Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3870 2023-11-02 18:36:31 |
2 references update Meta information modification 3870 2023-11-03 06:18:26 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Laera, N.; Malerba, P.; Vacanti, G.; Nardin, S.; Pagnesi, M.; Nardin, M. Trained Immunity in Atherosclerosis. Encyclopedia. Available online: https://encyclopedia.pub/entry/51111 (accessed on 07 July 2024).
Laera N, Malerba P, Vacanti G, Nardin S, Pagnesi M, Nardin M. Trained Immunity in Atherosclerosis. Encyclopedia. Available at: https://encyclopedia.pub/entry/51111. Accessed July 07, 2024.
Laera, Nicola, Paolo Malerba, Gaetano Vacanti, Simone Nardin, Matteo Pagnesi, Matteo Nardin. "Trained Immunity in Atherosclerosis" Encyclopedia, https://encyclopedia.pub/entry/51111 (accessed July 07, 2024).
Laera, N., Malerba, P., Vacanti, G., Nardin, S., Pagnesi, M., & Nardin, M. (2023, November 02). Trained Immunity in Atherosclerosis. In Encyclopedia. https://encyclopedia.pub/entry/51111
Laera, Nicola, et al. "Trained Immunity in Atherosclerosis." Encyclopedia. Web. 02 November, 2023.
Trained Immunity in Atherosclerosis
Edit

Coronary artery disease (CAD) is the leading cause of death worldwide. It is a result of the buildup of atherosclerosis within the coronary arteries. The role of the immune system in CAD is complex and multifaceted. The immune system responds to damage or injury to the arterial walls by initiating an inflammatory response. However, this inflammatory response can become chronic and lead to plaque formation. Neutrophiles, macrophages, B lymphocytes, T lymphocytes, and NKT cells play a key role in immunity response, both with proatherogenic and antiatherogenic signaling pathways.

atherosclerosis trained immunity

1. Introduction

Atherosclerosis is the leading cause of mortality worldwide and one of its main manifestations is ischemic heart disease, mostly determined by the involvement of the coronary arteries, known as coronary artery disease (CAD). CAD is primarily a result of the buildup of atherosclerosis within the coronary arteries, progressively leading to the narrowing of the vessel lumen, thus reducing the blood flow with consequent ischemia. Upon stimulation by deposited lipids and damaged endothelium, innate and adaptive immune cells are activated and recruited to initiate plaque development [1]. The role of the immune system in CAD has been shown during the last few years to be complex and multifaceted. Immunity plays a key role in several types of pathways. First, inflammation is a key part of atherosclerosis [2]. Both innate and adaptive immune responses are activated to remove dead and apoptotic cells, facilitate scar formation, and promote angiogenesis [3]. Innate immunity includes neutrophils and macrophages, which can directly phagocytose dead cells and debris. They also play a role in releasing cytokines and other molecules that promote an inflammatory response [4][5]. Adaptive immunity involves T cells and B cells, which can recognize and respond to specific antigens. This response can also shape the overall immune response to an injury [6][7]. However, in the context of atherosclerosis, this inflammatory response can become chronic and lead to plaque formation. Immune cells, particularly macrophages, play a pivotal role in the uptake of low-density lipoprotein (LDL), mostly oxidized-LDL (ox-LDL), leading to the formation of foam cells within the atherosclerotic plaque [5]

2. Immune Cells Involved in CAD

2.1. Neutrophils

Neutrophils are the most abundant leucocytes and play a significant role in mediating sterile inflammation and injury through a variety of mechanisms [4][8]. Previous experimental works have elucidated their roles in atherosclerotic diseases like CAD and its ensuing complications, i.e., acute coronary syndrome and heart failure [8]. Early aortic lesions and rupture- or erosion-prone atherosclerotic plaques show a significant presence of neutrophils [9]. A high peripheral neutrophil count directly relates to the degree of atherosclerosis in coronary arteries [10], infarct size, and declines in left ventricular ejection fraction [11][12]; the neutrophil-to-lymphocytes ratio raises clinical attention, due to its potential relationship with CAD [13]. Recent evidence also suggests a role for neutrophils in the activation of reparative processes [14]. In animal studies, for example, long-term depletion of neutrophils after myocardial infarction (MI) resulted in worsened cardiac function and increased fibrosis [14]. Neutrophils are paramount for innate immune response as they are the first responders in our defense against invading pathogenic microorganisms [4]. However, in sterile inflammatory conditions, activation of neutrophils may have detrimental effects on host tissues and therefore their homeostasis must be tightly regulated [15]. Interestingly, most clinically recognized cardiovascular risk factors contribute to enhanced granulopoiesis, e.g., the production of neutrophils in the bone marrow [4]. Once production bursts are triggered by those risk factors, neutrophils play a leading role in the initiation and evolution of unstable atherosclerotic plaques. At sites of disturbed blood flow and increased shear stress, they dysregulate vascular endothelial cells (ECs) and trigger leucocyte arrest [16], setting the stage for atherosclerosis [17]. Further release of granule proteins degrades the extracellular matrix (ECM), leading to extra-adhesion of monocytes, vascular hyperpermeability, and transfer of LDL particles [18]. Through the nucleotide-binding oligomerisation domain-like receptor pyrin domain-containing protein 3 (NLRP3) inflammasome signaling, activated neutrophils in the atheroma undergo neutrophil extracellular trap (NET) formation (NETosis), a type of cell death [19] using a network of ECM containing a variety of granule proteins. Their release decreases the stability of atherosclerotic plaques and contributes to thrombus formation through a variety of mechanisms [20][21][22][23]. Not only do neutrophils play an active role in the genesis of atherosclerotic plaques, they also mediate their consequences to the target organs. Rupture of an atherosclerotic plaque can lead to obstruction of blood circulation resulting in ischaemic death of tissues that immediately triggers an acute inflammatory response, led mostly by neutrophils [20]. Attracted by cellular debris released by dead cells, neutrophils massively infiltrate the infarct area within hours [21]. At the site of injury, activated neutrophils generate and release reactive oxygen species, proteases, and NETs [22] which promote cardiomyocyte apoptosis, degrade ECM [23], lead to leucocyte infiltration, and prime the NLRP3 inflammasome, which then stimulates granulopoiesis in the bone marrow, leading to a vicious circle of the neutrophil infiltration circle and maladaptive remodeling [22]. Neutrophils are also involved in the modulation of the healing and remodeling response: the protein S100A8/A9 in NETs activated macrophages to phagocyte dead cells [24]. The transcriptional profile of neutrophils changes from a pro-inflammatory profile to an anti-inflammatory profile, initiating the reparative process mostly by dedifferentiating cardiomyocytes and promoting the accumulation of reparative macrophages [25][26]. Finally, a subset of pro-angiogenesis neutrophils able to control blood vessel growth, a known mechanism of tissue regeneration after injury, has been recently identified [27][28].
This dichotomous and apparently paradoxical effect of neutrophil activity in atherosclerotic diseases highlights the difficulty of researching beneficial therapeutical strategies involving neutrophils and calls for an expertly tailored approach to the matter. In the short term, stunning inhibition of neutrophils such as through beta-adrenergic antagonists as metoprolol or through inhibition of S100A8/A9 has been shown to reduce infarct size and increase left ventricular ejection fraction [29][30]. However, long-lasting depletion of neutrophils or even long-term inhibition of S100A8/A9 resulted in worse cardiac function and increased fibrosis [24][31]. Identifying the right window to effectively suppress the inflammatory functions of neutrophils while retaining their reparative functions could pave the way for important therapeutical applications.

2.2. Macrophages

Just as for neutrophils, the role of macrophages in inflammation, particularly in its cardiovascular aspects, is multifaceted [5][32]. Many studies have shown their ability to trigger and drive robust and damaging inflammatory responses [33], while others have shown their involvement in tissue repair and even cardiac regeneration [5][34][35]. This seems to be because different macrophage populations mediate different responses [36]. Furthermore, due to their high plasticity, they can adopt different phenotypes in response to varying stimuli and environments, a process called polarization [32]. Classically, cardiac macrophages have been categorized according to inflammatory states and cell surface markers into pro-inflammatory (M1) and anti-inflammatory (M2) subsets [25].
Newly discovered subsets of macrophages with mixed M1/M2 cell surface markers have challenged the adequacy of the classification [37]. In vitro, culture of human macrophages revealed considerable deviation from the M1/M2 spectrum when in contact with a cardiovascular relevant stimulus, e.g., free fatty acids or high-density lipoprotein (HDL) [38]. Furthermore, in vivo studies of murine atherosclerosis showed how, for example, inflammatory macrophages express cell surface markers like CD206, typically used to define the M2 anti-inflammatory subset [39]. In other words, the traditional classification of macrophages into M1 and M2 phenotypes does not fully capture the diversity of the population in vivo.
An alternative approach would classify macrophages according to developmental lineage, transcriptional factors, and recruiting dynamic as well as cell surface markers [5]. Through this classification, new functionally distinct cardiac macrophage populations have been elucidated. The adult heart contains three distinct populations of macrophages to the expression of C-C chemokine receptor type 2 (CCR2) and major histocompatibility complex class II (MHC-II): CCR2-MHC-IIlow, CCR2-MHC-IIhigh, and CCR2+ MHC-IIhigh.
CCR2-MHC-IIlow and CCR2-MHC-IIhigh, are long-lived, derive from embryonic progenitors maintained through the proliferation of local macrophages in the heart, and represent the vast majority at steady state [5]. They show an enhanced capacity to phagocyte death cardiomyocytes and exhibit a low inflammatory profile [40][41], while CCR2-MHC-IIhigh can, in vitro, elicit an important inflammation response, for example through antigen-presenting cells (APC) to T-cells [41][42]. CCR2+ MHC-IIhigh are relatively short-lived and derive exclusively from circulating monocytes [5]. At a steady state, their function is still unclear. Following an MI, however, the presence of mitochondrial deoxyribonucleic acid (DNA) and alarmins from dying cardiomyocytes activates a vast array of proinflammatory genes in this subpopulation of macrophages, for example in the NLRP3 pathway, involved in neutrophil-associated inflammatory response, as previously stated [43]. In the first acute response to ischemic injury (first 4–7 days) CCR2-MHC-IIlow and CCR2-MHC-IIhigh continue to show their classical phagocytic, non-inflammatory function in the lesion area [25][44]. This triggers apoptosis of all resident macrophages and by 24 h post-MI they are almost completely absent [33]. At the same time, abundant blood monocytes infiltrate the lesion area and differentiate into pro-inflammatory CCR2+ MHC-IIhigh [33]. The fact that inhibition of monocyte extravasation into the cardiac tissue decreases macrophage numbers and improves cardiac physiology, highlights the importance of this population of macrophages in the adverse post-MI response [45]. In the reparative phase (days 5–14), monocytes differentiate into CCR2-MHC-IIhigh macrophages as opposed to CCR2+ MHC-IIhigh [46]. This switch in polarisation seems to be determined by changes in the local ischemic region: the infarct microenvironment is initially filled with early pro-M1 mediators, like interferon-γ (IFN-γ) and the granulocyte-macrophage colony-stimulating factor (GM-CSF), which trigger the initial differentiation into CCR2+ MHC-IIhigh macrophages and later with pro-M2 factors, like interleukin (IL) 10 and transform growth factor- β (TGF-β), stimulating differentiation into CCR2-MHC-IIhigh macrophages [47][48]. These promote angiogenesis and scar formation and regulate the ECM microenvironment [46][49][50], orchestrating the fine mechanisms leading to tissue modelling and healing.
The mechanistic aspects of this flexible macrophage polarization are still poorly understood [51]. In the past 5 years, several studies have suggested an extensive epigenetic and transcriptional crosstalk between pro-inflammatory and anti-inflammatory signaling [52][53]. Responding to local stimuli, macrophages not only react at a transcriptional level [54], mounting the real-time response but they also adopt unique and permissive epigenetic changes, creating a cellular memory [55]. This memory enables the cells to launch a faster response upon reactivation, changing the macrophage activation state [51]. This allows a potentially more efficient response to pathological stimuli [56][57] but makes the system also prone to the dysregulation responsible for the clinical disease [51][58]. The tight control by transcription factors and epigenetic modifiers makes these pathways in macrophages a promising therapeutic target for inflammation-driven diseases.

2.3. Natural Killer T Lymphocytes: New Actors for an Old Disease?

Natural Killer T (NKT) cells have been the subject of increasing research about their role in the immune response to atherosclerosis and CAD [59]. NKT cells can be broadly categorized into two main subsets: Type I (invariant) and Type II NKT cells; type I NKT cells are the most well-studied subset and are characterized by their invariant T-cell receptor (TCR) alpha chain combined with a limited set of beta chains. These cells recognize glycolipid antigens such as α-galactosylceramide; they can rapidly produce a wide spectrum of cytokines, making them versatile regulators of immune responses. Type II NKT cells are less well-defined than Type I NKT cells and exhibit more distinct TCRs. They recognize a broader range of lipid antigens, including sulfatides, phospholipids, and glycolipids. Their functions are less clear, but they may also influence immune responses in various contexts [60][61]. The activation of these lipid-reactive NKT cells involves the interaction between lipid antigens, both endogenous and exogenous, and the nonclassical major histocompatibility complex class I (MHC-I) molecules of the cluster of differentiation (CD) 1 family on APC [62]. These lipid antigens bind to specific TCRs on T-cell subsets, including NKT cells. The structures of molecules in the CD1 family have been studied to understand how these lipid antigens associate with them. Group I CD1 molecules present lipid antigens from microbes and self-lipids to T cells, while group II CD1 molecules, specifically CD1d, present lipids to NKT cells. These CD1-presenting molecules are found on APC-like dendritic cells (DC), macrophages, and B cells, all of which play a role in the development of atherosclerotic lesions [63][64].
The activation of NKT cells can occur through various pathways. When an antigen is presented by CD1d molecules, a subset of NKT cells called invariant NKT cells (iNKT) respond rapidly by releasing cytokines, like helper T cells [65][66]. These cytokines have the potential to influence the development of atherosclerotic lesions in multiple ways [67][68]. The cytokines secreted by activated iNKT cells within the lesion may affect the response of other cells involved in the immune system, both innate and adaptive. In addition to antigen presentation by CD1d molecules, NKT cells can also be activated through a CD1d-independent pathway: DC or macrophages can be activated by Toll-like receptor (TLR) ligands, which then produce cytokines like IL-12, IL-18, or type I interferons. These cytokines can activate NKT cells without the involvement of CD1d molecules [69]. Moreover, TLR2 and TLR4 activation have been linked with atherosclerosis [70][71]. The effects of iNKT cells have been observed in studies conducted on mice. These studies administer different diets to mice to examine the impact of varying levels of iNKT cell activation or quantity. For instance, in mouse models treated with a Western-type diet, increased iNKT cell activity leads to increased plaque formation on the aortic root [72]. Conversely, in mice with genetically iNKT-deficient cells, an opposite effect on atherosclerotic lesions was shown [73].
NKT cells therefore play a crucial role in the immune response to CAD and atherosclerosis. The activation of these cells can occur through various pathways, including antigen presentation by CD1d molecules and CD1d-independent pathways. The cytokines released by activated NKT cells can influence the development of atherosclerotic lesions and modulate the response of other immune cells. Further research is required to fully understand the complex mechanisms by which NKT cells contribute to these diseases.

2.4. B Lymphocytes

B cells play a multifaceted role in atherosclerosis, depending on cellular differentiation: this process leads to subtypes B1 and B2 cell formation [74][75][76]. When naive B cells are exposed to a complex set of stimuli, they undergo differentiation and become antibody-secreting cells, specifically plasma blasts and plasma cells. In dyslipidemia, activated endothelium lining atherosclerotic plaques allow different immunoglobulins to enter the plaque area [77]. B1 cells are associated with producing antibodies, including IgM antibodies, with anti-inflammatory properties [78]. B1 cells may have a protective role in atherosclerosis by reducing inflammation and promoting plaque stability through their antibody production [79][80][81]. Contrary, B2 cells were initially viewed to be proatherogenic after preferential B2-cell depletion using CD20-targeted antibodies [82][83]. However, recent studies provide marginal zone B cells can conduct protective effects potentially secreting IgM [84].
In atherosclerosis, B cells produce antibodies directed against specific antigens present within the plaques. The most well-studied antibody target is ox-LDL. When LDL cholesterol particles become oxidized, they produce Oxidation-Specific Epitopes (OSEs) that can be recognized by the immune system. Anti-ox-LDL antibodies can promote inflammation and contribute to plaque formation by facilitating the uptake of ox-LDL by macrophages, leading to the formation of foam cells [85][86][87]. Advanced stages of plaque formation give rise to artery tertiary lymphoid organs, such as those found in the adventitia, where plasma cells are formed within the plaque. This leads to the production of immunoglobulins in the adventitia. To support this, atherosclerotic plaques contain immunoglobulins specific to different OSEs [88]. A substantial number of IgM antibodies in our immune system can identify OSEs [89]. These epitopes can be found on ox-LDL, apoptotic cells, and microvesicles. They also inhibit the pro-inflammatory responses of macrophages triggered by microvesicles [90]. Additionally, when macrophages are triggered by microvesicles, the IgM antibodies also play a role in reducing the pro-inflammatory responses of the macrophages.
In contrast, IgG antibodies form immune complexes with ox-LDL, promoting inflammatory responses by macrophages [91]. IgE antibodies are known to have proatherogenic properties, as they stimulate macrophages and mast cells in both the plaque and perivascular area. Hamze et al. found that atherosclerotic plaques are rich in IgA and IgG, secreted by B cells during the inflammation process [92]. The involvement of IgA antibodies in atherosclerosis is still not well understood: a positive association between IgA and cardiovascular (CV) outcomes is reported, but functional roles have yet to be investigated [93].
B cells also produce various cytokines, including proatherogenic tumor necrosis factor -α (TNF-α) and antiatherogenic interleukin-10. Several studies on mouse models have described the protective role of B cells, remodeling the atheromatic plaque and increasing the lesion in case of B cell depletion [94][95]. However, there are different functional subsets of B cells, recognizing the heterogeneous population: both the proatherogenic and the antiatherogenic activities of various subsets are described [96][97]. While traditionally thought of as primarily involved in the production of antibodies, B cells also have antibody-independent pathways that influence the development and progression of atherosclerosis: the presence of B cells was characterized in the adventitia of atherosclerotic aortas but not in the atheromatous plaque [98]; this suggests a local immune response, associated with T cells, DC, and macrophages [99].

2.5. T Lymphocytes Subsets: Signalling and Mechanisms

T cells have been found in the blood vessel walls near various CV diseases. They can contribute to immune responses in two ways: directly, by producing cytokines and molecules that promote inflammation, and indirectly through the activation of B cells. The different subsets of T cells have distinct functions in CV diseases, depending on whether they produce pro-inflammatory or anti-inflammatory molecules. CD4+ T cells, when in a naive state, can be differentiated into several subsets: T helper 1 (TH1), TH2, TH17, or regulatory T (Treg) cells. TH1 cells are pro-atherogenic and act through the production of IFN-γ and TNF-α [100][101].
On the other hand, Treg cells have an anti-atherosclerotic effect by secreting IL-10 and transforming growth factor-β (TGF-β). In fact, studies have shown that IL-10 produced by Treg cells can slow down the progression of abdominal aortic aneurysm and the formation of artery blockages following angioplasty [101][102][103][104].
TH2 cells secrete molecules such as IL-4, IL-5, and IL-13. While TH2 cells and IL-4 may be associated with advanced atherosclerosis in mice lacking the apolipoprotein E (ApoE) gene [105], atherosclerosis decreased in mice lacking both the IL4 and Ldlr or IL4 and ApoE genes [106][107].
Lastly, there is inconsistent contrasting evidence on the impact of TH17 cells in atherosclerosis. TH17 cells produce cytokines like IL-17A and IL-17F. Blocking or inhibiting IL-17 in mice lacking the ApoE gene was found to promote the development of atherosclerosis [108][109][110]. However, mice lacking the ILl17a gene actually showed an accelerated formation of unstable atherosclerotic lesions compared to mice lacking only the ApoE gene [111].
Assorted studies suggest that CD4+ T-cells are crucially involved in left-ventricular (LV) remodeling during both ischemic [112] and non-ischemic [113] heart failure. Mice studies show activation of CD4+ T-cells post-MI is a controlled response designed to subside rapidly with scar formation to achieve complete immune resolution within 2 weeks post-MI. HF, on the other hand, is associated with a second wave of CD4+ T-cell activation, and their transmigration into the heart promotes LV remodeling, end-diastolic volume and end-systolic volume increasing, ejection fraction reduction, and progressive cardiac dysfunction [114][115].
CD8+ T cells play a role in the development of atherosclerosis. When these cells are activated, they release cytotoxins, perforin, and granzymes. The cytotoxins can induce programmed cell death, or apoptosis, in macrophages, vascular smooth muscle cells (VSMCs), and ECs. This contributes to the formation of vulnerable atherosclerotic lesions, which are areas of plaque that can rupture and lead to complications [116]. Furthermore, the absence of programmed cell death ligand (PDL)-1 and PDL-2 in mice has been shown to increase the development of atherosclerotic lesions in the aorta. It also leads to an increase in the numbers of CD4+ T cells and CD8+ T cells, suggesting that these cells are more involved in atherosclerosis when these molecules are lacking [117].
Moreover, T Lymphocytes crosstalk with other molecules, such as cyclophilins, have a new role in CAD: these proteins are released into the extracellular space in response to inflammatory stimuli. Gegunde et al. described the involvement of a cell surface receptor for extracellular cyclophilins in CAD, the CD147 receptor: patients with CAD had considerably higher levels of membrane expression of CD147, cyclophilin A, B, and C in T lymphocytes purified from these subjects, as well as pro-inflammatory interleukins [118].

3. Trained Immunity in Atherosclerosis: A New Proposal for a New Direction

When briefly exposed to certain stimuli, cells of the innate immune system such as monocytes, macrophages, DC, and NKT cells can develop a phenotype resembling immunologic memory, termed trained immunity [119]. Upon restimulation, trained cells manifest a long-term proinflammatory phenotype with an increased cytokine release, nonspecific with respect to the original stimulus. The persistent overactivation of these trained cells could contribute to the incessant vascular wall inflammation, a peculiar characteristic of atherosclerosis [120].
Previous works on trained immunity involved microorganisms and microbial products including the Bacillus Calmette–Guerin (BCG) vaccine, Candida albicansi, and its cell wall component β-glucan [121].
Exposure to these pathogens provokes an increased production of proinflammatory cytokines and chemokines in trained cells as a response to a secondary insult, even if different from the initial one.
It was later recognized that also endogenous, self-derived molecules such as ox-LDL, catecholamines, uric acid, and aldosterone can induce a persistent functional reprogramming of innate immune cells [122][123][124][125][126].

3.1. Trained Immunity in Infectious Disease

From an evolutionary perspective, trained immunity confers protection to the host against subsequent infection, although it can also be responsible for a maladaptive state. An exogenous stimulus such as BCG vaccination protects against lethal systemic C albicans infection in immunodeficient mice that lack adaptive immunity [127], as well as administration of β-glucan in mice confers protection against recurrent infections [128][129]. Similar evidence also exists in humans, with the profound decrease in infant mortality rates following BCG vaccination, not solely explained by the protection against tuberculosis [130].

3.2. Trained Immunity in Chronic Inflammatory Disease

Diversely from the protective effect against infections, trained immunity might be maladaptive in chronic inflammatory diseases in which innate immunity cells play a pivotal role in the pathophysiology. The detrimental effects of trained immunity are implicated in atherosclerosis, gout, neurodegenerative disorders, and transplant rejection [131][132][133], and in other inflammatory diseases such as rheumatoid arthritis and systemic lupus erythematosus [134][135].
Trained immunity could also be one of the mechanisms contributing to the epidemiological association between infectious burden and atherosclerotic CV diseases [136].
Different endogenous, nonmicrobial atherogenic stimuli have been recognized to induce trained immunity, such as ox-LDL and lipoprotein(a) (Lp(a)), but also catecholamines and high glucose concentration [123][137][138][139]. ox-LDL has a key role in atherogenic plaque formation thanks to its ability to activate immune cells and trigger foam cell formation. When exposed to a low concentration of ox-LDL and restimulated with a TLR agonist, macrophages produce higher quantities of atherogenic cytokines and chemokines, such as IL-6, MCP1 (monocyte chemoattractant protein 1), and TNF-α. Foam cell formation is similarly enhanced after exposure to ox-LDL, due to the overexpression of scavenger receptor-A (SR-A) and CD36 and downregulation of cholesterol efflux transporters adenosine triphosphate-binding cassette transporter-A1 and G1 [123][140].
Lp(a) is the main circulating carrier of oxidized phospholipids and plays a key role in atherogenesis [141][142]. Monocytes incubated with Lp(a) for 24 h show an increased proinflammatory cytokine production compared to untrained controls [143]. Monocytes isolated from patients with elevated Lp(a) levels also show a stronger trans-endothelial migration.
A pivotal role in diabetes and CV diseases is played by diet, and atherosclerosis-prone knock-out for LDL receptor mice display characteristics of trained immunity when fed a Western-type diet [144].
The characteristic increase in cytokine production in trained immunity has also been observed in patients with already established coronary atherosclerosis, familial hypercholesterolemia, and in patients with cerebral small vessel disease [145][146]. Similarly, hyperuricemia can induce long-term proinflammatory activation of innate immune cells [122][125].

References

  1. Libby, P. The Changing Landscape of Atherosclerosis. Nature 2021, 592, 524–533.
  2. Libby, P. Inflammation in Atherosclerosis. Arter. Thromb. Vasc. Biol. 2012, 32, 2045–2051.
  3. Hansson, G.K. Inflammation, Atherosclerosis, and Coronary Artery Disease. N. Engl. J. Med. 2005, 352, 1685–1695.
  4. Sreejit, G.; Johnson, J.; Jaggers, R.M.; Dahdah, A.; Murphy, A.J.; Hanssen, N.M.J.; Nagareddy, P.R. Neutrophils in Cardiovascular Disease: Warmongers, Peacemakers, or Both? Cardiovasc. Res. 2022, 118, 2596–2609.
  5. Lavine, K.J.; Pinto, A.R.; Epelman, S.; Kopecky, B.J.; Clemente-Casares, X.; Godwin, J.; Rosenthal, N.; Kovacic, J.C. The Macrophage in Cardiac Homeostasis and Disease: JACC Macrophage in CVD Series (Part 4). J. Am. Coll. Cardiol. 2018, 72, 2213–2230.
  6. Simons, K.H.; de Jong, A.; Jukema, J.W.; de Vries, M.R.; Arens, R.; Quax, P.H.A. T Cell Co-Stimulation and Co-Inhibition in Cardiovascular Disease: A Double-Edged Sword. Nat. Rev. Cardiol. 2019, 16, 325–343.
  7. Porsch, F.; Mallat, Z.; Binder, C.J. Humoral Immunity in Atherosclerosis and Myocardial Infarction: From B Cells to Antibodies. Cardiovasc. Res. 2021, 117, 2544–2562.
  8. Ley, K.; Hoffman, H.M.; Kubes, P.; Cassatella, M.A.; Zychlinsky, A.; Hedrick, C.C.; Catz, S.D. Neutrophils: New Insights and Open Questions. Sci. Immunol. 2018, 3, eaat4579.
  9. Silvestre-Roig, C.; Braster, Q.; Wichapong, K.; Lee, E.Y.; Teulon, J.M.; Berrebeh, N.; Winter, J.; Adrover, J.M.; Santos, G.S.; Froese, A.; et al. Externalized Histone H4 Orchestrates Chronic Inflammation by Inducing Lytic Cell Death. Nature 2019, 569, 236–240.
  10. Joseph, J.P.; Reyes, E.; Guzman, J.; O’Doherty, J.; McConkey, H.; Arri, S.; Kakkar, R.; Beckley, N.; Douiri, A.; Barrington, S.F.; et al. CXCR2 Inhibition—A Novel Approach to Treating CoronAry Heart DiseAse (CICADA): Study Protocol for a Randomised Controlled Trial. Trials 2017, 18, 473.
  11. Chia, S.; Nagurney, J.T.; Brown, D.F.M.; Raffel, O.C.; Bamberg, F.; Senatore, F.; Wackers, F.J.T.; Jang, I.K. Association of Leukocyte and Neutrophil Counts with Infarct Size, Left Ventricular Function and Outcomes after Percutaneous Coronary Intervention for ST-Elevation Myocardial Infarction. Am. J. Cardiol. 2009, 103, 333–337.
  12. Guasti, L.; Dentali, F.; Castiglioni, L.; Maroni, L.; Marino, F.; Squizzato, A.; Ageno, W.; Gianni, M.; Gaudio, G.; Grandi, A.M.; et al. Neutrophils and Clinical Outcomes in Patients with Acute Coronary Syndromes and/or Cardiac Revascularization: A Systematic Review on More than 34,000 Subjects. Thromb. Haemost. 2011, 106, 591–599.
  13. Verdoia, M.; Nardin, M.; Gioscia, R.; Negro, F.; Marcolongo, M.; Suryapranata, H.; Kedhi, E.; De Luca, G. Higher Neutrophil-to-Lymphocyte Ratio (NLR) Increases the Risk of Suboptimal Platelet Inhibition and Major Cardiovascular Ischemic Events among ACS Patients Receiving Dual Antiplatelet Therapy with Ticagrelor. Vasc. Pharmacol. 2020, 132, 106765.
  14. Prabhu, S.D.; Frangogiannis, N.G. The Biological Basis for Cardiac Repair after Myocardial Infarction. Circ. Res. 2016, 119, 91–112.
  15. Sreejit, G.; Abdel Latif, A.; Murphy, A.J.; Nagareddy, P.R. Emerging Roles of Neutrophil-Borne S100A8/A9 in Cardiovascular Inflammation. Pharmacol. Res. 2020, 161, 105212.
  16. Ortega-Gomez, A.; Salvermoser, M.; Rossaint, J.; Pick, R.; Brauner, J.; Lemnitzer, P.; Tilgner, J.; De Jong, R.J.; Megens, R.T.A.; Jamasbi, J.; et al. Cathepsin G Controls Arterial but Not Venular Myeloid Cell Recruitment. Circulation 2016, 134, 1176–1188.
  17. Zarbock, A.; Ley, K. Mechanisms and Consequences of Neutrophil Interaction with the Endothelium. Am. J. Pathol. 2008, 172, 1–7.
  18. Rasmuson, J.; Kenne, E.; Wahlgren, M.; Soehnlein, O.; Lindbom, L. Heparinoid Sevuparin Inhibits Streptococcus-Induced Vascular Leak through Neutralizing Neutrophil-Derived Proteins. FASEB J. 2019, 33, 10443–10452.
  19. Mawhin, M.-A.; Tilly, P.; Zirka, G.; Charles, A.-L.; Slimani, F.; Vonesch, J.-L.; Michel, J.-B.; Bäck, M.; Norel, X.; Fabre, J.-E. Neutrophils Recruited by Leukotriene B4 Induce Features of Plaque Destabilization during Endotoxaemia. Cardiovasc. Res. 2018, 114, 1656–1666.
  20. Arruda-Olson, A.M.; Reeder, G.S.; Bell, M.R.; Weston, S.A.; Roger, V.L. Neutrophilia Predicts Death and Heart Failure after Myocardial Infarction: A Community-Based Study. Circ. Cardiovasc. Qual. Outcomes 2009, 2, 656–662.
  21. Sreejit, G.; Abdel-Latif, A.; Athmanathan, B.; Annabathula, R.; Dhyani, A.; Noothi, S.K.; Quaife-Ryan, G.A.; Al-Sharea, A.; Pernes, G.; Dragoljevic, D.; et al. Neutrophil-Derived S100A8/A9 Amplify Granulopoiesis after Myocardial Infarction. Circulation 2020, 141, 1080–1094.
  22. Nagareddy, P.R.; Sreejit, G.; Abo-Aly, M.; Jaggers, R.M.; Chelvarajan, L.; Johnson, J.; Pernes, G.; Athmanathan, B.; Abdel-Latif, A.; Murphy, A.J. NETosis Is Required for S100A8/A9-Induced Granulopoiesis after Myocardial Infarction. Arter. Thromb. Vasc. Biol. 2020, 40, 2805–2807.
  23. Vasilyev, N.; Williams, T.; Brennan, M.L.; Unzek, S.; Zhou, X.; Heinecke, J.W.; Spitz, D.R.; Topol, E.J.; Hazen, S.L.; Pen, M.S. Myeloperoxidase-Generated Oxidants Modulate Left Ventricular Remodeling but Not Infarct Size after Myocardial Infarction. Circulation 2005, 112, 2812–2820.
  24. Marinković, G.; Koenis, D.S.; De Camp, L.; Jablonowski, R.; Graber, N.; De Waard, V.; De Vries, C.J.; Goncalves, I.; Nilsson, J.; Jovinge, S.; et al. S100A9 Links Inflammation and Repair in Myocardial Infarction. Circ. Res. 2020, 127, 664–676.
  25. Ma, Y.; Yabluchanskiy, A.; Iyer, R.P.; Cannon, P.L.; Flynn, E.R.; Jung, M.; Henry, J.; Cates, C.A.; Deleon-Pennell, K.Y.; Lindsey, M.L. Temporal Neutrophil Polarization Following Myocardial Infarction. Cardiovasc. Res. 2016, 110, 51–61.
  26. Lörchner, H.; Pöling, J.; Gajawada, P.; Hou, Y.; Polyakova, V.; Kostin, S.; Adrian-Segarra, J.M.; Boettger, T.; Wietelmann, A.; Warnecke, H.; et al. Myocardial Healing Requires Reg3β-Dependent Accumulation of Macrophages in the Ischemic Heart. Nat. Med. 2015, 21, 353–362.
  27. Massena, S.; Christoffersson, G.; Vågesjö, E.; Seignez, C.; Gustafsson, K.; Binet, F.; Hidalgo, C.H.; Giraud, A.; Lomei, J.; Weström, S.; et al. Identification and Characterization of VEGF-A-Responsive Neutrophils Expressing CD49d, VEGFR1, and CXCR4 in Mice and Humans. Blood 2015, 126, 2016–2026.
  28. Wang, J.; Hossain, M.; Thanabalasuriar, A.; Gunzer, M.; Meininger, C.; Kubes, P. Visualizing the Function and Fate of Neutrophils in Sterile Injury and Repair. Science 2017, 358, 111–116.
  29. García-Prieto, J.; Villena-Gutiérrez, R.; Gómez, M.; Bernardo, E.; Pun-García, A.; García-Lunar, I.; Crainiciuc, G.; Fernández-Jiménez, R.; Sreeramkumar, V.; Bourio-Martínez, R.; et al. Neutrophil Stunning by Metoprolol Reduces Infarct Size. Nat. Commun. 2017, 8, 14780.
  30. Marinković, G.; Grauen Larsen, H.; Yndigegn, T.; Szabo, I.A.; Mares, R.G.; De Camp, L.; Weiland, M.; Tomas, L.; Goncalves, I.; Nilsson, J.; et al. Inhibition of Pro-Inflammatory Myeloid Cell Responses by Short-Term S100A9 Blockade Improves Cardiac Function after Myocardial Infarction. Eur. Heart J. 2019, 40, 2713–2723.
  31. Horckmans, M.; Ring, L.; Duchene, J.; Santovito, D.; Schloss, M.J.; Drechsler, M.; Weber, C.; Soehnlein, O.; Steffens, S. Neutrophils Orchestrate Post-Myocardial Infarction Healing by Polarizing Macrophages towards a Reparative Phenotype. Eur. Heart J. 2017, 38, 187–197.
  32. Ma, Y.; Mouton, A.J.; Lindsey, M.L. Cardiac Macrophage Biology in the Steady-State Heart, the Aging Heart, and Following Myocardial Infarction. Transl. Res. 2018, 191, 15–28.
  33. Heidt, T.; Courties, G.; Dutta, P.; Sager, H.B.; Sebas, M.; Iwamoto, Y.; Sun, Y.; Da Silva, N.; Panizzi, P.; Van Der Lahn, A.M.; et al. Differential Contribution of Monocytes to Heart Macrophages in Steady-State and after Myocardial Infarction. Circ. Res. 2014, 115, 284–295.
  34. Glaros, T. Macrophages and Fibroblasts during Inflammation, Tissue Damage and Organ Injury. Front. Biosci. 2009, 14, 3988–3993.
  35. Frantz, S.; Nahrendorf, M. Cardiac Macrophages and Their Role in Ischaemic Heart Disease. Cardiovasc. Res. 2014, 102, 240–248.
  36. Nahrendorf, M.; Swirski, F.K.; Aikawa, E.; Stangenberg, L.; Wurdinger, T.; Figueiredo, J.L.; Libby, P.; Weissleder, R.; Pittet, M.J. The Healing Myocardium Sequentially Mobilizes Two Monocyte Subsets with Divergent and Complementary Functions. J. Exp. Med. 2007, 204, 3037–3047.
  37. Nahrendorf, M.; Swirski, F.K. Abandoning M1/M2 for a Network Model of Macrophage Function. Circ. Res. 2016, 119, 414–417.
  38. Xue, J.; Schmidt, S.V.; Sander, J.; Draffehn, A.; Krebs, W.; Quester, I.; DeNardo, D.; Gohel, T.D.; Emde, M.; Schmidleithner, L.; et al. Transcriptome-Based Network Analysis Reveals a Spectrum Model of Human Macrophage Activation. Immunity 2014, 40, 274–288.
  39. Cochain, C.; Vafadarnejad, E.; Arampatzi, P.; Pelisek, J.; Winkels, H.; Ley, K.; Wolf, D.; Saliba, A.E.; Zernecke, A. Single-Cell RNA-Seq Reveals the Transcriptional Landscape and Heterogeneity of Aortic Macrophages in Murine Atherosclerosis. Circ. Res. 2018, 122, 1661–1674.
  40. DeBerge, M.; Yeap, X.Y.; Dehn, S.; Zhang, S.; Grigoryeva, L.; Misener, S.; Procissi, D.; Zhou, X.; Lee, D.C.; Muller, W.A.; et al. MerTK Cleavage on Resident Cardiac Macrophages Compromises Repair after Myocardial Ischemia Reperfusion Injury. Circ. Res. 2017, 121, 930–940.
  41. Epelman, S.; Lavine, K.J.; Beaudin, A.E.; Sojka, D.K.; Carrero, J.A.; Calderon, B.; Brija, T.; Gautier, E.L.; Ivanov, S.; Satpathy, A.T.; et al. Embryonic and Adult-Derived Resident Cardiac Macrophages Are Maintained through Distinct Mechanisms at Steady State and during Inflammation. Immunity 2014, 40, 91–104.
  42. Epelman, S.; Lavine, K.J.; Randolph, G.J. Origin and Functions of Tissue Macrophages. Immunity 2014, 41, 21–35.
  43. Li, W.; Hsiao, H.; Higashikubo, R.; Saunders, B.T.; Bharat, A.; Goldstein, D.R.; Krupnick, A.S.; Gelman, A.E.; Lavine, K.J.; Kreisel, D. Heart-Resident CCR2+ Macrophages Promote Neutrophil Extravasation through TLR9/MyD88/CXCL5 Signaling. JCI Insight 2016, 1, e87315.
  44. Jung, K.; Kim, P.; Leuschner, F.; Gorbatov, R.; Kim, J.K.; Ueno, T.; Nahrendorf, M.; Yun, S.H. Endoscopic Time-Lapse Imaging of Immune Cells in Infarcted Mouse Hearts. Circ. Res. 2013, 112, 891–899.
  45. Sager, H.B.; Hulsmans, M.; Lavine, K.J.; Moreira, M.B.; Heidt, T.; Courties, G.; Sun, Y.; Iwamoto, Y.; Tricot, B.; Khan, O.F.; et al. Proliferation and Recruitment Contribute to Myocardial Macrophage Expansion in Chronic Heart Failure. Circ. Res. 2016, 119, 853–864.
  46. Hilgendorf, I.; Gerhardt, L.M.S.; Tan, T.C.; Winter, C.; Holderried, T.A.W.; Chousterman, B.G.; Iwamoto, Y.; Liao, R.; Zirlik, A.; Scherer-Crosbie, M.; et al. Ly-6 Chigh Monocytes Depend on Nr4a1 to Balance Both Inflammatory and Reparative Phases in the Infarcted Myocardium. Circ. Res. 2014, 114, 1611–1622.
  47. Christia, P.; Bujak, M.; Gonzalez-Quesada, C.; Chen, W.; Dobaczewski, M.; Reddy, A.; Frangogiannis, N.G. Systematic Characterization of Myocardial Inflammation, Repair, and Remodeling in a Mouse Model of Reperfused Myocardial Infarction. J. Histochem. Cytochem. 2013, 61, 555–570.
  48. Jung, M.; Ma, Y.; Iyer, R.P.; DeLeon-Pennell, K.Y.; Yabluchanskiy, A.; Garrett, M.R.; Lindsey, M.L. IL-10 Improves Cardiac Remodeling after Myocardial Infarction by Stimulating M2 Macrophage Polarization and Fibroblast Activation. Basic Res. Cardiol. 2017, 112, 33.
  49. Godwin, J.; Kuraitis, D.; Rosenthal, N. Extracellular Matrix Considerations for Scar-Free Repair and Regeneration: Insights from Regenerative Diversity among Vertebrates. Int. J. Biochem. Cell Biol. 2014, 56, 47–55.
  50. Mescher, A.L. Macrophages and Fibroblasts during Inflammation and Tissue Repair in Models of Organ Regeneration. Regeneration 2017, 4, 39–53.
  51. Kuznetsova, T.; Prange, K.H.M.; Glass, C.K.; de Winther, M.P.J. Transcriptional and Epigenetic Regulation of Macrophages in Atherosclerosis. Nat. Rev. Cardiol. 2020, 17, 216–228.
  52. Piccolo, V.; Curina, A.; Genua, M.; Ghisletti, S.; Simonatto, M.; Sabò, A.; Amati, B.; Ostuni, R.; Natoli, G. Opposing Macrophage Polarization Programs Show Extensive Epigenomic and Transcriptional Cross-Talk. Nat. Immunol. 2017, 18, 530–540.
  53. Czimmerer, Z.; Daniel, B.; Horvath, A.; Rückerl, D.; Nagy, G.; Kiss, M.; Peloquin, M.; Budai, M.M.; Cuaranta-Monroy, I.; Simandi, Z.; et al. The Transcription Factor STAT6 Mediates Direct Repression of Inflammatory Enhancers and Limits Activation of Alternatively Polarized Macrophages. Immunity 2018, 48, 75–90.e6.
  54. Li, Z.; Martin, M.; Zhang, J.; Huang, H.Y.; Bai, L.; Zhang, J.; Kang, J.; He, M.; Li, J.; Maurya, M.R.; et al. Krüppel-like Factor 4 Regulation of Cholesterol-25-Hydroxylase and Liver X Receptor Mitigates Atherosclerosis Susceptibility. Circulation 2017, 136, 1315–1320.
  55. Ostuni, R.; Piccolo, V.; Barozzi, I.; Polletti, S.; Termanini, A.; Bonifacio, S.; Curina, A.; Prosperini, E.; Ghisletti, S.; Natoli, G. Latent Enhancers Activated by Stimulation in Differentiated Cells. Cell 2013, 152, 157–171.
  56. Novakovic, B.; Habibi, E.; Wang, S.Y.; Arts, R.J.W.; Davar, R.; Megchelenbrink, W.; Kim, B.; Kuznetsova, T.; Kox, M.; Zwaag, J.; et al. β-Glucan Reverses the Epigenetic State of LPS-Induced Immunological Tolerance. Cell 2016, 167, 1354–1368.e14.
  57. Foster, S.L.; Hargreaves, D.C.; Medzhitov, R. Gene-Specific Control of Inflammation by TLR-Induced Chromatin Modifications. Nature 2007, 447, 972–978.
  58. Cheng, S.C.; Quintin, J.; Cramer, R.A.; Shepardson, K.M.; Saeed, S.; Kumar, V.; Giamarellos-Bourboulis, E.J.; Martens, J.H.A.; Rao, N.A.; Aghajanirefah, A.; et al. MTOR- and HIF-1α-Mediated Aerobic Glycolysis as Metabolic Basis for Trained Immunity. Science 2014, 345, 1250684.
  59. Getz, G.S.; Reardon, C.A. Natural Killer T Cells in Atherosclerosis. Nat. Rev. Cardiol. 2017, 14, 304–314.
  60. Liao, C.M.; Zimmer, M.I.; Wang, C.R. The Functions of Type I and Type II Natural Killer T Cells in Inflammatory Bowel Diseases. Inflamm. Bowel Dis. 2013, 19, 1330–1338.
  61. Constantinides, M.G.; Bendelac, A. Transcriptional Regulation of the NKT Cell Lineage. Curr. Opin. Immunol. 2013, 25, 161–167.
  62. Anderson, B.L.; Teyton, L.; Bendelac, A.; Savage, P.B. Stimulation of Natural Killer T Cells by Glycolipids. Molecules 2013, 18, 15662–15688.
  63. Hansson, G.K.; Hermansson, A. The Immune System in Atherosclerosis. Nat. Immunol. 2011, 12, 204–212.
  64. Yanaba, K.; Bouaziz, J.D.; Haas, K.M.; Poe, J.C.; Fujimoto, M.; Tedder, T.F. A Regulatory B Cell Subset with a Unique CD1dhiCD5+ Phenotype Controls T Cell-Dependent Inflammatory Responses. Immunity 2008, 28, 639–650.
  65. Tupin, E.; Kinjo, Y.; Kronenberg, M. The Unique Role of Natural Killer T Cells in the Response to Microorganisms. Nat. Rev. Microbiol. 2007, 5, 405–417.
  66. Van Kaer, L.; Parekh, V.V.; Wu, L. Invariant Natural Killer T Cells: Bridging Innate and Adaptive Immunity. Cell Tissue Res. 2011, 343, 43–55.
  67. Ait-Oufella, H.; Taleb, S.; Mallat, Z.; Tedgui, A. Recent Advances on the Role of Cytokines in Atherosclerosis. Arter. Thromb. Vasc. Biol. 2011, 31, 969–979.
  68. Kleemann, R.; Zadelaar, S.; Kooistra, T. Cytokines and Atherosclerosis: A Comprehensive Review of Studies in Mice. Cardiovasc. Res. 2008, 79, 360–376.
  69. Monteiro, M.; Almeida, C.F.; Caridade, M.; Ribot, J.C.; Duarte, J.; Agua-Doce, A.; Wollenberg, I.; Silva-Santos, B.; Graca, L. Identification of Regulatory Foxp3+ Invariant NKT Cells Induced by TGF-β. J. Immunol. 2010, 185, 2157–2163.
  70. Choi, S.H.; Sviridov, D.; Miller, Y.I. Oxidized Cholesteryl Esters and Inflammation. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2017, 1862, 393–397.
  71. Curtiss, L.K.; Tobias, P.S. Emerging Role of Toll-like Receptors in Atherosclerosis. J. Lipid Res. 2009, 50, S340–S345.
  72. VanderLaan, P.A.; Reardon, C.A.; Sagiv, Y.; Blachowicz, L.; Lukens, J.; Nissenbaum, M.; Wang, C.R.; Getz, G.S. Characterization of the Natural Killer T-Cell Response in an Adoptive Transfer Model of Atherosclerosis. Am. J. Pathol. 2007, 170, 1100–1107.
  73. Nakai, Y.; Iwabuchi, K.; Fujii, S.; Ishimori, N.; Dashtsoodol, N.; Watano, K.; Mishima, T.; Iwabuchi, C.; Tanaka, S.; Bezbradica, J.S.; et al. Natural Killer T Cells Accelerate Atherogenesis in Mice. Blood 2004, 104, 2051–2059.
  74. Upadhye, A.; Srikakulapu, P.; Gonen, A.; Hendrikx, S.; Perry, H.M.; Nguyen, A.; McSkimming, C.; Marshall, M.A.; Garmey, J.C.; Taylor, A.M.; et al. Diversification and CXCR4-Dependent Establishment of the Bone Marrow B-1a Cell Pool Governs Atheroprotective IgM Production Linked to Human Coronary Atherosclerosis. Circ. Res. 2019, 125, E55–E70.
  75. Prohaska, T.A.; Que, X.; Diehl, C.J.; Hendrikx, S.; Chang, M.W.; Jepsen, K.; Glass, C.K.; Benner, C.; Witztum, J.L. Massively Parallel Sequencing of Peritoneal and Splenic B Cell Repertoires Highlights Unique Properties of B-1 Cell Antibodies. J. Immunol. 2018, 200, 1702–1717.
  76. Baumgarth, N. B-1 Cell Heterogeneity and the Regulation of Natural and Antigen-Induced IgM Production. Front. Immunol. 2016, 7, 324.
  77. Farias-Itao, D.S.; Pasqualucci, C.A.; Nishizawa, A.; da Silva, L.F.F.; Campos, F.M.; Bittencourt, M.S.; da Silva, K.C.S.; Leite, R.E.P.; Grinberg, L.T.; Ferretti-Rebustini, R.E.d.L.; et al. B Lymphocytes and Macrophages in the Perivascular Adipose Tissue Are Associated with Coronary Atherosclerosis: An Autopsy Study. J. Am. Heart Assoc. 2019, 8, e013793.
  78. Kyaw, T.; Tay, C.; Krishnamurthi, S.; Kanellakis, P.; Agrotis, A.; Tipping, P.; Bobik, A.; Toh, B.H. B1a B Lymphocytes Are Atheroprotective by Secreting Natural IgM That Increases IgM Deposits and Reduces Necrotic Cores in Atherosclerotic Lesions. Circ. Res. 2011, 109, 830–840.
  79. Pattarabanjird, T.; Li, C.; McNamara, C. B Cells in Atherosclerosis: Mechanisms and Potential Clinical Applications. JACC Basic Transl. Sci. 2021, 6, 546–563.
  80. Adamo, L.; Rocha-Resende, C.; Mann, D.L. The Emerging Role of B Lymphocytes in Cardiovascular Disease. Annu. Rev. Immunol. 2020, 38, 99–121.
  81. Sage, A.P.; Tsiantoulas, D.; Binder, C.J.; Mallat, Z. The Role of B Cells in Atherosclerosis. Nat. Rev. Cardiol. 2019, 16, 180–196.
  82. Kyaw, T.; Tay, C.; Khan, A.; Dumouchel, V.; Cao, A.; To, K.; Kehry, M.; Dunn, R.; Agrotis, A.; Tipping, P.; et al. Conventional B2 B Cell Depletion Ameliorates Whereas Its Adoptive Transfer Aggravates Atherosclerosis. J. Immunol. 2010, 185, 4410–4419.
  83. Ait-Oufella, H.; Herbin, O.; Bouaziz, J.D.; Binder, C.J.; Uyttenhove, C.; Laurans, L.; Taleb, S.; Van Vré, E.; Esposito, B.; Vilar, J.; et al. B Cell Depletion Reduces the Development of Atherosclerosis in Mice. J. Exp. Med. 2010, 207, 1579–1587.
  84. Grasset, E.K.; Duhlin, A.; Agardh, H.E.; Ovchinnikova, O.; Hägglöf, T.; Forsell, M.N.; Paulsson-Berne, G.; Hansson, G.K.; Ketelhuth, D.F.J.; Karlsson, M.C.I. Sterile Inflammation in the Spleen during Atherosclerosis Provides Oxidation-Specific Epitopes That Induce a Protective B-Cell Response. Proc. Natl. Acad. Sci. USA 2015, 112, E2030–E2038.
  85. Friedman, P.; Hörkkö, S.; Steinberg, D.; Witztum, J.L.; Dennis, E.A. Correlation of Antiphospholipid Antibody Recognition with the Structure of Synthetic Oxidized Phospholipids. Importance of Schiff Base Formation and Aldol Condensation. J. Biol. Chem. 2002, 277, 7010–7020.
  86. Boullier, A.; Gillotte, K.L.; Hörkkö, S.; Green, S.R.; Friedman, P.; Dennis, E.A.; Witztum, J.L.; Steinberg, D.; Quehenberger, O. The Binding of Oxidized Low Density Lipoprotein to Mouse CD36 Is Mediated in Part by Oxidized Phospholipids That Are Associated with Both the Lipid and Protein Moieties of the Lipoprotein. J. Biol. Chem. 2000, 275, 9163–9169.
  87. Palinski, W.; Hörkkö, S.; Miller, E.; Steinbrecher, U.P.; Powell, H.C.; Curtiss, L.K.; Witztum, J.L. Cloning of Monoclonal Autoantibodies to Epitopes of Oxidized Lipoproteins from Apolipoprotein E-Deficient Mice: Demonstration of Epitopes of Oxidized Low Density Lipoprotein in Human Plasma. J. Clin. Investig. 1996, 98, 800–814.
  88. Yla-Herttuala, S.; Palinski, W.; Butler, S.W.; Picard, S.; Steinberg, D.; Witztum, J.L. Rabbit and Human Atherosclerotic Lesions Contain IgG That Recognizes Epitopes of Oxidized LDL. Arterioscler. Thromb. 1994, 14, 32–40.
  89. Chou, M.Y.; Fogelstrand, L.; Hartvigsen, K.; Hansen, L.F.; Woelkers, D.; Shaw, P.X.; Choi, J.; Perkmann, T.; Bäckhed, F.; Miller, Y.I.; et al. Oxidation-Specific Epitopes Are Dominant Targets of Innate Natural Antibodies in Mice and Humans. J. Clin. Investig. 2009, 119, 1335–1349.
  90. Imai, Y.; Kuba, K.; Neely, G.G.; Yaghubian-Malhami, R.; Perkmann, T.; van Loo, G.; Ermolaeva, M.; Veldhuizen, R.; Leung, Y.H.C.; Wang, H.; et al. Identification of Oxidative Stress and Toll-like Receptor 4 Signaling as a Key Pathway of Acute Lung Injury. Cell 2008, 133, 235–249.
  91. Ravandi, A.; Boekholdt, S.M.; Mallat, Z.; Talmud, P.J.; Kastelein, J.J.P.; Wareham, N.J.; Miller, E.R.; Benessiano, J.; Tedgui, A.; Witztum, J.L.; et al. Relationship of IgG and IgM Autoantibodies and Immune Complexes to Oxidized LDL with Markers of Oxidation and Inflammation and Cardiovascular Events: Results from the EPIC-Norfolk Study. J. Lipid Res. 2011, 52, 1829–1836.
  92. Hamze, M.; Desmetz, C.; Berthe, M.L.; Roger, P.; Boulle, N.; Brancherau, P.; Picard, E.; Guzman, C.; Tolza, C.; Guglielmi, P. Characterization of Resident B Cells of Vascular Walls in Human Atherosclerotic Patients. J. Immunol. 2013, 191, 3006–3016.
  93. Muscari, A.; Bozzoli, C.; Gerratana, C.; Zaca’, F.; Rovinetti, C.; Zauli, D.; La Placa, M.; Puddu, P. Association of Serum IgA and C4 with Severe Atherosclerosis. Atherosclerosis 1988, 74, 179–186.
  94. Major, A.S.; Fazio, S.; Linton, M.F. B-Lymphocyte Deficiency Increases Atherosclerosis in LDL Receptor-Null Mice. Arter. Thromb. Vasc. Biol. 2002, 22, 1892–1898.
  95. Caligiuri, G.; Nicoletti, A.; Poirierand, B.; Hansson, G.K. Protective Immunity against Atherosclerosis Carried by B Cells of Hypercholesterolemic Mice. J. Clin. Investig. 2002, 109, 745–753.
  96. Rosenfeld, S.M.; Perry, H.M.; Gonen, A.; Prohaska, T.A.; Srikakulapu, P.; Grewal, S.; Das, D.; McSkimming, C.; Taylor, A.M.; Tsimikas, S.; et al. B-1b Cells Secrete Atheroprotective IgM and Attenuate Atherosclerosis. Circ. Res. 2015, 117, e28–e39.
  97. Binder, C.J.; Hartvigsen, K.; Chang, M.K.; Miller, M.; Broide, D.; Palinski, W.; Curtiss, L.K.; Corr, M.; Witztum, J.L. IL-5 Links Adaptive and Natural Immunity Specific for Epitopes of Oxidized LDL and Protects from Atherosclerosis. J. Clin. Investig. 2004, 114, 427–437.
  98. RAMSHAW, A.L.; PARUMS, D.V. Immunohistochemical Characterization of Inflammatory Cells Associated with Advanced Atherosclerosis. Histopathology 1990, 17, 543–552.
  99. Houtkamp, M.A.; De Boer, O.J.; Van Der Loos, C.M.; Van Der Wal, A.C.; Becker, A.E. Adventitial Infiltrates Associated with Advanced Atherosclerotic Plaques: Structural Organization Suggests Generation of Local Humoral Immune Responses. J. Pathol. 2001, 193, 263–269.
  100. Engelbertsen, D.; Rattik, S.; Wigren, M.; Vallejo, J.; Marinkovic, G.; Schiopu, A.; Björkbacka, H.; Nilsson, J.; Bengtsson, E. IL-1R and MyD88 Signalling in CD4+ T Cells Promote Th17 Immunity and Atherosclerosis. Cardiovasc. Res. 2018, 114, 180–187.
  101. Tedgui, A.; Mallat, Z. Cytokines in Atherosclerosis: Pathogenic and Regulatory Pathways. Physiol. Rev. 2006, 86, 515–581.
  102. Ait-Oufella, H.; Salomon, B.L.; Potteaux, S.; Robertson, A.K.L.; Gourdy, P.; Zoll, J.; Merval, R.; Esposito, B.; Cohen, J.L.; Fisson, S.; et al. Natural Regulatory T Cells Control the Development of Atherosclerosis in Mice. Nat. Med. 2006, 12, 178–180.
  103. Adam, M.; Kooreman, N.G.; Jagger, A.; Wagenhäuser, M.U.; Mehrkens, D.; Wang, Y.; Kayama, Y.; Toyama, K.; Raaz, U.; Schellinger, I.N.; et al. Systemic Upregulation of IL-10 (Interleukin-10) Using a Nonimmunogenic Vector Reduces Growth and Rate of Dissecting Abdominal Aortic Aneurysm. Arter. Thromb. Vasc. Biol. 2018, 38, 1796–1805.
  104. Eefting, D.; Schepers, A.; De Vries, M.R.; Pires, N.M.M.; Grimbergen, J.M.; Lagerweij, T.; Nagelkerken, L.M.; Monraats, P.S.; Jukema, J.W.; van Bockel, J.H.; et al. The Effect of Interleukin-10 Knock-out and Overexpression on Neointima Formation in Hypercholesterolemic APOE*3-Leiden Mice. Atherosclerosis 2007, 193, 335–342.
  105. Zhou, X.; Paulsson, G.; Stemme, S.; Hansson, G.K. Hypercholesterolemia Is Associated with a T Helper (Th) 1/Th2 Switch of the Autoimmune Response in Atherosclerotic Apo E-Knockout Mice. J. Clin. Investig. 1998, 101, 1717–1725.
  106. King, V.L.; Szilvassy, S.J.; Daugherty, A. Interleukin-4 Deficiency Decreases Atherosclerotic Lesion Formation in a Site-Specific Manner in Female LDL Receptor-/- Mice. Arter. Thromb. Vasc. Biol. 2002, 22, 456–461.
  107. Davenport, P.; Tipping, P.G. The Role of Interleukin-4 and Interleukin-12 in the Progression of Atherosclerosis in Apolipoprotein E-Deficient Mice. Am. J. Pathol. 2003, 163, 1117–1125.
  108. Butcher, M.J.; Gjurich, B.N.; Phillips, T.; Galkina, E.V. The IL-17A/IL-17RA Axis Plays a Proatherogenic Role via the Regulation of Aortic Myeloid Cell Recruitment. Circ. Res. 2012, 110, 675–687.
  109. Smith, E.; Prasad, K.M.R.; Butcher, M.; Dobrian, A.; Kolls, J.K.; Ley, K.; Galkina, E. Blockade of Interleukin-17A Results in Reduced Atherosclerosis in Apolipoprotein E-Deficient Mice. Circulation 2010, 121, 1746–1755.
  110. Erbel, C.; Chen, L.; Bea, F.; Wangler, S.; Celik, S.; Lasitschka, F.; Wang, Y.; Böckler, D.; Katus, H.A.; Dengler, T.J. Inhibition of IL-17A Attenuates Atherosclerotic Lesion Development in ApoE-Deficient Mice. J. Immunol. 2009, 183, 8167–8175.
  111. Danzaki, K.; Matsui, Y.; Ikesue, M.; Ohta, D.; Ito, K.; Kanayama, M.; Kurotaki, D.; Morimoto, J.; Iwakura, Y.; Yagita, H.; et al. Interleukin-17A Deficiency Accelerates Unstable Atherosclerotic Plaque Formation in Apolipoprotein e-Deficient Mice. Arter. Thromb. Vasc. Biol. 2012, 32, 273–280.
  112. Santos-Zas, I.; Lemarié, J.; Tedgui, A.; Ait-Oufella, H. Adaptive Immune Responses Contribute to Post-Ischemic Cardiac Remodeling. Front. Cardiovasc. Med. 2019, 5, 198.
  113. Nevers, T.; Salvador, A.M.; Grodecki-Pena, A.; Knapp, A.; Velázquez, F.; Aronovitz, M.; Kapur, N.K.; Karas, R.H.; Blanton, R.M.; Alcaide, P. Left Ventricular T-Cell Recruitment Contributes to the Pathogenesis of Heart Failure. Circ. Heart Fail. 2015, 8, 776–787.
  114. Kumar, V.; Prabhu, S.D.; Bansal, S.S. CD4+ T-Lymphocytes Exhibit Biphasic Kinetics Post-Myocardial Infarction. Front. Cardiovasc. Med. 2022, 9, 992653.
  115. Kumar, V.; Rosenzweig, R.; Asalla, S.; Nehra, S.; Prabhu, S.D.; Bansal, S.S. TNFR1 Contributes to Activation-Induced Cell Death of Pathological CD4+ T Lymphocytes During Ischemic Heart Failure. JACC Basic Transl. Sci. 2022, 7, 1038–1049.
  116. Kyaw, T.; Winship, A.; Tay, C.; Kanellakis, P.; Hosseini, H.; Cao, A.; Li, P.; Tipping, P.; Bobik, A.; Toh, B.H. Cytotoxic and Proinflammatory CD8+ T Lymphocytes Promote Development of Vulnerable Atherosclerotic Plaques in ApoE-Deficient Mice. Circulation 2013, 127, 1028–1039.
  117. Gotsman, I.; Grabie, N.; Dacosta, R.; Sukhova, G.; Sharpe, A.; Lichtman, A.H. Proatherogenic Immune Responses Are Regulated by the PD-1/PD-L Pathway in Mice. J. Clin. Investig. 2007, 117, 2974–2982.
  118. Gegunde, S.; Alfonso, A.; Alvariño, R.; Alonso, E.; González-Juanatey, C.; Botana, L.M. Crosstalk between Cyclophilins and T Lymphocytes in Coronary Artery Disease. Exp. Cell Res. 2021, 400, 112514.
  119. Netea, M.G.; Domínguez-Andrés, J.; Barreiro, L.B.; Chavakis, T.; Divangahi, M.; Fuchs, E.; Joosten, L.A.B.; van der Meer, J.W.M.; Mhlanga, M.M.; Mulder, W.J.M.; et al. Defining Trained Immunity and Its Role in Health and Disease. Nat. Rev. Immunol. 2020, 20, 375–388.
  120. Moore, K.J.; Sheedy, F.J.; Fisher, E.A. Macrophages in Atherosclerosis: A Dynamic Balance. Nat. Rev. Immunol. 2013, 13, 709–721.
  121. van der Meer, J.W.M.; Joosten, L.A.B.; Riksen, N.; Netea, M.G. Trained Immunity: A Smart Way to Enhance Innate Immune Defence. Mol. Immunol. 2015, 68, 40–44.
  122. Crișan, T.O.; Cleophas, M.C.P.; Oosting, M.; Lemmers, H.; Toenhake-Dijkstra, H.; Netea, M.G.; Jansen, T.L.; Joosten, L.A.B. Soluble Uric Acid Primes TLR-Induced Proinflammatory Cytokine Production by Human Primary Cells via Inhibition of IL-1Ra. Ann. Rheum. Dis. 2016, 75, 755–762.
  123. Bekkering, S.; Quintin, J.; Joosten, L.A.B.; van der Meer, J.W.M.; Netea, M.G.; Riksen, N.P. Oxidized Low-Density Lipoprotein Induces Long-Term Proinflammatory Cytokine Production and Foam Cell Formation via Epigenetic Reprogramming of Monocytes. Arter. Thromb. Vasc. Biol. 2014, 34, 1731–1738.
  124. Bekkering, S.; Arts, R.J.W.; Novakovic, B.; Kourtzelis, I.; van der Heijden, C.D.C.C.; Li, Y.; Popa, C.D.; ter Horst, R.; van Tuijl, J.; Netea-Maier, R.T.; et al. Metabolic Induction of Trained Immunity through the Mevalonate Pathway. Cell 2018, 172, 135–146.e9.
  125. Joosten, L.A.B.; Crişan, T.O.; Bjornstad, P.; Johnson, R.J. Asymptomatic Hyperuricaemia: A Silent Activator of the Innate Immune System. Nat. Rev. Rheumatol. 2020, 16, 75–86.
  126. van der Heijden, C.D.C.C.; Keating, S.T.; Groh, L.; Joosten, L.A.B.; Netea, M.G.; Riksen, N.P. Aldosterone Induces Trained Immunity: The Role of Fatty Acid Synthesis. Cardiovasc. Res. 2019, 116, 317–328.
  127. Kleinnijenhuis, J.; Quintin, J.; Preijers, F.; Joosten, L.A.B.; Ifrim, D.C.; Saeed, S.; Jacobs, C.; van Loenhout, J.; de Jong, D.; Stunnenberg, H.G.; et al. Bacille Calmette-Guérin Induces NOD2-Dependent Nonspecific Protection from Reinfection via Epigenetic Reprogramming of Monocytes. Proc. Natl. Acad. Sci. USA 2012, 109, 17537–17542.
  128. Marakalala, M.J.; Williams, D.L.; Hoving, J.C.; Engstad, R.; Netea, M.G.; Brown, G.D. Dectin-1 Plays a Redundant Role in the Immunomodulatory Activities of β-Glucan-Rich Ligands in Vivo. Microbes Infect. 2013, 15, 511–515.
  129. Di Luzio, N.R.; Williams, D.L. Protective Effect of Glucan against Systemic Staphylococcus Aureus Septicemia in Normal and Leukemic Mice. Infect. Immun. 1978, 20, 804–810.
  130. Garly, M.-L.; Martins, C.L.; Balé, C.; Baldé, M.A.; Hedegaard, K.L.; Gustafson, P.; Lisse, I.M.; Whittle, H.C.; Aaby, P. BCG Scar and Positive Tuberculin Reaction Associated with Reduced Child Mortality in West Africa. Vaccine 2003, 21, 2782–2790.
  131. Braza, M.S.; van Leent, M.M.T.; Lameijer, M.; Sanchez-Gaytan, B.L.; Arts, R.J.W.; Pérez-Medina, C.; Conde, P.; Garcia, M.R.; Gonzalez-Perez, M.; Brahmachary, M.; et al. Inhibiting Inflammation with Myeloid Cell-Specific Nanobiologics Promotes Organ Transplant Acceptance. Immunity 2018, 49, 819–828.e6.
  132. Wendeln, A.-C.; Degenhardt, K.; Kaurani, L.; Gertig, M.; Ulas, T.; Jain, G.; Wagner, J.; Häsler, L.M.; Wild, K.; Skodras, A.; et al. Innate Immune Memory in the Brain Shapes Neurological Disease Hallmarks. Nature 2018, 556, 332–338.
  133. Bekkering, S.; Joosten, L.A.B.; van der Meer, J.W.M.; Netea, M.G.; Riksen, N.P. Trained Innate Immunity and Atherosclerosis. Curr. Opin. Lipidol. 2013, 24, 487–492.
  134. Arts, R.J.W.; Joosten, L.A.B.; Netea, M.G. The Potential Role of Trained Immunity in Autoimmune and Autoinflammatory Disorders. Front. Immunol. 2018, 9, 298.
  135. Grigoriou, M.; Banos, A.; Filia, A.; Pavlidis, P.; Giannouli, S.; Karali, V.; Nikolopoulos, D.; Pieta, A.; Bertsias, G.; Verginis, P.; et al. Transcriptome Reprogramming and Myeloid Skewing in Haematopoietic Stem and Progenitor Cells in Systemic Lupus Erythematosus. Ann. Rheum. Dis. 2020, 79, 242–253.
  136. Leentjens, J.; Bekkering, S.; Joosten, L.A.B.; Netea, M.G.; Burgner, D.P.; Riksen, N.P. Trained Innate Immunity as a Novel Mechanism Linking Infection and the Development of Atherosclerosis. Circ. Res. 2018, 122, 664–669.
  137. Thiem, K.; Keating, S.T.; Netea, M.G.; Riksen, N.P.; Tack, C.J.; van Diepen, J.; Stienstra, R. Hyperglycemic Memory of Innate Immune Cells Promotes In Vitro Proinflammatory Responses of Human Monocytes and Murine Macrophages. J. Immunol. 2021, 206, 807–813.
  138. Edgar, L.; Akbar, N.; Braithwaite, A.T.; Krausgruber, T.; Gallart-Ayala, H.; Bailey, J.; Corbin, A.L.; Khoyratty, T.E.; Chai, J.T.; Alkhalil, M.; et al. Hyperglycemia Induces Trained Immunity in Macrophages and Their Precursors and Promotes Atherosclerosis. Circulation 2021, 144, 961–982.
  139. van der Heijden, C.D.C.C.; Groh, L.; Keating, S.T.; Kaffa, C.; Noz, M.P.; Kersten, S.; van Herwaarden, A.E.; Hoischen, A.; Joosten, L.A.B.; Timmers, H.J.L.M.; et al. Catecholamines Induce Trained Immunity in Monocytes In Vitro and In Vivo. Circ. Res. 2020, 127, 269–283.
  140. van Tuijl, J.; Joosten, L.A.B.; Netea, M.G.; Bekkering, S.; Riksen, N.P. Immunometabolism Orchestrates Training of Innate Immunity in Atherosclerosis. Cardiovasc. Res. 2019, 115, 1416–1424.
  141. Boffa, M.B.; Koschinsky, M.L. Oxidized Phospholipids as a Unifying Theory for Lipoprotein(a) and Cardiovascular Disease. Nat. Rev. Cardiol. 2019, 16, 305–318.
  142. Nardin, M.; Verdoia, M.; Laera, N.; Cao, D.; De Luca, G. New Insights into Pathophysiology and New Risk Factors for ACS. J. Clin. Med. 2023, 12, 2883.
  143. van der Valk, F.M.; Bekkering, S.; Kroon, J.; Yeang, C.; Van den Bossche, J.; van Buul, J.D.; Ravandi, A.; Nederveen, A.J.; Verberne, H.J.; Scipione, C.; et al. Oxidized Phospholipids on Lipoprotein(a) Elicit Arterial Wall Inflammation and an Inflammatory Monocyte Response in Humans. Circulation 2016, 134, 611–624.
  144. Christ, A.; Günther, P.; Lauterbach, M.A.R.; Duewell, P.; Biswas, D.; Pelka, K.; Scholz, C.J.; Oosting, M.; Haendler, K.; Baßler, K.; et al. Western Diet Triggers NLRP3-Dependent Innate Immune Reprogramming. Cell 2018, 172, 162–175.e14.
  145. Noz, M.P.; ter Telgte, A.; Wiegertjes, K.; Joosten, L.A.B.; Netea, M.G.; de Leeuw, F.-E.; Riksen, N.P. Trained Immunity Characteristics Are Associated with Progressive Cerebral Small Vessel Disease. Stroke 2018, 49, 2910–2917.
  146. Bekkering, S.; van den Munckhof, I.; Nielen, T.; Lamfers, E.; Dinarello, C.; Rutten, J.; de Graaf, J.; Joosten, L.A.B.; Netea, M.G.; Gomes, M.E.R.; et al. Innate Immune Cell Activation and Epigenetic Remodeling in Symptomatic and Asymptomatic Atherosclerosis in Humans In Vivo. Atherosclerosis 2016, 254, 228–236.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 207
Revisions: 2 times (View History)
Update Date: 03 Nov 2023
1000/1000
Video Production Service