peter Nordbeck, Katharina Deiss, Sabrina Denzinger, Sebastian Herrmann, Erich Wettwer, Markus Weidendorfer, Daniel Becker, Florian Schäfer, Nicole Wagner, Süleyman Ergün, Joachim P Schmitt, Hugo A Katus, Frank Weidemann, Ursula Ravens, Christoph Maack, Lutz Hein, Georg Ertl, Oliver J Müller, Lars S Maier, Martin J Lohse & Kristina Lorenz Nat. Med. 21, 1298–1306 (2015); published online 19 October 2015; corrected after print 29 October 2015 Intheversionofthisarticleinitiallypublished,therearetypographicalerrorsinthelabelstoFig.3e–h:‘WT+AAV9-eGFP’and‘WT+AAV9-RKIPWT’ areincorrect,andshouldbe‘RKIP– +AAV9-eGFP’and‘RKIP– +AAV9-RKIPWT’,respectively.TheerrorshavebeencorrectedintheHTMLandPDF versions of the article. Erratum: Snail1-induced partial epithelial-to-mesenchymal transition drives renal fibrosis in mice and can be targeted to reverse established disease. M Teresa Grande, Berta Sánchez-Laorden, Cristina López-Blau, Cristina A De Frutos, Agnès Boutet, Miguel Arévalo, R Grant Rowe, Stephen J Weiss, José M López-Novoa & M Angela Nieto Nat. Med. 21, 989–997 (2015); published online 03 August 2015; corrected after print 4 September 2015 In the version of this article initially published, the word ‘genetically’ was added by the editor to the penultimate sentence of the abstract during proofing of the manuscript. However, the authors used not only a genetic knockout approach but also morpholino-induced inhibition of proper mRNA processing to target Snail1 expression. Thus, the word ‘genetically’ has been deleted from the sentence to better convey the findings of the report. The error has been corrected in the HTML and PDF versions of the article. Erratum: Genomic landscape of carcinogen-induced and genetically induced mouse skin squamous cell carcinoma Dany Nassar, Mathilde Latil, Bram Boeckx, Diether Lambrechts & Cédric Blanpain Nat. Med. 21, 946–954 (2015); published online 13 July 2015, corrected after print 6 August 2015 In the published article, the format was listed as Article, but this is a Resource. The error has been corrected in the HTML and PDF versions of the article. CO R R I G E N DA AN D E R R ATA NATURE MEDICINE VOLUME 22 | NUMBER 2 | FEBRUARY 2016 217 npg©2016NatureAmerica,Inc.Allrightsreserved.