Next Article in Journal
The Role of Lutheran/Basal Cell Adhesion Molecule in Hematological Diseases and Tumors
Previous Article in Journal
Female Psammomys obesus Are Protected from Circadian Disruption-Induced Glucose Intolerance, Cardiac Fibrosis and Adipocyte Dysfunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Inflammatory and Immune Mechanisms for Atherosclerotic Cardiovascular Disease in HIV

by
Laura Hmiel
1,†,
Suyu Zhang
2,†,
Laventa M. Obare
3,
Marcela Araujo de Oliveira Santana
4,
Celestine N. Wanjalla
3,
Boghuma K. Titanji
5,
Corrilynn O. Hileman
1 and
Shashwatee Bagchi
6,*
1
Department of Medicine, Division of Infectious Disease, MetroHealth Medical Center and Case Western Reserve University, Cleveland, OH 44109, USA
2
Department of Medicine, Emory University, Atlanta, GA 30322, USA
3
Division of Infectious Diseases, Vanderbilt University Medical Center, Nashville, TN 37232, USA
4
Department of Medicine, Washington University in St. Louis, St. Louis, MO 63110, USA
5
Division of Infectious Diseases, Emory University, Atlanta, GA 30322, USA
6
Division of Infectious Diseases, Washington University in St. Louis, St. Louis, MO 63110, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(13), 7266; https://doi.org/10.3390/ijms25137266
Submission received: 1 June 2024 / Revised: 26 June 2024 / Accepted: 28 June 2024 / Published: 1 July 2024

Abstract

:
Atherosclerotic vascular disease disproportionately affects persons living with HIV (PLWH) compared to those without. The reasons for the excess risk include dysregulated immune response and inflammation related to HIV infection itself, comorbid conditions, and co-infections. Here, we review an updated understanding of immune and inflammatory pathways underlying atherosclerosis in PLWH, including effects of viral products, soluble mediators and chemokines, innate and adaptive immune cells, and important co-infections. We also present potential therapeutic targets which may reduce cardiovascular risk in PLWH.

1. Introduction

Atherosclerotic cardiovascular disease (ASCVD) is an important cause of morbidity and mortality in persons living with human immunodeficiency virus (HIV) (PLWH), even when adjusting for known traditional risk factors [1]. The global burden of HIV-related cardiovascular disease (CVD) has tripled since the early 1990s, and PLWH are over twice as likely to develop CVD or myocardial infarction (MI) compared to people without HIV (PWoH) [2,3,4,5].
The reason for this heightened risk is multifactorial. While traditional CVD risk factors affect PLWH as well as PWoH, PLWH are more likely to be exposed to unique CVD risk factors. Prior reviews have extensively focused on CVD risk due to antiretroviral therapy (ART), especially older protease inhibitors such as lopinavir and ritonavir and the nucleoside reverse transcriptase inhibitor abacavir [6,7].
Inflammation and immune factors have been described in the pathogenesis of ASCVD in PWoH [8,9,10,11]. The interplay of inflammatory and immune pathways and HIV infection has been implicated as a possible explanation for the excess CVD risk in PLWH, and reviewed previously [12,13]. Moreover, the inflammatory response to HIV appears to be affected by co-infections, particularly viral infections like herpesviruses and hepatitis C virus (HCV) [14,15].
In this review, we provide an updated review of the inflammatory and immune mechanisms that contribute to ASCVD in PLWH, with a focus on the innate and adaptive immune mediators of atherosclerosis and the effects of co-pathogens such as HCV, cytomegalovirus (CMV), herpes simplex virus (HSV), and varicella zoster virus (VZV) on the mechanisms of ASCVD in HIV (Figure 1).

2. Regulation of Inflammation in PLWH

2.1. HIV-Related: Viral Replication, Viral Reservoir, Viral Proteins

Early in the HIV epidemic, depletion of CD4+ T cells was a major predictor of morbidity and mortality in PLWH. As ART improved, it became evident that risk of death was not explained by CD4+ T cell count alone but also depended on duration of infection and viremia [16]. HIV itself leads to a dysregulated immune response to bacterial and viral markers including lipopolysaccharide (LPS), resiquimod (R848), and polyinosinic:polycytidylic acid (Poly I:C) [17] while simultaneously causing intrinsic activation by HIV proteins. Furthermore, despite ART-mediated suppression of plasma HIV RNA, the virus persists in reservoirs where continued low-level production of viral proteins and RNA leads to production of pathogen-associated molecular patterns (PAMPs) and activation of pattern recognition receptors (PRR) like toll-like receptors (TLR), which promote pro-inflammatory cytokine signaling [18]. For example, HIV-1 gp120 induces interleukin-1β (IL-1β) release from macrophages through chemokine receptor 5 (CCR5)-mediated signaling pathways, and other proteins including p17, p24, and p41, all functioning as PAMPs that activate CCR5- and/or TLR-mediated pathways [19,20]. Other PRRs of note include retinoic acid-inducible gene I (RIG-I), cyclic GMP-AMP synthetase (cGAS), and interferon inducible protein 16 (IFI16). IFI16 and cGAS both recognize complementary HIV DNA and trigger pro-inflammatory stimulation of interferon genes (STING)-mediated cascades through transcription of the IFN-I gene, NF-κB pathways, and inflammasome activation with pyroptosis of affected T cells [21,22]. Levels of STING and cGAS gene expression subsequently decrease with viral suppression on ART [23]. HIV ssRNA itself activates multiple endosomal TLRs which promote continued inflammation, including TLR7-mediated plasmacytoid dendritic cell activation and TLR8-mediated T cell activation and differentiation to Th1/17 cells and HIV reactivation [24,25]. Furthermore, certain accessory HIV-1 proteins like Vpu and Vpr can modulate the immune response to infection in multiple ways, including downregulating interferon regulatory transcription factor 3 (IRF3) gene expression and subsequent interferon-based pathways, potentially allowing viral evasion of the innate immune system [21,26].

2.2. Alterations of the Microbiome

Gut-mucosal-associated T cell populations represent a significant portion of the lymphoid cells infected by HIV, and the gut mucosa becomes a primary location for viral replication even while on ART [27]. Microbial translocation occurs through epithelial cell damage resulting in systemic inflammation and immune dysregulation [28]. Higher HIV viral loads and lower CD4+ T cell counts correlate with elevated plasma concentrations of bacterial LPS, as well as increased host IL-6, interferon-α (IFN-α), tumor necrosis factor-α (TNF-α), and monocyte and T cell activation [28,29,30,31]. Beta-D-glucan, a fungal cell wall product, can also be elevated in PLWH and is associated with elevated TNF-α and IL-8, epithelial cell damage, and even directly with coronary atherosclerotic plaque [32,33].
HIV infection also affects the microbiome, leading to decreased microbial diversity and shifts in composition, with greater effects at lower CD4+ T cell counts [31]. In high-income western countries, this altered microbiome becomes enriched in opportunistic bacteria including Prevotella and Proteobacteria and diminished in more commensal protective bacteria such as Bacteroides [34]. This shift in microbiome is associated with enhanced local and systemic T cell and dendritic cell responses, as well as rise in pro-inflammatory cytokines associated with mortality including IL-6, TNF-α, IL-1β, and interferon-γ (IFN-γ) [35]. The same shift also leads to a rise in signaling molecules like soluble CD14 (sCD14), which binds LPS in the presence of LPS-binding protein and subsequently activates TLR-4 on myeloid cells [36,37]. While HIV affects the specific microbiome footprint, geography and sexual orientation have robust impacts as well [36].

2.3. Lipids

Dyslipidemia in PLWH typically presents as hypocholesterolemia with relatively lower high-density lipoprotein cholesterol (HDL-c) and higher triglyceride levels than controls [38,39,40], although some studies note lower or similar levels of triglycerides [41]. Sufficiently high HDL-c concentration and function are crucial for cholesterol clearance and are generally anti-atherogenic [42]. However, altered lipoprotein particles and changes in particle sizes (e.g., smaller HDL particles) may correlate better with overall cardiovascular risk in PLWH [43,44]. Lipoprotein (a) is a subset of low-density lipoprotein (LDL) that has been associated with CVD risk and inflammatory markers in PWoH. It has similarly been associated with increased systemic inflammatory markers like IL-6, high-sensitivity C-reactive protein (hs-CRP), and soluble TNFR-I and -II; markers of monocyte activation; and peri-coronary inflammation in PLWH [45]. Additionally, oxidized LDL (oxLDL), which is associated with surrogate measures of atherosclerosis, is present in higher concentrations in PLWH and induces atherogenic pathways as measured by increased sCD14 levels and monocyte activation [46,47].
HDL-mediated reverse cholesterol transport from macrophages to the liver is crucial for prevention of atherogenesis and is impaired in PLWH. HIV production of Nef protein, which is crucial to persistence of HIV infection, disrupts this process and promotes accumulation of cholesterol in macrophages, resulting in foam cells implicated in fatty streak formation [48,49]. Additionally, monocytes in PLWH exhibit lower gene expression of ATP binding cassette subfamily A member 1 (ABCA1), which is transcribed to an ATP-binding cassette transporter crucial for reverse cholesterol transcription [50]. These alterations in macrophage and monocyte function are discussed further in the section on innate immunity.

2.4. Substance Use: Heroin, Cocaine, Methamphetamine, Tobacco, Cannabis

Many PLWH have concomitant substance use disorders, with opioids use disorder (OUD) representing a major public health crisis. Aortic inflammation, which correlates with CVD and atherosclerosis, was found to be increased in PLWH who use heroin compared to both PWoH and PLWH who do not use heroin[51]. Heroin use similarly tends to be associated with markers of microbial translocation and systemic inflammation in PLWH, although this may be independent of HIV status [52]. PLWH who use opioids have demonstrated blunted innate immune responses of monocytes and T cells compared to PLWH without OUD, with reduced production of IL-10, IL-8, IL-6, IL-1α, and TNF-α when exposed to LPS [53].
Stimulants including cocaine and methamphetamines carry well-known cardiovascular risks. Methamphetamine use, especially intravenously, correlates with higher levels of inflammatory markers hs-CRP, IL-6, sTNFR-1, immune activation markers sCD163 and sCD14, and the microbial translocation marker LPS binding protein [54,55]. Similar inflammatory and immune-activating markers including hs-CRP and sCD14 are increased in PLWH who use stimulants such as cocaine or methamphetamines [55].
Tobacco smoking is disproportionately prevalent in PLWH, with between 40 and 70% estimated to smoke tobacco, and has an outsized effect on atherogenesis compared to PWoH [56,57]. Tobacco smoke impairs innate immune cell responses to pathogens while also upregulating activation of growth factors and cytokines including IL-6, IL-8, and TNF-α, promoting high-risk coronary plaque development in PLWH [57,58].
Cannabis use is common among PLWH and has overall anti-inflammatory effects through endogenous cannabinoid pathways, down-regulating inflammatory cell lines like activated T cells and improving gut barrier integrity [59,60]. However, cannabis use is associated with cardiovascular disease including stroke, even when adjusting for confounders such as tobacco smoking [60,61]. Further research into the inflammatory and immune mechanisms of cannabis on vascular health may elucidate the overall effect of cannabis on cardiovascular risk and inflammation after balancing various pathways.

3. Soluble Mediators

The human immune system comprises both cellular components as well as cytokines and other soluble mediators. Cytokines and soluble mediators play integral roles in regulating and driving the inflammation underlying atherosclerosis through downstream autocrine, paracrine, and endocrine effects. Of these, interleukins are critical mediators of HIV-associated atherosclerosis. Broadly characterized as pro- and anti-inflammatory, these cytokines along with other soluble mediators play a direct role in shaping the inflammatory and immune response through the innate and adaptive immune systems [12,62,63,64]. The effects of key systemic inflammatory mediators on atherosclerosis are summarized below.
ART-suppressed PLWH exhibit significantly higher levels of pro-inflammatory cytokines such as IL-1β, IL-6, and TNF-α compared to PWoH [65]. Soluble mediators IL-6, TNF-α, and D-dimer are associated with chronic HIV infection as well as various drivers of atherosclerosis including endothelial dysfunction, increased carotid intima-media thickness (cIMT), and transmigration of inflammatory subsets of monocytes [12,66,67,68,69]. Elevated levels of IL-6 and D-dimer have also been associated with greater risk of major adverse cardiac events (MACE) and CVD in PLWH [70,71,72,73,74,75]. In a study examining the effect of ART initiation on inflammation, women with HIV had persistently higher levels of IL-6, IL-2 receptors, and D-dimer, and had higher levels of subclinical atherosclerosis as measured by cIMT compared to women without HIV [68]. As in PWoH, hs-CRP correlates not only with general inflammation but also CVD, including type 1 MI, and overall mortality after CVD events [70,76]. However, hs-CRP has been variably predictive of CVD and all-cause mortality [71,77].
IL-10 is typically an anti-inflammatory cytokine associated with decreased plaque burden in PLWH [62,78,79]. In some studies, levels have been found to be comparable in PLWH and controls despite decreased plaque burden, which suggests that IL-10 function in the setting of HIV infection may have pronounced effects in local microenvironment, possibly mediated by differential milieu of immune cells or higher local tissue concentrations [62,78,79]. Such discordance of differential effects of plasma or serum levels with tissue level or clinical outcomes has been observed with other biomarkers previously as well [45,77,80].
IL-17a is a pro-inflammatory cytokine primarily produced by Th17 cells, which typically modulates inflammation through its effect on chemokine expression and leukocyte migration [81]. HIV infection depletes Th17 cells, and subsequently promotes the atherosclerotic kynurenine tryptophan catabolic pathway as measured by kynurenine–tryptophan ratio, both of which incompletely normalize after initiation of ART [82,83,84]. IL-17a additionally may promote atherogenesis through its effect on monocytes and macrophages [85]. In PLWH, IL-17a has been shown to be associated with endothelial dysfunction and is theorized as an important driver of inflammation leading to atherosclerosis [82].
IL-18 is another pro-inflammatory cytokine which has been demonstrated to be upregulated in PLWH and has been associated with subclinical atherosclerosis [86]. IL-18 is associated with monocyte and in vitro monocyte-derived macrophage (MDM)-induced inflammasome inflammation [87].
Chemokines and chemokine receptors are integral players in inflammation and inflammatory conditions [88]. Following interaction with their specific chemokine ligands, chemokine receptors trigger a flux in intracellular calcium ions that leads to cellular responses like the onset of chemotaxis of cells to a specific location within the body. Atherosclerosis involves the engagement of chemokines on the surface of the vascular endothelium and the recruitment of inflammatory cells into the vessel wall [89]. CCR5 is one important co-receptor for macrophage-tropic (M) and dual (M and T cell)-tropic HIV-1 [90], and C-X3-C motif chemokine receptor 1 (CX3CR1) is a minor co-receptor for HIV-1 [91]. CCR5+CD4+ T cells are upregulated in PLWH compared to PWoH, and higher proportions of CX3CR1+ CD4+ and CD8+ T cells have been noted in HIV [92,93]. Increased CCR2 expression on monocytes has also been observed in HIV [94]. Finally, expression of CXC motif chemokine ligand (CXCL10), a ligand which binds the chemokine receptor CXCR3, is increased in PLWH [28,95]. Among PLWH on stable ART, increased CCR5+CD8+ T cells, CCR5+CD4+ naïve and effector T cells, and monocyte markers were observed at least one year prior to first acute coronary syndrome (ACS) compared to matched controls who had not experienced ACS [96]. Also in ART-suppressed PLWH, CD16+ monocytes expressing CX3CR1 independently predicted cIMT [97], and higher frequency of CX3CR1+CD4+ T cells was associated with higher baseline and progression of cIMT regardless of HIV viral load or protease inhibitor use [98]. Elevated CXCL10 levels are known to be associated with increased ASCVD risk [99], and these levels return towards normal in PLWH virally suppressed on ART [28,95]. These observations suggest that CCR5, CX3CR1, CXCL10, and perhaps also CCR2 may have heightened significance in modulating atherosclerosis in HIV.
Table 1 lists important available or promising therapeutic agents targeting these soluble mediators of inflammation in atherosclerosis, as well as those targeting innate and adaptive immune cells discussed later.

4. Innate Immunity

The innate immune system plays an integral role in the initiation, establishment, and progression of ASCVD (Figure 2) [143]. Among a growing list of cells under investigation, special attention is paid to neutrophils, monocytes and macrophages, and dendritic cells (DCs) for their significant roles in inflammation and subsequent atherogenesis [10]. These innate immune cells and their dysregulation play critical roles in HIV pathogenesis and are increasingly recognized as having significant roles in HIV-mediated atherogenesis and ASCVD [144,145]. The mechanisms through which innate immune activation and dysregulation contribute to inflammation and subsequent atherogenesis in HIV infection are summarized below.

4.1. Neutrophils

Neutrophils are critical to the development of atherosclerosis in PWoH [146,147]. In PLWH, the role of neutrophil dysfunction in accelerating atherosclerosis has become a significant focus of research [148]. Chronic HIV infection results in elevated IL-8 levels and the expression of various integrins which upregulate and activate neutrophils. These activated neutrophils activate monocytes and MDMs through surface markers like CD11b/CD18 and CD11c/CD18 [148]. Moreover, neutrophils in PLWH both on and off ART display diminished chemotaxis, phagocytosis, bactericidal activity, and oxidative capacity during active viral infection. Concurrently, the dysregulated oxidative metabolism of neutrophils results in higher-than-normal reactive oxygen species (ROS) and oxidative stress, likely contributing to endothelial damage and intracellular stress [148,149,150,151]. This impairment in routine immune functions may contribute to chronic inflammation and endothelial damage caused by foreign pathogens. Finally, the role of neutrophils in gastrointestinal mucosal dysfunction and gut microbial translocation are currently being studied as independent factors in chronic inflammation and atherosclerosis [148]. Therapeutic approaches targeting neutrophil dysfunction in HIV-mediated atherosclerosis have focused on various aspects of neutrophil function, including more recently on the modulation of neutrophil extracellular traps as a therapeutic target for coronary artery disease.

4.2. Monocytes and Macrophages

In HIV infection, monocytes and macrophages contribute to atherosclerosis through altered innate immune signaling which activates a pro-inflammatory phenotype in monocytes and subsequently in macrophages. Chronic inflammation induces endothelial dysfunction, while monocyte and macrophage dysfunction can result in alterations in lipid processing and metabolism [122].
Several studies have shown that PLWH have elevated markers of myeloid cell activation [152]. Pro-inflammatory monocytes and MDM subsets (CD14+CD16+, CD38+, CD69+, CB11b+, sCD163+) in PLWH are linked to the development of carotid artery plaque [139,152,153,154,155]. The mechanisms through which HIV activates pro-inflammatory subsets of monocytes and MDMs are not fully elucidated but may be the result of direct HIV infection and circulating viral proteins [156]. Systemic inflammation with release of inflammatory cytokines and acute phase proteins, as well as chronic endotoxemia associated with HIV infection-mediated gut microbial translocation, also contribute to activation of myeloid cell subsets [153]. Monocyte function declines when ART is interrupted, causing HIV viremia. This suggests that in well-controlled HIV infection, monocytes and macrophages are not fully restored to normal but remain in a state of immune dysregulation [65].
Differential gene expression analysis of MDMs from PLWH has provided novel insights into the characteristics of monocytes and MDMs. A recent study examining peripheral blood monocytes identified extensive changes in innate immune signaling, and the expression of cell markers and immune receptors such as CD163, TLR4, and CD300e in monocytes and MDMs from PLWH [122]. These cells produced higher levels of inflammatory molecules associated with atherosclerosis, including TNF-α, ROS, and matrix metalloproteinases, in PLWH compared with cells in PWoH [122].
HIV infection causes endothelial dysfunction and alters monocyte migration and adhesion. This, combined with disrupted macrophage differentiation and lipid metabolism, leads to cholesterol accumulation in MDMs, forming foam cells and atheromas [153]. HIV infection can also directly stimulate increased expression of cell adhesion molecules such as CD11/CD18 integrins and intracellular adhesion molecule-1 (ICAM-1), which promote monocyte endothelial adhesion and extravascular dissemination [153,157]. Fibronectin fragments, created by proteases activated under inflammatory conditions, are present at high levels in PLWH and facilitate the trans-endothelial migration of HIV-infected mononuclear cells [158]. Additionally, research indicates that HIV-1 infection of macrophages impairs their ability to perform reverse trans-endothelial migration, a critical component of mononuclear cell immune surveillance and patrolling. This impairment may contribute to the formation of viral reservoirs, macrophage dysfunction, and the development of foam cells and chronic inflammation sites [50,159].
HIV-mediated macrophage dysfunction is a crucial area of focus for understanding HIV-associated atherosclerosis. Cellular and genomic studies have shown that macrophages from PLWH exhibit significant changes in innate immune signaling, genetic transcriptional modifications, and altered lipid processing [122,160,161]. HIV infection and associated viral proteins disrupt cholesterol efflux from macrophages, notably through downregulation of a membrane transporter protein called ABCA1 that is critical for cholesterol efflux and prevents excessive buildup of cholesterol within macrophages, which would lead to foam cell formation and atherosclerosis [153,160]. Chronic HIV infection and inflammation with increased levels of IL-6 downregulates lipoprotein lipase activity and increases macrophage uptake of lipids, further disrupting normal lipid metabolism and immunological homeostasis [162]. Additionally, an in vitro study demonstrated that HIV ssRNAs induced foam cell formation in MDMs in a dose-dependent manner, with increased TNF-α release triggered by TLR8 activation, suggesting a potential target for therapeutic intervention [163]. Current investigations of therapeutic strategies targeting macrophage dysfunction in HIV-induced atherosclerosis involve efforts to target and downregulate dysfunctional monocyte and macrophage recruitment.

4.3. Dendritic Cells

DCs represent another important subset of the innate immune system, as they have multiple roles in facilitating atherosclerosis [164,165]. DCs play an active role in recruitment and activation of the adaptive immune system and also directly transmit HIV to CD4+ T cells. Furthermore, it has been shown that HIV-exposed DCs are dysfunctional in their ability to present antigens and mature, leading to unregulated and persistent inflammation and cytokine production including IFN-α, IL-10, and IL-6. Furthermore, autophagy, which is critical to resolution and control of inflammation, is significantly impaired in dendritic cells infected with HIV [145,166].

4.4. NK Cells

Natural killer (NK) cells are lymphocytes that function in many ways as innate immune cells that recognize and eliminate infected or damaged cells [167]. The role of NK cells in atherosclerosis remains uncertain, and whether they possess pro-atherogenic or anti-atherogenic properties warrants further investigation. Single-cell transcriptomic analyses have revealed the heterogeneity of NK cells, suggesting that the balance between inflammatory and anti-inflammatory cells may influence atherosclerosis progression [168]. NK cells might contribute to atherosclerosis by promoting smooth muscle cell apoptosis and necrotic core formation in atherosclerotic plaque [169]. HIV infection is linked to impaired NK cell function, potentially accelerating atherosclerosis in PLWH. Preliminary data indicate that specific NK cell subsets in adolescents with perinatally acquired HIV infection may facilitate oxLDL uptake by vascular macrophages, contributing to atherosclerosis [170].

4.5. Platelets

Platelets play a central role in atherosclerosis [143,171,172]. HIV infection appears to heighten platelet activation and interactions with other components of the immune system, ultimately potentiating atherosclerosis [173,174]. Markers of platelet activation, cell adhesion, and signaling molecules which play central roles in leukocyte activation and activity (P-selectin and CD40 ligand) have been found to be elevated on platelets in HIV infection [173]. These activated platelets also have an increased propensity to form platelet–monocyte complexes, which are elevated in settings of MI [173,175]. During ART suppression, markers of platelet activation such as thromboxane2 synthesis and P-selectin remain elevated compared to healthy controls [176]. Recent studies suggest HIV-induced mitochondrial dysfunction as a potential driver of platelet dysfunction and apoptosis [177]. Additionally, thrombogenicity is noted to be significantly higher in PLWH compared to PWoH [178]. Lastly, some ARTs including ritonavir and abacavir have been associated with platelet hyperactivity, whereas raltegravir and rilpivirine appear to have antithrombotic effects on platelet activation, aggregation, and function [178,179,180,181].

5. Adaptive Immune System

The adaptive immune system plays a crucial role in the pathogenesis of ASCVD in PLWH (Figure 2). This system is responsible for recognizing and responding to specific antigens, including those present in atherosclerotic plaque. While complex, the adaptive immune response mainly involves antigen-driven activation and differentiation of T and B lymphocytes, as well as the production of antibodies and cytokines.

5.1. T Cell Lymphocytes

T cells, including CD4+ and CD8+ subsets, are essential in the adaptive immune response against HIV. Chronic immune activation and inflammation can lead to T-cell dysfunction and exhaustion, contributing to atherosclerosis [182,183]. Higher frequencies of activated CD4+ and CD8+ T cells are observed in untreated PLWH with high viral loads compared to those on ART [184,185].
Persistent T cell activation, despite ART, is pivotal in HIV progression. While ART reduces T cell activation, levels remain elevated, especially in CD8+ T cells [184,186,187]. Elevated levels of activated T cells (CD38+HLA-DR+) during ART predict CD4+ T cell recovery and mortality risk [188]. The role of T cell activation in ASCVD in PLWH remains debated, with studies showing conflicting associations between activated T cells and ASCVD events and subclinical disease [77,189,190]. Kaplan et al. noted a link between activated CD4+ T cells and increased carotid plaque risk in women from the Women’s Interagency HIV Study (WIHS) cohort [190]. Further research is needed to clarify these findings.

5.1.1. CD4+ T cells

CD4+ T cells play diverse roles in atherosclerosis. Low CD4+ counts in PLWH are associated with higher CVD risk and carotid lesions, independent of traditional CVD risk factors [191,192,193,194]. CD4+ T cells produce pro-inflammatory cytokines like IFN-γ and IL-17, contributing to endothelial dysfunction and atherosclerosis [82,195,196]. Naïve CD4+ T cells can differentiate into subtypes (including Th1, Th2, Th17, Treg), each with distinct effects on atherosclerosis [197,198]. Th1 cells have pro-inflammatory effects, while Th2, Th9, and Th17 effects are controversial [199]. Elevated Th17 cells and IL-17a levels are linked to endothelial dysfunction and CVD in PLWH [82,196]. Conversely, ART-treated PLWH with coronary plaques showed lower Th17 frequencies than those without plaques. Notably, Th17 depletion is a major cause of damage to the gut mucosal barrier, which leads to microbial translocation-associated systemic immune activation, which may explain this connection [200].

5.1.2. Regulatory T Cells (Tregs)

Tregs (CD4+CD25+FoxP3+) generally exhibit anti-inflammatory properties. In PLWH with coronary artery disease, Treg frequencies are elevated, but the cells are less differentiated and express lower levels of atheroprotective markers (CD39/CD73), potentially increasing atherosclerosis risk [199,201,202]. These Tregs also have a distinctive migratory capacity towards atherosclerotic plaques, possibly contributing to decreased plaque stability [202].

5.1.3. Cytotoxic CD4+ T cells

Cytotoxic CD4+ T cells express markers which have been implicated in atherosclerosis in PLWH, including CD57 (a marker of senescence), CX3CR1 (a chemokine receptor that homes cells to inflamed endothelium), and GPR56 (an adhesion g-coupled protein receptor that is expressed on exhausted T cells) [98,203,204,205]. These cells are elevated in PLWH with subclinical atherosclerosis and often overlap with senescent TEMRA (T effector re-expressing CD45RA) cells (CD28null and CD45RA+), which are pro-atherosclerotic [196,206,207,208,209,210,211].

5.1.4. CD8+ T cells

CD8+ T cells are also crucial in the pathogenesis of atherosclerosis. Expanded effector CD8+ T cells in ART-suppressed PLWH are linked to cardiovascular morbidity, with many cells expressing the CX3CR1 receptor [92,93,212,213]. CX3CR1 and its ligand CX3CL1, also elevated in PLWH, are associated with CVD morbidity and predict plaque rupture [214,215,216,217]. CD8+ T cells co-localize with CD68+ myeloid cells at sites of endothelial dysfunction in simian immunodeficiency virus (SIV)- and simian-human immunodeficiency virus (SHIV)-infected macaques [218]. Atherosclerotic plaques in PLWH are enriched in activated CD8+ T cells, which are potent cytokine producers and can promote monocyte procoagulant activity via TNF [92,93,219,220]. CX3CL1 expression by dysfunctional endothelium may attract CX3CR1+CD8+ T cells, influencing atherosclerosis progression in both PLWH and PWoH [218,220]. However, CD8+ T cells can reduce advanced atherosclerosis by limiting increases in Th1 cells and macrophages [221].

5.2. B Cells

While the role of B cells in atherosclerosis is well-studied in PWoH [142,222,223], their role in PLWH remains poorly understood [224]. B cells produce antibodies, which play a critical role in the humoral immune response [225]. HIV infection leads to impaired B cell function [226,227], potentially accelerating atherosclerosis in PLWH [224]. B cells are classified into B1 and B2 subsets, the roles of which have primarily been studied in murine models [228]. While B2 cells are the most well-studied subset of B cells, recent research has shown that B1 cells also play a critical role in atherosclerosis [229].

5.2.1. B1 Cells

B1 cells produce natural antibodies, mainly of the IgM isotype, without prior antigen exposure [223]. These antibodies recognize and bind to various antigens including oxLDL and play a protective role in atherosclerosis by promoting clearance of oxLDL from circulation [142,223,230]. B1 cells have not been studied in PLWH.

5.2.2. B2 Cells

B2 cells produce antibodies in response to specific antigens and participate in immune responses to pathogens. In atherosclerosis, B2 cells generate antibodies against oxLDL, forming immune complexes and activating pro-inflammatory pathways, contributing to atherosclerosis [231,232,233]. Antibodies against oxLDL and other plaque antigens may also help clear these antigens and prevent foam cell formation, conferring protection against atherosclerosis [234].

5.2.3. Immune Regulation

B cells also regulate immune responses through interactions with T cells. The cytokine APRIL (a proliferation-inducing ligand) enhances B cell activation and antibody production in HIV [235]. However, increased APRIL levels are associated with atherosclerosis in PLWH [141]. Another B cell-activating cytokine, B-cell activating factor (BAFF), is implicated in pathogenesis of atherosclerosis in HIV infection [236]. Elevated BAFF levels correlate with higher cardiovascular disease risk in PLWH [141]. Targeting B cell function and cytokine production may have therapeutic potential for preventing and treating atherosclerosis in PLWH [224].

6. Co-Pathogens

6.1. Herpesviruses: CMV, EBV, HSV, HHV-8, VZV

Several chronic pathogens prevalent in PLWH are thought to contribute to chronic systemic inflammation (Figure 3). The human herpesviridae family, which includes CMV, HSV-2, and VZV, causes chronic infections which are highly prevalent among PLWH. These viruses have been linked to the development of atherosclerosis [237]. Cross-sectional studies have associated CMV, HSV-2, and VZV with subclinical atherosclerosis in PLWH [189,238,239].

6.1.1. CMV

CMV seropositivity, which is more prevalent among PLWH than PWoH, is linked to an increased risk of ASCVD events and the expansion of potentially pro-atherosclerotic T cell subsets [208]. In a cohort of PLWH, CMV seropositivity was associated with an increased risk of ASCVD events, independent of traditional cardiovascular risk factors [98]. CMV is considered a non-traditional risk factor for atherogenesis, contributing to atherosclerosis by activating endothelial cells, an initial and critical step in the process [240]. HIV and CMV infections both have independent associations with inflammation, inflammation-related morbidities, and CVD risk, particularly in the elderly [15]. CMV infection impairs the nitric oxide synthase pathway, leading to endothelial dysfunction and atherosclerosis [241]. In immunocompetent hosts, CMV infection leads to expansion of pathogenic cytotoxic CMV-specific CD4+ T cells [203,206] and CD8+ T cells [189]. These cells are independently associated with increased cIMT in PLWH [189]. In individuals with CMV, the proportion of CMV-specific cells can constitute up to 30% of the total CD8+ T cell count [242]. While CMV DNA has been detected in atherosclerotic plaques in the general population [237,243,244], only a few studies have demonstrated this in PLWH.
In a small study of PLWH, valganciclovir was found to be effective in suppressing CMV DNA and reducing CD8 T cell activation, as indicated by decreased percentages of cells expressing CD38+HLA-DR+ [245]. This suggests that CMV replication contributes to immune activation in co-infected individuals and that pharmacological suppression of CMV may help reverse this process. However, it remains unclear whether persistent CMV suppression could have clinical benefits in terms of reducing the risk of ASCVD in PLWH. CMV-specific CD4+ T cells expressing CX3CR1 are linked to atherosclerosis, and treatments targeting CMV may impact CVD outcomes [204,246,247]. CMV treatment with letermovir in PLWH was found to affect multiple inflammatory biomarkers; however, effects of CMV suppression on long-term ASCVD outcomes have yet to be fully elucidated [248].

6.1.2. Epstein Barr Virus (EBV)

The role of EBV in HIV-related atherosclerosis remains controversial and complex, with limited research conducted in PLWH. While EBV DNA has been detected in atherosclerotic plaques in the general population [237,244], some studies have found no evidence linking EBV infection markers to coronary artery disease or endothelial dysfunction [249,250]. However, EBV infection can increase acute phase proteins, supporting the hypothesis of an indirect effect [249].
In PLWH with acute HIV infection, up to 80–90% of CD8+ T-lymphocytes are activated, including not only HIV-specific cells but also those specific to CMV and EBV [251]. This suggests a heightened immune response towards multiple pathogens, potentially contributing to the complex interactions between EBV and atherosclerosis in the context of HIV.

6.1.3. HSV-2 and Kaposi-Sarcoma-Associated Herpesvirus (HHV-8)

Herpesviruses including HSV-2 can induce thrombogenic and atherogenic alterations in the cellular structure of their host [252]. Kaposi-sarcoma-associated herpesvirus (HHV-8) targets the lymphatic and vascular systems and can induce atherogenesis. HHV-8-infected vascular endothelial cells can induce the expression of growth factors that cause angiogenesis, endothelial cell proliferation, enhanced vascular permeability, and cytokine production, potentially contributing to atherogenesis [253].
Co-infection with HHV-8 has been associated with increased inflammation and immune activation in virologically suppressed PLWH [254,255]. Lidón et al. found that HHV-8/HSV-2 co-infection was associated with the progression of subclinical atherosclerosis, measured by cIMT. However, when assessing the independent role of each virus, only HHV-8 had a significant relationship with cIMT progression. Participants co-infected with HHV-8 also had higher levels of inflammation measured with hs-CRP, which was associated with faster cIMT progression [255].
In a separate study among men who have sex with men (MSM) in the Multicenter AIDS Cohort Study, HSV-2 was independently associated with subclinical atherosclerosis. The study also found that that the presence of multiple herpesviruses may increase the risk of coronary artery calcium in MSM living with HIV [238]. These findings support the role of herpesviruses in the pathogenesis of atherosclerosis and highlight the importance of controlling herpesvirus infections in individuals at risk for CVD.

6.1.4. VZV

In the general population, several studies have demonstrated a possible correlation between herpes zoster and the incidence of ischemic cardiac and cerebral events, particularly within the first three months following reactivation [256,257,258]. Limited research has been conducted on the association between VZV and atherosclerosis in PLWH, despite the high frequency of VZV reactivation in PLWH both with and without significant CD4+ cell depletion [259,260,261]. In a cohort study of PLWH, VZV seropositivity was associated with an increased risk of ASCVD events, including MI and stroke [262]. Similarly, a study conducted by Masiá et al. investigated the relationship between different types of Herpesviridae and subclinical atherosclerosis by measuring cIMT and flow-mediated dilation (FMD) and found that PLWH with a higher serological response to VZV were more likely to have subclinical atherosclerosis, even after adjusting for the serological response to CMV [239].
The mechanisms by which VZV infection contributes to atherosclerosis are not fully understood, but several potential mechanisms have been proposed. VZV infection has been shown to activate the NF-κB pathway, leading to the production of pro-inflammatory cytokines such as IL-6 and TNF-α [239,263]. These cytokines can recruit immune cells to the arterial wall and promote plaque formation. VZV infection has also been shown to induce endothelial dysfunction, characterized by impaired vasodilation, increased permeability, and elevated expression of adhesion molecules [264]. As VZV is the only herpesvirus with a preventive vaccine, its role in atherosclerosis may be lessened in vaccinated PLWH, especially those with intact immune systems.

6.2. HCV

HCV is a common co-infection with HIV and currently affects an estimated 2.5% of the global population [265]. HCV lifecycle is closely related to lipid metabolism in persons with HCV, using triglyceride-rich very-low-density lipoproteins to create infective lipo-viro particles leading to dyslipidemia and hypolipidemia [266]. Chronic HCV has been associated with clinically significant inflammation, increased risk of ASCVD, and clinical consequences of atherosclerosis, including cerebrovascular accidents and overall cardiovascular mortality [267,268]. Moreover, treatment of chronic HCV infection improves mortality compared to those who remain untreated [269]. HCV infection alone promotes expression of atherogenic adhesive molecules soluble ICAM-1, vascular adhesion molecule-1 (VCAM-1), and E-selectin, as well as pro-inflammatory cytokines associated with CVD mortality like TNF-α [270,271,272,273].
Co-infection with HCV is a known risk factor for CVD in PLWH that, like HIV, is not accounted for in clinical calculators of ASCVD risk. HIV and HCV co-infection poses nearly twice the risk of coronary heart disease compared to HIV infection alone [274,275]. Compared to PWoH, co-infected PLWH had higher levels of many pro-inflammatory cytokines and chemokines including IL-1α, IL-1β, IL-6, IL-12p40, IL-12p70, and TNF-α [276].
HCV/HIV co-infection is also associated with lower levels of LDL but paradoxically higher levels of proprotein convertase subtilisin kexin 9 (PCSK-9) [277], a protein which decreases LDL receptor expression and thus increases circulating LDL concentrations, and likely has other pro-inflammatory and atherogenic roles in uninfected individuals [278]. PCSK-9 levels may be related to inflammatory mechanisms such as IL-6-mediated pathways, given higher levels of IL-6 measured in untreated co-infected PLWH compared with PLWH without HCV [277]. Treatment of HCV with direct-acting antivirals (DAA) has been reported to normalize levels of PCSK9 in co-infected PLWH back to those of PLWH without HCV, as measured about 43 weeks after end of treatment [278].
HCV eradication with DAA, with or without ribavirin-IFN, in HIV/HCV co-infected individuals reduced some other biomarkers of inflammation such as sE-selectin and sCD163, but other markers of inflammation including IL-6, IL-8, sCD14, sVCAM-1, and sICAM-1 and overall LDL levels were generally not affected by therapy [278,279]. Some studies of ribavirin-IFN therapy without DAA in HIV/HCV co-infected individuals have shown reduction in some inflammatory markers, such as sCD163 and sICAM-1 as well as oxLDL, at 24 weeks after treatment [280]. Notably, Chew et al. found that this difference was largely driven by hepatic inflammation, as it was largely attenuated by adjusting for ALT levels [268]. Much more research is needed to investigate the mechanisms driving the increased risk of atherosclerotic vascular disease in PLWH co-infected with HCV.

7. Summary and Future Directions

The inflammatory and immune mechanisms driving the ASCVD risk in HIV infection are multi-dimensional. First, there are numerous clinical and environmental factors that induce inflammation in PLWH, such as changes in the microbiota and lipid derangements, along with convergence with the substance use epidemic (Figure 1). Overall, there is a generalized state of chronic inflammation with elevated levels of key cytokines and chemokines (IL-6, TNF-α, IL-8, CCR5, CCR2, CX3CR1) and immune activation and dysfunction among numerous immune cell types of the innate and adaptive immune systems that are not functionally restored fully by ART-induced viral suppression (Figure 2). There are immune cell subsets that have been implicated in atherosclerosis but remain understudied in HIV infection and should be actively pursued (dendritic cells, NK cells, B cells). Furthermore, PLWH are often co-infected with co-pathogens that have been reported to promote atherogenic pathways that may be accentuated and accelerated with concomitant viral infections (Figure 3). This represents a much-needed area for future investigations, given the high prevalence of co-infections with CMV and/or HCV and HIV, thereby representing a significant public health burden globally. There is only a limited number of available therapeutic immunomodulators for the treatment of atherosclerosis, regardless of HIV status, though several promising targets exist that could be tested (Table 1). There is an urgent need to investigate immunomodulatory pathways that could be therapeutic targets in HIV-associated ASCVD, and to test these in well-designed clinical trials, given the disproportionately high burden of atherosclerosis in this population.

Author Contributions

Conceptualization, S.B.; project administration: S.B.; writing—original draft: L.H., S.Z. and L.M.O.; visualization: M.A.d.O.S., L.M.O., S.Z., L.H., B.K.T., C.O.H. and S.B.; writing—reviewing and editing: L.H., S.Z., L.M.O., C.N.W., B.K.T., C.O.H. and S.B.; supervision: C.N.W., B.K.T., C.O.H. and S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mathers, C.D.; Loncar, D. Projections of Global Mortality and Burden of Disease from 2002 to 2030. PLoS Med. 2006, 3, e442. [Google Scholar] [CrossRef] [PubMed]
  2. Shah, A.S.V.; Stelzle, D.; Lee, K.K.; Beck, E.J.; Alam, S.; Clifford, S.; Longenecker, C.T.; Strachan, F.; Bagchi, S.; Whiteley, W.; et al. Global Burden of Atherosclerotic Cardiovascular Disease in People Living with HIV: Systematic Review and Meta-Analysis. Circulation 2018, 138, 1100–1112. [Google Scholar] [CrossRef] [PubMed]
  3. Boccara, F.; Lang, S.; Meuleman, C.; Ederhy, S.; Mary-Krause, M.; Costagliola, D.; Capeau, J.; Cohen, A. HIV and Coronary Heart Disease: Time for a Better Understanding. J. Am. Coll. Cardiol. 2013, 61, 511–523. [Google Scholar] [CrossRef]
  4. Triant, V.A.; Lee, H.; Hadigan, C.; Grinspoon, S.K. Increased Acute Myocardial Infarction Rates and Cardiovascular Risk Factors among Patients with Human Immunodeficiency Virus Disease. J. Clin. Endocrinol. Metab. 2007, 92, 2506–2512. [Google Scholar] [CrossRef] [PubMed]
  5. Guaraldi, G.; Orlando, G.; Zona, S.; Menozzi, M.; Carli, F.; Garlassi, E.; Berti, A.; Rossi, E.; Roverato, A.; Palella, F. Premature Age-Related Comorbidities among HIV-Infected Persons Compared with the General Population. Clin. Infect. 2011, 53, 1120–1126. [Google Scholar] [CrossRef] [PubMed]
  6. Burrowes, S.; Cahill, P.; Kottilil, S.; Bagchi, S. Contribution of Antiretroviral Therapy to Cardiovascular Disease Risk in HIV-Infected Patients. Future Virol. 2016, 11, 509–527. [Google Scholar] [CrossRef]
  7. Jaschinski, N.; Greenberg, L.; Neesgaard, B.; Miró, J.M.; Grabmeier-Pfistershammer, K.; Wandeler, G.; Smith, C.; De Wit, S.; Wit, F.; Pelchen-Matthews, A.; et al. Recent Abacavir Use and Incident Cardiovascular Disease in Contemporary-Treated People with HIV. AIDS 2023, 37, 467–475. [Google Scholar] [CrossRef] [PubMed]
  8. Hansson, G.K. Inflammation, Atherosclerosis, and Coronary Artery Disease. N. Engl. J. Med. 2005, 352, 1685–1695. [Google Scholar] [CrossRef] [PubMed]
  9. Hansson, G.K.; Hermansson, A. The Immune System in Atherosclerosis. Nat. Immunol. 2011, 12, 204–212. [Google Scholar] [CrossRef]
  10. Libby, P.; Lichtman, A.H.; Hansson, G.K. Immune Effector Mechanisms Implicated in Atherosclerosis: From Mice to Humans. Immunity 2013, 38, 1092–1104. [Google Scholar] [CrossRef]
  11. Teague, H.L.; Ahlman, M.A.; Alavi, A.; Wagner, D.D.; Lichtman, A.H.; Nahrendorf, M.; Swirski, F.K.; Nestle, F.; Gelfand, J.M.; Kaplan, M.J.; et al. Unraveling Vascular Inflammation: From Immunology to Imaging. J. Am. Coll. Cardiol. 2017, 70, 1403–1412. [Google Scholar] [CrossRef] [PubMed]
  12. Hsue, P.Y.; Waters, D.D. HIV Infection and Coronary Heart Disease: Mechanisms and Management. Nat. Rev. Cardiol. 2019, 16, 745–759. [Google Scholar] [CrossRef] [PubMed]
  13. Lo, J.; Plutzky, J. The Biology of Atherosclerosis: General Paradigms and Distinct Pathogenic Mechanisms among HIV-Infected Patients. J. Infect. Dis. 2012, 205 (Suppl. 3), S368–S374. [Google Scholar] [CrossRef] [PubMed]
  14. Jeudy, J.; Patel, P.; George, N.; Burrowes, S.; Husson, J.; Chua, J.; Conn, L.; Weiss, R.G.; Bagchi, S. Assessment of Coronary Inflammation in Antiretroviral Treated People with HIV Infection and Active HIV/Hepatitis C Virus Co-Infection. AIDS 2022, 36, 399–407. [Google Scholar] [CrossRef] [PubMed]
  15. Freeman, M.L.; Lederman, M.M.; Gianella, S. Partners in Crime: The Role of CMV in Immune Dysregulation and Clinical Outcome during HIV Infection. Curr. HIV/AIDS Rep. 2016, 13, 10–19. [Google Scholar] [CrossRef] [PubMed]
  16. The Strategies for Management of Antiretroviral Therapy (SMART) Study Group. Inferior Clinical Outcome of the CD4+ Cell Count–Guided Antiretroviral Treatment Interruption Strategy in the SMART Study: Role of CD4+ Cell Counts and HIV RNA Levels during Follow-Up. J. Infect. Dis. 2008, 197, 1145–1155. [Google Scholar] [CrossRef] [PubMed]
  17. Hove-Skovsgaard, M.; Møller, D.L.; Hald, A.; Gerstoft, J.; Lundgren, J.; Ostrowski, S.R.; Nielsen, S.D. Improved Induced Innate Immune Response after cART Initiation in People with HIV. Front. Immunol. 2022, 13, 974767. [Google Scholar] [CrossRef] [PubMed]
  18. Kreider, E.F.; Bar, K.J. HIV-1 Reservoir Persistence and Decay: Implications for Cure Strategies. Curr. HIV/AIDS Rep. 2022, 19, 194–206. [Google Scholar] [CrossRef] [PubMed]
  19. Cheung, R.; Ravyn, V.; Wang, L.; Ptasznik, A.; Collman, R.G. Signaling Mechanism of HIV-1 Gp120 and Virion-Induced IL-1beta Release in Primary Human Macrophages. J. Immunol. 2008, 180, 6675–6684. [Google Scholar] [CrossRef]
  20. Henrick, B.M.; Yao, X.-D.; Rosenthal, K.L.; INFANT Study Team. HIV-1 Structural Proteins Serve as PAMPs for TLR2 Heterodimers Significantly Increasing Infection and Innate Immune Activation. Front. Immunol. 2015, 6, 426. [Google Scholar] [CrossRef]
  21. Altfeld, M.; Gale, M., Jr. Innate Immunity against HIV-1 Infection. Nat. Immunol. 2015, 16, 554–562. [Google Scholar] [CrossRef] [PubMed]
  22. Nissen, S.K.; Højen, J.F.; Andersen, K.L.D.; Kofod-Olsen, E.; Berg, R.K.; Paludan, S.R.; Østergaard, L.; Jakobsen, M.R.; Tolstrup, M.; Mogensen, T.H. Innate DNA Sensing Is Impaired in HIV Patients and IFI16 Expression Correlates with Chronic Immune Activation. Clin. Exp. Immunol. 2014, 177, 295–309. [Google Scholar] [CrossRef] [PubMed]
  23. de Lima, L.L.P.; de Olivera, A.Q.T.; Moura, T.C.F.; da Silva Graça Amoras, E.; Lima, S.S.; da Silva, A.N.M.R.; Queiroz, M.A.F.; Cayres-Vallinoto, I.M.V.; Ishak, R.; Vallinoto, A.C.R. STING and cGAS Gene Expressions Were Downregulated among HIV-1-Infected Persons after Antiretroviral Therapy. Virol. J. 2021, 18, 78. [Google Scholar] [CrossRef] [PubMed]
  24. Meås, H.Z.; Haug, M.; Beckwith, M.S.; Louet, C.; Ryan, L.; Hu, Z.; Landskron, J.; Nordbø, S.A.; Taskén, K.; Yin, H.; et al. Sensing of HIV-1 by TLR8 Activates Human T Cells and Reverses Latency. Nat. Commun. 2020, 11, 147. [Google Scholar] [CrossRef] [PubMed]
  25. Lv, T.; Cao, W.; Li, T. HIV-Related Immune Activation and Inflammation: Current Understanding and Strategies. J. Immunol. Res. 2021, 2021, 7316456. [Google Scholar] [CrossRef] [PubMed]
  26. Laguette, N.; Brégnard, C.; Hue, P.; Basbous, J.; Yatim, A.; Larroque, M.; Kirchhoff, F.; Constantinou, A.; Sobhian, B.; Benkirane, M. Premature Activation of the SLX4 Complex by Vpr Promotes G2/M Arrest and Escape from Innate Immune Sensing. Cell 2014, 156, 134–145. [Google Scholar] [CrossRef] [PubMed]
  27. Brenchley, J.M.; Schacker, T.W.; Ruff, L.E.; Price, D.A.; Taylor, J.H.; Beilman, G.J.; Nguyen, P.L.; Khoruts, A.; Larson, M.; Haase, A.T.; et al. CD4+ T Cell Depletion during All Stages of HIV Disease Occurs Predominantly in the Gastrointestinal Tract. J. Exp. Med. 2004, 200, 749–759. [Google Scholar] [CrossRef] [PubMed]
  28. Cassol, E.; Malfeld, S.; Mahasha, P.; van der Merwe, S.; Cassol, S.; Seebregts, C.; Alfano, M.; Poli, G.; Rossouw, T. Persistent Microbial Translocation and Immune Activation in HIV-1-Infected South Africans Receiving Combination Antiretroviral Therapy. J. Infect. Dis. 2010, 202, 723–733. [Google Scholar] [CrossRef] [PubMed]
  29. Ancuta, P.; Kamat, A.; Kunstman, K.J.; Kim, E.-Y.; Autissier, P.; Wurcel, A.; Zaman, T.; Stone, D.; Mefford, M.; Morgello, S.; et al. Microbial Translocation Is Associated with Increased Monocyte Activation and Dementia in AIDS Patients. PLoS ONE 2008, 3, e2516. [Google Scholar] [CrossRef]
  30. Jiang, W.; Lederman, M.M.; Hunt, P.; Sieg, S.F.; Haley, K.; Rodriguez, B.; Landay, A.; Martin, J.; Sinclair, E.; Asher, A.I.; et al. Plasma Levels of Bacterial DNA Correlate with Immune Activation and the Magnitude of Immune Restoration in Persons with Antiretroviral-Treated HIV Infection. J. Infect. Dis. 2009, 199, 1177–1185. [Google Scholar] [CrossRef]
  31. Nowak, P.; Troseid, M.; Avershina, E.; Barqasho, B.; Neogi, U.; Holm, K.; Hov, J.R.; Noyan, K.; Vesterbacka, J.; Svärd, J.; et al. Gut Microbiota Diversity Predicts Immune Status in HIV-1 Infection. AIDS 2015, 29, 2409–2418. [Google Scholar] [CrossRef] [PubMed]
  32. Isnard, S.; Fombuena, B.; Sadouni, M.; Lin, J.; Richard, C.; Routy, B.; Ouyang, J.; Ramendra, R.; Peng, X.; Zhang, Y.; et al. Circulating β-d-Glucan as a Marker of Subclinical Coronary Plaque in Antiretroviral Therapy-Treated People with Human Immunodeficiency Virus. Open Forum Infect. Dis. 2021, 8, ofab109. [Google Scholar] [CrossRef] [PubMed]
  33. Morris, A.; Hillenbrand, M.; Finkelman, M.; George, M.P.; Singh, V.; Kessinger, C.; Lucht, L.; Busch, M.; McMahon, D.; Weinman, R.; et al. Serum (1→3)-β-D-Glucan Levels in HIV-Infected Individuals Are Associated with Immunosuppression, Inflammation, and Cardiopulmonary Function. J. Acquir. Immune Defic. Syndr. 1999 2012, 61, 462–468. [Google Scholar] [CrossRef] [PubMed]
  34. Dillon, S.M.; Lee, E.J.; Kotter, C.V.; Austin, G.L.; Dong, Z.; Hecht, D.K.; Gianella, S.; Siewe, B.; Smith, D.M.; Landay, A.L.; et al. An Altered Intestinal Mucosal Microbiome in HIV-1 Infection Is Associated with Mucosal and Systemic Immune Activation and Endotoxemia. Mucosal Immunol. 2014, 7, 983–994. [Google Scholar] [CrossRef] [PubMed]
  35. Dillon, S.M.; Lee, E.J.; Kotter, C.V.; Austin, G.L.; Gianella, S.; Siewe, B.; Smith, D.M.; Landay, A.L.; McManus, M.C.; Robertson, C.E.; et al. Gut Dendritic Cell Activation Links an Altered Colonic Microbiome to Mucosal and Systemic T-Cell Activation in Untreated HIV-1 Infection. Mucosal Immunol. 2016, 9, 24–37. [Google Scholar] [CrossRef] [PubMed]
  36. Rocafort, M.; Gootenberg, D.B.; Luévano, J.M.; Paer, J.M.; Hayward, M.R.; Bramante, J.T.; Ghebremichael, M.S.; Xu, J.; Rogers, Z.H.; Munoz, A.R.; et al. HIV-Associated Gut Microbial Alterations Are Dependent on Host and Geographic Context. Nat. Commun. 2024, 15, 1055. [Google Scholar] [CrossRef] [PubMed]
  37. Dinh, D.M.; Volpe, G.E.; Duffalo, C.; Bhalchandra, S.; Tai, A.K.; Kane, A.V.; Wanke, C.A.; Ward, H.D. Intestinal Microbiota, Microbial Translocation, and Systemic Inflammation in Chronic HIV Infection. J. Infect. Dis. 2015, 211, 19–27. [Google Scholar] [CrossRef] [PubMed]
  38. Padmapriyadarsini, C.; Ramesh, K.; Sekar, L.; Ramachandran, G.; Reddy, D.; Narendran, G.; Sekar, S.; Chandrasekar, C.; Anbarasu, D.; Wanke, C.; et al. Factors Affecting High-Density Lipoprotein Cholesterol in HIV-Infected Patients on Nevirapine-Based Antiretroviral Therapy. Indian J. Med. Res. 2017, 145, 641–650. [Google Scholar] [PubMed]
  39. El-Sadr, W.M.; Mullin, C.M.; Carr, A.; Gibert, C.; Rappoport, C.; Visnegarwala, F.; Grunfeld, C.; Raghavan, S.S. Effects of HIV Disease on Lipid, Glucose and Insulin Levels: Results from a Large Antiretroviral-Naive Cohort. HIV Med. 2005, 6, 114–121. [Google Scholar] [CrossRef]
  40. Riddler, S.A.; Smit, E.; Cole, S.R.; Li, R.; Chmiel, J.S.; Dobs, A.; Palella, F.; Visscher, B.; Evans, R.; Kingsley, L.A. Impact of HIV Infection and HAART on Serum Lipids in Men. JAMA 2003, 289, 2978–2982. [Google Scholar] [CrossRef]
  41. Low, H.; Hoang, A.; Pushkarsky, T.; Dubrovsky, L.; Dewar, E.; Di Yacovo, M.-S.; Mukhamedova, N.; Cheng, L.; Downs, C.; Simon, G.; et al. HIV Disease, Metabolic Dysfunction and Atherosclerosis: A Three Year Prospective Study. PLoS ONE 2019, 14, e0215620. [Google Scholar] [CrossRef] [PubMed]
  42. Ouimet, M.; Barrett, T.J.; Fisher, E.A. HDL and Reverse Cholesterol Transport. Circ. Res. 2019, 124, 1505–1518. [Google Scholar] [CrossRef] [PubMed]
  43. Duprez, D.A.; Kuller, L.H.; Tracy, R.; Otvos, J.; Cooper, D.A.; Hoy, J.; Neuhaus, J.; Paton, N.I.; Friis-Moller, N.; Lampe, F.; et al. Lipoprotein Particle Subclasses, Cardiovascular Disease and HIV Infection. Atherosclerosis 2009, 207, 524–529. [Google Scholar] [CrossRef] [PubMed]
  44. Bucher, H.C.; Richter, W.; Glass, T.R.; Magenta, L.; Wang, Q.; Cavassini, M.; Vernazza, P.; Hirschel, B.; Weber, R.; Furrer, H.; et al. Small Dense Lipoproteins, Apolipoprotein B, and Risk of Coronary Events in HIV-Infected Patients on Antiretroviral Therapy: The Swiss HIV Cohort Study. J. Acquir. Immune Defic. Syndr. 1999 2012, 60, 135–142. [Google Scholar] [CrossRef] [PubMed]
  45. Zisman, E.; Hossain, M.; Funderburg, N.T.; Christenson, R.; Jeudy, J.; Burrowes, S.; Hays, A.G.; George, N.; Freeman, M.L.; Rebuck, H.; et al. Association of Lipoprotein (a) with Peri-Coronary Inflammation in Persons with and without HIV Infection. J. Clin. Lipidol. 2024, 18, e430–e443. [Google Scholar] [CrossRef] [PubMed]
  46. Zidar, D.A.; Juchnowski, S.; Ferrari, B.; Clagett, B.; Pilch-Cooper, H.A.; Rose, S.; Rodriguez, B.; McComsey, G.A.; Sieg, S.F.; Mehta, N.N.; et al. Oxidized LDL Levels Are Increased in HIV Infection and May Drive Monocyte Activation. J. Acquir. Immune Defic. Syndr. 1999 2015, 69, 154–160. [Google Scholar] [CrossRef] [PubMed]
  47. Hileman, C.O.; Turner, R.; Funderburg, N.T.; Semba, R.D.; McComsey, G.A. Changes in Oxidized Lipids Drive the Improvement in Monocyte Activation and Vascular Disease after Statin Therapy in HIV. AIDS 2016, 30, 65–73. [Google Scholar] [CrossRef] [PubMed]
  48. Mujawar, Z.; Rose, H.; Morrow, M.P.; Pushkarsky, T.; Dubrovsky, L.; Mukhamedova, N.; Fu, Y.; Dart, A.; Orenstein, J.M.; Bobryshev, Y.V.; et al. Human Immunodeficiency Virus Impairs Reverse Cholesterol Transport from Macrophages. PLoS Biol. 2006, 4, e365. [Google Scholar] [CrossRef]
  49. Cui, H.L.; Ditiatkovski, M.; Kesani, R.; Bobryshev, Y.V.; Liu, Y.; Geyer, M.; Mukhamedova, N.; Bukrinsky, M.; Sviridov, D. HIV Protein Nef Causes Dyslipidemia and Formation of Foam Cells in Mouse Models of Atherosclerosis. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2014, 28, 2828–2839. [Google Scholar] [CrossRef]
  50. Maisa, A.; Hearps, A.C.; Angelovich, T.A.; Pereira, C.F.; Zhou, J.; Shi, M.D.Y.; Palmer, C.S.; Muller, W.A.; Crowe, S.M.; Jaworowski, A. Monocytes from HIV-Infected Individuals Show Impaired Cholesterol Efflux and Increased Foam Cell Formation after Transendothelial Migration. AIDS 2015, 29, 1445–1457. [Google Scholar] [CrossRef]
  51. Hileman, C.O.; Durieux, J.C.; Janus, S.E.; Bowman, E.; Kettelhut, A.; Nguyen, T.-T.; Avery, A.K.; Funderburg, N.; Sullivan, C.; McComsey, G.A. Heroin Use Is Associated with Vascular Inflammation in Human Immunodeficiency Virus. Clin. Infect. Dis. 2023, 76, 375–381. [Google Scholar] [CrossRef]
  52. Hileman, C.O.; Bowman, E.R.; Gabriel, J.; Kettelhut, A.; Labbato, D.; Smith, C.; Avery, A.; Parran, T.; Funderburg, N.; McComsey, G.A. Impact of Heroin and HIV on Gut Integrity and Immune Activation. J. Acquir. Immune Defic. Syndr. 1999 2022, 89, 519–526. [Google Scholar] [CrossRef]
  53. Underwood, M.L.; Nguyen, T.; Uebelhoer, L.S.; Kunkel, L.E.; Korthuis, P.T.; Lancioni, C.L. Altered Monocyte Phenotype and Dysregulated Innate Cytokine Responses among People Living with HIV and Opioid-Use Disorder. AIDS 2020, 34, 177–188. [Google Scholar] [CrossRef]
  54. Miller, M.; Lee, J.-Y.; Fulcher, J.A.; Roach, M.E.; Dilworth, S.E.; Chahine, A.; Pallikkuth, S.; Fuchs, D.; Pahwa, S.; Carrico, A.W. Getting to the Point: Methamphetamine Injection Is Associated with Biomarkers Relevant to HIV Pathogenesis. Drug Alcohol Depend. 2020, 213, 108133. [Google Scholar] [CrossRef]
  55. Cherenack, E.M.; Chavez, J.V.; Martinez, C.; Hirshfield, S.; Balise, R.; Horvath, K.J.; Viamonte, M.; Jimenez, D.E.; Paul, R.; Dilworth, S.E.; et al. Stimulant Use, HIV, and Immune Dysregulation among Sexual Minority Men. Drug Alcohol Depend. 2023, 251, 110942. [Google Scholar] [CrossRef] [PubMed]
  56. Pacek, L.R.; Cioe, P.A. Tobacco Use, Use Disorders, and Smoking Cessation Interventions in Persons Living with HIV. Curr. HIV/AIDS Rep. 2015, 12, 413–420. [Google Scholar] [CrossRef] [PubMed]
  57. Kelly, S.G.; Plankey, M.; Post, W.S.; Li, X.; Stall, R.; Jacobson, L.P.; Witt, M.D.; Kingsley, L.; Cox, C.; Budoff, M.; et al. Associations between Tobacco, Alcohol, and Drug Use with Coronary Artery Plaque among HIV-Infected and Uninfected Men in the Multicenter AIDS Cohort Study. PLoS ONE 2016, 11, e0147822. [Google Scholar] [CrossRef] [PubMed]
  58. Calvo, M.; Laguno, M.; Martínez, M.; Martínez, E. Effects of Tobacco Smoking on HIV-Infected Individuals. AIDS Rev. 2015, 17, 47–55. [Google Scholar]
  59. Costiniuk, C.T.; Jenabian, M.-A. Cannabinoids and Inflammation: Implications for People Living with HIV. AIDS 2019, 33, 2273–2288. [Google Scholar] [CrossRef]
  60. Ellis, R.J.; Wilson, N.; Peterson, S. Cannabis and Inflammation in HIV: A Review of Human and Animal Studies. Viruses 2021, 13, 1581. [Google Scholar] [CrossRef]
  61. Lorenz, D.R.; Dutta, A.; Mukerji, M.; Holman, A.; Uno, H.; Gabuzda, D. Marijuana Use Impacts Midlife Cardiovascular Events in HIV-Infected Men. Clin. Infect. 2017, 65, 626–635. [Google Scholar] [CrossRef]
  62. Fourman, L.T.; Saylor, C.F.; Cheru, L.; Fitch, K.; Looby, S.; Keller, K.; Robinson, J.A.; Hoffmann, U.; Lu, M.T.; Burdo, T.; et al. Anti-Inflammatory Interleukin 10 Inversely Relates to Coronary Atherosclerosis in Persons with Human Immunodeficiency Virus. J. Infect. Dis. 2020, 221, 510–515. [Google Scholar] [CrossRef]
  63. León, R.; Reus, S.; López, N.; Portilla, I.; Sánchez-Payá, J.; Giner, L.; Boix, V.; Merino, E.; Torrús, D.; Moreno-Pérez, Ó.; et al. Subclinical Atherosclerosis in Low Framingham Risk HIV Patients. Eur. J. Clin. Investig. 2017, 47, 591–599. [Google Scholar] [CrossRef]
  64. Kearns, A.; Gordon, J.; Burdo, T.H.; Qin, X. HIV-1-Associated Atherosclerosis: Unraveling the Missing Link. J. Am. Coll. Cardiol. 2017, 69, 3084–3098. [Google Scholar] [CrossRef] [PubMed]
  65. Tilton, J.C.; Johnson, A.J.; Luskin, M.R.; Manion, M.M.; Yang, J.; Adelsberger, J.W.; Lempicki, R.A.; Hallahan, C.W.; McLaughlin, M.; Mican, J.M.; et al. Diminished Production of Monocyte Proinflammatory Cytokines during Human Immunodeficiency Virus Viremia Is Mediated by Type I Interferons. J. Virol. 2006, 80, 11486–11497. [Google Scholar] [CrossRef]
  66. Hileman, C.O.; Longenecker, C.T.; Carman, T.L.; Milne, G.L.; Labbato, D.E.; Storer, N.J.; White, C.A.; McComsey, G.A. Elevated D-Dimer Is Independently Associated with Endothelial Dysfunction: A Cross-Sectional Study in HIV-Infected Adults on Antiretroviral Therapy. Antivir. Ther. 2012, 17, 1345–1349. [Google Scholar] [CrossRef] [PubMed]
  67. Pontrelli, G.; Martino, A.M.; Tchidjou, H.K.; Citton, R.; Mora, N.; Ravà, L.; Tozzi, A.E.; Palma, P.; Muraca, M.; Franco, E.; et al. HIV Is Associated with Thrombophilia and High D-Dimer in Children and Adolescents. AIDS 2010, 24, 1145–1151. [Google Scholar] [CrossRef] [PubMed]
  68. Kaplan, R.C.; Landay, A.L.; Hodis, H.N.; Gange, S.J.; Norris, P.J.; Young, M.; Anastos, K.; Tien, P.C.; Xue, X.; Lazar, J.; et al. Potential Cardiovascular Disease Risk Markers among HIV-Infected Women Initiating Antiretroviral Treatment. J. Acquir. Immune Defic. Syndr. 1999 2012, 60, 359–368. [Google Scholar] [CrossRef]
  69. Chow, D.C.; Saiki, K.M.W.; Siriwardhana, C.; Lozano-Gerona, J.; Vanapruks, S.; Ogle, J.; Premeaux, T.A.; Ndhlovu, L.C.; Boisvert, W.A. Increased Transmigration of Intermediate Monocytes Associated with Atherosclerotic Burden in People with HIV on Antiretroviral Therapy. AIDS 2023, 37, 1177–1179. [Google Scholar] [CrossRef]
  70. Nordell, A.D.; McKenna, M.; Borges, Á.H.; Duprez, D.; Neuhaus, J.; Neaton, J.D.; INSIGHT SMART, ESPRIT Study Groups. SILCAAT Scientific Committee Severity of Cardiovascular Disease Outcomes among Patients with HIV Is Related to Markers of Inflammation and Coagulation. J. Am. Heart Assoc. 2014, 3, e000844. [Google Scholar] [CrossRef]
  71. Kuller, L.H.; Tracy, R.; Belloso, W.; De Wit, S.; Drummond, F.; Lane, H.C.; Ledergerber, B.; Lundgren, J.; Neuhaus, J.; Nixon, D.; et al. Inflammatory and Coagulation Biomarkers and Mortality in Patients with HIV Infection. PLoS Med. 2008, 5, e203. [Google Scholar] [CrossRef] [PubMed]
  72. Tenorio, A.R.; Zheng, Y.; Bosch, R.J.; Krishnan, S.; Rodriguez, B.; Hunt, P.W.; Plants, J.; Seth, A.; Wilson, C.C.; Deeks, S.G.; et al. Soluble Markers of Inflammation and Coagulation but Not T-Cell Activation Predict Non-AIDS-Defining Morbid Events during Suppressive Antiretroviral Treatment. J. Infect. Dis. 2014, 210, 1248–1259. [Google Scholar] [CrossRef]
  73. So-Armah, K.A.; Tate, J.P.; Chang, C.-C.H.; Butt, A.A.; Gerschenson, M.; Gibert, C.L.; Leaf, D.; Rimland, D.; Rodriguez-Barradas, M.C.; Budoff, M.J.; et al. Do Biomarkers of Inflammation, Monocyte Activation, and Altered Coagulation Explain Excess Mortality Between HIV Infected and Uninfected People? J. Acquir. Immune Defic. Syndr. 2016, 72, 206–213. [Google Scholar] [CrossRef] [PubMed]
  74. Wada, N.I.; Bream, J.H.; Martínez-Maza, O.; Macatangay, B.; Galvin, S.R.; Margolick, J.B.; Jacobson, L.P. Inflammatory Biomarkers and Mortality Risk among HIV-Suppressed Men: A Multisite Prospective Cohort Study. Clin. Infect. Dis. 2016, 63, 984–990. [Google Scholar] [CrossRef] [PubMed]
  75. Baker, J.V.; Sharma, S.; Grund, B.; Rupert, A.; Metcalf, J.A.; Schechter, M.; Munderi, P.; Aho, I.; Emery, S.; Babiker, A.; et al. Systemic Inflammation, Coagulation, and Clinical Risk in the START Trial. Open Forum Infect. Dis. 2017, 4, ofx262. [Google Scholar] [CrossRef] [PubMed]
  76. Graham, S.M.; Nance, R.M.; Chen, J.; Wurfel, M.M.; Hunt, P.W.; Heckbert, S.R.; Budoff, M.J.; Moore, R.D.; Jacobson, J.M.; Martin, J.N.; et al. Plasma Interleukin-6 (IL-6), Angiopoietin-2, and C-Reactive Protein Levels Predict Subsequent Type 1 Myocardial Infarction in Persons with Treated HIV Infection. J. Acquir. Immune Defic. Syndr. 2023, 93, 282–291. [Google Scholar] [CrossRef] [PubMed]
  77. Ford, E.S.; Greenwald, J.H.; Richterman, A.G.; Rupert, A.; Dutcher, L.; Badralmaa, Y.; Natarajan, V.; Rehm, C.; Hadigan, C.; Sereti, I. Traditional Risk Factors and D-Dimer Predict Incident Cardiovascular Disease Events in Chronic HIV Infection. AIDS 2010, 24, 1509–1517. [Google Scholar] [CrossRef] [PubMed]
  78. Werede, A.T.; Terry, J.G.; Nair, S.; Temu, T.M.; Shepherd, B.E.; Bailin, S.S.; Mashayekhi, M.; Gabriel, C.L.; Lima, M.; Woodward, B.O.; et al. Mean Coronary Cross-Sectional Area as a Measure of Arterial Remodeling Using Noncontrast CT Imaging in Persons with HIV. J. Am. Heart Assoc. 2022, 11, e025768. [Google Scholar] [CrossRef] [PubMed]
  79. Mallat, Z.; Besnard, S.; Duriez, M.; Deleuze, V.; Emmanuel, F.; Bureau, M.F.; Soubrier, F.; Esposito, B.; Duez, H.; Fievet, C.; et al. Protective Role of Interleukin-10 in Atherosclerosis. Circ. Res. 1999, 85, e17–e24. [Google Scholar] [CrossRef]
  80. Lu, M.T.; Ribaudo, H.; Foldyna, B.; Zanni, M.V.; Mayrhofer, T.; Karady, J.; Taron, J.; Fitch, K.V.; McCallum, S.; Burdo, T.H.; et al. Effects of Pitavastatin on Coronary Artery Disease and Inflammatory Biomarkers in HIV: Mechanistic Substudy of the REPRIEVE Randomized Clinical Trial. JAMA Cardiol. 2024, 9, 323–334. [Google Scholar] [CrossRef]
  81. Butcher, M.; Galkina, E. Current Views on the Functions of Interleukin-17A-Producing Cells in Atherosclerosis. Thromb. Haemost. 2011, 106, 787–795. [Google Scholar] [CrossRef] [PubMed]
  82. Wanjalla, C.N.; Temu, T.M.; Mashayekhi, M.; Warren, C.M.; Shepherd, B.E.; Gangula, R.; Fuseini, H.; Bailin, S.; Gabriel, C.L.; Gangula, P.; et al. Interleukin-17A Is Associated with Flow-Mediated Dilation and Interleukin-4 with Carotid Plaque in Persons with HIV. AIDS 2022, 36, 963–973. [Google Scholar] [CrossRef] [PubMed]
  83. Boyd, A.; Boccara, F.; Meynard, J.-L.; Ichou, F.; Bastard, J.-P.; Fellahi, S.; Samri, A.; Sauce, D.; Haddour, N.; Autran, B.; et al. Serum Tryptophan-Derived Quinolinate and Indole-3-Acetate Are Associated with Carotid Intima-Media Thickness and Its Evolution in HIV-Infected Treated Adults. Open Forum Infect. Dis. 2019, 6, ofz516. [Google Scholar] [CrossRef] [PubMed]
  84. Gelpi, M.; Hartling, H.J.; Ueland, P.M.; Ullum, H.; Trøseid, M.; Nielsen, S.D. Tryptophan Catabolism and Immune Activation in Primary and Chronic HIV Infection. BMC Infect. Dis. 2017, 17, 349. [Google Scholar] [CrossRef]
  85. 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. [Google Scholar] [CrossRef] [PubMed]
  86. Kearns, A.C.; Liu, F.; Dai, S.; Robinson, J.A.; Kiernan, E.; Tesfaye Cheru, L.; Peng, X.; Gordon, J.; Morgello, S.; Abuova, A.; et al. Caspase-1 Activation Is Related with HIV-Associated Atherosclerosis in an HIV Transgenic Mouse Model and HIV Patient Cohort. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 1762–1775. [Google Scholar] [CrossRef]
  87. Caocci, M.; Niu, M.; Fox, H.S.; Burdo, T.H. HIV Infection Drives Foam Cell Formation via NLRP3 Inflammasome Activation. Int. J. Mol. Sci. 2024, 25, 2367. [Google Scholar] [CrossRef] [PubMed]
  88. Charo, I.F.; Ransohoff, R.M. The Many Roles of Chemokines and Chemokine Receptors in Inflammation. N. Engl. J. Med. 2006, 354, 610–621. [Google Scholar] [CrossRef] [PubMed]
  89. Glass, C.K.; Witztum, J.L. Atherosclerosis. the Road Ahead. Cell 2001, 104, 503–516. [Google Scholar] [CrossRef]
  90. Dragic, T.; Litwin, V.; Allaway, G.P.; Martin, S.R.; Huang, Y.; Nagashima, K.A.; Cayanan, C.; Maddon, P.J.; Koup, R.A.; Moore, J.P.; et al. HIV-1 Entry into CD4+ Cells Is Mediated by the Chemokine Receptor CC-CKR-5. Nature 1996, 381, 667–673. [Google Scholar] [CrossRef]
  91. Garin, A.; Tarantino, N.; Faure, S.; Daoudi, M.; Lécureuil, C.; Bourdais, A.; Debré, P.; Deterre, P.; Combadiere, C. Two Novel Fully Functional Isoforms of CX3CR1 Are Potent HIV Coreceptors. J. Immunol. 2003, 171, 5305–5312. [Google Scholar] [CrossRef] [PubMed]
  92. Combadière, B.; Faure, S.; Autran, B.; Debré, P.; Combadière, C. The Chemokine Receptor CX3CR1 Controls Homing and Anti-Viral Potencies of CD8 Effector-Memory T Lymphocytes in HIV-Infected Patients. AIDS 2003, 17, 1279–1290. [Google Scholar] [CrossRef] [PubMed]
  93. Mudd, J.C.; Panigrahi, S.; Kyi, B.; Moon, S.H.; Manion, M.M.; Younes, S.-A.; Sieg, S.F.; Funderburg, N.T.; Zidar, D.A.; Lederman, M.M.; et al. Inflammatory Function of CX3CR1+ CD8+ T Cells in Treated HIV Infection Is Modulated by Platelet Interactions. J. Infect. Dis. 2016, 214, 1808–1816. [Google Scholar] [CrossRef] [PubMed]
  94. Eugenin, E.A.; Osiecki, K.; Lopez, L.; Goldstein, H.; Calderon, T.M.; Berman, J.W. CCL2/Monocyte Chemoattractant Protein-1 Mediates Enhanced Transmigration of Human Immunodeficiency Virus (HIV)-Infected Leukocytes across the Blood-Brain Barrier: A Potential Mechanism of HIV-CNS Invasion and NeuroAIDS. J. Neurosci. Off. J. Soc. Neurosci. 2006, 26, 1098–1106. [Google Scholar] [CrossRef] [PubMed]
  95. Zanni, M.V.; Toribio, M.; Robbins, G.K.; Burdo, T.H.; Lu, M.T.; Ishai, A.E.; Feldpausch, M.N.; Martin, A.; Melbourne, K.; Triant, V.A.; et al. Effects of Antiretroviral Therapy on Immune Function and Arterial Inflammation in Treatment-Naive Patients with Human Immunodeficiency Virus Infection. JAMA Cardiol. 2016, 1, 474–480. [Google Scholar] [CrossRef] [PubMed]
  96. Tarancon-Diez, L.; De Pablo-Bernal, R.S.; Álvarez-Rios, A.I.; Rosado-Sánchez, I.; Dominguez-Molina, B.; Genebat, M.; Pacheco, Y.M.; Jiménez, J.L.; Muñoz-Fernández, M.Á.; Ruiz-Mateos, E.; et al. CCR5+ CD8 T-Cell Levels and Monocyte Activation Precede the Onset of Acute Coronary Syndrome in HIV-Infected Patients on Antiretroviral Therapy. Thromb. Haemost. 2017, 117, 1141–1149. [Google Scholar] [CrossRef] [PubMed]
  97. Westhorpe, C.L.V.; Maisa, A.; Spelman, T.; Hoy, J.F.; Dewar, E.M.; Karapanagiotidis, S.; Hearps, A.C.; Cheng, W.-J.; Trevillyan, J.; Lewin, S.R.; et al. Associations between Surface Markers on Blood Monocytes and Carotid Atherosclerosis in HIV-Positive Individuals. Immunol. Cell Biol. 2014, 92, 133–138. [Google Scholar] [CrossRef] [PubMed]
  98. Sacre, K.; Hunt, P.W.; Hsue, P.Y.; Maidji, E.; Martin, J.N.; Deeks, S.G.; Autran, B.; McCune, J.M. A Role for Cytomegalovirus-Specific CD4+CX3CR1+ T Cells and Cytomegalovirus-Induced T-Cell Immunopathology in HIV-Associated Atherosclerosis. AIDS 2012, 26, 805–814. [Google Scholar] [CrossRef]
  99. van den Borne, P.; Quax, P.H.A.; Hoefer, I.E.; Pasterkamp, G. The Multifaceted Functions of CXCL10 in Cardiovascular Disease. BioMed Res. Int. 2014, 2014, 893106. [Google Scholar] [CrossRef]
  100. Ridker, P.M.; Everett, B.M.; Thuren, T.; MacFadyen, J.G.; Chang, W.H.; Ballantyne, C.; Fonseca, F.; Nicolau, J.; Koenig, W.; Anker, S.D.; et al. Antiinflammatory Therapy with Canakinumab for Atherosclerotic Disease. N. Engl. J. Med. 2017, 377, 1119–1131. [Google Scholar] [CrossRef]
  101. Hsue, P.Y.; Li, D.; Ma, Y.; Ishai, A.; Manion, M.; Nahrendorf, M.; Ganz, P.; Ridker, P.M.; Deeks, S.G.; Tawakol, A. IL-1β Inhibition Reduces Atherosclerotic Inflammation in HIV Infection. J. Am. Coll. Cardiol. 2018, 72, 2809–2811. [Google Scholar] [CrossRef] [PubMed]
  102. Broch, K.; Anstensrud, A.K.; Woxholt, S.; Sharma, K.; Tøllefsen, I.M.; Bendz, B.; Aakhus, S.; Ueland, T.; Amundsen, B.H.; Damås, J.K.; et al. Randomized Trial of Interleukin-6 Receptor Inhibition in Patients with Acute ST-Segment Elevation Myocardial Infarction. J. Am. Coll. Cardiol. 2021, 77, 1845–1855. [Google Scholar] [CrossRef] [PubMed]
  103. Funderburg, N.T.; Shive, C.L.; Chen, Z.; Tatsuoka, C.; Bowman, E.R.; Longenecker, C.T.; McComsey, G.A.; Clagett, B.M.; Dorazio, D.; Freeman, M.L.; et al. Interleukin 6 Blockade with Tocilizumab Diminishes Indices of Inflammation That Are Linked to Mortality in Treated Human Immunodeficiency Virus Infection. Clin. Infect. Dis. 2023, 77, 272–279. [Google Scholar] [CrossRef] [PubMed]
  104. Ridker, P.M.; Everett, B.M.; Pradhan, A.; MacFadyen, J.G.; Solomon, D.H.; Zaharris, E.; Mam, V.; Hasan, A.; Rosenberg, Y.; Iturriaga, E.; et al. Low-Dose Methotrexate for the Prevention of Atherosclerotic Events. N. Engl. J. Med. 2019, 380, 752–762. [Google Scholar] [CrossRef] [PubMed]
  105. Hsue, P.Y.; Ribaudo, H.J.; Deeks, S.G.; Bell, T.; Ridker, P.M.; Fichtenbaum, C.; Daar, E.S.; Havlir, D.; Yeh, E.; Tawakol, A.; et al. Safety and Impact of Low-Dose Methotrexate on Endothelial Function and Inflammation in Individuals with Treated Human Immunodeficiency Virus: AIDS Clinical Trials Group Study A5314. Clin. Infect. Dis. 2019, 68, 1877–1886. [Google Scholar] [CrossRef] [PubMed]
  106. Tardif, J.-C.; Kouz, S.; Waters, D.D.; Bertrand, O.F.; Diaz, R.; Maggioni, A.P.; Pinto, F.J.; Ibrahim, R.; Gamra, H.; Kiwan, G.S.; et al. Efficacy and Safety of Low-Dose Colchicine after Myocardial Infarction. N. Engl. J. Med. 2019, 381, 2497–2505. [Google Scholar] [CrossRef] [PubMed]
  107. Nidorf, S.M.; Fiolet, A.T.L.; Mosterd, A.; Eikelboom, J.W.; Schut, A.; Opstal, T.S.J.; The, S.H.K.; Xu, X.-F.; Ireland, M.A.; Lenderink, T.; et al. Colchicine in Patients with Chronic Coronary Disease. N. Engl. J. Med. 2020, 383, 1838–1847. [Google Scholar] [CrossRef]
  108. Bouabdallaoui, N.; Tardif, J.-C.; Waters, D.D.; Pinto, F.J.; Maggioni, A.P.; Diaz, R.; Berry, C.; Koenig, W.; Lopez-Sendon, J.; Gamra, H.; et al. Time-to-Treatment Initiation of Colchicine and Cardiovascular Outcomes after Myocardial Infarction in the Colchicine Cardiovascular Outcomes Trial (COLCOT). Eur. Heart J. 2020, 41, 4092–4099. [Google Scholar] [CrossRef]
  109. Hennessy, T.; Soh, L.; Bowman, M.; Kurup, R.; Schultz, C.; Patel, S.; Hillis, G.S. The Low Dose Colchicine after Myocardial Infarction (LoDoCo-MI) Study: A Pilot Randomized Placebo Controlled Trial of Colchicine Following Acute Myocardial Infarction. Am. Heart J. 2019, 215, 62–69. [Google Scholar] [CrossRef]
  110. Tong, D.C.; Quinn, S.; Nasis, A.; Hiew, C.; Roberts-Thomson, P.; Adams, H.; Sriamareswaran, R.; Htun, N.M.; Wilson, W.; Stub, D.; et al. Colchicine in Patients with Acute Coronary Syndrome: The Australian COPS Randomized Clinical Trial. Circulation 2020, 142, 1890–1900. [Google Scholar] [CrossRef]
  111. Hays, A.G.; Schär, M.; Barditch-Crovo, P.; Bagchi, S.; Bonanno, G.; Meyer, J.; Afework, Y.; Streeb, V.; Stradley, S.; Kelly, S.; et al. A Randomized, Placebo-Controlled, Double-Blinded Clinical Trial of Colchicine to Improve Vascular Health in People Living with HIV. AIDS 2021, 35, 1041–1050. [Google Scholar] [CrossRef] [PubMed]
  112. Chou, R.; Cantor, A.; Dana, T.; Wagner, J.; Ahmed, A.Y.; Fu, R.; Ferencik, M. Statin Use for the Primary Prevention of Cardiovascular Disease in Adults: Updated Evidence Report and Systematic Review for the US Preventive Services Task Force. JAMA 2022, 328, 754–771. [Google Scholar] [CrossRef] [PubMed]
  113. Grinspoon, S.K.; Fitch, K.V.; Zanni, M.V.; Fichtenbaum, C.J.; Umbleja, T.; Aberg, J.A.; Overton, E.T.; Malvestutto, C.D.; Bloomfield, G.S.; Currier, J.S.; et al. Pitavastatin to Prevent Cardiovascular Disease in HIV Infection. N. Engl. J. Med. 2023, 389, 687–699. [Google Scholar] [CrossRef] [PubMed]
  114. Longenecker, C.T.; Hileman, C.O.; Funderburg, N.T.; McComsey, G.A. Rosuvastatin Preserves Renal Function and Lowers Cystatin C in HIV-Infected Subjects on Antiretroviral Therapy: The SATURN-HIV Trial. Clin. Infect. Dis. 2014, 59, 1148–1156. [Google Scholar] [CrossRef]
  115. Aberg, J.A.; Sponseller, C.A.; Ward, D.J.; Kryzhanovski, V.A.; Campbell, S.E.; Thompson, M.A. Pitavastatin versus Pravastatin in Adults with HIV-1 Infection and Dyslipidaemia (INTREPID): 12 Week and 52 Week Results of a Phase 4, Multicentre, Randomised, Double-Blind, Superiority Trial. Lancet HIV 2017, 4, e284–e294. [Google Scholar] [CrossRef] [PubMed]
  116. Bedimo, R.J.; Mar, H.; Bosch, R.J.; Drechsler, H.; Cyktor, J.C.; Macatangay, B.J.C.; Lalama, C.; Rinaldo, C.; Collier, A.; Godfrey, C.; et al. Brief Report: No Evidence for an Association between Statin Use and Lower Biomarkers of HIV Persistence or Immune Activation/Inflammation during Effective ART. J. Acquir. Immune Defic. Syndr. 2019, 82, e27–e31. [Google Scholar] [CrossRef] [PubMed]
  117. Sabatine, M.S.; Giugliano, R.P.; Keech, A.C.; Honarpour, N.; Wiviott, S.D.; Murphy, S.A.; Kuder, J.F.; Wang, H.; Liu, T.; Wasserman, S.M.; et al. Evolocumab and Clinical Outcomes in Patients with Cardiovascular Disease. N. Engl. J. Med. 2017, 376, 1713–1722. [Google Scholar] [CrossRef] [PubMed]
  118. Leucker, T.M.; Gerstenblith, G.; Schär, M.; Brown, T.T.; Jones, S.R.; Afework, Y.; Weiss, R.G.; Hays, A.G. Evolocumab, a PCSK9-Monoclonal Antibody, Rapidly Reverses Coronary Artery Endothelial Dysfunction in People Living with HIV and People with Dyslipidemia. J. Am. Heart Assoc. 2020, 9, e016263. [Google Scholar] [CrossRef]
  119. O’Brien, M.P.; Hunt, P.W.; Kitch, D.W.; Klingman, K.; Stein, J.H.; Funderburg, N.T.; Berger, J.S.; Tebas, P.; Clagett, B.; Moisi, D.; et al. A Randomized Placebo Controlled Trial of Aspirin Effects on Immune Activation in Chronically Human Immunodeficiency Virus-Infected Adults on Virologically Suppressive Antiretroviral Therapy. Open Forum Infect. Dis. 2017, 4, ofw278. [Google Scholar] [CrossRef]
  120. Macatangay, B.J.C.; Jackson, E.K.; Abebe, K.Z.; Comer, D.; Cyktor, J.; Klamar-Blain, C.; Borowski, L.; Gillespie, D.G.; Mellors, J.W.; Rinaldo, C.R.; et al. A Randomized, Placebo-Controlled, Pilot Clinical Trial of Dipyridamole to Decrease Human Immunodeficiency Virus-Associated Chronic Inflammation. J. Infect. Dis. 2020, 221, 1598–1606. [Google Scholar] [CrossRef]
  121. Erbel, C.; Akhavanpoor, M.; Okuyucu, D.; Wangler, S.; Dietz, A.; Zhao, L.; Stellos, K.; Little, K.M.; Lasitschka, F.; Doesch, A.; et al. IL-17A Influences Essential Functions of the Monocyte/Macrophage Lineage and Is Involved in Advanced Murine and Human Atherosclerosis. J. Immunol. 2014, 193, 4344–4355. [Google Scholar] [CrossRef] [PubMed]
  122. Bowman, E.R.; Cameron, C.M.; Richardson, B.; Kulkarni, M.; Gabriel, J.; Cichon, M.J.; Riedl, K.M.; Mustafa, Y.; Cartwright, M.; Snyder, B.; et al. Macrophage Maturation from Blood Monocytes Is Altered in People with HIV, and Is Linked to Serum Lipid Profiles and Activation Indices: A Model for Studying Atherogenic Mechanisms. PLoS Pathog. 2020, 16, e1008869. [Google Scholar] [CrossRef] [PubMed]
  123. von Stebut, E.; Boehncke, W.-H.; Ghoreschi, K.; Gori, T.; Kaya, Z.; Thaci, D.; Schäffler, A. IL-17A in Psoriasis and Beyond: Cardiovascular and Metabolic Implications. Front. Immunol. 2019, 10, 3096. [Google Scholar] [CrossRef] [PubMed]
  124. Tenger, C.; Sundborger, A.; Jawien, J.; Zhou, X. IL-18 Accelerates Atherosclerosis Accompanied by Elevation of IFN-Gamma and CXCL16 Expression Independently of T Cells. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 791–796. [Google Scholar] [CrossRef] [PubMed]
  125. Francisci, D.; Pirro, M.; Schiaroli, E.; Mannarino, M.R.; Cipriani, S.; Bianconi, V.; Alunno, A.; Bagaglia, F.; Bistoni, O.; Falcinelli, E.; et al. Maraviroc Intensification Modulates Atherosclerotic Progression in HIV-Suppressed Patients at High Cardiovascular Risk. A Randomized, Crossover Pilot Study. Open Forum Infect. Dis. 2019, 6, ofz112. [Google Scholar] [CrossRef] [PubMed]
  126. Kelly, K.M.; Tocchetti, C.G.; Lyashkov, A.; Tarwater, P.M.; Bedja, D.; Graham, D.R.; Beck, S.E.; Metcalf Pate, K.A.; Queen, S.E.; Adams, R.J.; et al. CCR5 Inhibition Prevents Cardiac Dysfunction in the SIV/Macaque Model of HIV. J. Am. Heart Assoc. 2014, 3, e000874. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, K.; Hao, Z.; Zheng, H.; Wang, H.; Zhang, L.; Yan, M.; Tuerhong, R.; Zhou, Y.; Wang, Y.; Pang, T.; et al. Repurposing of Rilpivirine for Preventing Platelet Β3 Integrin-Dependent Thrombosis by Targeting c-Src Active Autophosphorylation. Thromb. Res. 2023, 229, 53–68. [Google Scholar] [CrossRef]
  128. Ytterberg, S.R.; Bhatt, D.L.; Mikuls, T.R.; Koch, G.G.; Fleischmann, R.; Rivas, J.L.; Germino, R.; Menon, S.; Sun, Y.; Wang, C.; et al. Cardiovascular and Cancer Risk with Tofacitinib in Rheumatoid Arthritis. N. Engl. J. Med. 2022, 386, 316–326. [Google Scholar] [CrossRef]
  129. Marconi, V.C.; Moser, C.; Gavegnano, C.; Deeks, S.G.; Lederman, M.M.; Overton, E.T.; Tsibris, A.; Hunt, P.W.; Kantor, A.; Sekaly, R.-P.; et al. Randomized Trial of Ruxolitinib in Antiretroviral-Treated Adults with Human Immunodeficiency Virus. Clin. Infect. Dis. 2022, 74, 95–104. [Google Scholar] [CrossRef]
  130. Syngle, A.; Garg, N.; Chauhan, K. Pos0205 Amelioration of Endothelial Dysfunction with Jak Inhibition in Rheumatoid Arthritis: Jak Cv-Risk Reduction Study. Ann. Rheum. Dis. 2021, 80, 318–319. [Google Scholar] [CrossRef]
  131. Lu, W.; Arraes, L.C.; Ferreira, W.T.; Andrieu, J.-M. Therapeutic Dendritic-Cell Vaccine for Chronic HIV-1 Infection. Nat. Med. 2004, 10, 1359–1365. [Google Scholar] [CrossRef] [PubMed]
  132. García, F.; Lejeune, M.; Climent, N.; Gil, C.; Alcamí, J.; Morente, V.; Alós, L.; Ruiz, A.; Setoain, J.; Fumero, E.; et al. Therapeutic Immunization with Dendritic Cells Loaded with Heat-Inactivated Autologous HIV-1 in Patients with Chronic HIV-1 Infection. J. Infect. Dis. 2005, 191, 1680–1685. [Google Scholar] [CrossRef] [PubMed]
  133. Mohamed, H.; Miller, V.; Jennings, S.R.; Wigdahl, B.; Krebs, F.C. The Evolution of Dendritic Cell Immunotherapy against HIV-1 Infection: Improvements and Outlook. J. Immunol. Res. 2020, 2020, 9470102. [Google Scholar] [CrossRef] [PubMed]
  134. Yi, S.; Zhang, X.; Sangji, H.; Liu, Y.; Allen, S.D.; Xiao, B.; Bobbala, S.; Braverman, C.L.; Cai, L.; Hecker, P.I.; et al. Surface Engineered Polymersomes for Enhanced Modulation of Dendritic Cells during Cardiovascular Immunotherapy. Adv. Funct. Mater. 2019, 29, 1904399. [Google Scholar] [CrossRef] [PubMed]
  135. Shi, Y.; Dong, M.; Wu, Y.; Gong, F.; Wang, Z.; Xue, L.; Su, Z. An Elastase-Inhibiting, Plaque-Targeting and Neutrophil-Hitchhiking Liposome against Atherosclerosis. Acta Biomater. 2024, 173, 470–481. [Google Scholar] [CrossRef] [PubMed]
  136. Mu, W.; Sharma, M.; Heymans, R.; Ritou, E.; Rezek, V.; Hamid, P.; Kossyvakis, A.; Sen Roy, S.; Grijalva, V.; Chattopadhyay, A.; et al. Apolipoprotein A-I Mimetics Attenuate Macrophage Activation in Chronic Treated HIV. AIDS 2021, 35, 543–553. [Google Scholar] [CrossRef] [PubMed]
  137. Wolska, A.; Reimund, M.; Sviridov, D.O.; Amar, M.J.; Remaley, A.T. Apolipoprotein Mimetic Peptides: Potential New Therapies for Cardiovascular Diseases. Cells 2021, 10, 597. [Google Scholar] [CrossRef]
  138. Walker, J.A.; Miller, A.D.; Burdo, T.H.; McGrath, M.S.; Williams, K.C. Direct Targeting of Macrophages with Methylglyoxal-Bis-Guanylhydrazone Decreases SIV-Associated Cardiovascular Inflammation and Pathology. J. Acquir. Immune Defic. Syndr. 2017, 74, 583–592. [Google Scholar] [CrossRef]
  139. Walker, J.A.; Beck, G.A.; Campbell, J.H.; Miller, A.D.; Burdo, T.H.; Williams, K.C. Anti-A4 Integrin Antibody Blocks Monocyte/Macrophage Traffic to the Heart and Decreases Cardiac Pathology in a SIV Infection Model of AIDS. J. Am. Heart Assoc. 2015, 4, e001932. [Google Scholar] [CrossRef]
  140. Moccia, M.; Albero, R.; Lanzillo, R.; Saccà, F.; De Rosa, A.; Russo, C.V.; Carotenuto, A.; Palladino, R.; Brescia Morra, V. Cardiovascular Profile Improvement during Natalizumab Treatment. Metab. Brain Dis. 2018, 33, 981–986. [Google Scholar] [CrossRef]
  141. Aranguren, M.; Doyon-Laliberté, K.; El-Far, M.; Chartrand-Lefebvre, C.; Routy, J.-P.; Barril, J.-G.; Trottier, B.; Tremblay, C.; Durand, M.; Poudrier, J.; et al. Subclinical Atherosclerosis Is Associated with Discrepancies in BAFF and APRIL Levels and Altered Breg Potential of Precursor-like Marginal Zone B-Cells in Long-Term HIV Treated Individuals. Vaccines 2022, 11, 81. [Google Scholar] [CrossRef] [PubMed]
  142. Sage, A.P.; Tsiantoulas, D.; Binder, C.J.; Mallat, Z. The Role of B Cells in Atherosclerosis. Nat. Rev. Cardiol. 2019, 16, 180–196. [Google Scholar] [CrossRef] [PubMed]
  143. Roy, P.; Orecchioni, M.; Ley, K. How the Immune System Shapes Atherosclerosis: Roles of Innate and Adaptive Immunity. Nat. Rev. Immunol. 2022, 22, 251–265. [Google Scholar] [CrossRef]
  144. Campbell, J.H.; Hearps, A.C.; Martin, G.E.; Williams, K.C.; Crowe, S.M. The Importance of Monocytes and Macrophages in HIV Pathogenesis, Treatment, and Cure. AIDS 2014, 28, 2175–2187. [Google Scholar] [CrossRef] [PubMed]
  145. Wacleche, V.S.; Tremblay, C.L.; Routy, J.-P.; Ancuta, P. The Biology of Monocytes and Dendritic Cells: Contribution to HIV Pathogenesis. Viruses 2018, 10, 65. [Google Scholar] [CrossRef]
  146. Soehnlein, O. Multiple Roles for Neutrophils in Atherosclerosis. Circ. Res. 2012, 110, 875–888. [Google Scholar] [CrossRef]
  147. Silvestre-Roig, C.; Braster, Q.; Ortega-Gomez, A.; Soehnlein, O. Neutrophils as Regulators of Cardiovascular Inflammation. Nat. Rev. Cardiol. 2020, 17, 327–340. [Google Scholar] [CrossRef]
  148. Hensley-McBain, T.; Klatt, N.R. The Dual Role of Neutrophils in HIV Infection. Curr. HIV/AIDS Rep. 2018, 15, 1–10. [Google Scholar] [CrossRef]
  149. Flø, R.W.; Naess, A.; Nilsen, A.; Harthug, S.; Solberg, C.O. A Longitudinal Study of Phagocyte Function in HIV-Infected Patients. AIDS 1994, 8, 771–777. [Google Scholar] [CrossRef]
  150. Elbim, C.; Prevot, M.H.; Bouscarat, F.; Franzini, E.; Chollet-Martin, S.; Hakim, J.; Gougerot-Pocidalo, M.A. Polymorphonuclear Neutrophils from Human Immunodeficiency Virus-Infected Patients Show Enhanced Activation, Diminished fMLP-Induced L-Selectin Shedding, and an Impaired Oxidative Burst after Cytokine Priming. Blood 1994, 84, 2759–2766. [Google Scholar] [CrossRef]
  151. Ellis, M.; Gupta, S.; Galant, S.; Hakim, S.; VandeVen, C.; Toy, C.; Cairo, M.S. Impaired Neutrophil Function in Patients with AIDS or AIDS-Related Complex: A Comprehensive Evaluation. J. Infect. Dis. 1988, 158, 1268–1276. [Google Scholar] [CrossRef] [PubMed]
  152. Hanna, D.B.; Lin, J.; Post, W.S.; Hodis, H.N.; Xue, X.; Anastos, K.; Cohen, M.H.; Gange, S.J.; Haberlen, S.A.; Heath, S.L.; et al. Association of Macrophage Inflammation Biomarkers with Progression of Subclinical Carotid Artery Atherosclerosis in HIV-Infected Women and Men. J. Infect. Dis. 2017, 215, 1352–1361. [Google Scholar] [CrossRef]
  153. Crowe, S.M.; Westhorpe, C.L.V.; Mukhamedova, N.; Jaworowski, A.; Sviridov, D.; Bukrinsky, M. The Macrophage: The Intersection between HIV Infection and Atherosclerosis. J. Leukoc. Biol. 2010, 87, 589–598. [Google Scholar] [CrossRef] [PubMed]
  154. Subramanian, S.; Tawakol, A.; Burdo, T.H.; Abbara, S.; Wei, J.; Vijayakumar, J.; Corsini, E.; Abdelbaky, A.; Zanni, M.V.; Hoffmann, U.; et al. Arterial Inflammation in Patients with HIV. JAMA 2012, 308, 379–386. [Google Scholar] [CrossRef] [PubMed]
  155. McKibben, R.A.; Margolick, J.B.; Grinspoon, S.; Li, X.; Palella, F.J.; Kingsley, L.A.; Witt, M.D.; George, R.T.; Jacobson, L.P.; Budoff, M.; et al. Elevated Levels of Monocyte Activation Markers Are Associated with Subclinical Atherosclerosis in Men with and Those without HIV Infection. J. Infect. Dis. 2015, 211, 1219–1228. [Google Scholar] [CrossRef] [PubMed]
  156. Appay, V.; Boutboul, F.; Autran, B. The HIV Infection and Immune Activation: “To Fight and Burn”. Curr. Infect. Dis. Rep. 2005, 7, 473–479. [Google Scholar] [CrossRef] [PubMed]
  157. Rossen, R.D.; Smith, C.W.; Laughter, A.H.; Noonan, C.A.; Anderson, D.C.; McShan, W.M.; Hurvitz, M.Y.; Orson, F.M. HIV-1-Stimulated Expression of CD11/CD18 Integrins and ICAM-1: A Possible Mechanism for Extravascular Dissemination of HIV-1-Infected Cells. Trans. Assoc. Am. Physicians 1989, 102, 117–130. [Google Scholar] [PubMed]
  158. Birdsall, H.H.; Porter, W.J.; Green, D.M.; Rubio, J.; Trial, J.; Rossen, R.D. Impact of Fibronectin Fragments on the Transendothelial Migration of HIV-Infected Leukocytes and the Development of Subendothelial Foci of Infectious Leukocytes. J. Immunol. 2004, 173, 2746–2754. [Google Scholar] [CrossRef]
  159. Westhorpe, C.L.V.; Zhou, J.; Webster, N.L.; Kalionis, B.; Lewin, S.R.; Jaworowski, A.; Muller, W.A.; Crowe, S.M. Effects of HIV-1 Infection in Vitro on Transendothelial Migration by Monocytes and Monocyte-Derived Macrophages. J. Leukoc. Biol. 2009, 85, 1027–1035. [Google Scholar] [CrossRef]
  160. Mukhamedova, N.; Hoang, A.; Dragoljevic, D.; Dubrovsky, L.; Pushkarsky, T.; Low, H.; Ditiatkovski, M.; Fu, Y.; Ohkawa, R.; Meikle, P.J.; et al. Exosomes Containing HIV Protein Nef Reorganize Lipid Rafts Potentiating Inflammatory Response in Bystander Cells. PLoS Pathog. 2019, 15, e1007907. [Google Scholar] [CrossRef]
  161. Renga, B.; Francisci, D.; D’Amore, C.; Schiaroli, E.; Carino, A.; Baldelli, F.; Fiorucci, S. HIV-1 Infection Is Associated with Changes in Nuclear Receptor Transcriptome, pro-Inflammatory and Lipid Profile of Monocytes. BMC Infect. Dis. 2012, 12, 274. [Google Scholar] [CrossRef] [PubMed]
  162. Yudkin, J.S.; Kumari, M.; Humphries, S.E.; Mohamed-Ali, V. Inflammation, Obesity, Stress and Coronary Heart Disease: Is Interleukin-6 the Link? Atherosclerosis 2000, 148, 209–214. [Google Scholar] [CrossRef] [PubMed]
  163. Bernard, M.A.; Han, X.; Inderbitzin, S.; Agbim, I.; Zhao, H.; Koziel, H.; Tachado, S.D. HIV-Derived ssRNA Binds to TLR8 to Induce Inflammation-Driven Macrophage Foam Cell Formation. PLoS ONE 2014, 9, e104039. [Google Scholar] [CrossRef]
  164. Zernecke, A. Dendritic Cells in Atherosclerosis: Evidence in Mice and Humans. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 763–770. [Google Scholar] [CrossRef] [PubMed]
  165. Engelen, S.E.; Robinson, A.J.B.; Zurke, Y.-X.; Monaco, C. Therapeutic Strategies Targeting Inflammation and Immunity in Atherosclerosis: How to Proceed? Nat. Rev. Cardiol. 2022, 19, 522–542. [Google Scholar] [CrossRef] [PubMed]
  166. Smed-Sörensen, A.; Loré, K. Dendritic Cells at the Interface of Innate and Adaptive Immunity to HIV-1. Curr. Opin. HIV AIDS 2011, 6, 405–410. [Google Scholar] [CrossRef] [PubMed]
  167. Tosello-Trampont, A.; Surette, F.A.; Ewald, S.E.; Hahn, Y.S. Immunoregulatory Role of NK Cells in Tissue Inflammation and Regeneration. Front. Immunol. 2017, 8, 301. [Google Scholar] [CrossRef]
  168. Yang, C.; Siebert, J.R.; Burns, R.; Gerbec, Z.J.; Bonacci, B.; Rymaszewski, A.; Rau, M.; Riese, M.J.; Rao, S.; Carlson, K.-S.; et al. Heterogeneity of Human Bone Marrow and Blood Natural Killer Cells Defined by Single-Cell Transcriptome. Nat. Commun. 2019, 10, 3931. [Google Scholar] [CrossRef] [PubMed]
  169. Martínez-Rodríguez, J.E.; Munné-Collado, J.; Rasal, R.; Cuadrado, E.; Roig, L.; Ois, A.; Muntasell, A.; Baro, T.; Alameda, F.; Roquer, J.; et al. Expansion of the NKG2C+ Natural Killer-Cell Subset Is Associated with High-Risk Carotid Atherosclerotic Plaques in Seropositive Patients for Human Cytomegalovirus. Arterioscler. Thromb. Vasc. Biol. 2013, 33, 2653–2659. [Google Scholar] [CrossRef]
  170. Alles, M.; Gunasena, M.; Kettelhut, A.; Ailstock, K.; Musiime, V.; Kityo, C.; Richardson, B.; Mulhern, W.; Tamilselvan, B.; Rubsamen, M.; et al. Activated NK Cells with Pro-Inflammatory Features Are Associated with Atherogenesis in Perinatally HIV-Acquired Adolescents. medRxiv 2023. [Google Scholar] [CrossRef]
  171. Nording, H.M.; Seizer, P.; Langer, H.F. Platelets in Inflammation and Atherogenesis. Front. Immunol. 2015, 6, 98. [Google Scholar] [CrossRef]
  172. Weber, C. Platelets and Chemokines in Atherosclerosis: Partners in Crime. Circ. Res. 2005, 96, 612–616. [Google Scholar] [CrossRef] [PubMed]
  173. O’Halloran, J.A.; Dunne, E.; Gurwith, M.; Lambert, J.S.; Sheehan, G.J.; Feeney, E.R.; Pozniak, A.; Reiss, P.; Kenny, D.; Mallon, P. The Effect of Initiation of Antiretroviral Therapy on Monocyte, Endothelial and Platelet Function in HIV-1 Infection. HIV Med. 2015, 16, 608–619. [Google Scholar] [CrossRef]
  174. Mayne, E.; Funderburg, N.T.; Sieg, S.F.; Asaad, R.; Kalinowska, M.; Rodriguez, B.; Schmaier, A.H.; Stevens, W.; Lederman, M.M. Increased Platelet and Microparticle Activation in HIV Infection: Upregulation of P-Selectin and Tissue Factor Expression. J. Acquir. Immune Defic. Syndr. 2012, 59, 340–346. [Google Scholar] [CrossRef]
  175. Sarma, J.; Laan, C.A.; Alam, S.; Jha, A.; Fox, K.A.A.; Dransfield, I. Increased Platelet Binding to Circulating Monocytes in Acute Coronary Syndromes. Circulation 2002, 105, 2166–2171. [Google Scholar] [CrossRef]
  176. Nkambule, B.B.; Mxinwa, V.; Mkandla, Z.; Mutize, T.; Mokgalaboni, K.; Nyambuya, T.M.; Dludla, P.V. Platelet Activation in Adult HIV-Infected Patients on Antiretroviral Therapy: A Systematic Review and Meta-Analysis. BMC Med. 2020, 18, 357. [Google Scholar] [CrossRef] [PubMed]
  177. Mesquita, E.C.; Hottz, E.D.; Amancio, R.T.; Carneiro, A.B.; Palhinha, L.; Coelho, L.E.; Grinsztejn, B.; Zimmerman, G.A.; Rondina, M.T.; Weyrich, A.S.; et al. Persistent Platelet Activation and Apoptosis in Virologically Suppressed HIV-Infected Individuals. Sci. Rep. 2018, 8, 14999. [Google Scholar] [CrossRef] [PubMed]
  178. Tunjungputri, R.N.; Van Der Ven, A.J.; Schonsberg, A.; Mathan, T.S.; Koopmans, P.; Roest, M.; Fijnheer, R.; Groot, P.G.D.E.; de Mast, Q. Reduced Platelet Hyperreactivity and Platelet-Monocyte Aggregation in HIV-Infected Individuals Receiving a Raltegravir-Based Regimen. AIDS 2014, 28, 2091–2096. [Google Scholar] [CrossRef] [PubMed]
  179. Loelius, S.G.; Lannan, K.L.; Blumberg, N.; Phipps, R.P.; Spinelli, S.L. The HIV Protease Inhibitor, Ritonavir, Dysregulates Human Platelet Function in Vitro. Thromb. Res. 2018, 169, 96–104. [Google Scholar] [CrossRef]
  180. Falcinelli, E.; Francisci, D.; Belfiori, B.; Petito, E.; Guglielmini, G.; Malincarne, L.; Mezzasoma, A.; Sebastiano, M.; Conti, V.; Giannini, S.; et al. In Vivo Platelet Activation and Platelet Hyperreactivity in Abacavir-Treated HIV-Infected Patients. Thromb. Haemost. 2013, 110, 349–357. [Google Scholar] [CrossRef]
  181. Satchell, C.S.; O’Halloran, J.A.; Cotter, A.G.; Peace, A.J.; O’Connor, E.F.; Tedesco, A.F.; Feeney, E.R.; Lambert, J.S.; Sheehan, G.J.; Kenny, D.; et al. Increased Platelet Reactivity in HIV-1-Infected Patients Receiving Abacavir-Containing Antiretroviral Therapy. J. Infect. Dis. 2011, 204, 1202–1210. [Google Scholar] [CrossRef] [PubMed]
  182. Trautmann, L.; Janbazian, L.; Chomont, N.; Said, E.A.; Gimmig, S.; Bessette, B.; Boulassel, M.-R.; Delwart, E.; Sepulveda, H.; Balderas, R.S.; et al. Upregulation of PD-1 Expression on HIV-Specific CD8+ T Cells Leads to Reversible Immune Dysfunction. Nat. Med. 2006, 12, 1198–1202. [Google Scholar] [CrossRef] [PubMed]
  183. Paiardini, M.; Müller-Trutwin, M. HIV-Associated Chronic Immune Activation. Immunol. Rev. 2013, 254, 78–101. [Google Scholar] [CrossRef]
  184. Hunt, P.W.; Martin, J.N.; Sinclair, E.; Bredt, B.; Hagos, E.; Lampiris, H.; Deeks, S.G. T Cell Activation Is Associated with Lower CD4+ T Cell Gains in Human Immunodeficiency Virus-Infected Patients with Sustained Viral Suppression during Antiretroviral Therapy. J. Infect. Dis. 2003, 187, 1534–1543. [Google Scholar] [CrossRef] [PubMed]
  185. Xia, H.; Jiang, W.; Zhang, X.; Qin, L.; Su, B.; Li, Z.; Sun, J.; Zhang, Y.; Zhang, T.; Lu, X.; et al. Elevated Level of CD4+ T Cell Immune Activation in Acutely HIV-1-Infected Stage Associates with Increased IL-2 Production and Cycling Expression, and Subsequent CD4+ T Cell Preservation. Front. Immunol. 2018, 9, 616. [Google Scholar] [CrossRef]
  186. Lederman, M.M.; Calabrese, L.; Funderburg, N.T.; Clagett, B.; Medvik, K.; Bonilla, H.; Gripshover, B.; Salata, R.A.; Taege, A.; Lisgaris, M.; et al. Immunologic Failure despite Suppressive Antiretroviral Therapy Is Related to Activation and Turnover of Memory CD4 Cells. J. Infect. Dis. 2011, 204, 1217–1226. [Google Scholar] [CrossRef] [PubMed]
  187. Hunt, P.W.; Landay, A.L.; Sinclair, E.; Martinson, J.A.; Hatano, H.; Emu, B.; Norris, P.J.; Busch, M.P.; Martin, J.N.; Brooks, C.; et al. A Low T Regulatory Cell Response May Contribute to Both Viral Control and Generalized Immune Activation in HIV Controllers. PLoS ONE 2011, 6, e15924. [Google Scholar] [CrossRef] [PubMed]
  188. Hunt, P.W.; Cao, H.L.; Muzoora, C.; Ssewanyana, I.; Bennett, J.; Emenyonu, N.; Kembabazi, A.; Neilands, T.B.; Bangsberg, D.R.; Deeks, S.G.; et al. Impact of CD8+ T-Cell Activation on CD4+ T-Cell Recovery and Mortality in HIV-Infected Ugandans Initiating Antiretroviral Therapy. AIDS 2011, 25, 2123–2131. [Google Scholar] [CrossRef] [PubMed]
  189. Hsue, P.Y.; Hunt, P.W.; Sinclair, E.; Bredt, B.; Franklin, A.; Killian, M.; Hoh, R.; Martin, J.N.; McCune, J.M.; Waters, D.D.; et al. Increased Carotid Intima-Media Thickness in HIV Patients Is Associated with Increased Cytomegalovirus-Specific T-Cell Responses. AIDS 2006, 20, 2275–2283. [Google Scholar] [CrossRef]
  190. Kaplan, R.C.; Sinclair, E.; Landay, A.L.; Lurain, N.; Sharrett, A.R.; Gange, S.J.; Xue, X.; Hunt, P.; Karim, R.; Kern, D.M.; et al. T Cell Activation and Senescence Predict Subclinical Carotid Artery Disease in HIV-Infected Women. J. Infect. Dis. 2011, 203, 452–463. [Google Scholar] [CrossRef]
  191. Lichtenstein, K.A.; Armon, C.; Buchacz, K.; Chmiel, J.S.; Buckner, K.; Tedaldi, E.M.; Wood, K.; Holmberg, S.D.; Brooks, J.T.; HIV Outpatient Study (HOPS) Investigators. Low CD4+ T Cell Count Is a Risk Factor for Cardiovascular Disease Events in the HIV Outpatient Study. Clin. Infect. Dis. 2010, 51, 435–447. [Google Scholar] [CrossRef]
  192. Kaplan, R.C.; Kingsley, L.A.; Gange, S.J.; Benning, L.; Jacobson, L.P.; Lazar, J.; Anastos, K.; Tien, P.C.; Sharrett, A.R.; Hodis, H.N. Low CD4+ T-Cell Count as a Major Atherosclerosis Risk Factor in HIV-Infected Women and Men. AIDS 2008, 22, 1615–1624. [Google Scholar] [CrossRef]
  193. Freiberg, M.S.; Chang, C.-C.H.; Kuller, L.H.; Skanderson, M.; Lowy, E.; Kraemer, K.L.; Butt, A.A.; Bidwell Goetz, M.; Leaf, D.; Oursler, K.A.; et al. HIV Infection and the Risk of Acute Myocardial Infarction. JAMA Intern. Med. 2013, 173, 614–622. [Google Scholar] [CrossRef] [PubMed]
  194. Ho, J.E.; Deeks, S.G.; Hecht, F.M.; Xie, Y.; Schnell, A.; Martin, J.N.; Ganz, P.; Hsue, P.Y. Initiation of Antiretroviral Therapy at Higher Nadir CD4+ T-Cell Counts Is Associated with Reduced Arterial Stiffness in HIV-Infected Individuals. AIDS 2010, 24, 1897–1905. [Google Scholar] [CrossRef] [PubMed]
  195. Wang, Y.; Li, W.; Zhao, T.; Zou, Y.; Deng, T.; Yang, Z.; Yuan, Z.; Ma, L.; Yu, R.; Wang, T.; et al. Interleukin-17-Producing CD4+ T Cells Promote Inflammatory Response and Foster Disease Progression in Hyperlipidemic Patients and Atherosclerotic Mice. Front. Cardiovasc. Med. 2021, 8, 667768. [Google Scholar] [CrossRef]
  196. Kundu, S.; Freiberg, M.S.; Tracy, R.P.; So-Armah, K.A.; Koethe, J.R.; Duncan, M.S.; Tindle, H.A.; Beckman, J.A.; Feinstein, M.J.; McDonnell, W.J.; et al. Circulating T Cells and Cardiovascular Risk in People with and without HIV Infection. J. Am. Coll. Cardiol. 2022, 80, 1633–1644. [Google Scholar] [CrossRef]
  197. Tse, K.; Tse, H.; Sidney, J.; Sette, A.; Ley, K. T Cells in Atherosclerosis. Int. Immunol. 2013, 25, 615–622. [Google Scholar] [CrossRef]
  198. Taleb, S.; Tedgui, A.; Mallat, Z. Adaptive T Cell Immune Responses and Atherogenesis. Curr. Opin. Pharmacol. 2010, 10, 197–202. [Google Scholar] [CrossRef] [PubMed]
  199. Saigusa, R.; Winkels, H.; Ley, K. T Cell Subsets and Functions in Atherosclerosis. Nat. Rev. Cardiol. 2020, 17, 387–401. [Google Scholar] [CrossRef]
  200. Wiche Salinas, T.R.; Zhang, Y.; Gosselin, A.; Rosario, N.F.; El-Far, M.; Filali-Mouhim, A.; Routy, J.-P.; Chartrand-Lefebvre, C.; Landay, A.L.; Durand, M.; et al. Alterations in Th17 Cells and Non-Classical Monocytes as a Signature of Subclinical Coronary Artery Atherosclerosis during ART-Treated HIV-1 Infection. Cells 2024, 13, 157. [Google Scholar] [CrossRef]
  201. Yero, A.; Shi, T.; Farnos, O.; Routy, J.-P.; Tremblay, C.; Durand, M.; Tsoukas, C.; Costiniuk, C.T.; Jenabian, M.-A. Dynamics and Epigenetic Signature of Regulatory T-Cells Following Antiretroviral Therapy Initiation in Acute HIV Infection. EBioMedicine 2021, 71, 103570. [Google Scholar] [CrossRef]
  202. Rothan, C.; Yero, A.; Shi, T.; Farnos, O.; Chartrand-Lefebvre, C.; El-Far, M.; Costiniuk, C.T.; Tsoukas, C.; Tremblay, C.; Durand, M.; et al. Antiretroviral Therapy-Treated HIV-Infected Adults with Coronary Artery Disease Are Characterized by a Distinctive Regulatory T-Cell Signature. AIDS 2021, 35, 1003–1014. [Google Scholar] [CrossRef]
  203. Chen, B.; Morris, S.R.; Panigrahi, S.; Michaelson, G.M.; Wyrick, J.M.; Komissarov, A.A.; Potashnikova, D.; Lebedeva, A.; Younes, S.-A.; Harth, K.; et al. Cytomegalovirus Coinfection Is Associated with Increased Vascular-Homing CD57+ CD4 T Cells in HIV Infection. J. Immunol. 2020, 204, 2722–2733. [Google Scholar] [CrossRef]
  204. Wanjalla, C.N.; Mashayekhi, M.; Bailin, S.; Gabriel, C.L.; Meenderink, L.M.; Temu, T.; Fuller, D.T.; Guo, L.; Kawai, K.; Virmani, R.; et al. Anticytomegalovirus CD4+ T Cells Are Associated with Subclinical Atherosclerosis in Persons with HIV. Arterioscler. Thromb. Vasc. Biol. 2021, 41, 1459–1473. [Google Scholar] [CrossRef]
  205. Wanjalla, C.N.; Gabriel, C.L.; Fuseini, H.; Bailin, S.S.; Mashayekhi, M.; Simmons, J.; Warren, C.M.; Glass, D.R.; Oakes, J.; Gangula, R.; et al. CD4+ T Cells Expressing CX3CR1, GPR56, with Variable CD57 Are Associated with Cardiometabolic Diseases in Persons with HIV. Front. Immunol. 2023, 14, 1099356. [Google Scholar] [CrossRef]
  206. Abana, C.O.; Pilkinton, M.A.; Gaudieri, S.; Chopra, A.; McDonnell, W.J.; Wanjalla, C.; Barnett, L.; Gangula, R.; Hager, C.; Jung, D.K.; et al. Cytomegalovirus (CMV) Epitope-Specific CD4+ T Cells Are Inflated in HIV+ CMV+ Subjects. J. Immunol. 2017, 199, 3187–3201. [Google Scholar] [CrossRef]
  207. Bernal, E.; Martinez, M.; Torres, A.; Guillamón, C.F.; Alcaraz, A.; Alcaraz, M.J.; Muñoz, A.; Valero, S.; Botella, C.; Campillo, J.A.; et al. T Cell Senescence Predicts Subclinical Atherosclerosis in HIV-Infected Patients Similarly to Traditional Cardiovascular Risk Factors. Antiviral Res. 2019, 162, 163–170. [Google Scholar] [CrossRef]
  208. Pera, A.; Caserta, S.; Albanese, F.; Blowers, P.; Morrow, G.; Terrazzini, N.; Smith, H.E.; Rajkumar, C.; Reus, B.; Msonda, J.R.; et al. CD28null Pro-Atherogenic CD4 T-Cells Explain the Link between CMV Infection and an Increased Risk of Cardiovascular Death. Theranostics 2018, 8, 4509–4519. [Google Scholar] [CrossRef]
  209. Liuzzo, G.; Biasucci, L.M.; Trotta, G.; Brugaletta, S.; Pinnelli, M.; Digianuario, G.; Rizzello, V.; Rebuzzi, A.G.; Rumi, C.; Maseri, A.; et al. Unusual CD4+CD28null T Lymphocytes and Recurrence of Acute Coronary Events. J. Am. Coll. Cardiol. 2007, 50, 1450–1458. [Google Scholar] [CrossRef]
  210. Zal, B.; Kaski, J.C.; Arno, G.; Akiyu, J.P.; Xu, Q.; Cole, D.; Whelan, M.; Russell, N.; Madrigal, J.A.; Dodi, I.A.; et al. Heat-Shock Protein 60-Reactive CD4+CD28null T Cells in Patients with Acute Coronary Syndromes. Circulation 2004, 109, 1230–1235. [Google Scholar] [CrossRef]
  211. Bailin, S.S.; Kundu, S.; Wellons, M.; Freiberg, M.S.; Doyle, M.F.; Tracy, R.P.; Justice, A.C.; Wanjalla, C.N.; Landay, A.L.; So-Armah, K.; et al. Circulating CD4+ TEMRA and CD4+ CD28- T Cells and Incident Diabetes among Persons with and without HIV. AIDS 2022, 36, 501–511. [Google Scholar] [CrossRef] [PubMed]
  212. Serrano-Villar, S.; Pérez-Elías, M.J.; Dronda, F.; Casado, J.L.; Moreno, A.; Royuela, A.; Pérez-Molina, J.A.; Sainz, T.; Navas, E.; Hermida, J.M.; et al. Increased Risk of Serious Non-AIDS-Related Events in HIV-Infected Subjects on Antiretroviral Therapy Associated with a Low CD4/CD8 Ratio. PLoS ONE 2014, 9, e85798. [Google Scholar] [CrossRef] [PubMed]
  213. Kulkarni, M.; Bowman, E.; Gabriel, J.; Amburgy, T.; Mayne, E.; Zidar, D.A.; Maierhofer, C.; Turner, A.N.; Bazan, J.A.; Koletar, S.L.; et al. Altered Monocyte and Endothelial Cell Adhesion Molecule Expression Is Linked to Vascular Inflammation in Human Immunodeficiency Virus Infection. Open Forum Infect. Dis. 2016, 3, ofw224. [Google Scholar] [CrossRef] [PubMed]
  214. McDermott, D.H.; Halcox, J.P.; Schenke, W.H.; Waclawiw, M.A.; Merrell, M.N.; Epstein, N.; Quyyumi, A.A.; Murphy, P.M. Association between Polymorphism in the Chemokine Receptor CX3CR1 and Coronary Vascular Endothelial Dysfunction and Atherosclerosis. Circ. Res. 2001, 89, 401–407. [Google Scholar] [CrossRef] [PubMed]
  215. Moatti, D.; Faure, S.; Fumeron, F.; Amara, M.E.-W.; Seknadji, P.; McDermott, D.H.; Debré, P.; Aumont, M.C.; Murphy, P.M.; de Prost, D.; et al. Polymorphism in the Fractalkine Receptor CX3CR1 as a Genetic Risk Factor for Coronary Artery Disease. Blood 2001, 97, 1925–1928. [Google Scholar] [CrossRef] [PubMed]
  216. Ikejima, H.; Imanishi, T.; Tsujioka, H.; Kashiwagi, M.; Kuroi, A.; Tanimoto, T.; Kitabata, H.; Ishibashi, K.; Komukai, K.; Takeshita, T.; et al. Upregulation of Fractalkine and Its Receptor, CX3CR1, Is Associated with Coronary Plaque Rupture in Patients with Unstable Angina Pectoris. Circ. J. 2010, 74, 337–345. [Google Scholar] [CrossRef] [PubMed]
  217. Panigrahi, S.; Freeman, M.L.; Funderburg, N.T.; Mudd, J.C.; Younes, S.A.; Sieg, S.F.; Zidar, D.A.; Paiardini, M.; Villinger, F.; Calabrese, L.H.; et al. SIV/SHIV Infection Triggers Vascular Inflammation, Diminished Expression of Krüppel-like Factor 2 and Endothelial Dysfunction. J. Infect. Dis. 2016, 213, 1419–1427. [Google Scholar] [CrossRef] [PubMed]
  218. Panigrahi, S.; Chen, B.; Fang, M.; Potashnikova, D.; Komissarov, A.A.; Lebedeva, A.; Michaelson, G.M.; Wyrick, J.M.; Morris, S.R.; Sieg, S.F.; et al. CX3CL1 and IL-15 Promote CD8 T Cell Chemoattraction in HIV and in Atherosclerosis. PLoS Pathog. 2020, 16, e1008885. [Google Scholar] [CrossRef] [PubMed]
  219. Grivel, J.-C.; Ivanova, O.; Pinegina, N.; Blank, P.S.; Shpektor, A.; Margolis, L.B.; Vasilieva, E. Activation of T Lymphocytes in Atherosclerotic Plaques. Arterioscler. Thromb. Vasc. Biol. 2011, 31, 2929–2937. [Google Scholar] [CrossRef]
  220. Freeman, M.L.; Panigrahi, S.; Chen, B.; Juchnowski, S.; Sieg, S.F.; Lederman, M.M.; Funderburg, N.T.; Zidar, D.A. CD8+ T-Cell-Derived Tumor Necrosis Factor Can Induce Tissue Factor Expression on Monocytes. J. Infect. Dis. 2019, 220, 73–77. [Google Scholar] [CrossRef]
  221. van Duijn, J.; Kritikou, E.; Benne, N.; van der Heijden, T.; van Puijvelde, G.H.; Kröner, M.J.; Schaftenaar, F.H.; Foks, A.C.; Wezel, A.; Smeets, H.; et al. CD8+ T-Cells Contribute to Lesion Stabilization in Advanced Atherosclerosis by Limiting Macrophage Content and CD4+ T-Cell Responses. Cardiovasc. Res. 2019, 115, 729–738. [Google Scholar] [CrossRef]
  222. Adamo, L.; Rocha-Resende, C.; Mann, D.L. The Emerging Role of B Lymphocytes in Cardiovascular Disease. Annu. Rev. Immunol. 2020, 38, 99–121. [Google Scholar] [CrossRef]
  223. Pattarabanjird, T.; Li, C.; McNamara, C. B Cells in Atherosclerosis: Mechanisms and Potential Clinical Applications. JACC Basic Transl. Sci. 2021, 6, 546–563. [Google Scholar] [CrossRef]
  224. Obare, L.M.; Bonami, R.H.; Doran, A.C.; Wanjalla, C.N. B Cells and Atherosclerosis: A HIV Perspective. J. Cell. Physiol. 2024, 239, 31270. [Google Scholar] [CrossRef]
  225. LeBien, T.W.; Tedder, T.F. B Lymphocytes: How They Develop and Function. Blood 2008, 112, 1570–1580. [Google Scholar] [CrossRef]
  226. Moir, S.; Fauci, A.S. B Cells in HIV Infection and Disease. Nat. Rev. Immunol. 2009, 9, 235–245. [Google Scholar] [CrossRef]
  227. Moir, S.; Ho, J.; Malaspina, A.; Wang, W.; DiPoto, A.C.; O’Shea, M.A.; Roby, G.; Kottilil, S.; Arthos, J.; Proschan, M.A.; et al. Evidence for HIV-Associated B Cell Exhaustion in a Dysfunctional Memory B Cell Compartment in HIV-Infected Viremic Individuals. J. Exp. Med. 2008, 205, 1797–1805. [Google Scholar] [CrossRef]
  228. Pillai, S.; Cariappa, A. The Follicular versus Marginal Zone B Lymphocyte Cell Fate Decision. Nat. Rev. Immunol. 2009, 9, 767–777. [Google Scholar] [CrossRef]
  229. 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. [Google Scholar] [CrossRef]
  230. 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. [Google Scholar] [CrossRef]
  231. 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. [Google Scholar] [CrossRef] [PubMed]
  232. 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. [Google Scholar] [CrossRef] [PubMed]
  233. Kyaw, T.; Tay, C.; Hosseini, H.; Kanellakis, P.; Gadowski, T.; MacKay, F.; Tipping, P.; Bobik, A.; Toh, B.-H. Depletion of B2 but Not B1a B Cells in BAFF Receptor-Deficient ApoE Mice Attenuates Atherosclerosis by Potently Ameliorating Arterial Inflammation. PLoS ONE 2012, 7, e29371. [Google Scholar] [CrossRef]
  234. Samson, S.; Mundkur, L.; Kakkar, V.V. Immune Response to Lipoproteins in Atherosclerosis. Cholesterol 2012, 2012, 571846. [Google Scholar] [CrossRef] [PubMed]
  235. Liu, Y.; Li, X.; Han, Y.; Qiu, Z.; Song, X.; Li, B.; Zhang, H.; Wang, H.; Feng, K.; Liu, L.; et al. High APRIL Levels Are Associated with Slow Disease Progression and Low Immune Activation in Chronic HIV-1-Infected Patients. Front. Med. 2020, 7, 299. [Google Scholar] [CrossRef]
  236. Tsiantoulas, D.; Sage, A.P.; Göderle, L.; Ozsvar-Kozma, M.; Murphy, D.; Porsch, F.; Pasterkamp, G.; Menche, J.; Schneider, P.; Mallat, Z.; et al. B Cell-Activating Factor Neutralization Aggravates Atherosclerosis. Circulation 2018, 138, 2263–2273. [Google Scholar] [CrossRef] [PubMed]
  237. Ibrahim, A.I.; Obeid, M.T.; Jouma, M.J.; Moasis, G.A.; Al-Richane, W.L.; Kindermann, I.; Boehm, M.; Roemer, K.; Mueller-Lantzsch, N.; Gärtner, B.C. Detection of Herpes Simplex Virus, Cytomegalovirus and Epstein-Barr Virus DNA in Atherosclerotic Plaques and in Unaffected Bypass Grafts. J. Clin. Virol. 2005, 32, 29–32. [Google Scholar] [CrossRef]
  238. Hechter, R.C.; Budoff, M.; Hodis, H.N.; Rinaldo, C.R.; Jenkins, F.J.; Jacobson, L.P.; Kingsley, L.A.; Taiwo, B.; Post, W.S.; Margolick, J.B.; et al. Herpes Simplex Virus Type 2 (HSV-2) as a Coronary Atherosclerosis Risk Factor in HIV-Infected Men: Multicenter AIDS Cohort Study. Atherosclerosis 2012, 223, 433–436. [Google Scholar] [CrossRef] [PubMed]
  239. Masiá, M.; Robledano, C.; Ortiz de la Tabla, V.; Antequera, P.; López, N.; Gutiérrez, F. Increased Carotid Intima-Media Thickness Associated with Antibody Responses to Varicella-Zoster Virus and Cytomegalovirus in HIV-Infected Patients. PLoS ONE 2013, 8, e64327. [Google Scholar] [CrossRef]
  240. Castellon, X.; Bogdanova, V. Chronic Inflammatory Diseases and Endothelial Dysfunction. Aging Dis. 2016, 7, 81–89. [Google Scholar] [CrossRef]
  241. Weis, M.; Kledal, T.N.; Lin, K.Y.; Panchal, S.N.; Gao, S.Z.; Valantine, H.A.; Mocarski, E.S.; Cooke, J.P. Cytomegalovirus Infection Impairs the Nitric Oxide Synthase Pathway: Role of Asymmetric Dimethylarginine in Transplant Arteriosclerosis. Circulation 2004, 109, 500–505. [Google Scholar] [CrossRef] [PubMed]
  242. van den Berg, S.P.H.; Pardieck, I.N.; Lanfermeijer, J.; Sauce, D.; Klenerman, P.; van Baarle, D.; Arens, R. The Hallmarks of CMV-Specific CD8 T-Cell Differentiation. Med. Microbiol. Immunol. 2019, 208, 365–373. [Google Scholar] [CrossRef] [PubMed]
  243. Heybar, H.; Alavi, S.M.; Farashahi Nejad, M.; Latifi, M. Cytomegalovirus Infection and Atherosclerosis in Candidate of Coronary Artery Bypass Graft. Jundishapur J. Microbiol. 2015, 8, e15476. [Google Scholar] [CrossRef] [PubMed]
  244. Shi, Y.; Tokunaga, O. Herpesvirus (HSV-1, EBV and CMV) Infections in Atherosclerotic Compared with Non-Atherosclerotic Aortic Tissue. Pathol. Int. 2002, 52, 31–39. [Google Scholar] [CrossRef] [PubMed]
  245. Hunt, P.W.; Martin, J.N.; Sinclair, E.; Epling, L.; Teague, J.; Jacobson, M.A.; Tracy, R.P.; Corey, L.; Deeks, S.G. Valganciclovir Reduces T Cell Activation in HIV-Infected Individuals with Incomplete CD4+ T Cell Recovery on Antiretroviral Therapy. J. Infect. Dis. 2011, 203, 1474–1483. [Google Scholar] [CrossRef] [PubMed]
  246. Wanjalla, C.N.; McDonnell, W.J.; Barnett, L.; Simmons, J.D.; Furch, B.D.; Lima, M.C.; Woodward, B.O.; Fan, R.; Fei, Y.; Baker, P.G.; et al. Adipose Tissue in Persons with HIV Is Enriched for CD4+ T Effector Memory and T Effector Memory RA+ Cells, Which Show Higher CD69 Expression and CD57, CX3CR1, GPR56 Co-Expression with Increasing Glucose Intolerance. Front. Immunol. 2019, 10, 408. [Google Scholar] [CrossRef]
  247. National Institute of Allergy and Infectious Diseases (NIAID). Randomized, Controlled Trial to Evaluate the Anti-Inflammatory Efficacy of Letermovir (Prevymis) in Adults with Human Immunodeficiency Virus (HIV)-1 and Asymptomatic Cytomegalovirus (CMV) Who Are on Suppressive ART and Its Effect on Chronic Inflammation, HIV Persistence, and Other Clinical Outcomes; NIAID: Bethesda, MD, USA, 2024. [Google Scholar]
  248. Weibel, S.G.; Kitch, D.W.; Beck-Engeser, G.B.; Meneses, M.; Letendre, S.L.; Dube, M.; Fox, L.; Koethe, J.; Hsue, P.Y.; Tawakol, A.A.; et al. Suppressing Asymptomatic CMV with Letermovir Reshapes Cardiometabolic Proteome in Treated HIV. In Proceedings of the Conference on Retroviruses and Opportunistic Infections, Denver, CO, USA, 3–6 March 2024. [Google Scholar]
  249. De Backer, J.; Mak, R.; De Bacquer, D.; Van Renterghem, L.; Verbraekel, E.; Kornitzer, M.; De Backer, G. Parameters of Inflammation and Infection in a Community Based Case-Control Study of Coronary Heart Disease. Atherosclerosis 2002, 160, 457–463. [Google Scholar] [CrossRef]
  250. Lobzin, I.V.; Boĭtsov, S.A.; Filippov, A.E.; Linchak, R.M.; Mangutov, D.A. Effect of respiratory infections on the clinical course of coronary artery disease. Klin. Med. 2005, 83, 22–26. [Google Scholar]
  251. Papagno, L.; Spina, C.A.; Marchant, A.; Salio, M.; Rufer, N.; Little, S.; Dong, T.; Chesney, G.; Waters, A.; Easterbrook, P.; et al. Immune Activation and CD8+ T-Cell Differentiation towards Senescence in HIV-1 Infection. PLoS Biol. 2004, 2, E20. [Google Scholar] [CrossRef]
  252. Sutherland, M.R.; Raynor, C.M.; Leenknegt, H.; Wright, J.F.; Pryzdial, E.L. Coagulation Initiated on Herpesviruses. Proc. Natl. Acad. Sci. USA 1997, 94, 13510–13514. [Google Scholar] [CrossRef]
  253. Grahame-Clarke, C.; Alber, D.G.; Lucas, S.B.; Miller, R.; Vallance, P. Association between Kaposi’s Sarcoma and Atherosclerosis: Implications for Gammaherpesviruses and Vascular Disease. AIDS 2001, 15, 1902–1904. [Google Scholar] [CrossRef]
  254. Masiá, M.; Robledano, C.; Ortiz de la Tabla, V.; Antequera, P.; Lumbreras, B.; Hernández, I.; Gutiérrez, F. Coinfection with Human Herpesvirus 8 Is Associated with Persistent Inflammation and Immune Activation in Virologically Suppressed HIV-Infected Patients. PLoS ONE 2014, 9, e105442. [Google Scholar] [CrossRef] [PubMed]
  255. Lidón, F.; Padilla, S.; García, J.A.; Fernández, M.; García, J.; Ortiz de la Tabla, V.; Gutiérrez, F.; Masiá, M. Contribution of Human Herpesvirus 8 and Herpes Simplex Type 2 to Progression of Carotid Intima-Media Thickness in People Living with HIV. Open Forum Infect. Dis. 2019, 6, ofz041. [Google Scholar] [CrossRef] [PubMed]
  256. Curhan, S.G.; Kawai, K.; Yawn, B.; Rexrode, K.M.; Rimm, E.B.; Curhan, G.C. Herpes Zoster and Long-Term Risk of Cardiovascular Disease. J. Am. Heart Assoc. 2022, 11, e027451. [Google Scholar] [CrossRef]
  257. Nagel, M.A.; Bubak, A.N. Varicella Zoster Virus Vasculopathy. J. Infect. Dis. 2018, 218, S107–S112. [Google Scholar] [CrossRef] [PubMed]
  258. Yawn, B.P.; Wollan, P.C.; Nagel, M.A.; Gilden, D. Risk of Stroke and Myocardial Infarction After Herpes Zoster in Older Adults in a US Community Population. Mayo Clin. Proc. 2016, 91, 33–44. [Google Scholar] [CrossRef] [PubMed]
  259. Ba’Omar, M.; Chhetri, S.; Pandak, N.; Khamis, F.; Al-Balushi, Z.; Al-Hajri, A.; Abouelhamd, H.; Al-Fahdi, Z. Varicella Zoster Virus Vasculopathy; An HIV Adult Presenting with Multiple Strokes. IDCases 2022, 30, e01641. [Google Scholar] [CrossRef] [PubMed]
  260. Marais, G.; Naidoo, M.; McMullen, K.; Stanley, A.; Bryer, A.; van der Westhuizen, D.; Bateman, K.; Hardie, D.R. Varicella-Zoster Virus Reactivation Is Frequently Detected in HIV-Infected Individuals Presenting with Stroke. J. Med. Virol. 2022, 94, 2675–2683. [Google Scholar] [CrossRef] [PubMed]
  261. Ku, H.-C.; Tsai, Y.-T.; Konara-Mudiyanselage, S.-P.; Wu, Y.-L.; Yu, T.; Ko, N.-Y. Incidence of Herpes Zoster in HIV-Infected Patients Undergoing Antiretroviral Therapy: A Systematic Review and Meta-Analysis. J. Clin. Med. 2021, 10, 2300. [Google Scholar] [CrossRef]
  262. Fiordelisi, D.; Poliseno, M.; De Gennaro, N.; Milano, E.; Santoro, C.R.; Segala, F.V.; Franco, C.F.; Manco Cesari, G.; Frallonardo, L.; Guido, G.; et al. Varicella-Zoster Virus Reactivation and Increased Vascular Risk in People Living with HIV: Data from a Retrospective Cohort Study. Viruses 2023, 15, 2217. [Google Scholar] [CrossRef]
  263. Wang, J.P.; Kurt-Jones, E.A.; Shin, O.S.; Manchak, M.D.; Levin, M.J.; Finberg, R.W. Varicella-Zoster Virus Activates Inflammatory Cytokines in Human Monocytes and Macrophages via Toll-like Receptor 2. J. Virol. 2005, 79, 12658–12666. [Google Scholar] [CrossRef] [PubMed]
  264. Eleftheriou, D.; Moraitis, E.; Hong, Y.; Turmaine, M.; Venturini, C.; Ganesan, V.; Breuer, J.; Klein, N.; Brogan, P. Microparticle-Mediated VZV Propagation and Endothelial Activation: Mechanism of VZV Vasculopathy. Neurology 2020, 94, e474–e480. [Google Scholar] [CrossRef] [PubMed]
  265. Petruzziello, A.; Marigliano, S.; Loquercio, G.; Cozzolino, A.; Cacciapuoti, C. Global Epidemiology of Hepatitis C Virus Infection: An up-Date of the Distribution and Circulation of Hepatitis C Virus Genotypes. World J. Gastroenterol. 2016, 22, 7824–7840. [Google Scholar] [CrossRef] [PubMed]
  266. André, P.; Komurian-Pradel, F.; Deforges, S.; Perret, M.; Berland, J.L.; Sodoyer, M.; Pol, S.; Bréchot, C.; Paranhos-Baccalà, G.; Lotteau, V. Characterization of Low- and Very-Low-Density Hepatitis C Virus RNA-Containing Particles. J. Virol. 2002, 76, 6919–6928. [Google Scholar] [CrossRef] [PubMed]
  267. Lee, K.K.; Stelzle, D.; Bing, R.; Anwar, M.; Strachan, F.; Bashir, S.; Newby, D.E.; Shah, J.S.; Chung, M.H.; Bloomfield, G.S.; et al. Global Burden of Atherosclerotic Cardiovascular Disease in People with Hepatitis C Virus Infection: A Systematic Review, Meta-Analysis, and Modelling Study. Lancet Gastroenterol. Hepatol. 2019, 4, 794–804. [Google Scholar] [CrossRef] [PubMed]
  268. Chew, K.W.; Hua, L.; Bhattacharya, D.; Butt, A.A.; Bornfleth, L.; Chung, R.T.; Andersen, J.W.; Currier, J.S. The Effect of Hepatitis C Virologic Clearance on Cardiovascular Disease Biomarkers in Human Immunodeficiency Virus/Hepatitis C Virus Coinfection. Open Forum Infect. Dis. 2014, 1, ofu104. [Google Scholar] [CrossRef]
  269. Babiker, A.; Jeudy, J.; Kligerman, S.; Khambaty, M.; Shah, A.; Bagchi, S. Risk of Cardiovascular Disease Due to Chronic Hepatitis C Infection: A Review. J. Clin. Transl. Hepatol. 2017, 5, 343–362. [Google Scholar] [CrossRef]
  270. Roed, T.; Kristoffersen, U.S.; Knudsen, A.; Wiinberg, N.; Lebech, A.-M.; Almdal, T.; Thomsen, R.W.; Kjær, A.; Weis, N. Increased Prevalence of Coronary Artery Disease Risk Markers in Patients with Chronic Hepatitis C—A Cross-Sectional Study. Vasc. Health Risk Manag. 2014, 10, 55–62. [Google Scholar] [CrossRef]
  271. Al-Jiffri, O.H. Adhesive Molecules and Inflammatory Markers among Hepatitis C Virus Saudi Patients. Electron. J. Gen. Med. 2017, 14, 89–93. [Google Scholar] [CrossRef]
  272. Kaplanski, G.; Farnarier, C.; Payan, M.J.; Bongrand, P.; Durand, J.M. Increased Levels of Soluble Adhesion Molecules in the Serum of Patients with Hepatitis C. Correlation with Cytokine Concentrations and Liver Inflammation and Fibrosis. Dig. Dis. Sci. 1997, 42, 2277–2284. [Google Scholar] [CrossRef]
  273. Babiker, A.; Hassan, M.; Muhammed, S.; Taylor, G.; Poonia, B.; Shah, A.; Bagchi, S. Inflammatory and Cardiovascular Diseases Biomarkers in Chronic Hepatitis C Virus Infection: A Review. Clin. Cardiol. 2019, 43, 222–234. [Google Scholar] [CrossRef] [PubMed]
  274. Bagchi, S.; Burrowes, S.A.; Fantry, L.E.; Hossain, M.B.; Tollera, G.H.; Kottilil, S.; Pauza, C.D.; Miller, M.; Baumgarten, M.; Redfield, R.R. Factors Associated with High Cardiovascular Risk in a Primarily African American, Urban HIV-Infected Population. SAGE Open Med. 2017, 5, 2050312117725644. [Google Scholar] [CrossRef] [PubMed]
  275. Freiberg, M.S.; Chang, C.-C.H.; Skanderson, M.; McGinnis, K.; Kuller, L.H.; Kraemer, K.L.; Rimland, D.; Goetz, M.B.; Butt, A.A.; Rodriguez Barradas, M.C.; et al. The Risk of Incident Coronary Heart Disease among Veterans with and without HIV and Hepatitis C. Circ. Cardiovasc. Qual. Outcomes 2011, 4, 425–432. [Google Scholar] [CrossRef] [PubMed]
  276. Kushner, L.E.; Wendelboe, A.M.; Lazzeroni, L.C.; Chary, A.; Winters, M.A.; Osinusi, A.; Kottilil, S.; Polis, M.A.; Holodniy, M. Immune Biomarker Differences and Changes Comparing HCV Mono-Infected, HIV/HCV Co-Infected, and HCV Spontaneously Cleared Patients. PLoS ONE 2013, 8, e60387. [Google Scholar] [CrossRef] [PubMed]
  277. Kohli, P.; Ganz, P.; Ma, Y.; Scherzer, R.; Hur, S.; Weigel, B.; Grunfeld, C.; Deeks, S.; Wasserman, S.; Scott, R.; et al. HIV and Hepatitis C-Coinfected Patients Have Lower Low-Density Lipoprotein Cholesterol Despite Higher Proprotein Convertase Subtilisin Kexin 9 (PCSK9): An Apparent “PCSK9-Lipid Paradox”. J. Am. Heart Assoc. 2016, 5, e002683. [Google Scholar] [CrossRef] [PubMed]
  278. Gandhi, M.M.; Nguyen, K.-L.; Lake, J.E.; Liao, D.; Khodabakhshian, A.; Guerrero, M.; Shufelt, C.L.; Bairey Merz, C.N.; Jordan, W.C.; Daar, E.S.; et al. Proprotein Convertase Subtisilin/Kexin 9 Levels Decline with Hepatitis C Virus Therapy in People with HIV/Hepatitis C Virus and Correlate with Inflammation. AIDS 2024, 38, 317–327. [Google Scholar] [CrossRef]
  279. Carrero, A.; Berenguer, J.; Hontañón, V.; Navarro, J.; Hernández-Quero, J.; Galindo, M.J.; Quereda, C.; Santos, I.; Téllez, M.J.; Ortega, E.; et al. Effects of Eradication of HCV on Cardiovascular Risk and Preclinical Atherosclerosis in HIV/HCV-Coinfected Patients. J. Acquir. Immune Defic. Syndr. 2020, 83, 292–300. [Google Scholar] [CrossRef]
  280. Masiá, M.; Robledano, C.; López, N.; Escolano, C.; Gutiérrez, F. Treatment for Hepatitis C Virus with Pegylated Interferon-α plus Ribavirin Induces Anti-Atherogenic Effects on Cardiovascular Risk Biomarkers in HIV-Infected and -Uninfected Patients. J. Antimicrob. Chemother. 2011, 66, 1861–1868. [Google Scholar] [CrossRef]
Figure 1. HIV-associated atherosclerosis results from a culmination of numerous intersecting pathways. HIV-related factors include viral proteins, viral ssRNA, viral replication, antiretroviral therapy, and the microbiome. Clinical and environmental co-factors include hypertension, dyslipidemia, diabetes mellitus, obesity, maladaptive anthropometry and fat composition, liver inflammation/metabolic dysfunction-associated fatty liver disease (MAFLD), and concurrent substance use. Important co-infections and co-pathogens include cytomegalovirus (CMV), varicella zoster virus (VZV), herpes simplex virus 2 (HSV-2), Epstein–Barr Virus (EBV), and hepatitis C virus (HCV). These HIV-related, clinical, and environmental co-factors and co-infections induce inflammation and immune dysfunction. Perturbations of cytokines, chemokines, and the innate and adaptive immune systems together result in immune dysregulation and inflammation, which lead to athrosclerosis and clinical atherosclerotic cardiovascular disease, cerebrovascular disease, peripheral artery disease, and aortic aneurysms. Abbreviations; ABC = abacavir; CCL = chemokine ligand; CCR = chemokine receptor; CXCL = CXC motif chemokine ligand; CXCR = CXC motif chemokine receptor; CX3CL = C-X3-C motif chemokine ligand; CX3CR = C-X3-C motif chemokine receptor; gp = glycoprotein; cGAS = cyclic GMP-AMP synthetase; HDL = high-density lipoprotein; IFI16 = interferon inducible protein 16; IFN = interferon; IL = interleukin; INSTI = integrase strand transfer inhibitor; LDL = low-density lipoprotein; Lp(a) = lipoprotein (a); LPS = lipopolysaccharide; MAFLD= metabolic-associated fatty liver disease; NK = natural killer; oxLDL = oxidized low-density lipoprotein; PIs = protease inhibitors; ssRNA = single-stranded ribonucleic acid; TAF = tenofovir alafenamide; tat = trans-Activator of transcription; TIA = transient ischemic attack; TLR = toll-like receptorl TNF = tumor necrosis factor.
Figure 1. HIV-associated atherosclerosis results from a culmination of numerous intersecting pathways. HIV-related factors include viral proteins, viral ssRNA, viral replication, antiretroviral therapy, and the microbiome. Clinical and environmental co-factors include hypertension, dyslipidemia, diabetes mellitus, obesity, maladaptive anthropometry and fat composition, liver inflammation/metabolic dysfunction-associated fatty liver disease (MAFLD), and concurrent substance use. Important co-infections and co-pathogens include cytomegalovirus (CMV), varicella zoster virus (VZV), herpes simplex virus 2 (HSV-2), Epstein–Barr Virus (EBV), and hepatitis C virus (HCV). These HIV-related, clinical, and environmental co-factors and co-infections induce inflammation and immune dysfunction. Perturbations of cytokines, chemokines, and the innate and adaptive immune systems together result in immune dysregulation and inflammation, which lead to athrosclerosis and clinical atherosclerotic cardiovascular disease, cerebrovascular disease, peripheral artery disease, and aortic aneurysms. Abbreviations; ABC = abacavir; CCL = chemokine ligand; CCR = chemokine receptor; CXCL = CXC motif chemokine ligand; CXCR = CXC motif chemokine receptor; CX3CL = C-X3-C motif chemokine ligand; CX3CR = C-X3-C motif chemokine receptor; gp = glycoprotein; cGAS = cyclic GMP-AMP synthetase; HDL = high-density lipoprotein; IFI16 = interferon inducible protein 16; IFN = interferon; IL = interleukin; INSTI = integrase strand transfer inhibitor; LDL = low-density lipoprotein; Lp(a) = lipoprotein (a); LPS = lipopolysaccharide; MAFLD= metabolic-associated fatty liver disease; NK = natural killer; oxLDL = oxidized low-density lipoprotein; PIs = protease inhibitors; ssRNA = single-stranded ribonucleic acid; TAF = tenofovir alafenamide; tat = trans-Activator of transcription; TIA = transient ischemic attack; TLR = toll-like receptorl TNF = tumor necrosis factor.
Ijms 25 07266 g001
Figure 2. Innate and adaptive immune cells contribute in multiple ways to atherosclerosis via inflammatory pathways. The left side of the figure shows soluble mediators released by innate and adaptive immune cells that contribute to vascular inflammation. The right side of the figure shows how cells are drawn to, interact with, and damage blood vessel walls directly and contribute to plaque formation and destabilization. Abbreviations: APRIL = a proliferation-inducing ligand; B = B cell, including B1 and B2 subtypes; CD = cluster of differentiation; CX3CR= C-X3-C motif chemokine receptor; DC = dendritic cell; ICAM = intercellular adhesion molecule; IFN= interferon; IL= interleukin; Mφ = macrophage; MDM = monocyte-derived macrophage; NK = natural killer cell; ox-LDL= oxidized low-density lipoprotein; ROS = reactive oxygen species; Tcyt = cytotoxic T cell, including CD4+ and CD8+ subtypes; Th = T helper cell, including Th1, Th2, Th9, and Th17 subtypes; Treg = regulatory T cell.
Figure 2. Innate and adaptive immune cells contribute in multiple ways to atherosclerosis via inflammatory pathways. The left side of the figure shows soluble mediators released by innate and adaptive immune cells that contribute to vascular inflammation. The right side of the figure shows how cells are drawn to, interact with, and damage blood vessel walls directly and contribute to plaque formation and destabilization. Abbreviations: APRIL = a proliferation-inducing ligand; B = B cell, including B1 and B2 subtypes; CD = cluster of differentiation; CX3CR= C-X3-C motif chemokine receptor; DC = dendritic cell; ICAM = intercellular adhesion molecule; IFN= interferon; IL= interleukin; Mφ = macrophage; MDM = monocyte-derived macrophage; NK = natural killer cell; ox-LDL= oxidized low-density lipoprotein; ROS = reactive oxygen species; Tcyt = cytotoxic T cell, including CD4+ and CD8+ subtypes; Th = T helper cell, including Th1, Th2, Th9, and Th17 subtypes; Treg = regulatory T cell.
Ijms 25 07266 g002
Figure 3. Multiple chronic viral infections are linked to atherosclerosis. Specific viruses such as CMV, HSV-2, and VZV have been associated with subclinical atherosclerosis in PLWH. CMV infection contributes to atherosclerosis through activation of endothelial cells, leading to chronic inflammation and endothelial dysfunction. HCV promotes chronic inflammation, immune activation, and direct vascular injury. HCV and CMV promote LDL oxidation. HSV-2 and HHV-8 can induce thrombogenic and atherogenic alterations in cellular structure, contributing to atherosclerosis. VZV infection may induce chronic inflammation and endothelial dysfunction, contributing to atherosclerosis. EBV’s role in HIV-related atherosclerosis remains controversial, with complex and contradictory associations. Abbreviations: CMV= cytomegalovirus; CTL = cytotoxic T lymphocyte; CX3CL1 = chemokine (C-X3-C motif) ligand 1; EBV= Epstein Barr virus; EC = endothelial cell; HCV= hepatitis C virus; HHV-8= human herpesvirus-8; HSV-2= herpes simplex virus-2; LMP-1 = latent membrane protein 1; NF-κb = nuclear factor kappa B; VEGF = vascular endothelial growth factor.
Figure 3. Multiple chronic viral infections are linked to atherosclerosis. Specific viruses such as CMV, HSV-2, and VZV have been associated with subclinical atherosclerosis in PLWH. CMV infection contributes to atherosclerosis through activation of endothelial cells, leading to chronic inflammation and endothelial dysfunction. HCV promotes chronic inflammation, immune activation, and direct vascular injury. HCV and CMV promote LDL oxidation. HSV-2 and HHV-8 can induce thrombogenic and atherogenic alterations in cellular structure, contributing to atherosclerosis. VZV infection may induce chronic inflammation and endothelial dysfunction, contributing to atherosclerosis. EBV’s role in HIV-related atherosclerosis remains controversial, with complex and contradictory associations. Abbreviations: CMV= cytomegalovirus; CTL = cytotoxic T lymphocyte; CX3CL1 = chemokine (C-X3-C motif) ligand 1; EBV= Epstein Barr virus; EC = endothelial cell; HCV= hepatitis C virus; HHV-8= human herpesvirus-8; HSV-2= herpes simplex virus-2; LMP-1 = latent membrane protein 1; NF-κb = nuclear factor kappa B; VEGF = vascular endothelial growth factor.
Ijms 25 07266 g003
Table 1. Current and novel therapeutic interventions targeting the immune system for ASCVD prevention and MACE reduction.
Table 1. Current and novel therapeutic interventions targeting the immune system for ASCVD prevention and MACE reduction.
Therapeutic Interventions Targeting the Immune System That Have Been Assessed for ASCVD Prevention and MACE Reduction
Therapeutic InterventionTarget Cytokine/Cell Line/Inflammatory MarkerMechanism of ActionEvidence in Persons without HIVExperience in Persons with HIVReferences
Canakinumab IL-1βMonoclonal antibody neutralizes IL-1β signaling In the CANTOS trial, interleukin-1β blockade with canakinumab at a dose of 150 mg every 3 months led to a significantly lower rate of recurrent CVD events than placebo, independent of lipid-level lowering.8 weeks of canakinumab treatment significantly lowered inflammatory biomarkers, including hs-CRP and sCD163. Additionally, leukopoiesis as measured by bone marrow FDG-PET signal and arterial inflammation decreased.[100,101]
TocilizumabIL-6 receptor (IL-6R)Monoclonal antibody targets and blocks IL-6R In the ASSAIL-MI randomized control trial (RCT), tocilizumab increased salvage of viable myocardium in persons with acute STEMI. In a small RCT, tocilizumab was found to be safe and decreased IL-6, but significantly increased cholesterol levels.[102,103]
Methotrexate (MTX) IL-1β, IL-6, CRPModulates inflammation by promoting adenosine uptake and inhibiting transmethylation reactions, subsequently reducing inflammatory biomarkers like CRP, IL-6, and TNF-αIn the CIRT trial, among persons with stable atherosclerosis, low-dose MTX did not reduce levels of IL-1β, IL-6, or CRP and did not result in fewer CVD events than placebo.In a small study of PLWH with increased ASCVD risk, low-dose MTX had more adverse safety events than placebo. Additionally, low-dose MTX had no significant effect on endothelial function or inflammatory markers, but was associated with a significant decrease in CD8+ T cells. [104,105]
ColchicineIL-8, NLR3 inflammasome related cytokines (e.g., IL1-β)Diminishes neutrophil inflammatory function and migration In the COLCOT and LoDOCo2 trials,
colchicine led to a significantly lower risk of ischemic CVD events than placebo.
In the COPS Trial, no significant reductions in acute coronary syndromes or non-cardiometabolic stroke were observed, indicating that effects may be dependent on specific study population.
A small study examining colchicine’s effect on coronary endothelial function and serum inflammatory markers in PLWH without coronary artery disease (CAD) did not show improvement on endothelial function or inflammatory markers.[106,107,108,109,110,111]
Statins HMG-CoA reductase enzyme Multiple: inhibits cholesterol synthesis in liver; inhibits NF-κB, an important transcription regulatory protein in inflammatory response; activates NOS gene transcription to stimulate production of nitric oxideUnited States Preventive Services Task Force (USPSTF) recommends statins for primary prevention of ASCVD in adults aged 40 to 75 years with a 10-year ASCVD risk of 10% or greater and a nuanced approach among those with 10-year ASCVD risk of 7.5% to 10% based on multiple randomized trials and meta-analyses of cohort studies demonstrating benefit in reducing ASCVD.REPRIEVE trial found that pitavastatin reduced risk of MACE by 35% compared to placebo in PLWH with low-moderate ASCVD risk. The REPRIEVE mechanistic sub-study demonstrated reduction in markers of vascular inflammation in those who received statin. Similar findings have been reported in smaller observational studies like INTREPID, SATURN-HIV, and ACTG A5255. [80,112,113,114,115,116]
PCSK9 inhibitors LDL receptors Binds to PCSK9 protein preventing its interaction with LDL receptors, and stops their degradation, allowing more uptake of cholesterol from the blood into the liver PCSK9 inhibition with evolocumab with concurrent statin therapy lowered LDL cholesterol levels by an additional 15–20% and reduced the risk of cardiovascular events.In a small study, evocolumab reversed coronary artery dysfunction, but had no effect on reducing markers.[117,118]
Aspirin Platelets Cyclo-oxygenase-1 pathway Based on multiple clinical studies, USPSTF guidelines recommend aspirin use for the primary prevention of CVD events in adults aged 40 to 59 years who have a 10% or greater 10-year CVD risk, with moderate certainty; has a small net benefit. Aspirin treatment for 12 weeks had no major impact on sCD14, IL-6, sCD163, D-dimer, T-cell or monocyte activation, or flow-mediated dilation.[119]
Dipyridamole Platelets Inhibits adenosine re-uptake Older anti-platelet agent not currently recommended for primary prevention of ASCVD. In a small RCT, dipyridamole increased extracellular adenosine levels and decreased T-cell activation significantly among PLWH virally suppressed on ART.[120]
Novel Therapeutic Strategies Targeting the Immune System to Reduce ASCVD and MACE
Therapeutic InterventionTarget Cytokine/Cell Line/Inflammatory MarkerMechanism of ActionEvidence in Persons without HIVExperience in Persons with HIVReferences
IL-17a BlockadeIL-17aInhibit effect of IL-17a on monocyte and macrophage function, chemokine expression, and leukocyte migration Functional blockade of IL-17a in vitro has been shown to prevent atherosclerotic lesion progression and promote plaque stabilization in murine models of ASCVD.
Clinical studies of persons treated for psoriasis with IL-17a antagonists have shown reduction in early indicators of CVD.
Not widely studied.[121,122,123]
IL-18 antagonists IL-18Inhibit effects of IL-18 on monocyte and macrophage activation, thus reducing production of interferon gamma and enhancing Th1 responses Pro-atherogenic effects of IL-18 demonstrated in murine model of atherosclerosis. HIV-infected macrophages had increased foam cell formation, expression of NLRP3 inflammasome components, and production of downstream cytokines including IL-18.[86,87,124]
Maraviroc CCR5 chemokine receptor CCR5 receptor antagonistNot studied for ASCVD prevention or MACE reduction.In non-human primate studies of simian immunodeficiency virus (SIV), maraviroc
led to fewer activated CD163+ macrophages in the heart and preserved diastolic function.
ART intensification studies with maraviroc showed decreased expression of vascular cellular adhesion molecule-1 (VCAM-1) compared to controls.
[125,126]
RilpivirinePlatelet activation markers (P-selectin, CD40 ligand)Inhibits thrombin and ADP-triggered platelet aggregation, degranulation, and activationNot studied for ASCVD prevention or MACE reduction.Rilpivirine was found to prevent the progression of thrombotic CVD by inhibiting thrombin and ADP-triggered platelet aggregation, degranulation, and activation without increasing risk of bleeding in murine models of ASCVD. [127]
Janus kinase (JAK) inhibitors JAK-STAT signaling pathway Inhibits specific enzymes in the JAK family including JAK1, JAK2, JAK3, and TYK2, which are involved in the signaling pathways that regulate hematopoiesis and immune activationTofacitinib has been shown to increase risk of MACE in persons with rheumatoid arthritis (RA) compared to TNF-α antagonists in a RCT. In a smaller study, tofacitinib reduced markers of endothelial dysfunction among persons with RA. Ruxolitinib has been shown to be safe in a phase 2 study and reduced IL-6 and sCD14 in ART-treated PLWH. [128,129,130]
HIV vaccinesDendritic cells (DCs)Inactivated HIV-1 virus introduced to autologous monocyte-derived DCs, leading to
enhanced HIV-1-specific T cell responses
Not studied for ASCVD prevention or MACE reduction. Potential decrease in HIV-mediated chronic inflammation. [131,132,133]
Anti-inflammatory nanoparticles and antigenic peptidesAtheroma-specific dendritic dells Induces an anti-inflammatory milieu facilitated by DCs Targeted delivery of engineered anti-inflammatory nanoparticles and antigenic peptides to atheroma-specific DCs has been shown to reduce atherosclerotic lesions and induce a more anti-inflammatory milieu. Not studied. [134]
Sivelestat Neutrophil elastase inhibitorInhibits neutrophil extracellular trap (NET) formationDelivering sivelestat utilizing a plaque-targeting, neutrophil-hitchhiking liposome (cRGD-SVT-Lipo) reduced CAD plaque area and increased plaque stabilization in a murine model of ASCVD.Not studied. [135]
MCC950 NLRP3 inflammasomeDownregulates NLRP3 pro-inflammatory activity, which is implicated in foam cell formation and pro-inflammatory cytokine production Selective inhibition of NLRP3 with MCC950 downregulates inflammasome pro-inflammatory activity.Inhibition of NLRP3 with MCC950 in HIV-infected macrophages reduced foam cell formation in vitro.[87]
Apolipoprotein A-I mimetic peptidesMacrophagesReduces biomarkers of macrophage activationSmall and short-term early stage trials, mostly based on imaging endpoints, have shown a favorable effect of rapidly reducing atherosclerotic plaque volume. However, larger clinical studies have failed to show significant benefit. Apolipoprotein A-I mimetic peptides reduced biomarkers of macrophage activation in a murine model of treated HIV. [136,137]
Methylglyoxal-bis-guanylhydrazone (MGBG)MacrophagesInhibits biosynthesis of polyamines, which are necessary for macrophage activation, proliferation and differentiationNot extensively studied for ASCVD prevention or MACE reduction. In a nonhuman primate model of SIV, MGBG has been demonstrated to reduce cardiovascular inflammation, carotid artery intima-media thickness, and fibrosis.[138]
Natalizumab α4 integrinMonoclonal antibody against α4 integrin decreases macrophage traffic to cardiac tissuePersons receiving natalizumab for treatment of multiple sclerosis have beneficial changes in lipid profiles and antioxidant uric acid levels.Decreased trafficking of macrophages to cardiac tissue and decreased cardiac pathology in nonhuman primate model of SIV.[139,140]
B cell activation factors APRIL (a proliferation-inducing ligand), BAFF (B-cell activating factor)Reduces B cell activation and associated cytokines Not extensively studied for ASCVD prevention or MACE reduction. Increased levels of APRIL have been associated with the development of atherosclerosis. Increased levels of BAFF are associated with increased risk of ASCVD. [141,142]
ART = antiretroviral therapy; ASCVD = atherosclerotic cardiovascular disease; BAFF = B-cell activating factor; CD = cluster of differentiation; COX = cyclooxygenase; CVD = cardiovascular disease; DC = dendritic cells; FDG-PET = 18-fluoro-deoxyglucose positron emission tomography; HIV = human immunodeficiency virus; HMG-CoA = β-hydroxy β-methylglutamyl coenzyme A; hs-CRP = high-sensitivity C-reactive protein; JAK-STAT = janus-kinase/signal transducers and activators of transcription; LDL = low-density lipoprotein; MACE = major adverse cardiovascular events; MTX = methotrexate; NET = neutrophil extracellular traps; NF-kappa β = nuclear factor kappa-beta; NLRP3 = nucleotide-binding domain, leucine-rich-containing family pyrin domain-containing 3; NOS = nitric oxide synthetase; PBMCs = peripheral blood mononuclear cells; PCSK9 = proprotein convertase subtilisin/kexin type 9; PLWH = persons living with HIV; RA = rheumatoid arthritis; RCT = randomized controlled trial; sCD = soluble cluster of differentiation; SIV = simian immunodeficiency virus; TNF = tumor necrosis factor; USPSTF = United States Preventive Services Task Force; VCAM-1 = vascular adhesion molecule-1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hmiel, L.; Zhang, S.; Obare, L.M.; Santana, M.A.d.O.; Wanjalla, C.N.; Titanji, B.K.; Hileman, C.O.; Bagchi, S. Inflammatory and Immune Mechanisms for Atherosclerotic Cardiovascular Disease in HIV. Int. J. Mol. Sci. 2024, 25, 7266. https://doi.org/10.3390/ijms25137266

AMA Style

Hmiel L, Zhang S, Obare LM, Santana MAdO, Wanjalla CN, Titanji BK, Hileman CO, Bagchi S. Inflammatory and Immune Mechanisms for Atherosclerotic Cardiovascular Disease in HIV. International Journal of Molecular Sciences. 2024; 25(13):7266. https://doi.org/10.3390/ijms25137266

Chicago/Turabian Style

Hmiel, Laura, Suyu Zhang, Laventa M. Obare, Marcela Araujo de Oliveira Santana, Celestine N. Wanjalla, Boghuma K. Titanji, Corrilynn O. Hileman, and Shashwatee Bagchi. 2024. "Inflammatory and Immune Mechanisms for Atherosclerotic Cardiovascular Disease in HIV" International Journal of Molecular Sciences 25, no. 13: 7266. https://doi.org/10.3390/ijms25137266

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

Article Metrics

Back to TopTop