Next Article in Journal
Triple-Networked Hybrid Hydrogels Reinforced with Montmorillonite Clay and Graphene Nanoplatelets for Soft and Hard Tissue Regeneration
Next Article in Special Issue
Memory Macrophages
Previous Article in Journal
Simultaneous Inhibition of Histone Deacetylases and RNA Synthesis Enables Totipotency Reprogramming in Pig SCNT Embryos
Previous Article in Special Issue
Gut Barrier Damage and Gut Translocation of Pathogen Molecules in Lupus, an Impact of Innate Immunity (Macrophages and Neutrophils) in Autoimmune Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Macrophage-, Dendritic-, Smooth Muscle-, Endothelium-, and Stem Cells-Derived Foam Cells in Atherosclerosis

by
Malgorzata Kloc
1,2,3,*,
Jacek Z. Kubiak
4,5 and
Rafik M. Ghobrial
1,2
1
Transplant Immunology, The Houston Methodist Research Institute, Houston, TX 77030, USA
2
Department of Surgery, The Houston Methodist Hospital, Houston, TX 77030, USA
3
Department of Genetics, MD Anderson Cancer Center, The University of Texas, Houston, TX 77030, USA
4
Dynamics and Mechanics of Epithelia Group, Faculty of Medicine, Institute of Genetics and Development of Rennes, University of Rennes, CNRS, UMR 6290, 35043 Rennes, France
5
Laboratory of Molecular Oncology and Innovative Therapies, Department of Oncology, Military Institute of Medicine, 04-141 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 14154; https://doi.org/10.3390/ijms232214154
Submission received: 19 October 2022 / Revised: 10 November 2022 / Accepted: 14 November 2022 / Published: 16 November 2022

Abstract

:
Atherosclerosis is an inflammatory disease depending on the buildup, called plaque, of lipoproteins, cholesterol, extracellular matrix elements, and various types of immune and non-immune cells on the artery walls. Plaque development and growth lead to the narrowing of the blood vessel lumen, blocking blood flow, and eventually may lead to plaque burst and a blood clot. The prominent cellular components of atherosclerotic plaque are the foam cells, which, by trying to remove lipoprotein and cholesterol surplus, also participate in plaque development and rupture. Although the common knowledge is that the foam cells derive from macrophages, studies of the last decade clearly showed that macrophages are not the only cells able to form foam cells in atherosclerotic plaque. These findings give a new perspective on atherosclerotic plaque formation and composition and define new targets for anti-foam cell therapies for atherosclerosis prevention. This review gives a concise description of foam cells of different pedigrees and describes the main mechanisms participating in their formation and function.

1. Introduction

Atherosclerosis is a chronic inflammatory disease depending, among other factors (genetics, microbiome, pollution, stress, high blood pressure, smoking, sleep disorders, physical inactivity), on the accumulation of lipoproteins, which carry cholesterol, in the arterial wall [1,2]. The arterial wall contains several layers. The most internal (luminal) layer, called tunica intima, is composed of the endothelial layer, connective tissue, and internal elastic layer. The tunica media contains connective tissue and smooth muscle cells. The most external layer, called tunica externa or adventitia, is composed of an external elastic membrane, connective tissue, and fibroblasts (Figure 1). The onset of atherosclerosis and other cardiovascular disease starts from the hemodynamic, inflammatory, or lipoprotein-exposure injury, resulting in loss of integrity, homeostasis, and barrier function of the luminal endothelium [3,4]. Such a classical and commonly accepted view of atherosclerosis initiated from the luminal side of the wall is called the “inside-out” theory [5]. However, recently, more and more studies support the “outside-in” mechanism, where the perivascular adipose tissue (PVAT) surrounding the blood vessels plays a crucial role in atherosclerosis development [5,6]. PVAT, like other adipose tissues, is an endocrine organ. It produces hormones, growth factors, cytokines, and adipokines, which modulate energy production, glucose and lipid metabolism, and inflammation. The surplus of lipids and triglycerides stored in adipose tissue leads to hypertrophy and adipocyte dysfunction. Enlarged adipocytes promote the accumulation of inflammatory factors in PVAT and, eventually, adipocyte apoptosis. Cell-free DNA and fatty acids released from dying adipocytes induce stress pathways leading to the arterial wall cells injury, vascular dysfunction, and promoting the development of atherosclerosis [5,6]. No matter if the initial trigger comes from the inside or outside of the artery wall, upon crossing the damaged endothelial barrier and entry into the intima of the arterial wall, lipoproteins undergo a series of modifications, including oxidation of low-density lipoproteins (LDLs) to oxLDLs, which major bioactive component is the phospholipid lysophosphatidylcholine (LPC). The oxLDLs cause damage to the endothelium and induce apoptosis. In response to signals released from damaged endothelial cells, the immune cells, such as macrophages, infiltrate these sites trying to eliminate oxLDL, and as a result, participate in the buildup of fat-laden atherosclerotic plaques at the vessel wall interior [1,7,8]. Growing plaque gradually damages the artery wall, causing artery narrowing and blockage of blood flow. Eventually, the plaque can rupture, leading to a blood clot (Figure 1).
The atherosclerotic plaque has many acellular and cellular components [7,9]. The acellular components include lipids (cholesterol, cholesteryl esters, and phospholipids) and an extracellular matrix containing fibronectin, collagen, and proteoglycans. Cellular components include smooth muscle cells, macrophages, dendritic cells, mesenchymal stem cells, and T cells [9,10]. Macrophages, smooth muscle cells, dendritic cells, stem cells, and endothelial cells of the vessel wall internalize lipids to become foam cells, accelerating plaque development and atherosclerosis. One of the important factors related to foam cell formation and progression of inflammatory processes in atherosclerosis is a dysregulation of mitochondrial functions and mitophagy [11]. Studies of atherosclerotic patients showed the presence of several mutations in the leukocyte mitochondrial genome. Some of these mutations are related to oxidative phosphorylation, oxygen consumption, and defective mitophagy. Accumulation of such atherosclerosis-related mutations of mitochondrial DNA will stimulate pro-inflammatory responses against damaged or dysfunctional mitochondria leading to atherosclerosis development [11,12].
Foam cells are not limited to atherosclerosis, but also form during the inflammatory response to some pathogens and may initiate the development and progression of many cancers [13,14,15,16,17]. For example, in tuberculosis (TB), the infection of lung macrophages with Mycobacterium tuberculosis induces them to form lipid-laden foam cells. Infected macrophages recruit T cells and granulocytes. All these cells, together with the encasing fibrotic ring, form cellular foci called granulomata. Although phenotypically, the TB foam cells resemble the foam cells in atherosclerosis, they have different lipid composition (predominantly triglycerides and not cholesterol) and function. The lipid droplets in TB foam cells form not only due to increased uptake of extracellular lipids, but also metabolic changes and response to infection. TB foam cells produce inflammatory molecules, including TNFα and IL-1β, which fight the pathogen [15]. Many cancers contain tumor-specific macrophages (TAMs) [18]. Although the lipid-laden TAMs present in some tumors are often described as foam cells, they have lower lipid content than atherosclerotic foam cells. Another difference is that the accumulated lipids are mainly triglycerides, glycerophospholipids, and sphingomyelins, rather than cholesterol. Based on these differences, some authors discriminate them from the atherosclerotic foam cell by calling them “lipid-laden macrophages” [19].

2. What Are the Origin and Mechanisms of Foam Cell Formation?

Although the formation of foam cells differs in many details depending on cell pedigree, place of recruitment, and environment, the basic molecular mechanisms are the same or very similar. The first step in foam cell formation is the internalization of lipoproteins such as beta very low-density lipoprotein (βVLDL), acetylated low-density lipoprotein (ActLDL), and oxidized low-density lipoprotein (OxLDL) [20,21]. Studies showed that the particles of desialylated LDL are much smaller than the native LDL, which allows them easier penetration of the arterial wall [21]. The endothelium of the healthy arterial wall forms a tight barrier permeable only to water and small molecules below 6 nm in diameter. Larger molecules are transported across endothelium via transcellular transport called transcytosis [22,23]. There are two forms of transcytosis: receptor-mediated (indirect) and caveola-mediated (direct). Caveolae are small (50–100 nanometer) indentations of cellular membrane associated with caveolin (Cav-1) protein. Caveolae engulf LDL particles and transport them within caveola vesicles through the cytoplasm. Although the direct transcytosis was believed to be independent of receptors, recent studies showed that caveolae are associated with the scavenger receptor SRB-1 and endothelium-restricted TGF-β-type 1 receptor ALK1, which can help in the internalization of LDL by the caveolae [22,23]. In the indirect transcytosis, lipoproteins are recognized by and bind to the scavenger receptors (SRs) class B (CD36) and scavenger receptor-A (SR-A) at the cell surface [24,25,26,27,28,29,30,31,32]. The influx of lipids upregulates the expression of SRs and downregulates the expression of efflux transporters such as ATP-binding cassette A1 and G1 (ABCA1 and ABCG1) [33,34]. Receptor-bound lipoproteins are internalized via endocytosis [28,35,36]. Subsequently, scavenged lipoproteins enter the endosomal/lysosomal degradation pathway [37,38]. Lysosomal acid lipase (LPL) hydrolyzes cholesteryl esters (CE) into unesterified free cholesterol (FC), which translocates to the endoplasmic reticulum (ER) [39,40]; (Figure 2).
Macrophage-, dendritic cell-, smooth muscle cell- endothelial cell-, and stem cell-derived precursor of foam cell internalizes lipoproteins via a family of scavenger receptors (SR). The influx of lipoproteins upregulates the expression of SRs, leading to an enhanced accumulation of lipoproteins. SR-bound lipoproteins are internalized via endocytosis and subsequently enter the lysosomal pathway. Lysosomal enzymes hydrolyze cholesteryl esters into unesterified free cholesterol, which translocates to the endoplasmic reticulum (ER). In the ER, free cholesterol is re-esterified to form the lipid droplets that are released into the cytoplasm. Lipids accumulated in the foam cell precursor down-regulate expression of the reverse cholesterol transport (RCT) system proteins, which are responsible for cholesterol efflux from the cell. The increased influx and decreased efflux of lipoproteins lead to the development of lipid-laden foam cells. Foam cells produce various pro-inflammatory factors, which recruit immune cells, and metalloproteinase enzymes, which degrade the extracellular matrix. Eventually, these deteriorative changes in cell metabolism and homeostasis induce one of the cell-death programs (apoptosis, necroptosis, autophagy, or pyroptosis) that causes foam cell death.
Once in the ER, the free cholesterol (FC) undergoes re-esterification by acyl-CoA: cholesterol acyltransferase (ACAT1) [32]. The final product is stored as lipid droplets in the cell cytoplasm, giving the cell a foamy appearance. The foam cells secrete pro-inflammatory factors, which recruit additional immune cells promoting plaque development [41,42]. Additionally, the foam cells have a profibrotic activity. They release extracellular matrix metalloproteinases (MMPs), which degrade extracellular matrix proteins leading to tissue remodeling, atherosclerotic plaques instability, and rupture caused by release of proteases and cell necrosis [42,43]. The fate of foam cells is programmed (apoptosis) or unprogrammed (necroptosis) cell death, a lysosome-dependent degradation (autophagy), or inflammatory lytic programmed cell death (pyroptosis) [2,44,45,46,47,48,49,50]. Below, we describe the formation and function of the foam cells of different ancestry (Figure 3).

2.1. Macrophage-Derived Foam Cells

During atherosclerosis, in response to the signals, such as monocyte chemoattractant protein-1 (MCP-1) chemokine (C-C motif) ligands 3, 4, and 5 (CCL3, CCL4, CCL5) released from the endothelial cells damaged by the accumulating lipoproteins, monocytes migrate to the damaged subendothelial sites in the attempt to remove oxLDL. Once in the subendothelial space, which is rich in macrophage colony-stimulating factor (M-CSF) and granulocyte/macrophage colony-stimulating factor (GM-CSF), they differentiate into macrophages and dendritic cells (DCs), respectively. Macrophages differentiate further into different subtypes, including M1 and M2. It is still not clear which macrophage subtype converts into foam cells, but both M1 and M2 macrophages can do so [7,51]. Macrophages engulf modified lipoproteins via a set of scavenger receptors, such as SR-AI and SR-AII, with high affinity for oxLDL and actLDL. Other scavenger receptors involved in the internalization of modified lipoproteins and macrophage-derived foam cell formation are CD36, lectin-like oxLDL receptor-1 (LOX-1), and scavenger receptor for phosphatidylserine and oxidized lipoprotein (SR-PSOX), identical to chemokine CXCL16 [52,53]. The SP-PSOX is a specific receptor for oxLDL. Expression of SR-PSOX in endothelial cells and macrophages is regulated by pro-inflammatory factors interferon-gamma (IFN-y) and tumor necrosis factor-alpha (TNF-α) [50]. Once inside the macrophage, the modified lipoproteins enter the lysosomal pathway. In the lysosomes they are hydrolyzed into free cholesterol and fatty acids. Subsequently, free cholesterol is re-esterified into cholesterol esters by acylcholesterol transferase 1 (ACAT-1). Finally, the cholesterol esters emerge from the cisternae of the endoplasmic reticulum as the lipid droplets, which will be stored in the cytoplasm [7,54,55,56]. In normal cells, the homeostasis of cholesterol is tightly controlled by the balance between cholesterol influx and efflux [57]. However, the macrophages on the path to becoming the foam cells have highly dysregulated mechanisms controlling cholesterol efflux. Although some cholesterol can exit the cell by passive diffusion, the active pathway controlling the efflux and preventing excessive accumulation of cholesterol is called the reverse cholesterol transport (RCT) system [57,58]. The main participants of RCT, which mediate active transport of phospholipids and cholesterol out of the cell, are the Scavenger Receptor Class B type 1 (SR-B1), ATP-binding cassette transporter 1 (ABCA1), and ATP-binding cassette sub-family G member-1 (ABCG1) [25,59,60]. Recent studies of macrophage-derived foam cells showed that inhibition of miR-200b-3p reduces lipid accumulation and increases cholesterol efflux by targeting ABCA1 [61]. The expression of phospholipids and cholesterol exporters is regulated by the lipid-activated transcription factors: Liver X receptor (LXR) and peroxisome proliferator-activated receptor (PPAR). Studies of human macrophages showed that M2 macrophages under express ABCA1 and LXR. This reduces cholesterol efflux and enhances lipoprotein accumulation, which may lead to the formation of foam cells in atherosclerosis [62]. The role of LXR/PPAR signaling reaches beyond cholesterol homeostasis. It also plays a role in inflammation and immunity [63]. These attributes make the transporters and transcription factors involved in cholesterol homeostasis a good target for pharmacological manipulation to provide novel therapeutic approaches for the modulation of the foam cell formation/inflammatory processes in atherosclerosis.

2.2. Dendritic-Derived Foam Cells

Dendritic cells (DCs) were discovered by Steinman and Cohn [64] in 1973 in mouse spleen. Most information on dendritic cell ontogeny derives from mice studies. The macrophage and DC precursor (MDP) located in the bone marrow will differentiate into common monocyte precursor (cMoP), and two different precursors of DCs: conventional DCs (cDCs) and common DC (CDPs), which will generate plasmacytoid DCs (pDCs). The pDCs will terminally differentiate within the bone marrow while pre-DCs will exit the bone marrow, migrate through circulation, and settle in lymphoid and non-lymphoid organs. At destination, they differentiate, under the influence of specific transcription factors, into several subsets (expressing different markers) of cDCs [65]. The main function of DCs is initiation of antigen-specific adaptive immune responses while maintaining self-antigen tolerance. In mice and healthy young humans, DCs locate in the arterial intima especially in the areas prone to atherosclerosis, and heavily accumulate in atherosclerotic lesions [66]. The resident vascular DCs have been designated as ‘vascular DCs’ (VDCs). Vascular DCs, especially those concentrated in the subendothelial space, have a unique tubulovesicular system consisting hypertrophied Golgi and endoplasmic reticulum compartments [67]. The accumulation of DCs in the arterial intima depends on CCL2, CCL5, and CX3CL1 chemokines, and adhesion molecules such as P-selectin, E-selectin and vascular cell adhesion molecule-1 (VCAM-1). In mouse model, the removal of CX3CL1 receptor CX3CR1 prevents accumulation of DCs in the arterial wall [67,68]. Mouse studies showed that in even short, lasting several days, hypercholesterolemia, the arterial intima DCs accumulate lipids acquiring foam-cell phenotype and initiating plaque formation [69]. The main form of low-density lipoprotein (LDL) in the atherosclerotic plaque is aggregated LDL (agLDL), which is bound to the subendothelial matrix. In macrophages, the internalization of agLDL starts with the invaginations of the cell surface in which extracellular agLDL is digested by exocytosed lysosomal enzymes. They called this highly acidic hydrolytic compartment a lysosomal synapse [70]. In the following studies, Haka et al. [71] described how the different subsets of human and mouse DCs interact with agLDL to form foam cells. They found that maturing monocyte-derived DCs upregulate their ability for exophagy and degradation of agLDL, which is followed by lipid accumulation and the formation of foam cells. Studies of in vitro generated monocyte-derived DCs showed that lipid accumulation and formation of foamy DCs is regulated by a proinflammatory cytokine Interleukin (IL)-17A [70]. IL-17A is not only involved in the innate and adaptive immune response, but also in chronic inflammatory and immunometabolic diseases. Exposure of DCs to IL-71A dramatically (2–12 times) increased the amount of internalized cholesterol, cholesteryl esters, triglycerides, and phospholipids. Electron microscopy analysis showed that DCs accumulated lipid droplets by capturing extracellular lipids. Increased lipid uptake was facilitated by the IL-17A-dependent overexpression of scavenger receptors MSRI and CD68, fatty acid transport protein FATP1, and low-density lipoprotein receptor LRP [72]. The transcriptional programs regulating lipid homeostasis and metabolism in foam DCs and foam macrophages are controlled by the nuclear receptor LXR-a/NR1H3, which affects expression of target genes involved in lipid metabolism such as cholesterol transporter ABCA1. Salvatore et al. [72] showed that foamy DCs expressed NR1H3 and ABCA1 at much higher level than regular DCs. Interestingly, they also showed that the foamy DCs retain, like regular DCs, the ability to stimulate proliferation of allogeneic T cells. Although animal and human studies suggest that DCs beside a modulating adaptive immune response control cholesterol metabolism and lipid uptake, the exact molecular mechanisms of DCs engagement in development and progression of atherosclerosis are still very vague.
For many years, a commonly accepted paradigm was that foam cells derive from myeloid cell lineage. However, this paradigm was called into question by the findings that the vascular smooth muscle cells, endothelial cells, and stem/progenitor cells also express scavenger receptors, cholesterol transporters, and pro-inflammatory factors, suggesting that they can accumulate lipids.

2.3. Smooth Muscle Cell-Derived Foam Cells

Smooth muscle cells (SMCs) of the arterial wall play a major role in the deposition of cholesterol in atherosclerotic plaques. In human atherosclerosis at least 50% of all foam cells are SMC-derived [7,73]. SMC lineage cells tracking studies in mice showed that they represent at least 1/3 of all cells present in atherosclerotic plaque [73,74]. Studies on mouse and human arterial SMCs showed that internalization of cholesterol leads to the loss of expression of SMC specific genes and the induction of macrophage specific and proinflammatory markers (CD68 and Mac2), and that eliminating cholesterol excess reverses these changes [73,75,76]. A key regulator of SMC conversion process into macrophage-like cells is regulated by the Kruppel-like factor 4 (KLF-4). SMC-specific knock down of KLF-4 prevents suppression of SMC-specific genes resulting in fewer macrophage-like cells [74]. Interestingly, the acquisition of macrophage specific markers does not transform SMCs into real macrophages as they remain deficient in phagocytosis and efferocytosis [73,76,77]. It seems that the mechanisms of lipid uptake and storage by SMCs and macrophages are quite different. While macrophages and dendritic cells uptake agLDL via acidified lysosomal synapse (see above; [70]), the SMCs use LDL receptor-related protein 1 (LRP1) pathway, which does not involve endocytosis [78]. The SMC-derived foam cells accumulate lipids because they do not express the cholesterol exporter, the ATP-binding cassette transporter A1 (ABCA1), which allows phospholipid and cholesterol exit from the cell by binding free Apolipoprotein A-I (apoA-I) to these lipids [73,79]. Some studies indicate that SMC-derived foam cells are less effective than myeloid-derived foam cells in coping with lipid-loading [7]. The future challenge is to further identify the molecular and cellular specificities of SMCs-derived foam cells and commonalities and differences with macrophage-, dendritic-, and endothelial-derived foam cells. This will allow for the development of novel anti-SMC-derived foam cells therapies.

2.4. Endothelial Cell-Derived Foam Cells

The endothelial cells (ECs) of the vessel wall are the first cells experiencing the impact of hyperlipidemic blood serum. The membrane of hyperlipidemia-activated ECs alters, becoming more permeable to lipoproteins. Gradually, ECs, by accumulating more and more lipid droplets, transform into foam cells expressing adhesion molecules (VCAM-1, VLA-4). Additionally, in response to increased accumulation of oxLDL, they increase the synthesis of cytoprotective (anti-apoptotic and antioxidant) heat shock proteins (Hsp27, Hsp70, and Hsp90) [80]. Corrêa et al. [2], showed that the main component of oxLDL, the phospholipid lysophosphatidylcholine (LPC), induces endothelial cells and macrophages to form foam cells through the NOD-, LRR- and pyrin domain-containing protein 3 (NLRP3) inflammasome-mediated pathway. The NLRP3 is an intracellular sensor protein that detects danger signals, resulting in the formation and activation of the NLRP3 inflammasome and the release of pro-inflammatory cytokine IL-1β [81]. The information on endothelia-derived foam cells is limited so further studies on the subject are necessary.

2.5. Stem/Progenitor Cells-Derived Foam Cells

The walls of healthy and diseased arteries contain a subpopulation of multipotent mesenchymal stem cells (MSCs), also called stem/progenitor cells (SPCs), which have tissue repair/regeneration potential [82,83,84]. There are two sources of MSCs located in the arterial wall. The first source is the bone marrow-derived stem cells, which, after entering circulation, infiltrate various injured organs and tissues, including a wall of diseased arteries, where they promote regeneration processes. The second source is the mesenchymal stem cells residing in the wall of healthy arteries and microvessels. These resident MSCs also can enter circulation and eventually infiltrate other organs and tissues, becoming the progenitor cells for regeneration and repair [10]. Recently, Jiang et al. [85], showed that the subpopulation of CD34+ mesenchymal cells, which do not derive from the bone marrow, can acquire endothelial cell fate in the injured artery, while the bone marrow derived MSCs differentiate into immune cells. Additionally, the CD34+ cells isolated from vascular adventitia can also differentiate into endothelial cells, thus enhancing repair processes. Further studies on the CD34+ progenitor cells showed that human platelets recruit these cells through the adhesion receptors P-selectin/PSGL-1 and beta1- and beta2-integrin pathways, and induce their differentiation into foam cells and endothelial cells. The formation of platelet-induced foam cells is inhibited by HMG coenzyme A reductase inhibitors and the agonists of peroxisome proliferator-activated receptor-alpha and -gamma, which decrease the expression/secretion of matrix metalloproteinase-9 (MMP-9) [86]. Another study on CD34+ progenitor cells showed that the migration and differentiation of these cells into foam cells and macrophages is regulated by the chemokine stromal cell-derived factor 1 (SDF-1). Co-culturing of activated platelets expressing SDF-1 with human CD34+ progenitor cells resulted in platelets’ phagocytosis by CD34+ cells, and the conversion of CD34+ cells into CD68+macrophages and foam cells. The anti-SDF-1 antibody inhibited platelet-induced foam cell formation and MMP-9 expression. Authors suggest that the phagocytosis of platelets by CD34+ progenitor cells in SDF-1dependent manner and their differentiation into foam cells contributes to the progression of atherosclerosis [87]. Chen et al. [88] studied adventitial SPCs, expressing stem cell antigen-1 (Sca-1+), in the mouse model of atherosclerosis. They showed that transplanted adventitial SPCs infiltrated arterial intima, and by converting into foam cells expanded the size of neointima. They also showed that the Sca-1+ SPCs express CCR2 and CXCR2 receptors and that they migrate into neointima in response to the respective chemokine ligands CCL2 (chemokine (C-C motif) ligand 2) and CXCL1 (chemokine (C-X-C motif) ligand 1). Due to the possible importance of stem cells in the development and progression of atherosclerosis and the lack of information on the exact mechanisms and molecules involved in formation of foam cells from stem/progenitor cells, further studies are urgently needed on this subject.
It is well established that at least 50% of all foam cells derive from smooth muscle cells, but there is very little information on the proportion of endothelial cells and stem cells that become foam cells. Although the proportion of endothelial- and stem cell-derived is likely low, this requires further study.

3. Conclusions and Future Directions

It is quite obvious that not only all lipid-laden cells, but also the foam cells, differ in origin and function. The multi-progenitor origin of foam cells requires further studies on the progenitor-specific versus common mechanisms regulating foam cell formation, development, and exact function. Independently of the origin, all foam cells have hyperactive lipoprotein influx and defective efflux machinery; therefore, one of the potential therapeutic approaches would be trying to down-regulate or shut down scavenger receptors responsible for the influx, and upregulate the expression of efflux transporters. Another approach would be to manipulate the pro-inflammatory cytokines, such as interleukin-17A, which besides being involved in the inflammatory responses, increase the amount of lipoproteins internalized by cells. The caveat is that scavenger receptors, efflux transporters, and inflammatory cytokines are necessary for the normal function and homeostasis of the non-foam cells. Thus, the future challenge will be to find importers, exporters, and cytokines specific to the particular type of foam cells. Another recently promoted approach derived from the “outside-in” theory, or atherosclerosis initiation and development, would be preventive targeting of adipose perivascular tissue. There is also a recent focus on mitochondrial dysfunction as a therapeutic target for the treatment of atherosclerosis. The mutation of mitochondrial DNA and associated mitochondrial dysfunction in the cells of the arterial wall are involved in atherosclerosis development and progression. Some of these mutations are already used as atherosclerosis markers, and several centers study the effect of antioxidant therapies for improving mitochondrial function, and mitochondria-targeting therapies in a clinical setting.
However, these novel therapeutic targets and approaches require further studies on the factors and mechanisms involved.

Author Contributions

Conceptualization, M.K.; writing, M.K., J.Z.K., and R.M.G.; figures, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Some drawings used to make figures were from the Servier Medical ART: SMART, smart.servier.com (accessed on 1 November 2022).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbCA1ATP-binding cassette A1 efflux transporter
AbCG1ATP-binding cassette G1 efflux transporter
ACAT1acyl-CoA: cholesterol acyltransferase
ActLDLacetylated low-density lipoprotein
agLDLaggregated low-density lipoprotein
apoA-Iapolipoprotein A-I
βVLDLbeta very low-density lipoprotein
CCL2C-C motif chemokine ligand 2
CCL3C-C motif chemokine ligand 3
CCL4C-C motif chemokine ligand 4
CCL5C-C motif chemokine ligand 5
CXCL16C-C motif chemokine ligand 16
CCR2CC motif chemokine receptor 2
CD34cluster of differentiation 34 protein
CD36scavenger receptor class B
CD68cluster of differentiation 68 protein
cDCconventional DC
CDPcommon DC precursor
CEcholesteryl ester
CX3CL1C-X3-C motif chemokine ligand 1
CX3CR1C-X3-C motif chemokine receptor1
CXCR2C-X-C motif chemokine receptor 2
DCdendritic cell
ECendothelial cell
ERendoplasmic reticulum
FATP1fatty acid transport protein-1
FCfree cholesterol
GM-CSFgranulocyte/macrophage colony-stimulating factor
Hspheat shock protein
IL-17Ainterleukin-17A
IL-1βinterleukin-1β
IFN-yinterferon-gamma
KLF-4Kruppel-like factor 4
LDLlow-density lipoprotein
LOX-1lectin-like oxLDL receptor-1
LPCphospholipid lysophosphatidylcholine
LPLlysosomal acid lipase
LRP1low-density lipoprotein receptor-related protein 1
LPRlow-density lipoprotein receptor
LXRliver X receptor
M1pro-inflammatory macrophage subtype
M2ant-inflammatory macrophage subtype
Mac2galectin-3 protein
MCP-1monocyte chemoattractant protein-1
M-CSFmacrophage colony-stimulating factor
MDPmacrophage and dendritic cell precursor
MHG3-hydroxy-3-methyl-glutaryl-coenzyme A reductase
MMPmatrix metalloproteinase
MMP-9matrix metalloproteinase-9
MSCmesenchymal stem cell
MSRImacrophage scavenger receptor type I
NLRP3NOD-, LRR- and pyrin domain-containing protein 3
NR1H3nuclear receptor subfamily 1 group H member 3
oxLDLoxidized low-density lipoprotein
pDCplasmacytoid DC
PPARperoxisome proliferator-activated receptor
PSGL-1P-selectin glycoprotein ligand-1
RCTreverse cholesterol transport system
PVATperivascular adipose tissue
SCa-1stem cell antigen-1
SDF-1stromal cell-derived factor 1
SMCsmooth muscle cell
SPCstem/progenitor cell
SRscavenger receptor
SR-Ascavenger receptor class A
SR-AIscavenger receptor class A type I
SR-AIIscavenger receptor class A type II
SR-PSOXscavenger receptor for phosphatidylserine and oxidized lipoprotein
TAMtumor-specific macrophage
TBtuberculosis
TNFαtumor necrosis factor alpha
VCAM-1vascular cell adhesion molecule-1
VDCvascular DC
VLA-4very late antigen-4 integrin

References

  1. Libby, P. The changing landscape of atherosclerosis. Nature 2021, 592, 524–533. [Google Scholar] [CrossRef] [PubMed]
  2. Corrêa, R.; Silva, L.F.F.; Ribeiro, D.J.S.; Almeida, R.D.N.; Santos, I.O.; Corrêa, L.H.; de Sant’Ana, L.P.; Assunção, L.S.; Bozza, P.T.; Magalhães, K.G. Lysophosphatidylcholine Induces NLRP3 Inflammasome-Mediated Foam Cell Formation and Pyroptosis in Human Monocytes and Endothelial Cells. Front. Immunol. 2020, 10, 2927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Stanek, A.; Fazeli, B.; Bartuś, S.; Sutkowska, E. The Role of Endothelium in Physiological and Pathological States: New Data. BioMed Res. Int. 2018, 2018, 1098039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Xu, S.; Ilyas, I.; Little, P.J.; Li, H.; Kamato, D.; Zheng, X.; Luo, S.; Li, Z.; Liu, P.; Han, J.; et al. Endothelial Dysfunction in Atherosclerotic Cardiovascular Diseases and Beyond: From Mechanism to Phar-macotherapies. Pharmacol. Rev. 2021, 73, 924–967. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, H.W.; Shi, H.; Winkler, M.A.; Lee, R.; Weintraub, N.L. Perivascular Adipose Tissue and Vascular Perturbation/Atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 2569–2576. [Google Scholar] [CrossRef] [PubMed]
  6. Stanek, A.; Brożyna-Tkaczyk, K.; Myśliński, W. The Role of Obesity-Induced Perivascular Adipose Tissue (PVAT) Dysfunction in Vascular Homeostasis. Nutrients 2021, 13, 3843. [Google Scholar] [CrossRef]
  7. Maguire, E.M.; Pearce, S.W.A.; Xiao, Q. Foam cell formation: A new target for fighting atherosclerosis and cardiovascular disease. Vascul. Pharmacol. 2019, 112, 54–71. [Google Scholar] [CrossRef]
  8. Taki, A.; Kermani, A.; Ranjbarnavazi, S.M.; Pourmodheji, A. Overview of Different Medical Imaging Techniques for the Identification of Coronary Atherosclerotic Plaques. In Computing and Visualization for Intravascular Imaging and Computer-Assisted Stenting; Balocco, S., Zuluaga, M.A., Zahnd, G., Lee, S., Demirci, S., Eds.; The Elsevier and MICCAI Society Book Series; Academic Press: Cambridge, MA, USA, 2017; Chapter 4; pp. 79–106. ISBN 9780128110188. [Google Scholar] [CrossRef]
  9. Susser, L.I.; Rayner, K.J. Through the layers: How macrophages drive atherosclerosis across the vessel wall. J. Clin. Investig. 2022, 132, e157011. [Google Scholar] [CrossRef]
  10. Abedin, M.; Tintut, Y.; Demer, L.L. Mesenchymal stem cells and the artery wall. Circ. Res. 2004, 95, 671–676. [Google Scholar] [CrossRef]
  11. Bezsonov, E.E.; Sobenin, I.A.; Orekhov, A.N. Immunopathology of Atherosclerosis and Related Diseases: Focus on Molecular Biology. Int. J. Mol. Sci. 2021, 22, 4080. [Google Scholar] [CrossRef]
  12. Salnikova, D.; Orekhova, V.; Grechko, A.; Starodubova, A.; Bezsonov, E.; Popkova, T.; Orekhov, A. Mitochondrial Dysfunction in Vascular Wall Cells and Its Role in Atherosclerosis. Int. J. Mol. Sci. 2021, 22, 8990. [Google Scholar] [CrossRef] [PubMed]
  13. Guerrini, V.; Gennaro, M.L. Foam Cells: One Size Doesn’t Fit All. Trends Immunol. 2019, 40, 1163–1179. [Google Scholar] [CrossRef] [PubMed]
  14. Gui, Y.; Zheng, H.; Cao, R.Y. Foam Cells in Atherosclerosis: Novel Insights Into Its Origins, Consequences, and Molecular Mechanisms. Front. Cardiovasc. Med. 2022, 9, 845942. [Google Scholar] [CrossRef] [PubMed]
  15. Agarwal, P.; Gordon, S.; Martinez, F.O. Foam Cell Macrophages in Tuberculosis. Front. Immunol. 2021, 12, 775326. [Google Scholar] [CrossRef] [PubMed]
  16. Zheng, B.; Yu, L.; Hu, J.; Xu, H.; Wang, J.; Shi, Y.; Luo, X.; Yan, W. Expression of PD-L1 in mononuclear cells, multinucleated cells, and foam cells in tenosynovial giant cell tumors. Int. J. Clin. Exp. Pathol. 2019, 12, 876–884. [Google Scholar] [PubMed]
  17. Murray, P.J. Cancer metastasis linked to macrophage size, shape, and metabolism. J. Exp. Med. 2020, 217, e20201259. [Google Scholar] [CrossRef]
  18. Kloc, M.; Li, X.C.; Ghobrial, R.M. Are Macrophages Responsible for Cancer Metastasis? J. Immunol. Biol. 2016, 1, 1. [Google Scholar] [CrossRef]
  19. Lee-Rueckert, M.; Lappalainen, J.; Kovanen, P.T.; Escola-Gil, J.C. Lipid-Laden Macrophages and Inflammation in Atherosclerosis and Cancer: An Integrative View. Front. Cardiovasc. Med. 2022, 9, 777822. [Google Scholar] [CrossRef]
  20. Jones, N.L.; Allen, N.S.; Willingham, M.C.; Lewis, J.C. Modified LDLs induce and bind to membrane ruffles on macrophages. Anat. Rec. 1999, 255, 44–56. [Google Scholar] [CrossRef]
  21. Kattoor, A.J.; Kanuri, S.H.; Mehta, J.L. Role of Ox-LDL and LOX-1 in Atherogenesis. Curr. Med. Chem. 2019, 26, 1693–1700. [Google Scholar] [CrossRef]
  22. Zhang, X.; Sessa, W.C.; Fernández-Hernando, C. Endothelial Transcytosis of Lipoproteins in Atherosclerosis. Front. Cardiovasc. Med. 2018, 5, 130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Zhang, X.; Fernández-Hernando, C. Transport of LDLs into the arterial wall: Impact in atherosclerosis. Curr. Opin. Lipidol. 2020, 31, 279–285. [Google Scholar] [CrossRef] [PubMed]
  24. Summerhill, V.I.; Grechko, A.V.; Yet, S.F.; Sobenin, I.A.; Orekhov, A.N. The Atherogenic Role of Circulating Modified Lipids in Atherosclerosis. Int. J. Mol. Sci. 2019, 20, 3561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Shen, W.J.; Azhar, S.; Kraemer, F.B. SR-B1: A Unique Multifunctional Receptor for Cholesterol Influx and Efflux. Annu. Rev. Physiol. 2018, 80, 95–116. [Google Scholar] [CrossRef] [PubMed]
  26. Luchetti, F.; Crinelli, R.; Nasoni, M.G.; Benedetti, S.; Palma, F.; Fraternale, A.; Iuliano, L. LDL receptors, caveolae and cholesterol in endothelial dysfunction: oxLDLs accomplices or victims? Br. J. Pharmacol. 2021, 178, 3104–3114. [Google Scholar] [CrossRef] [PubMed]
  27. Fuenzalida, B.; Kallol, S.; Lüthi, M.; Albrecht, C.; Leiva, A. The polarized localization of lipoprotein receptors and cholesterol transporters in the syncytiotrophoblast of the placenta is reproducible in a monolayer of primary human trophoblasts. Placenta 2021, 105, 50–60. [Google Scholar] [CrossRef]
  28. Sukhorukov, V.N.; Khotina, V.A.; Chegodaev, Y.S.; Ivanova, E.; Sobenin, I.A.; Orekhov, A.N. Lipid Metabolism in Macrophages: Focus on Atherosclerosis. Biomedicines 2020, 8, 262. [Google Scholar] [CrossRef] [PubMed]
  29. Suzuki, H.; Kurihara, Y.; Takeya, M.; Kamada, N.; Kataoka, M.; Jishage, K.; Ueda, O.; Sakaguchi, H.; Higashi, T.; Suzuki, T.; et al. A role for macrophage scavenger receptors in atherosclerosis and susceptibility to infection. Nature 1997, 386, 292–296. [Google Scholar] [CrossRef]
  30. Febbraio, M.; Podrez, E.A.; Smith, J.D.; Hajjar, D.P.; Hazen, S.L.; Hoff, H.F.; Sharma, K.; Silverstein, R.L. Targeted disruption of the class B scavenger receptor CD36 protects against atherosclerotic lesion development in mice. J. Clin. Investig. 2000, 105, 1049–1056. [Google Scholar] [CrossRef]
  31. Kunjathoor, V.V.; Febbraio, M.; Podrez, E.A.; Moore, K.J.; Andersson, L.; Koehn, S.; Rhee, J.S.; Silverstein, R.; Hoff, H.F.; Freeman, M.W. Scavenger receptors class A-I/II and CD36 are the principal receptors responsible for the uptake of modified low density lipoprotein leading to lipid loading in macrophages. J. Biol. Chem. 2002, 277, 49982–49988. [Google Scholar] [CrossRef] [Green Version]
  32. Kuchibhotla, S.; Vanegas, D.; Kennedy, D.J.; Guy, E.; Nimako, G.; Morton, R.E.; Febbraio, M. Absence of CD36 protects against atherosclerosis in ApoE knock-out mice with no additional protection provided by absence of scavenger receptor A I/II. Cardiovasc. Res. 2008, 78, 185–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zgurskaya, H.I. Introduction: Transporters, Porins, and Efflux Pumps. Chem. Rev. 2021, 121, 5095–5097. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, F.; Gu, H.M.; Zhang, D.W. Caveolin-1 and ATP binding cassette transporter A1 and G1-mediated cholesterol efflux. Cardiovasc. Hematol. Disord. Drug Targets 2014, 14, 142–148. [Google Scholar] [CrossRef] [PubMed]
  35. Lippincott-Schwartz, J. Coated-pit dynamics. Nature 1999, 398, 753. [Google Scholar] [CrossRef]
  36. Doherty, G.J.; McMahon, H.T. Mechanisms of endocytosis. Annu. Rev. Biochem. 2009, 78, 857–902. [Google Scholar] [CrossRef] [Green Version]
  37. Meng, Y.; Heybrock, S.; Neculai, D.; Saftig, P. Cholesterol Handling in Lysosomes and Beyond. Trends Cell Biol. 2020, 30, 452–466. [Google Scholar] [CrossRef]
  38. Zhang, H. Lysosomal acid lipase and lipid metabolism: New mechanisms, new questions, and new therapies. Curr. Opin. Lipidol. 2018, 29, 218–223. [Google Scholar] [CrossRef]
  39. Kockx, M.; Dinnes, D.L.; Huang, K.Y.; Sharpe, L.J.; Jessup, W.; Brown, A.J.; Kritharides, L. Cholesterol accumulation inhibits ER to Golgi transport and protein secretion: Studies of apolipoprotein E and VSVGt. Biochem. J. 2012, 447, 51–60. [Google Scholar] [CrossRef] [Green Version]
  40. Widenmaier, S.B.; Snyder, N.A.; Nguyen, T.B.; Arduini, A.; Lee, G.Y.; Arruda, A.P.; Saksi, J.; Bartelt, A.; Hotamisligil, G.S. NRF1 Is an ER Membrane Sensor that Is Central to Cholesterol Homeostasis. Cell 2017, 171, 1094–1109.e15. [Google Scholar] [CrossRef]
  41. Chistiakov, D.A.; Bobryshev, Y.V.; Orekhov, A.N. Macrophage-mediated cholesterol handling in atherosclerosis. J. Cell. Mol. Med. 2016, 20, 17–28. [Google Scholar] [CrossRef] [Green Version]
  42. Galkina, E.; Ley, K. Immune and inflammatory mechanisms of atherosclerosis. Annu. Rev. Immunol. 2009, 27, 165–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Westhorpe, C.L.; Dufour, E.M.; Maisa, A.; Jaworowski, A.; Crowe, S.M.; Muller, W.A. Endothelial cell activation promotes foam cell formation by monocytes following transendothelial migration in an in vitro model. Exp. Mol. Pathol. 2012, 93, 220–226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Uitz, E.; Bahadori, B.; McCarty, M.F.; Moghadasian, M.H. Practical strategies for modulating foam cell formation and behavior. World J. Clin. Cases 2014, 2, 497–506. [Google Scholar] [CrossRef] [PubMed]
  45. Lin, J.; Kakkar, V.; Lu, X. Impact of matrix metalloproteinases on atherosclerosis. Curr. Drug Targets 2014, 15, 442–453. [Google Scholar] [CrossRef] [PubMed]
  46. Hegyi, L.; Skepper, J.N.; Cary, N.R.; Mitchinson, M.J. Foam cell apoptosis and the development of the lipid core of human atherosclerosis. J. Pathol. 1996, 180, 423–429. [Google Scholar] [CrossRef]
  47. Lei, S.; Chen, J.; Song, C.; Li, J.; Zuo, A.; Xu, D.; Li, T.; Guo, Y. CTRP9 alleviates foam cells apoptosis by enhancing cholesterol efflux. Mol. Cell. Endocrinol. 2021, 522, 111138. [Google Scholar] [CrossRef]
  48. Chen, X.; Deng, Z.; Feng, J.; Chang, Q.; Lu, F.; Yuan, Y. Necroptosis in Macrophage Foam Cells Promotes Fat Graft Fibrosis in Mice. Front. Cell Dev. Biol. 2021, 9, 651360. [Google Scholar] [CrossRef]
  49. Robichaud, S.; Rasheed, A.; Pietrangelo, A.; Doyoung Kim, A.; Boucher, D.M.; Emerton, C.; Vijithakumar, V.; Gharibeh, L.; Fairman, G.; Mak, E.; et al. Autophagy Is Differentially Regulated in Leukocyte and Nonleukocyte Foam Cells During Atherosclerosis. Circ. Res. 2022, 130, 831–847. [Google Scholar] [CrossRef]
  50. Liu, J.; Wang, C.; Li, J.; Yu, Y.; Liu, Y.; Liu, H.; Peng, Q.; Guan, X. Autophagy blockage promotes the pyroptosis of ox-LDL-treated macrophages by modulating the p62/Nrf2/ARE axis. J. Physiol. Biochem. 2021, 77, 419–429. [Google Scholar] [CrossRef]
  51. Stöger, J.L.; Gijbels, M.J.; van der Velden, S.; Manca, M.; van der Loos, C.M.; Biessen, E.A.; Daemen, M.J.; Lutgens, E.; de Winther, M.P. Distribution of macrophage polarization markers in human atherosclerosis. Atherosclerosis 2012, 225, 461–468. [Google Scholar] [CrossRef] [Green Version]
  52. Mushenkova, N.V.; Bezsonov, E.E.; Orekhova, V.A.; Popkova, T.V.; Starodubova, A.V.; Orekhov, A.N. Recognition of Oxidized Lipids by Macrophages and Its Role in Atherosclerosis Development. Biomedicines 2021, 9, 915. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, L.; Liu, H.J.; Li, T.J.; Yang, Y.; Guo, X.L.; Wu, M.C.; Rui, Y.C.; Wei, L.X. Lentiviral vector-mediated siRNA knockdown of SR-PSOX inhibits foam cell formation in vitro. Acta Pharmacol. Sin. 2008, 29, 847–852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Olzmann, J.A.; Carvalho, P. Dynamics and functions of lipid droplets. Nat. Rev. Mol. Cell Biol. 2019, 20, 137–155. [Google Scholar] [CrossRef] [PubMed]
  55. McLaren, J.E.; Michael, D.R.; Ashlin, T.G.; Ramji, D.P. Cytokines, macrophage lipid metabolism and foam cells: Implications for cardiovascular disease therapy. Prog. Lipid Res. 2011, 50, 331–347. [Google Scholar] [CrossRef]
  56. Walther, T.C.; Chung, J.; Farese, R.V., Jr. Lipid Droplet Biogenesis. Annu. Rev. Cell Dev. Biol. 2017, 33, 491–510. [Google Scholar] [CrossRef] [Green Version]
  57. Yu, X.H.; Zhang, D.W.; Zheng, X.L.; Tang, C.K. Cholesterol transport system: An integrated cholesterol transport model involved in atherosclerosis. Prog. Lipid Res. 2019, 73, 65–91. [Google Scholar] [CrossRef]
  58. Ben-Aicha, S.; Badimon, L.; Vilahur, G. Advances in HDL: Much More than Lipid Transporters. Int. J. Mol. Sci. 2020, 21, 732. [Google Scholar] [CrossRef] [Green Version]
  59. Matsuo, M. ABCA1 and ABCG1 as potential therapeutic targets for the prevention of atherosclerosis. J. Pharmacol. Sci. 2022, 148, 197–203. [Google Scholar] [CrossRef]
  60. Subramaniam, K.; Babu, L.A.; Shah, N. A case of premature and recurrent myocardial infarction associated with ABCA.1 gene mutation. J. Postgrad. Med. 2021, 67, 29–32. [Google Scholar] [CrossRef]
  61. Wu, Y.T.; Li, J.B.; Lin, H.Q.; Zhang, G.X.; Hong, C.M.; Li, M.; Guo, Z.J.; Yang, Y.B. Inhibition of miR-200b-3p alleviates lipid accumulation and promotes cholesterol efflux by targeting ABCA1 in macrophage-derived foam cells. Exp. Ther. Med. 2021, 22, 831. [Google Scholar] [CrossRef]
  62. Chinetti-Gbaguidi, G.; Baron, M.; Bouhlel, M.A.; Vanhoutte, J.; Copin, C.; Sebti, Y.; Derudas, B.; Mayi, T.; Bories, G.; Tailleux, A.; et al. Human atherosclerotic plaque alternative macrophages display low cholesterol handling but high phagocytosis because of distinct activities of the PPARγ and LXRα pathways. Circ. Res. 2011, 108, 985–995. [Google Scholar] [CrossRef] [PubMed]
  63. Kidani, Y.; Bensinger, S.J. LXR and PPAR as integrators of lipid homeostasis and immunity. as integrators of lipid homeostasis and immunity. Immunol. Rev. 2012, 249, 72–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Steinman, R.M.; Cohn, Z.A. Identification of a novel cell type in peripheral lymphoid organs of mice. I. Morphology, quantitation, tissue distribution. J. Exp. Med. 1973, 137, 1142–1162. [Google Scholar] [CrossRef] [Green Version]
  65. Schraml, B.U.; Reis e Sousa, C. Defining dendritic cells. Curr. Opin. Immunol. 2015, 32, 13–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zernecke, A. Dendritic cells in atherosclerosis: Evidence in mice and humans. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 763–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Bobryshev, Y.V. Dendritic cells and their role in atherogenesis. Lab Investig. 2010, 90, 970–984. [Google Scholar] [CrossRef] [Green Version]
  68. Liu, P.; Yu, Y.R.; Spencer, J.A.; Johnson, A.E.; Vallanat, C.T.; Fong, A.M.; Patterson, C.; Patel, D.D. CX3CR1 deficiency impairs dendritic cell accumulation in arterial intima and reduces atherosclerotic burden. Arterioscler. Thromb. Vasc. Biol. 2008, 28, 243–250. [Google Scholar] [CrossRef] [Green Version]
  69. Paulson, K.E.; Zhu, S.N.; Chen, M.; Nurmohamed, S.; Jongstra-Bilen, J.; Cybulsky, M.I. Resident intimal dendritic cells accumulate lipid and contribute to the initiation of atherosclerosis. Circ. Res. 2010, 106, 383–390. [Google Scholar] [CrossRef] [Green Version]
  70. Haka, A.S.; Grosheva, I.; Chiang, E.; Buxbaum, A.R.; Baird, B.A.; Pierini, L.M.; Maxfield, F.R. Macrophages create an acidic extracellular hydrolytic compartment to digest aggregated lipoproteins. Mol. Biol. Cell 2009, 20, 4932–4940. [Google Scholar] [CrossRef]
  71. Haka, A.S.; Singh, R.K.; Grosheva, I.; Hoffner, H.; Capetillo-Zarate, E.; Chin, H.F.; Anandasabapathy, N.; Maxfield, F.R. Monocyte-Derived Dendritic Cells Upregulate Extracellular Catabolism of Aggregated Low-Density Lipoprotein on Maturation, Leading to Foam Cell Formation. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 2092–2103. [Google Scholar] [CrossRef] [Green Version]
  72. Salvatore, G.; Bernoud-Hubac, N.; Bissay, N.; Debard, C.; Daira, P.; Meugnier, E.; Proamer, F.; Hanau, D.; Vidal, H.; Aricò, M.; et al. Human monocyte-derived dendritic cells turn into foamy dendritic cells with IL-17A. J. Lipid Res. 2015, 56, 1110–1122. [Google Scholar] [CrossRef] [Green Version]
  73. Dubland, J.A.; Francis, G.A. So Much Cholesterol: The unrecognized importance of smooth muscle cells in atherosclerotic foam cell formation. Curr. Opin. Lipidol. 2016, 27, 155–161. [Google Scholar] [CrossRef] [PubMed]
  74. Shankman, L.S.; Gomez, D.; Cherepanova, O.A.; Salmon, M.; Alencar, G.F.; Haskins, R.M.; Swiatlowska, P.; Newman, A.A.; Greene, E.S.; Straub, A.C.; et al. KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis. Nat. Med. 2015, 21, 628–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. 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 2003, 100, 13531–13536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Vengrenyuk, Y.; Nishi, H.; Long, X.; Ouimet, M.; Savji, N.; Martinez, F.O.; Cassella, C.P.; Moore, K.J.; Ramsey, S.A.; Miano, J.M.; 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. 2015, 35, 535–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Huff, M.W.; Pickering, J.G. Can a vascular smooth muscle-derived foam-cell really change its spots? Arterioscler. Thromb. Vasc. Biol. 2015, 35, 492–495. [Google Scholar] [CrossRef] [Green Version]
  78. Costales, P.; Fuentes-Prior, P.; Castellano, J.; Revuelta-Lopez, E.; Corral-Rodríguez, M.Á.; Nasarre, L.; Badimon, L.; Llorente-Cortes, V. K Domain CR9 of Low Density Lipoprotein (LDL) Receptor-related Protein 1 (LRP1) Is Critical for Aggregated LDL-induced Foam Cell Formation from Human Vascular Smooth Muscle Cells. J. Biol. Chem. 2015, 290, 14852–14865. [Google Scholar] [CrossRef] [Green Version]
  79. Wang, N.; Silver, D.L.; Thiele, C.; Tall, A.R. ATP-binding cassette transporter A1 (ABCA1) functions as a cholesterol efflux regulatory protein. J. Biol. Chem. 2001, 276, 23742–23747. [Google Scholar] [CrossRef] [Green Version]
  80. Ivan, L.; Antohe, F. Hyperlipidemia induces endothelial-derived foam cells in culture. J. Recept. Signal. Transduct. Res. 2010, 30, 106–114. [Google Scholar] [CrossRef]
  81. Swanson, K.V.; Deng, M.; Ting, J.P.Y. The NLRP3 inflammasome: Molecular activation and regulation to therapeutics. Nat. Rev. Immunol. 2019, 19, 477–489. [Google Scholar] [CrossRef]
  82. Uccelli, A.; Moretta, L.; Pistoia, V. Mesenchymal stem cells in health and disease. Nat. Rev. Immunol. 2008, 8, 726–736. [Google Scholar] [CrossRef] [PubMed]
  83. Ding, D.C.; Shyu, W.C.; Lin, S.Z. Mesenchymal stem cells. Cell Transplant. 2011, 20, 5–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kloc, M.; Uosef, A.; Leśniak, M.; Kubiak, J.Z.; Ghobrial, R.M. Reciprocal interactions between mesenchymal stem cells and macrophages. Int. J. Dev. Biol. 2020, 64, 465–469. [Google Scholar] [CrossRef] [PubMed]
  85. Jiang, L.; Chen, T.; Sun, S.; Wang, R.; Deng, J.; Lyu, L.; Wu, H.; Yang, M.; Pu, X.; Du, L.; et al. Nonbone Marrow CD34+ Cells Are Crucial for Endothelial Repair of Injured Artery. Circ. Res. 2021, 129, e146–e165. [Google Scholar] [CrossRef] [PubMed]
  86. Daub, K.; Langer, H.; Seizer, P.; Stellos, K.; May, A.E.; Goyal, P.; Bigalke, B.; Schönberger, T.; Geisler, T.; Siegel-Axel, D.; et al. Platelets induce differentiation of human CD34+ progenitor cells into foam cells and endothelial cells. FASEB J. 2006, 20, 2559–2561. [Google Scholar] [CrossRef] [Green Version]
  87. Stellos, K.; Seizer, P.; Bigalke, B.; Daub, K.; Geisler, T.; Gawaz, M. Platelet aggregates-induced human CD34+ progenitor cell proliferation and differentiation to macrophages and foam cells is mediated by stromal cell derived factor 1 in vitro. Semin. Thromb. Hemost. 2010, 36, 139–145. [Google Scholar] [CrossRef]
  88. Chen, Y.; Wong, M.M.; Campagnolo, P.; Simpson, R.; Winkler, B.; Margariti, A.; Hu, Y.; Xu, Q. Adventitial stem cells in vein grafts display multilineage potential that contributes to neointimal formation. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 1844–1851. [Google Scholar] [CrossRef]
Figure 1. Artery wall composition and Atherosclerosis progression. (1) Layers of the arterial wall. The tunica intima is the most internal (luminal) layer composed of the endothelial layer, connective tissue, and internal elastic layer. The tunica media contains connective tissue and smooth muscle cells. The tunica externa or adventitia is the most external layer composed of external elastic membrane, connective tissue, and fibroblasts. (2) Stages of atherosclerosis progression. (A) Lipid accumulation/Early stage: Hyperlipidemia leads to the influx of lipoproteins into the intima of arterial wall. Oxidation of LDL to oxLDL results in endothelial cell damage and loss of inter-endothelial cell adhesion molecules. (B) The fatty streak stage: Endothelial transcytosis and leaky endothelial barrier allow infiltration of LDL into the arterial wall. LDL accumulated and retained in the intima of the arterial wall-manifest forms a fatty streak. (C) Early plaque stage: Some endothelial cells undergo redifferentiation into mesenchymal cells, which initiate fibrotic processes. (D) Advanced/ruptured plaque stages: injured endothelial cells produce inflammatory factors, which recruit immune and non-immune cells to the plaque. Some of these recruited cells become foam cells. Eventually, foam cell death promotes plaque instability, rupture, and the development of blood clots.
Figure 1. Artery wall composition and Atherosclerosis progression. (1) Layers of the arterial wall. The tunica intima is the most internal (luminal) layer composed of the endothelial layer, connective tissue, and internal elastic layer. The tunica media contains connective tissue and smooth muscle cells. The tunica externa or adventitia is the most external layer composed of external elastic membrane, connective tissue, and fibroblasts. (2) Stages of atherosclerosis progression. (A) Lipid accumulation/Early stage: Hyperlipidemia leads to the influx of lipoproteins into the intima of arterial wall. Oxidation of LDL to oxLDL results in endothelial cell damage and loss of inter-endothelial cell adhesion molecules. (B) The fatty streak stage: Endothelial transcytosis and leaky endothelial barrier allow infiltration of LDL into the arterial wall. LDL accumulated and retained in the intima of the arterial wall-manifest forms a fatty streak. (C) Early plaque stage: Some endothelial cells undergo redifferentiation into mesenchymal cells, which initiate fibrotic processes. (D) Advanced/ruptured plaque stages: injured endothelial cells produce inflammatory factors, which recruit immune and non-immune cells to the plaque. Some of these recruited cells become foam cells. Eventually, foam cell death promotes plaque instability, rupture, and the development of blood clots.
Ijms 23 14154 g001
Figure 2. Foam cell formation and fate.
Figure 2. Foam cell formation and fate.
Ijms 23 14154 g002
Figure 3. Sources of foam cell progenitors and foam cells. Bone marrow is the primary source of macrophage/dendritic cell progenitors and stem cells. Macrophage/dendritic cell progenitors differentiate into common dendritic cell precursors, which give rise to plasmacytoid dendritic cells and pre-conventional dendritic cells and into the monocytes, which form macrophages and myeloid dendritic cells. Stem cells differentiate into mesenchymal stem cells, some of which enter various tissues and organs, including arterial wall. Bone-marrow-derived and wall-resident stem cells, macrophages, dendritic cells, smooth muscle cells, and endothelial cells can form foam cells.
Figure 3. Sources of foam cell progenitors and foam cells. Bone marrow is the primary source of macrophage/dendritic cell progenitors and stem cells. Macrophage/dendritic cell progenitors differentiate into common dendritic cell precursors, which give rise to plasmacytoid dendritic cells and pre-conventional dendritic cells and into the monocytes, which form macrophages and myeloid dendritic cells. Stem cells differentiate into mesenchymal stem cells, some of which enter various tissues and organs, including arterial wall. Bone-marrow-derived and wall-resident stem cells, macrophages, dendritic cells, smooth muscle cells, and endothelial cells can form foam cells.
Ijms 23 14154 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kloc, M.; Kubiak, J.Z.; Ghobrial, R.M. Macrophage-, Dendritic-, Smooth Muscle-, Endothelium-, and Stem Cells-Derived Foam Cells in Atherosclerosis. Int. J. Mol. Sci. 2022, 23, 14154. https://doi.org/10.3390/ijms232214154

AMA Style

Kloc M, Kubiak JZ, Ghobrial RM. Macrophage-, Dendritic-, Smooth Muscle-, Endothelium-, and Stem Cells-Derived Foam Cells in Atherosclerosis. International Journal of Molecular Sciences. 2022; 23(22):14154. https://doi.org/10.3390/ijms232214154

Chicago/Turabian Style

Kloc, Malgorzata, Jacek Z. Kubiak, and Rafik M. Ghobrial. 2022. "Macrophage-, Dendritic-, Smooth Muscle-, Endothelium-, and Stem Cells-Derived Foam Cells in Atherosclerosis" International Journal of Molecular Sciences 23, no. 22: 14154. https://doi.org/10.3390/ijms232214154

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop