ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Pathogenesis of atherosclerosis

Pathogenesis of atherosclerosis
Literature review current through: Jan 2024.
This topic last updated: Dec 05, 2023.

INTRODUCTION — Atherosclerosis is a pathologic process that causes disease of the coronary, cerebral, and peripheral arteries and the aorta [1,2]. Forms of accelerated arteriopathies, such as restenosis following percutaneous coronary intervention with stenting and coronary transplant vasculopathy differ in pathogenesis and are discussed separately. (See "Intracoronary stent restenosis" and "Heart transplantation in adults: Cardiac allograft vasculopathy pathogenesis and risk factors".)

EPIDEMIOLOGY — Atherosclerosis can begin in childhood with the development of fatty streaks (table 1). The lesions of atherosclerosis advance with aging [3,4]. The following points demonstrate the frequency of atherosclerosis in Western populations and its progression with age:

In an autopsy study of 2876 men and women aged 15 to 34 years who died of non-cardiac causes, all individuals had aortic fatty streaks [5].

In another autopsy study of 760 young (age 15 to 34 years) victims of accidents, suicides, or homicides, advanced coronary atheromata were seen in 2 and 0 percent of men and women aged 15 to 19, and 20 and 8 percent of men and women aged 30 to 34 [4].

In a study of 260 peri-renal aortic patches collected during organ transplantation, the first three decades of life were characterized by intimal thickening and xanthomas. The fourth, fifth, and sixth decades were characterized by more complicated plaques of pathological intimal thickening, early and late fibroatheromas with thin fibrous caps, ruptured plaques, healed rupture, and fibrotic calcified plaques [6].

Using intracoronary ultrasound, one in six United States teenagers have abnormal intimal thickening [3].

HISTOLOGY

Fatty streaks — The first phase in atherosclerosis histologically occurs as focal thickening of the intima with accumulation of lipid-laden macrophages (foam cells) and extracellular matrix (table 1 and figure 1) [7]. Smooth muscle cells can also populate the intima, some of which may arise from hematopoietic stem cells [8], migrate, and proliferate. Lipids accumulate early in fatty streak formation yielding both intracellular lipid and extracellular deposits, which produce the fatty streak. Biglycan, a small dermatan sulfate proteoglycan detected in the intima of atherosclerotic coronary artery segments, can bind and trap lipoproteins, including very low density lipoproteins and low density lipoprotein [9]. The fatty streak can also contain T lymphocytes. (See 'Inflammation' below.) Foam cells constitute the hallmark of the early atheroma.

As these lesions expand, more smooth muscle cells accumulate in the intima. The smooth muscle cells within the deep layer of the fatty streak can undergo apoptosis, which associates with further macrophage accumulation and microvesicles that can calcify, perhaps contributing to the transition of fatty streaks into atherosclerotic plaques (table 1) [10].

Fibrous cap — Fibrous cap atheromas are defined as plaques with a well-defined lipid core covered by a fibrous cap, which may be relatively acellular (made of dense collagen) or may be rich in smooth cells. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".)

Vasa vasorum — The vasa vasorum form a network of micro-vessels that originates primarily from the adventitial layer of large arteries. These vessels supply oxygen and nutrients to the outer layers of the arterial wall [11]. As atherosclerotic plaques develop and expand, they acquire their own microvascular network, extending from the adventitia through the media and into the thickened intima [12]. These thin-walled vessels are prone to disruption, leading to hemorrhage within the substance of the plaque and contributing to the progression of coronary atherosclerosis [13].

Fibrous plaque — The fibrous plaque evolves from the fatty streak via accumulation of connective tissue with an increased number of smooth muscle cells filled with lipids and often a deeper extracellular lipid pool.

Advanced lesions — More advanced lesions often contain a necrotic lipid-rich core, and eventually calcified regions (table 1) [14].

Atheroma formation associates with coronary artery remodeling (figure 2) [15]. Positive remodeling involves expansion of the plaque and external elastic membrane area due to a compensatory increase in local vessel size, while negative remodeling refers to a smaller external elastic membrane area at the lesion site due to the local shrinkage of vessel size (image 1A-B).

Positive remodeling acts as a compensatory mechanism in early coronary artery disease, preventing luminal loss despite plaque accumulation. However, arterial remodeling consequent to plaque formation and expansion is associated with abnormal arterial physiology as well as the development of clinical symptoms [15]. Positive remodeling is seen with complex, unstable plaques in patients presenting with unstable angina; in contrast, negative remodeling is associated with obstructive plaques in patients with stable angina [16]. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".)

Intraplaque hemorrhage — Intraplaque hemorrhage mainly results from plaque neovascularization and increased neovessel permeability [17,18], is a common feature of advanced atherosclerotic lesions, and a critical element leading to accelerated plaque progression [19-22], plaque instability [23-25], and ischemic vascular events [26-29].

Cholesterol crystals — Cholesterol crystals (CCs) are taken up by macrophages and other cell types, which subsequently alter cellular function and survival [30]. The effect and impact of external CCs on various cell types have also been studied in neutrophils and endothelial cells. The presence of CCs has been implicated in the activation of the complement system and in coagulation events.

PATHOGENESIS — Multiple factors contribute to the pathogenesis of atherosclerosis, including endothelial dysfunction, dyslipidemia, inflammatory, and immunologic factors, plaque rupture, and smoking. (See "Overview of established risk factors for cardiovascular disease" and "Cardiovascular risk of smoking and benefits of smoking cessation".)

Endothelial dysfunction — The endothelium forms an active biologic interface between the blood and all other tissues. The single layer of continuous endothelium lining arteries forms a unique thromboresistant layer between blood and potentially thrombogenic subendothelial tissues. The endothelium also modulates tone, growth, hemostasis, and inflammation throughout the circulatory system. Endothelial vasodilator dysfunction is an initial step in atherosclerosis and is felt to be caused principally by loss of endothelium-derived nitric oxide [31]. This issue is discussed in detail separately and will only briefly be reviewed here. (See "Coronary endothelial dysfunction: Clinical aspects".)

Endothelial dysfunction is associated with many of the traditional risk factors for atherosclerosis, including hypercholesterolemia, diabetes, hypertension, cigarette smoking. In particular, endothelial dysfunction is induced by oxidized low density lipoprotein (LDL) and in some respects can be considered as a final common pathway (table 2). It can be improved with correction of hyperlipidemia by diet or by therapy with a statin (HMG-coenzyme A reductase inhibitor), which increases the bioavailability of nitric oxide [32,33], with angiotensin converting enzyme inhibitors [34], or with high doses of antioxidants such as vitamin C or flavonoids contained in red wine and purple grape juice [35,36]. However, clinical benefits of these therapies have only been demonstrated convincingly for statins. (See "Coronary endothelial dysfunction: Clinical aspects" and "Mechanisms of benefit of lipid-lowering drugs in patients with coronary heart disease".)

Inflammation — Evidence of inflammation in atherosclerotic lesions has been noted from the earliest histologic observations and inflammation is central to understanding the pathogenesis of atherosclerosis [37-40]. Macrophages that have taken up oxidized LDL release a variety of inflammatory substances, cytokines, and growth factors [41,42]. Among the many molecules that have been implicated are: monocyte chemotactic protein (MCP)-1 [43,44]; intercellular adhesion molecule (ICAM)-1 [43]; macrophage and granulocyte-macrophage colony stimulating factors [45,46]; CD40 ligand ; interleukin (IL)-1, IL-3, IL-6, IL-8, and IL-18 [47-49]; and tumor necrosis factor alpha [50-52]. The role of IL-6 is discussed separately. (See "Overview of established risk factors for cardiovascular disease", section on 'Interleukin-6'.)

Evidence supporting the importance of inflammation in the pathogenesis of atherosclerosis comes from the observation that markers of increased or decreased systemic inflammation associated with the risk of atherosclerosis. Furthermore, the Cankanumab Anti-inflammatory Thrombosis Outcomes Study (CANTOS) has demonstrated that inhibition of interleukin-1 beta with canakinumab substantially lowered the inflammatory biomarkers hsCRP and IL-6 without changing atherogenic lipids in patients with a prior myocardial infarction. The risk of a composite of cardiovascular death, nonfatal myocardial infarction, and non-fatal stroke fell by 15 percent (p = 0.021) with canakinumab 150 mg subcutaneous injection every three months [53]. Some of these markers are discussed in the next sections.

Serum high sensitivity CRP (hsCRP) — hsCRP is one of the downstream inflammatory markers. It consistently associates with the increased risk of atherosclerotic cardiovascular disease independent of cholesterol level [54,55], although genetic data do not support its function as a causal risk factor. It has been considered as a useful marker to identify individuals with increased vascular inflammation [53,56]. The role of C-reactive protein in cardiovascular disease is discussed in detail separately. (See "C-reactive protein in cardiovascular disease".)

Lp-PLA2 — Lipoprotein-associated phospholipase A2 (Lp-PLA2) is a macrophage-secreted enzyme that may perpetuate plaque inflammation and whose elevated levels predict a 40 to 400 percent (averaging about 100 percent) increased risk of myocardial infarction (MI) and stroke in population studies fully adjusted for other cardiovascular disease risk factors. Clinical trials with an inhibitor of Lp-PLA2 [57] did not show improved outcomes. (See "Prevention of cardiovascular disease events in those with established disease (secondary prevention) or at very high risk", section on 'Therapies with uncertain or no benefit'.)

Cytokines — Cytokines may participate in the pathogenesis of atherosclerosis [58]. Mediators such as interleukin-1 or tumor necrosis factor-alpha have a multitude of atherogenic effects. Basic science studies [59-61] have identified that the proinflammatory cytokine interleukin-1ß plays multiple roles in the development of atherosclerotic plaque and the favorable effects of inhibiting interleukin-1ß signaling in animals with experimental atherosclerosis [62,63]. The cytokines enhance the expression of cell surface molecules such as ICAM-1, VCAM-1, CD40, and selectins on endothelial cells, smooth muscle cells, and macrophages. Pro-inflammatory cytokines can also induce cell proliferation, contribute to the production of reactive oxygen species, stimulate matrix metalloproteinases, and induce tissue factor expression. Other cytokines, such as interleukin-4 and interleukin-10, are antiatherogenic. Still others, such as interferon-gamma, can promote experimental atherogenesis. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".)

Leukocyte activation — Leukocytes infiltrate and accumulate in the atherosclerotic lesion, providing evidence for the role of local inflammation [64]. Inflammatory cells, including macrophages and T-lymphocytes, have often been found at the immediate site of intimal rupture or erosion of a thrombosed coronary artery in patients who died of acute myocardial infarction [65]. A systemic evaluation of immune/inflammatory cells in 114 aortic atherosclerotic specimens has described the inflammatory footprint in plaque instability:

Scattered CD4 and CD8 cells with a T memory subtype were found in normal aorta, early, nonprogressive lesions.

The total number of T cells increased in progressive lesions.

A further increase in medial and adventitial T cells was found upon progression to vulnerable lesions.

Formation of adventitial lymphoid-like structures containing B cells and plasma cells, a process accompanied by transient expression of CXCL13, was identified at this critical stage of progression to vulnerable lesions.

A dramatic decrease of T-cell subsets, disappearance of lymphoid-like structures, and loss of CXCL13 expression were characterized among post-ruptured lesions [66].

Evidence in support of systemic inflammation comes from a study in which the inflammatory mRNA profile of circulating leukocytes was tested in 524 men with a prior MI and 628 controls [67]. The patients, compared to controls, had mRNA profiles showing increased levels of many inflammatory mRNAs.

Toll-like receptor 4 — Further evidence for the role of inflammation comes from study of polymorphisms in the toll-like receptor 4 gene that confer differences in the inflammatory response to Gram negative pathogens and perhaps other ligands [68]. A particular polymorphism of this gene, Asp299Gly, presents in 7 percent of an Italian population, diminishes receptor cycling, and is associated with diminished inflammatory response to Gram-negative pathogens.

Carriers of the Asp299Gly polymorphism, compared to patients with only wild-type alleles, have reduced circulating levels of a variety of inflammatory markers, including CRP, adhesion molecules, and IL-6, and a reduced incidence of carotid atherosclerosis (odds ratio 0.54) as detected by ultrasonography. They have an increased rate of serious bacterial infections.

Pregnancy-associated plasma protein-A (PAPP-A) — PAPP-A is a high molecular weight, zinc-binding metalloproteinase that is thought to degrade the proteins that maintain the integrity of the protective fibrous cap of atherosclerotic plaques [69]. PAPP-A can be produced by osteoblasts, fibroblasts, endothelial, vascular smooth muscle cells, and monocytes/macrophages [70-74], and has been identified in vulnerable coronary plaques but not in stable ones [75-77].

Dyslipidemia — Lipid abnormalities play a critical role in the development of atherosclerosis [42,78-83]. Early experiments in animals demonstrated accelerated atherosclerosis with a high cholesterol diet. Subsequent epidemiologic studies conducted in countries around the world showed an increasing incidence of atherosclerosis when serum cholesterol concentrations were above 150 mg/dL (3.9 mmol/L) (figure 3).

The role of the different lipid particles in the development of atherosclerosis is discussed elsewhere. (See "Lipoprotein classification, metabolism, and role in atherosclerosis".)

It is useful, however, to summarize the major observations.

High levels of LDL cholesterol [83,84] are particularly important risk factors for atherosclerosis (algorithm 1) [78].

Cholesterol accumulates in the lipid-laden macrophages (foam cells), and in the lipid core, of atherosclerotic plaque [85]. Oxidative modification of LDL facilitates macrophage uptake via unregulated macrophage scavenger receptors (among them, CD36, also called scavenger receptor B) and for accelerated accumulation of cholesterol [86,87]. Macrophage uptake of LDL cholesterol may initially be an adaptive response, which prevents LDL-induced endothelial injury [88]. However, cholesterol accumulation in foam cells leads to mitochondrial dysfunction, apoptosis, and necrosis, with resultant release of cellular proteases, inflammatory cytokines, and prothrombotic molecules [88]. A large body of the accumulative evidence in the past 50 years has demonstrated that lowering LDL-C can reduce cardiovascular events and the lower levels of LDL-C achieved are associated with a better clinical outcome [89,90]. (See "Low-density lipoprotein cholesterol-lowering therapy in the primary prevention of cardiovascular disease" and "Management of low density lipoprotein cholesterol (LDL-C) in the secondary prevention of cardiovascular disease".)

HDL, in contrast to LDL, has putative antiatherogenic properties that include reverse cholesterol transport, maintenance of endothelial function, and protection against thrombosis. There is an inverse relationship between plasma HDL-cholesterol levels and cardiovascular risk (table 3). Values above 75 mg/dL (1.9 mmol/L) are associated with a longevity syndrome. Values above 60 mg/dL (1.5 mmol/L) count as a negative risk factor in the Framingham Risk Assessment [91]. However, cardiovascular disease event reduction from increasing HDL-cholesterol has not been established, particularly in patients with well-controlled LDL-cholesterol levels [92-94]. A Mendelian randomization study showed that raised plasma HDL-cholesterol levels through some genetic mechanisms are not associated with lower risk of MI [95]. Studies showed that HDL-cholesterol efflux had a close relationship with clinical atherosclerosis [96] and CV events [97]. These data challenge the concept that raising HDL-cholesterol levels will uniformly translate into a reduction in the risk of CV events. One explanation for the lack of benefit from therapies that raise HDL-cholesterol levels tested so far is that they may not improve the reverse cholesterol transport role of HDL. (See "HDL cholesterol: Clinical aspects of abnormal values".) Genetic studies have cast doubt on the causal relationship between low HDL levels and increased risk of atherosclerotic events. Thus, HDL may serve as a biomarker of risk, but no evidence yet shows that HDL-cholesterol is a modifiable risk factor.

Current epidemiologic and genetic evidence support a causal role for triglyceride-rich lipoproteins, particularly those that contain apolipoprotein C3.

An analysis in 60,608 individuals found that nonfasting remnant cholesterol, which is nonfasting total cholesterol minus HDL cholesterol and minus LDL cholesterol, associates causally with ischemic heart disease and with low-grade inflammation [98,99]. (See "Hypertriglyceridemia in adults: Management".)

Oxidized lipoproteins including LDL [42], HDL [100], remnant lipoproteins [101], and phospholipid [102] have been thought to cause disruption of the endothelial cell surface, and promote inflammatory via cytokine release from macrophages, reduced cholesterol efflux, and the development of atherosclerosis. It may also play a role in plaque instability. Levels of oxidized LDL are increased in patients with an acute coronary syndrome and are positively correlated with the severity of the syndrome [103,104]. In addition, antibodies to oxidized LDL have been found in human atherosclerotic plaques and in the plasma of patients with atherosclerosis [105,106]. Measurement of oxidized lipoproteins remains for research to better understand the pathogenesis and has not been standardized or validated as a clinically useful biomarker. Furthermore, antioxidant strategies have consistently failed to improve cardiovascular outcomes in clinical trials.

Observational and genetic studies support a causal role for lipoprotein(a), which is LDL covalently linked to apolipoprotein (a). (See "Lipoprotein(a)".)

The LDL moiety of Lp(a) promotes atherosclerosis, whereas the plasminogen-like apo(a) molecule may favor thrombus accumulation by interfering with fibrinolysis [107,108]. Additional functions of Lp(a) include initiation of signaling pathways in macrophages [109] and vascular endothelial cells [110,111], resulting in proatherogenic changes in cell phenotype and gene expression. Lp(a) binding to macrophages could lead to foam cell formation and localization of Lp(a) at atherosclerotic plaques [112]. Lp(a) associates with increased residual cardiovascular event risk in JUPITER [111] and AIM-HIGH [113], which suggest that Lp(a) remains a risk factor in subjects with aggressive LDL-cholesterol lowering. (See "Lipoprotein(a)".)

Hypertension — Hypertension is a major risk factor for the development of atherosclerosis, particularly in the coronary and cerebral circulations [114-116]. It can increase arterial wall tension, potentially leading to disturbed repair processes and aneurysm formation [2].

Smoking — Cigarette smoking is another major risk factor [115,116] and it impacts all phases of atherosclerosis from endothelial dysfunction to acute clinical events, the latter being largely thrombotic [117]. The following observations have been made:

In humans, cigarette smoke exposure impairs endothelium-dependent vasodilation, perhaps through decreased nitric oxide (NO) availability [118-120].

Cigarette smoking is associated with an increased level of multiple inflammatory markers, including C-reactive protein, interleukin-6, and tumor necrosis factor alpha in both male and female smokers [121-124].

Cigarette smoking may decrease availability of platelet-derived NO, decrease platelet sensitivity to exogenous NO (both of which may lead to increased activation and adhesion) [125,126], increase fibrinogen levels [127,128], and decrease fibrinolysis [129].

Cigarette smoking increases oxidative modification of LDL [130] and decreases the plasma activity of paraoxonase, an enzyme that protects against LDL oxidation [131]. The triglyceride/HDL abnormalities seen among smokers have been suggested to be related to insulin resistance [132].

Diabetes — In addition to atherogenic effects from the diabetes-related dyslipidemia (elevated triglycerides, low level of HDL cholesterol, and small/dense LDL particles), as mentioned above, many clinical and experimental studies reveal that high levels of insulin precede development of arterial diseases [133,134]. Macrophages express most insulin signaling molecules except IR substrate 1 (IRS1) and glucose transporter type 4 [135,136]. Even though insulin activates the IR/IRS2/PI3K/Akt pathway in macrophages as in other types of insulin-responsive cells, few studies have investigated the biological functions of insulin signaling in macrophages.

Atherosclerosis and type 2 diabetes share similar pathological mechanisms, including elevation in cytokines like MCP-1 and interleukin-6 (IL-6), which contribute to underlying inflammation of both [137,138]. A published study [137] evaluated if insulin affects macrophage foam cell formation and found that insulin increased the expression of CD36 and decreased ABCA-1 expression, which may promote cholesterol accumulation in human monocyte-derived macrophages. The study also showed that low concentration of adiponectin increased the phosphorylation of Akt (Ser436) by the same degree as insulin and had the same modulating effect on CD36 and ABCA-1 as insulin. Clinically, the overall risk increase conferred by type 2 diabetes is driven by accelerated progression of pre-existing atherosclerosis to clinical cardiovascular events [139].

Tissue factor — Tissue factor is the primary initiator of coagulation and, in advanced atherosclerosis, is found in the plaque. (See "Overview of hemostasis".) Tissue factor, as well as other factors such as enhanced platelet activity, may contribute to the development of thrombosis following plaque disruption. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".) Tissue factor also plays a role in the progression of atherosclerosis via coagulation-dependent and coagulation-independent mechanisms. In one study, tissue factor overexpression increased neointimal area and plaque size by increasing mural thrombus and smooth muscle cell migration and accelerating endothelial regrowth over the plaque after rupture [140].

Angiotensin II — The renin-angiotensin-aldosterone system plays an important role in inflammatory response regulation by recruiting inflammatory cells to the injured site and stimulating the production of cytokines such as IL-6, TNF-a, and COX-2 [141]. Angiotensin II stimulates the production of reactive oxygen species [142,143] and lowers nitric oxide production [144], leading to endothelial dysfunction. Increased plasma concentrations of angiotensin II promote the development and severity of atherosclerosis, particularly when combined with hyperlipidemia [145]. Angiotensin II may modulate vascular smooth muscle cell proliferation and the production of extracellular matrix [146,147].

Endothelin-1 — Endothelin-1 may contribute to the pathogenesis of atherosclerosis at all stages, even when the plaque is clinically imperceptible [148,149]. Endothelin-1 is a potent vasoconstrictor as well as a mitogen for vascular smooth muscle cells, stimulating their migration and growth. Oxidized LDL can stimulate its production and enhance its vasoconstrictor effects [150]. (See "Pathophysiology of heart failure: Neurohumoral adaptations", section on 'Endothelin'.)

Endothelin-1 immunoreactivity is ubiquitous within the intracellular and extracellular compartments of human coronary atherosclerotic tissue and is released from these sites in response to mechanical stress [151]. In addition, endothelin-converting enzyme-1, the final enzyme for endothelin-1 production, is expressed in smooth muscle cells and macrophages of human coronary atherosclerotic lesions at all stages of development [149,152].

Adhesion molecules — Intercellular adhesion molecule-1 (ICAM-1) and vascular cell adhesion molecule-1 (VCAM-1) are cell surface glycoproteins induced at endothelial sites of inflammation that mediate the adherence of leukocytes to endothelium. (See "Leukocyte-endothelial adhesion in the pathogenesis of inflammation".) A low level of ICAM-1 is expressed on normal endothelial cells and is seen in normal arterial segments, while VCAM-1 expression occurs only with inflammation and is present in the microvessels of human atherosclerotic lesions [153].

The expression of both ICAM-1 and VCAM-1 increases in atherosclerotic lesions, but VCAM-1 appears to be more important in the initiation of atherosclerosis [154]. P-selectin, a platelet and endothelial cell receptor that mediates adhesion between vascular cells, also may promote migration of inflammatory cells into early and advanced atherosclerotic lesions [155,156].

Monocyte adhesion to endothelial cells is reduced by L-arginine, the precursor of nitric oxide [157] and alpha tocopherol (vitamin E), and increased by androgens due in part to enhanced endothelial cell-surface expression of VCAM-1 [158]. Antibodies that block adhesion molecules may ameliorate elements of the inflammatory response in atherosclerotic plaques [159].

Flow characteristics — The frequent occurrence of atheroma at sites of bends, branches, and bifurcations of coronary arteries suggests that altered blood flow and low shear stress can play a role in the development of atherosclerosis. Disturbed flow can alter endothelial cell function [160] in a manner that impairs atheroprotective functions [41]. These changes may be mediated by inhibition of the release of nitric oxide from the endothelial cells [161,162].

Atherosclerosis preferentially affects regions of coronary arteries that experience low shear stress or disturbed flow [163]. In addition, low shear stress is associated with an increase in intima medial thickness of the common carotid artery in healthy men [164].

Mitochondrial DNA damage — The hypothesis that mitochondrial DNA (mtDNA) injury might lead to plaque development and vulnerability has been proposed and tested in experimental animal models [165,166]. In the hyperlipidemic ApoE knockout mouse model, mtDNA lesions found in the aortas preceded the development of aortic plaques [167]. More mtDNA lesions and larger plaque size were detected in the aortas of mice expressing mitochondrial mutation [167]. In patients, association of leukocyte mtDNA with atherosclerotic plaque vulnerability was examined in the Virtual Histology in Vulnerable Atherosclerosis (VIVA) trial [168]. In this study, mtDNA lesions were found to be uniquely associated with thin-cap fibroatheroma, which are associated with a high risk of cardiovascular events [168,169]. Taken together, these findings suggest a possible role of mtDNA injury in plaque progression and disruption.

Genetic associations — The genetic influences on atherosclerosis formation, progression, and atherosclerotic vascular events has attracted much attention. Two major genetic study approaches have been undertaken for a better understanding of the molecular mechanism of atherosclerosis. The first is the candidate gene approach, in which genes in known atherogenesis pathways are tested for their role in atherosclerosis in vitro, in vivo, and in association studies [170-172]. The second approach is conducting genome-wide linkage studies to find atherogenesis-regulating quantitative trait loci. This method has the potential to find new atherosclerosis genes. The availability of whole genome sequences in humans and mice, especially the abundant single nucleotide polymorphism and haplotype information, has made it possible to perform genome-wide association studies, another unbiased approach, to identify disease genes relatively quickly as compared with traditional genetic methods [173].

The complex pathophysiological processes that occur in atherosclerosis probably do not arise from a single or small number of genes. In addition, environmental factors, and their interactions with genes, likely participate. In the general population, genetic polymorphisms occur in many genes in the pathways of lipid metabolism, inflammation, and thrombogenesis.

Genetic and genomic studies have helped to identify newer therapeutic targets in atherosclerosis. For example, discovery of genetic variants of PCSK9 [174-180] in regulating LDL-C levels has led to a rapid development of PCSK9 inhibitor as a potential therapy for LDL-C lowering and reducing cardiovascular events. Furthermore, genetic manipulations in animal models will accelerate the pace of atherosclerosis research [181]. (See "PCSK9 inhibitors: Pharmacology, adverse effects, and use".)

Infection — Chronic infection may contribute to the pathogenesis of atherosclerosis. The major organisms that have been reported are Chlamydia pneumoniae [182], cytomegalovirus (CMV) [183-192], coxsackie B virus infection [193], and Helicobacter pylori [194-196]. (See "Overview of possible risk factors for cardiovascular disease", section on 'Infection'.)

In addition to individual infections, the total pathogen burden, ie, the number of pathogens to which an individual has been exposed, may be an important risk factor for atherosclerosis [197-200]. Pathogen burden with atherosclerotic lesions has been directly assessed by testing for bacterial rDNA signatures by polymerase chain reaction in specimens obtained during catheter-based atherectomy [201]. Bacterial DNA was found in all patients with a mean of 12 species per lesion; bacterial DNA was not seen in control material or any unaffected coronary arteries. It is possible that secondary colonization may accelerate disease progression.

Chronic infection could act by a number of mechanisms, including direct vascular injury and induction of a systemic inflammatory state. (See 'Inflammation' above.)

Despite the attractiveness of infection as a potential important contributing factor, clinical evidence to support its role is lacking, and antibiotic therapy does not reduce the risk of recurrent myocardial infarction.

Effect of vaccination — Vaccination against influenza has been evaluated for a potential benefit in preventing cardiovascular disease [202-204]. Several studies have found a beneficial effect of influenza vaccination on cardiovascular events [202,203]. While these studies are of interest and do document a beneficial effect of influenza vaccination in older adults, they do not establish a causative relationship between influenza infection and the pathogenesis of atherosclerosis. (See "Geriatric health maintenance", section on 'Influenza vaccine' and "Seasonal influenza vaccination in adults".)

Gut microbiota — An important contribution of gut microbiota in pathobiological processes in cardiovascular diseases has been suggested [205].

Studies support a potential link between dietary nutrients, such as choline and carnitine (found in red meat, egg yolks, and full fat dairy foods), and their gut microbiota-dependent metabolism to trimethylamine N-oxide (TMAO), and atherosclerosis [205,206]. Prospective cohort studies have shown that increased plasma TMAO levels predict an elevated risk of major adverse cardiovascular events such as MI, stroke, or death [207-212]. High TMAO levels were found to increase the expression of proinflammatory genes including inflammatory cytokines, adhesion molecules, and chemokines [211,213-216]. These data support the notion that TMAO promotes inflammation and adds to elevated cardiovascular risk.

ROLE OF PLAQUE RUPTURE AND EROSION — Atherosclerosis is generally asymptomatic until the plaque stenosis exceeds 70 or 80 percent of the luminal diameter, which can produce a reduction in flow, as with coronary blood flow to myocardium. These stenotic lesions can produce typical symptoms of angina pectoris. Progression of atherosclerotic plaques involves two distinct processes: a chronic one that leads to luminal narrowing slowly, and an acute one that causes rapid luminal obstruction associated with plaque hemorrhage and/or luminal thrombosis.

Acute coronary and cerebrovascular syndromes (unstable angina, myocardial infarction, sudden death, and stroke) are typically due to rupture or erosion of plaques leading to thrombosis, although these plaques may have less than 50 percent stenosis [217-219]. In a review [220] of 22 autopsy studies in which 1847 coronary arteries were examined, plaque rupture was the main cause of coronary thrombosis regardless of clinical presentation (myocardial infarction: 79 percent; sudden coronary death: 65 percent), age (>60 years: 77 percent; <60 years: 64 percent; unknown: 73 percent), sex (men: 76 percent; women: 55 percent), and continent (Europe: 72 percent; United States: 68 percent; Asia 81 percent).

Plaque erosion is identified when serial sectioning of the thrombosed arterial segment fails to reveal plaque rupture [221]. Typically, the endothelium is missing at the erosion site, and the exposed intima consists predominantly of vascular smooth muscle cells and proteoglycans. Clinical studies of optical coherence tomography performed in patients presenting with acute myocardial infarction showed that the incidence of fibrous cap disruption was 73 percent, whereas plaque erosion was 23 percent and patients with plaque rupture had a higher incidence of thin cap fibroatheroma of 83 percent [222]; results were very similar to autopsy studies.

Plaque rupture or erosion may also be silent; repeated silent ruptures and thrombosis, followed by wound healing, may cause progression of atherosclerosis, with an increase in plaque burden and percent stenosis and negative arterial remodeling [223]. In studies of patients with acute coronary syndromes who underwent imaging with intravascular ultrasound (and were treated with percutaneous coronary intervention of the culprit lesion), recurrent major adverse cardiovascular events were predicted by the finding of culprit and non-culprit lesions with thin-cap fibroatheromas [168,169].

A recent review examined the mechanisms of atherosclerotic plaque healing, their role in the progression of atherosclerotic disease and in the development of acute coronary syndromes, as well as the potential therapeutic implications of the healing process [224].

The pathology and pathogenesis of the vulnerable plaque and its role in acute coronary syndromes are discussed elsewhere. (See "Mechanisms of acute coronary syndromes related to atherosclerosis".)

INFORMATION FOR PATIENTS — UpToDate offers two types of patient education materials, “The Basics” and “Beyond the Basics.” The Basics patient education pieces are written in plain language, at the 5th to 6th grade reading level, and they answer the four or five key questions a patient might have about a given condition. These articles are best for patients who want a general overview and who prefer short, easy-to-read materials. Beyond the Basics patient education pieces are longer, more sophisticated, and more detailed. These articles are written at the 10th to 12th grade reading level and are best for patients who want in-depth information and are comfortable with some medical jargon.

Here are the patient education articles that are relevant to this topic. We encourage you to print or e-mail these topics to your patients. (You can also locate patient education articles on a variety of subjects by searching on “patient info” and the keyword(s) of interest.)

Basics topic (See "Patient education: Atherosclerosis (The Basics)".)

SUMMARY

Multifactorial pathogenesis – Multiple factors contribute to the pathogenesis of atherosclerosis and its complications, including endothelial dysfunction, inflammatory and immunologic factors, plaque rupture or erosion, and the traditional risk factors of hypertension, diabetes, dyslipidemia, and smoking. Despite the attractiveness of infection as a potential contributing factor, clinical evidence to support its role is lacking. (See 'Pathogenesis' above.)

Childhood origin – Atherosclerosis begins in childhood with the development of fatty streaks. The advanced lesions of atherosclerosis occur with increasing frequency with aging. (See 'Epidemiology' above.)

Histologic stages – The histologic stages of atherosclerosis include fatty streak, fibrous cap, fibrous plaques, and advances lesions. With the availability of advanced vascular imaging techniques, the plaque histological characteristics can be identified in vivo. (See 'Histology' above.)

ACKNOWLEDGMENT — The UpToDate editorial staff acknowledges Xue-Qiao Zhao, MD, FACC, who contributed to earlier versions of this topic review.

  1. Faxon DP, Fuster V, Libby P, et al. Atherosclerotic Vascular Disease Conference: Writing Group III: pathophysiology. Circulation 2004; 109:2617.
  2. Libby P, Ridker PM, Hansson GK. Progress and challenges in translating the biology of atherosclerosis. Nature 2011; 473:317.
  3. Tuzcu EM, Kapadia SR, Tutar E, et al. High prevalence of coronary atherosclerosis in asymptomatic teenagers and young adults: evidence from intravascular ultrasound. Circulation 2001; 103:2705.
  4. McGill HC Jr, McMahan CA, Zieske AW, et al. Association of Coronary Heart Disease Risk Factors with microscopic qualities of coronary atherosclerosis in youth. Circulation 2000; 102:374.
  5. Strong JP, Malcom GT, McMahan CA, et al. Prevalence and extent of atherosclerosis in adolescents and young adults: implications for prevention from the Pathobiological Determinants of Atherosclerosis in Youth Study. JAMA 1999; 281:727.
  6. van Dijk RA, Virmani R, von der Thüsen JH, et al. The natural history of aortic atherosclerosis: a systematic histopathological evaluation of the peri-renal region. Atherosclerosis 2010; 210:100.
  7. Davies MJ, Woolf N, Rowles PM, Pepper J. Morphology of the endothelium over atherosclerotic plaques in human coronary arteries. Br Heart J 1988; 60:459.
  8. Sata M, Saiura A, Kunisato A, et al. Hematopoietic stem cells differentiate into vascular cells that participate in the pathogenesis of atherosclerosis. Nat Med 2002; 8:403.
  9. O'Brien KD, Olin KL, Alpers CE, et al. Comparison of apolipoprotein and proteoglycan deposits in human coronary atherosclerotic plaques: colocalization of biglycan with apolipoproteins. Circulation 1998; 98:519.
  10. Kockx MM, De Meyer GR, Muhring J, et al. Apoptosis and related proteins in different stages of human atherosclerotic plaques. Circulation 1998; 97:2307.
  11. Ritman EL, Lerman A. The dynamic vasa vasorum. Cardiovasc Res 2007; 75:649.
  12. Mulligan-Kehoe MJ, Simons M. Vasa vasorum in normal and diseased arteries. Circulation 2014; 129:2557.
  13. Virmani R, Narula J, Farb A. When neoangiogenesis ricochets. Am Heart J 1998; 136:937.
  14. Stary HC, Chandler AB, Dinsmore RE, et al. A definition of advanced types of atherosclerotic lesions and a histological classification of atherosclerosis. A report from the Committee on Vascular Lesions of the Council on Arteriosclerosis, American Heart Association. Circulation 1995; 92:1355.
  15. Schoenhagen P, Ziada KM, Vince DG, et al. Arterial remodeling and coronary artery disease: the concept of "dilated" versus "obstructive" coronary atherosclerosis. J Am Coll Cardiol 2001; 38:297.
  16. Schoenhagen P, Ziada KM, Kapadia SR, et al. Extent and direction of arterial remodeling in stable versus unstable coronary syndromes : an intravascular ultrasound study. Circulation 2000; 101:598.
  17. Sluimer JC, Kolodgie FD, Bijnens AP, et al. Thin-walled microvessels in human coronary atherosclerotic plaques show incomplete endothelial junctions relevance of compromised structural integrity for intraplaque microvascular leakage. J Am Coll Cardiol 2009; 53:1517.
  18. Virmani R, Kolodgie FD, Burke AP, et al. Atherosclerotic plaque progression and vulnerability to rupture: angiogenesis as a source of intraplaque hemorrhage. Arterioscler Thromb Vasc Biol 2005; 25:2054.
  19. Takaya N, Yuan C, Chu B, et al. Presence of intraplaque hemorrhage stimulates progression of carotid atherosclerotic plaques: a high-resolution magnetic resonance imaging study. Circulation 2005; 111:2768.
  20. Sun J, Underhill HR, Hippe DS, et al. Sustained acceleration in carotid atherosclerotic plaque progression with intraplaque hemorrhage: a long-term time course study. JACC Cardiovasc Imaging 2012; 5:798.
  21. Simpson RJ, Akwei S, Hosseini AA, et al. MR imaging-detected carotid plaque hemorrhage is stable for 2 years and a marker for stenosis progression. AJNR Am J Neuroradiol 2015; 36:1171.
  22. Underhill HR, Yuan C, Yarnykh VL, et al. Predictors of surface disruption with MR imaging in asymptomatic carotid artery stenosis. AJNR Am J Neuroradiol 2010; 31:487.
  23. Altaf N, Goode SD, Beech A, et al. Plaque hemorrhage is a marker of thromboembolic activity in patients with symptomatic carotid disease. Radiology 2011; 258:538.
  24. van Dijk AC, Truijman MT, Hussain B, et al. Intraplaque Hemorrhage and the Plaque Surface in Carotid Atherosclerosis: The Plaque At RISK Study (PARISK). AJNR Am J Neuroradiol 2015; 36:2127.
  25. Takaya N, Yuan C, Chu B, et al. Association between carotid plaque characteristics and subsequent ischemic cerebrovascular events: a prospective assessment with MRI--initial results. Stroke 2006; 37:818.
  26. Singh N, Moody AR, Gladstone DJ, et al. Moderate carotid artery stenosis: MR imaging-depicted intraplaque hemorrhage predicts risk of cerebrovascular ischemic events in asymptomatic men. Radiology 2009; 252:502.
  27. Saam T, Hetterich H, Hoffmann V, et al. Meta-analysis and systematic review of the predictive value of carotid plaque hemorrhage on cerebrovascular events by magnetic resonance imaging. J Am Coll Cardiol 2013; 62:1081.
  28. Gupta A, Baradaran H, Schweitzer AD, et al. Carotid plaque MRI and stroke risk: a systematic review and meta-analysis. Stroke 2013; 44:3071.
  29. Hellings WE, Peeters W, Moll FL, et al. Composition of carotid atherosclerotic plaque is associated with cardiovascular outcome: a prognostic study. Circulation 2010; 121:1941.
  30. Baumer Y, Mehta NN, Dey AK, et al. Cholesterol crystals and atherosclerosis. Eur Heart J 2020; 41:2236.
  31. Kitta Y, Obata JE, Nakamura T, et al. Persistent impairment of endothelial vasomotor function has a negative impact on outcome in patients with coronary artery disease. J Am Coll Cardiol 2009; 53:323.
  32. Harrison DG, Armstrong ML, Freiman PC, Heistad DD. Restoration of endothelium-dependent relaxation by dietary treatment of atherosclerosis. J Clin Invest 1987; 80:1808.
  33. John S, Schlaich M, Langenfeld M, et al. Increased bioavailability of nitric oxide after lipid-lowering therapy in hypercholesterolemic patients: a randomized, placebo-controlled, double-blind study. Circulation 1998; 98:211.
  34. Mancini GB, Henry GC, Macaya C, et al. Angiotensin-converting enzyme inhibition with quinapril improves endothelial vasomotor dysfunction in patients with coronary artery disease. The TREND (Trial on Reversing ENdothelial Dysfunction) Study. Circulation 1996; 94:258.
  35. Levine GN, Frei B, Koulouris SN, et al. Ascorbic acid reverses endothelial vasomotor dysfunction in patients with coronary artery disease. Circulation 1996; 93:1107.
  36. Stein JH, Keevil JG, Wiebe DA, et al. Purple grape juice improves endothelial function and reduces the susceptibility of LDL cholesterol to oxidation in patients with coronary artery disease. Circulation 1999; 100:1050.
  37. Paoletti R, Gotto AM Jr, Hajjar DP. Inflammation in atherosclerosis and implications for therapy. Circulation 2004; 109:III20.
  38. Hansson GK. Inflammation, atherosclerosis, and coronary artery disease. N Engl J Med 2005; 352:1685.
  39. Libby P, Ridker PM, Hansson GK, Leducq Transatlantic Network on Atherothrombosis. Inflammation in atherosclerosis: from pathophysiology to practice. J Am Coll Cardiol 2009; 54:2129.
  40. Weber C, Noels H. Atherosclerosis: current pathogenesis and therapeutic options. Nat Med 2011; 17:1410.
  41. Berliner JA, Navab M, Fogelman AM, et al. Atherosclerosis: basic mechanisms. Oxidation, inflammation, and genetics. Circulation 1995; 91:2488.
  42. Steinberg D, Witztum JL. Oxidized low-density lipoprotein and atherosclerosis. Arterioscler Thromb Vasc Biol 2010; 30:2311.
  43. Gawaz M, Neumann FJ, Dickfeld T, et al. Activated platelets induce monocyte chemotactic protein-1 secretion and surface expression of intercellular adhesion molecule-1 on endothelial cells. Circulation 1998; 98:1164.
  44. Tsao PS, Wang B, Buitrago R, et al. Nitric oxide regulates monocyte chemotactic protein-1. Circulation 1997; 96:934.
  45. Takahashi M, Kitagawa S, Masuyama JI, et al. Human monocyte-endothelial cell interaction induces synthesis of granulocyte-macrophage colony-stimulating factor. Circulation 1996; 93:1185.
  46. Rajavashisth T, Qiao JH, Tripathi S, et al. Heterozygous osteopetrotic (op) mutation reduces atherosclerosis in LDL receptor- deficient mice. J Clin Invest 1998; 101:2702.
  47. Rectenwald JE, Moldawer LL, Huber TS, et al. Direct evidence for cytokine involvement in neointimal hyperplasia. Circulation 2000; 102:1697.
  48. Brizzi MF, Formato L, Dentelli P, et al. Interleukin-3 stimulates migration and proliferation of vascular smooth muscle cells: a potential role in atherogenesis. Circulation 2001; 103:549.
  49. Simonini A, Moscucci M, Muller DW, et al. IL-8 is an angiogenic factor in human coronary atherectomy tissue. Circulation 2000; 101:1519.
  50. Li H, Freeman MW, Libby P. Regulation of smooth muscle cell scavenger receptor expression in vivo by atherogenic diets and in vitro by cytokines. J Clin Invest 1995; 95:122.
  51. Tintut Y, Patel J, Parhami F, Demer LL. Tumor necrosis factor-alpha promotes in vitro calcification of vascular cells via the cAMP pathway. Circulation 2000; 102:2636.
  52. Kaartinen M, Penttilä A, Kovanen PT. Mast cells in rupture-prone areas of human coronary atheromas produce and store TNF-alpha. Circulation 1996; 94:2787.
  53. Ridker PM, Everett BM, Thuren T, et al. Antiinflammatory Therapy with Canakinumab for Atherosclerotic Disease. N Engl J Med 2017; 377:1119.
  54. Ridker PM, Cushman M, Stampfer MJ, et al. Inflammation, aspirin, and the risk of cardiovascular disease in apparently healthy men. N Engl J Med 1997; 336:973.
  55. Ridker PM, Hennekens CH, Buring JE, Rifai N. C-reactive protein and other markers of inflammation in the prediction of cardiovascular disease in women. N Engl J Med 2000; 342:836.
  56. Ridker PM, Danielson E, Fonseca FA, et al. Rosuvastatin to prevent vascular events in men and women with elevated C-reactive protein. N Engl J Med 2008; 359:2195.
  57. Davidson MH, Corson MA, Alberts MJ, et al. Consensus panel recommendation for incorporating lipoprotein-associated phospholipase A2 testing into cardiovascular disease risk assessment guidelines. Am J Cardiol 2008; 101:51F.
  58. Young JL, Libby P, Schönbeck U. Cytokines in the pathogenesis of atherosclerosis. Thromb Haemost 2002; 88:554.
  59. Kirii H, Niwa T, Yamada Y, et al. Lack of interleukin-1beta decreases the severity of atherosclerosis in ApoE-deficient mice. Arterioscler Thromb Vasc Biol 2003; 23:656.
  60. Shimokawa H, Ito A, Fukumoto Y, et al. Chronic treatment with interleukin-1 beta induces coronary intimal lesions and vasospastic responses in pigs in vivo. The role of platelet-derived growth factor. J Clin Invest 1996; 97:769.
  61. Duewell P, Kono H, Rayner KJ, et al. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature 2010; 464:1357.
  62. Elhage R, Maret A, Pieraggi MT, et al. Differential effects of interleukin-1 receptor antagonist and tumor necrosis factor binding protein on fatty-streak formation in apolipoprotein E-deficient mice. Circulation 1998; 97:242.
  63. Isoda K, Sawada S, Ishigami N, et al. Lack of interleukin-1 receptor antagonist modulates plaque composition in apolipoprotein E-deficient mice. Arterioscler Thromb Vasc Biol 2004; 24:1068.
  64. Falk E. Pathogenesis of atherosclerosis. J Am Coll Cardiol 2006; 47:C7.
  65. van der Wal AC, Becker AE, van der Loos CM, Das PK. Site of intimal rupture or erosion of thrombosed coronary atherosclerotic plaques is characterized by an inflammatory process irrespective of the dominant plaque morphology. Circulation 1994; 89:36.
  66. van Dijk RA, Duinisveld AJ, Schaapherder AF, et al. A change in inflammatory footprint precedes plaque instability: a systematic evaluation of cellular aspects of the adaptive immune response in human atherosclerosis. J Am Heart Assoc 2015; 4.
  67. Wettinger SB, Doggen CJ, Spek CA, et al. High throughput mRNA profiling highlights associations between myocardial infarction and aberrant expression of inflammatory molecules in blood cells. Blood 2005; 105:2000.
  68. Kiechl S, Lorenz E, Reindl M, et al. Toll-like receptor 4 polymorphisms and atherogenesis. N Engl J Med 2002; 347:185.
  69. Laursen LS, Overgaard MT, Nielsen CG, et al. Substrate specificity of the metalloproteinase pregnancy-associated plasma protein-A (PAPP-A) assessed by mutagenesis and analysis of synthetic peptides: substrate residues distant from the scissile bond are critical for proteolysis. Biochem J 2002; 367:31.
  70. Lawrence JB, Oxvig C, Overgaard MT, et al. The insulin-like growth factor (IGF)-dependent IGF binding protein-4 protease secreted by human fibroblasts is pregnancy-associated plasma protein-A. Proc Natl Acad Sci U S A 1999; 96:3149.
  71. Ortiz CO, Chen BK, Bale LK, et al. Transforming growth factor-beta regulation of the insulin-like growth factor binding protein-4 protease system in cultured human osteoblasts. J Bone Miner Res 2003; 18:1066.
  72. Conover CA, Harrington SC, Bale LK. Differential regulation of pregnancy associated plasma protein-A in human coronary artery endothelial cells and smooth muscle cells. Growth Horm IGF Res 2008; 18:213.
  73. Li W, Li H, Gu F. CRP and TNF-α  induce PAPP-A expression in human peripheral blood mononuclear cells. Mediators Inflamm 2012; 2012:697832.
  74. Sangiorgi G, Mauriello A, Bonanno E, et al. Pregnancy-associated plasma protein-a is markedly expressed by monocyte-macrophage cells in vulnerable and ruptured carotid atherosclerotic plaques: a link between inflammation and cerebrovascular events. J Am Coll Cardiol 2006; 47:2201.
  75. Bayes-Genis A, Conover CA, Overgaard MT, et al. Pregnancy-associated plasma protein A as a marker of acute coronary syndromes. N Engl J Med 2001; 345:1022.
  76. Iversen KK, Dalsgaard M, Teisner AS, et al. Pregnancy-associated plasma protein-A, a marker for outcome in patients suspected for acute coronary syndrome. Clin Biochem 2010; 43:851.
  77. Iversen KK, Dalsgaard M, Teisner AS, et al. Usefulness of pregnancy-associated plasma protein A in patients with acute coronary syndrome. Am J Cardiol 2009; 104:1465.
  78. Gordon T, Castelli WP, Hjortland MC, et al. High density lipoprotein as a protective factor against coronary heart disease. The Framingham Study. Am J Med 1977; 62:707.
  79. Romm PA, Green CE, Reagan K, Rackley CE. Relation of serum lipoprotein cholesterol levels to presence and severity of angiographic coronary artery disease. Am J Cardiol 1991; 67:479.
  80. Manninen V, Tenkanen L, Koskinen P, et al. Joint effects of serum triglyceride and LDL cholesterol and HDL cholesterol concentrations on coronary heart disease risk in the Helsinki Heart Study. Implications for treatment. Circulation 1992; 85:37.
  81. Criqui MH, Heiss G, Cohn R, et al. Plasma triglyceride level and mortality from coronary heart disease. N Engl J Med 1993; 328:1220.
  82. Assmann G, Schulte H. Relation of high-density lipoprotein cholesterol and triglycerides to incidence of atherosclerotic coronary artery disease (the PROCAM experience). Prospective Cardiovascular Münster study. Am J Cardiol 1992; 70:733.
  83. Summary of the second report of the National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults (Adult Treatment Panel II). JAMA 1993; 269:3015.
  84. Report of the National Cholesterol Education Program Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults. The Expert Panel. Arch Intern Med 1988; 148:36.
  85. Iuliano L, Mauriello A, Sbarigia E, et al. Radiolabeled native low-density lipoprotein injected into patients with carotid stenosis accumulates in macrophages of atherosclerotic plaque : effect of vitamin E supplementation. Circulation 2000; 101:1249.
  86. Podrez EA, Febbraio M, Sheibani N, et al. Macrophage scavenger receptor CD36 is the major receptor for LDL modified by monocyte-generated reactive nitrogen species. J Clin Invest 2000; 105:1095.
  87. Febbraio M, Hajjar DP, Silverstein RL. CD36: a class B scavenger receptor involved in angiogenesis, atherosclerosis, inflammation, and lipid metabolism. J Clin Invest 2001; 108:785.
  88. Tabas I. Consequences of cellular cholesterol accumulation: basic concepts and physiological implications. J Clin Invest 2002; 110:905.
  89. Silverman MG, Ference BA, Im K, et al. Association Between Lowering LDL-C and Cardiovascular Risk Reduction Among Different Therapeutic Interventions: A Systematic Review and Meta-analysis. JAMA 2016; 316:1289.
  90. Sabatine MS, Giugliano RP, Keech AC, et al. Evolocumab and Clinical Outcomes in Patients with Cardiovascular Disease. N Engl J Med 2017; 376:1713.
  91. National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults (Adult Treatment Panel III). Third Report of the National Cholesterol Education Program (NCEP) Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults (Adult Treatment Panel III) final report. Circulation 2002; 106:3143.
  92. AIM-HIGH Investigators, Boden WE, Probstfield JL, et al. Niacin in patients with low HDL cholesterol levels receiving intensive statin therapy. N Engl J Med 2011; 365:2255.
  93. Schwartz GG, Olsson AG, Abt M, et al. Effects of dalcetrapib in patients with a recent acute coronary syndrome. N Engl J Med 2012; 367:2089.
  94. HPS2-THRIVE Collaborative Group. HPS2-THRIVE randomized placebo-controlled trial in 25 673 high-risk patients of ER niacin/laropiprant: trial design, pre-specified muscle and liver outcomes, and reasons for stopping study treatment. Eur Heart J 2013; 34:1279.
  95. Voight BF, Peloso GM, Orho-Melander M, et al. Plasma HDL cholesterol and risk of myocardial infarction: a mendelian randomisation study. Lancet 2012; 380:572.
  96. Khera AV, Cuchel M, de la Llera-Moya M, et al. Cholesterol efflux capacity, high-density lipoprotein function, and atherosclerosis. N Engl J Med 2011; 364:127.
  97. Rohatgi A, Khera A, Berry JD, et al. HDL cholesterol efflux capacity and incident cardiovascular events. N Engl J Med 2014; 371:2383.
  98. Varbo A, Benn M, Tybjærg-Hansen A, et al. Remnant cholesterol as a causal risk factor for ischemic heart disease. J Am Coll Cardiol 2013; 61:427.
  99. Varbo A, Benn M, Tybjærg-Hansen A, Nordestgaard BG. Elevated remnant cholesterol causes both low-grade inflammation and ischemic heart disease, whereas elevated low-density lipoprotein cholesterol causes ischemic heart disease without inflammation. Circulation 2013; 128:1298.
  100. Khine HW, Teiber JF, Haley RW, et al. Association of the serum myeloperoxidase/high-density lipoprotein particle ratio and incident cardiovascular events in a multi-ethnic population: Observations from the Dallas Heart Study. Atherosclerosis 2017; 263:156.
  101. Nakajima K, Nakano T, Tanaka A. The oxidative modification hypothesis of atherosclerosis: the comparison of atherogenic effects on oxidized LDL and remnant lipoproteins in plasma. Clin Chim Acta 2006; 367:36.
  102. Berliner JA, Leitinger N, Tsimikas S. The role of oxidized phospholipids in atherosclerosis. J Lipid Res 2009; 50 Suppl:S207.
  103. Ehara S, Ueda M, Naruko T, et al. Elevated levels of oxidized low density lipoprotein show a positive relationship with the severity of acute coronary syndromes. Circulation 2001; 103:1955.
  104. Tsimikas S, Bergmark C, Beyer RW, et al. Temporal increases in plasma markers of oxidized low-density lipoprotein strongly reflect the presence of acute coronary syndromes. J Am Coll Cardiol 2003; 41:360.
  105. Inoue T, Uchida T, Kamishirado H, et al. Clinical significance of antibody against oxidized low density lipoprotein in patients with atherosclerotic coronary artery disease. J Am Coll Cardiol 2001; 37:775.
  106. Bui MN, Sack MN, Moutsatsos G, et al. Autoantibody titers to oxidized low-density lipoprotein in patients with coronary atherosclerosis. Am Heart J 1996; 131:663.
  107. Boffa MB, Koschinsky ML. Update on lipoprotein(a) as a cardiovascular risk factor and mediator. Curr Atheroscler Rep 2013; 15:360.
  108. Anglés-Cano E, de la Peña Díaz A, Loyau S. Inhibition of fibrinolysis by lipoprotein(a). Ann N Y Acad Sci 2001; 936:261.
  109. Seimon TA, Nadolski MJ, Liao X, et al. Atherogenic lipids and lipoproteins trigger CD36-TLR2-dependent apoptosis in macrophages undergoing endoplasmic reticulum stress. Cell Metab 2010; 12:467.
  110. Cho T, Jung Y, Koschinsky ML. Apolipoprotein(a), through its strong lysine-binding site in KIV(10'), mediates increased endothelial cell contraction and permeability via a Rho/Rho kinase/MYPT1-dependent pathway. J Biol Chem 2008; 283:30503.
  111. Cho T, Romagnuolo R, Scipione C, et al. Apolipoprotein(a) stimulates nuclear translocation of β-catenin: a novel pathogenic mechanism for lipoprotein(a). Mol Biol Cell 2013; 24:210.
  112. Zioncheck TF, Powell LM, Rice GC, et al. Interaction of recombinant apolipoprotein(a) and lipoprotein(a) with macrophages. J Clin Invest 1991; 87:767.
  113. Albers JJ, Slee A, O'Brien KD, et al. Relationship of apolipoproteins A-1 and B, and lipoprotein(a) to cardiovascular outcomes: the AIM-HIGH trial (Atherothrombosis Intervention in Metabolic Syndrome with Low HDL/High Triglyceride and Impact on Global Health Outcomes). J Am Coll Cardiol 2013; 62:1575.
  114. DAWBER TR, MEADORS GF, MOORE FE Jr. Epidemiological approaches to heart disease: the Framingham Study. Am J Public Health Nations Health 1951; 41:279.
  115. Yusuf S, Hawken S, Ounpuu S, et al. Effect of potentially modifiable risk factors associated with myocardial infarction in 52 countries (the INTERHEART study): case-control study. Lancet 2004; 364:937.
  116. Frohlich J, Al-Sarraf A. Cardiovascular risk and atherosclerosis prevention. Cardiovasc Pathol 2013; 22:16.
  117. Ambrose JA, Barua RS. The pathophysiology of cigarette smoking and cardiovascular disease: an update. J Am Coll Cardiol 2004; 43:1731.
  118. Mayhan WG, Sharpe GM. Effect of cigarette smoke extract on arteriolar dilatation in vivo. J Appl Physiol (1985) 1996; 81:1996.
  119. Mayhan WG, Patel KP. Effect of nicotine on endothelium-dependent arteriolar dilatation in vivo. Am J Physiol 1997; 272:H2337.
  120. Ota Y, Kugiyama K, Sugiyama S, et al. Impairment of endothelium-dependent relaxation of rabbit aortas by cigarette smoke extract--role of free radicals and attenuation by captopril. Atherosclerosis 1997; 131:195.
  121. Tracy RP, Psaty BM, Macy E, et al. Lifetime smoking exposure affects the association of C-reactive protein with cardiovascular disease risk factors and subclinical disease in healthy elderly subjects. Arterioscler Thromb Vasc Biol 1997; 17:2167.
  122. Bermudez EA, Rifai N, Buring JE, et al. Relation between markers of systemic vascular inflammation and smoking in women. Am J Cardiol 2002; 89:1117.
  123. Mendall MA, Patel P, Asante M, et al. Relation of serum cytokine concentrations to cardiovascular risk factors and coronary heart disease. Heart 1997; 78:273.
  124. Tappia PS, Troughton KL, Langley-Evans SC, Grimble RF. Cigarette smoking influences cytokine production and antioxidant defences. Clin Sci (Lond) 1995; 88:485.
  125. Ichiki K, Ikeda H, Haramaki N, et al. Long-term smoking impairs platelet-derived nitric oxide release. Circulation 1996; 94:3109.
  126. Sawada M, Kishi Y, Numano F, Isobe M. Smokers lack morning increase in platelet sensitivity to nitric oxide. J Cardiovasc Pharmacol 2002; 40:571.
  127. Kannel WB, D'Agostino RB, Belanger AJ. Fibrinogen, cigarette smoking, and risk of cardiovascular disease: insights from the Framingham Study. Am Heart J 1987; 113:1006.
  128. Smith FB, Lee AJ, Fowkes FG, et al. Hemostatic factors as predictors of ischemic heart disease and stroke in the Edinburgh Artery Study. Arterioscler Thromb Vasc Biol 1997; 17:3321.
  129. Barua RS, Ambrose JA, Saha DC, Eales-Reynolds LJ. Smoking is associated with altered endothelial-derived fibrinolytic and antithrombotic factors: an in vitro demonstration. Circulation 2002; 106:905.
  130. Heitzer T, Ylä-Herttuala S, Luoma J, et al. Cigarette smoking potentiates endothelial dysfunction of forearm resistance vessels in patients with hypercholesterolemia. Role of oxidized LDL. Circulation 1996; 93:1346.
  131. Nishio E, Watanabe Y. Cigarette smoke extract inhibits plasma paraoxonase activity by modification of the enzyme's free thiols. Biochem Biophys Res Commun 1997; 236:289.
  132. Reaven G, Tsao PS. Insulin resistance and compensatory hyperinsulinemia: the key player between cigarette smoking and cardiovascular disease? J Am Coll Cardiol 2003; 41:1044.
  133. Haffner SM, Stern MP, Hazuda HP, et al. Hyperinsulinemia in a population at high risk for non-insulin-dependent diabetes mellitus. N Engl J Med 1986; 315:220.
  134. Rönnemaa T, Laakso M, Pyörälä K, et al. High fasting plasma insulin is an indicator of coronary heart disease in non-insulin-dependent diabetic patients and nondiabetic subjects. Arterioscler Thromb 1991; 11:80.
  135. Welham MJ, Bone H, Levings M, et al. Insulin receptor substrate-2 is the major 170-kDa protein phosphorylated on tyrosine in response to cytokines in murine lymphohemopoietic cells. J Biol Chem 1997; 272:1377.
  136. Malide D, Davies-Hill TM, Levine M, Simpson IA. Distinct localization of GLUT-1, -3, and -5 in human monocyte-derived macrophages: effects of cell activation. Am J Physiol 1998; 274:E516.
  137. Park YM, R Kashyap S, A Major J, Silverstein RL. Insulin promotes macrophage foam cell formation: potential implications in diabetes-related atherosclerosis. Lab Invest 2012; 92:1171.
  138. Fernández-Real JM, Ricart W. Insulin resistance and chronic cardiovascular inflammatory syndrome. Endocr Rev 2003; 24:278.
  139. Saely CH, Rein P, Vonbank A, et al. Type 2 diabetes and the progression of visualized atherosclerosis to clinical cardiovascular events. Int J Cardiol 2013; 167:776.
  140. Hasenstab D, Lea H, Hart CE, et al. Tissue factor overexpression in rat arterial neointima models thrombosis and progression of advanced atherosclerosis. Circulation 2000; 101:2651.
  141. Poznyak AV, Bharadwaj D, Prasad G, et al. Renin-Angiotensin System in Pathogenesis of Atherosclerosis and Treatment of CVD. Int J Mol Sci 2021; 22.
  142. Kovanen PT. Mast Cells as Potential Accelerators of Human Atherosclerosis-From Early to Late Lesions. Int J Mol Sci 2019; 20.
  143. Poznyak AV, Wu WK, Melnichenko AA, et al. Signaling Pathways and Key Genes Involved in Regulation of foam Cell Formation in Atherosclerosis. Cells 2020; 9.
  144. Sata M, Fukuda D. Crucial role of renin-angiotensin system in the pathogenesis of atherosclerosis. J Med Invest 2010; 57:12.
  145. Daugherty A, Manning MW, Cassis LA. Angiotensin II promotes atherosclerotic lesions and aneurysms in apolipoprotein E-deficient mice. J Clin Invest 2000; 105:1605.
  146. Potter DD, Sobey CG, Tompkins PK, et al. Evidence that macrophages in atherosclerotic lesions contain angiotensin II. Circulation 1998; 98:800.
  147. Strawn WB, Chappell MC, Dean RH, et al. Inhibition of early atherogenesis by losartan in monkeys with diet-induced hypercholesterolemia. Circulation 2000; 101:1586.
  148. Lerman A, Edwards BS, Hallett JW, et al. Circulating and tissue endothelin immunoreactivity in advanced atherosclerosis. N Engl J Med 1991; 325:997.
  149. Ihling C, Szombathy T, Bohrmann B, et al. Coexpression of endothelin-converting enzyme-1 and endothelin-1 in different stages of human atherosclerosis. Circulation 2001; 104:864.
  150. Mathew V, Cannan CR, Miller VM, et al. Enhanced endothelin-mediated coronary vasoconstriction and attenuated basal nitric oxide activity in experimental hypercholesterolemia. Circulation 1997; 96:1930.
  151. Hasdai D, Holmes DR Jr, Garratt KN, et al. Mechanical pressure and stretch release endothelin-1 from human atherosclerotic coronary arteries in vivo. Circulation 1997; 95:357.
  152. Minamino T, Kurihara H, Takahashi M, et al. Endothelin-converting enzyme expression in the rat vascular injury model and human coronary atherosclerosis. Circulation 1997; 95:221.
  153. Iiyama K, Hajra L, Iiyama M, et al. Patterns of vascular cell adhesion molecule-1 and intercellular adhesion molecule-1 expression in rabbit and mouse atherosclerotic lesions and at sites predisposed to lesion formation. Circ Res 1999; 85:199.
  154. Cybulsky MI, Iiyama K, Li H, et al. A major role for VCAM-1, but not ICAM-1, in early atherosclerosis. J Clin Invest 2001; 107:1255.
  155. Johnson RC, Chapman SM, Dong ZM, et al. Absence of P-selectin delays fatty streak formation in mice. J Clin Invest 1997; 99:1037.
  156. Dong ZM, Brown AA, Wagner DD. Prominent role of P-selectin in the development of advanced atherosclerosis in ApoE-deficient mice. Circulation 2000; 101:2290.
  157. Adams MR, Jessup W, Hailstones D, Celermajer DS. L-arginine reduces human monocyte adhesion to vascular endothelium and endothelial expression of cell adhesion molecules. Circulation 1997; 95:662.
  158. McCrohon JA, Jessup W, Handelsman DJ, Celermajer DS. Androgen exposure increases human monocyte adhesion to vascular endothelium and endothelial cell expression of vascular cell adhesion molecule-1. Circulation 1999; 99:2317.
  159. Patel SS, Thiagarajan R, Willerson JT, Yeh ET. Inhibition of alpha4 integrin and ICAM-1 markedly attenuate macrophage homing to atherosclerotic plaques in ApoE-deficient mice. Circulation 1998; 97:75.
  160. Hagiwara H, Mitsumata M, Yamane T, et al. Laminar shear stress-induced GRO mRNA and protein expression in endothelial cells. Circulation 1998; 98:2584.
  161. Tsao PS, Buitrago R, Chan JR, Cooke JP. Fluid flow inhibits endothelial adhesiveness. Nitric oxide and transcriptional regulation of VCAM-1. Circulation 1996; 94:1682.
  162. Gimbrone MA Jr, García-Cardeña G. Vascular endothelium, hemodynamics, and the pathobiology of atherosclerosis. Cardiovasc Pathol 2013; 22:9.
  163. Nwasokwa ON, Weiss M, Gladstone C, Bodenheimer MM. Effect of coronary artery size on the prevalence of atherosclerosis. Am J Cardiol 1996; 78:741.
  164. Gnasso A, Carallo C, Irace C, et al. Association between intima-media thickness and wall shear stress in common carotid arteries in healthy male subjects. Circulation 1996; 94:3257.
  165. Yu E, Mercer J, Bennett M. Mitochondria in vascular disease. Cardiovasc Res 2012; 95:173.
  166. Davidson SM, Yellon DM. Mitochondrial DNA damage, oxidative stress, and atherosclerosis: where there is smoke there is not always fire. Circulation 2013; 128:681.
  167. Yu E, Calvert PA, Mercer JR, et al. Mitochondrial DNA damage can promote atherosclerosis independently of reactive oxygen species through effects on smooth muscle cells and monocytes and correlates with higher-risk plaques in humans. Circulation 2013; 128:702.
  168. Calvert PA, Obaid DR, O'Sullivan M, et al. Association between IVUS findings and adverse outcomes in patients with coronary artery disease: the VIVA (VH-IVUS in Vulnerable Atherosclerosis) Study. JACC Cardiovasc Imaging 2011; 4:894.
  169. Stone GW, Maehara A, Lansky AJ, et al. A prospective natural-history study of coronary atherosclerosis. N Engl J Med 2011; 364:226.
  170. Lusis AJ, Mar R, Pajukanta P. Genetics of atherosclerosis. Annu Rev Genomics Hum Genet 2004; 5:189.
  171. Hansson GK, Libby P. The immune response in atherosclerosis: a double-edged sword. Nat Rev Immunol 2006; 6:508.
  172. Arnett DK, Baird AE, Barkley RA, et al. Relevance of genetics and genomics for prevention and treatment of cardiovascular disease: a scientific statement from the American Heart Association Council on Epidemiology and Prevention, the Stroke Council, and the Functional Genomics and Translational Biology Interdisciplinary Working Group. Circulation 2007; 115:2878.
  173. Chen Y, Rollins J, Paigen B, Wang X. Genetic and genomic insights into the molecular basis of atherosclerosis. Cell Metab 2007; 6:164.
  174. Cameron J, Holla ØL, Ranheim T, et al. Effect of mutations in the PCSK9 gene on the cell surface LDL receptors. Hum Mol Genet 2006; 15:1551.
  175. Abifadel M, Rabès JP, Devillers M, et al. Mutations and polymorphisms in the proprotein convertase subtilisin kexin 9 (PCSK9) gene in cholesterol metabolism and disease. Hum Mutat 2009; 30:520.
  176. Horton JD, Cohen JC, Hobbs HH. PCSK9: a convertase that coordinates LDL catabolism. J Lipid Res 2009; 50 Suppl:S172.
  177. Cohen J, Pertsemlidis A, Kotowski IK, et al. Low LDL cholesterol in individuals of African descent resulting from frequent nonsense mutations in PCSK9. Nat Genet 2005; 37:161.
  178. Zhao Z, Tuakli-Wosornu Y, Lagace TA, et al. Molecular characterization of loss-of-function mutations in PCSK9 and identification of a compound heterozygote. Am J Hum Genet 2006; 79:514.
  179. Cohen JC, Boerwinkle E, Mosley TH Jr, Hobbs HH. Sequence variations in PCSK9, low LDL, and protection against coronary heart disease. N Engl J Med 2006; 354:1264.
  180. Benn M, Nordestgaard BG, Grande P, et al. PCSK9 R46L, low-density lipoprotein cholesterol levels, and risk of ischemic heart disease: 3 independent studies and meta-analyses. J Am Coll Cardiol 2010; 55:2833.
  181. Daugherty A, Tabas I, Rader DJ. Accelerating the pace of atherosclerosis research. Arterioscler Thromb Vasc Biol 2015; 35:11.
  182. Joshi R, Khandelwal B, Joshi D, Gupta OP. Chlamydophila pneumoniae infection and cardiovascular disease. N Am J Med Sci 2013; 5:169.
  183. Adler SP, Hur JK, Wang JB, Vetrovec GW. Prior infection with cytomegalovirus is not a major risk factor for angiographically demonstrated coronary artery atherosclerosis. J Infect Dis 1998; 177:209.
  184. Blum A, Giladi M, Weinberg M, et al. High anti-cytomegalovirus (CMV) IgG antibody titer is associated with coronary artery disease and may predict post-coronary balloon angioplasty restenosis. Am J Cardiol 1998; 81:866.
  185. Ridker PM, Hennekens CH, Stampfer MJ, Wang F. Prospective study of herpes simplex virus, cytomegalovirus, and the risk of future myocardial infarction and stroke. Circulation 1998; 98:2796.
  186. Sorlie PD, Nieto FJ, Adam E, et al. A prospective study of cytomegalovirus, herpes simplex virus 1, and coronary heart disease: the atherosclerosis risk in communities (ARIC) study. Arch Intern Med 2000; 160:2027.
  187. Zhu J, Nieto FJ, Horne BD, et al. Prospective study of pathogen burden and risk of myocardial infarction or death. Circulation 2001; 103:45.
  188. Rupprecht HJ, Blankenberg S, Bickel C, et al. Impact of viral and bacterial infectious burden on long-term prognosis in patients with coronary artery disease. Circulation 2001; 104:25.
  189. Nieto FJ, Adam E, Sorlie P, et al. Cohort study of cytomegalovirus infection as a risk factor for carotid intimal-medial thickening, a measure of subclinical atherosclerosis. Circulation 1996; 94:922.
  190. Melnick JL, Hu C, Burek J, et al. Cytomegalovirus DNA in arterial walls of patients with atherosclerosis. J Med Virol 1994; 42:170.
  191. Zhou YF, Guetta E, Yu ZX, et al. Human cytomegalovirus increases modified low density lipoprotein uptake and scavenger receptor mRNA expression in vascular smooth muscle cells. J Clin Invest 1996; 98:2129.
  192. Zhou YF, Shou M, Guetta E, et al. Cytomegalovirus infection of rats increases the neointimal response to vascular injury without consistent evidence of direct infection of the vascular wall. Circulation 1999; 100:1569.
  193. Epstein SE, Zhou YF, Zhu J. Infection and atherosclerosis: emerging mechanistic paradigms. Circulation 1999; 100:e20.
  194. Patel P, Mendall MA, Carrington D, et al. Association of Helicobacter pylori and Chlamydia pneumoniae infections with coronary heart disease and cardiovascular risk factors. BMJ 1995; 311:711.
  195. Gunn M, Stephens JC, Thompson JR, et al. Significant association of cagA positive Helicobacter pylori strains with risk of premature myocardial infarction. Heart 2000; 84:267.
  196. Danesh J, Peto R. Risk factors for coronary heart disease and infection with Helicobacter pylori: meta-analysis of 18 studies. BMJ 1998; 316:1130.
  197. Espinola-Klein C, Rupprecht HJ, Blankenberg S, et al. Impact of infectious burden on extent and long-term prognosis of atherosclerosis. Circulation 2002; 105:15.
  198. Prasad A, Zhu J, Halcox JP, et al. Predisposition to atherosclerosis by infections: role of endothelial dysfunction. Circulation 2002; 106:184.
  199. Anderson JL, Carlquist JF, Muhlestein JB, et al. Evaluation of C-reactive protein, an inflammatory marker, and infectious serology as risk factors for coronary artery disease and myocardial infarction. J Am Coll Cardiol 1998; 32:35.
  200. Mayr M, Kiechl S, Willeit J, et al. Infections, immunity, and atherosclerosis: associations of antibodies to Chlamydia pneumoniae, Helicobacter pylori, and cytomegalovirus with immune reactions to heat-shock protein 60 and carotid or femoral atherosclerosis. Circulation 2000; 102:833.
  201. Ott SJ, El Mokhtari NE, Musfeldt M, et al. Detection of diverse bacterial signatures in atherosclerotic lesions of patients with coronary heart disease. Circulation 2006; 113:929.
  202. Gurfinkel EP, de la Fuente RL, Mendiz O, Mautner B. Influenza vaccine pilot study in acute coronary syndromes and planned percutaneous coronary interventions: the FLU Vaccination Acute Coronary Syndromes (FLUVACS) Study. Circulation 2002; 105:2143.
  203. Nichol KL, Nordin J, Mullooly J, et al. Influenza vaccination and reduction in hospitalizations for cardiac disease and stroke among the elderly. N Engl J Med 2003; 348:1322.
  204. Naghavi M, Barlas Z, Siadaty S, et al. Association of influenza vaccination and reduced risk of recurrent myocardial infarction. Circulation 2000; 102:3039.
  205. Wang Z, Klipfell E, Bennett BJ, et al. Gut flora metabolism of phosphatidylcholine promotes cardiovascular disease. Nature 2011; 472:57.
  206. Koeth RA, Wang Z, Levison BS, et al. Intestinal microbiota metabolism of L-carnitine, a nutrient in red meat, promotes atherosclerosis. Nat Med 2013; 19:576.
  207. Tang WH, Wang Z, Levison BS, et al. Intestinal microbial metabolism of phosphatidylcholine and cardiovascular risk. N Engl J Med 2013; 368:1575.
  208. Tang WH, Wang Z, Fan Y, et al. Prognostic value of elevated levels of intestinal microbe-generated metabolite trimethylamine-N-oxide in patients with heart failure: refining the gut hypothesis. J Am Coll Cardiol 2014; 64:1908.
  209. Wang Z, Tang WH, Buffa JA, et al. Prognostic value of choline and betaine depends on intestinal microbiota-generated metabolite trimethylamine-N-oxide. Eur Heart J 2014; 35:904.
  210. Li XS, Obeid S, Klingenberg R, et al. Gut microbiota-dependent trimethylamine N-oxide in acute coronary syndromes: a prognostic marker for incident cardiovascular events beyond traditional risk factors. Eur Heart J 2017; 38:814.
  211. Haghikia A, Li XS, Liman TG, et al. Gut Microbiota-Dependent Trimethylamine N-Oxide Predicts Risk of Cardiovascular Events in Patients With Stroke and Is Related to Proinflammatory Monocytes. Arterioscler Thromb Vasc Biol 2018; 38:2225.
  212. Li XS, Obeid S, Wang Z, et al. Trimethyllysine, a trimethylamine N-oxide precursor, provides near- and long-term prognostic value in patients presenting with acute coronary syndromes. Eur Heart J 2019; 40:2700.
  213. Seldin MM, Meng Y, Qi H, et al. Trimethylamine N-Oxide Promotes Vascular Inflammation Through Signaling of Mitogen-Activated Protein Kinase and Nuclear Factor-κB. J Am Heart Assoc 2016; 5.
  214. Ma G, Pan B, Chen Y, et al. Trimethylamine N-oxide in atherogenesis: impairing endothelial self-repair capacity and enhancing monocyte adhesion. Biosci Rep 2017; 37.
  215. Sun X, Jiao X, Ma Y, et al. Trimethylamine N-oxide induces inflammation and endothelial dysfunction in human umbilical vein endothelial cells via activating ROS-TXNIP-NLRP3 inflammasome. Biochem Biophys Res Commun 2016; 481:63.
  216. Boini KM, Hussain T, Li PL, Koka S. Trimethylamine-N-Oxide Instigates NLRP3 Inflammasome Activation and Endothelial Dysfunction. Cell Physiol Biochem 2017; 44:152.
  217. Falk E, Shah PK, Fuster V. Coronary plaque disruption. Circulation 1995; 92:657.
  218. Little WC, Constantinescu M, Applegate RJ, et al. Can coronary angiography predict the site of a subsequent myocardial infarction in patients with mild-to-moderate coronary artery disease? Circulation 1988; 78:1157.
  219. Ambrose JA, Tannenbaum MA, Alexopoulos D, et al. Angiographic progression of coronary artery disease and the development of myocardial infarction. J Am Coll Cardiol 1988; 12:56.
  220. Falk E, Nakano M, Bentzon JF, et al. Update on acute coronary syndromes: the pathologists' view. Eur Heart J 2013; 34:719.
  221. Virmani R, Kolodgie FD, Burke AP, et al. Lessons from sudden coronary death: a comprehensive morphological classification scheme for atherosclerotic lesions. Arterioscler Thromb Vasc Biol 2000; 20:1262.
  222. Kubo T, Imanishi T, Takarada S, et al. Assessment of culprit lesion morphology in acute myocardial infarction: ability of optical coherence tomography compared with intravascular ultrasound and coronary angioscopy. J Am Coll Cardiol 2007; 50:933.
  223. Burke AP, Kolodgie FD, Farb A, et al. Healed plaque ruptures and sudden coronary death: evidence that subclinical rupture has a role in plaque progression. Circulation 2001; 103:934.
  224. Vergallo R, Crea F. Atherosclerotic Plaque Healing. N Engl J Med 2020; 383:846.
Topic 13603 Version 29.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟