Next Article in Journal
Applications of Probiotics and Their Potential Health Benefits
Previous Article in Journal
Heterogeneity of Phase II Enzyme Ligands on Controlling the Progression of Human Gastric Cancer Organoids as Stem Cell Therapy Model
Previous Article in Special Issue
scTIGER: A Deep-Learning Method for Inferring Gene Regulatory Networks from Case versus Control scRNA-seq Datasets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Independent and Interactive Roles of Immunity and Metabolism in Aortic Dissection

1
School of Medicine, Chongqing University, Chongqing 400044, China
2
Department of Biochemistry and Molecular Biology, Third Military Medical University (Army Medical University), Chongqing 400038, China
3
Department of Cardiac Surgery, The First Affiliated Hospital of Third Military Medical University (Army Medical University), Chongqing 400038, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(21), 15908; https://doi.org/10.3390/ijms242115908
Submission received: 7 October 2023 / Revised: 22 October 2023 / Accepted: 27 October 2023 / Published: 2 November 2023

Abstract

:
Aortic dissection (AD) is a cardiovascular disease that seriously endangers the lives of patients. The mortality rate of this disease is high, and the incidence is increasing annually, but the pathogenesis of AD is complicated. In recent years, an increasing number of studies have shown that immune cell infiltration in the media and adventitia of the aorta is a novel hallmark of AD. These cells contribute to changes in the immune microenvironment, which can affect their own metabolism and that of parenchymal cells in the aortic wall, which are essential factors that induce degeneration and remodeling of the vascular wall and play important roles in the formation and development of AD. Accordingly, this review focuses on the independent and interactive roles of immunity and metabolism in AD to provide further insights into the pathogenesis, novel ideas for diagnosis and new strategies for treatment or early prevention of AD.

Graphical Abstract

1. Introduction

Aortic dissection (AD) is a dangerous cardiovascular disease. The incidence rate in the general population is 3–5 cases/100,000 persons, and the incidence rate in the middle-aged and elderly population can be as high as 10 cases/100,000 persons. The mortality rate of untreated type A acute AD (AAD) patients after symptom onset is 1% to 2% per hour [1,2,3,4]. Without clinical intervention, the mortality rate is as high as 50% within 48 h after onset. The long-term survival of patients who survive an acute aortic event is reduced, with an increased risk of death that extends to 90 days from diagnosis [5,6]. The aortic wall is composed of the intima, media and adventitia from the inside to the outside. The inner membrane is composed of endothelial cells (ECs) and the subendothelial layer. The media is mainly composed of elastic fibers, collagen fibers and vascular smooth muscle cells (VSMCs). The adventitia of the aorta is composed of loose connective tissue, which contains elastic fibers, collagen fibers and fibroblasts. Under physiological conditions, the three aortic membranes stick closely to each other and jointly bear the pressure of blood flow in the blood vessels. However, when there is a break in the inner membrane, the impact of blood can further tear and expand the gap, causing the three-layer membrane to separate and resulting in the formation of AD. The pathogenesis of AD is not completely clear. The accepted view is that, in response to some internal and external factors, an entry forms in the inner membrane, and blood enters the media through the entry, separating the inner and outer membranes. The occurrence and development of this event are mainly caused by vascular inflammation, extravascular matrix degradation and VSMC apoptosis. Hypertension, atherosclerosis, Marfan syndrome and smoking are all related risk factors for AD [7,8,9]. Among them, approximately two-thirds of patients with dissection have high blood pressure, 50% of AD patients under 40 years of age have Marfan syndrome and approximately 20% of patients have thoracic aortic aneurysm (AA) or a family history of AD [10].
Apart from genetics, lifestyle habits and background diseases, it is important to focus on the cellular and molecular mechanisms of AD, and abnormal immunity and metabolism are two essential aspects involved in AD initiation and progression [11,12,13]. Recent clinical and basic studies have shown that extravascular matrix degradation and VSMC apoptosis are exacerbated as the degree of inflammation increases [14,15]. This finding suggests that the inflammatory response may play an important role in the early onset of AD and can activate multiple pathological processes to further promote the onset of AD. Furthermore, immune cells such as macrophages, lymphocytes and neutrophils are often detected in the media and adventitia of AD tissues [16,17]. Moreover, abnormal carbohydrate, lipid and amino acid metabolism has gained increasing attention recently. Immunometabolism has gradually become a new research hotspot in the medical field and has been widely used in the research of diabetes, cancer and cardiovascular disease. Choosing a more reasonable strategy to regulate immune metabolism and maintain natural immune homeostasis is critical. Immunity and metabolism exert modulatory effects independently and interactively. Infiltrated immune cells not only promote the secretion of matrix metalloproteinases (MMPs) and adhesion molecules but also release reactive oxygen species (ROS), leading to changes in the metabolism in the microenvironment and VSMC apoptosis, which ultimately leads to the development of AD [18,19,20]. In contrast, alterations in immune and parenchymal cell metabolism change the immune microenvironment and immune cell function, which are associated with the development of AD [21,22]. Therefore, this review elucidates the pathological mechanism of AD development through a novel perspective of immunity and metabolism to provide a new perspective for comprehensive cognition, diagnosis and treatment of AD.

2. Immunity in AD

2.1. Innate Immunity and Associated Signaling in AD

2.1.1. Neutrophils

Neutrophils have phagocytic and chemotactic functions and play a role in the inflammatory cascade after AAD. Lauren et al. showed that neutrophil infiltration into the adventitia occurred within 12 h after AD and peaked from 12 to 24 h [23]. Moreover, clinical studies have shown that the neutrophil-to-lymphocyte ratio (NLR) can be used as a prognostic predictor of AD [24,25]. Apart from this, neutrophils are the main cells that secrete MMPs. Some researchers have shown that the accumulation of MMP-9 leads to disorders of ECM metabolism, promoting inflammation and the degradation of elastic fibers and accelerating the expansion and rupture of the dissection [26]. Moreover, neutrophil-derived IL-6 is a potent enhancer of post-AAD adventitial inflammation that leads to increased CRP levels following aortic rupture [27]. Additionally, activated neutrophils can release neutrophil extracellular traps (NETs), serine proteases, histone proteases and ROS intermediates that can promote the progression of vascular inflammation and AD [28,29]. Yang et al. measured the expression levels of three NET markers in AAD with different severities and confirmed that NET levels are a reliable factor in the diagnosis of AAD and are related to the severity and prognosis of the disease [30]. Laridan et al. also reached the same conclusion in a mouse model, in which the presence of NETs could be detected in the early stage of abdominal aortic aneurysm (AAA). Treatment of induced mice with cloamidine, a NET inhibitor, significantly reduced AAA formation [31]. The new study demonstrates that neutrophil elastase (NE) promotes thoracic aortic dissection (TAD) development and aortic dissection by enhancing inflammatory cell recruitment/migration into the vessel wall, increasing aortic inflammation and promoting SMC phenotypic switching. Importantly, this study also highlights the therapeutic value of the NE-TBL1x signaling axis in patients with high-risk TAD [32]. Research has shown that depleting neutrophils with antibodies or reducing neutrophil infiltration into vascular tissue in mice could reduce MMP production and significantly reduce the incidence of AD [33].

2.1.2. Monocytes

Monocytes are a heterogeneous cell population that can be categorized into three subpopulations with different phenotypic and functional properties: “classical” (CD14++CD16), “intermediate” (CD14++CD16+) and “nonclassical” (CD14+CD16+). Furthermore, classical monocytes are significantly increased, whereas intermediate monocytes are significantly decreased in AAD [34]. Factors such as oxidative stress, cytokines, viral or bacterial infections, high blood sugar or high low-density lipoprotein (LDL) levels can activate the vascular endothelium, induce monocyte recruitment, promote the secretion of the chemokines CCL2, CCL5 and CCL7 and upregulate adhesion molecules such as intercellular cell adhesion molecule-1 (ICAM-1) and vascular cell adhesion molecule-1 (VCAM-1), which can promote the rolling of circulating blood mononuclear cells, adhesion to ECs and migration into the tissue [35,36]. Monocytes that migrate from the blood to the tissues play a phagocytic role, produce inflammatory factors and then differentiate into tissue macrophages and dendritic cells [37,38]. It has been reported that IL-6 can induce monocytes to differentiate into macrophages and can mediate Th17 production through the IL-6 signal transducer and activator of transcription 3 (STAT3) pathway [39]. Activated monocytes and platelet receptor glycoprotein Ib alpha (GPIbα) participate in local thrombin amplification through coagulation factor XI (FXI), thereby promoting the development of vascular inflammation and hypertension and accelerating the progression of AD [40]. Additionally, low-density lipoprotein receptor-related protein 8 (LRP8) is released from monocytes and can induce vascular inflammation and endothelial dysfunction, leading to AD [41].

2.1.3. Macrophages

Pathophysiological processes, including macrophage migration and infiltration into the aorta, differentiation or phenotype transformation, the secretion of various inflammatory factors and the release of extracellular matrix (ECM)-degrading proteins, are closely related to the occurrence and development of AD [42]. Inflammatory factors in the peripheral blood and vascular wall of patients with Stanford type A dissection were examined by Del Porto et al., and the results showed that macrophages were the main inflammatory cells that infiltrated aortic tissues, the ECM was degraded, and the immune response played an important role in this process [43]. Consistently, Darrell et al. showed that the infiltration of macrophages in AAD tissue was significantly higher than that in chronic AD tissue [44]. M1 (proinflammatory) and M2 (anti-inflammatory) macrophages play different roles in the pathological process of AD [45]. Cytotoxic M1 macrophages release a variety of proinflammatory factors, ROS and nitric oxide synthase (NOS) to further tissue damage; however, M2 macrophages exert anti-inflammatory effects by inhibiting T cell proliferation and the phagocytosis of apoptotic neutrophils and reducing the production of proinflammatory factors [46]. Recent studies have shown that IL-6 activates STAT3 through negative regulation of cytokine signal transduction inhibitor 3 and changes the gene expression programs of macrophages and VSMCs, thereby regulating the inflammatory response and affecting AD [47]. In contrast, IL-5 overexpression in macrophages attenuates the inflammatory response and SMC apoptosis, thereby inhibiting the development of AD [48]. In addition, researchers found that DNA from damaged SMCs was phagocytosed by macrophages and activated the STING pathway via interferon regulatory factor 3 to induce the expression of MMP-9, which promoted the degradation of the ECM to accelerate AD [49]. Additionally, adenosine deaminase acting on RNA (ADAR1) is involved in macrophage activation and AAA formation through degradation of Drosha protein [50].

2.1.4. Mast Cells

Mast cells promote the development and progression of AD through the release of inflammatory factors. Gao et al. found that mast cells could secrete the proinflammatory factors interferon-γ (IFN-γ) and MMPs, which cause VSMC apoptosis and vascular remodeling, leading to the formation of AA [51]. In addition, Springer et al. showed that in vivo mast cells could partially and transiently regulate systemic IL-6 homeostasis [52]. Swedenborg et al. also proved that mast cells play an important role in AA. Studies indicate that mast cells are the major source of cathepsin G in AA and promote the activation of the renin–angiotensin system through cathepsin G and chymase, which contribute to VSMC apoptosis and the release of proteolytic enzymes, thereby accelerating AA [53].

2.2. Adaptive Immunity and Associated Signaling in AD

2.2.1. T Cells/Th Cells

T cells/Th cells have been shown to induce VSMC apoptosis and MMP synthesis [54], which suggests that these cells play a role in the development of AD [55,56]. Clinical studies have shown high levels of CD3+, CD4+, CD8+ and CD45+ T cells in the aortic tissue of AD patients, indicating that T cell activation is involved in the development of AD [57]. CD4+ T cells can differentiate into different subpopulations under different conditions, such as T helper 1 (Th1), Th2, Th17 and T regulatory (Treg) cells [58]. In addition, the numbers of Th1, Th9, Th17 and Th22 cells and the expression of their transcription factors were elevated in AD patients, while the levels of Th2 and Treg cells and their transcription factors were decreased [56]. This finding suggested that Th2 cells and Treg cells may protect against the development of AD. Several clinical studies have shown that Treg cells play a critical role in inflammation associated with arterial injury in a variety of cardiovascular diseases, such as AAD and symptomatic carotid artery stenosis [59,60]. Studies have confirmed that Treg cells can play both anti-inflammatory and proinflammatory roles, which may be attributed to the heterogeneity of Treg cells. For instance, IL-10-producing Treg cells may protect against aortic wall rupture through anti-inflammatory effects on AAD, whereas CD25+ Treg cells are increased and mediate inflammation in patients with symptomatic carotid stenosis [61,62]. Furthermore, it was confirmed that Th17 and Treg cells share common precursor cells, and both cells require transforming growth factor-β (TGF-β) signaling for their initial differentiation, but they perform opposite functions, in which IL-6 plays an essential role [63]. IL-6 and IL-17 produced by Th17 cells are significantly elevated in AD tissue, and IL-6 can promote the transformation of Th0 cells into Th17 cells [64]. Therefore, inhibiting IL-6 to reduce Th0 cells to Th17 cells has a positive impact on controlling AAD onset and progression [65].

2.2.2. B Cells

B cells play a key role in adaptive immunity and participate in the pathogenesis of AD, mainly through their ability to detect and process antigens [66,67]. Saraff et al. revealed that in an angiotensin II (Ang II)-induced AD model, B cells mainly infiltrated the media of the vessel wall. During AD development, B cells are stimulated by antigens to proliferate and differentiate into plasma cells, which synthesize and secrete antibodies and circulate in the blood to perform humoral immune functions. However, the specific role of B cells in AD is still unclear. Schaheen et al. discovered that a lack of mature B cells protects against the formation of AAAs induced by elastase [68]. However, Akshaya et al. showed that there was no significant difference in the incidence of AAs between mice lacking mature B cells and wild-type mice [69]. At present, the role of B cells in AA is still controversial, and more research on their function in AD or AA is needed.

2.3. Interactions of Immune Cells in AD

Local inflammatory responses and systemic inflammatory responses in the vessel wall are triggered after AD occurs; this process involves interactions among immune cells. The high expression of chemokines and granulocyte colony-stimulating factor in AD tissue after local inflammation leads to neutrophil infiltration and accelerates the occurrence of lesions in the aortic media [70]. NETs facilitate the secretion of inflammatory factors by macrophages and indirectly activate Th17 cells to promote inflammation [71]. Furthermore, among inflammatory factors, IL-6 can induce the differentiation of monocytes into macrophages and mediate the production of Th17 through the IL-6–STAT3 pathway; Th17 cells can promote the chemotaxis, adhesion and migration of monocytes by mediating inflammatory responses [72]; IL-8 has a role in regulating neutrophil mobilization and migration [73]; IL-1β can promote the secretion of MMPs by immune cells through the phosphorylation of p38 [74]; and exosomes from monocytes are transformed into active macrophages through the NF-κB signaling pathway which continue to secrete a variety of inflammatory substances, proteolytic enzymes and ECM-degrading proteins and continue to damage blood vessels [75]. In addition, immune cells cause changes in the microenvironment, such as the accumulation of homocysteine (Hcy). Liu et al. demonstrated that Hcy stimulates adventitial fibroblasts to produce IL-6 and monocyte chemoattractant protein-1 (MCP-1), thereby recruiting more THP-1 cells to infiltrate the vessel wall [76]. Shao et al. also proved that Hcy induces the secretion of anti-beta 2 glycoprotein I (anti-β2GPI) antibodies by B cells through the toll-like receptor 4 (TLR-4) pathway and then promotes the accumulation of IgG in AAA tissue. In addition, Hcy can polarize inflammatory macrophages and upregulate MMP-2 and MMP-9 expression to promote inflammatory reactions and matrix degradation [77]. In summary, inflammatory cells play an important role in the progression of AD. The direct and indirect interactions between various subsets of immune cells provide a complex inflammatory microenvironment for the initiation and development of AD.

3. Metabolisms in AD

3.1. Carbohydrate Metabolism

Glucose is an essential source of energy for metabolism. Under normal oxygen conditions, cells prefer glycolysis to the oxidative phosphorylation (OXPHOX) process, known as “Warburg effect”. Lactic acid accumulation and OXPHOX damage caused by the Warburg effect are common in parenchymal cells in AD [78,79]. The inflammatory tissue of AD is often hypoxic [80,81]. Under this condition, the expression of hypoxia-inducible factor (HIF), especially O2-sensitive HIF-1α, is enhanced, which plays an important role in the Warburg effect [82,83,84]. Migrating VSMCs have increased glucose transporter 1 (GLUT1) and hexokinase 2 (HK2) expression to promote glycolysis [85,86]. Lactate dehydrogenase (LDHA) is also upregulated, stimulating VSMC proliferation, migration and synthetic phenotypic transformation. In vivo, the inhibitor of LDH and lactate reduced the degradation of elastic fibers, collagen deposition and MMP2/9 production and inhibited the phenotypic transformation of VSMC so that the progression of AD was delayed [87,88]. The NR1D1–ACO2 axis interferes with aberrant tricarboxylic acid (TCA) cycle functions. α-KG, the substrate of ACO2, supplementation is regarded as an effective therapeutic approach for AD which can increase mitochondrial ETC complex expression to prevent AAA formation [89]. Hypoxic conditions cause a shift in glucose flux towards the pentose phosphate pathway (PPP). High levels of NADPH generated via the PPP as part of an antioxidant mechanism protect VSMCs from apoptosis [90]. Fan et al. indicated that a downregulation of TCA cycle metabolites may contribute to lung function damage in AAD [91]. Based on these findings, there is potential for using metabolic therapies to reduce mortality and improve prognosis in AAD in the future. Any targeted intervention that inhibits the Warburg effect or the generation of metabolic intermediates will be a new direction in the treatment of AD.

3.2. Lipid Metabolism

It is becoming increasingly clear that high lipid levels predispose individuals to the development of clinical cardiovascular diseases. When the β oxidation of fatty acids (FAs) is suppressed, less acetyl coenzyme A (acetyl-CoA) enters the TCA, thereby inhibiting OXPHOX [92]. Increased concentrations of non-esterified FAs induce endoplasmic reticulum stress, mitochondrial dysfunction, inflammation and cytokine release, which are causative factors of AD [93]. Observational studies found obvious dyslipidemia in AD patients, and the in-hospital mortality of AD patients was correlated with dyslipidemia [94]. The secretion of arachidonic acid (ARA) and increased synthesis of prostaglandin E2 (PGE2) are associated with endothelial dysfunction and vascular inflammation in patients with MFS [95,96]. Likewise, oleic acid (OA) produces the same effects by activating inducible NOS (iNOS), leading to AA production [97]. In contrast, nitro-oleic acid (nitro-OA) inhibits the ERK1/2 and Smad2 pathways and NF-κB overactivation, which significantly reduces aortic dilatation in AD [98]. Oxidized low-density lipoprotein (ox-LDL) activates caspase 1, leading to the activation of IL-1β, and mature IL-1β is released by human VSMCs, thereby triggering vascular disease [99]. Treatment with ferroptosis driven by the excessive lipid peroxidation accumulation can decrease morbidity in AD models [100]. Considering the different roles of various lipid types in promoting health and reducing disease risk, there is growing interest in lipid metabolism. Lipids can not only be diagnostic markers of AD but also target molecules for AD therapy. Statins are considered the most efficient drugs for the treatment of hyperlipidemia. Steinmetz et al. found that statins were beneficial for the risk of AA expansion after using them to treat an AA animal model [101]. Therefore, it was hypothesized that certain lipid interventions could prevent or slow the progression of AD.

3.3. Amino Acid Metabolism

Aberrant and dysregulated signaling associated with amino acid metabolism is closely implicated in AD. Wang et al. showed that plasma aminograms were significantly altered in patients with AD [102]. According to epidemiological studies, hyperhomocysteinemia (HHcy) is a hallmark of vascular injury. HHcy stimulates the secretion of IL-6 and MCP-1 and the subsequent recruitment of monocytes/macrophages and promotes adventitial fibroblasts’ transformation into myofibroblasts [103]. Vitamin B restored TGF-β signaling and promoted lysyl-oxidase-mediated collagen maturation in aortic media. Vitamin B can serve as a synergistic drug for the treatment of MFS [104]. S-adenosyl-L-homocysteine cysteine (SAH) is a more sensitive biomarker of cardiovascular disease than Hcy. High levels of SAH in plasma contribute to endothelial dysfunction [105]. Overexpression of solute carrier family 1 member 5 (SLC1A5), a key glutamine (GLN) transporter, results in mTORC1 activation and VSMC proliferation; these changes are attenuated after additional treatments with GLN metabolism inhibitor BPTES [106,107]. Gallina found that AMPA-type glutamate receptors mediate VSMC transformation into the contractile phenotype [108]. In addition, mutations in the glutamate ammonia ligase (GLUL) gene and high glutamate levels increase the risk of cardiovascular disease [109]. Glutamate is converted to proline during the production of anti-inflammatory cytokines such as IL-10 to promote tissue and inflammatory repair [110]. In addition, clinical studies have shown that elevated plasma succinic acid levels can reliably distinguish AAD patients from patients with other diseases [111]. ECs play a critical role in the pathogenesis of AAD by regulating protein S-sulfhydration, and higher plasma H2S levels are associated with a lower risk of AAD [112]. Considering the findings of the current study, the amino acid profile is expected to provide a novel, noninvasive and objective diagnosis for AD.

3.4. Interconnection and Cross-Regulation of Metabolism in AD

The metabolism of carbohydrates, lipids and amino acids is linked through a common intermediate product (acetyl-CoA) and a common metabolic pathway, the TCA cycle, to form a whole system; therefore, when there is a metabolic disorder in any one or more of these factors, there is crosstalk with other pathway disorders. As we mentioned previously, the Warburg effect is not only associated with the glycolytic pathway; “errors” in the glycolytic pathway can bring about “mistakes” in amino acid and lipid metabolism. This series of changes promotes the production of ROS, NADPH and NO, and the production of these substances results in various adverse consequences through various pathways, such as parenchymal cell disruption and dysfunction and the activation of inflammatory factors and inflammasomes. These adverse consequences can eventually promote the occurrence and development of AD.

4. Synergism or Antagonism between Immunity and Metabolism in AD

4.1. Metabolism Shapes the Immune Microenvironment of AD

Immune cells and parenchymal cells need to constantly regulate their own metabolism to perform their corresponding functions in AD. During this process, the production of both cytokines and metabolites alters the microenvironment, further affecting metabolism; for example, the preferential production of lactate via the Warburg effect promotes neutrophil activity [113]. Nox-derived ROS, which are maintained by NADPH produced via the PPP, can effectively induce the formation of NETs [114,115]. The production of α-KG via GLN catabolism for Jmjd3-dependent epigenetic reprogramming is important for alternative activation of M2 macrophages [116]. Lipoprotein a (Lp(a)) mediates a proinflammatory response through oxidized phospholipid (OxPL), which is recognized as a danger-associated molecular pattern by pattern recognition receptors on innate immune cells and induces monocyte trafficking to induce inflammation in the arterial wall [117]. In addition, the acidic environment produced by aerobic glycolysis accelerates the uptake of lipoproteins by macrophages, and more carbon from gluconeogenesis is bound to FAs and sterols, which contributes to the further accumulation of triglycerides (TGs) and reduces TG catabolism in activated macrophages; a series of microenvironmental changes can enhance the proinflammatory effects of macrophages, exacerbating the progression of AD [118]. There have been reports showing that GLN can increase the expression of the NADPH oxidase components p22 (phox), gp91 (phox) and p47 (phox) and meditate the production of superoxide by neutrophils [119]. Antagonizing GLN uptake is considered a novel approach for the treatment of AD [120,121].

4.2. Regulatory Effects of Immune Cells on Metabolic Reprogramming in AD Parenchymal Cells

Different immune cell subpopulations differentiate due to metabolic disorders and, in turn, regulate the metabolic reprogramming of parenchymal cells [122]. When AD is accompanied by inflammation, disturbed carbohydrate, lipid and amino acid metabolism in immune cells results in oxidative stress, mitochondrial stress and endoplasmic reticulum stress, and these series of changes prompt cells to produce ROS, NADPH and NO and activate inflammatory vesicles, accelerating EC, VSMC and other parenchymal cell migration and apoptosis, as well as enhancing ECM degradation [123,124,125,126]. Lian et al. revealed that macrophages induced HIF-1α activation in response to fumarate accumulation, which triggers vascular inflammation, metalloproteinase degradation and elastic plate rupture through increased deintegrin and extracellular mesenchymal protein structural domain 17 (ADAM17) expression in humans and mouse models [127,128]. The researchers also identified an enzyme secreted by inflammatory cells called FLp-PLA2, which can hydrolyze glycerophospholipids to produce bioactive lipids, leading to vascular endothelial dysfunction and exacerbating vascular inflammation, and is clinically regarded as an early warning indicator of AD [129,130]. Additionally, in M1-type macrophages, interruption of the TCA cycle causes the accumulation of the intermediate product succinate, which enhances the inflammatory response and further regulates metabolic reprogramming in parenchymal cells [112]. In addition, ox-LDL can promote the glycolytic capacity of macrophages, thereby promoting cholesterol uptake by VSMCs. Cholesterol can regulate Nox isoforms and redox signaling in VSMCs to provoke vascular disease [131]. In addition, T cells regulate EC function during the inflammatory response by secreting a variety of immunomodulatory molecules and cytokines, such as IL-10 and TGF-β, which have anti-inflammatory effects; IL-4 and IL-13 promote the synthesis of FAs and phospholipids, which protect against complement-induced injury and are potential molecules for AD therapy, and maintain mitochondrial function [132].

5. Perspectives

With increasing exploration of research strategies and technologies, elucidating the occurrence and development of AD at the cellular and molecular levels will be conducive to understanding its pathogenesis, diagnosis, treatment and prognosis. To date, several studies focusing on immunity and metabolism have revealed the underlying cellular and molecular mechanisms of AD and mainly included three aspects: How do the heterogeneity and multiple functions of immune cells shape inflammation in AD? How do immune cells interact with parenchymal cells to cause vascular remodeling in AD? How does metabolism at the cell and whole-body level affect the pathological course of AD? Although some features are still unclear, these studies provide a variety of insights (Figure 1 and Graphical Abstract). In the future, to achieve better prevention and treatment of AD, two issues should be addressed: on the one hand, the key molecular networks orchestrating immunity and metabolism must be identified, which will elucidate AD pathogenesis and provide potential treatment targets; on the other hand, specific circulating immune cell subsets or their released factors and specific circulating metabolites will provide early warning or diagnostic biomarkers for AD.

Author Contributions

Conceptualization, W.H. and S.-S.D.; writing review original draft preparation, S.L. and W.H.; writing article review and editing, J.L. and W.C.; funding acquisition, J.L., W.H. and S.-S.D. All authors have read and agreed to the published version of the manuscript.

Funding

The present study was supported by grants from the National Natural Science Foundation of China (no. 82201941, no. 82370483 and no. 81670407), Chongqing University innovation group project for Shuang-Shuang Dai and Chongqing Natural Science Foundation (2023NSCQ-MSX2767).

Acknowledgments

We should thank Sanjiu Yu for assistance with the analysis of clinic data in some references.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rylski, B.; Schilling, O.; Czerny, M. Acute aortic dissection: Evidence, uncertainties, and future therapies. Eur. Heart J. 2023, 44, 813–821. [Google Scholar] [CrossRef]
  2. Juraszek, A.; Czerny, M.; Rylski, B. Update in aortic dissection. Trends Cardiovas Med. 2022, 32, 456–461. [Google Scholar] [CrossRef]
  3. Stombaugh, D.K.; Mangunta, V.R. Aortic Dissection. Anesthesiol. Clin. 2022, 40, 685–703. [Google Scholar] [CrossRef]
  4. Zhu, Y.; Lingala, B.; Baiocchi, M.; Tao, J.J.; Toro Arana, V.; Khoo, J.W.; Woo, Y.J. Type A Aortic Dissection-Experience Over 5 Decades: JACC Historical Breakthroughs in Perspective. J. Am. Coll. Cardiol. 2020, 76, 1703–1713. [Google Scholar] [CrossRef]
  5. Sayed, A.; Munir, M.; Bahbah, E.I. Aortic Dissection: A Review of the Pathophysiology, Management and Prospective Advances. Curr. Cardiol. Rev. 2021, 17, e1507047109. [Google Scholar] [CrossRef]
  6. Braverman, A.C.; Mittauer, E.; Harris, K.M.; Evangelista, A.; Pyeritz, R.E.; Brinster, D.; Conklin, L.; Suzuki, T.; Fanola, C.; Ouzounian, M.; et al. Clinical Features and Outcomes of Pregnancy-Related Acute Aortic Dissection. JAMA Cardiol. 2021, 6, 58–66. [Google Scholar] [CrossRef]
  7. Zhou, Z.; Cecchi, A.C.; Prakash, S.K.; Milewicz, D.M. Risk Factors for Thoracic Aortic Dissection. Genes 2022, 13, 1814. [Google Scholar] [CrossRef]
  8. Rombouts, K.B.; van Merrienboer, T.; Ket, J.; Bogunovic, N.; van der Velden, J.; Yeung, K.K. The role of vascular smooth muscle cells in the development of aortic aneurysms and dissections. Eur. J. Clin. Investig. 2022, 52, e13697. [Google Scholar] [CrossRef]
  9. Okita, Y. Aortic dissection is more violent in the young. Eur. J. Cardio-Thorac. 2023, 63, ezad209. [Google Scholar] [CrossRef]
  10. Roman, M.J.; Devereux, R.B. Aortic Dissection Risk in Marfan Syndrome. J. Am. Coll. Cardiol. 2020, 75, 854–856. [Google Scholar] [CrossRef]
  11. Ostberg, N.P.; Zafar, M.A.; Ziganshin, B.A.; Elefteriades, J.A. The Genetics of Thoracic Aortic Aneurysms and Dissection: A Clinical Perspective. Biomolecules 2020, 10, 182. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, F.; Wei, T.; Liu, L.; Hou, F.; Xu, C.; Guo, H.; Zhang, W.; Ma, M.; Zhang, Y.; Yu, Q.; et al. Role of Necroptosis and Immune Infiltration in Human Stanford Type A Aortic Dissection: Novel Insights from Bioinformatics Analyses. Oxid. Med. Cell Longev. 2022, 2022, 6184802. [Google Scholar] [CrossRef] [PubMed]
  13. Magoon, R.; Shri, I.; Jose, J. The malnutritional facet of inflammatory prognostication in acute aortic dissection. J. Cardiac Surg. 2022, 37, 1458–1459. [Google Scholar] [CrossRef] [PubMed]
  14. Postnov, A.; Suslov, A.; Sobenin, I.; Chairkin, I.; Sukhorukov, V.; Ekta, M.B.; Khotina, V.; Afanasiev, M.; Chumachenko, P.; Orekhov, A. Thoracic Aortic Aneurysm: Blood Pressure and Inflammation as Key Factors in the Development of Aneurysm Dissection. Curr. Pharm. Design. 2021, 27, 3122–3127. [Google Scholar] [CrossRef]
  15. Skotsimara, G.; Antonopoulos, A.; Oikonomou, E.; Papastamos, C.; Siasos, G.; Tousoulis, D. Aortic Wall Inflammation in the Pathogenesis, Diagnosis and Treatment of Aortic Aneurysms. Inflammation 2022, 45, 965–976. [Google Scholar] [CrossRef]
  16. Gao, L.; Jin, B.; Shen, J.; Zhang, X. Effect of Dexmedetomidine on Inflammatory Response in Aortic Dissection. Heart Surg. Forum. 2022, 25, E829–E832. [Google Scholar] [CrossRef]
  17. Adachi, Y.; Ueda, K.; Nomura, S.; Ito, K.; Katoh, M.; Katagiri, M.; Yamada, S.; Hashimoto, M.; Zhai, B.; Numata, G.; et al. Beiging of perivascular adipose tissue regulates its inflammation and vascular remodeling. Nat. Commun. 2022, 13, 5117. [Google Scholar] [CrossRef]
  18. Shao, Y.; Li, G.; Huang, S.; Li, Z.; Qiao, B.; Chen, D.; Li, Y.; Liu, H.; Du, J.; Li, P. Effects of Extracellular Matrix Softening on Vascular Smooth Muscle Cell Dysfunction. Cardiovasc. Toxicol. 2020, 20, 548–556. [Google Scholar] [CrossRef]
  19. Zhang, C.; Niu, K.; Ren, M.; Zhou, X.; Yang, Z.; Yang, M.; Wang, X.; Luo, J.; Shao, Y.; Zhang, C.; et al. Targeted Inhibition of Matrix Metalloproteinase-8 Prevents Aortic Dissection in a Murine Model. Cells 2022, 11, 3218. [Google Scholar] [CrossRef]
  20. Qiu, L.; Yi, S.; Yu, T.; Hao, Y. Sirt3 Protects Against Thoracic Aortic Dissection Formation by Reducing Reactive Oxygen Species, Vascular Inflammation, and Apoptosis of Smooth Muscle Cells. Front. Cardiovasc. Med. 2021, 8, 675647. [Google Scholar] [CrossRef]
  21. Oduro, P.K.; Zheng, X.; Wei, J.; Yang, Y.; Wang, Y.; Zhang, H.; Wang, Q. The cGAS-STING signaling in cardiovascular and metabolic diseases: Future novel target option for pharmacotherapy. Acta Pharm. Sin. B. 2022, 12, 50–75. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, L.; Zhang, L.; Zhao, Z.; Liu, Y.; Wang, J.; Niu, M.; Sun, X.; Zhao, X. Metabolic syndrome and its components are associated with hypoxemia after surgery for acute type A aortic dissection: An observational study. J. Cardiothorac. Surg. 2022, 17, 151. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, L.; Burke, A. Acute medial dissection of the ascending aorta: Evolution of reactive histologic changes. Am. J. Surg. Pathol. 2013, 37, 1275–1282. [Google Scholar] [CrossRef] [PubMed]
  24. Erdolu, B.; As, A.K. C-Reactive Protein and Neutrophil to Lymphocyte Ratio Values in Predicting Inhospital Death in Patients with Stanford Type A Acute Aortic Dissection. Heart Surg. Forum. 2020, 23, E488–E492. [Google Scholar] [CrossRef] [PubMed]
  25. Li, S.; Yang, J.; Dong, J.; Guo, R.; Chang, S.; Zhu, H.; Li, Z.; Zhou, J.; Jing, Z. Neutrophil to lymphocyte ratio and fibrinogen values in predicting patients with type B aortic dissection. Sci. Rep. 2021, 11, 11366. [Google Scholar] [CrossRef]
  26. He, W.; Yu, S.; Li, H.; He, P.; Xiong, T.; Yan, C.; Zhang, J.; Chen, S.; Guo, M.; Tan, X.; et al. Comparison and Evaluation of Two Combination Modes of Angiotensin for Establishing Murine Aortic Dissection Models. J. Cardiovasc. Transl. 2023. [Google Scholar] [CrossRef]
  27. Sano, M. Complexity of Inflammation in the Trajectory of Vascular Disease: Interleukin 6 and Beyond. Ann. Vasc. Dis. 2023, 16, 8–16. [Google Scholar] [CrossRef]
  28. Pan, G.; Liao, M.; Dai, Y.; Li, Y.; Yan, X.; Mai, W.; Liu, J.; Liao, Y.; Qiu, Z.; Zhou, Z. Inhibition of Sphingosine-1-Phosphate Receptor 2 Prevents Thoracic Aortic Dissection and Rupture. Front. Cardiovasc. Med. 2021, 8, 748486. [Google Scholar] [CrossRef]
  29. Frangou, E.; Vassilopoulos, D.; Boletis, J.; Boumpas, D.T. An emerging role of neutrophils and NETosis in chronic inflammation and fibrosis in systemic lupus erythematosus (SLE) and ANCA-associated vasculitides (AAV): Implications for the pathogenesis and treatment. Autoimmun. Rev. 2019, 18, 751–760. [Google Scholar] [CrossRef]
  30. Yang, S.; Xiao, Y.; Du, Y.; Chen, J.; Ni, Q.; Guo, X.; Xue, G.; Xie, X. Diagnostic and Prognostic Value of Neutrophil Extracellular Trap Levels in Patients With Acute Aortic Dissection. Front. Cardiovasc. Med. 2021, 8, 683445. [Google Scholar] [CrossRef]
  31. 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] [PubMed]
  32. Yang, M.; Zhou, X.; Pearce, S.W.; Yang, Z.; Chen, Q.; Niu, K.; Liu, C.; Luo, J.; Li, D.; Shao, Y.; et al. Causal Role for Neutrophil Elastase in Thoracic Aortic Dissection in Mice. Arter. Throm Vas. 2023, 43, 1900–1920. [Google Scholar] [CrossRef] [PubMed]
  33. Chang, T.T.; Liao, L.Y.; Chen, J.W. Inhibition on CXCL5 reduces aortic matrix metalloproteinase 9 expression and protects against acute aortic dissection. Vasc. Pharmacol. 2021, 141, 106926. [Google Scholar] [CrossRef] [PubMed]
  34. Cifani, N.; Proietta, M.; Taurino, M.; Tritapepe, L.; Del, P.F. Monocyte Subsets, Stanford-A Acute Aortic Dissection, and Carotid Artery Stenosis: New Evidences. J. Immunol. Res. 2019, 2019, 9782594. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, K.; Saaoud, F.; Yu, S.; Drummer, I.V.C.; Shao, Y.; Sun, Y.; Yang, X. Monocyte Adhesion Assays for Detecting Endothelial Cell Activation in Vascular Inflammation and Atherosclerosis. Methods Mol. Biol. 2022, 2419, 169–182. [Google Scholar]
  36. Jardine, L.; Wiscombe, S.; Reynolds, G.; McDonald, D.; Fuller, A.; Green, K.; Filby, A.; Forrest, I.; Ruchaud-Sparagano, M.-H.; Scott, J.; et al. Lipopolysaccharide inhalation recruits monocytes and dendritic cell subsets to the alveolar airspace. Nat. Commun. 2019, 10, 1999. [Google Scholar] [CrossRef]
  37. Shen, W.-Y.; Li, H.; Zha, A.-H.; Luo, R.-Y.; Zhang, Y.-L.; Luo, C.; Dai, R.-P. Platelets reprogram monocyte functions by secreting MMP-9 to benefit postoperative outcomes following acute aortic dissection. Iscience 2023, 26, 106805. [Google Scholar] [CrossRef]
  38. Tomida, S.; Aizawa, K.; Nishida, N.; Aoki, H.; Imai, Y.; Nagai, R.; Suzuki, T. Indomethacin reduces rates of aortic dissection and rupture of the abdominal aorta by inhibiting monocyte/macrophage accumulation in a murine model. Sci. Rep. 2019, 9, 10751. [Google Scholar] [CrossRef]
  39. Cai, Y.L.; Wang, Z.W. The expression and significance of IL-6, IFN-γ, SM22α, and MMP-2 in rat model of aortic dissection. Eur. Rev. Med. Pharmaco. 2017, 21, 560–568. [Google Scholar]
  40. Koltsova, E.K.; Sundd, P.; Zarpellon, A.; Ouyang, H.; Mikulski, Z.; Zampolli, A.; Ruggeri, Z.M.; Ley, K. Genetic deletion of platelet glycoprotein Ib alpha but not its extracellular domain protects from atherosclerosis. Thromb. Haemost. 2014, 112, 1252–1263. [Google Scholar] [CrossRef]
  41. Lagrange, J.; Finger, S.; Kossmann, S.; Garlapati, V.; Ruf, W.; Wenzel, P. Angiotensin II Infusion Leads to Aortic Dissection in LRP8 Deficient Mice. Int. J. Mol. Sci. 2020, 21, 4916. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, X.; Zhang, H.; Cao, L.; He, Y.; Ma, A.; Guo, W. The Role of Macrophages in Aortic Dissection. Front Physiol. 2020, 11, 54. [Google Scholar] [CrossRef] [PubMed]
  43. Del Porto, F.; Proietta, M.; Tritapepe, L.; Miraldi, F.; Koverech, A.; Cardelli, P.; Tabacco, F.; De Santis, V.; Vecchione, A.; Mitterhofer, A.P.; et al. Inflammation and immune response in acute aortic dissection. Ann. Med. 2010, 42, 622–629. [Google Scholar] [CrossRef] [PubMed]
  44. Ortega, R.; Collado, A.; Selles, F.; Gonzalez-Navarro, H.; Sanz, M.-J.; Real, J.T.; Piqueras, L. SGLT-2 (Sodium-Glucose Cotransporter 2) Inhibition Reduces Ang II (Angiotensin II)-Induced Dissecting Abdominal Aortic Aneurysm in ApoE (Apolipoprotein E) Knockout Mice. Arter. Throm Vas. 2019, 39, 1614–1628. [Google Scholar] [CrossRef]
  45. Liu, X.; Liu, J.; Li, Y.; Zhang, H. The Correlation between the Inflammatory Effects of Activated Macrophages in Atherosclerosis and Aortic Dissection. Ann. Vasc. Surg. 2022, 85, 341–346. [Google Scholar] [CrossRef]
  46. Zeng, T.; Yuan, J.; Gan, J.; Liu, Y.; Shi, L.; Lu, Z.; Xue, Y.; Xiong, R.; Huang, M.; Yang, Z.; et al. Thrombospondin 1 Is Increased in the Aorta and Plasma of Patients With Acute Aortic Dissection. Can. J. Cardiol. 2019, 35, 42–50. [Google Scholar] [CrossRef]
  47. Aoki, H.; Majima, R.; Hashimoto, Y.; Hirakata, S.; Ohno-Urabe, S. Ying and Yang of Stat3 in pathogenesis of aortic dissection. J. Cardiol. 2021, 77, 471–474. [Google Scholar] [CrossRef]
  48. Ren, W.; Wang, Z.; Wang, J.; Wu, Z.; Ren, Q.; Yu, A.; Ruan, Y. IL-5 overexpression attenuates aortic dissection by reducing inflammation and smooth muscle cell apoptosis. Life Sci. 2020, 241, 117144. [Google Scholar] [CrossRef]
  49. Luo, W.; Wang, Y.; Zhang, L.; Ren, P.; Zhang, C.; Li, Y.; Azares, A.R.; Zhang, M.; Guo, J.; Ghaghada, K.B.; et al. Critical Role of Cytosolic DNA and Its Sensing Adaptor STING in Aortic Degeneration, Dissection, and Rupture. Circulation 2020, 141, 42–66. [Google Scholar] [CrossRef]
  50. Cai, D.; Sun, C.; Murashita, T.; Que, X.; Chen, S.Y. ADAR1 Non-Editing Function in Macrophage Activation and Abdominal Aortic Aneurysm. Circ. Res. 2023, 132, e78–e93. [Google Scholar] [CrossRef]
  51. Gao, R.; Liu, D.; Guo, W.; Ge, W.; Fan, T.; Li, B.; Wang, J. Meprin-α (Mep1A) enhances TNF-α secretion by mast cells and aggravates abdominal aortic aneurysms. Brit J. Pharmacol. 2020, 177, 2872–2885. [Google Scholar] [CrossRef] [PubMed]
  52. Springer, J.M.; Raveendran, V.V.; Zhang, M.; Funk, R.; Smith, D.D.; Maz, M.; Dileepan, K.N. Mast Cell Degranulation Decreases Lipopolysaccharide-Induced Aortic Gene Expression and Systemic Levels of Interleukin-6 In Vivo. Mediat. Inflamm. 2019, 2019, 3856360. [Google Scholar] [CrossRef]
  53. Swedenborg, J.; Mäyränpää, M.I.; Kovanen, P.T. Mast cells: Important players in the orchestrated pathogenesis of abdominal aortic aneurysms. Arter. Throm Vas. 2011, 31, 734–740. [Google Scholar] [CrossRef] [PubMed]
  54. Tian, Z.; Zhang, P.; Li, X.; Jiang, D. Analysis of immunogenic cell death in ascending thoracic aortic aneurysms based on single-cell sequencing data. Front. Immunol. 2023, 14, 1087978. [Google Scholar] [CrossRef] [PubMed]
  55. Watanabe, R.; Hashimoto, M. Vasculitogenic T Cells in Large Vessel Vasculitis. Front. Immunol. 2022, 13, 923582. [Google Scholar] [CrossRef]
  56. Song, M.; Deng, L.; Shen, H.; Zhang, G.; Shi, H.; Zhu, E.; Xia, Q.; Han, H. Th1, Th2, and Th17 cells are dysregulated, but only Th17 cells relate to C-reactive protein, D-dimer, and mortality risk in Stanford type A aortic dissection patients. J. Clin. Lab. Anal. 2022, 36, e24469. [Google Scholar] [CrossRef]
  57. He, R.; Guo, D.-C.; Estrera, A.L.; Safi, H.J.; Huynh, T.T.; Yin, Z.; Cao, S.-N.; Lin, J.; Kurian, T.; Buja, L.M.; et al. Characterization of the inflammatory and apoptotic cells in the aortas of patients with ascending thoracic aortic aneurysms and dissections. J. Thorac. Cardiov Sur. 2006, 131, 671–678. [Google Scholar] [CrossRef]
  58. Liu, Y.; Zou, L.; Tang, H.; Li, J.; Liu, H.; Jiang, X.; Jiang, B.; Dong, Z.; Fu, W. Single-Cell Sequencing of Immune Cells in Human Aortic Dissection Tissue Provides Insights Into Immune Cell Heterogeneity. Front. Cardiovasc. Med. 2022, 9, 791875. [Google Scholar] [CrossRef]
  59. Lu, S.; White, J.V.; Nwaneshiudu, I.; Nwaneshiudu, A.; Monos, D.S.; Solomides, C.C.; Oleszak, E.L.; Platsoucas, C.D. Human abdominal aortic aneurysm (AAA): Evidence for an autoimmune antigen-driven disease. Autoimmun. Rev. 2022, 21, 103164. [Google Scholar] [CrossRef]
  60. Li, J.; Xia, N.; Li, D.; Wen, S.; Qian, S.; Lu, Y.; Gu, M.; Tang, T.; Jiao, J.; Lv, B.; et al. Aorta Regulatory T Cells with a Tissue-Specific Phenotype and Function Promote Tissue Repair through Tff1 in Abdominal Aortic Aneurysms. Adv. Sci. 2022, 9, e2104338. [Google Scholar] [CrossRef]
  61. Del Porto, F.; Cifani, N.; Proietta, M.; Dezi, T.; Tritapepe, L.; Raffa, S.; Micaloni, A.; Taurino, M. Lag3(+) regulatory T lymphocytes in critical carotid artery stenosis. Clin. Exp. Med. 2019, 19, 463–468. [Google Scholar] [CrossRef] [PubMed]
  62. Kuan, R.; Agrawal, D.K.; Thankam, F.G. Treg cells in atherosclerosis. Mol. Biol. Rep. 2021, 48, 4897–4910. [Google Scholar] [CrossRef] [PubMed]
  63. Lee, G.R. The Balance of Th17 versus Treg Cells in Autoimmunity. Int. J. Mol. Sci. 2018, 19, 730. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, Z.; Liu, Z.; Wang, K.; Ye, Z.; Xiong, Y.; Zhang, B.; Liao, J.; Zeng, L.; Zeng, H.; Liu, G.; et al. Reduced Number and Activity of Circulating Endothelial Progenitor Cells in Acute Aortic Dissection and Its Relationship with IL-6 and IL-17. Front. Cardiovasc. Med. 2021, 8, 628462. [Google Scholar] [CrossRef]
  65. Ju, X.; Ijaz, T.; Sun, H.; Ray, S.; Lejeune, W.; Lee, C.; Recinos, A.; Guo, D.-C.; Milewicz, D.M.; Tilton, R.G.; et al. Interleukin-6-signal transducer and activator of transcription-3 signaling mediates aortic dissections induced by angiotensin II via the T-helper lymphocyte 17-interleukin 17 axis in C57BL/6 mice. Arter. Throm Vas. 2013, 33, 1612–1621. [Google Scholar] [CrossRef]
  66. Gao, Y.; Wang, Z.; Zhao, J.; Sun, W.; Guo, J.; Yang, Z.; Tu, Y.; Yu, C.; Pan, L.; Zheng, J. Involvement of B cells in the pathophysiology of β-aminopropionitrile-induced thoracic aortic dissection in mice. Exp. Anim. 2019, 68, 331–339. [Google Scholar] [CrossRef]
  67. Hou, Y.; Li, Y.; Liu, B.; Wan, H.; Liu, C.; Xia, W. Research Progress on B Cells and Thoracic Aortic Aneurysm/Dissection. Ann. Vasc. Surg. 2022, 82, 377–382. [Google Scholar] [CrossRef]
  68. Schaheen, B.; Downs, E.A.; Serbulea, V.; Almenara, C.C.; Spinosa, M.; Su, G.; Zhao, Y.; Srikakulapu, P.; Butts, C.; McNamara, C.A.; et al. B-Cell Depletion Promotes Aortic Infiltration of Immunosuppressive Cells and Is Protective of Experimental Aortic Aneurysm. Arter. Throm Vas. 2016, 36, 2191–2202. [Google Scholar] [CrossRef]
  69. Meher, A.K.; Johnston, W.F.; Lu, G.; Pope, N.H.; Bhamidipati, C.M.; Harmon, D.B.; Su, G.; Zhao, Y.; McNamara, C.A.; Upchurch, G.R.; et al. B2 cells suppress experimental abdominal aortic aneurysms. Am. J. Pathol. 2014, 184, 3130–3141. [Google Scholar] [CrossRef]
  70. Belambri, S.A.; Rolas, L.; Raad, H.; Hurtado-Nedelec, M.; Dang, P.M.; El-Benna, J. NADPH oxidase activation in neutrophils: Role of the phosphorylation of its subunits. Eur. J. Clin. Investig. 2018, 48 (Suppl. S2), e12951. [Google Scholar] [CrossRef]
  71. Ravindran, M.; Khan, M.A.; Palaniyar, N. Neutrophil Extracellular Trap Formation: Physiology, Pathology, and Pharmacology. Biomolecules 2019, 9, 365. [Google Scholar] [CrossRef] [PubMed]
  72. Liu, H.; Xiao, T.; Zhang, L.; Huang, Y.; Shi, Y.; Ji, Q.; Liu, L. Effects of circulating levels of Th17 cells on the outcomes of acute Stanford B aortic dissection patients after thoracic endovascular aortic repair: A 36-month follow-up study a cohort study. Medicine 2019, 98, e18241. [Google Scholar] [CrossRef] [PubMed]
  73. Matsushima, K.; Yang, D.; Oppenheim, J.J. Interleukin-8: An evolving chemokine. Cytokine 2022, 153, 155828. [Google Scholar] [CrossRef] [PubMed]
  74. Wenjing, F.; Tingting, T.; Qian, Z.; Hengquan, W.; Simin, Z.; Agyare, O.K.; Shunlin, Q. The role of IL-1β in aortic aneurysm. Clin. Chim. Acta 2020, 504, 7–14. [Google Scholar] [CrossRef]
  75. Tang, N.; Sun, B.; Gupta, A.; Rempel, H.; Pulliam, L. Monocyte exosomes induce adhesion molecules and cytokines via activation of NF-κB in endothelial cells. Faseb J. 2016, 30, 3097–3106. [Google Scholar] [CrossRef]
  76. Liu, Z.; Luo, H.; Zhang, L.; Huang, Y.; Liu, B.; Ma, K.; Feng, J.; Xie, J.; Zheng, J.; Hu, J.; et al. Hyperhomocysteinemia exaggerates adventitial inflammation and angiotensin II-induced abdominal aortic aneurysm in mice. Circ. Res. 2012, 111, 1261–1273. [Google Scholar] [CrossRef]
  77. Shao, F.; Miao, Y.; Zhang, Y.; Han, L.; Ma, X.; Deng, J.; Jiang, C.; Kong, W.; Xu, Q.; Feng, J.; et al. B cell-derived anti-beta 2 glycoprotein I antibody contributes to hyperhomocysteinaemia-aggravated abdominal aortic aneurysm. Cardiovasc. Res. 2020, 116, 1897–1909. [Google Scholar] [CrossRef]
  78. Werle, M.; Kreuzer, J.; Höfele, J.; Elsässer, A.; Ackermann, C.; Katus, H.A.; Vogt, A.M. Metabolic control analysis of the Warburg-effect in proliferating vascular smooth muscle cells. J. Biomed. Sci. 2005, 12, 827–834. [Google Scholar] [CrossRef]
  79. Kornberg, M.D.; Bhargava, P.; Kim, P.M.; Putluri, V.; Snowman, A.M.; Putluri, N.; Snyder, S.H. Dimethyl fumarate targets GAPDH and aerobic glycolysis to modulate immunity. Science 2018, 360, 449–453. [Google Scholar] [CrossRef]
  80. Zhang, K.; Qi, Y.; Wang, M.; Chen, Q. Long non-coding RNA HIF1A-AS2 modulates the proliferation, migration, and phenotypic switch of aortic smooth muscle cells in aortic dissection via sponging microRNA-33b. Bioengineered 2022, 13, 6383–6395. [Google Scholar] [CrossRef]
  81. Wan, H.; Liu, D.; Liu, B.; Sha, M.; Xia, W.; Liu, C. Bioinformatics analysis of aging-related genes in thoracic aortic aneurysm and dissection. Front. Cardiovasc. Med. 2023, 10, 1089312. [Google Scholar] [CrossRef] [PubMed]
  82. Bao, X.; Zhang, J.; Huang, G.; Yan, J.; Xu, C.; Dou, Z.; Sun, C.; Zhang, H. The crosstalk between HIFs and mitochondrial dysfunctions in cancer development. Cell Death Dis. 2021, 12, 215. [Google Scholar] [CrossRef] [PubMed]
  83. Kashihara, T.; Mukai, R.; Oka, S.-I.; Zhai, P.; Nakada, Y.; Yang, Z.; Mizushima, W.; Nakahara, T.; Warren, J.S.; Abdellatif, M.; et al. YAP mediates compensatory cardiac hypertrophy through aerobic glycolysis in response to pressure overload. J. Clin. Investig. 2022, 132. [Google Scholar] [CrossRef] [PubMed]
  84. Bierhansl, L.; Conradi, L.C.; Treps, L.; Dewerchin, M.; Carmeliet, P. Central Role of Metabolism in Endothelial Cell Function and Vascular Disease. Physiology 2017, 32, 126–140. [Google Scholar] [CrossRef]
  85. Zhou, Q.; Xu, J.; Liu, M.; He, L.; Zhang, K.; Yang, Y.; Yang, X.; Zhou, H.; Tang, M.; Lu, L.; et al. Warburg effect is involved in apelin-13-induced human aortic vascular smooth muscle cells proliferation. J. Cell Physiol. 2019, 234, 14413–14421. [Google Scholar] [CrossRef]
  86. Chiong, M.; Morales, P.E.; Torres, G.; Gutiérrez, T.; García, L.; Ibacache, M.; Michea, L. Influence of glucose metabolism on vascular smooth muscle cell proliferation. Vasa 2013, 42, 8–16. [Google Scholar] [CrossRef]
  87. Kim, J.-H.; Bae, K.-H.; Byun, J.-K.; Lee, S.; Kim, J.-G.; Lee, I.K.; Jung, G.-S.; Lee, Y.M.; Park, K.-G. Lactate dehydrogenase-A is indispensable for vascular smooth muscle cell proliferation and migration. Biochem. Biophys. Res. Commun. 2017, 492, 41–47. [Google Scholar] [CrossRef]
  88. Wu, X.; Ye, J.; Cai, W.; Yang, X.; Zou, Q.; Lin, J.; Zheng, H.; Wang, C.; Chen, L.; Li, Y. LDHA mediated degradation of extracellular matrix is a potential target for the treatment of aortic dissection. Pharmacol. Res. 2022, 176, 106051. [Google Scholar] [CrossRef]
  89. Sun, L.-Y.; Lyu, Y.-Y.; Zhang, H.-Y.; Shen, Z.; Lin, G.-Q.; Geng, N.; Wang, Y.-L.; Huang, L.; Feng, Z.-H.; Guo, X.; et al. Nuclear Receptor NR1D1 Regulates Abdominal Aortic Aneurysm Development by Targeting the Mitochondrial Tricarboxylic Acid Cycle Enzyme Aconitase-2. Circulation 2022, 146, 1591–1609. [Google Scholar] [CrossRef]
  90. Alamri, A.; Burzangi, A.S.; Coats, P.; Watson, D.G. Untargeted Metabolic Profiling Cell-Based Approach of Pulmonary Artery Smooth Muscle Cells in Response to High Glucose and the Effect of the Antioxidant Vitamins D and E. Metabolites 2018, 8, 87. [Google Scholar] [CrossRef]
  91. Fan, L.; Meng, K.; Meng, F.; Wu, Y.; Lin, L. Metabolomic characterization benefits the identification of acute lung injury in patients with type A acute aortic dissection. Front. Mol. Biosci. 2023, 10, 1222133. [Google Scholar] [CrossRef]
  92. Kim, B.; Arany, Z. Endothelial Lipid Metabolism. Csh Perspect. Med. 2022, 12, a041162. [Google Scholar] [CrossRef] [PubMed]
  93. Zechner, R.; Zimmermann, R.; Eichmann, T.O.; Kohlwein, S.D.; Haemmerle, G.; Lass, A.; Madeo, F. FAT SIGNALS—Lipases and lipolysis in lipid metabolism and signaling. Cell Metab. 2012, 15, 279–291. [Google Scholar] [CrossRef] [PubMed]
  94. Li, R.; Zhang, C.; Du, X.; Chen, S. Genetic Association between the Levels of Plasma Lipids and the Risk of Aortic Aneurysm and Aortic Dissection: A Two-Sample Mendelian Randomization Study. J. Clin. Med. 2023, 12, 1991. [Google Scholar] [CrossRef] [PubMed]
  95. Lindholt, J.S.; Kristensen, K.L.; Burillo, E.; Martinez-Lopez, D.; Calvo, C.; Ros, E.; Sala-Vila, A. Arachidonic Acid, but Not Omega-3 Index, Relates to the Prevalence and Progression of Abdominal Aortic Aneurysm in a Population-Based Study of Danish Men. J. Am. Heart Assoc. 2018, 7, e007790. [Google Scholar] [CrossRef]
  96. Liu, B.; Kong, J.; An, G.; Zhang, K.; Qin, W.; Meng, X. Regulatory T cells protected against abdominal aortic aneurysm by suppression of the COX-2 expression. J. Cell Mol. Med. 2019, 23, 6766–6774. [Google Scholar] [CrossRef]
  97. Lamers, D.; Schlich, R.; Greulich, S.; Sasson, S.; Sell, H.; Eckel, J. Oleic acid and adipokines synergize in inducing proliferation and inflammatory signalling in human vascular smooth muscle cells. J. Cell Mol. Med. 2011, 15, 1177–1188. [Google Scholar] [CrossRef]
  98. Nettersheim, F.S.; Lemties, J.; Braumann, S.; Geißen, S.; Bokredenghel, S.; Nies, R.; Hof, A.; Winkels, H.; Freeman, B.A.; Klinke, A.; et al. Nitro-oleic acid reduces thoracic aortic aneurysm progression in a mouse model of Marfan syndrome. Cardiovasc. Res. 2022, 118, 2211–2225. [Google Scholar] [CrossRef]
  99. Al, M.M.; Patel, P.A.; Chamberlain, J.; Francis, S. OxLDL induces IL-1β release from human EC and VSMC via different caspase-1 dependent mechanisms. Vasc. Biol. 2022, 4, 11–18. [Google Scholar]
  100. Chen, Y.; Yi, X.; Huo, B.; He, Y.; Guo, X.; Zhang, Z.; Zhong, X.; Feng, X.; Fang, Z.-M.; Zhu, X.-H.; et al. BRD4770 functions as a novel ferroptosis inhibitor to protect against aortic dissection. Pharmacol. Res. 2022, 177, 106122. [Google Scholar] [CrossRef]
  101. Steinmetz, E.F.; Buckley, C.; Shames, M.L.; Ennis, T.L.B.; Vanvickle-Chavez, S.J.B.; Mao, D.; Goeddel, L.A.; Hawkins, C.J.; Thompson, R.W. Treatment with simvastatin suppresses the development of experimental abdominal aortic aneurysms in normal and hypercholesterolemic mice. Ann. Surg. 2005, 241, 92–101. [Google Scholar] [CrossRef] [PubMed]
  102. Wang, L.; Liu, S.; Yang, W.; Yu, H.; Zhang, L.; Ma, P.; Wu, P.; Li, X.; Cho, K.; Xue, S.; et al. Plasma Amino Acid Profile in Patients with Aortic Dissection. Sci. Rep. 2017, 7, 40146. [Google Scholar] [CrossRef] [PubMed]
  103. Zhou, Y.; Wang, T.; Fan, H.; Liu, S.; Teng, X.; Shao, L.; Shen, Z. Research Progress on the Pathogenesis of Aortic Aneurysm and Dissection in Metabolism. Curr. Probl. Cardiol. 2023, 49, 102040. [Google Scholar] [CrossRef] [PubMed]
  104. Huang, T.H.; Chang, H.H.; Guo, Y.R.; Chang, W.C.; Chen, Y.F. Vitamin B Mitigates Thoracic Aortic Dilation in Marfan Syndrome Mice by Restoring the Canonical TGF-β Pathway. Int. J. Mol. Sci. 2021, 22, 11737. [Google Scholar] [CrossRef] [PubMed]
  105. Dai, X.; Liu, S.; Cheng, L.; Huang, T.; Guo, H.; Wang, D.; Xia, M.; Ling, W.; Xiao, Y. Epigenetic Upregulation of H19 and AMPK Inhibition Concurrently Contribute to S-Adenosylhomocysteine Hydrolase Deficiency-Promoted Atherosclerotic Calcification. Circ. Res. 2022, 130, 1565–1582. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, C.-Y.; Hu, Y.-C.; Zhang, Y.; Ma, W.-D.; Song, Y.-F.; Quan, X.-H.; Guo, X.; Wang, C.-X. Glutamine switches vascular smooth muscle cells to synthetic phenotype through inhibiting miR-143 expression and upregulating THY1 expression. Life Sci. 2021, 277, 119365. [Google Scholar] [CrossRef]
  107. Osman, I.; He, X.; Liu, J.; Dong, K.; Wen, T.; Zhang, F.; Yu, L.; Hu, G.; Xin, H.-B.; Zhang, W.; et al. TEAD1 (TEA Domain Transcription Factor 1) Promotes Smooth Muscle Cell Proliferation Through Upregulating SLC1A5 (Solute Carrier Family 1 Member 5)-Mediated Glutamine Uptake. Circ. Res. 2019, 124, 1309–1322. [Google Scholar] [CrossRef]
  108. Gallina, A.L.; Rykaczewska, U.; Wirka, R.C.; Caravaca, A.S.; Shavva, V.S.; Youness, M.; Karadimou, G.; Lengquist, M.; Razuvaev, A.; Paulsson-Berne, G.; et al. AMPA-Type Glutamate Receptors Associated With Vascular Smooth Muscle Cell Subpopulations in Atherosclerosis and Vascular Injury. Front. Cardiovasc. Med. 2021, 8, 655869. [Google Scholar] [CrossRef]
  109. Zheng, Y.; Hu, F.B.; Ruiz-Canela, M.; Clish, C.B.; Dennis, C.; Salas-Salvado, J.; Hruby, A.; Liang, L.; Toledo, E.; Corella, D.; et al. Metabolites of Glutamate Metabolism Are Associated With Incident Cardiovascular Events in the PREDIMED PREvención con DIeta MEDiterránea (PREDIMED) Trial. J. Am. Heart Assoc. 2016, 5, e003755. [Google Scholar] [CrossRef]
  110. Zhang, J.; Chen, C.; Li, L.; Zhou, H.J.; Li, F.; Zhang, H.; Yu, L.; Chen, Y.; Min, W. Endothelial AIP1 Regulates Vascular Remodeling by Suppressing NADPH Oxidase-2. Front. Physiol. 2018, 9, 396. [Google Scholar] [CrossRef]
  111. Cui, H.; Chen, Y.; Li, K.; Zhan, R.; Zhao, M.; Xu, Y.; Lin, Z.; Fu, Y.; He, Q.; Tang, P.C.; et al. Untargeted metabolomics identifies succinate as a biomarker and therapeutic target in aortic aneurysm and dissection. Eur. Heart J. 2021, 42, 4373–4385. [Google Scholar] [CrossRef] [PubMed]
  112. Luo, S.; Kong, C.; Zhao, S.; Tang, X.; Wang, Y.; Zhou, X.; Li, R.; Liu, X.; Tang, X.; Sun, S.; et al. Endothelial HDAC1-ZEB2-NuRD Complex Drives Aortic Aneurysm and Dissection Through Regulation of Protein S-Sulfhydration. Circulation 2023, 147, 1382–1403. [Google Scholar] [CrossRef] [PubMed]
  113. Khatib-Massalha, E.; Bhattacharya, S.; Massalha, H.; Biram, A.; Golan, K.; Kollet, O.; Kumari, A.; Avemaria, F.; Petrovich-Kopitman, E.; Gur-Cohen, S.; et al. Lactate released by inflammatory bone marrow neutrophils induces their mobilization via endothelial GPR81 signaling. Nat. Commun. 2020, 11, 3547. [Google Scholar] [CrossRef] [PubMed]
  114. Cichon, I.; Ortmann, W.; Kolaczkowska, E. Metabolic Pathways Involved in Formation of Spontaneous and Lipopolysaccharide-Induced Neutrophil Extracellular Traps (NETs) Differ in Obesity and Systemic Inflammation. Int. J. Mol. Sci. 2021, 22, 7718. [Google Scholar] [CrossRef] [PubMed]
  115. Alarcón, P.; Manosalva, C.; Conejeros, I.; Carretta, M.D.; Muñoz-Caro, T.; Silva, L.M.R.; Taubert, A.; Hermosilla, C.; Hidalgo, M.A.; Burgos, R.A. d(-) Lactic Acid-Induced Adhesion of Bovine Neutrophils onto Endothelial Cells Is Dependent on Neutrophils Extracellular Traps Formation and CD11b Expression. Front. Immunol. 2017, 8, 975. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, P.-S.; Wang, H.; Li, X.; Chao, T.; Teav, T.; Christen, S.; Di Conza, G.; Cheng, W.-C.; Chou, C.-H.; Vavakova, M.; et al. α-ketoglutarate orchestrates macrophage activation through metabolic and epigenetic reprogramming. Nat. Immunol. 2017, 18, 985–994. [Google Scholar] [CrossRef]
  117. Van Der Valk, F.M.; Bekkering, S.; Kroon, J.; Yeang, C.; Van den Bossche, J.; Van Buul, J.D.; Ravandi, A.; Nederveen, A.J.; Verberne, H.J.; Scipione, C.; et al. Oxidized Phospholipids on Lipoprotein(a) Elicit Arterial Wall Inflammation and an Inflammatory Monocyte Response in Humans. Circulation 2016, 134, 611–624. [Google Scholar] [CrossRef]
  118. Wei, X.; Song, H.; Yin, L.; Rizzo, M.G.; Sidhu, R.; Covey, D.F.; Ory, R.S.D.S.; Semenkovich, C.F. Fatty acid synthesis configures the plasma membrane for inflammation in diabetes. Nature 2016, 539, 294–298. [Google Scholar] [CrossRef]
  119. Pithon-Curi, T.C.; Levada, A.C.; Lopes, L.R.; Doi, S.Q.; Curi, R. Glutamine plays a role in superoxide production and the expression of p47phox, p22phox and gp91phox in rat neutrophils. Clin. Sci. 2002, 103, 403–408. [Google Scholar] [CrossRef]
  120. Durante, W. The Emerging Role of l-Glutamine in Cardiovascular Health and Disease. Nutrients. 2019, 11, 2092. [Google Scholar] [CrossRef]
  121. Cruzat, V.; Macedo, R.M.; Noel, K.K.; Curi, R.; Newsholme, P. Glutamine: Metabolism and Immune Function, Supplementation and Clinical Translation. Nutrients 2018, 10, 1564. [Google Scholar] [CrossRef] [PubMed]
  122. Makowski, L.; Chaib, M.; Rathmell, J.C. Immunometabolism: From basic mechanisms to translation. Immunol. Rev. 2020, 295, 5–14. [Google Scholar] [CrossRef] [PubMed]
  123. Mittal, M.; Siddiqui, M.R.; Tran, K.; Reddy, S.P.; Malik, A.B. Reactive oxygen species in inflammation and tissue injury. Antioxid. Redox Sign. 2014, 20, 1126–1167. [Google Scholar] [CrossRef] [PubMed]
  124. Hu, X.; Jiang, W.; Wang, Z.; Li, L.; Hu, Z. NOX1 Negatively Modulates Fibulin-5 in Vascular Smooth Muscle Cells to Affect Aortic Dissection. Biol. Pharm. Bull. 2019, 42, 1464–1470. [Google Scholar] [CrossRef]
  125. Badran, A.; Nasser, S.A.; Mesmar, J.; El-Yazbi, A.F.; Bitto, A.; Fardoun, M.M.; Baydoun, E.; Eid, A.H. Reactive Oxygen Species: Modulators of Phenotypic Switch of Vascular Smooth Muscle Cells. Int. J. Mol. Sci. 2020, 21, 8764. [Google Scholar] [CrossRef]
  126. Fukai, T.; Ushio-Fukai, M. Cross-Talk between NADPH Oxidase and Mitochondria: Role in ROS Signaling and Angiogenesis. Cells 2020, 9, 1849. [Google Scholar] [CrossRef]
  127. Lian, G.; Li, X.; Zhang, L.; Zhang, Y.; Sun, L.; Zhang, X.; Jiang, C. Macrophage metabolic reprogramming aggravates aortic dissection through the HIF1α-ADAM17 pathway. Ebiomedicine 2019, 49, 291–304. [Google Scholar] [CrossRef]
  128. Wang, T.; Liu, H.; Lian, G.; Zhang, S.Y.; Wang, X.; Jiang, C. HIF1α-Induced Glycolysis Metabolism Is Essential to the Activation of Inflammatory Macrophages. Mediat. Inflamm. 2017, 2017, 9029327. [Google Scholar] [CrossRef]
  129. Lv, S.-L.; Zeng, Z.-F.; Gan, W.-Q.; Wang, W.-Q.; Li, T.-G.; Hou, Y.-F.; Yan, Z.; Zhang, R.-X.; Yang, M. Lp-PLA2 inhibition prevents Ang II-induced cardiac inflammation and fibrosis by blocking macrophage NLRP3 inflammasome activation. Acta Pharmacol. Sin. 2021, 42, 2016–2032. [Google Scholar] [CrossRef]
  130. Acosta, S.; Taimour, S.; Gottsäter, A.; Persson, M.; Engström, G.; Melander, O.; Zarrouk, M.; Nilsson, P.M.; Smith, J.G. Lp-PLA(2) activity and mass for prediction of incident abdominal aortic aneurysms: A prospective longitudinal cohort study. Atherosclerosis 2017, 262, 14–18. [Google Scholar] [CrossRef]
  131. Anagnostopoulou, A.; Camargo, L.L.; Rodrigues, D.; Montezano, A.C.; Touyz, R.M. Importance of cholesterol-rich microdomains in the regulation of Nox isoforms and redox signaling in human vascular smooth muscle cells. Sci. Rep. 2020, 10, 17818. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, W.; Hou, S.; Han, M.; Zheng, Y.; Ma, R.; Liu, Z.; Lyu, Y.; Wu, X.; Liu, R. Research progress in metabolic reprogramming of endothelial cells and T cells and potential therapeutic targets for chronic inflammation. Xi Bao Yu Fen Zi Mian Yi Xue Za Zhi 2021, 37, 1038–1044. [Google Scholar] [PubMed]
Figure 1. Schematic diagram of the interaction and influence between the immune metabolic and parenchymal cells. Innate immunity and adaptive immunity continuously interact to produce an immune response that changes the immune microenvironment; meanwhile, the release of inflammatory factors, proteases, etc., results in abnormal metabolism of parenchymal cells. The glycolysis of glucose in VSMCs becomes the main metabolic pathway rather than OXPHOX, known as the Warburg effect. The excessive accumulation of FFAs and SFA in the lipid metabolism process of ECs can result in oxidative stress. Amino acid products of the TCA cycle are reduced in fibroblasts, whereas HCY accounts for major metabolism which can lead to OXPHPX dysfunction. These metabolic abnormalities eventually lead to apoptosis, necrosis and phenotypic transformation of parenchymal cells. In the process of degrading proteins, ROS and superoxide produced will also cause metabolic reprogramming of immune cells, which will cause abnormality and heterogeneity of immune cells so that they perpetuate inflammation and damage of aortic wall. This vicious cycle will eventually accelerate the process of AD. (NETs: neutrophil extracellular traps, ROS: reactive oxygen species, OXPHOX: oxidative phosphorylation, USFA: unsaturated fatty acid, FAs: fatty acid, SFA: saturated fatty acid, FFAs: free fatty acid, TCA: tricarboxylic acid cycle, Hcy: homocysteine, VSMCs: vascular smooth muscle cells, ECs: endothelial cells).
Figure 1. Schematic diagram of the interaction and influence between the immune metabolic and parenchymal cells. Innate immunity and adaptive immunity continuously interact to produce an immune response that changes the immune microenvironment; meanwhile, the release of inflammatory factors, proteases, etc., results in abnormal metabolism of parenchymal cells. The glycolysis of glucose in VSMCs becomes the main metabolic pathway rather than OXPHOX, known as the Warburg effect. The excessive accumulation of FFAs and SFA in the lipid metabolism process of ECs can result in oxidative stress. Amino acid products of the TCA cycle are reduced in fibroblasts, whereas HCY accounts for major metabolism which can lead to OXPHPX dysfunction. These metabolic abnormalities eventually lead to apoptosis, necrosis and phenotypic transformation of parenchymal cells. In the process of degrading proteins, ROS and superoxide produced will also cause metabolic reprogramming of immune cells, which will cause abnormality and heterogeneity of immune cells so that they perpetuate inflammation and damage of aortic wall. This vicious cycle will eventually accelerate the process of AD. (NETs: neutrophil extracellular traps, ROS: reactive oxygen species, OXPHOX: oxidative phosphorylation, USFA: unsaturated fatty acid, FAs: fatty acid, SFA: saturated fatty acid, FFAs: free fatty acid, TCA: tricarboxylic acid cycle, Hcy: homocysteine, VSMCs: vascular smooth muscle cells, ECs: endothelial cells).
Ijms 24 15908 g001
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

Li, S.; Li, J.; Cheng, W.; He, W.; Dai, S.-S. Independent and Interactive Roles of Immunity and Metabolism in Aortic Dissection. Int. J. Mol. Sci. 2023, 24, 15908. https://doi.org/10.3390/ijms242115908

AMA Style

Li S, Li J, Cheng W, He W, Dai S-S. Independent and Interactive Roles of Immunity and Metabolism in Aortic Dissection. International Journal of Molecular Sciences. 2023; 24(21):15908. https://doi.org/10.3390/ijms242115908

Chicago/Turabian Style

Li, Siyu, Jun Li, Wei Cheng, Wenhui He, and Shuang-Shuang Dai. 2023. "Independent and Interactive Roles of Immunity and Metabolism in Aortic Dissection" International Journal of Molecular Sciences 24, no. 21: 15908. https://doi.org/10.3390/ijms242115908

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